You are on page 1of 4

PHYSICAL REVIEW A 85, 063613 (2012)

Spin-Seebeck effect in a strongly interacting Fermi gas

C. H. Wong, H. T. C. Stoof, and R. A. Duine


Institute for Theoretical Physics, Utrecht University, Leuvenlaan 4, 3584 CE Utrecht, The Netherlands
(Received 21 March 2012; published 15 June 2012)
We study the spin-Seebeck effect in a strongly interacting, two-component Fermi gas and propose an experiment
to measure this effect by relatively displacing spin-up and spin-down atomic clouds in a trap using spin-dependent
temperature gradients. We compute the spin-Seebeck coefficient and related spin-heat transport coefficients as
functions of temperature and interaction strength. We find that, when the interspin scattering length becomes
larger than the Fermi wavelength, the spin-Seebeck coefficient changes sign as a function of temperature, and
hence so does the direction of the spin separation. We compute this zero-crossing temperature as a function of
interaction strength and in particular in the unitary limit for the interspin scattering.

DOI: 10.1103/PhysRevA.85.063613 PACS number(s): 67.85.−d, 67.10.Fj, 67.10.Jn

I. INTRODUCTION
Spin caloritronics, the study of coupled spin and heat
∇Ts
transport, is a rapidly developing subfield of spintronics [1]. In
particular, the spin-dependent generalization of the Seebeck
effect, called the spin-Seebeck effect, has been intensively
studied in the solid-state environment [2]. Recently, there has
been broad interest in exploring spintronic phenomena in cold
atomic systems [3–5]. Spin transport in a strongly interacting,
two-component Fermi gas was investigated experimentally in
Ref. [6]. It is the purpose of this article to study the associated
heat transport, i.e., thermo-spin effects, in a similar setting.
In the ordinary Seebeck effect in metals, an electrochemical xs
potential gradient is generated by applying a temperature
gradient. Similarly, for a gas with two spin states, the spin- FIG. 1. (Color online) Spin-up and spin-down atomic clouds are
Seebeck coefficient Ss determines the spin chemical potential spatially relatively displaced in the presence of a spin temperature
µs generated by a spin temperature gradient ∇Ts through the gradient. The distance between the center of mass of the different
relation ∇µs = Ss ∇Ts , where µs ≡ µ+ − µ− and Ts ≡ T+ − spin components, denoted by xs , is proportional to the spin-Seebeck
T− , µσ and Tσ being the spin-dependent chemical potential and coefficient.
temperatures of the spin σ atoms, respectively, and we label
spin components by + and −. To measure the spin-Seebeck zeros of Ss as a function of kF a in Fig. 3(d). The temperature
coefficient, we propose relatively displacing the center of of the zero crossing reaches a universal value, T0 , in the
mass of spin-up and spin-down atom clouds in a harmonic unitary limit for the interspin scattering length kF a → ∞. We
trap by applying a spin-dependent temperature gradient, for find T0 # 0.378 TF in our calculation, where TF is the Fermi
example, by selectively heating one spin component with a temperature.
laser, as illustrated in Fig. 1. The locations x± of the center The thermodynamic reciprocal of the spin-Seebeck effect is
of mass of the spin-up and spin-down atoms are shifted to the spin-Peltier effect, in which a spin-dependent heat current
the minimum of µ± + V , where V is the trapping potential, proportional to the spin-Seebeck coefficient is induced by a
resulting in a spin separation, xs = x+ − x− = Ss ∇Ts /mω2 , spin current. This effect will heat up spin-up and spin-down
where m is the mass of the atoms and ω is the trap frequency components differently and provides another way to measure
in the direction of the temperature gradients. For an order of Ss . Furthermore, as discussed in Ref. [5], the spin-Seebeck
magnitude estimate, we take Ss # 0.01kB , as verified below. effect contributes to the total dissipation so that Ss can also be
For ∇Ts = 10−5 K/cm [7], ω = 2π × 1.46, we find xs # 1 measured through the heating.
mm, which is well within experimental resolution. We note that the spin-Seebeck coefficient was calculated for
We have computed the spin-Seebeck coefficient for a a weakly interacting Bose gas in Ref. [5], but the Bose gas is
two-component Fermi gas, plotted in Fig. 2, as a function unstable toward the formation of molecules for large scattering
of temperature and for several values of the interaction lengths, which makes the strongly interacting regime more
strength kF a, where kF is the Fermi wave vector and a difficult to realize experimentally.
is the interspin scattering length. As seen from the figure,
for weak interactions (kF a < 1), Ss is small and negative,
II. PHENOMENOLOGY
while for strong interactions (kF a ! 1), Ss is larger and
its sign changes as a function of temperature. In terms of We are specifically interested in phenomena due to spin-
the experiment mentioned above, this means that the spin drag, the transfer of momentum between different spins due
displacement changes direction as a function of temperature, to interspin scattering, which allows one to generate currents in
which is an interesting qualitative effect. We also plot the one spin species by applying forces on the other, and we define

1050-2947/2012/85(6)/063613(4) 063613-1 ©2012 American Physical Society


C. H. WONG, H. T. C. STOOF, AND R. A. DUINE PHYSICAL REVIEW A 85, 063613 (2012)

Ss
s
s
kB kF k B kF F
1.2
0.05 0.1
kF a .1

0.04 kF a 1 0.6 0.06

T 0.02
0.03 kF a 10
0.1 T
0.5 1.2 2 TF 0.2 1 2 TF
0.02
(a) (b)
T0
0.01
ZsT TF
T TF 0.004 1
0.5 1.0 1.5 2.0

0.01 0.6
0.001
FIG. 2. (Color online) The Seebeck coefficient plotted as a T 0.2
function of the reduced temperature T /TF , where TF is the Fermi kF a
0.2 1 2 TF 2 6 10
temperature, for values of kF a representing weak (kF a = 0.1) (c) (d)
and strong (kF a = 1) coupling and approaching the unitary limit
(kF a = 10). FIG. 3. (Color online) (a) A plot of the spin conductivity
normalized as h̄&σs , (b) the spin-heat conductivity normalized as
a set of spin-heat transport coefficients which captures these h̄&κs /kB 'F , and (c) the figure of merit Zs T , for kF a = 0.1 (thick
effects as follows. We consider a Fermi gas with two different blue), kF a = 1 (dashed purple), and kF a = 10 (dotted yellow). (d) A
spin states selected from a larger half-integer spin multiplet, plot of the temperatures T0 relative to TF where Ss = 0 as a function
which we will call “spin up” (+) and “spin down” (−), for of kF a.
equal spin-up and spin-down densities n+ = n− ≡ n, i.e., in
the absence of spin polarization, and apply equal and opposite A spin-dependent temperature gradient can only be estab-
forces and temperature gradients for the two spin species, i.e., lished when intraspin scattering is much greater than interspin
F+ = −F− and ∇T+ = −∇T− . In linear response, the ensuing scattering. For fermions, the first nonvanishing intraspin
spin current and spin-heat current defined by js = j+ − j− and scattering amplitude is p wave. To make this large, one can tune
qs = q+ − q− , respectively, are given by the p-wave scattering length by a Feshbach resonance [12].
! " ! "! " Taking the unitary limit for intraspin scattering, the intra- and
js 1 Ss Fs interspin differential cross sections are given by
= σs T S κs (1 + Z T ) , (1)
qs s σs s −∇Ts
dσ ++ dσ −− 9(p̂r · p̂'r )2
where Fs ≡ F+ − F− is the spin force, Ts = T+ − T− is = = ,
d( d( (pr /h̄)2
the spin temperature, T is the equilibrium temperature, σs (2)
is the spin conductivity, κs is the spin-heat conductivity (at dσ +− a2
= ,
zero spin current), Zs T = σs Ss2 T /κs , and Onsager reciprocity d( 1 + (pr a/h̄)2
is explicitly included in the matrix above. We note that Fs
where pr = |pr | is the relative momentum of incoming parti-
is the thermodynamic force which includes forces coming
cles with momenta p1 and p2 , defined by pr = (p1 − p2 )/2,
from pressure gradients, i.e., Fs = fsext − ∇ps /n, where fsext is
the hat superscripts denotes unit vectors, and a is the interspin
the external spin force, and ps = p+ − p− is the difference
s-wave scattering length.
in pressures of the spin-up and spin-down atoms. These
coefficients, computed with the Boltzmann equation described
below, are plotted in Fig. 3 as functions of T /TF for several III. CALCULATION OF TRANSPORT COEFFICIENTS
values of kF a. Next, we present the computation of Ss using the Boltzmann
As is well known, the spin conductivity σs rapidly increases equation. We parametrize the nonequilibrium, steady-state
at low temperatures due to Pauli blocking. Our result for distribution by
σs , plotted in Fig. 3(a), includes corrections due to spin-heat
coupling, but they are negligibly small, so that one can safely npσ (r) = fpσ (r) − ∂' fp0 φpσ (r), (3)
take σs = nτs /m with τs being the spindrag relaxation time [8]
measured in Ref. [6] and calculated in Ref. [9]. The downturn where fp0 = {exp[('p − µ)/kB T ] + 1}−1 is the equilibrium
of Ss at low temperatures is a quantum mechanical effect Fermi distribution, µ is the chemical potential, 'p = p2 /2m,
that also occurs for bosons, where, in contrast to fermions, fpσ (r,t) = (exp {['p − µσ (r)]/kB Tσ (r)} + 1)−1 is the local
the spin conductivity decreases sharply at low temperatures equilibrium distribution, ∂' fp0 = −fp0 (1 − fp0 )/kB T , and φpσ
due to bosonic enhancement of scattering [10]. The spin-heat is determined by solving the Boltzmann equation for the spin
conductivity κs is plotted in Fig. 3(b), where it is seen to distribution nps = np+ − np− in linear response,
increase with increasing T . The dimensionless figure of merit ! "
Zs T [Fig. 3(c)] determines the thermodynamic efficiency of 'p − w(T )
∂' fp0 vp = Cp [φ], (4)
engines based on thermo-spin effects [11]. kB T

063613-2
SPIN-SEEBECK EFFECT IN A STRONGLY INTERACTING . . . PHYSICAL REVIEW A 85, 063613 (2012)

m 3
s ns px, 0, 0
s 2 kB x Ts kF T djs
0.2 kF x Ts T d
0.2
0.6 0.1 0.005
px
0.1 2 1 1 2 pF 1 2 3 4 F
0.1 0.005
0.1 a a 0.2
1 3 5 1 3 5 0.010
(a) (b) (a) (b)
Ss ZsT
kB FIG. 5. (Color online) A plot of (a) the perturbed spin distribution
0.008 per spin temperature gradient, normalized as δns /(−∂x T /kF T )
0.08
along the px axis. A plot of (b) the spin current density along
0.005
0.04 the px axis, per d' per spin temperature gradient, normalized as
0.002 (djs /d')(h̄/kF (∂x Ts /T )), for kF a = 10 and temperatures at which
a a
Ss is negative (T = 0.3TF , blue thick line) and positive (T = 0.5TF ,
0.01 1 3 5 1 3 5
purple dashed line).
(c) (d)

FIG. 4. (Color online) Plots of the transport coefficients in the


where we define +++ [φp ] = φp3 + φp4 − φp1 − φp2 and
high-temperature limit, T ( TF , as a function of a/&. (a) The spin ++− [φp ] = φp3 − φp4 − φp1 + φp2 for an arbitrary function
conductivity normalized as h̄&σs . (b) The spin-heat conductivity φp , and in the integrand momentum conservation is satisfied:
normalized as m&3 κs /2π h̄kB , (c) Ss /kB , and (d) Zs T . p1 + p2 = p3 + p4 .
From Eqs. (5), (6), and (1), it follows that the Seebeck
where w(T ) = µ + T s is the enthalpy per particle and s coefficient is given in terms of b0 and b1 , and the spin
is the entropy per particle [13]. We defined φps ≡ kB φ p · conductivity σs by
(−∇Ts ) and expressed the linearized collision integral in the $ %
n w(T )
Boltzmann equation for the spin distribution as (∂nps /∂t)coll ≡ Ss (a,T ) = b0 (a,T ) + b1 (a,T ) . (10)
σs kB T
Cp [φ p ] · (−∇Ts ). The spin current is given by
# The other transport coefficients are calculated similarly. In
d 3p order to make the numerics more tractable, we have omitted
js = − ∂' fp0 vp φps . (5)
(2π h̄)3 the angular dependence in dσ++ /d( [c.f. Eq. (2)]. On the other
hand, in the high-temperature limit, the integrals in Eq. (9) in-
We solve Eq. (4) using the method described in Ref. [5].
cluding the angular factor can be done analytically [14], which
Applying the temperature gradient along the x axis, we
allows us to calculate the coefficients in the high-temperature
parametrize the response by a power series,
$ ! "% limit. The result is +shown in Fig. 4 plotted as a function
'p of a/&, where & = 2π h̄2 /mkB T is the thermal deBroglie
φps (a,T ) = b0 (a,T ) + b1 (a,T ) px (−kB ∂x Ts ).
kB T wavelength. In this limit, the dependence of σs on the scattering
(6) length a can be understood from classical considerations. The
The coefficients b0 and b1 , determined by the approximate spin conductivity is approximately related to the interspin
solution to Eq. (4), are given by collision time τ+− by σs ∝ nτ+− and 1/τ+− = nσ̄+− vT , where
& ' & ' σ̄+− is the interspin cross section and vT ∝ 1/& is the
b0 (a,T ) 3nl(T ) −C01 (a,T ) average thermal velocity. In the limit a/& → 0, σ̄+− ∝ a 2 ,
= , (7)
b1 (a,T ) 2
C00 C11 − C01 C00 (a,T ) thus &σs ∝ (&/a)2 , and when a/& → ∞, σ̄+− ∝ &2 , thus
where &σs ∝ 1, in agreement with our result.
! "2 The behavior of the transport coefficients depends crucially
35 f7/2 (z) w(T ) on the shape of the perturbed spin distribution δnps =
l(T ) = − , (8) −∂' fp0 φps and the associated spin current density, which
4 f3/2 (z) kB T
we plot for kF a = 10 in Fig. 5. The positive (negative)
z = eµ/kB T is the fugacity, fn (z) = −Lin (−z), Lin (z) are the parts of δnpσ may be regarded as particles (holes) having
polylogarithmic functions, and Cnm are the 2 × 2 matrix group velocities ±p/m. Since δnpσ = −δnp−σ , every spin-up
elements of the collision integral, particle is matched with a spin-down hole with the same
# momentum, resulting in the spin current. Thus, the sign of
1 dp1 dp2 |p1 − p2 | 0 0
Cnm = fp1 fp2 Ss is determined by the relative number of particles or holes
kB T (2π h̄)6 m induced in response to the spin temperature gradient.
#
1 ( )( ) Next, we give a criterion that determines the sign of the
× d(r 1 − fp03 1 − fp04
4 spin-Seebeck coefficient and show that it does not depend on
* &dσ+σ ' the form of the intraspin scattering at all. First, we note that
× ++σ [('p /kBT ) p] · ++σ [('p /kB T ) p] ,
n m
in our solution given by Eq. (7), we always have b0 < 0 and
d(r
σ =± b1 > 0 because l(T ) > 0, Cnm > 0, and det Ĉ > 0 [15]. Thus
(9) for positive momenta, b1 (b0 ) corresponds to particles (holes)

063613-3
C. H. WONG, H. T. C. STOOF, AND R. A. DUINE PHYSICAL REVIEW A 85, 063613 (2012)

created above (below) the Fermi surface. Inspecting Eq. (10), a 2 , when kF a , 1 or T /TF , 1, to momentum-dependent
we find that the criteria to have Ss " 0 is scattering, dσ/d( ∼ p˜r −2 , when kF a # 1 and T # TF .
, , , ,
, b0 , , C01 , w
, ,=, ,
,b , ,C , ! k T (11) IV. DISCUSSIONS AND OUTLOOK
1 00 B

and the opposite inequality for Ss > 0. Thus Ss is positive when We note that as the temperature is lowered, one expects
b1 becomes large enough to violate Eq. (11) [16]. Furthermore, to enter the Fermi-liquid regime, for Tc < T , TF , where
it turns out that the only integral in Eq. (9) containing intraspin Tc is the temperature for the superfluid transition, and one
scattering that is nonvanishing is C11 , which does not enter in should use Fermi-liquid scattering amplitudes in the collision
Eq. (11). integral [9]. Near and above Tc , one should also take into
The temperature dependence of b0 and b1 follows from account the effects of pairing correlations on the interspin
the temperature dependence of the collision matrix elements interaction [17]. Both these regimes can be analyzed with the
in Eq. (9),√which are given by integrals nonvanishing only formalism presented in this paper and will be relegated to
for pr ∼ 4π h̄/&. Therefore, it is useful to express the future work.
differential cross section
√ Eq. (2) in terms of the rescaled
momentum p̃r = (&/ 4π h̄)pr ,
ACKNOWLEDGMENTS
dσ +− 1 (kF a)2
= 2 . (12) This work was supported by Stichting voor Fundamenteel
d( kF 1 + (kF a)2 (T /TF )p̃r , Onderzoek der Materie (FOM), the Netherlands Organization
From Eq. (12) we see that the change in the sign of Ss is for Scientific Research (NWO), and the European Research
related to the crossover from hard-sphere scattering, dσ/d( # Council (ERC) under the Seventh Framework Program (FP7).

[1] G. E. W. Bauer, A. H. MacDonald, and S. Maekawa, Solid State [9] G. M. Bruun, New J. Phys. 13, 035005 (2011).
Communications 150, 459 (2010). [10] H. J. van Driel, R. A. Duine, and H. T. C. Stoof, Phys. Rev. Lett.
[2] M. Hatami et al., Solid State Commun. 150, 480 (2010); C. M. 105, 155301 (2010).
Jaworski et al., Nat. Mater. 9, 898 (2010); A. Slachter et al., Nat. [11] G. Mahan, Solid State Physics (Academic Press, San Diego,
Phys. 6, 879 (2010). 1997), Vol. 51, pp. 81–157.
[3] Y. J. Lin, K. Jimenez-Garcia, and I. B. Spielman, Nature [12] C. A. Regal, C. Ticknor, J. L. Bohn, and D. S. Jin, Phys. Rev.
(London) 471, 83 (2011); T. D. Stanescu, C. Zhang, and Lett. 90, 053201 (2003); K. B. Gubbels and H. T. C. Stoof, ibid.
V. Galitski, Phys. Rev. Lett. 99, 110403 (2007). 99, 190406 (2007).
[4] R. A. Duine and H. T. C. Stoof, Phys. Rev. Lett. 103, 170401 [13] Since s ∼ T /TF , for degenerate fermions, one usually takes
(2009). w → µ.
[5] C. H. Wong, H. J. van Driel, R. Kittinaradorn, H. T. C. Stoof, [14] Specifically, Cnm can be expressed in terms of incomplete -
and R. A. Duine, Phys. Rev. Lett. 108, 075301 (2012). functions.
[6] A. Sommer et al., Nature (London) 472, 201 (2011). [15] The eigenvalues of the collision matrix are negative because
[7] R. Meppelink, R. van Rooij, J. M. Vogels, and P. van der Straten, they correspond to relaxation times, and therefore, for the 2 × 2
Phys. Rev. Lett. 103, 095301 (2009). collision matrix, the determinant must be positive.
[8] We note that the measured values of the transport coefficients [16] Note that the Seebeck coefficient vanishes in a simple relax-
should be compared with the trap-averaged values, which differ ation time approximation given by φps = τs (w − 'p )vx ∂x Ts /T ,
slightly from the results presented here, but may readily be corresponding to the case b0 /b1 = w/kB T in Eq. (11).
computed using our Boltzmann formalism. [17] G. M. Bruun and H. Smith, Phys. Rev. A 72, 043605 (2005).

063613-4

You might also like