Analytical limits for cold-atom Bose gases with tunable interactions

11
LA-UR-11-02248 Analytical limits for cold atom Bose gases with tunable interactions Bogdan Mihaila, 1 Fred Cooper, 1, 2 John F. Dawson, 3 Chih-Chun Chien, 1 and Eddy Timmermans 1 1 Los Alamos National Laboratory, Los Alamos, NM 87545 2 Santa Fe Institute, Santa Fe, NM 87501 3 Department of Physics, University of New Hampshire, Durham, NH 03824 (Dated: December 21, 2013, 11:52 P.M.) We discuss the equilibrium properties of dilute Bose gases using a non-perturbative formalism based on auxiliary fields related to the normal and anomalous densities. We show analytically that for a dilute Bose gas of weakly-interacting particles at zero temperature, the leading-order auxiliary field (LOAF) approximation leads to well-known analytical results. Close to the critical point the LOAF predictions are the same as those obtained using an effective field theory in the large-N approximation. We also report analytical approximations for the LOAF results in the unitarity limit, which compare favorably with our numerical results. LOAF predicts that the equation of state for the Bose gas in the unitarity limit is E/(pV ) = 1, unlike the case of the Fermi gas when E/(pV )=3/2. PACS numbers: 03.75.Hh, 05.30.Jp, 67.85.Bc I. INTRODUCTION Recently we introduced a new theoretical framework for the study of a dilute gas of Bose particles with tunable interactions[1] based on a loop expansion of the one-particle irreducible (1-PI) effective action in terms of composite-field propagators. The auxiliary field (AF) formalism makes use of the Hubbard-Stratonovitch transformation[2] to rewrite the Lagrangian in terms of auxiliary fields related to the normal and anomalous densities. Employing general quantum field theoretical methods[37], the AF formalism is part of a continuing effort in the community to apply methods traditionally used in high-energy physics[818] to the study of ultra- cold atomic gases[19]. For an interacting dilute Bose gas, the leading- order auxiliary field (LOAF) approximation is a non- perturbative, conserving and gapless approximation that describes a large interval of values of the coupling constant, satisfies Goldstone’s theorem and yields a second-order phase transition to a Bose-Einstein conden- sate (BEC) regime. In contrast with other resumma- tion schemes, such as the large-N expansion[20, 21] or the functional renormalization techiques[1618], here we treat the normal and anomalous densities on equal foot- ing. LOAF produces the same slope of the linear depar- ture of the critical temperature from the noninteracting limit derived by Baym et al.[21] using a large-N expan- sion for the critical theory. Unlike the large-N expansions developed by Baym et al., the LOAF approximation can be used at all temperatures. Furthermore, one can sys- tematically improve upon the LOAF approximation by calculating the 1-PI action order-by-order corrections. The broken U (1) symmetry Ward identities guarantee the preservation Goldstone’s theorem order-by-order[5]. The detailed derivation of the LOAF approximation was discussed recently for the case of dilute Bose[22] and Fermi[23] atomic gases. Unlike the case of Fermi gases where the LOAF approximation is equivalent to the stan- dard Bardeen-Cooper-Schrieffer (BCS) ansatz [8, 9], in the case of Bose gases the LOAF approximations leads to yet unexplored possibilities. Therefore it is important to study analytically the LOAF predictions for Bose gases in limiting cases such as the case of weakly-interacting sys- tems and also in the unitarity limit, which corresponds to the strongly-interacting regime where the s-wave scat- tering length, a 0 , is much larger than the inter-particle distance. In the unitarity limit the properties of the sys- tem have a universal character[24]. The intrinsic non- perturbative character of the LOAF approximation may become particularly relevant because the development of novel cold atom technology that produce stable, flat potentials bound by a sharp edge[25, 26] leads to the prospect of studying finite temperature properties of di- lute gases, such as the BEC transition temperature, T c , superfluid to normal fluid ratio, depletion, and specific heat, at fixed particle density ρ. In this paper we focus on the study of the LOAF pre- dictions in the broken-symmetry phase. We make contact with existing analytical approximations in the weakly- interacting limit, such as those discussed in the textbook of Fetter and Walecka[27] and the analytical results ob- tained close to the critical temperature by Kita[2830] using the related Luttinger-Ward functional. We will also show that the analytical techniques developed here can be used to study the LOAF predictions in the unitarity limit. This paper is organized as follows: In Sec. II, we briefly review the derivation of the LOAF equations. The LOAF effective potential and the derivation of thermodynamic properties are outlined in Sec. III. In Sec. IV we special- ize to the study of the interacting Bose properties in the broken-symmetry phase. In Sec. V we study the zero- temperature LOAF results in the weakly-coupling limit and we compare with the weakly interacting Bose gas theory discussed by Fetter and Walecka[27]. The zero- temperature analysis suggests the scaling of the LOAF equations discussed in Sec. VI. In Sec. VII we discuss the arXiv:1107.3750v1 [cond-mat.quant-gas] 19 Jul 2011

Transcript of Analytical limits for cold-atom Bose gases with tunable interactions

LA-UR-11-02248

Analytical limits for cold atom Bose gases with tunable interactions

Bogdan Mihaila,1 Fred Cooper,1, 2 John F. Dawson,3 Chih-Chun Chien,1 and Eddy Timmermans1

1Los Alamos National Laboratory, Los Alamos, NM 875452Santa Fe Institute, Santa Fe, NM 87501

3Department of Physics, University of New Hampshire, Durham, NH 03824(Dated: December 21, 2013, 11:52 P.M.)

We discuss the equilibrium properties of dilute Bose gases using a non-perturbative formalismbased on auxiliary fields related to the normal and anomalous densities. We show analytically thatfor a dilute Bose gas of weakly-interacting particles at zero temperature, the leading-order auxiliaryfield (LOAF) approximation leads to well-known analytical results. Close to the critical point theLOAF predictions are the same as those obtained using an effective field theory in the large-Napproximation. We also report analytical approximations for the LOAF results in the unitaritylimit, which compare favorably with our numerical results. LOAF predicts that the equation ofstate for the Bose gas in the unitarity limit is E/(pV ) = 1, unlike the case of the Fermi gas whenE/(pV ) = 3/2.

PACS numbers: 03.75.Hh, 05.30.Jp, 67.85.Bc

I. INTRODUCTION

Recently we introduced a new theoretical frameworkfor the study of a dilute gas of Bose particles withtunable interactions[1] based on a loop expansion ofthe one-particle irreducible (1-PI) effective action interms of composite-field propagators. The auxiliary field(AF) formalism makes use of the Hubbard-Stratonovitchtransformation[2] to rewrite the Lagrangian in terms ofauxiliary fields related to the normal and anomalousdensities. Employing general quantum field theoreticalmethods[3–7], the AF formalism is part of a continuingeffort in the community to apply methods traditionallyused in high-energy physics[8–18] to the study of ultra-cold atomic gases[19].

For an interacting dilute Bose gas, the leading-order auxiliary field (LOAF) approximation is a non-perturbative, conserving and gapless approximation thatdescribes a large interval of values of the couplingconstant, satisfies Goldstone’s theorem and yields asecond-order phase transition to a Bose-Einstein conden-sate (BEC) regime. In contrast with other resumma-tion schemes, such as the large-N expansion[20, 21] orthe functional renormalization techiques[16–18], here wetreat the normal and anomalous densities on equal foot-ing. LOAF produces the same slope of the linear depar-ture of the critical temperature from the noninteractinglimit derived by Baym et al.[21] using a large-N expan-sion for the critical theory. Unlike the large-N expansionsdeveloped by Baym et al., the LOAF approximation canbe used at all temperatures. Furthermore, one can sys-tematically improve upon the LOAF approximation bycalculating the 1-PI action order-by-order corrections.The broken U(1) symmetry Ward identities guaranteethe preservation Goldstone’s theorem order-by-order[5].

The detailed derivation of the LOAF approximationwas discussed recently for the case of dilute Bose[22] andFermi[23] atomic gases. Unlike the case of Fermi gaseswhere the LOAF approximation is equivalent to the stan-

dard Bardeen-Cooper-Schrieffer (BCS) ansatz [8, 9], inthe case of Bose gases the LOAF approximations leads toyet unexplored possibilities. Therefore it is important tostudy analytically the LOAF predictions for Bose gases inlimiting cases such as the case of weakly-interacting sys-tems and also in the unitarity limit, which correspondsto the strongly-interacting regime where the s-wave scat-tering length, a0, is much larger than the inter-particledistance. In the unitarity limit the properties of the sys-tem have a universal character[24]. The intrinsic non-perturbative character of the LOAF approximation maybecome particularly relevant because the developmentof novel cold atom technology that produce stable, flatpotentials bound by a sharp edge[25, 26] leads to theprospect of studying finite temperature properties of di-lute gases, such as the BEC transition temperature, Tc,superfluid to normal fluid ratio, depletion, and specificheat, at fixed particle density ρ.

In this paper we focus on the study of the LOAF pre-dictions in the broken-symmetry phase. We make contactwith existing analytical approximations in the weakly-interacting limit, such as those discussed in the textbookof Fetter and Walecka[27] and the analytical results ob-tained close to the critical temperature by Kita[28–30]using the related Luttinger-Ward functional. We will alsoshow that the analytical techniques developed here canbe used to study the LOAF predictions in the unitaritylimit.

This paper is organized as follows: In Sec. II, we brieflyreview the derivation of the LOAF equations. The LOAFeffective potential and the derivation of thermodynamicproperties are outlined in Sec. III. In Sec. IV we special-ize to the study of the interacting Bose properties in thebroken-symmetry phase. In Sec. V we study the zero-temperature LOAF results in the weakly-coupling limitand we compare with the weakly interacting Bose gastheory discussed by Fetter and Walecka[27]. The zero-temperature analysis suggests the scaling of the LOAFequations discussed in Sec. VI. In Sec. VII we discuss the

arX

iv:1

107.

3750

v1 [

cond

-mat

.qua

nt-g

as]

19

Jul 2

011

2

LOAF properties close to the critical temperature in theweakly-interacting limit and we compare with the resultsobtained by Kita using a related approximation[30]. An-alytical approximations of the LOAF predictions in theunitarity limit are discussed in Sec. VIII. We conclude inSec. IX.

II. LEADING ORDER AUXILIARY FIELD(LOAF) FORMALISM

The detailed derivation of the LOAF approximationfor the case of dilute Bose gases was discussed recentlyin Ref. 22. For completeness, we will review next thesalient aspects of the AF-formalism derivation.

In dilute bosonic gas systems, the classical action isgiven by

S[φ, φ∗ ] =

∫dx L[φ, φ∗ ] , (1)

with dx ≡ dtd3x and the Lagrangian density

L[φ, φ∗ ] =i~2

[φ∗(x) (∂t φ(x))− (∂t φ∗(x))φ(x) ]

− φ∗(x)(−~2∇2

2m− µ

)φ(x)− λ0

2|φ(x)|4 . (2)

Here, µ is the chemical potential and λ0 is the bare cou-pling constant. In the auxiliary field formalism we usethe Hubbard-Stratonovitch transformation[2] to elimi-nate the quartic interaction in Eq. (2) by introducing thereal and complex auxiliary fields (AF), χ(x) and A(x),related to the normal and anomalous densities. We addto Eq. (2) the AF Lagrangian density[3–5]

Laux[φ, φ∗, χ,A,A∗] =1

2λ0

(χ(x)−

√2λ0 |φ(x)|2

)2

− 1

2λ0

∣∣∣A(x)− λ0 φ2(x)

∣∣∣2 . (3)

Then, the action becomes

S[Φ, J ] = S[φa, χ,A,A∗, ja, s, S] (4)

= −1

2

∫∫dxdx′ φa(x)G−1a

b[χ,A](x, x′)φb(x′)

+

∫dx [

χ2(x)− |A(x)|2]/(2λ0)− s(x)χ(x)

+ S∗(x)A(x) + S(x)A∗(x) + j∗(x)φ(x) + j(x)φ∗(x),

with

G−1ab[χ,A] (5)

= δ(x, x′)

(−γ∇2− i~ ∂t + χ′ −A(x)

−A∗(x) −γ∇2 + i~ ∂t + χ′

),

where we introduced the notations γ = ~2/(2m) and

χ′ =√

2χ(x)− µ , (6)

together with a two-component notation, φa(x) =φ(x), φ∗(x) , for a = 1, 2. Φ(x) and J(x) signify thefive-component fields and currents.

The generating functional for connected graphs is

Z[J ] = eiW [J]/~ = N∫

DΦ eiS[Φ;J]/~ , (7)

with S[Φ; J ] given by Eq. (4). Performing the path inte-gral over the fields φa, we obtain the effective action forχ,A,A∗, as

Seff[χ; J, ε] =1

2

∫∫dx dx′ ja(x)G[χ]ab(x, x

′) ja(x) (8)

+

∫dx χi(x)χi(x)

2λ0− Si(x)χi(x)− ~

2iTr ln[G−1 ]

,

where

χi(x) =χ(x), A(x)/

√2, A∗(x)/

√2, (9)

Si(x) =s(x), S(x)/

√2, S∗(x)/

√2. (10)

Next, we expand the effective action about the station-ary points, χi0(x), defined by δSeff[χ; j]/δχi(x) = 0. Weobtain the “gap” equations:

χ0(x)

λ0=√

2[|φ0(x)|2 +

~2i

TrG(x, x)]

+ s(x) , (11)

A0(x)

λ0=[φ2

0(x) +~iG2

1(x, x)]

+ S(x) , (12)

where we introduced the notations φa0 [χ0](x) as

φa0 [χ0](x) =

∫dx′G[χ0]ab(x, x

′) jb(x′) . (13)

Both χ0 and A0 include self-consistent fluctuations.Expanding the effective action about the stationary

point, we write

Seff[χ; J ] = Seff[χ0; J ] +1

2

∫∫d4x d4x′D−1

ij [χ0](x, x′)

× [χi(x)− χi0(x)] [χj(x′)− χj0(x′)] + · · · , (14)

where D−1ij (x, x′) is given by the second-order derivatives,

D−1ij [χ0](x, x′) =

δ2 Seff[χa]

δχi(x) δχj(x′)

∣∣∣∣χ0

, (15)

evaluated at the stationary points. By keeping the gaus-sian fluctuations and Legendre transforming, the one-particle irreducible (1-PI) graphs generating functional

Γ[Φ] =

∫dx jα(x)φα(x)−W [J ] (16)

=1

2

∫∫dxdx′ φa(x)G−1[χ]ab(x, x

′)φb(x′)

−∫

dx χi(x)χi(x)

2λ0− ~

2iTr

ln[G−1[χ](x, x) ]

− ~2i

Tr ln[D−1ii [Φ](x, x) ]

+ · · · ,

3

is the negative of the classical action plus self-consistentone-loop corrections in the φa and χi propagators. Thelast term in Eq. (16) is next-to-leading order [5], and isnot included in the leading-order auxiliary field (LOAF)approximation. Hence, the static part of the effectiveaction per unit volume is

Veff[Φ] = χ′ |φ|2 − 1

2(A∗ φ2 +Aφ∗ 2)

− χ2 − |A|2

2λ0+

~2i

Tr ln[G−1[χ] ] . (17)

In the imaginary time formalism, the last term in Eq. (17)becomes

~2i

Tr ln[G−1[χ] ] =

∫d3k

(2π)3

ωk2

+1

βln[ 1− e−βωk ]

,

(18)where the dispersion relation is given by

ω2k = (εk + χ′)2 − |A|2 , (19)

with εk = γk2. At the minimum of the effective potential,we also have

δVeff[Φ]

δφ∗

∣∣∣φ0

= χ′ φ0 −Aφ∗0 = 0 . (20)

Using the U(1) gauge symmetry, we choose φ0 to be realin the broken-symmetry phase. Then, A is real and thedispersion, ω2

k = εk(εk + 2A), represents the Goldstonetheorem.

III. THERMODYNAMICS IN THE LOAFAPPROXIMATION

Using standard regularization techniques for effective-field theories[31, 32], the renormalized effective potential

Veff = (χ′ −A)ρ0 −(χ′ + µ)2

4λ+A2

2λ(21)

+

∫d3k

(2π)3

[1

2

(ωk − εk − χ′ +

A2

2εk

)+ T ln(1− e−ωk/T )

],

represents the grand potential per unit volume,

Veff = Ω[T, µ,V]/V . (22)

Here, ρ0 = φ20 is the condensate density, and the renor-

malized coupling constant is related to the s-wave scat-tering length by λ = 8πγ a0. The values of χ′ and A areobtained by solving self-consistently the gap equations

ρ = ρ0 +

∫d3k

(2π)3

(εk + χ′

2ωk− 1

2

)+

∫d3k

(2π)3

εk + χ′

ωknω ,

(23)

ρ0 =A

λ−∫

d3k

(2π)3

( A

2ωk− A

2εk

)−∫

d3k

(2π)3

A

ωknω , (24)

where nω = 1/(eωk/T − 1) is the Bose-Einstein particledistribution and we assume kB = 1 units.

Using the grand potential, Ω[T, µ,V], we calculate thetotal number of particles

N [T, µ,V] = −∂µΩ|T,V , (25)

the pressure

p[T, µ] = −∂VΩ|T,µ , (26)

entropy

S[T, µ,V] = −∂TΩ|µ,V . (27)

The energy is obtained as

E = Ω + TS + µN . (28)

Hence, the physical density is given by

ρ = −∂µVeff =1

2λ(χ′ + µ) , (29)

the pressure is

p = −Veff[ρ0, µ, T ] , (30)

and the entropy density, s = S/V, is

s =

∫d3k

(2π)3

[(1 + nω) ln(1 + nω)− nω lnnω

],

=

∫d3k

(2π)3

[− ln(1− e−ωk/T ) +

ωk/T

eωk/T − 1

]. (31)

The energy density, ε = E/V, is obtained as

ε = −p+ Ts+ µρ . (32)

IV. BROKEN SYMMETRY PHASE

In the following we will focus on the broken symmetryregion of the phase diagram, i.e. the regime where thedensity of the BEC condensate is nonzero, ρ0 6= 0. In thebroken-symmetry phase, T < Tc, we have equal normaland anomalous densities, χ′ = A = ∆, and the dispersionrelation becomes

ω2k = εk(εk + 2∆) . (33)

We note that in the long-wavelength limit, Eq. (33) re-duces to the linear dispersion relation

ωk ≈ ~k√

m, k → 0 , (34)

with the characteristic velocity (zero sound),√

∆/m.Comparing (33) with Eq. (21.11) in Ref. 27, i.e.

ωk ≈ ~k√ρ0 V (0)

m, k → 0 , (35)

4

we find that the parameter χ′ = A = ∆ in LOAF playsthe role of ρV (0) in the weakly interacting Bose gas the-ory discussed by Fetter and Walecka, with the zerothmoment of the potential

V (0) = 4π

∫V (r) r2 dr ↔ ∆

ρ0. (36)

In the broken-symmetry phase we can calculate explic-itly the temperature independent integrals in Eqs. (21),(23) and (24), as

∫d3k

(2π)3

(εk + ∆

2ωk− 1

2

)=

√2

3A∆3/2 , (37)∫

d3k

(2π)3

( ∆

2ωk− ∆

2εk

)= −

√2A∆3/2 , (38)∫

d3k

(2π)3

1

2

(ωk − εk − χ′ +

A2

2εk

)=

8√

2

15A∆5/2 , (39)∫

d3k

(2π)3

1

2

( εkωk

+∆

εk− 1)

=4√

2

3A∆3/2 . (40)

Here we introduced the notation A = (4π2)−1γ−3/2. Wehave also∫

d3k

(2π)3

εk + ∆

ωknω = AT 3/2

∫ ∞0

ε+ δ

ω(δ)

√ε dε

eω(δ) − 1, (41)∫

d3k

(2π)3

ωknω = AT 3/2

∫ ∞0

δ

ω(δ)

√ε dε

eω(δ) − 1, (42)∫

d3k

(2π)3ln(1− e−ωk/T )=AT 3/2

∫ ∞0

√ε dε ln

[1− e−ω(δ)

],

(43)

with δ = ∆/T . Hence, the gap equations (23) and (24)become

ρ = ρ0 +

√2

3A∆3/2 +AT 3/2

∫ ∞0

ε+ δ

ω(δ)

√ε dε

eω(δ) − 1,

(44)

ρ0 =∆

λ+√

2A∆3/2 −AT 3/2

∫ ∞0

δ

ω(δ)

√ε dε

eω(δ) − 1,

(45)

and the pressure (26) reads

p =λρ2 − 1

2λ∆2 − 8

√2

15A∆5/2 (46)

+AT 5/2

∫ ∞0

√ε dε ln

[1− e−ω(δ)

].

The entropy density is given by

s = AT 3/2

∫ ∞0

√ε dε

− ln

[1− e−ω(δ)

]+

ω(δ)

eω(δ) − 1

. (47)

V. ZERO TEMPERATURE PROPERTIES INTHE WEAKLY-INTERACTING LIMIT

With the above results we can study the propertiesof the zero-temperature Bose gas. For T = 0, the gapequations combine to give

∆0

λ ρ= 1− 4

√2

3

∆3/20 , (48)

ρ0

ρ= 1−

√2

3

∆3/20 , (49)

p0 = λρ2 − 1

2λ∆2

0 −8√

2

15A∆

5/20 . (50)

We also have ε0 = −p0 +µ0ρ, with µ0 = 2λρ−∆0. Herewe note: λρ = (8π ξ) γ ρ2/3, λA = (2/π)γ−1/2a0, withthe dimensionless parameter, ξ = ρ1/3a0.

It is convenient to introduce the following rescaled vari-ables

∆ =∆

λρ, µ =

µ

λρ, p =

p

λρ, ε =

ε

λρ. (51)

Then, we obtain

∆0 = 1− 32

3√πξ3/2 ∆

3/20 , (52)

ρ0/ρ = 1− 8

3√πξ3/2 ∆

3/20 , (53)

p0/ρ = 1− 1

2∆2

0 −64

15√πξ3/2 ∆

5/20 , (54)

and

ε0 = −p0 + µ0ρ , (55)

with

µ0 = 2− ∆0 . (56)

Iterating Eq. (52), we obtain the following successiveapproximations for ∆0,

∆0 ≈ 1− y , (57)

≈ 1− y (1− y)3/2 ,

≈ 1− y[1− y (1− y)3/2

]3/2, · · ·

with y = 32/(3√π) ξ3/2. Hence, we obtain

∆0 = 1− 32

3√πξ3/2 +

512

3πξ3 + · · · , (58)

and

µ0 = 1 +32

3√πξ3/2 − 512

3πξ3 + · · · . (59)

The coefficient of the ξ3/2 term above is the same as inEq. (22.20) in Ref. 27. Substituting Eqs. (57) in Eq. (53),

5

0.80

1.00

1 orderst

2 ordernd

0.95

0.96

0.97

0.98

0.99

0.84

0.88

0.92

0.96

0.50

0.54

0.56

0.60

0.0 0.02 0.04 0.06 0.08 0.10

p

/()

0

2

0.58

0.52

FIG. 1. (Color online) Comparisons of the exact zero-temperature values of the auxiliary field, ∆0, condensate frac-tion, ρ0/ρ, pressure, p0, and their respective first- and second-

order approximations, as a function of ξ = ρ1/3a0, in theweakly-interacting regime. The LOAF approximations for theauxiliary field, ∆0, condensate fraction, ρ0/ρ, and pressure,p0, in the weakly-interacting regime, are given in Eqs. (58),(61) and (63), respectively.

we obtain

ρ0/ρ ≈ 1− y

4(1− y)3/2 , (60)

≈ 1− y

4

[1− y (1− y)3/2

]3/2,

≈ 1− y

4

1− y

[1− y (1− y)3/2

]3/23/2

, · · ·

which gives

ρ0/ρ = 1− 8

3√πξ3/2 +

128

3πξ3 + · · · . (61)

The coefficient of the ξ3/2 term above is the same as inEq. (22.14) in Ref. 27. Similarly, from Eqs. (54) and (57),we obtain

p0/ρ = 1− 1

2

[(1− y)2 +

4

5y (1− y)5/2

], · · · (62)

which gives

p0/ρ =1

2+

32

5√πξ3/2 − 1024

9πξ3 + · · · , (63)

and

ε0/ρ =1

2+

64

15√πξ3/2 − 512

9πξ3 + · · · . (64)

Again, the coefficient of the ξ3/2 term above is the sameas in Eq. (22.19) in Ref. 27.

We note that the next-to-leading order correction tothe zero-temperature energy was calculated by Wu [33],yielding a second-order logarithmic term that cannot becaptured by LOAF, which is only a one-loop approx-imation. The second-order logarithmic correction waslater confirmed by Hugenhltz and Pines [34] and bySawada [35].

For illustrative purposes, in Fig. 1 we depict the exactzero-temperature values of the auxiliary field, ∆0, con-densate fraction, ρ0/ρ, pressure, p0, and their respectivefirst- and second-order approximations, as a function ofξ = ρ1/3a0, in the weakly-interacting regime.

By construction [22], in the weak coupling limitthe LOAF approximation agrees with the Bogoliubovapproximation[36, 37], which represents the leading-order low-density approximation of the theory. In ad-dition, the related Popov approximation can be obtainedfrom Eqs. (23) and (24) by setting A = χ′ = λρ0 andneglecting the quantum fluctuations in the anomalousdensity. We showed in Ref. 22 that the LOAF and the“gapless” Popov approximation results become qualita-tively similar in the weak coupling limit, even though theorder of the phase transitions remains different. As a con-sequence, the LOAF results in the weak-coupling limitdiscussed above agree with the Bogoliubov and Popovapproximations discussed for instance in the Andersen’sreview article[37].

VI. RESCALED EQUATIONS

The above zero-temperature results suggest that λρ isone of the two characteristic energy scales of the BECsystem. At finite temperature, we find the second en-ergy scale is given by T0 = 4πγ[ρ/ζ(3/2)]2/3, the criticaltemperature of the non-interacting Bose gas. It appearsthat T0 represents the temperature scale of the BEC sys-tem, and it is convenient to supplement the set of scaledvariables given in Eq. (51), by introducing the scaled tem-

perature, T = T/T0. In this context, we note the useful

results λρ/T0 = c ξ and (A/ρ)T3/20 = 2 [

√π ζ(3/2)]−1,

with c = 2 ζ23 (3/2).

In terms of the rescaled variables, we have

1 =ρ0

ρ+

8

3√πξ3/2 ∆3/2 + T 3/2 I1(δ) ,

ρ0

ρ= ∆ +

8√πξ3/2 ∆3/2 − T 3/2 I2(δ) , (65)

6

with the notations δ = ∆/T = c ξ (∆/T ), and

I1(δ) =2√

π ζ(3/2)

∫ ∞0

ε+ δ√ε+ 2δ

e√ε(ε+2δ) − 1

, (66)

I2(δ) =2√

π ζ(3/2)

∫ ∞0

δ√ε+ 2δ

e√ε(ε+2δ) − 1

, (67)

I3(δ) =2

√π ζ

53 (3/2)

∫ ∞0

√ε dε ln(1− e−

√ε(ε+2δ)) . (68)

The gap equations combine to give

∆ = 1− 32

3√πξ3/2 ∆3/2 − T 3/2

[I1(δ)− I2(δ)

], (69)

ρ0

ρ= 1− 8

3√πξ3/2 ∆3/2 − T 3/2 I1(δ) , (70)

p/ρ = 1− 1

2∆2 − 64

15√πξ3/2∆5/2 − 1

2ξT 5/2 I3(δ) . (71)

We introduce the rescaled entropy density as

s =T0

λρs , (72)

given as, see Eq. (47),

s

ρ=

1

2ξT 3/2

[I4(δ)− I3(δ)

], (73)

where we introduced the notation

I4(δ) =2

√π ζ

53 (3/2)

∫ ∞0

√ε dε

ω(δ)

eω(δ) − 1. (74)

The energy density is

ε = −p+ T s+ µρ . (75)

VII. CRITICAL PROPERTIES IN THEWEAKLY-INTERACTING LIMIT

In the following we will follow closely the approachoutlined by Kita in Ref. 30. It is important to note thatdespite the fact that Eqs. (69), (70) and (71) are differ-ent from Kita’s Eqs. (41)-(43), in the weakly-interactinglimit they become the same. Therefore, in the broken-symmetry phase, for T < Tc, our results match closelyKita’s results. Differences are simply due to the fact thatwe found better second-order approximations of the in-tegrals (66), (67) and (68).

In the weakly-interacting limit, we have δ 0 and thetemperature-dependent integrals can be approximated as(see App. A)

I1(δ) ≈ 1− b1 δ1/2+ b2 δ , (76)

I2(δ) ≈ b1 δ1/2 − b′2 δ , (77)

I3(δ) ≈ 1

ζ23 (3/2)

(b′0 + δ − 4b1

3δ3/2 +

3b14δ2), (78)

with b′0 = −ζ(5/2)/ζ(3/2), b1 =√

2π/ζ(3/2), b2 = b1/2,b′2 = b1. As indicated above, our approximations of theintegrals (66), (67) and (68) differ from Kita’s approxi-mations – see Eqs. (48a), (48b), and (48c) in Ref. 30 –at the second order in the b2 and b′2 coefficients.

Then, the gap equations (65) read

ρ0

ρ= 1− T 3

2

(1− b1x+ b2x

2), (79)

ρ0

ρ=x2T

cξ− T 3

2

(b1x− b′2x2

), (80)

where we introduced the variable x2 = δ = c ξ (∆/T ),

with c = 2 ζ23 (3/2). Combining Eqs. (79) and (80), we

obtain

x(T ) =b1(cξ)T

12

1 + (b2 + b′2)(cξ)T12

(81)

×

1 +

√1− T

32 − 1

b21(cξ)T 2

[1 + (b2 + b′2)(cξ)T

12

],

and we can write Eq. (80) as

ρ0

ρ=

1 + b′2(cξ)T12

cξTx[x− b1(cξ)T

12

1 + b′2(cξ)T12

]. (82)

At the critical point we have ρ0 = 0. Then, from Eq. (82),we obtain

xc ≡ x(Tc) =b1(cξ)T

12c

1 + b′2(cξ)T12c

. (83)

The expansion of Tc in powers of ξ is obtained fromEqs. (81) and (83) as

Tc =1 +2

3b21c ξ +

2

3b21c[7

6b21c− (b2 + b′2)c

]ξ2 + · · · .

(84)

which gives

Tc − T0

T0=

ζ43 (3/2)

ξ (85)

+8π

ζ43 (3/2)

[14π

3ζ43 (3/2)

− 3√

ζ13 (3/2)

]ξ2 + · · · .

Here, the linear coefficient is ≈ 2.33, whereas thequadratic coefficient is ≈ −3.23. The result for thelinear coefficient in ξ is the same as that obtain byBaym et al. using the large-N expansion for the criticaltheory[20, 21]. Kita also obtained this linear coefficient,see Eq. (52) in Ref. 30.

We note that the large-N expansion results obtainedby Baym et al. for the critical theory[20, 21] were laterimproved by Kleinert[38] and Kastening[39, 40] usingfive-, six- and seven-loop variational perturbation the-ory, respectively. At the seven-loop order, Kastening[40]

7

0.00

0.04

0.08

0.12

c

0.00

0.05

0.10

0.15

0.20

(T-T

)/T

c0

0

3

4

5

6

7

p/(

)c

5/3

0.0 0.02 0.04 0.06 0.08 0.10

1 orderst

2 ordernd

FIG. 2. (Color online) Comparisons of the exact critical valuesof the auxiliary field, ∆c, the ratio (Tc−T0)/T0, and pressure,pc, and their respective first- and second-order approxima-tions, as a function of ξ = ρ1/3a0, in the weakly-interactingregime. The LOAF approximations for the critical values ofthe auxiliary field, ∆c, the ratio (Tc − T0)/T0, and pressure,pc, in the weakly-interacting regime, are given in Eqs. (87),(85) and (92), respectively.

calculated a value (Tc − T0)/T0 = (1.27 ± 0.11)ξ, whichis in excellent agreement with Monte Carlo lattice field-theory results[41–43]. The quadratic coefficient was cal-culated by Arnold, Moore and Tomasik[44], yielding alsoa second-order logarithmic correction that cannot be cap-tured by LOAF, which is only a one-loop approxima-tion. Their result, (Tc − T0)/T0 ≈ (1.32 ± 0.02)ξ +[19.7518 ln(ξ) + (75.7 ± 0.4)]ξ2, indicates that two-loopcontributions are also important in determining the valueof the quadratic coefficient, as the LOAF result is toosmall.

Substituting Tc from Eq. (84) in Eq. (83), we obtain

xc = b1c ξ + b1c(1

3b21c− b′2c

)ξ2 + · · · . (86)

and

∆c

T0= b21c

2 ξ2 + b21c2(4

3b21c− 2b′2c

)ξ3 + · · · . (87)

In the latter, the quadratic coefficient is 8π/ζ23 (3/2) ≈

13.25, whereas the coefficient of the cubic term is ≈−34.76. The leading-order approximation was also ob-tained by Kita, see Eq. (53) in Ref. 30.

The temperature dependence of ∆ close to Tc is derived

from

x(T )− x(Tc) (88)

= −b1b2c2 ξ +

√b21c

2 ξ2 +3

2c ξ(

1− T

Tc

)+ · · · ,

as

∆(T )

T0− ∆(Tc)

T0(89)

= b21c2 ξ2 +

√4b41c

4 ξ4 + 6b21c3 ξ3(

1− T

Tc

)+ · · · .

The Eq. (88) above is the same as Kita’s Eq. 54. Simi-larly, we obtain

ρ0

ρ= b21c ξ +

√b41c

2 ξ2 +3

2b21c ξ

(1− T

Tc

)+ · · · . (90)

The critical pressure in the weakly-interacting limit isobtained from

pc/ρ ≈ 1− T5/2c

2ξζ23 (3/2)

(b′0 + x− 4b1

3x3 +

3b14x4). (91)

This gives

pc/ρ =ζ(5/2)

2ζ53 (3/2)

1

ξ+[1 +

5

6

ζ(5/2)

ζ53 (3/2)

b21c]

(92)

− b21c2

2ζ23 (3/2)

1− 5

3

ζ(5/2)

ζ(3/2)

[5

3b21 − (b2 + b′2)

]ξ + · · · .

In the noninteracting limit, we obtain pc(T0)/(γρ5/3) =

4π ζ(5/2)/ζ53 (3/2) ≈ 3.4. The coefficient of the linear

term is 8π[1 + (10π/3) ζ(5/2)/ζ3(3/2)

]≈ 44.94, whereas

the coefficient of the quadratic contribution is ≈ −6.42.For illustrative purposes, in Fig. 2 we depict the exact

critical values of the auxiliary field, ∆c, the ratio (Tc −T0)/T0, and pressure, pc, and their respective first- andsecond-order approximations, as a function of ξ = ρ1/3a0,in the weakly-interacting regime.

VIII. CRITICAL PROPERTIES IN THEUNITARITY LIMIT

The LOAF approximation is a non-perturbative ap-proximation. Therefore one can solve Eqs. (69) and (70)for arbitrary values of coupling constant related to thedimensionless parameter, ξ = ρ1/3a0, to calculate thevalues of the condensate fraction, auxiliary field, ∆, andall thermodynamical variables derivable from the grandpotential related to the pressure (71). For illustrativepurposes, in Fig. 3 we depict the zero-temperature val-ues of the auxiliary field ∆0, condensate fraction, ρ0/ρ,and chemical potential µ0, as a function of ξ, whereas inFig. 4 we show the dependence of the critical values of theauxiliary field, ∆c, the ratio (Tc − T0)/T0, and chemical

8

0.75

0.80

0.85

0.90

0.95

0/

0)

(/

1.0

1.2

1.4

1.6

1.8

100 101 102 10310-2 10-1

0.0

0.2

0.4

0.6

0.8

1.2

0T 0

1.0

3/4

1.147

2.0

FIG. 3. (Color online) Zero-temperature values of the auxil-iary field, ∆0, condensate fraction, ρ0/ρ, and chemical poten-

tial, µ0, as a function of ξ = ρ1/3a0. The condensate fraction,ρ0/ρ, and chemical potential, µ0/(λρ), plots depicted in thebottom two panels, were first shown in Fig. 2 of Ref. 22. Thereader is directed to Ref. 22 for further studies and deriva-tions. Here, we emphasize that the numerical values of theauxiliary field, ∆0, condensate fraction, ρ0/ρ, and chemicalpotential µ0, in the unitarity limit, compare well with the ex-act solutions given in Eq. (96), (97) and Eq. (98), respectively.

potential, µc, as a function of ξ = ρ1/3a0. From Fig. 4we conclude that in the unitarity limit (i.e. in the limitξ → ∞), we have ∆c/T0 → 0.873 and Tc/T0 → 1.396.

In the previous sections we focused on analytic approx-imations of the solutions to the gap equations, Eqs. (69)and (70), for the weakly-interacting limit. These approxi-mations were derived by dropping the terms proportionalto ∆3/2 and higher in Eqs. (69), (70) and (71), and byapproximating the integrals (66), (67) and (68) as de-scribed in App. A. This approach is made possible bythe fact that the critical value of the auxiliary field, ∆ issmall in the weakly-interacting limit, and ∆ 1 leadsto δ = ∆/T 1 as T > 1.

A similar approach is possible for obtaining analyti-cal approximations for the critical values of the auxiliaryfield, ∆c, and the critical temperature Tc, in the unitaritylimit. This approach requires improving the approxima-tions to the integrals (66), (67) and (68) described inApp. A, by supplementing the approximation of I2(δ)by the term (19/64)b1δ

3/2, and by adding the term−(99/256)b1δ

3/2 to the approximation of I1(δ) − I2(δ).This allows us to take into account the terms proportionalto ∆3/2 in Eqs. (69), (70) and (71). We will describe thisprocedure next.

0.0

0.2

0.4

0.6

0.8

1.0

0.873

0.0

0.1

0.2

0.3

0.4

0.396

cT 0

T/)T-

T(c

00

100 101 102 10310-2 10-1

c)

(/

1.70

1.75

1.80

1.85

1.90

1.952.0

FIG. 4. (Color online) Critical values of the auxiliary field,∆c, the ratio (Tc − T0)/T0, and chemical potential, µc, as a

function of ξ = ρ1/3a0. The critical auxiliary field, ∆c, andthe ratio (Tc − T0)/T0, depicted in the top two panels, werereported previously in Fig. 6 of Ref. 22, and further discus-sions regarding these plots can be found in Ref. 22. Here, weemphasize that the numerical values of the critical auxiliaryfield, ∆c, and the ratio (Tc − T0)/T0, in the unitarity limit,compare well with the semi-analytical approximations givenin Table 1. The numerical value of the chemical potential µc,agrees with the exact result obtained in Eq. (98).

As already discussed, the unitarity limit correspondsto the strongly-interacting limit, ξ →∞. In this regime,the energy scale λρ diverges, and the only scale remain-ing in the problem is the temperature scale, T0. Hence,in the unitarity limit the gap equations are scaled by in-troducing the variable, ∆ = ∆/T0. From Eqs. (44) and(45), we obtain

1 =ρ0

ρ+

2

3

√2

π

∆3/2

ζ(3/2)+ T 3/2 I1(δ) , (93)

ρ0

ρ= 2

√2

π

∆3/2

ζ(3/2)− T 3/2 I2(δ) , (94)

with δ = ∆/T = ∆/T . By adding the gap equations, weobtain also

1 =8

3

√2

π

∆3/2

ζ(3/2)+ T 3/2

[I1(δ)− I2(δ)

]. (95)

At zero temperature, T = 0, we can solve (95) to ob-tain the value of the auxiliary field in the unitarity limit(UL) at zero temperature, as

∆0/T0 =[ 3

8

√π

2ζ(3/2)

]2/3≈ 1.147 . (96)

9

TABLE I. Exact and approximate values of the criticalvalues of the auxiliary field, ∆c, and the critical temperature,Tc, in the unitarity limit.

Exact 1st order 2nd order 3rd order√δc 0.790 1.253 0.694 0.789

∆c/T0 0.873 4.619 0.653 0.875

Tc/T0 1.396 2.941 1.356 1.407

and from (94) we obtain the UL condensate fraction is

ρ0

ρ=

3

4. (97)

Alternatively, the UL condensate depletion fractionis 1/4.

The critical temperature, δc in the unitarity limit isobtained by solving Eq. (94) with ρ0 = 0. Assuming theexpansion of I2(δ) as a series powers of δ is truncatedby dropping terms proportional to δ3/2 and higher, wefind δc as the solution of a quadratic equation. Thenwe can calculate Tc = Tc/T0 from Eq. (95). Table Isummarizes approximations for the critical values of theauxiliary field, ∆c, and the critical temperature Tc, inthe unitarity limit. These approximations are comparedwith the numerical “exact” values. The errors of thethird-order approximations relative to the exact valuesare less than half of a percent.

Finally, we can show that the equation of state in theunitarity limit is independent of temperature. We begin,by calculating the UL asymptotes of the chemical poten-tial and pressure. For arbitrary temperature, Eqs. (56)and (71) give

µ = 2λρ , p = λρ2 . (98)

Then, from Eq. (55) we find the UL equation of state atzero temperature is

E0/(p0V ) = ε0/p0 = ε0/p0 = 1 . (99)

For finite temperature we use instead Eqs. (75) and (73)and obtain

E/(pV ) = ε/p = ε/p = 1 . (100)

This result is different than in the Fermi gas case,where the equation of state in the unitarity limit is,E/(pV )=3/2 [23, 45].

IX. CONCLUSIONS

In summary, in this paper we discussed analytical ap-proximations of the properties of dilute Bose gases us-ing the LOAF approximation in the weakly-interactingand the strongly-interacting (unitarity) limit. We focus

deliberately on the case of Bose gases in the broken-symmetry phase, in order to make contact with ana-lytical results already existing in the literature. Theweakly-interacting results at zero-temperature are shownto be identical with those found in the weakly interact-ing Bose gas theory discussed by Fetter and Walecka[27],whereas close to the critical temperature, the LOAF re-sults are similar to those obtained by Kita using the re-lated Luttinger-Ward functional[28–30]. In obtaining ourresults, we have improved the analytical approximationof temperature-dependent integrals, first discussed byKita[30]. These approximations were then applied to theanalytical study of the LOAF predictions in the unitaritylimit and found to give good agreement with our numeri-cal results. LOAF predicts that the equation of state forthe Bose gas in the unitarity limit is E/(pV ) = 1, unlikethe case of the Fermi gas when E/(pV ) = 3/2.

ACKNOWLEDGMENTS

Work performed in part under the auspices of the U.S.Department of Energy. The authors would like to thankE. Mottola for useful discussions and the Santa Fe Insti-tute for its hospitality during this work.

Appendix A: Approximations of certain integrals

For completeness, in this appendix we will derivethe first- and second-order approximations of the inte-grals (66), (67) and (68). Our approach follows closelythe discussion in Kita’s paper, see Ref. 30. Our first-orderapproximations are the same as Kita’s, but we differ atthe second order.

Consider the integral

I1(z) =2√

π ζ(3/2)

∫ ∞0

ε+ z

ω

√ε dε

eω − 1, (A1)

with ω =√ε(ε+ 2z). In the weakly-coupling limit we

have z → 0, as λ→ 0, and we seek a power expansion ofI1(z) in powers of z. We have

ε+ z

ω

√ε

eω − 1=

√ε

eε − 1−√ε eε

(eε − 1)2+O(ε2) . (A2)

Upon integration with respect to ε, we find that only thefirst term converges,

b0 =2√

π ζ(3/2)

∫ ∞0

√ε dε

eε − 1= 1 . (A3)

Therefore, we can write I1(z) = b0 + I1,∞(z), with theremainder

I1,∞(z) =2√

π ζ(3/2)

∫ ∞0

dε(ε+ z

ω

√ε

eω − 1−√ε

eε − 1

).

(A4)

10

The leading-order approximation of I1,∞(z) is obtainedby expanding out the exponentials in the denominator tofirst order. We obtain

I(1)1,∞(z) =

2√π ζ(3/2)

∫ ∞0

( ε+ z

ε+ 2z− 1) dε√

ε= −b1

√z ,

(A5)

with b1 =√

2π/ζ(3/2). Note that it is important to treaton equal footing the two exponentials in Eq. (A4) in orderto obtain the correct divergence subtraction.

The second-order correction to I1,∞(z) corresponds tothe term linear in z in the power expansion of

ε+ z

ε+ 2z

1√ε

( 1

1 + 12

√ε(ε+ 2z)

− 1). (A6)

This gives

I(2)1,∞(z) =

2√π ζ(3/2)

∫ ∞0

z√ε

(ε+ 2)2dε = b2 z , (A7)

with b2 = b1/2. Therefore, the second-order approxima-tion of I1(z) is

I1(z) =2√

π ζ(3/2)

∫ ∞0

ε+ z

ω

√ε dε

eω − 1≈ 1− b1

√z + b2 z .

(A8)The first- and second-order approximations of I1(z) areillustrated in Fig. 5. The coefficient b2 = 0.559 calcu-lated by Kita[30] results in a worse second-order approx-imations.

Similarly, we can evaluate the expansion in powers ofz of the integral

I2(z) =2√

π ζ(3/2)

∫ ∞0

z

ω

√ε dε

eω − 1. (A9)

In this case, there is no analytic part identified after per-forming the z expansion of the integrand in I2(z). There-fore, we have I2(z) = I2,∞(z), with the remainder

I2,∞(z) =2√

π ζ(3/2)

∫ ∞0

z√ε+ 2z

eω − 1. (A10)

0.0 0.2 0.3 0.4 0.5z

0.4

0.6

0.8

1.0

2 ordernd

1 orderst

exact

0.10.2

I(z

)1

FIG. 5. (Color online) Comparisons of the exact z depen-dence of I1(z), and the first- and second-order approximationsin powers of z.

2 ordernd

1 orderst

exact

0.1

0.2

0.3

0.4

0.0

I(z

)2

0.00 0.10 0.15 0.20z

0.05

FIG. 6. (Color online) Comparisons of the exact z depen-dence of I2(z), and the first- and second-order approximationsin powers of z.

-0.28

-0.26

-0.24

-0.22

-0.20

-0.18

0.0 0.2 0.3 0.4z

0.1

I(z

)3

2 ordernd

1 orderst

exact

FIG. 7. (Color online) Comparisons of the exact z depen-dence of I3(z), and the first- and second-order approximationsin powers of z.

The first-order approximation of I2,∞(z) is obtained byexpanding out the exponential in the denominator to firstorder. We obtain

I(1)2,∞(z) =

2√π ζ(3/2)

∫ ∞0

z

ε+ 2z

dε√ε

= b′1√z , (A11)

with b′1 = b1. The second-order correction to I2,∞(z) cor-responds to the term linear in z in the power expansionof

z

ε+ 2z

1√ε

( 1

1 + 12

√ε(ε+ 2z)

− 1). (A12)

This gives

I(2)2,∞(z) = − 2√

π ζ(3/2)

∫ ∞0

z

ε+ 2

dε√ε

= − b′2 z , (A13)

with b′2 = b1. Therefore, the second-order approximationof I2(z) is

I2(z) =2√

π ζ(3/2)

∫ ∞0

z

ω

√ε dε

eω − 1≈ b1

√z− b′2 z. (A14)

The first- and second-order approximations of I2(z) areillustrated in Fig. 6. The coefficient b′2 = 1.12 calculated

11

by Kita[30] results in a worse second-order approxima-tions.

Finally, we also need an approximation for the integral

I3(z) =2

√π ζ

53 (3/2)

∫ ∞0

√ε dε ln(1− e−ω) . (A15)

We note that the derivative of I3(z) with respect to z canbe written in terms of I1(z) and I2(z), as

dI3(z)

dz=

2√π ζ

53 (3/2)

∫ ∞0

ε

ω

√ε dε

eω − 1∝ I1(z)− I2(z) .

(A16)

Therefore, we can write

I3(z) =1

ζ23 (3/2)

b′0 +

∫ z

0

dz′[I1(z′)− I2(z′)

], (A17)

where b′0 = −ζ(5/2)/ζ(3/2). Thus, we obtain the second-order approximation of I3(z) as

I3(z) ≈ 1

ζ23 (3/2)

(b′0 + z − 4b1

3z3/2 +

3b14z2). (A18)

The first- and second-order approximations of I3(z) areillustrated in Fig. 7.

[1] F. Cooper, C.-C. Chien, B. Mihaila, J. F. Dawson, andE. Timmermans, Phys. Rev. Lett. 105, 240402 (2010).

[2] J. Hubbard, Phys. Rev. Lett. 3, 77 (1959); R. L.Stratonovich, Doklady 2, 416 (1958).

[3] S. Coleman, R. Jackiw, and H. D. Politzer, Phys. Rev.D 10, 2491 (1974).

[4] R. G. Root, Phys. Rev. D 10, 3322 (1974).[5] C. Bender, F. Cooper, and G. Guralnik, Ann. Phys. 109,

165 (1977).[6] J. W. Negele and H. Orland, Quantum Many-Particle

Systems (Addison-Wesley, New York, NY, 1988).[7] E. A. Calzetta and B.-L. B. Hu, Nonequilibrium quan-

tum field theory (Camb. U. Press, Cambridge, England,2008).

[8] C. A. R. Sa de Melo, M. Randeria, and J. R. Engelbrecht,Phys. Rev. Lett. 71, 3202 (1993).

[9] J. R. Engelbrecht, M. Randeria, and C. A. R. Sade Melo,Phys. Rev. B 55, 15153 (1997).

[10] E. Braaten and A. Nieto, Phys. Rev. B 56, 14745 (1997).[11] A. M. Rey, B. L. Hu, E. Calzetta, A. Roura, and C. W.

Clark, Phys. Rev. A 69, 033610 (2004).[12] T. Gasenzer, J. Berges, M. G. Schmidt, and M. Seco,

Phys. Rev. A 72, 063604 (2005).[13] K. Temme and T. Gasenzer, Phys. Rev. A 74, 053603

(2006).[14] J. Berges and T. Gasenzer, Phys. Rev. A 76, 033604

(2007).[15] I. Tikhonenkov, J. R. Anglin, and A. Vardi, Phys. Rev.

A 75, 013613 (2007).[16] S. Friederich, H. C. Krahl, and C. Wetterich, Phys. Rev.

B 81, 235108 (2010).[17] S. Floerchinger and C. Wetterich, Phys. Rev. A 77,

053603 (2008).[18] S. Floerchinger, M. Scherer, S. Diehl, and C. Wetterich,

Phys. Rev. B 78, 174528 (2008).[19] J. O. Andersen, Rev. Mod. Phys. 76, 599 (2004).[20] G. Baym, J.-P. Blaizot, M. Holzmann, F. Laloe, and

D. Vautherin, Phys. Rev. Lett. 83, 1703 (1999).[21] G. Baym, J.-P. Blaizot, and J. Zinn-Justin, Europhys.

Lett. 49, 150 (2000).[22] F. Cooper, B. Mihaila, J. F. Dawson, C.-C. Chien, and

E. Timmermans, Phys. Rev. A 83, 053622 (2011).[23] B. Mihaila, J. F. Dawson, F. Cooper, C.-C. Chien, and

E. Timmermans, Phys. Rev. A 83, 053637 (2011).[24] Y. Shin, C. H. Schunck, A. Schirotzek, and W. Ketterle,

Phys. Rev. Lett. 99, 090403 (2007).[25] K. Henderson, H. Kelkar, T. C. Lee, B. Gutirez-Medina,

and M. G. Raizen, Europhys. Lett. 75, 392 (2006).[26] K. Henderson, C. Ryu, C. MacCormic, and M. Boshier,

New J. Phys. 11, 043030 (2009).[27] A. L. Fetter and J. D. Walecka, Quantum theory of many-

particle systems (McGraw-Hill, New York, NY, 1971).[28] T. Kita, J. Phys. Soc. Japan 74, 1891 (2005).[29] T. Kita, J. Phys. Soc. Japan 74, 3397 (2005).[30] T. Kita, J. Phys. Soc. Japan 75, 044603 (2006).[31] T. Papenbrock and G. F. Bertsch, Phys. Rev. C 59, 2052

(1999).[32] A. Bulgac and Y. Yu, Phys. Rev. Lett. 88, 042504 (2002).[33] T. T. Wu, Phys. Rev. 115, 1390 (1959).[34] N. M. Hugenholtz and D. Pines, Phys. Rev. 116, 489

(1959).[35] K. Sawada, Phys. Rev. 116, 1344 (1959).[36] N. N. Bogoliubov, J. Phys. USSR 11, 23 (1947).[37] J. O. Andersen, Rev. Mod. Phys. 76, 599 (2004).[38] H. Kleinert, MODERN PHYSICS LETTERS B 17, 1011

(2003).[39] B. Kastening, Phys. Rev. A 68, 061601 (2003).[40] B. Kastening, Phys. Rev. A 69, 043613 (2004).[41] P. Arnold and G. D. Moore, Phys. Rev. Lett. 87, 120401

(2001).[42] V. A. Kashurnikov, N. V. Prokof’ev, and B. V. Svis-

tunov, Phys. Rev. Lett. 87, 120402 (2001).[43] P. Arnold and G. D. Moore, Phys. Rev. E 64, 066113

(2001).[44] P. Arnold, G. Moore, and B. Tomasik, Phys. Rev. A 65,

013606 (2001).[45] T.-L. Ho, Phys. Rev. Lett. 92, 090402 (2004).