<<

arXiv:cond-mat/0509373v2 [cond-mat.other] 15 Dec 2005 h trciesd ( side attractive the et ilas eal osuytesprudBSphase BCS superfluid the experi- study to future ( able that be also suggest will temperatures reached ments low currently the are but region, which crossover the on trated ntetmeauerne0 approach, range this temperature In the super- [8]. in trapped temperature non-zero in at modes gases collective fluid describe to used been temperature. also zero is at it done not [7], are is experiments experiments which the that in clear condition, fulfilled latter trap necessarily the the not from if Apart ap- temperature local-density zero justify proximation. to at smooth sufficiently superfluid is potential a which for 6], a valid 5, on is [4, based hydrodynamics” or usually “superfluid theoretical [2], called is The theory off experiments 3]. switched such [1, of been cloud interpretation the has of trap oscillations the collective the after of cloud expansion the atom like observables dynamical at atoms Fermionic looked trapped ultracold with experiments cent low. extremely be will ie h eprtr isblwacranciia tem- critical pro- certain superfluid, a is below system lies perature temperature the the that vided between expects distance of mean one which the side atoms, pairs with BCS compared Cooper the large form very on atoms are the as where of well crossover, (BEC) as the condensate molecules, Bose-Einstein bound a tightly forms system the iesd ( side sive n h tmao cteiglength scattering the atom-atom chang- tuning thus the to resonance, by Feshbach ing is a crossover around experiments BEC-BCS field these magnetic so-called for the motivation main study The ture. where eylwtmeaue fteodrof order the of temperatures low very et ihtapdfrincaoslike atoms fermionic trapped with ments < a eyrcnl,Lna’ w-udhdoyaishas hydrodynamics two-fluid Landau’s recently, Very re- some superfluidity, for signals find to order In u oipoe oln ehius urn experi- current techniques, cooling improved to Due and 0 T F T > a c = k ni o h xeiet aeconcen- have experiments the now Until . ASnmes 03.75.Ss,03.75.Kk,67.40.Bz sem numbers: PACS the dampi solving for Landau method strong numerical and a approximations. temperature suggest we critical addition the In diffe and th the approximations, zero are quadrupo simplifying which the at reproduced, some and are medium of calculations uniform spite mechanical a In in wave simp trap. two sound monic to a applied found: and version be out linearized worked can The is equilibrium, one. from critical tions hydr the superfluid at of equation tem cases Vlasov transition limiting well-known superfluid two the the between and zero between temperatures F | ǫ edrv eilsia rnpr qain o rpe a trapped a for equations transport semiclassical derive We )truhteuiaylmt( limit unitary the through 0) a F ≪ | .INTRODUCTION I. /k B ) lhuhisciia temperature critical its although 1), < a eoe h eeeaytempera- degeneracy the denotes yaiso rpe em a nteBSphase BCS the in gas Fermi trapped a of Dynamics ) nteBCsd,where side, BEC the On 0). T < T < ntttd hsqeNc´ar,F946OsyCee,Fra C´edex, Orsay Nucl´eaire, F-91406 Physique de Institut c eti fraction certain a , a T 6 rmterepul- the from ihe ra n ee Schuck Peter and Urban Michael ior Li ≈ a 0 . 40 ∞ → 03 reach K T F to ) [1], fteaosi o uefli u om normal-fluid a forms but superfluid density not with is component atoms the of ihdensity with scranytu nteBSpae where phase, BCS the in This true phase. certainly normal is the of behavior hydrodynamic ensure oee,cno etknfrgatd nRf 9 twas scattering it the where [9] limit, Ref. unitary length In the condition, in This granted. even that for equilibrium. found taken colli- local be enough in cannot undergo always however, atoms be the to that sions assumed also is It qain,weealqatte r oa ie,functions (i.e., local coordinate spatial are the quantities of hydrodynamical all in to where modes Contrary collective equations, [4]. of regime gases frequencies collisionless Fermi the the atomic re- predict trapped to to and to order order [18], applied in in nuclei also was atomic e.g., it in cently physics, resonances used nuclear was giant in latter describe The success equation. for great Vlasov method the with by Fermi semiclassical given normal a A is in gas deformation surface body. Fermi the elastic of treating but an possibility hydrodynamically, like behave the not more a of does of Because gas component in collisionless normal atoms 17]. the fermionic deformations, [16, trapped Fermi-surface also phase of has case BCS model the the two-fluid to the applied Recently been pa- order 15]. liquid the [14, of describe structure rameter spin superconductivity to the of approaches of because devel- complicated Similar theory been the 13]. has in 12, It 3 case 11, too. [10, this regime, for collisionless oped the in ful equilibrium local reach to this oscillation. impossible Ω, the frequency is during trap it the that of order implies collective the the of of colliding are frequencies the oscillations before Since atom trap an atom. the another i.e., with in Ω, oscillations frequency 1 several trap rate the performs collision than the Colli- smaller that much regime. context is collisionless this in so-called means the that sionless assume in safely is can system one the that small so are temperature esol emninda el lhuhte r more are they although well, as mentioned be should He eeteesteie fatofli oe suse- is model two-fluid a of idea the Nevertheless fteeeutos ai o ml devia- small for valid equations, these of etfeuniso h uduoemode quadrupole the of frequencies rent a dnmc tzr eprtr n the and temperature zero at odynamics anqaiaierslso quantum of results qualitative main e eaue hs qain interpolate equations These perature. eeape hr nltclsolutions analytical where examples le iegs h olso aemgtb o o to low too be might rate collision the diverges, oi em a nteBSpaeat phase BCS the in gas Fermi tomic cascleutoswtotfurther without equations iclassical ga nemdaetemperatures. intermediate at ng ρ s eectto nashrclhar- spherical a in excitation le = ρ − ρ nce n r ρ n ny,teVao qainre- equation Vlasov the only), tl eaelk superfluid. a like behave still hl h eann atoms remaining the while , k F | a | n the and /τ 2 quires a phase-space description (i.e., the quantities are II. FORMALISM functions of r and p). The aim of the present article is to derive a hydrodynamical equation for the superfluid com- A. Derivation of a transport equation ponent coupled to a Vlasov equation for the normal com- ponent, interpolating between superfluid hydrodynamics In this subsection we will derive a quasiparticle trans- at T = 0 and the usual Vlasov equation at T = Tc. In 3 port equation for a superfluid gas of trapped fermionic principle, as it was done in the theory of liquid He [14], atoms in the BCS phase. Throughout this article we one could also think of including a collision term into will assume that the two spin states and are equally this equation, in order to treat systems which are nei- populated, which allows us to remove↑ the spin↓ degree of ther collisionless nor hydrodynamical, but somewhere in freedom from the beginning. However, the generalization between. However, in the present article we will restrict to include the spin, which in fact would be necessary, e.g., ourselves to the collisionless case. in order to describe spin waves or systems with unequal populations, is straight-forward. In order to be in the Like the semiclassical description of the ground state BCS phase, the atoms must have an attractive interac- (Thomas-Fermi approximation), the semiclassical de- tion, i.e., a negative scattering length a < 0, which on scription of the dynamics of the system can be expected the other hand must be weak enough for the BCS ap- to become more and more accurate if the number of proximation to be valid. atoms in the trap increases. This was the main moti- Let us start by writing down the TDHFB equations vation for us to develop the semiclassical approach pre- [18]. To that end we define the non-local normal and sented here. A fully quantum-mechanical description anomalous density matrices, of the collective modes of a trapped Fermi gas can be ′ † ′ † ′ obtained, e.g., by the quasiparticle random-phase ap- ̺(r, r )= ψ (r )ψ↑(r) = ψ (r )ψ↓(r) , (1) h ↑ i h ↓ i proximation (QRPA), corresponding to the lineariza- ′ ′ ′ κ(r, r )= ψ↑(r )ψ↓(r) = ψ↓(r )ψ↑(r) , (2) tion of the time-dependent Bogoliubov-de Gennes equa- h i −h i tions around equilibrium. The latter are also known where ψ is the field operator. The single-particle hamil- as time-dependent Hartree-Fock-Bogoliubov (TDHFB) tonian (minus the chemical potential µ) reads equations, especially in nuclear physics. QRPA calcula- ~2∇2 tions become tremendously difficult and time-consuming h = + Vext (r)+ gρ(r) µ , (3) if the number of particles increases. At present, they are − 2m − restricted to systems of 104 atoms [7, 19, 20], while ∼ where m is the atomic mass, Vext (r) is the potential of the numbers of atoms in the experiments are at least the trap, gρ(r) is the mean-field potential. The coupling ten times larger. In addition, all present QRPA calcu- constant g is related to the atom-atom scattering length lations are done for the case of spherically symmetric a by g =4π~2a/m and the density per spin state ρ(r) is traps, while the traps used in the experiments are gener- just equal to the local part of the density matrix, ally not spherical. The numerical solution of the QRPA equations without spherical symmetry seems to be almost ρ(r)= ̺(r, r) . (4) unfeasible, unless one reduces drastically the number of particles. Therefore semiclassical approaches are at the According to the usual regularization prescription [21], moment the only way to perform calculations for large the pairing gap is related to the anomalous density by numbers of atoms in realistic trap geometries. d s s ∆(r)= g lim sκ r + , r . (5) − s→0 ds  2 − 2 Our article is organized as follows. In Sec. II we will Combining all quantities in the 2 2 matrices present the formalism. Having derived a quasiparticle × transport equation in Sec. II A, an important point will h ∆ ̺ κ = , = − , (6) be to work out the linearized version of this equation in H  ∆† h¯ R  κ† 1 ̺¯ order to apply it to oscillations around the equilibrium − − − − state. This is done in Sec. II B. In Sec. II C we will where̺ ¯ and h¯ denote the time-reversed operators to ̺ show explicitly that our equations indeed reproduce su- and h, respectively, the TDHFB equation can be written perfluid hydrodynamics and the Vlasov equation in the in the compact form [18] limits of zero and critical temperature, respectively. The i~ ˙ = , . (7) next part, Sec. III, is devoted to two simple examples R H R for which our equations can be solved more or less an-   In analogy to the derivation of the Vlasov equation in alytically. The first example, discussed in Sec. III A, is the normal phase from the Hartree-Fock equation [18], it a sound wave in a uniform gas. The second one, de- is useful to introduce the Wigner transform of the density scribed in Sec. III B, concerns a quadrupole oscillation of matrix, a harmonically trapped gas with some additional simpli- fications. Finally, in Sec. IV we will summarize and draw ~ s s ̺(r, p)= d3se−ip·s/ ̺ r + , r . (8) our conclusions. Z 2 − 2  3

It is appealing, although strictly speaking not correct, to operators̺ ¯ and h¯, and the Wigner transforms of the interprete the function ̺(r, p) as a distribution function adjoint operators κ† and ∆†. To that end we recall the of particles in phase space. In a completely analogous general relations way we define the Wigner transform of the anomalous density matrix, κ(r, p), and the Wigner transform of the A¯(r, p)= A(r, p) , [A†](r, p)= A∗(r, p) , (12) − hamiltonian, h(r, p), which is equal to the classical hamil- tonian which are valid for an arbitrary operator A. The use- fulness of the Wigner transform lies in the fact that, to p2 first order in an expansion into powers of ~, the Wigner h(r, p)= + Vext (r)+ gρ(r) µ . (9) 2m − transform of the product of two operators A and B can be obtained according to (For the sake of readability we are using the same sym- bol for the operators and their Wigner transforms, but i~ [AB](r, p) A(r, p)B(r, p)+ A(r, p),B(r, p) , whenever there is a risk of confusion we will write down ≈ 2 { } the arguments.) Eqs. (4) and (5) can be written in terms (13) of the Wigner transformed quantities as follows: where , denotes te Poisson bracket {· ·} d3p ∂A ∂B ∂A ∂B ρ(r)= ̺(r, p) , (10) A, B = . (14) Z (2π~)3 { } ∂r ∂p − ∂p ∂r i=Xx,y,z  i i i i  d3p ∆(r) ∆(r)= g κ(r, p) . (11) ~ 3 2 Applying this product rule to the Wigner transform of − Z (2π )  − p /m the TDHFB equation (7), one obtains four coupled equa- We also need the Wigner transforms of the time-reversed tions:

i~̺˙ = i~ h,̺ +2i Im(∆∗κ) i~ Re ∆∗,κ , (15a) { } − { } i~ i~ i~κ˙ = (h + h¯)κ + h h,κ¯ + ∆(̺ +̺ ¯ 1) ∆,̺ ̺¯ , (15b) 2 { − } − − 2 { − } i~ i~ i~κ˙ ∗ = (h + h¯)κ∗ + h h,κ¯ ∗ ∆∗(̺ +¯̺ 1) ∆∗,̺ ̺¯ , (15c) − 2 { − }− − − 2 { − } i~̺¯˙ = i~ h,¯ ̺¯ +2i Im(∆∗κ)+ i~ Re ∆∗,κ . (15d) − { } { }

In order to proceed further, it is useful to separate the transformed gap ∆˜ is real, we have completely sepa- the collective superfluid motion from the dynamics due rated the collective motion of the superfluid component to quasiparticle excitations. This can be achieved by a from the motion due to quasiparticle excitations. A for- gauge transformation, ψ˜(r) = ψ(r) exp[iφ(r)] [13, 14]. mal argument for the necessity of this choice of the gauge According to their definitions, the normal and anoma- is given in Ref. [14]. lous density matrices behave very differently under this From now on we will suppose that ∆˜ is real. Splitting transformation. The corresponding Wigner transforms ̺˜ and h˜ into time-even and time-odd parts, are given by

̺˜(r, p)= ̺[r, p ~∇φ(r)] , (16) 1 1 ̺˜ev = ̺˜ + ̺˜¯ , ̺˜od = ̺˜ ̺˜¯ , (20) − 2 2 − κ˜(r, p)= κ(r, p) e2iφ(r) . (17)   1 ¯ p2 (~∇φ)2 h˜ev = h˜ + h˜ = + + V µ ~φ˙ , (21) If the hamiltonian and the gap are changed according to 2 2m 2m − −  ~ ˜ 1 ˜ ˜¯ ~ 2 hod = h h = p ∇φ , (22) [p ∇φ(r)] 2 − −m · h˜(r, p)= − ~φ˙(r)+ V (r) µ , (18)  2m − − ∆(˜ r) = ∆(r) e2iφ(r) , (19) andκ ˜ into real and imaginary parts, the equation of motion of the gauge transformed quan- tities looks exactly like Eq. (7). The superfluid velocity κ˜re = Re˜κ , κ˜im = Im˜κ , (23) is proportional to the gradient of the phase of the gap. Hence, if we choose the gauge transformation such that one can rewrite the gauge transformed version of the sys- 4 tem of equations (15) as follows: From now on we will suppress the index “(0)” and (0) (0) simply write̺ ˜ev ,̺ ˜od , andκ ˜re instead of̺ ˜ev ,̺ ˜od , and ~̺˜˙ev = ~ h˜ev , ̺˜od + ~ h˜od , ̺˜ev +2∆˜˜ κim , (24a) (0) { } { } κ˜re . The next step is to exploit Eq. (28d) in order to ~̺˜˙od = ~ h˜ev , ̺˜ev + ~ h˜od , ̺˜od ~ ∆˜ , κ˜re , (24b) reduce the number of unknown functions. To that end { } { }− { } we introduce a new phase-space function νev (r, p), the ~κ˜˙ re =2h˜ev κ˜im + ~ h˜od , κ˜re ~ ∆˜ , ̺˜od , (24c) { }− { } so-called “quasiparticle distribution function”, which is ~κ˜˙ im = 2h˜ev κ˜re + ∆(1˜ 2˜̺ev )+ ~ h˜od , κ˜im . (24d) defined in such a way that the two members of Eq. (28d) − − { } are equal to h˜ev ∆(1˜ 2νev )/Eev . In other words,̺ ˜ev For a semiclassical ~ expansion it seems disturbing that − ~ andκ ˜re can be expressed in terms of this function νev as these equations mix different orders in . However, it follows: is possible to decouple the equations of motion for the leading-order quantities from those of the higher-order 1 h˜ev ̺˜ev = (1 2νev ) , (32) ones. In order to show this, we expand̺ ˜ andκ ˜ into 2 − 2Eev − ~ powers of . Since the Eqs. (24) themselves are only ∆ valid up to order ~, it does not make sense to go beyond κ˜re = (1 2νev ) . (33) the first order in this series. From Eqs. (24a) and (24c) 2Eev − it is evident thatκ ˜im must be suppressed by one power In fact, the definition of νev has been chosen such that of ~ with respect to the other quantities. We therefore these relations resemble the well-known expressions for write ̺ and κ in equilibrium, where νev has to be replaced by (0) (1) the Fermi distribution function for , f(E) ̺˜ev od =̺ ˜ev od + ~̺˜ev od + , (25) , , , ··· (see Sec. II B). With the help of Eqs. (32) and (33) the (0) (1) κ˜re =κ ˜re + ~κ˜re + , (26) remaining two equations, (28b) and (30), take the rather ··· (1) simple form κ˜im = ~κ˜im + . (27) ··· ̺˜˙od = Eev ,νev + h˜od , ̺˜od , (34a) Inserting these expansions into Eqs. (24a) – (24d) and { } { } ̺˜˙ev = Eev , ̺˜od + h˜od ,νev . (34b) retaining only the leading order in each equation [order { } { } ~ in the case of Eqs. (24a) – (24c), order 1 in the case of Since the first of these equations is purely time-odd while Eq. (24d)], one obtains the second one is purely time-even, we can add both equa- (0) (0) (0) (1) tions without any loss of information. The result can be ̺˜˙ev = h˜ev , ̺˜od + h˜od , ̺˜ev +2∆˜˜ κim , (28a) { } { } written as (0) (0) (0) (0) ̺˜˙od = h˜ev , ̺˜ev + h˜od , ̺˜od ∆˜ , κ˜re , (28b) { } { }−{ } ν˙ = E,ν , (35) (0) (1) (0) (0) κ˜˙ re =2h˜ev κ˜im + h˜od , κ˜re ∆˜ , ̺˜od , (28c) { } { }−{ } (0) (0) where we have introduced the new functions 2h˜ev κ˜re = ∆(1˜ 2˜̺ev ) . (28d) − ev od ev ˜od (1) ν = ν +̺ ˜ , E = E + h . (36) Only one of the higher-order quantities, namelyκ ˜im , ap- pears in these equations, but it can be expressed in terms Eq. (35) resembles very much the usual Vlasov equa- of the leading-order quantities, e.g., with the help of Eq. tion for the normal Fermi gas, which can be written as (28a): ̺˙ = h,̺ . One just has to replace the distribution function{ ̺}by the quasiparticle distribution function ν (1) 1 (0) (0) (0) and the hamiltonian h by the quasiparticle energy E. κ˜im = ̺˜˙ev h˜ev , ̺˜od h˜od , ̺˜ev . (29) ˜ −{ }−{ } 2∆  It should be mentioned that Eq. (35) or similar kinetic equations have already been derived in the literature sev- By taking a linear combination of Eqs. (28a) and (28c) (1) eral times. Probably for the first time it was given by one can eliminateκ ˜im . The resulting equation reads Betbeder-Matibet and Nozi`eres [13] in a linearized form for small deviations from equilibrium. In order to be (0) (0) (0) (0) h˜ev ̺˜˙ev ∆˜ κ˜˙ re = Eev Eev , ̺˜od + h˜ev h˜od , ̺˜ev self-contained, we gave here our own way to arrive at Eq. − { } { } (0) (35). ∆˜ h˜od , κ˜re , (30) − { } In order to obtain a closed system of equations, Eq. where we have introduced the abbreviation (35) must be complemented by an equation for the so-far unknown phase φ. As stated above, the phase is fixed ˜ ˜2 ˜ 2 by the requirement that the gauge transformed gap ∆ is Eev = hev + ∆ . (31) ˜ q real, i.e., Im ∆ = 0. With the help of the relation (29) and of the gap equation (11), this can be rewritten as Eqs. (28b), (28d), and (30) form a system of three cou- (0) pled equations for the three leading-order quantities̺ ˜ev , d3p (0) (0) ̺˜˙ev h˜ev , ̺˜od h˜od , ̺˜ev =0 . (37) ̺˜od , andκ ˜re . Z (2π~)3 −{ }−{ }  5

As we will see in a moment, this is nothing but the etc. Our aim is to calculate the small deviations from continuity equation. This observation confirms earlier equilibrium induced by the perturbation V1ext , which will statements in the literature that the continuity equation be marked by an index “1”. To that end we linearize the should be used for the determination of the phase [13]. In equation of motion (35) for the quasiparticle distribution order to derive the continuity equation from Eq. (37), we function: write down explicitly the poisson brackets and integrate by parts. In this way we obtain ′ ν˙1 E0,ν1 = f (E0) E1, E0 , (47) d3p p ~∇φ −{ } { } ̺˜˙ + ∇ ̺˜ =0 . (38) ~ 3 − Z (2π )  · m  where f ′(E )= df/dE . We also linearize the continuity Using Eq. (16) and changing the integration variable ac- 0 0 equation (38): cording to p p + ~∇φ, this can be transformed into the usual continuity→ equation, ∇ ~ ρ˙(r)+ j(r)=0 , (39) ρ˙ (r)+ ∇ j (r) ∇ ρ (r)∇φ (r)=0 , (48) · 1 · 1ν − m · 0 1 with with d3p p j(r)= ̺(r, p) . (40) Z (2π~)3 m d3p p j (r)= ν1(r, p) . (49) 1ν Z (2π~)3 m

B. Linearization around equilibrium

From now on we will assume that the external potential In order to have a closed system of equations, we must Vext can be written as express E1(r, p) and ρ1(r) in terms of equilibrium quan- tities, the perturbation V1ext (r), and the unknown func- Vext = V0ext + V1ext , (41) tions ν1(r, p) and φ1(r, p). Linearizing E(r, p), one ob- tains where V0ext is time-independent and V1ext can be consid- ered as a small perturbation. The equilibrium quantities corresponding to the static potential V0ext will be marked h0 ∆0 E1 = h˜1ev + ∆˜ 1 + h˜1od , (50) by an index “0”, e.g., E0 E0

ν0(r, p)= f[E0(r, p)] , (42)

φ0(r)=0 , (43) with where f(E) denotes the Fermi function ˜ ~ ˙ 1 h1ev (r, p)= V1ext (r)+ gρ1(r) φ1(r) , (51) f(E)= (44) − E/(kB T ) ~ e +1 h˜ od (r, p)= p ∇φ (r) . (52) 1 −m · 1 and

2 2 E0(r, p)= h0(r, p) + ∆0(r) , (45) q The most difficult part is to derive the expressions for 2 p ρ (r) and ∆˜ (r). We start by linearizing Eqs. (32) and h (r, p)= + V ext (r)+ gρ (r) µ , (46) 1 1 0 2m 0 0 − (33):

h0 1 2f(E0) 2 ev ev ext ~ ˙ ˜ ̺˜1 = ν1 + − 3 [ ∆0(V1 + gρ1 φ1)+ h0∆0∆1] , (53) E0 2E0 − −

∆0 1 2f(E0) 2 1 2f(E0) re ev ext ~ ˙ ˜ ˜ κ˜1 = ν1 − 3 [h0∆0(V1 + gρ1 φ1) + ∆0∆1]+ − ∆1 . (54) − E0 − 2E0 − 2E0

According to Eqs. (10) and (11), ρ1 and ∆˜ 1 can be obtained by integrating Eqs. (53) and (54) over p. This gives a 6 coupled system of two linear equations,

ρ (r)= ρ (r) A(r)[V ext (r)+ gρ (r) ~φ˙ (r)] + B(r)∆˜ (r) , (55a) 1 1ν − 1 1 − 1 1 ∆˜ (r) = ∆ (r)+ gB(r)[V ext (r)+ gρ (r) ~φ˙ (r)] + [gA(r) + 1]∆˜ (r) , (55b) 1 1ν 1 1 − 1 1

where the gap equation (11) for the equilibrium case has One can show that ϕ = 0for T = 0 and ϕ = 1 for ∆0 = 0. been used in the derivation of the last term, and the In all other cases, the function ϕ must be evaluated nu- following abbreviations have been introduced: merically. From its definition one can see that ϕ depends 1.2 3 d p h0(r, p) r ev r p ρ1ν ( )= 3 ν1 ( , ) , (56) 1 Z (2π~) E0(r, p) 3 0.8 d p ∆0(r) ∆ (r)= g ν ev (r, p) , (57) 1ν 3 1 ϕ 0.6 Z (2π~) E0(r, p) d3p 1 2f[E (r, p)] 0.4 A(r) = ∆2(r) 0 , (58) 0 ~ 3 − 3 Z (2π ) 2E0 (r, p) 0.2 d3p 1 2f[E (r, p)] 0 B(r) = ∆ (r) h (r, p) 0 . (59) 0 ~ 3 0 − 3 0 0.2 0.4 0.6 0.8 1 1.2 Z (2π ) 2E0 (r, p) T / Tc Below we will show that the coefficient B is negligible compared with the coefficient A. In the limit B 0 FIG. 1: Temperature dependence of the function ϕ defined in the two equations (55a) and (55b) are decoupled and→ can Eq. (64). immediately be solved for ρ1 and ∆˜ 1:

ρ1ν (r) A(r)[V1ext (r) ~φ˙1(r)] ρ1(r)= − − , (60) 1+ gA(r) on r only through the dimensionless parameter T/Tc(r), where Tc(r)=0.57∆0(r; T = 0)/kB is the local critical ∆1ν (r) ∆˜ 1(r)= . (61) temperature [31]. For illustration, the numerical result gA(r) for ϕ as a function of this parameter is shown in Fig. 1. We will now calculate the coefficients A and B for the If one applies the same method to the coefficient B, case that both ∆0(r) and kB T are small compared with one obtains B = 0. This is because the integrand in Eq. the local Fermi energy [30] (59) is odd in ξ if one neglects the energy dependence of 2 the density of states. Already from this argument one pF (r) ǫ (r)= = µ V ext (r) gρ (r) . (62) can conclude that the coefficient B must be suppressed F 2m − 0 − 0 by at least one power of ∆0/ǫF or T/ǫF . Indeed, after a In this case, the relevant contributions to the integrals rather lengthy and delicate analysis one finds (58) and (59) come from momenta near the Fermi sur- face. As usual, the integrals over p can be simplified by ∆ (r) mp (r) 1 transforming them into integrals over the energy vari- B(r)= 0 F [2 + ϕ(r)] . (65) 2 2~3 able ξ = p /2m ǫF (r) and approximating the den- 2ǫF (r)  2π − g  sity of states by− its value at the Fermi energy, i.e., p2dp mp (r)dξ. For the coefficient A, one obtains ≈ F in this way This is the justification for neglecting the coefficient B when solving Eqs. (55a) and (55b). mp (r) A(r)= F [1 ϕ(r)] , (63) 2π2~3 − Finally, let us put everything together and give a con- cise summary of the system of equations which has to where the function ϕ describes the temperature depen- be solved. First of all, there is the equation of motion dence: (47) for the quasiparticle distribution function. After ξ2 the Poisson bracket on the r.h.s. has been written down ϕ(r)= dξ f ′(E) . (64) explicitly, it can be transformed into 2 2 2 − Z E E=√ξ +∆0(r)

7

′ ~ ˙ ~ ˙ f (E0) ∇V1ext + gρ1ν φ1 ∆0 ∇∆0(V1ext + gρ1ν φ1) h0 ∇∆0∆1ν ν˙1 E0,ν1 = p − + 2 p − + 2 p −{ } − m h − · 1+ gA E0 · 1+ gA E0 · gA ~ h0 2 h0 ∆0 + (p ∇) φ1 ~ ∇(V0ext + gρ0)+ ∇∆0 ∇φ1 . (66) m E0 · − E0 E0  · i The second equation is the continuity equation

ρ˙ (r) A(r)[V˙ ext (r) ~φ¨ (r)] ~ 1ν − 1 − 1 + ∇ j (r) ∇ ρ (r)∇φ (r)=0 . (67) 1+ gA(r) · 1ν − m · 0 1

The definitions of ρ1ν , ∆1ν , and j1ν in terms of ν1 are and linearizing Eqs. (69) and (70) around equilibrium, given by Eqs. (56), (57), and (49). one obtains ~ ρ˙ (r) ∇ ρ (r)∇φ (r)=0 , (74) 1 − m · 0 1 C. Limiting cases 2π2~3 ~φ˙(r)= V ext (r)+ + g ρ (r) . (75) 1 mp (r) 1 We are now going to check that our equations repro-  F  duce superfluid hydrodynamics and the Vlasov equation Solving Eq. (75) for ρ1 and inserting the result into Eq. in the cases T = 0 and T Tc, respectively. In the limit (74), one reproduces exactly Eq. (68). This can be seen ≥ of zero temperature, Eq. (66) becomes extremely simple as an alternative to the recent derivation of superfluid since f(E) = 0 and therefore the r.h.s. of Eq. (66) van- hydrodynamics from the underlying microscopic theory ishes identically. The corresponding solution is of course in Ref. [22]. ν1 = 0 [32], which implies ρ1ν = ∆1ν = j =0. As a The analysis of the other limit, T T , is more diffi- 1ν ≥ c consequence, the continuity equation (67) reduces to cult. In this limit, the gap ∆0 vanishes and consequently

E0(r, p)= h0(r, p) , (76) V˙1ext (r) ~φ¨1(r) ~ − + ∇ ρ (r)∇φ (r)=0 . (68) | | 2π2~3 m · 0 1 ν1ev (r, p) = sgn[p pF (r)]˜̺1ev (r, p) . (77) + g − mpF (r) In addition, one has ϕ(r) = 1, A(r) = 0, ρ1(r)= ρ1ν (r), and ∆˜ 1(r) = ∆1ν (r) = 0. Using these relations, and Here we have used the explicit expression for A(r) and considering separately the two cases ppF (i.e., h0 > 0), one can convince oneself that How does Eq. (68) compare to superfluid hydrodynam- Eqs. (66) and (67) reduce to ics? The continuity and Euler equations of superfluid ′ hydrodynamics can be written as [5]: f (h0) ̺˜˙1 h0, ̺˜1 = p ∇(V1ext + gρ1 ~φ˙1) −{ } m  − · − ρ˙(r)+ ∇ ρ(r)v(r)=0 , (69) ~ 2 · + (p ∇) φ ~[∇(V ext + gρ )] ∇φ . (78) v2(r) Vext (r) µloc(r) m 1 0 0 1 v˙ (r)= ∇ + + , (70) · − ·  −  2 m m  and where v(r) denotes the velocity field and µloc(r) is the ~ ρ˙ (r)+ ∇ j (r) ∇ ρ (r)∇φ (r)=0 . (79) local chemical potential, which in the BCS phase (∆ 1 · 1ν − m · 0 1 ≪ ǫF ) is related to the density ρ(r) by the Thomas-Fermi As we will see in a moment, these two equations are not relation independent of each other. Hence, they do not allow 2 to determine̺ ˜1(r, p) and φ1(r) in a unique way. This pF (r) µloc(r)= + gρ(r) , (71) ˜ 2m is in fact very reasonable since the condition Im ∆=0 fixing the phase φ becomes meaningless above Tc, where with ∆˜ = 0, and therefore the function φ should be completely arbitrary in this case. The relevant physical quantity, ~ 2 1/3 pF (r)= [6π ρ(r)] . (72) which of course should be unique, is ̺1(r, p). Linearizing Eq. (16) and using ̺0(r, p)= f[h0(r, p)], we can express Writing the irrotational velocity field in the form ̺˜1(r, p) in terms of ̺1(r, p) as follows: ~ ~ v(r)= ∇φ(r) (73) ̺˜ (r, p)= ̺ (r, p) f ′[h (r, p)]p ∇φ . (80) −m 1 1 − m 0 · 1 8

If we insert this into Eq. (78), all terms containing the a uniform medium. This case has already been studied phase φ1 drop out, and we are left with by Leggett [12] many years ago (except for the numerical evaluation of the integrals) by using the standard tech- ′ p ̺˙1 h0,̺1 = f (h0) ∇(V1ext + gρ1) . (81) niques of normal and anomalous Green’s functions. The −{ } m · purpose of the present subsection is therefore to check This is nothing but the linearized form of the Vlasov that our apparently very complicated equations (66) and equation, (67) correctly interpolate between the limits of zero and critical temperature. ̺˙1 h0,̺1 = h1,̺0 , (82) Since the medium is assumed to be uniform, the equi- −{ } { } librium quantities do not depend on r. We consider an with h1 = V1ext + gρ1. It remains to check that the con- excitation operator of the form tinuity equation (79) is satisfied for arbitrary functions φ1, if ̺1 fulfills Eq. (81). To that end, we multiply Eq. ik·r−iωt ext r ˆ ext (81) by p and integrate over p, which leads to the usual V1 ( ; t)= V1 e . (85) continuity equation As usual, in order to ensure that the perturbation van- ρ˙ (r)+ ∇ j (r)=0 . (83) ishes for t , one can assume that ω has an in- 1 · 1 finitesimal positive→ −∞ imaginary part. From translational With the help of Eq. (80) the current j1 can be written invariance it is clear that all quantities describing the as deviations from equilibrium will also have the form of a ~ plane wave, with the same wave vector k and frequency j (r)= j (r) ρ (r)∇φ (r) . (84) ω as the excitation. Like Vˆ ext , the amplitudes will be 1 1ν m 0 1 1 − marked by a hat over the corresponding symbol. Con- Inserting this into Eq. (83), we indeed recover Eq. (79). cerning the phase φ1, it turns out to be convenient to Since we did not make any assumptions about the func- parametrize it in the form tion φ1(r), we conclude that it is completely arbitrary, as it should be. ˆ i k·r− φ (r; t)= φ˙ ei iωt . (86) 1 1 ω The Poisson bracket on the l.h.s. of Eq. (66) now becomes III. SIMPLE EXAMPLES

A. Sound wave in a uniform system h0 p k ik·r−iωt E0,ν1 = i · νˆ1e , (87) { } − E0 m In this subsection we are considering a particularly simple excitation, namely a sound wave traveling through and Eq. (66) can easily be solved forν ˆ1:

ˆ ˆ ext ~ ˙ ˆ ′ p k h0 h0 V1 φ1 + gρˆ1ν ∆0 ∆1ν p k ˆ f (E0) · − + · ~φ˙1 − mω E0 E0 1+ gA − E0 gA mω νˆ1 =   . (88) h0 p k 1 · − E0 mω

Of course, the quantitiesρ ˆ1ν and ∆ˆ 1ν on the r.h.s. de- grals which lead to the coupling between the equations pend themselves onν ˆ1. Therefore the next step consists forρ ˆ1ν and ∆ˆ 1ν are zero within this approximation, i.e., in inserting this expression forν ˆ1 into Eqs. (56) and (57). they are of higher order in ∆/ǫF or T/ǫF and can be The integrals over the angle between p and k can be eval- neglected. The resulting equation forρ ˆ1ν reads uated in closed form. For the remaining integrals over p, we will again exploit the fact that the gap and the tem- perature are much smaller than the Fermi energy, as we 2 did already in Sec. II B. We thus replace p dp by mpF dξ, and in the integrand we replace p by p , except for h F 0 ˆ ~ ˆ˙ 2 2 mpF V1ext φ1 + gρˆ1ν ˆ and E0, which must be replaced by ξ and ξ + ∆0, ρˆ = I (s)+ ~φ˙ I (s) . 1ν 2~3 − 2 1 0 respectively. Like the coefficient B in Sec. II B,p the inte- −2π  1+ gA  (89) 9

ˆ Here we have introduced the abbreviation We will now express ˆj1ν in terms of Vˆ1ext , φ˙1, andρ ˆ1ν n by inserting Eq. (88) into Eq. (49). The integration over ′ ξ sE ξ In(s)= dξf (E) 1 arctanh , p is done as explained above for the case ofρ ˆ1ν , and the − Z  E  h − ξ sE i result reads (90) where E = ξ2 + ∆2, and s denotes the dimensionless 0 ˆ ~ ˆ˙ ratio of the sound velocity c = ω/k and the Fermi velocity mcpF V1ext φ + gρˆ1ν p ˆj = − I (s) 1ν 2~3 0 vF = pF /m, −2π  1+ gA ~ ˆ˙ c mω ˆ ρn φ1 s = = . (91) + ~φ˙1I−2(s) . (93) vF pF k  − mc

Although not marked explicitly, In(s) depends not only In the last term, we have introduced the “normal density” on s but also on the ratio T/Tc (analogously to the func- of the system, ρn, which is given by tion ϕ). Note that the integrals In(s) have a branch cut along the real axis from s = 1 to s = 1. The infinites- ′ − ρn = ρ0 ρs = ρ0 dξf (E0) . (94) imal imaginary part of ω, i.e., of s, fixes the sign of the − − Z imaginary part of In(s). Until now we have one equation for two unknown quan- Correspondingly, ρs is the “superfluid density”. Note ˆ that the ratios ρ /ρ and ρ /ρ depend only on one pa- tities,ρ ˆ1ν and φ˙1. The second equation can be obtained n 0 s 0 from the continuity equation (67). It is evident that the rameter, namely T/Tc. The numerical results for ρn/ρ0 and ρs/ρ0 as functions of T/Tc are shown in Fig. 2. current j1ν flows in longitudinal direction, such that it can be written in the form Inserting Eq. (93) into the continuity equation (67), one obtains the second equation which is needed for de- k j (r; t)= ˆj eik·r−iωt . (92) terminingρ ˆ and φˆ˙ : 1ν 1ν k 1ν 1

ˆ ~ ˆ˙ ˆ ~ ˆ˙ ~ ˆ˙ ρˆ1ν A(V1ext φ1) mpF V1ext φ + gρˆ1ν ˆ ρs φ1 − − + − I (s)+ ~φ˙ I (s) =0 . (95) 2~3 0 1 −2 2 1+ gA 2π  1+ gA  − mc ˆ In principle we could now solve Eqs. (89) and (95) for the two unknown variablesρ ˆ1ν and φ˙1. However, it is more transparent to use the amplitude of the total density oscillations, ρˆ1, as variable instead of the auxiliary quantity ρ1ν . Expressingρ ˆ1ν in terms ofρ ˆ1 with the help of Eq. (60), we rewrite Eqs. (89) and (95) as

gmpF mpF ˆ mpF 1+ [1 ϕ + I (s)] ρˆ [1 ϕ + I (s) I (s)]~φ˙ = [1 ϕ + I (s)]Vˆ ext , (96a) 2~3 2 1 2~3 2 0 1 2~3 2 1  2π −  − 2π − − −2π − gmpF mpF 1 ρs ˆ mpF 1+ I (s) ρˆ + I (s) I (s) ~φ˙ = I (s)Vˆ ext . (96b) 2~3 0 1 2~3 −2 0 2 1 2~3 0 1  2π  2π  − − 3s ρ0  −2π

It is straight-forward to solve this 2 2 system of equa- expressed in the form × tions forρ ˆ1. Let us introduce the response function Π, defined such that Π0(s; T/Tc) Π(s; T/Tc; kF a)= , (99) 2kF a mp 1 Π (s; T/T ) ρˆ = F Π(s; T/T ; k a)Vˆ . (97) − π 0 c 1 2π2~3 c F 1 where Π0 is the response function in the limit kF a 0: The first term has been factored out in order to make → Π dimensionless. From the system of equations (96) one 1 ρs 2 (1 ϕ + I2) 2 I−2 + I0 can see that Π is a function of s and the two parameters − 3s ρ0 − Π0 =   . (100) T/Tc and 1 ρs 1 ϕ + I + I− 2I − 3s2 ρ − 2 2 − 0 gmp 0 k a = F . (98) F 4π~3 Note that these expressions coincide exactly with the quantum mechanical result in the long-wavelength and The explicit expression for Π can most conveniently been low-frequency limit as given by Eqs. (68) and (69) of Ref. 10

1.2 in the integrand vanish, i.e., ϕ = ρn/ρ0 = In(s) = 0, and the response function reduces to 1

ρ 1 / ρ / ρ ρ / ρ s 0.8 s n Π(s; 0; kF a)= . (101)

ρ 2 3s 1 2kF a/π 0.6 − − ,

ρ This means that the excitation spectrum is a δ function / 0.4 n at ρ 0.2 1 2k a s = + F , (102) 0 r3 3π 0 0.2 0.4 0.6 0.8 1 1.2 T / T corresponding to the hydrodynamic speed of sound. c In the other limit, T T , one has ϕ = ρ /ρ = 1. → c n 0 FIG. 2: Temperature dependence of ρn/ρ0 (solid line) and The integrals In(s) reduce to ρs/ρ0 (dashed line). 1 I (s; T/T 1)=1 s arctanh , (103) n c ≥ − s 6 independent of n, since the factors ξ/E in the integrand − kF a = 0.25 T/Tc = 0.8 of Eq. (90) can be replaced by 1. As a consequence, the 5 T/Tc = 0.9 T/T = 0.99 two equations of the system (96) become identical and 4 c

(s) T/T = 1 ˆ c the coefficients in front of φ˙1 vanish, in accordance with Π 3 the more general arguments of Sec. IIC. Solving forρ ˆ1 Im

− 2 gives 1 1 1 s arctanh 0 − − s Π(s; T/Tc 1; kF a)=   , 0 0.2 0.4 0.6 0.8 1 1.2 ≥ 2kF a 1 s = c / v 1+ 1 s arctanh F π  − s (104) FIG. 3: Excitation spectrum − Im Π of a uniform medium as in agreement with the usual result of Landau’s Fermi- function of the reduced sound velocity s = c/vF for different liquid theory for the case of a pure s-wave interaction. temperatures. If the interaction was repulsive (a> 0), Eq. (104) would have a pole at s > 1, corresponding to the propagation of zero sound. However, here we are considering the case [12]. In order to see this, it is sufficient to observe that of an attractive interaction, where zero sound does not after integration over the solid angle the quantities α, ζ, exist. Instead there is a continuous spectrum of particle- and η defined in Eq. (65) of Ref. [12] can be expressed hole excitations ranging from s =0 to s = 1. in terms of our integrals as α = (1 ϕ + I2 I0)/2, Our numerical results shown in Fig. 3 can be inter- 2 − − ζ = [ρs/(3ρ0)+ s (I0 I−2)]/2, and η = I0. In our preted as follows. At zero temperature, there exists a − − case of a pure s-wave interaction, the Landau parame- collective hydrodynamic sound, which is undamped (at ters in Eq. (68) of Ref. [12] are given by F0 = 2kF a/π least within the present theoretical treatment). As the and F1 = 0. Then the quantities K1 and Q of Ref. [12] temperature increases, a normal component consisting correspond to our Π and Π0, respectively. As stated in of thermally excited quasiparticles builds up. However, Ref. [12], the long-wavelength and low-frequency limit is at temperatures where ρn is already considerably differ- valid if ~ω, vF ~k ∆. Since our semiclassical result co- ent from zero, the hydrodynamic sound is still practi- ≪ incides with this limit of the quantum mechanical result, cally undamped. The reason for this is that all thermally we conclude that this is the condition for the validity of excited quasiparticles contribute equally to ρn, whereas our semiclassical theory. A calculation of the response only those quasiparticles whose velocity dE/dp vF ξ/E function beyond the long-wavelength and low-frequency is at least equal to the sound velocity c contribute≈ to limit can be found in Ref. [23]. the Landau damping. At sufficiently high temperature, The excitation spectrum of the system is characterized the Landau damping becomes very strong and the hy- by the imaginary part of Π, which is plotted in Fig. 3 for drodynamic sound ceases to exist. What remains is a k a = 0.25 and several temperatures between 0.8 T continuum of particle-hole excitations, and the interac- F − c and Tc. One can see that at 0.8 Tc the excitation spec- tion manifests itself only in the rounded edge near s = 1. trum exhibits a peak near s 0.5 which becomes broader and finally disappears when≈ the temperature approaches Tc. B. Quadrupole mode in a spherical trap In order to interprete this behavior, let us again con- sider the two limits T 0 and T Tc. In the zero- Our main motivation for developing the present semi- → → ′ temperature case, all integrals containing the term f (E0) classical approach was to apply it to the case of trapped 11 atomic Fermi gases. The simplest example which comes the time dependence of ν1 as well as the explicit time to our mind is the quadrupole oscillation of a Fermi gas dependence of F will be harmonic, too. We will denote in a spherical trap. Even in this case, the r dependence of the amplitudes by a hat over the corresponding symbols. iωt the equilibrium quantities induced by the trap potential Multiplying Eq. (111) by e , one finds thatν ˆ1 is given makes our equations very complicated. Since in this first by the Fourier integral investigation we are interested in problems which can be ∞ solved analytically, we will apply two additional simpli- iωτ νˆ1(r, p)= dτ e Fˆ[R(r, p; τ), P(r, p; τ)] . fying approximations, which allow us to obtain explicit Z0 − − − solutions. A numerical method for solving our equations (113) without additional approximations will be proposed at For the purpose of illustration we want to discuss a the end of this subsection. simple case in which the classical trajectories are ana- Let us start with the linearized equation (66), which lytically known. We make two approximations: First, has the form we replace the r-dependent gap ∆0(r) by a constant ∆0. This approximation implies that ν1ev is odd in ξ and ν˙1(r, p; t) E0(r, p),ν1(r, p; t) = F (r, p; t) . (105) ∆˜ 1ν can be neglected, as it was the case in the preced- −{ } ing subsection. Second, we will neglect effects from the For its solution we adopt the Green function method used Hartree mean-field as well in the equilibrium state as in in Ref. [24] to solve the linearized Vlasov equation for the deviations from equilibrium. The second approxima- nuclear giant resonances, which is formally very similar to tion, which is by far not as unrealistic as the first one, our problem. The starting point is to write the solution amounts to neglecting all gρ0 and gρ1ν terms and replac- of Eq. (105) in the form ing the denominators 1 + gA in Eq. (66) by 1. The trap potential is assumed to be a spherical harmonic oscilla- ′ 3 ′ 3 ′ ′ ′ ′ ν1(r, p; t)= dt d r d p G(r, p, r , p ; t t ) tor, Z Z Z − ′ ′ ′ 1 2 2 F (r , p ; t ) , (106) V ext = mΩ r . (114) × 0 2 where G is the Green function of the differential operator It is evident that the equations of motion (108) conserve on the l.h.s. of Eq. (105), satisfying E0. However, if ∆0 is a constant, this implies that h0 is ∂ ∂E ∂ ∂E ∂ conserved, too, and the solutions of Eqs. (108) are closely 0 0 G(r, p, r′, p′; t) related to the those of the ordinary harmonic oscillator. ∂t − ∂r ∂p − ∂p ∂r h i=Xxyz  i i i i i Indeed, it is straight-forward to show that the trajectories = δ(t)δ(r r′)δ(p p′) . (107) are given by − − h Ωt p h Ωt Denoting by R(r, p; t) and P(r, p; t) the solutions of the R(r, p; t)= r cos 0 + sin 0 , (115a) classical equations of motion E0 mΩ E0 h0Ωt h0Ωt ∂E0(R, P) ∂E0(R, P) P(r, p; t)= p cos mΩr sin . (115b) R˙ i = , P˙i = (108) E0 − E0 ∂Pi − ∂Ri Since h0 and E0 are constants of the motion, they can satisfying the initial conditions R(r, p;0) = r and likewise be evaluated at (r, p) or (R, P). P(r, p;0) = p, one can show that Due to our approximations, the function F [given by the r.h.s. of Eq. (66)] reduces to G(r, p, r′, p′; t)= θ(t)δ[r R(r′, p′; t)]δ[p P(r′, p′; t)] . − − (109) 2 ′ h0 p ˆ ∇ ˆ ext ~ ˆ fulfills Eq. (107). Due to time-reversal symmetry and F = f (E0) 2 (V1 + i ωφ1) Liouville’s theorem, this Green function can be rewritten − h − E0 m · 2 as h0 p 2 h0 + ∇ ~φˆ1 Ω r ∇~φˆ1 . (116) ′ ′ ′ ′ E0 m ·  − E0 · i G(r, p, r , p ; t)= θ(t)δ[r R(r, p; t)]δ[p +P(r, p; t)] . − − − (110) As excitation we choose the quadrupole operator The latter form renders the phase-space integrals in Eq. 2 Vˆ ext (r)= αmΩ (r r) , (117) (106) trivial. Changing the time integration variable ac- 1 ⊗ 20 cording to τ = t t′, one obtains − where, explicitly, ∞ 2v w v w v w ν1(r, p; t)= dτF [R(r, p; τ), P(r, p; τ); t τ] . z z x x y y (v w)20 = (1µ1ν 20)vµwν = − − . Z0 − − − − ⊗ | √6 (111) Xµν In the case of a harmonic perturbation, (118) The prefactor in Eq. (117) has been chosen such that −iωt V1ext (r; t)= Vˆ1ext (r)e , (112) the coefficient α is dimensionless. Due to the spherical 12 symmetry of the trap, the angular dependence of φˆ1 must choice of the coefficient β. Quadratic ans¨atze like Eq. be of the same quadrupolar form as that of Vˆ1ext , but the (119), corresponding to a superfluid velocity field which radial dependence could in principle be different. Here is linear in the coordinates, have frequently been used we make the ansatz that φˆ1 is proportional to Vˆ1ext , i.e., (see, e.g., Ref. [5]) for the calculation of the frequencies of collective modes in the limit of superfluid hydrodynamics (T = 0). mΩ φˆ (r)= β (r r) , (119) 1 ~ ⊗ 20 Inserting Eqs. (117) and (119) into Eq. (116) and using and we will show afterwards that with this ansatz for φˆ the explicit form of the trajectories, Eq. (115), we can the continuity equation can be satisfied by an appropriate evaluate the Fourier integral in Eq. (113), with the result

2 2 2 ω ∆0 h0 ∆0 ω iβ 2 α 2 2 4β 2 + iα ′  (p p)20 2 h0 Ω E0 − E0 h0  E0 Ω νˆ1 = 2f (E0) ⊗ mΩ (r r)20 2 2 + Ω(r p)20 2 β 2 2 , (120) − m − ⊗ E0 ω h ⊗ E − ω h   4 0 0  4 0   2 2  2 2   Ω − E0  Ω − E0 

Now we have to calculate the corresponding current ˆj and to satisfy the continuity equation, and the corresponding density oscillationsρ ˆ1ν . As detailed in the preceding sub- solution for the coefficient β reads section, this is accomplished by integrating pνˆ1/m and 2 1 ϕ +2I (z) νˆ1, respectively, over p. Replacing in the integrals p dp β = izα − 22 . (126) 2 2 2 by mp (r)dξ, p by p (r), h by ξ, and E by ξ2 + ∆ , (1 ϕ)(z 2)+2I22(z)(z 4) F F 0 0 0 − − − we obtain p This expression can be used to obtainρ ˆ1ν . Here we will ˆj (r)= ρ0(r)Ω β[ϕ 4I22(z)] iαzI20(z) ∇(r r)20 , immediately give the result for the amplitude of the total 1ν − − ⊗  (121) density oscillations, i.e.,ρ ˆ =ρ ˆ A(Vˆ ext + i~ωφˆ ), 1 1ν − 1 1 m2Ω2p (r) which we write in the form ρˆ (r)= F [αI (z) iβzI (z)](r r) , 1ν 2~3 40 22 20 2 2 π − ⊗ m Ω pF (r) (122) ρˆ1(r)= α (r r)20Π(z) (127) π2~3 ⊗ with the abbreviations with ω z = , (123) Ω [1 ϕ +2I22(z)][1 ϕ +4I22(z)] Π(z)= I (z)+ . i j 20 − 2 − 2 ′ ξ ∆0 1 (1 ϕ)(z 2)+2I22(z)(z 4) Iij (z)= dξf (E) , (124) − − − − Z Ei+j z2 4ξ2/E2 (128) − Before discussing numerical results, let us again study where E = ξ2 + ∆2. From its definition it is evident the two extreme cases T = 0 and T T . In the zero- 0 ≥ c that I40 = I20p I22, such that it is sufficient to evaluate temperature limit, all integrals ϕ and I are zero, and − ij two of these integrals numerically. The functions Iij (z) hence the response function becomes have a branch cut along the real axis from z = 2 to − 1 z = 2. Remember that ω, and therefore also z, is assumed Π(z; T/T =0)= , (129) c z2 2 to have an infinitesimal positive imaginary part, fixing − the sign of the imaginary part of Iij (z). i.e., it has a single pole at the hydrodynamical frequency As stated above, the coefficient β must be determined by the continuity equation (67). Due to our approx- imation to neglect the Hartree field, the denominator ω = √2Ω . (130) 1+ gA(r) in the first term of Eq. (67) can be replaced by 1, and the Fermi can be given in closed form: In the case T T , i.e., in the normal phase, we have ≥ c ϕ = 1, and in the definition (124) we can replace ∆0 and E by 0 and ξ, respectively, such that we obtain 1 0 p (r)= 2m µ mΩ2r2 . (125) F r − 2 1   I20(z; T/Tc 1) = , (131a) ˆ ≥ z2 4 Inserting the results for j1ν andρ ˆ1ν into the continuity − I (z; T/T 1)=0 . (131b) equation, one finds that the ansatz (119) indeed allows 22 c ≥ 13

frequency in the normal phase is exactly equal to z = 2 is 12 T/Tc = 0.5 T/T = 0.7 a consequence of neglecting the Hartree mean field. How- 10 c ever, in the range of validity of our theory (k a 1), T/Tc = 0.8 F T/T = 0.9 the Hartree mean field cannot shift the frequency| | ≪ very (z) 8 c

Π T/Tc = 0.95 much (in Ref. [7], e.g., the frequency is shifted from 2 to 6 T/Tc = 0.99 Im 2.2). We therefore believe that this effect is not very − 4 important.≈ More problematic is the constant-gap approx- 2 imation which we needed for the analytical solution of the 0 equations of motion (108). Because of this approxima- 0 0.5 1 1.5 2 tion, there are no quasiparticles having energies below z = ω / Ω ∆0, and as a consequence, the Landau damping sets in at rather high temperatures. In the full calculation, how- FIG. 4: Quadrupole excitation spectrum − ImΠ of a harmon- ever, the lowest-lying quasiparticles are those whose wave ically trapped gas as a function of the excitation frequency (in functions are localized near the surface, where the gap is units of the trap frequency) for different temperatures. The small, and which have much smaller energies than the spatial dependence of the gap as well as the Hartree mean- central value of the gap. Therefore within the full cal- field have been neglected. culation the Landau damping is already quite important at very low temperatures. In the semiclassical formal- ism these low-lying quasiparticles can be understood as Thus the response function reduces to quasiparticles bouncing back and forth between the po- 1 tential wells created by the trap potential and the spa- Π(z; T/T 1) = . (132) c ≥ z2 4 tially varying gap (Andreev reflection) [20, 25]. The in- − clusion of this effect would require a numerical solution Like in the zero-temperature case, we have a single pole, of the equations of motion (108). √ but now at a frequency which is higher by a factor of 2. This leads us to a possible numerical method for solv- The reason for the difference of the two frequencies is as ing even the original (i.e., not linearized) kinetic equation follows. In the superfluid phase, the momentum distri- (35). In nuclear physics, the Vlasov equation (usually bution stays spherical during the oscillation. In contrast complemented by a collision term) is routinely solved by to this, in the normal phase, the momentum distribution the so-called test-particle method, e.g., in order to sim- is deformed in the opposite direction as the density in ulate heavy- collisions [26]. Recently this method has coordinate space. This deformation of the Fermi sphere also been applied to the solution of the Vlasov equation costs kinetic energy, which increases the restoring force with collision term for trapped atomic Fermi gases [27] and thereby the frequency of the oscillation. and of a Vlasov-like equation for trapped fermion-boson At intermediate temperatures 0

[1] M. Bartenstein et al., Phys. Rev. Lett. 92, 203201 (2004). 89, 250402 (2002). [2] K.M. O’Hara, et al., Science 298, 2179 (2002). [5] M. Cozzini and S. Stringari, Phys. Rev. Lett. 91, 070401 [3] J. Kinast, S.L. Hemmer, M.E. Gehm, A. Turlapov, and (2003). J.E. Thomas, Phys. Rev. Lett. 92, 150402 (2004); J. Ki- [6] S. Stringari, Europhys. Lett. 65, 749 (2004). nast, A. Turlapov, and J.E. Thomas, Phys. Rev. A 70, [7] M. Grasso, E. Khan, and M. Urban, Phys. Rev. A 72, 051401(R) (2004). 043617 (2005). [4] C. Menotti, P. Pedri, and S. Stringari, Phys. Rev. Lett. [8] E. Taylor and A. Griffin, preprint cond-mat/0509040. 15

[9] P. Massignan, G.M. Bruun, and H. Smith, Phys. Rev. A 17, 49 (2001). 71, 033607 (2005). [24] H. Kohl, P. Schuck, and S. Stringari, Nucl. Phys. A 459, [10] J.R. Schrieffer, Theory of superconductivity (Benjamin, 265 (1986). New York, 1964). [25] G.M. Bruun and H. Heiselberg, Phys. Rev. A 65, 053407 [11] A. J. Leggett, Phys. Rev. 140, A1869 (1965). (2002). [12] A. J. Leggett, Phys. Rev. 147, 119 (1966). [26] G.F. Bertsch and S. Das Gupta, Phys. Rep. 160, 190 [13] O. Betbeder-Matibet and P. Nozi`eres, Ann. Phys. (N.Y.) (1988). 51, 392 (1969). [27] F. Toschi, P. Vignolo, S. Succi, and M.P. Tosi, Phys. Rev. [14] J.W. Serene and D. Rainer, Phys. Rep. 101, 221 (1983). A 67, 041605(R) (2003); F. Toschi, P. Capuzzi, S. Succi, [15] D. Vollhardt and P. W¨olfle, The Superfluid Phases of P. Vignolo, and M.P. Tosi, J. Phys. B 37, S91 (2004). Helium 3 (Taylor & Francis, London, 1990) [28] T. Maruyama, H. Yabu, T. Suzuki, preprint [16] M. Urban and P. Schuck, Phys. Rev. A 67, 033611 (2003); cond-mat/0505357 (2005). M. Urban, Phys. Rev. A 71, 033611 (2005). [29] B. Jackson, P. Pedri, and S. Stringari, Europhys. Lett. [17] N. Nygaard, Ph.D. thesis, University of Maryland, 2004. 67, 524 (2004). [18] P. Ring and P. Schuck, The Nuclear Many-Body Problem [30] Note that in the case of a trapped system, the condition (Springer-Verlag, Berlin, 1980). ∆0(r) ≪ ǫF (r) is automatically fulfilled everywhere in [19] G.M. Bruun and B.R. Mottelson, Phys. Rev. Lett. 87, the trap if it is valid at the center. 270403 (2001). [31] It is evident that ϕ is a function of ∆0(r)/(kB T ), but [20] Y. Ohashi and A. Griffin, preprint cond-mat/0503641 ∆0(r) can in turn be written as kB Tc(r) times a universal (2005). function of T/Tc(r). [21] G. Bruun, Y. Castin, R. Dum, and K. Burnett, Eur. [32] This trivial solution is unique if we assume that the sys- Phys. J. D 7, 433 (1999). tem was in equilibrium at the moment when the time- [22] G. Tonini and Y. Castin, preprint cond-mat/0504612 dependent perturbation was switched on. (2005). [23] A. Minguzzi, G. Ferrari, and Y. Castin, Eur. Phys. J. D