Mesoscopic Fermi gas in a harmonic trap
J. Schneider and H. Wallis
arXiv:cond-mat/9710086v2 [cond-mat.stat-mech] 20 Oct 1997
Max-Planck-Institut für Quantenoptik, Hans-Kopfermann-Straße 1, D-85748 Garching, Germany
(submitted to Phys. Rev. A, February 24, 2018)
investigate the case of an isotropic harmonic trap. We
analyze the influence of the shell structure on the chemical potential and the specific heat as a function of number
and temperature. For small particle number, density distributions deviating from the Thomas-Fermi distribution
are obtained. In Sec. III the anisotropic trap is considered with respect to chemical potential, specific heat and
density. The deviations from the isotropic trap are discussed. In Sec. IV the results are summarized in view of
experimental realizations, and the validity of the assumptions underlying our theoretical approach is discussed.
We study the thermodynamical properties of a mesoscopic
Fermi gas in view of recent possibilities to trap ultracold
atoms in a harmonic potential. We focus on the effects of
shell closure for finite small atom numbers. The dependence
of the chemical potential, the specific heat and the density
distribution on particle number and temperature is obtained.
Isotropic and anisotropic traps are compared. Possibilities of
experimental observations are discussed.
I. INTRODUCTION
II. PROPERTIES OF THE FERMI GAS IN AN
ISOTROPIC HARMONIC TRAP
The recent realizations of Bose-Einstein condensation
in dilute atomic vapors [1–3] have not only stimulated
many investigations on Bose atoms but also studies of
degenerate Fermi gases. As opposed to charged Fermi
gases, i.e. nucleons or electrons in solids, the effects of
Fermi statistics in neutral atomic gases occur at much
lower temperatures (typically below 10−7 K) and at densities which allow a treatment as dilute quantum gas because of the weak interatomic interactions. Due to the
selection rules for collisions, spin-polarized fermionic alkali atoms, like 6 Li or 40 K in magnetic traps, remain
metastable in the regime of quantum statistical degeneracy. Due to Fermi statistics the lowest scattering channel
(s-wave scattering) is closed for atoms in identical magnetic sublevels. Therefore an ultracold spin-polarized
Fermi gas will be less influenced by interactions than the
ultracold Bose gases [1–3]. Also a BCS transition as studied in [4,5] is naturally excluded here. The only remaining interaction is the magnetic dipole-dipole interaction
between the atoms. An estimate of its contribution to
the mean-field [6] yields hVDD i = (h̄ad /M )hr−3 i where
M is the atomic mass and ad < 10−10 m. This effect is
neglected here.
The purpose of the present paper is rather to consider
the stationary features of an ideal Fermi gas in isotropic
or anisotropic harmonic traps. Since our results are based
on the numerical calculation of the state sum without
further approximation, they are complementary to the
recent paper of Butts and Rokhsar [7], where a continuous spectrum and Thomas-Fermi approximation were
used. That treatment becomes exact in the limit of large
particle numbers. In contrast, we here focus on the effects of small particle numbers, where the shell structure
still affects the behavior of the many-particle system.
The outline of the paper is as follows. In Sec.II we
A. Degeneracies and the Fermi edge
We first study the isotropic harmonic trap with potential energy
V =
M ω2 2
x + y2 + z 2 ,
2
(2.1)
and frequency ω, because of its distinct features compared with the anisotropic trap studied below. In this
case the degeneracy of states with equal energy
Eν = (ν + 3/2)h̄ω
is given by
gν =
1
(ν + 1)(ν + 2),
2
(2.2)
i.e. equal to the number of simple partitions of ν as a
sum of three integers ν = νx + νy + νz . Since gν gives the
degeneracy of a shell of energy Eν , one finds the total
number Sα of quantum states with energy smaller than
Eα as the sum over the shells 0 ≤ ν ≤ α,
α
X
g ν = Sα .
(2.3)
ν=0
The sums
Sα =
1
(α + 1)(α + 2)(α + 3)
6
(2.4)
define the sequence {Sα } = {1, 4, 10, 20, 35, 56, ...} and so
forth. Note that each oscillator state is assumed to be
filled with a single Fermion since only one spin orientation
is confined by the magnetic trap.
1
For simplicity, our calculations are done using the
grand canonical ensemble [8]. The thermal occupation
of a state with energy Eν at a temperature kB T = 1/β
is given by Fermi-Dirac statistics as
nν =
z −1
1
,
exp(βh̄ων) + 1
for small temperatures. Whereas the solid lines correspond to the Thomas-Fermi approximation, the three
other curves were obtained numerically by truncating the
sum in Eq. (2.6) at sufficiently high ν. They display a
step-like variation that becomes increasingly smoother
for higher temperatures. The step function will appear
to be broken into smaller steps in the anisotropic oscillator case studied below. Here, the steps occur whenever
a shell is saturated and νF acquires the next higher integer value. µ converges to a certain (“plateau”) value
h̄ω(νF + 3/2) in the limit T → 0 for all N which do not
coincide with a “magic” number Sα . However, if a shell
is closed (N = SνF ), µ takes the value µ = h̄ω(νF + 2),
which is very close to the value of the Thomas-Fermi result at SνF (solid line in Fig. 1). As can be shown by
asymptotic expansion, the two curves intersect approximately at N = Sα respectively N = (Sα + Sα−1 )/2, i.e.
at total or half filling of shells.
This information is displayed in detail in Fig. 2 giving
the dependence of µ on the temperature around the value
N = S7 = 120. At T = 0, the N = 119 curve still
approaches the previous plateau value µ/h̄ω = 7 + 3/2,
whereas the N = 121 curve has to approach the value
µ/h̄ω = 8 + 3/2.
The temperature dependence of µ can be calculated
analytically in the limits of high and low temperature.
The high temperature region of Fig. 2 is well described
by the Sommerfeld-like formula [7]
2 !
π 2 kT
(2.12)
µ̃(T ) = ẼF 1 −
3 ẼF
(2.5)
where the fugacity z is determined from the condition
∞
X
gν nν = N.
(2.6)
ν=0
The definition of the fugacity z = exp β(µ − (3/2)h̄ω)
absorbs the zero-point energy. N is the total number of
particles in the trap. For a given particle number one
can determine the Fermi energy
EF = (νF + 3/2)h̄ω,
(2.7)
where νF is the shell up to which the trap levels are
filled with particles at temperature T = 0. In this limit
(β → ∞) the Fermi-Dirac distribution approaches a step
function. Note that in the case of a mesoscopic ensemble
the zero-temperature equation
∞
X
gν θ(νF − ν) = N
(2.8)
ν=0
does not have a solution for each N , but only for the
discrete set of total particle numbers N ∈ {Sα }. For
N 6∈ {Sα } we may still define νF = ⌈xF ⌉ as the smallest integer equal or greater than the exact solution of
Eq. (2.8),
xF = A +
1
− 2,
3A
where the factor π 2 /3 replaces the factor of π 2 /12 of the
usual case of fermions in a box. We note that for high
temperatures the exact result for N = 119 approaches
the Sommerfeld approximation for N = 120 (see Fig. 2).
In the low temperature regime, the variation of µ can
be analyzed in analogy to the chemical potential of electrons in an intrinsic semi-conductor. We first consider
the “magic” numbers N ∈ {Sα }. Let N> (T ) be the number of atoms excited to states above EF and N< (T ) the
number of unoccupied states (“holes”) at or below EF
(2.9)
where A is given by
1/3
p
.
A = 3N + 9N 2 − 1/27
(2.10)
The expression EF = h̄ω(⌈xF ⌉+3/2) has to be compared
to the Thomas-Fermi approximation for the Fermi edge,
1/3
ẼF = h̄ω(6N )
.
∞
X
gν
−1 exp(βh̄ων) + 1
z
ν=νF +1
νF
X
1
.
N< (T ) =
gν 1 − −1
z exp(βh̄ων) + 1
ν=0
N> (T ) =
(2.11)
B. Calculation of the chemical potential
(2.13)
(2.14)
For low temperatures, i.e. for
We now turn to the determination of the most important properties of the ideal gas. The fugacity resp.
the chemical potential are obtained by a numerical solution of Eq. (2.6) for given temperature and particle
number which is exact inasmuch as it does not invoke
the Thomas-Fermi approximation. The results are then
analyzed in certain limits below. Fig. 1 shows the dependence of the chemical potential µ on the atom number N
kB T ≪ EνF +1 − µ
and
kB T ≪ µ − EF ,
(2.15)
the number of “particles” and “holes” can be approximated as
N> (T ) ≈ Σ> e−β(EνF +1 −µ)
N< (T ) ≈ Σ< e
2
−β(µ−EF )
,
(2.16)
(2.17)
where
C. Specific heat
∞
X
Σ> =
Σ< =
gν e−β(Eν −EνF +1 )
ν=νF +1
νF
X
gν e−β(EF −Eν )
The discontinuity of the chemical potential manifests
itself most drastically in the specific heat of the gas. It
is calculated from the total energy
(2.18)
(2.19)
U (T ) =
ν=0
are essentially Boltzmann sums. On combination of the
above equations one arrives at
N> (T ) · N< (T ) = Σ> Σ< e−βh̄ω .
via
kT
C̃(T )
= π2
.
Nk
h̄ω(6N )1/3
(2.27)
whereas the classical high temperature limit equals
Ccl /N k = 3. Here, we determine C(T ) for finite N from
the state sum and compare it to the Sommerfeld approximation. The results are shown in Fig. 3, where the N 1/3
scaling is already included on the ordinate. In the limit
of ultra-low temperatures the finite-size effects result in
a deviation from the linear Sommerfeld prediction. For
higher temperature the specific heat C(T ) approaches the
Sommerfeld result (2.27).
At very low temperatures C(T ) remains zero instead of
increasing linearly. This is consistent with the assumptions leading to Eq. (2.23) which are confirmed by the calculation of the state sums. In the ultra-low temperature
regime no states above the Fermi energy are populated
due to the energy gap, the total energy does not increase
and C(T ) equals zero. This explanation seems to be correct also for the case of closed shells (N = 84, 120, 9880)
where Eq. (2.23) does not hold. Note that the range
where C(T ) remains zero is the same as the range of
validity of the linear approximation for µ(T ).
At intermediate temperatures a strong non-monotonic
N -dependence of the specific heat at constant T occurs,
roughly at those temperatures where the linear approximation ceases to be applicable. The origin of this behavior is revealed in Fig. 4. Each time a shell closure occurs,
C(N ) runs through a maximum. At these points, the
system can access a new, totally empty shell, at the expense of adding the gap energy to the new particles. On
the contrary, the minima occur half way between successive shell closures. In Fig. 5, the total heat capacity is
plotted versus N without a rescaling.
Fig. 3 shows another interesting detail: the two limiting curves for totally filled shells (N = 84, 120, 9880)
resp. half filled shells (N = 102, 142, 10270) do not
depend on νF . Up to kB T /h̄ω ≈ 0.5 the function
(6N )1/3 C/(N k) does not seem to depend on N explicitly
for the values considered here, but only on the relative
filling of the Fermi shell. This is related to the fact that
the dependence µ(T ) shown in Fig. 2 repeats itself around
each value for νF .
(2.22)
Assuming that ∆N is a constant for very low temperatures, one can solve Eq. (2.22) for the chemical potential
g
3
νF
µ(T ) = h̄ω(νF + ) − kT ln
−1 .
(2.23)
2
∆N
This expression varies linearly with T for non-vanishing
∆N , with its slope changing sign at ∆N = gνF /2. If the
highest shell is less than half filled (∆N < gνF /2), µ(T )
decreases linearly from µ(0) = EνF – if it is more than
half filled, it increases linearly from the µ(0) = EνF . The
exact result then approaches the Sommerfeld curve. The
range of validity ∆T of the linear approximation can be
roughly determined by equating
h̄ω
≡ k∆T ln(gF − 1)
2
(2.25)
C(T ) =
Thus µ(0) lies in the middle of the “gap” between EF
and EF + h̄ω, like in an intrinsic semi-conductor where
the valence band is filled at zero temperature and the
chemical potential lies in the middle of the bandgap. It
shows a slow linear decrease with increasing temperature,
governed by the small factor ln(Σ> /Σ< ).
If on the other hand N 6∈ {Sα }, one can calculate µ(T )
from the following approximation. For very low temperatures the Fermi function is well approximated by nν = 1
for ν < νF resp. nν = 0 for ν > νF . The number of
occupied states in the Fermi shell ∆N = N − SνF −1 then
reads approximately
gνF
.
z −1 exp(βh̄ωνF ) + 1
gν h̄ων
−1 exp(βh̄ων) + 1
z
ν=0
∂U (T )
.
(2.26)
∂T
The usual Sommerfeld approximation for low temperatures yields
(2.20)
From this condition the chemical potential can be determined. As for N ∈ {Sα } the Fermi shell is totally filled
at T = 0, N> (T ) must equal N< (T ) in that case, and
one obtains from Eq. (2.16–2.20) the low temperature
behavior
Σ>
kB T
.
(2.21)
ln
µ(T ) = h̄ω(νF + 2) −
2
Σ<
∆N =
∞
X
(2.24)
since the maximum deviation from the Sommerfeld approximation equals h̄ω/2 at ∆N = 1 (see Fig.2). For
νF = 7 one obtains this range as kB ∆T /h̄ω ≤ 0.14. For
larger values of ∆N the slope is smaller and the validity
range may be larger.
3
partitions of νr = νx + νy ). We use the notation (νr , νz )
for these states.
The N -dependence of quantities at zero temperature
turns out to have more features than in the isotropic
case. For example only in the oblate case (λ > 1) the
notion of shell closures still exists because only then it is
energetically favorable to fill up a transversal shell with
degeneracy gνr before populating a higher longitudinal
state. However, these new structures occur on a smaller
scale of particle numbers (due to the smaller degeneracy
factors) and might be less accessible in experiments with
a finite uncertainty of the atom number.
Formulas analogous to Eq. (2.4) can only be given as
sums. If λ < 1, one can count all states up to a certain
excitation (αr , αz ) by
D. Density distributions
Density distributions in traps can be measured quite
easily. In an isotropic oscillator, one expects radially
symmetric distributions. The radial wavefunctions unr ,l
(cf. e.g. [9]) are numbered by a radial quantum number
nr and angular momentum l. The corresponding energy
is Enr ,l = h̄ω(2(nr − 1) + l + 3/2) so ν = 2(nr − 1) + l. To
compute the total density one has to sum up the squared
wavefunctions weighted correctly with nν
ρ(r) =
∞
X
ν=0
[ν/2]+1
nν (T )
X 2l + 1
|unr ,l (r)|2 ,
4π
n =1
(2.28)
r
where l = ν − 2(nr − 1). The factor (2l + 1)/(4π) is due
to the summation over all states with m = −l, . . . , l. In
Fig. 6, ρ(r) is displayed for different particle numbers and
temperatures
p and scaled with the size of the trap groundstate σ = h̄/(M ω). The zero temperature result from
the Thomas-Fermi approximation (cf. [7], see Eq. (3.6),
λ = 1) is also shown.
For N = 120 (closed shell, νF = 7), one observes a
central minimum that disappears at N = 142 (half filled
shell, νF = 8). This shell gets totally filled at N = 165
where ρ(r) has a maximum at r = 0. The N -dependence
of the density at r = 0 is due to the fact that shells with
odd ν do not contribute to ρ(0) because they are made
up of odd angular momentum states which all have zero
density at the origin. The curves for kB T = 0.1h̄ω are
almost indistinguishable from the T = 0 curves. The
minima and maxima are still visible at kB T = 0.25h̄ω
but disappear for temperatures above kB T = h̄ω. For
not too high temperatures the density approaches the
Thomas-Fermi result. Interestingly, ρ(r) is almost equal
to this approximation for half filled Fermi shells.
αr +[αz λ]
Sαr ,αz =
νr =0
(3.3)
(3.4)
The Sommerfeld formula Eq. (2.12) for the chemical potential holds equally in the anisotropic case.
In the following, we give some of the results for
anisotropic traps. We concentrate on the case of a heavily
deformed cigar-shaped trap, λ ≪ 1 (in fact λ = 0.076 like
in [10]). We first consider the dependence of the chemical potential on the temperature in the low-temperature
regime. The graph of µ(T ) for 1000 particles displayed
in Fig. 7 shows a very intriguing feature: it starts linearly at T = 0 as predicted by Eq. (2.23) but then goes
through a local maximum. The highest occupied state
at T = 0 is (5, 22). So gνr = 6 and with Eq. (3.3)
∆N = S5,22 − N = 5. The next higher state is (6, 9).
The correction to the linear approximation including this
state basically shows that it is responsible for the local
maximum. Indeed, the energy difference between the two
states corresponds to kB T = 0.157h̄ωλ which is roughly
at the local minimum (cf. arrow in Fig. 7). Thus, at the
maximum of µ(T ) the next higher level above the Fermi
level becomes thermally accessible and µ decreases.
As in the isotropic case, the specific heat shows deviations from the Sommerfeld approximation at temperatures where the linear approximation for µ(T ) begins to
fail. If T is fixed to a value in that region, the graph
of (6N λ)1/3 C/(N k) as a function of particle number
(Fig. 8) exhibits structure on two scales of the particle
number, considerably more complex than the isotropic
analogue in Fig. 4. The big jumps take place whenever
there are enough particles to access a new shell of the
Experimentally realized magnetic traps are usually at
least slightly anisotropic. In this section we therefore
study a deformed oscillator with a potential
(3.1)
i.e. we allow for prolate and oblate ellipsoid iso-energy
surfaces. Accordingly, the energy eigenvalues are
1
Eνr ,νz = h̄ω(νr + 1 + λ(νz + )),
2
αr − νr
+ αz + 1
λ
ẼF = h̄ω(6N λ)1/3 .
A. Chemical potential and heat capacity
M ω2 2
x + y 2 + λ2 z 2 ,
2
gνr
where [x] denotes the largest integer less than or equal
to x. Thus, one needs N = Sαr ,αz particles to populate
all states up to Eαr ,αz . For λ > 1, there is an analogous
expression. In general, the exact Fermi energy can only
be found by searching the lowest state (νr , νz ) with N ≤
Sνr ,νz . However, the Thomas-Fermi approximation for
the Fermi energy is only slightly modified [7]
III. THE ANISOTROPIC CASE
V =
X
(3.2)
where νr and νz count radial and longitudinal excitations,
respectively. For given νr there are still gνr = νr + 1
degenerate states with different numbers of excitations
in the two degenerate transversal directions (number of
4
the density profiles is interchanged, the abovementioned
oscillations occur in the transversal density profile and
disappear in the longitudinal direction.
transversal oscillator. The arrows denote the values Sαr ,0
for αr = 5, 6, 7. Between two such particle numbers (say
Sαr ,0 , Sαr +1,0 ) there are 13 major peaks, corresponding
to 1/λ ≈ 13 longitudinal states being filled before the
next transversal shell can be reached. The finer substructure (see inset of Fig. 8) can also be explained easily: if one starts with the state (αr , 0) the next state is
(0, [αr /λ] + 1) followed by (1, [(αr − 1)/λ] + 1), etc. Consequently, there should be αr maxima before (αr , 1) is
reached. In case 1/λ is integer this substructure disappears.
IV. CONCLUSION
Our calculations have shown that quantum statistical
effects on the easily accessible observables of a trapped
Fermi gas are restricted to the regime of rather small
atom number, e.g. below N = 1000. For larger atom
numbers quantum statistical effects can be more easily
understood in terms of a local density approximation
[7,11]. Our work has been carried out using the grand
canonical ensemble. It is known from the ideal Bose gas
that the grand canonical and the canonical ensemble give
differing predictions [8]. However the deviations are often
smaller than the difference between the interacting and
the interaction-free case. Problems like artificial fluctuations in a grand canonical ensemble of bosons will not
occur for fermions. For that reason we don’t dwell on a
comparison of the different ensembles here.
The present study of an ideal Fermi gas showed some
remarkable effects of the shell structure in the harmonic
potential visible e.g. in Fig. 1. For a real Fermi gas the
atom-atom interactions will introduce an additional dependence of the chemical potential on the particle number which will smoothen out the steps in Fig. 1. A quantitative prediction of this effect depends on the relative
magnitude of atom-atom interactions and the experimentally controllable energy gap h̄ω. As mentioned in the introduction, an effective suppression of the interactions in
the ultracold regime can be attributed to the suppression
of s-wave scattering for Fermions in identical substates.
Therefore the ideal gas behavior might still be visible
provided the harmonic potential is steep enough.
BCS-like behavior as discussed in [4,5] requires two
spin states to be trapped. The effects of the harmonic
potential in that situation have been allowed for in local
density approximation in [5], i.e. for large atom numbers.
Because of the sensitivity of the BCS transition to the
difference of the atom numbers in both spin states, it
however remains a challenge to observe a BCS transition
experimentally. By contrast shell effects as discussed in
this paper should be visible in a suitable range of small
atom numbers and sufficiently large trap frequencies.
B. Density distributions
In order to calculate the density distribution we make
use of the transverse symmetry and obtain
ρ(r, z) = 2
∞
X
[νr /2]
nνr ,νz (T )
νr ,νz =0
X
|ũnr ,νr −2nr (r) · ψnz (z)|2 ,
nr =0
(3.5)
where ũnr ,νr −2nr (r) is the radial wavefunction of a 2D
harmonic oscillator (r2 = x2 + y 2 ) with magnetic quantum number |m| = νr − 2nr and ψnz (z) is the wavefunction of the one-dimensional harmonic oscillator. The
overall factor 2 allows for the twofold degeneracy of a
state with given νz , Nr , |m|. nνr ,νz (T ) is the Fermi-Dirac
occupation number analogous to Eq. (2.5). The numerical results should again be compared to the ThomasFermi approximation at T = 0 [7]
Nλ 8
ρ(r, z) = 3 2
RF π
3/2
r2 + λ2 z 2
1−
,
RF2
(3.6)
where RF = (48N λ)1/6 σ is the so called “Fermi radius”
of the density distribution.
We restrict ourself to a plot of ρ(r, 0) resp. ρ(0, z)
(cf. Fig. 9). In transversal direction the density at zero
temperature shows only very slight deviations from the
Thomas-Fermi result. In contrast, the density in longitudinal direction shown on the right exhibits oscillations around this approximation. These oscillations are
mainly due to a step-like behavior of the occupation number Nνz of the oscillator states in longitudinal direction,
which is due to the filling of new transversal states. The
steps have a width of 1/λ. In addition, Nνz decreases
much slower (average derivative ∝ λ) than the occupation number Nνx for the transversal states (average
derivative ∝ 1/λ, here we consider ρ(x, y = z = 0)). As
a result, the longitudinal density profile ρ(0, z) receives a
bigger contribution from higher states than the transversal one, so one can still see the various maxima of high
oscillator states in the first case but not in the latter.
We finally note that for oblate traps the effects are
basically the same as in prolate traps. The behavior of
ACKNOWLEDGMENTS
We appreciate stimulating discussions with A. Schenzle and C. Zimmermann. H.W. acknowledges financial
support by the DFG under Grant Nr. Wa 727/6.
5
[1] M.H. Anderson, J.R. Ensher, M.R. Matthews, C.E. Wieman, and E.A. Cornell, Science 269, 198 (1995).
[2] C.C. Bradley, C.A. Sackett, J.J. Tollet, and R.G. Hulet,
Phys. Rev. Lett. 75 1687 (1995).
[3] K.B.
Davis,
M.-O.
Mewes,
M.R.
Andrews, N.J. van Druten, D.S. Durfee, D.M. Kurn, and
W. Ketterle, Phys. Rev. Lett. 75, 3969 (1995).
[4] H.T.C. Stoof, M. Houbiers, C.A. Sackett, and
R.G. Hulet, Phys. Rev. Lett. 76 10 (1995).
[5] M. Houbiers, R. Ferwerda, H.T.C. Stoof, W.I. McAlexander, C.A. Sackett, and R.G. Hulet, Report No. condmat/9707006.
[6] H.T.C. Stoof, communication at the ESF-Workshop on
Bose-Einstein-Condensation, Castelvecchio Pascoli, July
1997.
[7] D.A. Butts and D.S. Rokhsar, Phys. Rev. A 55, 4346
(1997).
[8] The difference between different statistical ensembles for bosons are investigated in H.D. Politzer,
Phys. Rev. A 54, 5048 (1996), C. Herzog and M. Olshanii, Phys. Rev. A 55, 3254 (1997) and C. Weiss and
M. Wilkens, J. Modern Optics (1997), in press.
[9] J.P. Blaizot and G. Ripka, Quantum Theory of Finite
Systems, (MIT Press, Cambridge, 1986), chapt. 6
[10] M.-O. Mewes, M.R. Andrews, N.J. van Druten,
D.M. Kurn, D.S. Durfee, C.G. Townsend, and W. Ketterle, Phys. Rev. Lett. 77 988 (1996).
[11] J. Oliva, Phys. Rev. B 39, 4204 (1989).
9.5
h !
9.25
9
8.75
8.5
8.25
0
0.2
0.4
0.6
kT
h !
0.8
1.2
1
FIG. 2. This figure shows µ(T ) for the isotropic trap. The
Sommerfeld approximation is for N = S7 = 120 but agrees
very well with the numerical curve for N = 119. This occurs
also for other values of N .
8
1=3
C=(N k )
7
(6N )
11
10
9
8
h ! 7
6
5
4
3
2
1
0
N = 119
N = 120
N = 121
linear approx.
Sommmerfeld
6
5
4
N = 84, 120, 9880
3
N = 102, 142, 10270
2
N = 90
1
0
0
N )1=3
(6
50
100
N
150
0.2
0.3
0.4
kT
h !
0.5
0.6
0.7
0.8
FIG. 3. C(T ) appropriately scaled. The linear curve is the
Sommerfeld result Eq. (2.27). The deviation at large kB T /h̄ω
from the linear behavior occurs only for small N , because for
high temperatures (6N )1/3 C(T )/(N k) → 3(6N )1/3 due to
the equipartition theorem.
kT
h ! = 0:002
kT = 0:04
h !
kT = 0:08
h !
0
0.1
200
FIG. 1. µ(N ) shows a step-like behavior following the continuous approximation in Eq. (2.11). The dashed and dotted
curves are displaced vertically by −1 resp. −2.
(6N )
1=3
C=(N k )
5
= 0:45
4
= 0:3
3
= 0:15
2
1
0
6
0
50
100
150
200
N
250
300
350
400
101.51
FIG. 4. Specific heat as a function of the particle number at different temperatures and scaled with (6N )1/3 .
τ = kB T /h̄ω denotes the temperature in units of the level
spacing. The arrows point to N = Sν for ν = 1, . . . , 11
(N = 4, 10, 20, 35, 56, 84, 120, 165, 220, 286, 364).
101.5
h ! 101.49
140
101.48
101.5
101.47
101.48
101.46
101.46
101.45
101.44
C=k
120
100
= 0:45
101.44
= 0:3
80
20
0
50
100
150
200
N
250
300
350
C=(Nk )(6N)1=3
1.6
3
N = 165
1
0.8
N = 142
N = 120
0.6
0.6
0.8
1
1.2
1.4
0.1
1.8
1.2
0.4
400
FIG. 5. Total specific heat as a function of the particle
number at different temperatures. Again, τ = kB T /h̄ω.
1.4
0.2
0.2
FIG. 7. µ(T ) for a cigar-shaped trap with λ = 0.076 and
N = 1000. The arrow denotes the energy difference between
the Fermi level and the next higher state. The inset displays
the linear and next higher approximation taking into account
the states directly below and above the Fermi level.
40
0
0
0.1
kT
h !
= 0:15
60
0
0.08
760
0.06
780
800
820
840
0.04
0.02
0.4
0.2
0
0
1
2
3
r=
4
0
400
5
500
600
700
800
N
900 1000 1100 1200
FIG. 8. Specific heat for a cigar-shaped trap as a function of N at kB T /h̄ω = 0.044. The arrows point to
Sαr ,0 (αr = 5, 6, 7). The inset shows the number range
750 < N < 850 in more detail.
FIG. 6. Spatial density ρ in a isotropic trap as a function
of the distance to the trap center forpdifferent numbers of
fermions. The scaling parameters σ = h̄/(M ω) is the width
of the ground state. The unbroken lines denote kB T = 0.1h̄ω,
the dotted kB T = 0.25h̄ω and the dashed ones are obtained
from the Thomas-Fermi approximation at T = 0.
7
N = 1000
3
1
= 0:076
0.8
0.6
0.4
0.2
0
5
4
3
2
r=
1
0
10 20 30 40 50
z=
FIG. 9. Density ρ for a cigar-shaped trap. The left part
shows the density in transversal direction for z = 0, the graph
on the right is the density in longitudinal direction on the
symmetry axis (x = y = 0). The unbroken lines are numerical results for T = 0, the dashed lines come from the
Thomas-Fermi approximation. Note that the “Fermi radius”
in longitudinal direction is about 1/λ ≈ 13 times larger than
the transversal one.
8