0801.2500/chap3_quantanalysis.tex
1: \section{Quantitative analysis of density distributions}
2: \label{c:analysis}
3: 
4: The purpose of imaging and image processing is to record density
5: distributions of the atomic cloud, either trapped or during
6: ballistic expansion.  All our knowledge about the properties of
7: cold atom systems  comes from the analysis of such images.  They
8: are usually compared to the results of models of the atomic gas.
9: Some models are exact (for the ideal gas), others are
10: phenomenological or approximations.  Many important models for
11: bosonic atoms have been presented in our 1999 Varenna notes. Here
12: we discuss important models for fermions, which allow us to infer
13: properties of the system from recorded (column) density
14: distributions.
15: 
16: 
17: \subsection{Trapped atomic gases}
18: 
19: \subsubsection{Ideal Bose and Fermi gases in a harmonic trap}
20: 
21: The particles in an atom trap are isolated from the surroundings,
22: thus the atom number $N$ and total energy content $E_{\rm tot}$ of
23: the atomic cloud is fixed. However, it is convenient to consider
24: the system to be in contact with a reservoir, with which it can
25: exchange particles and energy (grand canonical ensemble). For
26: non-interacting particles with single-particle energies $E_i$, the
27: average occupation of state $i$ is
28: \begin{eqnarray}
29:     \left<n_i\right> =  \frac{1}{e^{(E_i - \mu)/k_B T} \mp 1}
30:     \label{e:BoseFermidist}
31: \end{eqnarray}
32: with the upper sign for bosons, the lower sign for fermions. These
33: are the Bose-Einstein and Fermi-Dirac distributions, respectively.
34: For a fixed number of particles $N$ one chooses the chemical potential $\mu$ such that $N
35: = \left<N\right> = \sum_i \left<n_i\right>$.
36: 
37: Let us now apply these distributions to particles confined in a
38: harmonic trap, with trapping potential
39: \begin{equation}
40: V(\vect{r}) = \frac{1}{2} m (\omega_x^2 x^2 + \omega_y^2 y^2 +
41: \omega_z^2 z^2) \label{e:potential}
42: \end{equation}
43: We assume that the thermal energy $k_B T \equiv 1/\beta$ is much
44: larger than the quantum mechanical level spacings $\hbar
45: \omega_{x,y,z}$ (Thomas-Fermi approximation). In this case, the
46: occupation of a phase space cell
47: $\left\{\vect{r},\vect{p}\right\}$ (which is the phase-space
48: density times $h^3$) is given by Eq.~\ref{e:BoseFermidist}
49: \begin{eqnarray}
50: f(\vect{r},\vect{p}) = \frac{1}{e^{(\frac{\vect{p}^2}{2m} +
51: V(\vect{r}) - \mu)/k_B T} \mp 1} \label{e:FermiBose}
52: \end{eqnarray}
53: The density distribution of the thermal gas is
54: \begin{eqnarray}
55: n_{th}(\vect{r}) &=& \Intp{p}\; f(\vect{r},\vect{p})\nonumber\\
56: &=& \pm \frac{1}{\lambda_{ dB}^3}\, {\rm Li}_{3/2}\left(\pm
57: e^{\beta\left(\mu - V(\vect{r})\right)}\right) \label{e:density}
58: \end{eqnarray}
59: where $\sqrt{\frac{2\pi \hbar^2}{m k_B T}}$ is the de Broglie
60: wavelength. ${\rm Li}_n(z)$ is the $n^{th}$-order
61: Polylogarithm, defined as
62: \begin{eqnarray}
63:     {\rm Li}_n(z)\; \equiv\; \frac{1}{\pi^n} \int {\rm d}^{2n}r \frac{1}{e^{\vect{r}^2}/z - 1}\; \stackrel{n\ne 0}{=}\; \frac{1}{\Gamma(n)}\int_0^\infty {\rm d}q \frac{q^{n-1}}{e^q/z - 1}
64: \label{e:polylog}
65: \end{eqnarray}
66: where the first integral is over $2n$ dimensions, $\vect{r}$ is
67: the radius vector in $2n$ dimensions, $n$ is any positive
68: half-integer or zero and $\Gamma(n)$ is the Gamma-function
69: \footnote{The Polylogarithm appears naturally in integrals over
70: Bose-Einstein or Fermi-Dirac distributions.  Some
71: authors~\cite{huan87} use different functions for bosons
72: $g_n(z)={\rm Li}_n(z)$ and for fermions
73: $f_n(z)=-{\rm Li}_n(-z)$.  The Polylogarithm can be
74: expressed as a sum ${\rm Li}_n(z) = \sum_{k=1}^\infty
75: \frac{z^k}{k^n}$ which is often used as the definition of
76: the Polylogarithm.  This expression is valid for all complex
77: numbers $n$ and $z$ where $|z|\le 1$. The definition
78: given in the text is valid for all $z\le l$.
79: 
80: Special cases: ${\rm Li}_0(z) = \frac{1}{1/z - 1}$,
81: ${\rm Li_1}(z) = -\ln(1-z)$. $f(\vect{r},\vect{p})$ can
82: be written as $\pm{\rm Li}_0(\pm
83: \exp[\beta(\mu-\frac{\vect{p}^2}{2m} - V(\vect{r}))])$. When integrating density distributions to obtain column densities, a useful formula is:
84: \begin{equation}
85:     \int_{-\infty}^\infty dx \;{\rm Li}_n(z\,e^{- x^2}) = \sqrt{\pi}\; {\rm Li}_{n+1/2}(z)
86:     \label{e:polyintegral}.
87: \end{equation}
88:  Limiting values: ${\rm Li}_n(z) \stackrel{z \ll 1}{\rightarrow} z$ and $-{\rm Li}_n(-z) \stackrel{z\rightarrow\infty}{\rightarrow} \frac{1}{\Gamma(n+1)}\; \ln^n(z)$.}.
89: Note that expression~\ref{e:density} is correct for any potential
90: $V(\vect{r})$. The constraint on the number of thermal particles
91: is
92: \begin{equation}
93: N_{th} = \Int{r} \; n_{th}(\vect{r})
94: \end{equation}
95: For a harmonic potential (~\ref{e:potential}), we obtain
96: \begin{equation}
97: N_{th} = \pm \left(\frac{k_B T}{\hbar \bar{\omega}}\right)^3 {\rm
98: Li}_3(\pm\,e^{\beta\mu}) \label{e:numberofatoms}
99: \end{equation}
100: with $\bar{\omega} = (\omega_x \omega_y \omega_z)^{1/3}$ the
101: geometric mean of the trapping frequencies.
102: 
103: In the classical limit at high temperature, we recover the
104: Maxwell-Boltzmann result of a gaussian distribution,
105: \begin{equation}
106:     n_{cl}(\vect{r}) = \frac{N}{\pi^{3/2} \sigma_x \sigma_y \sigma_z} e^{- \sum_i x_i^2/\sigma_{x_i}^2} \qquad {\rm with} \; \sigma_{x,y,z}^2 = \frac{2 k_B T}{m \omega_{x,y,z}^2}
107: \end{equation}
108: 
109: The regime of quantum degeneracy is reached when $\lambda_{dB}
110: \approx n ^{-1/3}$, or when the temperature $T \approx T_{\rm
111: deg}$.  The  degeneracy temperature $T_{\rm deg} =
112: \frac{\hbar^2}{2m k_B} n^{2/3}$ is around or below one $\mu \rm K$
113: for typical experimental conditions.
114: 
115: For {\bf bosons}, it is at this point that the ground state
116: becomes macroscopically occupied and the condensate forms. The
117: density profile of the ideal gas condensate is given by the square
118: of the harmonic oscillator ground state wave function:
119: \begin{equation}
120: n_c(\vect{r}) = \frac{N_0}{\pi^{3/2} d_x d_y d_z} e^{-\sum_i
121: x_i^2/d_{x_i}^2}
122: \end{equation}
123: where $d_{x_i} = \sqrt{\frac{\hbar}{m \omega_{x_i}}}$ are the
124: harmonic oscillator lengths. The density profile of the thermal,
125: non-condensed component can be obtained from Eq.~\ref{e:density}
126: if the chemical potential $\mu$ is known. As the number of
127: condensed bosons $N_0$ grows to be significantly larger than 1,
128: the chemical potential $\mu \approx - \frac{k_B T}{N_0}$ (from
129: Eq.~\ref{e:BoseFermidist} for $E_0 = 0$) will be much closer to
130: the ground state energy than the first excited harmonic oscillator
131: state. Thus we set $\mu = 0$ in the expression for the
132: non-condensed density $n_{th}$ and number $N_{th}$ and obtain
133: \begin{eqnarray}
134: n_{th}(\vect{r}) &=& \frac{1}{\lambda_{dB}^3} {\rm Li}_{3/2}(e^{-V(\vect{r})/k_B T})\\
135: N_{th} &=& N (T/T_C)^3\qquad \mbox{for $T<T_C$}
136: \end{eqnarray}
137: with the critical temperature for Bose-Einstein condensation in a
138: harmonic trap
139: \begin{equation}
140: T_C \equiv \hbar \bar{\omega}\; (N / \zeta(3))^{1/3} = 0.94 \;
141: \hbar \bar{\omega} N^{1/3}
142: \end{equation}
143: where $\zeta(3) = {\rm Li}_3(1) \approx 1.202$. At $T=T_C$, the
144: condition for Bose condensation is fulfilled in the center of the
145: trap, $n = {\rm Li}_{3/2}(1)/\lambda_{dB}^3 =
146: 2.612/\lambda_{dB}^3$. For lower temperatures, the maximum density
147: of the thermal cloud is ``quantum saturated'' at the critical value
148: $n_{th} = 2.612/\lambda_{dB}^3 \propto T^{3/2}$. The condensate
149: fraction in a harmonic trap is given by
150: \begin{equation}
151: N_0/N = 1 - (T/T_C)^3
152: \end{equation}
153: For $T/T_C = 0.5$ the condensate fraction is already about 90\%.
154: 
155: For {\bf fermions}, the occupation of available phase space cells
156: smoothly approaches unity without any sudden transition:
157: \begin{equation}
158:   f(\vect{r},\vect{p}) = \frac{1}{e^{(\frac{\vect{p}^2}{2m} + V(\vect{r}) - \mu)/k_B T} + 1} \stackrel{T \rightarrow 0} \rightarrow \left\{%
159: \begin{array}{ll}
160:     1, & \hbox{$\frac{\vect{p}^2}{2m} + V(\vect{r}) < \mu$} \\
161:     0, & \hbox{$\frac{\vect{p}^2}{2m} + V(\vect{r}) > \mu$} \\
162: \end{array}%
163: \right. \label{e:fermiphasespace}
164: \end{equation}
165: Accordingly, also the density profile changes smoothly from its
166: gaussian form at high temperatures to its zero temperature shape:
167: \begin{eqnarray}
168: n_F(\vect{r}) &=& \Intp{p} \, f(\vect{r},\vect{p}) \stackrel{T\rightarrow 0}{\rightarrow} \int_{\left|\vect{p}\right|< \sqrt{2m(\mu-V(\vect{r}))}} \frac{{\rm d}^3\vect{p}}{(2\pi\hbar)^3}\nonumber\\
169: &=& \frac{1}{6\pi^2} \left(\frac{2m}{\hbar^2}\right)^{3/2}
170: \left(\mu - V(\vect{r})\right)^{3/2}.
171: \end{eqnarray}
172: 
173: From Eq.~\ref{e:fermiphasespace} we observe that at zero
174: temperature, $\mu$ is the energy of the highest occupied state of
175: the non-interacting Fermi gas, also called the Fermi energy $E_F$.
176: The (globally) largest momentum is $p_F \equiv \hbar k_F \equiv
177: \sqrt{2 m E_F}$, the Fermi momentum. {\it Locally}, at position
178: $\vect{r}$ in the trap, it is $p_F(\vect{r}) \equiv \hbar
179: k_F(\vect{r}) \equiv \sqrt{2 m \epsilon_F(\vect{r})} \equiv \hbar
180: (6\pi^2 n_F(\vect{r}))^{1/3}$ with the local Fermi energy
181: $\epsilon_F(\vect{r})$ which equals $\mu(\vect{r},T=0) = E_F -
182: V(\vect{r})$. The value of $E_F$ is fixed by the number of
183: fermions $N$, occupying the $N$ lowest energy states of the trap.
184: For a harmonic trap we obtain
185: \begin{eqnarray}
186:     N &=& \Int{r} \; n_F(\vect{r}) = \frac{1}{6} \left(\frac{E_F}{\hbar \bar{\omega}}\right)^3\nonumber\\
187: \Rightarrow    E_F &=& \hbar \bar{\omega} (6 N)^{1/3}
188:     \label{e:Ferminumber}
189: \end{eqnarray}
190: and for the zero-temperature profile
191: \begin{eqnarray}
192: n_F(\vect{r}) &=& \frac{8}{\pi^2} \frac{N}{R_{Fx} R_{Fy} R_{Fz}}
193: \; \left[\max \left(1 - \sum_i
194: \frac{x_i^2}{R_{Fi}^2},0\right)\right]^{3/2}
195: \label{e:Fermidensity}
196: \end{eqnarray}
197: with the Fermi radii $R_{F{x,y,z}} = \sqrt{\frac{2 E_F}{m
198: \omega_{x,y,z}^2}}$. The profile of the degenerate Fermi gas has a
199: rather flat top compared to the gaussian profile of a thermal
200: cloud, as the occupancy of available phase space cells saturates
201: at unity.
202: 
203: At finite $T \lesssim T_F$, we can understand the shape of the
204: cloud by comparing $k_B T$ with the local Fermi energy
205: $\epsilon_F(\vect{r})$. For the outer regions in the trap where $k_B
206: T \gg \epsilon_F(\vect{r})$, the gas shows a classical (Boltzmann)
207: density distribution $n(\vect{r}) \propto e^{-\beta V(\vect{r})}$. In
208: the inner part of the cloud where $k_B T \ll
209: \epsilon_F(\vect{r})$, the density is of the zero-temperature form $n(\vect{r})
210: \propto (E_F - V(\vect{r}))^{3/2}$. The Polylogarithm smoothly
211: interpolates between the two regimes. We notice here the
212: difficulty of thermometry for very cold Fermi clouds: Temperature
213: only affects the far wings of the density distribution. While for
214: thermal clouds above $T_F$, the size of the cloud is a direct
215: measure of temperature, for cold Fermi clouds one needs to extract
216: the temperature from the shape of the distribution's wings.
217: 
218: Note that the validity of the above derivation required the Fermi
219: energy $E_F$ to be much larger than the level spacing $\hbar
220: \omega_{x,y,z}$. For example, in very elongated traps and for low
221: atom numbers one can have a situation where this condition is
222: violated in the tightly confining radial dimensions.
223: 
224: \subsubsection{Trapped, interacting Fermi mixtures at zero temperature}
225: \label{s:trappedmixtures}
226: \begin{figure}
227:     \centering
228:     \includegraphics[width=4in]{figs_quantanalysis/phasediagram.eps}
229:     \caption{Phase diagram of interacting Fermi mixtures in a
230:     harmonic trap, as a function of temperature and interaction
231:     strength $1/k_F a$. Shown is the critical temperature $T_C$
232:     for the formation of a superfluid as a function of $1/k_F a$
233:     (full line) as well as the characteristic temperature $T^*$ at
234:     which fermion pairs start to form (dashed line),
235:     after~\cite{pera04temp}. The shading indicates that pair
236:     formation is a smooth process, not a phase transition.}
237:     \label{f3:phasediagram}
238: \end{figure}
239: 
240: We now consider the case of $N$ fermionic atoms equally populating
241: two hyperfine states (``spin up'' and ``spin down''). Atoms in
242: different spin states interact via $s$-wave collisions
243: characterized by the scattering length $a$. A dimensionless
244: parameter measuring the strength and sign of the interaction
245: strength is $1/k_F a$, essentially the ratio of the interparticle
246: spacing to the scattering length. For weak attractive
247: interactions, $1/k_F a \rightarrow -\infty$, the ground state of
248: the system is a BCS superfluid (see chapter~\ref{c:BECBCStheory}).
249: As the magnitude of the scattering length increases to a point
250: where $a \rightarrow \mp \infty$ diverges (thus $1/k_F a
251: \rightarrow 0$), a  two-body molecular bound state enters the
252: interparticle potential. For weak repulsive interactions, $1/k_F a
253: \rightarrow +\infty$, the ground state of the system is then a
254: Bose-Einstein condensate of weakly-interacting molecules of mass
255: $M = 2m$, in which two fermions of opposite spin are tightly
256: bound.
257: 
258: Fig.~\ref{f3:phasediagram} summarizes the different regimes within
259: this BEC-BCS crossover. We see that the character of the
260: Fermi mixture drastically changes as a function of temperature and
261: interaction strength. For temperatures $T \gg T^*$ fermions are
262: unpaired, and a free Fermi mixture exists on the BEC- and the
263: BCS-side of the phase diagram. On resonance, the mixture might
264: still be strongly interacting even at high temperatures, thus
265: possibly requiring an effective mass description of the
266: interacting gas. The density distribution will have the same shape
267: as a free Fermi gas at all interaction strengths. Below $T^*$,
268: fermion pairs start to form. On the BEC-side, where fermions are
269: tightly bound, the thermal distribution should now be that of a
270: gas of bosons with mass $M = 2m$. As a consequence, the cloud will
271: shrink. Below $T_C$, we will finally observe a superfluid,
272: condensed core, surrounded by a thermal cloud of molecules in the
273: BEC-limit, or of unpaired fermions in the BCS-limit.
274: 
275: In general, the calculation of density
276: distributions in the strongly interacting regime is a difficult
277: affair. Simple expressions for the densities can be derived for
278: superfluid gases at zero temperature, for molecular gases on the
279: ``BEC''-side at large and positive $1/k_F a$, for weakly interacting
280: Fermi gases on the ``BCS''-side for large and negative $1/k_F a$,
281: and in the classical limit at high temperatures.
282: 
283: \paragraph{BEC limit}
284: 
285: The molecular Bose-Einstein condensate is described by a many-body
286: wave function $\psi(\vect{r})$ which obeys the {\rm
287: Gross-Pitaevskii equation}~\cite{peth02bec}
288: \begin{equation}
289: \left(-\frac{\hbar^2 \nabla^2}{2M} + V_M(\vect{r}) + g
290: \left|\psi(\vect{r},t)\right|^2\right) \psi(\vect{r},t) = i \hbar
291: \frac{\partial}{\partial t} \psi(\vect{r},t)
292: \label{e:grosspitaevskii}
293: \end{equation}
294: where $V_M(\vect{r})$ is the trapping potential experienced by the
295: molecules, and $g = \frac{4\pi \hbar^2 a_M}{M}$ describes the
296: intermolecular interactions. We can identify $\left|\psi\right|^2$
297: with the condensate density $n_c$, which for weak interactions and
298: at zero temperature equals the density of molecules $n_M$. The
299: validity of Eq.~\ref{e:grosspitaevskii} is limited to {\it weakly
300: interacting} gases of molecules, for which the gas parameter $n_M
301: a_M^3 \approx (\frac{k_F a}{6.5})^3 \ll 1$. In typical experiments
302: on BECs of bosonic atoms, the corresponding condition is very well
303: fulfilled. For a sodium BEC with $n \approx 10^{14} \, \rm
304: cm^{-3}$ and $a = 3.3\,\rm nm$, we have $n a^3 \approx 4 \cdot
305: 10^{-6}$. However, for molecular condensates near a Feshbach
306: resonance,  this condition can be easily violated (see
307: chapter~\ref{c:expobservation}).
308: 
309: In equilibrium, the ground-state wave function is $\psi(\vect{r},t)
310: =  e^{-i \mu_M t/\hbar} \psi(\vect{r})$, where $\mu_M$ is the
311: ground state energy and is identified with the molecular chemical
312: potential, and $\psi(\vect{r})$ is a solution of the stationary
313: Gross-Pitaevskii equation
314: \begin{equation}
315: \left(-\frac{\hbar^2 \nabla^2}{2M} + V_M(\vect{r}) + g
316: \left|\psi(\vect{r})\right|^2\right) \psi(\vect{r}) = \mu_M
317: \psi(\vect{r}) \label{e:statgrosspitaevskii}
318: \end{equation}
319: In the {\it ideal gas limit}, $g n_c \ll \hbar \omega_{x, y, z}$,
320: we recover the harmonic oscillator result for the condensate's
321: density distribution $n_c(\vect{r})$. In the {\it Thomas-Fermi
322: limit}, on the other hand, interactions dominate over the kinetic
323: energy of the condensate wave function, $g n_c \gg \hbar
324: \omega_{x,y,z}$. Already for weakly interacting alkali gases, this
325: condition is very well fulfilled, with typical interaction
326: energies of $g n_c \sim k_B \times 150 \,\rm nK$ and
327: $\hbar\omega_r \approx k_B \times 5 \,\rm nK$. In this
328: approximation we obtain the condensate density $n_c(\vect{r}) =
329: \left|\psi(\vect{r})\right|^2$:
330: \begin{equation}
331:     n_c(\vect{r}) = \max \left(\frac{\mu_M - V_M(\vect{r})}{g},0\right)
332: \end{equation}
333: Thus, a condensate in the Thomas-Fermi approximation ``fills in''
334: the bottom of the trapping potential up to an energy $\mu_M$,
335: which is determined by the total number of molecules, $N_M = N/2 =
336: \Int{r} \; n_c(\vect{r})$. Taking $V_M(\vect{r}) = 2 V(\vect{r})$
337: with the harmonic trapping potential for single atoms in
338: Eq.~\ref{e:potential}, one obtains a parabolic density profile,
339: \begin{equation}
340:     n_c(\vect{r}) = \frac{15}{8\pi} \frac{N_M}{R_x R_y R_z} \max \left(1 - \sum_i \frac{x_i^2}{R_i^2},0\right)
341:     \label{e:BECparabola}
342: \end{equation}
343: where the {\it Thomas-Fermi radii} $R_i = \sqrt{\frac{2 \mu_M}{M
344: \omega_i^2}}$ give the half-lengths of the trapped condensate
345: where the density vanishes. The chemical potential is given by
346: \begin{equation}
347:     \mu_M = \frac{1}{2} \hbar \bar{\omega} \left(\frac{15 N_M a_M}{\bar{d}_{h.o.}}\right)^{2/5}
348:     \label{e:chempot}
349: \end{equation}
350: where $\bar{d}_{h.o.} = (d_x d_y d_z)^{1/3} = \sqrt{\hbar/M
351: \bar{\omega}}$ is the geometric mean of the harmonic oscillator
352: lengths for molecules.
353: 
354: Interactions thus have a major effect on the shape of the
355: Bose-Einstein condensate, changing the density profile from the
356: gaussian harmonic oscillator ground state wave function to a broad
357: parabola, as a result of the interparticle repulsion. The
358: characteristic size of the condensate is no longer given by the
359: harmonic oscillator length but by the generally much larger
360: Thomas-Fermi radius $R_{x,y,z} = d_{x,y,z}
361: \sqrt{\frac{\bar{\omega}}{\omega_{x,y,z}}} \left(\frac{15 N_M
362: a_M}{\bar{d}_{h.o.}}\right)^{1/5}$. Also the aspect ratio changes,
363: for example in the $x$-$y$ plane from $\frac{d_x}{d_y} =
364: \sqrt{\frac{\omega_y}{\omega_x}}$ to $\frac{R_x}{R_y} =
365: \frac{\omega_y}{\omega_x}$. Nevertheless, weakly interacting
366: condensates are still considerably smaller in size than a thermal
367: cloud at $k_B T > \mu_M$, and more dense. This leads to the clear
368: separation between the dense condensate in the center of the cloud
369: and the large surrounding thermal cloud, the ``smoking gun'' for
370: Bose-Einstein condensation (both in the trapped and in the
371: expanding cloud, see section~\ref{s:expansion} below). In the case
372: of strong interactions, when the chemical potential $\mu_M$
373: becomes comparable to $k_B T_C$, this direct signature of
374: condensation will be considerably weaker. In this regime we also
375: have to account for the mutual repulsion between the thermal cloud
376: and the condensate (see section~\ref{s:molecularclouds} below).
377: 
378: \paragraph{BCS limit}
379: 
380: In the weakly interacting BCS limit ($1/k_F a \rightarrow
381: -\infty$), pairing of fermions and superfluidity have very small
382: effects on the density profile of the gas. The sharp Fermi surface
383: in $k$-space at $k_F$ is modified only in an exponentially narrow
384: region of width $\sim k_F \exp(-\frac{\pi}{2 k_F |a|})$. The
385: density, i.e. the integral over occupied $k$-states, is thus
386: essentially identical to that of a non-interacting Fermi gas. The
387: result is Eq.~\ref{e:Fermidensity} with the number of spin-up
388: (spin-down) atoms $N_{\uparrow,\downarrow} = N/2$ and Fermi energy
389: $E_{F} = \hbar \bar{\omega} (6
390: N_{\uparrow,\downarrow})^{1/3} = \hbar \bar{\omega} (3 N)^{1/3}$.
391: As one approaches the strongly interacting regime $1/k_F a \approx
392: -1$, it is conceivable that the formation of the superfluid leaves
393: a distinct trace in the density profile of the gas, as this is the
394: situation in the BEC-limit, and the crossover between the two
395: regimes is smooth. Indeed, several theoretical studies have
396: predicted kinks in the density profiles signalling the onset of
397: superfluidity~\cite{holl01,pera04temp,ho04uni,staj05dens}. We were able to observe
398: such a direct signature of condensation on resonance ($1/k_F a =
399: 0$) and on the BCS-side ($1/k_F a < 0$) in unequal Fermi mixtures
400: (see section~\ref{s:imbalance}). In equal mixtures, we detected a
401: faint but distinct deviation from the Thomas-Fermi profile on
402: resonance (see~\ref{s:directunitarity}).
403: 
404: \paragraph{Unitarity}
405: \label{s:unitarity}
406: 
407: The regime on resonance ($1/k_F a = 0$) deserves special
408: attention. The scattering length diverges and leaves the
409: interparticle distance $n^{-1/3} \sim 1/k_F$ as the only relevant
410: length scale. Correspondingly, the only relevant energy scale is
411: the Fermi energy $\epsilon_F = \hbar^2 k_F^2 / 2 m$. The regime is
412: thus said to be universal. The chemical potential $\mu$ can then
413: be written as a universal constant times the Fermi energy: $\mu =
414: \xi \epsilon_F$. In the trapped case, we can use this relation
415: locally  (local density approximation) and relate the local
416: chemical potential $\mu(\vect{r}) = \mu - V(\vect{r})$ to the
417: local Fermi energy $\epsilon_F(\vect{r}) \equiv \hbar^2
418: k_F(\vect{r})^2 / 2 m \equiv \frac{\hbar^2}{2 m} (6\pi^2
419: n_{\uparrow}(\vect{r}))^{2/3}$, where $n_{\uparrow}(\vect{r})$ is the density of atoms in one spin state. We then directly obtain a relation for the density profile $n_{\uparrow U}(\vect{r})$ of the unitary Fermi gas:
420: $$n_{\uparrow U}(\vect{r}) = \frac{1}{6\pi^2} \left(\frac{2m}{\xi \hbar^2}\right)^{3/2} \left(\mu - V(\vect{r})\right)^{3/2}.$$
421: The constraint from the number of particles in spin up, $N_\uparrow = N/2$, determines $\mu$:
422: \begin{eqnarray}
423:     N_\uparrow &=& \Int{r} \; n_{\uparrow U}(\vect{r}) = \frac{1}{6} \left(\frac{\mu}{\sqrt{\xi} \hbar \bar{\omega}}\right)^3\nonumber\\
424: \Rightarrow    \mu &=& \sqrt{\xi} E_F.
425: \end{eqnarray}
426: The density profile becomes
427: \begin{eqnarray}
428: n_{\uparrow U}(\vect{r}) &=& \frac{8}{\pi^2} \frac{N_\uparrow}{R_{Ux} R_{Uy} R_{Uz}}
429: \; \left[\max \left(1 - \sum_i
430: \frac{x_i^2}{R_{Ui}^2},0\right)\right]^{3/2}
431: \label{e:Unitaritydensity}
432: \end{eqnarray}
433: with the radii $R_{U{x,y,z}} = \xi^{1/4} R_{F{x,y,z}}$.
434: Table~\ref{tab:zerotemptrappeddensity} summarizes the various
435: density profiles of interacting Fermi mixtures.
436: 
437: Remarkably, the functional form $n_{\uparrow U}(\vect{r}) \propto
438: (\mu-V(\vect{r}))^{3/2}$ is identical to that of a non-interacting
439: Fermi gas.  The underlying reason is that the equation of state
440: $\mu \propto n^{2/3}$ has the same power-law form as for
441: non-interacting fermions. The universal constant $\xi$ simply
442: rescales the radii (by a factor $\xi^{1/4}$) and the central
443: density (by a factor $\xi^{-3/4}$).  One thus has direct
444: experimental access to the universal constant $\xi$ by measuring
445: the size of the cloud at unitarity (see section~\ref{s:energymeasurements}).
446: 
447: \begin{table}
448:   \centering
449:   \begin{tabular}{p{0.02\linewidth}c|c|c|cp{0.3\linewidth}}
450:     % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
451: &     &  BEC-limit  &   Unitarity  & BCS-limit \\
452: &$\frac{1}{k_F a}$ & $\infty$ & 0& $-\infty$\\ \hline
453: &    $\gamma$ (in $\mu \propto n^\gamma$) & 1 & 2/3 & 2/3 \\
454: &    $n_\uparrow(\vect{r})/n_\uparrow(\vect{0})$ & $1 - \sum_i \frac{x_i^2}{R_{i}^2}$ & $(1 - \sum_i \frac{x_i^2}{R_{Ui}^2})^{3/2}$ & $(1 - \sum_i \frac{x_i^2}{R_{Fi}^2})^{3/2}$ \\
455: &$n_\uparrow(\vect{0})$ & $\frac{15}{8\pi} \frac{N_\uparrow}{R_x R_y R_z}$ &
456: $\frac{8}{\pi^2}  \frac{N_\uparrow}{R_{Ux} R_{Uy} R_{Uz}}$ &
457: $\frac{8}{\pi^2}  \frac{N_\uparrow}{R_{Fx} R_{Fy} R_{Fz}}$ \\
458: &Radii & $R_i = \sqrt{\frac{2 \mu_M}{M \omega_i^2}}$ & $R_{Ui} =
459: \xi^{1/4} R_{Fi}$ & $R_{Fi} = \sqrt{\frac{2 E_F}{m \omega_i^2}}$
460: \end{tabular}
461:   \caption{Zero-temperature density profiles of a trapped, interacting Fermi mixture in the BEC-BCS crossover. The density is zero when the expressions are not positive. For definitions see the text.}
462:   \label{tab:zerotemptrappeddensity}
463: \end{table}
464: 
465: 
466: \subsection{Expansion of strongly interacting Fermi mixtures}
467: \label{s:expansion}
468: 
469: Intriguingly rich physics can be uncovered by the simple release
470: of ultracold gases from their confining trap. From the size of the
471: expanded cloud and the known time-of-flight one directly obtains
472: the energy content of the gas: the temperature in the case of
473: thermal clouds, the Fermi energy for non-interacting degenerate
474: Fermi gases, the mean-field energy for Bose-Einstein condensates.
475: In the case of free ballistic expansion, where no collisions occur
476: during expansion, the density distribution of the expanded cloud
477: directly reveals the original momentum distribution of the
478: particles in the trap. Thermal clouds will become spherical after
479: ballistic expansion, reflecting their isotropic momentum
480: distribution in the trap. The expansion of Bose-Einstein
481: condensates is not ballistic but mean-field driven, leading to
482: superfluid hydrodynamic expansion. As mean-field energy is
483: preferentially released in the direction(s) of tight confinement,
484: this allows for the famous ``smoking gun'' signature of
485: Bose-Einstein condensation: inversion of the condensate's aspect
486: ratio after expansion out of an anisotropic trap. In strongly
487: interacting gases the normal, uncondensed cloud can be
488: collisionally dense, and will expand according to classical
489: hydrodynamics. As particles will preferentially leave the cloud
490: along the narrower dimensions, where they undergo fewer
491: collisions, this also leads to an inversion of the cloud's initial
492: aspect ratio. It is thus no longer a ``smoking gun'' for
493: condensation, but merely for strong interactions. Expansion is
494: also useful to measure correlations in momentum
495: space~\cite{grei05corr}. Finally, in the case of harmonic
496: trapping, expansion of a superfluid cloud can often be described
497: as a ``magnifying glass'', a mere scaling of the density
498: distribution in the trap. This allows for example to observe
499: quantized vortices~\cite{zwie05vort}, which are too small to be
500: observable in the trap.  In this section, we show how quantitative
501: information can be derived from images of expanding clouds.
502: 
503: 
504: \subsubsection{Free ballistic expansion}
505: \label{s:ballistic}
506: 
507: Let us consider the expansion of a non-condensed thermal cloud. If
508: the mean free path $\lambda_c$ between collisions is longer than
509: the size of the trapped cloud $R$, we can neglect collisions
510: during expansion, which is hence ballistic. The collision rate is
511: $\Gamma = n \sigma v$, with density $n$, collisional cross section
512: $\sigma$ and thermal (root mean square) velocity $v$, which gives
513: $\lambda_c = v/\Gamma = 1/n \sigma$. As $R = v/\omega$ for a
514: harmonic trap, the condition $\lambda_c \gg R$ is equivalent to
515: having $\Gamma \ll \omega$, that is, the mean time interval
516: between collisions should be larger than a period of oscillation
517: in the trap.
518: 
519: This condition can be fulfilled for the cloud of uncondensed
520: molecules in the BEC limit where $1/k_F a \gg 1$ and collisions
521: are negligible (this has been the case also for atomic BECs with
522: the exception of very large thermal clouds,
523: see~\cite{stam98coll,shva03hydro}), and for the cloud of unpaired
524: fermions in the BEC- and in the BCS-limit for $k_F |a| \ll 1$ (the
525: exact criterion is still $\Gamma \ll \omega$). For molecules with
526: mass $M$, we need to replace $m \rightarrow M$ in the following
527: discussion.
528: 
529: In the ballistic case, a particle initially at point $\vect{r}_0$
530: in the trap, will reach point $\vect{r} = \vect{r}_0 +
531: \frac{\vect{p}_0}{m} t$ after expansion time $t$. We obtain the
532: density at point $\vect{r}$ at time $t$ by adding the
533: contributions from particles at all points $\vect{r}_0$ that had
534: the correct initial momentum $\vect{p}_0 = m(\vect{r} -
535: \vect{r}_0)/t$. In terms of the semi-classical distribution
536: $f(\vect{r},\vect{p})$, Eq.~\ref{e:FermiBose}, this is
537: \begin{eqnarray}
538: n(\vect{r},t) &=& \Int{r}_0 \Intp{p_0} \; f(\vect{r_0},\vect{p_0})\; \delta\left(\vect{r} - \vect{r_0} - \frac{\vect{p}_0}{m}t\right) \nonumber\\
539: &=& \Intp{p_0}\; f\left(\vect{r}-\frac{\vect{p}_0}{m}t,\vect{p}_0\right) \nonumber\\
540: &=& \Intp{p_0}\; \left\{\exp\left[\beta\frac{\vect{p}_0^2}{2m} +
541: \beta V\left(\vect{r}-\frac{\vect{p}_0}{m}t\right) - \beta\mu\right] \mp
542: 1\right\}^{-1} \label{e:expandeddensity}
543: \end{eqnarray}
544: The integral can be carried out analytically in the case of a
545: harmonic potential (Eq.~\ref{e:potential}):
546: \begin{eqnarray}
547: n(\vect{r},t) &=& \Intp{p_0}\; \left\{\exp\left[\beta\sum_i \left(1+\omega_i^2 t^2\right) \frac{p_{0i}^2}{2m} +\beta\sum_i\frac{1}{2}m\frac{\omega_i^2 x_i^2}{1+\omega_i^2 t^2} - \beta\mu\right] \mp 1\right\}^{-1} \nonumber \\
548: &=& \pm \frac{1}{\lambda_{dB}^3}\prod_i
549: \frac{1}{\sqrt{1+\omega_i^2 t^2}}\; {\rm Li}_{3/2}\left[\pm
550: \exp\left(\beta\mu - \beta\sum_i\frac{1}{2}m\frac{\omega_i^2
551: x_i^2}{1+\omega_i^2 t^2}\right)\right] \nonumber
552: \end{eqnarray}
553: Note that this has the same form as the density distribution in
554: the trap, but with spatial dimension $i={x,y,z}$ rescaled by the
555: factor $b_i(t) = \sqrt{1 + \omega_i^2 t^2}$. Ballistic expansion
556: of a thermal (bosonic or fermionic) cloud from a harmonic trap is
557: thus a scaling transformation:
558: \begin{eqnarray}
559:     n(\vect{r},t) = \frac{1}{\mathcal{V}(t)}\;n\left(\frac{x}{b_x(t)},\frac{y}{b_y(t)},\frac{z}{b_z(t)},t =0\right)
560: \label{e:scalingtrafo}
561: \end{eqnarray}
562: where the unit volume scales as $\mathcal{V}(t) = b_x b_y b_z$.
563:  After an expansion time long compared to the trapping periods ($t \gg 1/\omega_i$), we have
564: \begin{eqnarray}
565: n(\vect{r},t \gg 1/\omega_i) &=& \pm \frac{1}{\lambda_{
566: dB}^3}\frac{1}{(\bar{\omega} t)^3}\, {\rm Li}_{3/2}\left[\pm
567: \exp\left(\beta\mu -
568: \beta\frac{1}{2}m\frac{\vect{r}^2}{t^2}\right)\right]
569: \label{e:expdensitylongtimes}
570: \end{eqnarray}
571: As expected, we obtain an isotropic density profile, reflecting
572: the original isotropic momentum distribution of the trapped gas.
573: Importantly, the {\it shape} of the density profile, i.e. its
574: variation with $\vect{r}$, becomes insensitive to the trapping
575: potential. Eq.~\ref{e:expdensitylongtimes} thus holds for a
576: general trapping geometry, for expansion times long compared to
577: the longest trapping period. Even if the trapping potential is not
578: known in detail, one can still determine the cloud's temperature
579: and even decide whether the gas is degenerate. Note that the {\it
580: momentum distribution} at point $\vect{r}$ after long expansion
581: times $t \gg 1/\omega_i$ has become {\it anisotropic}:
582: \begin{eqnarray}
583: f(\vect{r},\vect{p_0},t) &=& \Int{r}_0 \; f(\vect{r_0},\vect{p_0})\; \delta\left(\vect{r} - \vect{r_0} - \frac{\vect{p}_0}{m}t\right) \nonumber\\
584: &=&  f\left(\vect{r}-\frac{\vect{p}_0}{m}t,\vect{p}_0\right) \nonumber\\
585: \stackrel{t \gg 1/\omega_i}{\rightarrow} &=&
586: \left(\exp\left[\beta\left(\sum_i \omega_i^2 t^2
587: \frac{\left(p_{0i} - m \frac{x_i}{t}\right)^2}{2m}
588: +\frac{1}{2}\,m\frac{\vect{r}^2}{t^2} - \mu\right)\right] \mp
589: 1\right)^{-1} \label{e:ballisticmomentum}
590: \end{eqnarray}
591: The momentum distribution at point $\vect{r}$ is ellipsoidal,
592: centered at $\bar{\vect{p}} = m\frac{\vect{r}}{t}$, and with
593: characteristic widths $\Delta p_i \propto m\frac{\Delta x_i}{t}
594: \propto \frac{1}{\omega_i}$ directly mirroring the ellipsoidal
595: atomic distribution in the trap.
596: 
597: \paragraph{Ballistic expansion into a saddle potential}
598: \label{s:saddleballistic}
599: 
600: In many experiments, atoms are released from an optical trap, but
601: magnetic fields (Feshbach fields) are still left on.  In general,
602: these magnetic fields are inhomogeneous, either due to technical
603: limitations, or deliberately, e.g. in case of the optical-magnetic
604: hybrid trap discussed in section~\ref{s:opticaltrap}.  We focus here on the
605: important case of a magnetic field created by pair of coils which
606: generates a saddle point potential.
607: 
608: So we assume that at $t>0$, the gas is not released into free
609: space, but into a new potential.  We define $V(\vect{r},t>0) =
610: \frac{1}{2} m \left(\omega_{Sx}^2 x^2 + \omega_{Sz}^2 y^2 +
611: \omega_{Sz}^2 z^2\right)$, and can describe expansion into
612: anticonfining potentials with imaginary frequencies. For
613: example, for the magnetic saddle potentials relevant for the MIT
614: experiments, the radial dimension is anticonfining and
615: $\omega_{Sx,y} = {\rm i} \frac{1}{\sqrt{2}}\omega_{Sz}$. In the
616: potential $V(\vect{r},t>0)$, particles with initial position
617: $\vect{r_0}$ and momentum $\vect{p_0}$ will reach the point
618: $\vect{r}$ with $x_i = \cos(\omega_{Si} t) x_{0i} +
619: \frac{1}{\omega_{Si}} \sin(\omega_{Si} t) \frac{p_{0i}}{m}$ after
620: expansion time $t$. The calculation of the density profile is
621: fully analogous to the case of free expansion, after the change of
622: variables $x_{0i} \rightarrow \tilde{x}_{0i}/\cos(\omega_{Si}t)$
623: and the substitution $t \rightarrow
624: \sin(\omega_{Si}t)/\omega_{Si}$. We again obtain a scaling
625: transformation, Eq.~\ref{e:scalingtrafo}, but for this case with
626: scaling parameters $b_i(t) = \sqrt{\cos^2(\omega_{Si}t) +
627: \frac{\omega_i^2}{\omega_{Si}^2} \sin^2(\omega_{Si}t)}$. For
628: expansion into the magnetic saddle potential, this gives
629: $b_\perp(t) = \sqrt{\cosh^2(\frac{1}{\sqrt{2}}\omega_{Sz}t) +
630: \frac{2\omega_\perp^2}{\omega_{Sz}^2}
631: \sinh^2(\frac{1}{\sqrt{2}}\omega_{Sz}t)}$ and $b_z(t) =
632: \sqrt{\cos^2(\omega_{Sz}t) + \frac{\omega_z^2}{\omega_{Sz}^2}
633: \sin^2(\omega_{Sz}t)}$. For the MIT trap, the initial axial
634: trapping potential is dominated by the magnetic field curvature,
635: while the initial radial potential is almost entirely due to the
636: optical trap. After switching off the optical trap, we have
637: $\omega_{Sz} = \omega_z$ and $\omega_{Sx} = {\rm
638: i}\frac{1}{\sqrt{2}}\omega_z$. In this case, $b_z(t) = 1$ and the
639: cloud expands only into the radial direction.
640: 
641: \subsubsection{Collisionally hydrodynamic expansion}
642: \label{s:collhydro}
643: 
644: If the mean free path $\lambda_c$ is short compared to the cloud
645: size, the gas is in the hydrodynamic regime, and collisions during
646: expansion can no longer be neglected.  Collisions will tend to
647: reestablish local thermal equilibrium, in particular an isotropic
648: momentum distribution. For anisotropic traps, this directly leads
649: to anisotropic expansion, in strong contrast to the ballistic
650: case: Particles trying to escape in one direction suffer
651: collisions that redistribute their momenta equally in all
652: directions. The escape is hindered more for the weakly confined
653: directions where the cloud is long initially and particles can
654: undergo more collisions. For cylindrically symmetric clouds, this
655: leads to an inversion of the aspect ratio of the cloud during
656: expansion.
657: 
658: Hydrodynamic expansion can take place for $1/k_F |a| < 1$, which
659: includes (for $a>0$) strongly interacting clouds of
660: uncondensed molecules, and (for $a<0$) a strongly interacting,
661: normal Fermi mixture.  There is no sharp boundary between
662: molecular hydrodynamics and fermionic hydrodynamics, since $1/k_F
663: |a| < 1$ is the strongly interacting regime where many-body
664: physics dominates and the single-particle description (molecules
665: in one limit, unbound fermions in the other) is no longer valid.
666: 
667: In the hydrodynamic regime, the evolution of the gas is governed
668: by the continuity equation for the density $n(\vect{r},t)$ and,
669: neglecting friction (viscosity), the Euler equation for the
670: velocity field $\vect{v}(\vect{r},t)$:
671: \begin{eqnarray}
672: \frac{\partial n}{\partial t} + \nabla \cdot (n \vect{v}) &=& 0 \\
673: m \frac{{\rm d}\vect{v}}{{\rm d}t} = m \frac{\partial
674: \vect{v}}{\partial t} + m (\vect{v} \cdot \nabla) \vect{v} &=& -
675: \nabla V(\vect{r},t) - \frac{1}{n} \nabla P(\vect{r},t)
676: \end{eqnarray}
677: where $P$ is the pressure. Friction is
678: negligible deep in the hydrodynamic regime, when the mean free
679: path approaches zero.  The Euler equation is simply Newton's law
680: for the collection of gas particles at $\vect{r}$. In steady
681: state, we recover the equilibrium solution
682: \begin{equation}
683: \nabla P_0(\vect{r}) = n_0(\vect{r}) \nabla \mu_0(\vect{r}) = -
684: n_0(\vect{r}) \nabla V(\vect{r},0) \label{e:initialpressure}
685: \end{equation}
686: where we have used the expression for the local chemical potential
687: $\mu_0(\vect{r}) = \mu - V(\vect{r})$.
688: \paragraph{Scaling solution for harmonic potentials}
689: In the case of free expansion, the potential $V(\vect{r},t)$ is
690: the initial harmonic trapping potential for $t<0$, with radial and
691: axial trapping frequencies $\omega_{\perp}(0)$ and
692: $\omega_{z}(0)$, and zero for $t>0$. We can more generally
693: consider here an arbitrary time variation $\omega_\perp(t)$ and
694: $\omega_z(t)$ of the trapping frequencies. For this case, the
695: Euler equation allows a simple scaling solution for the
696: coordinates and velocities~\cite{kaga97bose}
697: \begin{eqnarray}
698: x_i(t) &=& b_i(t)\, x_{0i} \nonumber\\
699: v_i(t) &=& \frac{\dot{b}_i}{b_i}\, x_i(t)
700: \end{eqnarray}
701: with initial conditions $b_i(0) = 1$ and $\dot{b}_i(0) = 0$. The
702: unit volume scales as $\mathcal{V}(t) = b_x\,b_y\,b_z$, the
703: density varies as $n(\vect{r},t) = n_0(\vect{r}_0)/\mathcal{V}$,
704: where the fluid element at initial position $\vect{r}_0$ has
705: propagated to $\vect{r}$ at time $t$.
706: 
707: \paragraph{Pressure}
708: The thermodynamic properties of a simple fluid or gas only depend
709: on three variables, that are, in the grand canonical description,
710: the temperature $T$, the chemical potential $\mu$ and the volume
711: $V$. From the grand canonical partition function Z, one obtains in
712: this case the pressure $P = k_B T \frac{\ln Z}{V}$. For a
713: non-interacting, ideal gas of bosons or fermions, the average
714: energy is $E = \frac{3}{2} k_B T \ln Z$, leading to the relation
715: $PV = \frac{2}{3} E$. This equation is no longer true for an
716: interacting gas, for example the van der Waals gas. It is very
717: remarkable, then, that this relation nevertheless holds also for
718: the strongly interacting, unitary gas on resonance, for all
719: temperatures~\cite{ho04uni,thom05virial}~\footnote{On resonance,
720: universality requires that the energy $E = N \epsilon_F f(T/T_F)$
721: with a universal function $f$. Entropy can only be a function of
722: $T/T_F$, so adiabaticity requires this ratio to be constant. The
723: pressure is then $P = -\partial E/\partial V|_{S,N} = -N f(T/T_F)
724: \partial \epsilon_F/\partial V = \frac{2}{3} E/V$.}. Under an
725: adiabatic expansion, the energy $E$ changes according to ${\rm d}E
726: = - P {\rm d}V$. Hence $\frac{3}{2}{\rm d} (PV) = \frac{3}{2}
727: (V{\rm d} P + P{\rm d} V) = - P {\rm d}V$, which leads to the law
728: $P V^{5/3} = {\rm const}$ for adiabatic expansion. The pressure
729: thus scales as $\mathcal{V}^{-5/3}$, and the force, using
730: Eq.~\ref{e:initialpressure},
731: \begin{eqnarray}
732:     -\frac{1}{n} \frac{\partial}{\partial x_i} P(\vect{r},t) &=& -\frac{\mathcal{V}}{n_0} \frac{1}{b_i} \frac{\partial}{\partial x_{0i}} \frac{P_0(\vect{r}_0)}{\mathcal{V}^{5/3}} = \frac{1}{b_i\mathcal{V}^{2/3}} \frac{\partial}{\partial x_{i0}} V(\vect{r}_0,0)\nonumber \\
733: &=& \frac{1}{b_i\mathcal{V}^{2/3}} m \omega_{i}^2(0) x_{i0}
734: \end{eqnarray}
735: 
736: The Euler equations then reduce to equations for the scaling
737: parameters $b_i(t)$, which can be solved numerically:
738: \begin{eqnarray}
739:     \label{e:eulerclass}
740:     \ddot{b}_i = -\omega_i^2(t)\, b_i + \frac{\omega_{i}^2(0)}{b_i\mathcal{V}^{2/3}}
741: \end{eqnarray}
742: 
743: In the following section we will see that superfluid hydrodynamics
744: leads to very similar scaling equations, with the exponent $2/3$
745: for the volume scaling parameter $\mathcal{V}$ replaced by the
746: parameter $\gamma$ in the equation of state of the superfluid
747: $\mu(n) = n^\gamma$. The discussion of free expansion, the
748: long-time behavior, inversion of the aspect ratio etc. will be
749: identical for superfluid hydrodynamics, so we defer the topic until
750: the next section.
751: 
752: \paragraph{From ballistic to hydrodynamic expansion}
753: 
754: The regime in between ballistic, collisionless expansion and pure
755: hydrodynamic, collisional expansion can be treated approximately.
756: For the effects of interactions on a classical gas,
757: see~\cite{guer99osc,pedr03}, for the case of Fermi gases with
758: attractive interactions, see~\cite{meno02}.
759: 
760: \subsubsection{Superfluid hydrodynamic expansion}
761: 
762: In the simplest (scalar) case, a superfluid is described by a
763: macroscopic, complex order parameter $\psi(\vect{r},t) =
764: \sqrt{n(\vect{r},t)} e^{{\rm i} \phi(\vect{r},t)}$ parameterized by
765: the {\it superfluid density} $n(\vect{r},t)$ and a phase
766: $\phi(\vect{r},t)$. The dynamics of the order parameter are well
767: described by a time-dependent Schr\"odinger equation of the type
768: \begin{equation}
769:     {\rm i} \hbar \frac{\partial}{\partial t} \psi(\vect{r},t) = \left(-\frac{\hbar^2}{2m}\nabla^2 + V(\vect{r},t) + \mu(n(\vect{r},t))\right)\psi(\vect{r},t)
770: \label{e:GPequation}
771: \end{equation}
772: where $\mu(n)$ is the chemical potential given by the equation of
773: state of the superfluid. In the case of weakly interacting BECs,
774: this is the Gross-Pitaevskii equation for the condensate
775: wave function from section~\ref{s:trappedmixtures}. For fermionic superfluids, a formally similar equation is the Ginzburg-Landau equation, which is however valid only close to $T_C$.
776: Rewriting Eq.~\ref{e:GPequation}
777: in terms of the superfluid density $n$ and velocity $\vect{v}$,
778: neglecting the curvature $\nabla^2 \sqrt{n}$ of the magnitude of
779: $\psi$ and using the fact that the superfluid is
780: irrotational $\nabla \times \vect{r} = 0$, we arrive again at the
781: continuity equation and the Euler equation for classical inviscous
782: flow:
783: \begin{eqnarray}
784: \frac{\partial n}{\partial t} + \nabla \cdot (n \vect{v}) &=& 0\\
785: m \frac{\partial \vect{v}}{\partial t} + m (\vect{v} \cdot \nabla)
786: \vect{v} &=& - \nabla \left(V+ \mu(n)\right)
787: \end{eqnarray}
788: The validity of these hydrodynamic equations is restricted to superfluids whose
789: healing length is much smaller than the sample size and thus, for fermionic superfluids in
790: a harmonic trap, for a superfluid gap larger than
791: the harmonic oscillator energies $\hbar
792: \omega_{x,y,z}$~\cite{meno02}.
793: For a power-law equation of state $\mu(n) \propto n^\gamma$, the
794: equations allow a scaling solution for (possibly time-varying)
795: harmonic potentials. The scaling parameters $b_i(t)$ are
796: given by the differential equations~\cite{cast96,kaga97bose,meno02,hu04coll}
797: \begin{eqnarray}
798:     \ddot{b}_i = -\omega_i^2(t)\, b_i + \frac{\omega_{i}^2(0)}{b_i\mathcal{V}^{\gamma}}
799:     \label{e:euler}
800: \end{eqnarray}
801: Important limiting cases in the BEC-BCS crossover are:
802: \begin{itemize}
803: \item[-] BEC-limit ($1/k_F a \gg 1$): Here, the mean-field repulsion between molecules leads to a chemical potential per fermion $\mu(n) = \frac{\pi\hbar^2 a_M n}{m}$, so $\gamma = 1$.
804: \item[-] BCS-limit ($1/k_F a \ll -1$): In the BCS-limit, the dominant contribution to the chemical potential comes from the kinetic energy of the constituent fermions, given by the Fermi energy. So here $\mu(n) = \epsilon_F \propto n^{2/3}$ and $\gamma = 2/3$.
805: \item[-] Unitarity limit ($1/k_F a = 0$): In the unitarity limit, the only remaining energy scale is the Fermi energy. One necessarily has $\mu(n) \propto \epsilon_F \propto n^{2/3}$ and $\gamma = 2/3$, just as in the BCS-limit.
806: \end{itemize}
807: 
808: Note that the scaling laws for the BCS- and the unitarity limit~\cite{cast04scal}
809: are identical to those found for a collisionally hydrodynamic gas
810: in section~\ref{s:collhydro}.
811: For a derivation of superfluid hydrodynamics in the BCS-limit, we refer the reader to the contribution of Y. Castin to these lecture notes.
812: 
813: \begin{figure}
814:     \centering
815:     \includegraphics[width=3in]{figs_quantanalysis/gamma.eps}
816:     \caption{The exponent $\gamma$ as a function of the interaction parameter $1/k_F a$. $\gamma$ approximately describes the superfluid equation of state $\mu(n) \sim n^\gamma$ in the BEC-BCS crossover. A similar figure can be found in~\cite{hu04coll}.}
817:     \label{f:gamma}
818: \end{figure}
819: 
820: The Leggett ansatz (see section~\ref{s:crossoverwavefunction}) allows to interpolate
821: between the BEC- and the BCS-regime and gives a chemical potential
822: $\mu(n)$ that correctly captures the physics in the two limits.
823: With its help, we can define an effective exponent $\gamma =
824: \frac{n}{\mu}\frac{\partial \mu}{\partial n}$ and write $\mu(n)
825: \simeq n^\gamma$, assuming that $\gamma$ varies slowly with the
826: interaction parameter $1/k_F a$. This exponent, shown in
827: Fig.~\ref{f:gamma}, attains the correct limiting values in the
828: BEC- and the BCS-limit, as well as on resonance, so we may use it
829: for the present purpose as an approximate description of the gas'
830: equation of state throughout the crossover.
831: 
832: \paragraph{In-trap density profile}
833: The in-trap density profile of the superfluid at zero temperature can be deduced from
834: the Euler equations in steady state. Neglecting kinetic energy
835: $\frac{1}{2}\, m\, \vect{v}^2$ (Thomas-Fermi approximation), the
836: equation simply reads $V(\vect{r}) + \mu(n(\vect{r})) = {\rm
837: const.} = \mu(n(\vect{0}))$. For the power-law equation of state
838: $\mu(n)\propto n^\gamma$, we directly obtain
839: \begin{equation}
840:     n(\vect{r}) \propto \left(\mu(n(\vect{0})) - V(\vect{r})\right)^{1/\gamma}
841: \end{equation}
842: for $\mu(n(\vect{0})) > V(\vect{r})$ and zero otherwise. For a BEC
843: and harmonic trapping, we recover the inverted parabola,
844: Eq.~\ref{e:BECparabola}, for a BCS superfluid in the limit of weak
845: interactions the density distribution of an ideal Fermi gas,
846: Eq.~\ref{e:Fermidensity}. Note that in the crossover $1/k_F |a|
847: \lesssim 1$, the correct calculation of the density profile is
848: less straightforward, as the parameter $1/k_F(\vect{r}) a$ depends
849: on position, and the equation of state varies across the cloud.
850: The power-law approach, using a fixed $\gamma =
851: \gamma(1/k_F(\vect{0}) a)$, will only provide an approximate
852: description. Fortunately, on resonance evidently $1/k_F(\vect{r})a
853: = 0$ across the entire cloud, and the power-law equation of state
854: becomes exact at $T=0$.
855: 
856: \paragraph{Free expansion out of a cylindrically symmetric trap}
857: 
858: In this case $\omega_i(t>0) = 0$, and $\omega_x(0) = \omega_y(0)
859: \equiv \omega_\perp$. We have
860: \begin{eqnarray}
861: \label{e:cylindricscaling1}
862: \ddot{b}_\perp &=& \frac{\omega_\perp^2}{b_\perp^{2\gamma+1}
863: b_z^{\gamma}}\\ \ddot{b}_z &=&
864: \frac{\omega_z^2}{b_\perp^{2\gamma} b_z^{\gamma+1}}
865: \label{e:cylindricscaling2}
866: \end{eqnarray}
867: The MIT trap is cigar-shaped, with an aspect ratio of short to
868: long axes $\epsilon = \omega_z/\omega_\perp \ll 1$. In such a
869: case, expansion is fast in the radial, initially tightly confined
870: dimensions, whereas it is slow in the $z$-direction. For times
871: short compared to $\tau_\epsilon = \frac{1}{\omega_z \epsilon}$,
872: many axial trapping periods, we can set $b_z \approx 1$ on the
873: right side of Eqs.~\ref{e:cylindricscaling1} and \ref{e:cylindricscaling2}, decoupling the
874: transverse from the axial expansion. For $\gamma = 1$, the case of
875: a Bose-Einstein condensate of tightly bound molecules, the
876: simplified equations for $t\ll\tau_\epsilon$ have an analytic
877: solution~\cite{cast96,kaga97bose}: $b_\perp(t) = \sqrt{1 +
878: \omega_\perp^2 t^2}$ and $b_z(t) = 1 + \epsilon^2
879: \left(\omega_\perp t \arctan(\omega_\perp t) - \ln\sqrt{1 +
880: \omega_\perp^2 t^2}\right)$.  For long times $t$, the
881: expansion is linear in time: $b_\perp(t) = \omega_\perp t$ for $t \gg 1/\omega_\perp$ and
882: $b_z(t) = (\pi/2) \epsilon^2 \omega_\perp t$ for $t \gg \tau_\epsilon$. Note that the radial
883: expansion accidentally follows the same scaling law as that of a
884: ballistically expanding normal cloud.
885: 
886: The general behavior of the expanding gas is the same for all
887: relevant $\gamma$. Driven either by repulsive interactions
888: (BEC-case) or by kinetic energy (BCS-case), the gas first expands
889: radially at constant acceleration $\ddot{R}_\perp(t\ll
890: 1/\omega_\perp) = R_\perp(0) \omega_\perp^2$, and over a radial
891: trapping period reaches a final expansion velocity $\dot{R}(t\gg
892: 1/\omega_\perp) \approx \omega_\perp R_\perp(0)$. The axial size
893: grows as $b_z(t) -1 \approx \epsilon^2 \omega_\perp t$, leading to
894: an inversion of the cloud's aspect ratio from initially $\epsilon$
895: to $\sim 1/\epsilon$. This inversion is in contrast to the
896: isotropic aspect ratio of a ballistically expanding gas, and is
897: thus  characteristic for hydrodynamic expansion, which can be of
898: collisional {\it or} of superfluid origin. Fig.~\ref{f:aspect} and
899: table~\ref{t:expansionsummary} summarize the time evolution of the
900: cloud's radii and aspect ratios for $\gamma = 1$ (BEC) and $\gamma
901: = 2/3$ (BCS and Unitarity), while Fig.~\ref{f:velocityplot}
902: compares the long-time behavior of the velocities and aspect
903: ratios across the BEC-BCS crossover. For expansion out of an
904: elongated cigar-shaped trap and $\gamma = 2/3$, which holds in the
905: BCS-limit, at unitarity, but also for a collisionally hydrodynamic
906: gas, the asymptotic expansion velocity is $v_\perp =
907: \sqrt{\frac{3}{2}} \omega_\perp R_\perp(0) \approx 1.22\;
908: \omega_\perp R_\perp(0)$. This can be understood by noting that
909: the cloud's kinetic energy, initially distributed isotropically,
910: is released only into the radial direction during hydrodynamic
911: expansion, so $\frac{1}{2} m v_\perp^2 = \frac{3}{2} \mu =
912: \frac{3}{4} m \omega_\perp^2 R_\perp(0)^2$.
913: 
914: \begin{figure}
915:     \centering
916:     \includegraphics[width=3in]{figs_quantanalysis/aspect.eps}
917:     \caption{Aspect ratio $\epsilon(t) = R_x(t)/R_z(t)$ as a function of time for the MIT trap ($\epsilon = 1/6$) in ballistic, collisional or superfluid hydrodynamic expansion ($\gamma = 2/3$) and superfluid hydrodynamic expansion of a molecular BEC ($\gamma = 1$). }
918:     \label{f:aspect}
919: \end{figure}
920: 
921: 
922: \begin{figure}
923:     \centering
924:     \includegraphics[width=3in]{figs_quantanalysis/velplot.eps}
925:     \caption{Asymptotic velocities and aspect ratio for hydrodynamic expansion out of a very elongated cigar-shaped trap ($\epsilon = \omega_z/\omega_x \ll 1$), as a function of the interaction parameter $1/k_F a$. The dashed lines show the asymptotic values in the BCS-limit. A similar figure can be found in~\cite{hu04coll}.}
926:     \label{f:velocityplot}
927: \end{figure}
928: 
929: \begin{table}
930:   \centering
931:   \begin{tabular}{p{0.03\linewidth}r|c|c|cp{0.25\linewidth}}
932:   % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
933: &   & Ballistic & Hydrodynamic (BEC) & (BCS, unitarity, collisional) \\ \hline
934: &  $b_\perp(t)$ & $\sqrt{1+\omega_\perp^2 t^2}$ & $\sqrt{1+\omega_\perp^2 t^2}$ & $\sim 1.22\; \omega_\perp t$ \\
935: &  $b_z(t)$ & $\sqrt{1+\omega_z^2 t^2}$ & $1 + \epsilon^2 (\omega_\perp t \arctan(\omega_\perp t)$ & $\sim 2.05 \;\frac{\pi}{2}\epsilon^2 \omega_\perp t$ \\
936: && & $\qquad \quad- \ln\sqrt{1 + \omega_\perp^2 t^2})$ & \\
937: &  AR & 1 & $\frac{2}{\pi} \frac{1}{\epsilon}$ & $\sim 0.60\; \frac{2}{\pi} \frac{1}{\epsilon}$
938: \end{tabular}
939:   \caption{Comparison between ballistic and hydrodynamic expansion.
940: Formulas for hydrodynamic expansion assume a long cigar-shaped
941: trap ($\epsilon = \omega_z/\omega_x \ll 1$), formulas for the
942: aspect ratio (AR) and for the BCS, Unitarity, collisional limit
943: give the asymptotic behavior. The formula for BEC-expansion is valid at short times $t \ll \omega_\perp/\omega_z^2$, but also captures the correct long time limit.}\label{t:expansionsummary}
944: \end{table}
945: 
946: 
947: \paragraph{Hydrodynamic expansion into a saddle potential}
948: 
949: As discussed in section~\ref{s:saddleballistic}, expansion may not occur into free
950: space, but into an inhomogeneous magnetic field which is often
951: described by a saddle potential.
952: 
953: The Euler equations (\ref{e:euler}) now read for $t>0$
954: \begin{equation}
955:     \ddot{b}_i = -\omega_{S,i}^2\, b_i + \frac{\omega_{i}^2(0)}{b_i\mathcal{V}^{\gamma}}
956: \end{equation}
957: Here, $\omega_{S,i}$ are the real or imaginary frequencies
958: characterizing the saddle point potential. These equations
959: typically need to be solved numerically. For a Bose-Einstein
960: condensate of molecules expanding from long cigar-shaped traps
961: ($\epsilon \ll 1$), the {\it radial} equation again allows for an
962: analytic solution identical to that for a ballistically expanding,
963: non-interacting gas. One obtains
964: \begin{equation}
965: b_\perp(t) = \sqrt{\cosh^2\left(\frac{1}{\sqrt{2}}\omega_{Sz}t\right) +
966: \frac{2\omega_\perp^2}{\omega_{Sz}^2}
967: \sinh^2\left(\frac{1}{\sqrt{2}}\omega_{Sz}t\right)}.
968: \end{equation}
969: However, the {\it axial} cloud size behaves drastically different
970: from a non-interacting cloud. For $\omega_{Sz} = \omega_z$, the
971: axial cloud size of a non-interacting gas would never change
972: ($b_z(t) = 1$), whereas a hydrodynamic gas, released into the
973: radial dimensions, will start to {\it shrink} axially under the
974: influence of the confining axial potential. The cloud's energy
975: (interaction energy for a BEC, kinetic energy for a BCS
976: superfluid) escapes radially, hence there is not sufficient
977: pressure to maintain the axial cloud size.
978: 
979: Further discussions of superfluid hydrodynamics and scaling transformations can be found in the contributions of Y. Castin and S. Stringari to these proceedings.
980: 
981: \subsection{Fitting functions for trapped and expanded Fermi gases}
982: 
983: In the preceding sections we derived the 3D density distribution
984: of a Fermi mixture in various regimes. However, all imaging
985: techniques record column densities, density profiles integrated
986: along the line of sight (the $z$-axis in the following).
987: 
988: For condensed gases, where $n(\vect{r}) \approx n(\vect{0})\, (1 -
989: V(\vect{r})/\mu)^{1/\gamma}$, one obtains the column density
990: \begin{equation}
991:     n_{\rm 2D, c}(x,y) =  n_c\left(1 - \frac{x^2}{R_x^2} - \frac{y^2}{R_y^2}\right)^{\frac{1}{\gamma}+\frac{1}{2}}
992: \end{equation}
993: For thermal Bose (molecular) and Fermi clouds, we have
994: \begin{equation}
995: \label{e:coldens}
996:     n_{\rm 2D}(x,y) = n_{\rm 2D,0}\, {\rm Li}_2\left(\pm \exp\left[\beta\mu - \beta\frac{1}{2}m(\omega_x^2 x^2 + \omega_y^2 y^2)  \right]\right)/\,{\rm Li}_2\left(\pm\, e^{\beta\mu}\right)
997: \end{equation}
998: 
999: In the following, we will discuss the fitting functions valid in the different regimes of interaction, and the derived quantities.
1000: 
1001: \subsubsection{Non-interacting Fermi gases}
1002: \paragraph{Cloud size}
1003: In the classical regime at $T/T_F \gg 1$, the characteristic cloud
1004: size is given by the gaussian radius $\sigma_i = \sqrt{\frac{2 k_B
1005: T}{m\omega_i^2}}$. In the degenerate regime, however, the cloud
1006: size saturates at the Fermi radius $R_{Fi} = \sqrt{\frac{2
1007: E_F}{m\omega_i^2}}$. It is thus convenient to define a fit
1008: parameter that interpolates between the two limits:
1009: \begin{eqnarray}
1010:     R_i^2 &=& \frac{2 k_B T}{m\omega_i^2} \; f(e^{\mu\beta}) \rightarrow \left\{%
1011: \begin{array}{ll}
1012:     \sigma_i, & T/T_F \gg 1 \\
1013:     R_{Fi}, & T/T_F \ll 1 \\
1014: \end{array}%
1015: \right.    \nonumber \\
1016:     \mbox{where } f(x) &=& \frac{{\rm Li}_1(-x)}{{\rm Li}_0(-x)} = \frac{1 + x}{x} \ln(1+x)
1017: \end{eqnarray}
1018: For all temperatures, $R_i$ is thus directly related to the
1019: physical size of the cloud, and thus a better choice as a fit
1020: parameter than $\sigma_i$, which goes to zero at $T=0$, or
1021: $\sqrt{\frac{2\mu}{m\omega_i^2}}$, which goes to zero around
1022: $T/T_F = 0.57$. Numerically, using $R_i$ is easier to implement
1023: than using the root mean square radius of the cloud
1024: \begin{equation}
1025:     \left<x_i^2\right> = \frac{k_B T}{m \omega_i^2}\; \frac{{\rm Li}_4(-e^{\mu\beta})}{{\rm Li}_3(-e^{\mu\beta})}
1026: \end{equation}
1027: 
1028: \paragraph{Fitting function}
1029: 
1030: The fit function used for the density profiles of Fermi clouds is
1031: then in 2D
1032: \begin{equation}
1033:     n_{\rm 2D}(x,y) = n_{\rm 2D,0}\, \frac{{\rm Li}_2\left(\pm \exp\left[q - \left(\frac{x^2}{R_x^2} + \frac{y^2}{R_y^2}\right)\,f(e^q)\right]\right) }{{\rm Li}_2\left(\pm e^q\right)}
1034: \end{equation} and for 1D
1035: \begin{equation}
1036:     n_{\rm 1D}(x) = n_{\rm 1D,0}\, \frac{{\rm Li}_{5/2}\left(\pm \exp\left[q -\frac{x^2}{R_x^2} \,f(e^q)\right]\right) }{{\rm Li}_{5/2}\left(\pm e^q\right)}.
1037: \end{equation}
1038: The parameter $q = \mu\beta$, the logarithm of the fugacity,
1039: determines the {\it shape} of the cloud. For a small fugacity
1040: (large and negative $q$), the above functions reduce to the simple
1041: gaussian distribution of thermal clouds. For high fugacity (large
1042: and positive $q$), they tend to the
1043: zero-temperature distribution $n_{\rm 2D,0}(1 -
1044: \frac{x^2}{R_{Fx}^2})^2$ (in 2D) and $n_{\rm 1D,0}(1 -
1045: \frac{x^2}{R_{Fx}^2})^{5/2}$ (in 1D).
1046: 
1047: \paragraph{Derived quantities}
1048: 
1049: {\it Degeneracy}
1050: The degeneracy parameter $T/T_F$ can be calculated by combining
1051: Eq.~\ref{e:numberofatoms} with Eq.~\ref{e:Ferminumber}:
1052: \begin{equation}
1053:     \frac{T}{T_F} = \left[-6\, {\rm Li}_3(-e^{q})\right]^{-1/3}
1054:     \label{e:Degeneracy}
1055: \end{equation}
1056: This parameter depends only on the {\it shape} of the cloud. A
1057: characteristic point where shape deviations due to quantum
1058: statistics start to play a role is the point where $\mu$ changes
1059: sign, and we see from Eq.~\ref{e:Degeneracy} that this occurs at
1060: $T/T_F \approx 0.57$. Many non-ideal aspects of imaging, such as
1061: finite resolution, out of focus imaging, saturation, heating of
1062: the cloud by the probe pulse etc., tend to wash out the
1063: non-gaussian features of a highly degenerate Fermi cloud and hence
1064: lead to a larger value of $T/T_F$. However, dispersive effects due
1065: to non-resonant imaging light can potentially mimic sharp edges of
1066: the cloud, which the fitting routine would then falsely interpret
1067: to result from a very low $T/T_F$. It is clear that care has to be
1068: taken when determining the degeneracy parameter from the shape of
1069: the cloud alone.
1070: 
1071: \paragraph{Temperature}
1072: \label{s:temperature}
1073: The size of the cloud and the shape parameter $q$ give the
1074: temperature as
1075: \begin{equation}
1076:     k_B T = \frac{1}{2} m \omega_i^2 \frac{R_i^2}{b_i(t)^2} \frac{1}{f(e^q)}
1077:     \label{e:thermometry}
1078: \end{equation}
1079: where we have used the expansion factor $b_i(t)$ from
1080: section~\ref{s:expansion}. We recall that $b_i(t) =
1081: \sqrt{1+\omega_i^2 t^2}$ for the free expansion of a
1082: non-interacting Fermi gas. For low temperatures $T \ll T_F$,
1083: $f(e^{\mu\beta}) \rightarrow \mu\beta = \mu/k_B T$ and $R_i =
1084: b_i(t)\,R_{Fi}$. In this case, temperature only affects the wings
1085: of the density distribution, where the {\it local}
1086: $T/T_F(\vect{r})$ is still large. In fact,
1087: \begin{equation}
1088:     n_{\rm 1D}(x) \propto \left\{%
1089: \begin{array}{ll}
1090:     (1 - \frac{x^2}{R_{Fx}^2})^{5/2} & \hbox{for } x \ll R_{Fx} \\
1091:     e^{-\frac{x^2}{\sigma_x^2}} & \hbox{for } x \gg R_{Fx}, \\
1092: \end{array}%
1093: \right.
1094: \end{equation}
1095: and we see that temperature only affects the cloud's wings beyond
1096: the zero-temperature Fermi radius. Thermometry of very low
1097: temperature Fermi clouds is thus difficult, limited by the
1098: signal-to-noise ratio in the low-density wings of the
1099: distribution. This is different from thermometry of thermal clouds
1100: at high temperature $T \gg T_F$, where the entire size of the
1101: cloud $\sigma_i$ directly gives the temperature.
1102: 
1103: Because of the sensitivity to the cloud's wings, thermometry is
1104: more robust when the full 2D distribution is used for the fit.
1105: Alternatively, one can rely on the known trapping geometry plus
1106: the local density approximation and perform an average over the
1107: elliptical equipotential lines in the $x$-$y$ plane (line of sight
1108: integration necessarily mixes points at different values of the
1109: potential energy.) As the number of points included in the average
1110: grows with the distance from the cloud's center, the
1111: signal-to-noise will actually be best in the wings. Such an
1112: average is superior to a simple integration along the $x$-axis,
1113: for example, as this will more strongly mix regions that have
1114: different local $T/T_F$.
1115: 
1116: The ideal gases (Fermi, Bose, Boltzmann) are the only systems for
1117: which we have an exact description.  Therefore, they are
1118: attractive as a thermometer, when brought in contact with strongly
1119: interacting systems.  This concept has been recently carried out
1120: by determining absolute temperatures for imbalanced Fermi gases at
1121: unitarity~\cite{shin07phasediagram}.  In these systems, for sufficiently
1122: high imbalance, the majority cloud extends beyond the minority
1123: cloud, and is (locally) an ideal gas.  Therefore, in Ref.~\cite{shin07phasediagram} the spatial
1124: wings of these clouds could be fitted with the functions for the
1125: ideal Fermi gas discussed in this section, and absolute
1126: temperatures for the superfluid phase transition could be
1127: determined. The fitting of the majority wings had to be done with in-trap
1128: profiles, which required to address the effect of anharmonicities of the
1129: optical trapping potential.  Usually, for thermometry, ballistic
1130: expansion is preferable since velocity distributions are
1131: independent of the shape of the trapping potential.  However, in
1132: the case of imbalanced Fermi gases, the atoms in the wings
1133: can collide with the strongly interacting core during expansion,
1134: modifying their velocity distribution.
1135: 
1136: Another way to perform ideal gas thermometry is done by converting
1137: the sample to a non-interacting system by sweeping sufficiently
1138: far away from the Feshbach resonance.  If such magnetic field
1139: sweeps are adiabatic, they conserve entropy (but not temperature).
1140: By fitting the spatial profiles of the non-interacting cloud, the
1141: entropy $S$ of the strongly interacting system can be determined.
1142: If it is possible to vary the energy of the strongly interacting
1143: system in a controlled way (e.g. by using the virial theorem at
1144: unitarity~\cite{ho04uni,thom05virial} or by providing controlled heating~\cite{kina05heat}), one
1145: can determine the derivative $dS/dE$ which is equal to the inverse
1146: absolute temperature.  So far this method could be implemented
1147: only for a balanced Fermi system at unitarity~\cite{luo07entropy} and, due to the need
1148: of determining a derivative, could only provide temperatures
1149: averaged over a range of energies.
1150: 
1151: \paragraph{Number of atoms and Fermi energy}
1152: 
1153: The number of atoms in the observed spin state can be obtained
1154: from the total absorption recorded in the cloud's CCD image. The
1155: transmission of resonant light at pixel $(x,y)$ is given by
1156: $\tilde{T}(x,y) = e^{-\sigma_0 \int n_{\rm 3D}(x,y,z)\,{\rm d}z}$,
1157: where $\sigma_0$ is the resonant atom-photon cross section for
1158: light absorption. Thus, the number of atoms is
1159: \begin{equation}
1160:     N_\uparrow = \frac{A}{M \sigma_0} \sum_{pixels} - \ln(\tilde{T}(x,y))
1161: \end{equation}
1162: where $A$ is the area per pixel and $M$ the optical magnification.
1163: 
1164: Typically, the fitting functions are applied to the optical
1165: density $\sigma_0 \int_z n_{\rm 3D}(x,y,z) = -
1166: \ln\left(\tilde{T}(x,y)\right)$. The fit parameter $n_{\rm 2D,0}$
1167: thus measures the peak optical density of the cloud, while the
1168: radii $\tilde{R}_x$ and $\tilde{R}_y$ have units of camera pixels.
1169: The number of atoms described by the fitting function is thus
1170: given by
1171: \begin{eqnarray}
1172:     N_{\rm fit} &=& \frac{A}{M \sigma_0}\pi\, n_{\rm 2D,0}\, \tilde{R}_x \tilde{R}_y \frac{{\rm Li}_3(-e^{q})}{{\rm Li}_2(-e^{q})}\frac{{\rm Li}_0(-e^{q})}{{\rm Li}_1-e^{q})} \nonumber \\
1173: &\rightarrow& \frac{A}{\sigma_0}\left\{%
1174: \begin{array}{ll}
1175:     \frac{\pi}{3} \, n_{\rm 2D,0} \, \tilde{R}_{Fx} \tilde{R}_{Fy}, & T \ll T_F \\
1176:     \pi\, n_{\rm 2D,0} \, \tilde{\sigma}_{x} \tilde{\sigma}_{y}, & T \gg T_F. \\
1177: \end{array}%
1178: \right.
1179: \end{eqnarray}
1180: 
1181: From the number of atoms and the trapping frequencies, one can
1182: calculate the Fermi energy $k_B T_F$:
1183: \begin{equation}
1184: k_B T_F = \hbar \bar{\omega} (6 N_{\rm fit})^{1/3}
1185: \end{equation}
1186: 
1187: An independent determination of the Fermi energy is provided
1188: by the measured (physical) size of the cloud $R_i$ for highly
1189: degenerate clouds. For $T\rightarrow 0$, $R_i \approx b_i(t)
1190: R_{Fi}$ and thus $ k_B T_F = \frac{1}{2} m \omega_i^2
1191: \frac{R_i^2}{b_i(t)^2} $.  As only the trapping frequencies and
1192: the magnification of the imaging system enter into this equation,
1193: this relation allows a calibration of the light absorption cross
1194: section which may be reduced from the resonant cross section by
1195: detuning, non-ideal polarization of the probe light, and
1196: saturation.
1197: 
1198: For arbitrary temperature, the shape parameter $q$ enters the
1199: relation for the Fermi energy:
1200: \begin{equation}
1201: k_B T_F = k_B T \frac{T_F}{T} = \frac{1}{2} m \omega_i^2
1202: \frac{R_i^2}{b_i(t)^2} \frac{\left(-6\, {\rm
1203: Li}_3(-e^{q})\right)^{1/3}}{f(e^q)}
1204: \end{equation}
1205: 
1206: \subsubsection{Resonantly interacting Fermi gases}
1207: \label{s:resonantgases}
1208: 
1209: The calculation of density profiles of interacting gases is
1210: delicate. Already above the superfluid transition temperature,
1211: attractive interactions lead to a shrinking of the cloud. Since
1212: interactions (parameterized by the local $k_F a$) vary across the
1213: cloud, there is a priori no simple analytical function describing
1214: interacting Fermi gases. Experimentally, it turns out that the
1215: difference in the {\it shape} of a (balanced) interacting and a
1216: non-interacting Fermi mixture is minute around resonance and on
1217: the BCS-side. Especially for the resonant case ($1/k_F a = 0$),
1218: this has led to the common practice of using the shape of the
1219: non-interacting Fermi gas as  fitting function, and quote an
1220: effective temperature $\tilde{T}$ and effective degeneracy
1221: $\frac{\tilde{T}}{T_F}$ of resonantly interacting
1222: clouds~\cite{ohar02science,kina05heat}. In fact, universality on
1223: resonance implies that the gas' chemical potential must be
1224: $\mu(\vect{r}) = \xi(T/T_F) \epsilon_F(\vect{r})$, with a
1225: universal function $\xi(T/T_F)$ which only depends on the reduced
1226: temperature $T/T_F$~\cite{ho04uni}. The zero-temperature limit of
1227: $\xi \equiv \xi_0$ has been subject of extensive experimental and
1228: theoretical studies (see section~\ref{s:energymeasurements}), and its
1229: value is $\xi(0) \approx 0.42$. At $T=0$, we have for a trapped
1230: gas $\mu(\vect{r}) = \mu_0 - V(\vect{r}) = \xi_0
1231: \epsilon_F(\vect{r}) \propto n^{2/3}(\vect{r})$. The density
1232: profile will then have the exact same shape as a non-interacting
1233: Fermi gas, with a renormalized Fermi temperature. However, for
1234: finite temperature, $\xi(T/T_F)$ differs from the temperature
1235: dependence of a non-interacting gas~\cite{bulg06TC}, and there is
1236: no a priori reason that the shape of the cloud at unitarity should
1237: be similar to that of a non-interacting Fermi gas. It turns out
1238: that the difference is very small.
1239: 
1240: \begin{figure}[h]
1241: \begin{center}
1242: \includegraphics[width=5.3in]{figs_quantanalysis/dilemma.eps}
1243: \caption[The absence of a clear signature of condensation in the
1244: spatial profile of strongly interacting Fermi gases]{The absence
1245: of a signature of condensation in the spatial profile of strongly
1246: interacting Fermi gases.  Shown are high-resolution images of spin
1247: up atoms in a resonantly interacting, equal mixture of spin up and
1248: spin down for different temperatures. The lower graphs show
1249: azimuthally averaged radial profiles (noise level well below 1\%
1250: of the maximum optical density). All three clouds are very well
1251: fit using a finite-temperature Thomas-Fermi distribution (with
1252: fugacity $e^{\mu / k_B \widetilde{T}}$, central density $n_0$ and
1253: mean square radius $<r^2>$ as free parameters, see
1254: Eq.~\ref{e:coldens}).   However, the
1255: empirical temperatures of $\widetilde{T}/T_F = 0.22$ (a), $0.13$
1256: (b) and $0.075$ (c) determined from the profiles' wings indicate
1257: that at least clouds $b$ and $c$ should be in the superfluid
1258: regime. Trap parameters $\nu_r = 162$ Hz, $\nu_z = 22.8$ Hz, 10 ms
1259: time of flight, expansion factor 13.9,  atom numbers $N$ per spin
1260: state were $10.2$ (a), 9.5 (b) and 7.5 $\times
1261: 10^6$.}\label{f:dilemma}
1262: \end{center}
1263: \end{figure}
1264: 
1265: The shape similarity was an important issue in the quest for
1266: superfluidity in Fermi gases. In the case of weakly interacting
1267: BECs, condensation is apparent from the sudden appearance of a
1268: dense, central core in midst of a large thermal cloud. In contrast
1269: to that, Fermi gases do not show such a  signature, at least at
1270: first sight (see Fig.~\ref{f:dilemma}), and different detection
1271: methods for superfluidity were explored.
1272: 
1273: The only loophole that may allow seeing a signature of superfluidity in the
1274: spatial profile of balanced Fermi gases would be a rapid variation of $\xi(T/T_F)$ around the critical temperature $T_C$. This would
1275: translate into a sudden variation of the density at the interface
1276: between the normal and superfluid region, e.g. where the gas is
1277: locally critical, $T = T_C(\vect{r})$.
1278: 
1279: 
1280: We have indeed found a faint signature of condensation in density
1281: profiles of the unitary gas on resonance after expansion. These
1282: results will be presented in section~\ref{s:directunitarity}. Note
1283: that the observation of such a feature in the density profiles
1284: draws into question the common practice of determining an
1285: ``effective temperature'' from density profiles at unitarity using
1286: the ideal gas fitting function. In contrast to balanced Fermi
1287: mixtures, a striking signature of condensation can be observed in
1288: the density profiles of mixtures with {\it imbalanced} populations
1289: of spin up and spin down fermions. This will be discussed in
1290: section~\ref{s:directimbalance}.
1291: 
1292: \subsubsection{Molecular clouds}
1293: \label{s:molecularclouds}
1294: In partially condensed molecular gases that are weakly
1295: interacting, one can neglect the mutual repulsion between the
1296: condensate and the surrounding thermal cloud of molecules. The
1297: density distribution is typically well-fit by a bimodal sum of an
1298: inverted parabola for the condensate
1299: \begin{equation}
1300:     n_c(x,y) = n_{c0}\left(1 - \frac{x^2}{R_{cx}^2} - \frac{y^2}{R_{cy}^2}\right)
1301:     \label{e:condensatefit}
1302: \end{equation}
1303: and a Bose-function for the thermal cloud, as in
1304: Eq.~\ref{e:coldens}, with the parameter $q = \mu\beta$ often left as an adjustable parameter (instead of fixing it via the condensate's chemical potential $\mu = g n_{c0}$):
1305: \begin{equation}
1306:     n_{th}(x,y) = n_{th0} \, {\rm Li}_2\left(\exp\left[q-\frac{x^2}{R_{th,x}^2} - \frac{y^2}{R_{th,y}^2}\right]\right)/{\rm Li}_2(\exp\left[q\right])
1307: \label{e:bimodalfit}
1308: \end{equation}
1309: For practical purposes, this is often simplified by a gaussian, as if $q \ll 0$. Then $n_{th} \approx n_{th0} \, \exp\left(-\frac{x^2}{R_{th,x}^2} - \frac{y^2}{R_{th,y}^2}\right)$.
1310: 
1311: Once the condensate mean-field $2 g n_c$ (with $g =
1312: \frac{4\pi\hbar^2 a_M}{M}$) experienced by thermal molecules is no
1313: longer small compared to $k_B T$, the mutual repulsion can no
1314: longer be neglected. The thermal molecules will then experience a
1315: ``Mexican-hat'' potential, the sum of the confining harmonic
1316: potential $V_M(\vect{r})$, and the repulsion from the condensate $2
1317: g n_c(\vect{r})$ and from the surrounding thermal cloud, $2 g
1318: n_{th}(\vect{r})$. The thermal molecules themselves will in turn
1319: repel the condensate. The situation can be captured by two coupled
1320: equations for the condensate (in Thomas-Fermi approximation) and
1321: the thermal cloud:
1322: \begin{eqnarray}
1323:     g n_c(\vect{r}) &=& {\rm Max}\left(\mu - V_M(\vect{r}) - 2 g n_{th}(\vect{r}),0\right) \nonumber \\
1324:     n_{th}(\vect{r}) &=& \frac{1}{\lambda_M^3}\,{\rm Li}_{3/2}(e^{\beta\left(\mu - V_M(\vect{r}) - 2 g n_c(\vect{r}) - 2 g n_{th}(\vect{r})\right)}) \nonumber \\
1325: &=& \frac{1}{\lambda_M^3}\,{\rm
1326: Li}_{3/2}(e^{-\beta\left|\mu - V_M(\vect{r}) - 2 g
1327: n_{th}(\vect{r})\right|})
1328: \end{eqnarray}
1329: where $\lambda_M$ is the thermal de Broglie wavelength for molecules. In the case of weakly interacting Bose gases, one can neglect the mean field term $2 g n_{th}(\vect{r})$~\cite{nara98semi}.
1330: 
1331: Note that these coupled equations are only an approximative way to
1332: describe the strongly correlated gas. The mean field approximation
1333: for the thermal molecules neglects phonons and other collective
1334: excitations.  The above equations can be solved numerically. In
1335: the limit of strong interactions, the condensate almost fully
1336: expels the thermal molecules from the trap center, so that the
1337: thermal cloud forms a shell around the condensate.
1338: 
1339: The practical implication of this discussion is that there is no
1340: simple analytic expression for the density distribution of partially condensed
1341: clouds in the strongly interacting regime. For fitting purposes
1342: one may still choose for example the bimodal fit of
1343: Eq.~\ref{e:bimodalfit}, but one must be aware that quantities like
1344: the ``condensate fraction'' thus obtained depend on the model
1345: assumed in the fit.   For tests of many-body calculations, the
1346: full density distributions should be compared to those predicted
1347: by theory.
1348: 
1349: \paragraph{Derived quantities}
1350: {\it Temperature.}
1351: 
1352: For weakly interacting Bose gases, Eq.~\ref{e:bimodalfit} holds
1353: and the temperature is given by
1354: \begin{equation}
1355:     k_B T = \frac{1}{2} m \omega_i^2 \frac{R_{th,i}^2}{b_i(t)^2}
1356: \end{equation}
1357: where $b_i(t) = \sqrt{1 + \omega_i^2 t^2}$ is the expansion factor
1358: of the thermal gas. To ensure model-independent results, only the
1359: thermal gas should be included in the fit, not the condensed core.
1360: For strongly interacting clouds, temperature can in principle
1361: still be obtained from the thermal wings of the {\it trapped}
1362: molecular distribution, which is gaussian at distances $\vect{r}$
1363: for which $V_M(\vect{r}) \gg \mu$.  However, a
1364: possible systematic correction can occur in expansion due to
1365: interactions of molecules in the wings with the core, which may be
1366: either condensed or strongly interacting.
1367: 
1368: Note that unless the whole cloud is deep in the hydrodynamic
1369: regime, there is no simple scaling law for the expansion of such
1370: strongly interacting molecular gases.  Absolute thermometry of
1371: strongly interacting, balanced gases is still a challenging
1372: problem.
1373: 
1374: 
1375: {\it Chemical potential}
1376: 
1377: In a confining potential, and at zero temperature, the chemical
1378: potential is given by the size of the condensate, as $V(\vect{r})
1379: = \mu$. It can be expressed by the fit parameters according to
1380: Eq.~\ref{e:condensatefit} as
1381: \begin{equation}
1382:     \mu = \frac{1}{2} m \omega_i^2 \frac{R_{c,i}^2}{b_i(t)^2}
1383: \end{equation}
1384: with $b_i(t)$ the expansion factor for superfluid hydrodynamic
1385: expansion into direction $i$. At finite temperatures and for
1386: strong interactions, the thermal cloud will mostly reside outside
1387: the condensate and can affect the actual or fitted condensate
1388: size.
1389: 
1390: {\it Condensate fraction}
1391: 
1392: In the field of dilute atomic gases, the condensate fraction is a
1393: key quantity to characterize the superfluid regime.  In contrast
1394: to superfluid helium and superconductors, gaseous condensates can
1395: be directly observed in a dramatic way. However, unless
1396: interactions are negligible,  the determination of the condensate
1397: fraction is  model dependent. For weakly interacting gases (or
1398: those obtained after a rapid ramp into the weakly interacting
1399: regime), the density distribution can typically be well fit with
1400: the bimodal distribution of Eqs.~\ref{e:condensatefit} and
1401: \ref{e:bimodalfit}. A robust way to define a ``condensate
1402: fraction'' is then to ascribe the total number of molecules in the
1403: narrower distribution to the condensate. For strong interactions
1404: however, the mean-field repulsion of thermal and condensed
1405: molecules (see above) will lead to the expulsion of a large part
1406: of thermal molecules from the condensate. In addition, low-energy
1407: excitations such as phonons, as well as quantum depletion will
1408: modify the non-condensed fraction at the position of the
1409: condensate, and the fitted condensate fraction depends on the form
1410: of the fitting function for the bimodal fit.  In these cases, it
1411: is better to directly compare density distributions with
1412: theoretical predictions.
1413: