1: \documentclass[pra,aps,twocolumn,showpacs]{revtex4}
2: \usepackage{graphicx}
3: \usepackage{dcolumn}
4: \newcommand{\nn}{\nonumber}
5:
6: \begin{document}
7: \def\be{\begin{equation}}
8: \def\ee{\end{equation}}
9: \def\bearr{\begin{eqnarray}}
10: \def\eearr{\end{eqnarray}}
11: \def\la{\langle}
12: \def\ra{\rangle}
13: \def\l{\left}
14: \def\r{\right}
15:
16: \title{Dynamic structure factor of Fermi superfluid in the BEC-BCS crossover}
17:
18: \author{Tarun Kanti Ghosh}
19: \affiliation
20: {Institute for Theoretical Physics, Heinrich-Heine Duesseldorf University,
21: 40225, Duesseldorf, Germany}
22:
23: \date{\today}
24:
25: \begin{abstract}
26: We consider cigar shaped Fermi superfluid in the BEC-BCS crossover.
27: Using polytropic form of equation of state, we derive low energy multibranch
28: bosonic excitations and the corresponding density fluctuations in three different
29: regimes along the crossover, namely
30: weak-coupling BCS, unitarity and molecular BEC regimes.
31: Bragg spectroscopy can be used to probe the multibranch nature of the low
32: energy bosonic excitations by measuring dynamic structure factor.
33: Therefore, we calculate dynamic structure factor in those three different regimes.
34: In the Bragg spectroscopy, an actual observable is momentum imparted to
35: the superfluid due to the Bragg potential. We also present results of the momentum
36: imparted to the superfluid due to the Bragg pulses.
37: \end{abstract}
38:
39: \pacs{03.75.Ss,03.75.Kk,32.80.Lg}
40:
41: \maketitle
42:
43: \section{Introduction}
44: The crossover from Bose-Einstein condensation (BEC) to Bardeen-Cooper-Schrieffer (BCS)
45: state has drawn renewed interest in past few years due to rapid experimental progress
46: in cold atomic two-component Fermi gases. Several experimental groups
47: \cite{greiner,jochim, zw,hara,regal,barten,bourdel,chin}
48: achieved the BEC-BCS crossover in the two-component atomic Fermi gases due to the
49: magnetized Feshbach resonance mechanism \cite{houb,stwa,ties}.
50: The fermionic system becomes molecular BEC for strong repulsive interaction and
51: transform into the BCS states when the interaction is
52: attractive. Near the resonance, the zero energy $s$-wave scattering length $a$
53: exceeds the interparticle spacing and the interparticle interactions
54: are unitarity limited and universal.
55:
56:
57: As in the case of bosonic clouds the frequencies of collective
58: modes of Fermi gases can be measured to high accuracy. The collective
59: oscillation frequencies of a trapped gas can provide crucial information
60: on the equation of state of the system.
61: The experimental results on the collective frequencies of the lowest axial
62: and radial breathing modes on ultra cold gases of ${}^6$Li across the
63: Feshbach resonance have also become available \cite{freq1,freq2}.
64: Since the weak-coupling BCS and unitarity limits are characterized by
65: the same collective oscillation frequency, it is an interesting to find
66: out another observable which makes a clear identification of these two regimes
67: and to better characterize two kinds of superfluid.
68:
69:
70: All the experiments of cold atomic Fermi gases are done in a cigar shaped
71: geometry in which the atomic density is inhomogeneous in the radial
72: plane and quasi-homogeneous along the symmetry axis. The axial excitations
73: of a cigar shaped Fermi superfluid can be divided into two regimes:
74: short wavelength excitations whose wavelength is much smaller than
75: the axial size and long wavelength excitations whose wavelength is
76: equal or larger than the axial size of the system.
77: In the later case, the axial excitations are discrete and the lowest
78: axial breathing mode frequency has been measured \cite{freq2}.
79: In the former case, the short wavelength
80: axial phonons with different number of radial modes of a cigar-shaped
81: Fermi superfluid give rise to the multibranch spectrum \cite{zaremba}.
82: These are similar to the electromagnetic wave propagation in a wave guide.
83: In the BCS side of the resonance, the low energy bosonic excitations, apart
84: from the gaped fermionic excitations, of the Fermi superfluid are called
85: multibranch Bogoliubov-Anderson (BA) modes.
86: In the usual electronic superconductors, the BA phonon mode is absent due to
87: the long-range Coulomb interaction.
88:
89: In this work we find that the low energy multibranch modes in the BCS limit are
90: different from that of in the unitarity limit,
91: although discrete radial and axial mode frequencies are same due to the same exponent
92: in the equation of state in the unitarity and BCS regimes.
93: Therefore, these multibranch modes can be used to identify and characterize the
94: weak-coupling BCS and unitarity states.
95:
96: It is an interesting to study how one can probe such bosonic modes in
97: the current available experimental setup. In fact, these multibranch low-energy
98: bosonic modes could be observed by measuring dynamic structure factor (DSF) in a
99: Bragg scattering experiment.
100: Bragg spectroscopy of a trapped atomic system has proven to be an important
101: tool for probing many bulk properties such as dynamic structure factor \cite{phonon},
102: verification of multibranch Bogoliubov excitation spectrum in the usual
103: atomic BEC \cite{mbs2}, correlation functions and momentum distributions of a phase
104: fluctuating Bose gases \cite{ric,gerbier}.
105: There has been a suggestion of Bragg scattering experiment to study the
106: fermionic excitations at zero temperature \cite{deb} as well as at temperature
107: close to the critical temperature \cite{baym}.
108:
109: In this work, we also calculate DSF in the various regimes of the crossover, namely
110: the weak-coupling BCS, unitarity and molecular BEC regimes. In actual Bragg scattering
111: experiments, the measured response function is momentum transferred
112: $ P_z(t) $ to the superfluid by the Bragg pulses. The momentum
113: transferred is directly related to the dynamic structure factor when time duration of
114: the Bragg pulses is long enough. Therefore, we also study momentum transferred to the
115: superfluid by the Bragg pulses.
116: In order to probe the multibranch nature of the modes, we also estimate required
117: values of wave vector and time duration of the two-photon Bragg pulses in the Bragg
118: scattering experiments.
119:
120: Recently, there is a measurement of sound velocity along the crossover \cite{sound}.
121: The measured sound velocity do not match with the theoretical prediction in
122: the molecular BEC regime, however it matches very well in other regimes of the crossover
123: \cite{tkgsound}.
124: One can also estimate the sound velocity in the three different regimes along the
125: crossover by measuring slope of the phonon mode.
126: It may resolve the puzzle of mismatch of the measured sound velocity with that of
127: the theoretical prediction in the molecular BEC state.
128:
129:
130: This paper is organized as follows. In section II, we provide quantized hydrodynamic
131: description of the Fermi superfluid along the crossover. In section III, we calculate
132: the dynamic structure factor along the crossover. In section IV, we discuss the
133: possible Bragg scattering experiment in this system and present results of the momentum
134: transferred to the system due to the Bragg pulses.
135: We also discuss a summary and conclusions in section V.
136:
137: \section{quantized hydrodynamic theory of Fermi superfluid}
138: We use hydrodynamic model with the Weizsacker quantum pressure term
139: to describe low-energy dynamics of a two-component Fermi superfluid at
140: zero temperature. This system can be well described by a time-dependent
141: non-linear Schrodinger equation as follows \cite{kim}:
142: \be \label{sch}
143: i\hbar \frac{\partial \hat \psi ({\bf r},t)}{\partial t} =
144: [- \frac{\hbar^2}{2M} \nabla^2 + V_{\rm ext}(r,z) + \bar \mu(n)]
145: \hat \psi ({\bf r},t),
146: \ee
147: where the non-linear term
148: $ \bar \mu(n) = \frac{\partial}{\partial n} [n \epsilon (n)] $ is the
149: chemical potential in a uniform system and $ \epsilon (n) $ is the
150: ground state energy per particle. Here, $ M $ is the mass of a Fermi
151: atom and $ V_{\rm ext}(r,z) = (M/2)(\omega_{r}^2 r^2 +
152: \omega_z^2 z^2) $ is an external harmonic trap potential with
153: $ \omega_{r} >> \omega_z $. We have taken $\omega_z/\omega_r = 0.1 $ and
154: total number of atoms $ N=2.0 \times 10^{6}$ in all numerical calculations.
155: On the basis of quantum Monte Carlo data of Astrakharchik {\em et al.}
156: \cite{astra},
157: Manini and Salasnich \cite{manini} proposed a very useful analytical
158: fitting expression for $ \epsilon (n) $ (see Eq. (6) and Table 1 of Ref.
159: \cite{manini}).
160: The Hamiltonian corresponds to Eq. (\ref{sch}) can be written as
161: $$
162: H = \int d {\bf r} \hat \psi^*({\bf r},t)[- \frac{\hbar^2}{2M} \nabla^2
163: + V_{\rm ext}(r,z) + \bar \mu(n)] \hat \psi ({\bf r},t).
164: $$
165: Using phase ($\theta $)-density ($n$) representation of the order parameter of
166: the composite bosons:
167: $ \hat \psi ({\bf r},t) = \sqrt{\hat n ({\bf r},t)} e^{i\hat \theta ({\bf r},t)} $,
168: the above Hamiltonian becomes,
169:
170: \bearr
171: \hat H & = & \int d {\bf r}[\sqrt{\hat n}(- \frac{\hbar^2}{2M} \nabla^2 \sqrt{\hat n})
172: + \frac{1}{2} M {\hat {\bf v}}^2 n + V_{\rm ext}(r,z) \hat n \nonumber \\
173: & + & \bar \mu(n) \hat n].
174: \eearr
175:
176: Linearizing the density and phase around their equilibrium values:
177: $ \hat n({\bf r},t) = n_0({\bf r}) + \delta \hat n({\bf r},t) $ and
178: $ {\hat {\bf v}}({\bf r},t) = \delta {\hat {\bf v}}({\bf r},t) $,
179: where $ {\hat {\bf v}} = (\hbar/M) \nabla \hat \theta $ is the superfluid
180: velocity.
181: Keeping upto quadratic fluctuations, the above Hamiltonian reads
182:
183: \be
184: H = H_0 + \int d{\bf r} [\frac{M \delta {\hat {\bf v}}^2}{2}
185: + \frac{\partial \bar \mu}{\partial n}|_{n_0} \delta \hat n^2],
186: \ee
187: where $ H_0 $ is the ground state energy. By using time-dependent
188: Heisenberg equations of motion for the density and velocity fluctuations,
189: one can get continuity and Euler's equations which are given by
190:
191: \be \label{den}
192: \delta \dot {\hat n} = - \nabla \cdot[n_0(r) \delta \hat {\bf v}],
193: \ee
194: and
195: \be \label{vel}
196: M \delta \dot {\hat {\bf v}} = - \nabla [ \frac{\partial \bar \mu}
197: {\partial n}|_{n_0} \hat \delta n].
198: \ee
199:
200: Since the Hamiltonian is quadratic in terms of the fluctuation operators,
201: it can be diagonalized
202: by using the following standard canonical transformations:
203: \be \label{dencan}
204: \delta \hat n (r,z,t) = \sum_{j,k} [ A_{j,k} \psi_{j,k}(r)
205: \hat \alpha_{j,k}
206: e^{i(kz- \omega_j(k) t)} + h. c.],
207: \ee
208: and
209: \be
210: \delta \hat \theta (r,z,t) = \sum_{j,k} [ B_{j,k} \psi_{j,k}(r) \hat \alpha_{j,k}
211: e^{i(kz- \omega_j(k) t)} + h. c.].
212: \ee
213: Here, $j$ is a set of two quantum numbers: radial quantum number,
214: $n_r$ and the angular quantum number, $m$. Also, $ k$ is the axial wave vector.
215: The density and phase fluctuations satisfy the following equal-time commutator
216: relation: $ [\delta \hat n(r), \delta \hat \theta(r^{\prime})]
217: = i \delta (r - r^{\prime}) $.
218: One can easily show that
219: \be
220: A_{j,k} = i \sqrt{\frac{\hbar \omega_j(k)}{2 \frac{\partial \bar \mu}
221: {\partial n}|_{n_0}}},
222: B_{j,k} = \sqrt{\frac{2 \frac{\partial \bar \mu}{\partial n}|_{n_0}}
223: {\hbar \omega_j(k)}}
224: \ee
225: We are assuming that $ \psi_{j,k}(r) $ satisfies the orthonormal
226: conditions:
227: $ \int d {\bf r} \psi_{j,k}^*(r) \psi_{j^{\prime},k}(r) =
228: \delta_{jj^{\prime}} $
229: and $ \sum_j \psi_{j,k}(r) \psi_{j,k}^*({ r}^{\prime}) =
230: \delta (r - r^{\prime}) $.
231:
232: We assume power-law form of the equation of state as
233: $ \bar \mu(n) = C n^{\gamma} $. Here, $C$ depends on interaction strength and
234: the effective polytropic index $ \gamma $ is a function of a dimensionless parameter
235: $ y = 1/k_F a$, where $k_F$ is the Fermi wave vector and $ a $ is the scattering
236: length between Fermi atoms of different components. The
237: weak-coupling BCS ($y<<-1$)
238: and the unitarity ($y=0$) states are described by the same exponent $ \gamma = 2/3 $ with
239: different values of $C$. For the weak-coupling BCS regime,
240: $ C_{\rm bcs} \simeq (3\pi^2)^{2/3}(3\hbar^2/10M) $ and
241: for the unitarity regime, $C_{\rm uni} = 0.44 C_{\rm bcs} $.
242: The molecular
243: BEC state ($y>>1$) is described by $\gamma = 1$ and
244: $ C_{\rm bec} = 4 \pi \hbar^2 a_m/M $, where
245: $ a_m = 0.6 a$ is the molecular scattering length \cite{gvs}.
246: In our calculation we have
247: assumed $ a_m = 1.0 \times 10^{-8} m $.
248: The power-law form of the equation of state is being used successfully to study the
249: Fermi superfluid along the crossover \cite{manini,poly1,poly2,bulgac,astra1}.
250: At equilibrium, the density profile takes the form
251: $ n_0(r) = (\mu/C)^{1/\gamma}( 1- \tilde r^2)^{1/\gamma} $,
252: where $ \tilde r = r/R_{0} $,
253: $ R_{0} = \sqrt{2 \mu/M \omega_{r}^2} $ and $ \mu $ is the chemical potential
254: in the non-uniform system, which can be obtained from the normalization condition.
255:
256:
257: Taking first-order time-derivative of Eq. (\ref{den}) and using Eq. (\ref{vel}),
258: the second-order equation of motion for the density fluctuation is
259: given by
260: \be \label{den0}
261: \frac{\partial^2 \delta n}{\partial t^2} =
262: \nabla \cdot [n_0(r) \nabla \frac{\partial \bar \mu(n) }{\partial n}|_{n =n_0}
263: \delta n ].
264: \ee
265:
266: Using the polytropic form of the equation of state and Eq. (\ref{dencan}),
267: then Eq. (\ref{den0}) reduces to the following equation:
268:
269: \bearr \label{den1}
270: - \tilde \omega_{j}^2(k) \psi_{j,k}(r) & = & [\frac{\gamma}{2} \nabla_{\tilde r}
271: \cdot [(1- \tilde r^2)^{1/\gamma}
272: \nabla_{\tilde r}(1- \tilde r^2)^{1-1/\gamma}] \nonumber \\
273: & - & \frac{\gamma}{2} \tilde k^2 (1-\tilde r^2)] \psi_{j,k}(r),
274: \eearr
275: where $ \tilde \omega = \omega/\omega_r $ and $ \tilde k = k R_{0} $.
276:
277: For $ k = 0 $, it reduces to a two-dimensional eigenvalue problem and the solutions
278: of it can be obtained analytically. The energy spectrum is
279: given by
280: \be
281: \tilde \omega_{j}^2 = |m| + 2 n_r [\gamma(n_r + |m|) + 1].
282: \ee
283: The corresponding orthogonal eigenfunction is given by
284: \be
285: \psi_{j} \propto (1-\tilde r^2)^{1/\gamma -1} \tilde r^{|m|}
286: P_{n_r}^{(1/\gamma -1, |m|)} (2\tilde r^2 -1) e^{im\phi},
287: \ee
288: where $ P_{n}^{(a,b)}(x) $ is a Jacobi polynomial of order $n$ and $\phi $ is
289: the polar angle.
290:
291: The solution of Eq. (\ref{den1})
292: can be obtained for arbitrary value of $k$ by
293: numerical diagonalization.
294: For $ k \neq 0 $, we expand the density fluctuation as
295: \be
296: \psi_{j,k}(r,\phi) = \sum_{j} b_{j} \psi_{j}(r,\phi).
297: \ee
298:
299: Substituting the above expansion into Eq. (\ref{den1}), we obtain,
300: \bearr \label{density2}
301: 0 & = & [\tilde \omega_{j}(k)^2 - [|m| + 2 n_r ( \gamma (n_r +|m|) +1)]
302: \nonumber \\ & - &
303: \frac{\gamma}{2} \tilde k^2] b_{j} +
304: \frac{\gamma}{2} \tilde k^2 \sum_{j^{\prime}} M_{j j^{\prime}}
305: b_{j^{\prime}}.
306: \eearr
307: Here, the matrix element $ M_{j j^{\prime}} $ is given by
308: \be \label{matrix}
309: M_{j j^{\prime}} = \int d^2 \tilde r \psi_{j}
310: \tilde r^2 \psi_{j^{\prime}}.
311: \ee
312: The above eigenvalue problem (Eq. (\ref{density2})) is block diagonal with
313: no overlap between the subspaces of different angular momentum, so that the
314: solutions to Eq.(\ref{density2}) can be obtained separately in each angular
315: momentum subspace. We can obtain all low energy multibranch spectrum on the
316: both sides of the Feshbach resonance including the unitarity
317: limit from Eq. (\ref{density2}). Equations (\ref{density2}) and (\ref{matrix})
318: show that the spectrum depends on the average over the radial coordinate and the
319: coupling between the axial mode and transverse modes within a given angular
320: momentum symmetry. Particularly, the coupling is important for large values of
321: $k $.
322: We are interested to study $m=0$ states since these states are
323: excited in the Bragg scattering experiments due to axial symmetry of the system.
324: We show low energy multibranch modes of $m=0$ states in three different regimes
325: in Fig. 1. The
326: top three panels of Fig. 1 show the multibranch spectrum in the three different
327: regimes. To compare these spectrum, we have plotted all those spectrum of
328: different regimes in a single frame, which is shown in the bottom panel of Fig. 1.
329: The lowest branch corresponds to the Bogoliubov axial mode with no radial nodes.
330: This mode has the usual form $ \omega = c_s k $ at low momenta, where
331: $c_s $ is the sound velocity. In the limit of small $k$, the other branches have
332: free-particle dispersion due to the gaped nature of these modes.
333:
334: The discrete radial and axial modes are same in the unitarity and BCS regimes
335: since the exponent $ \gamma $ are the same for both the regimes. However, the
336: multibranch modes are different in the unitarity and BCS regimes in spite of
337: the same exponent in the equation of state. This is due to the different radial
338: sizes in those two regimes for a given number of atoms and the trap potential.
339: Therefore, these low energy axial propagation of discrete radial modes can be
340: used to characterize different regimes of the superfluid.
341: The density fluctuations for a fixed value of the axial momentum are plotted in
342: Fig. 2. The density fluctuations corresponds to the multibranch modes are also
343: different in the different regimes. In Fig. 2, the magnitude of the density fluctuations
344: are given in an arbitrary unit since these are the linear fluctuations.
345: \begin{figure}[ht]
346: \includegraphics[width=8.0cm]{specall.eps}
347: \caption{(Color online) Plots of the low energy multibranch modes in the three different
348: regimes.}
349: \end{figure}
350:
351:
352: \section{dynamic structure factor}
353: The dynamic structure factor is the Fourier transformation of density-density
354: correlation functions and it is given as
355:
356: $$
357: S(k,\omega) = \int d {\bf r} d {\bf r}^{\prime} dt
358: <\delta \hat n^{\dag}({\bf r},t) \delta \hat n ({\bf r}^{\prime},0)>
359: e^{i{\bf k} \cdot ({\bf r} - {\bf r}^{\prime})} e^{i\omega t}
360: $$
361:
362: This can be written as
363: \be \label{delta}
364: S(k,\omega) = \sum_j S_{j}(k) \delta (\omega - \omega_j(k)),
365: \ee
366: where the weight factor $ S_{j}(k) $ is given by
367: \be
368: S_{j}(k) = \frac{\hbar \omega_j(k)}{2 \frac{\partial \bar \mu}{\partial n}|_{n_0} }
369: |\psi_{j}(k)|^2.
370: \ee
371: Here,
372: $ \psi_j(k) = \int d {\bf r} e^{-i{\bf k} \cdot {\bf r}} \psi_{j,k}({r})$ is the
373: Fourier transform of the eigenfunctions $ \psi_{j,k}({r})$.
374: The weight factors are plotted in Fig. 3.
375: The weight factors $ S_{j}(k) $
376: determine how many modes are excited for a given value of $k$.
377: For example, when $ ka_{r} = 1.0 $, $ n_r $ = \{0, 1\}, $ n_r $ = \{0, 1, 2 \} and
378: $ n_r $ = \{0, 1, 2, 3\} modes are excited in the molecular BEC, unitarity and
379: weak-coupling BCS regimes, respectively.
380: The harmonic oscillator length is defined as $ a_r = \sqrt{\hbar/M \omega_r} $.
381: To excite many other low-energy modes, the wave vector in the Bragg potential must be large.
382: From Fig. 3, it is clear that $ S_0 (k) \sim k $ and $ S_{n_r>0} (k) $ is almost constant
383: in the limit of small $k$. Therefore, $ \omega_0 (k) \sim k $ and
384: $ \omega_{n_r>0} (k) \sim k^2 $ when $k $ is very small.
385: Therefore, our analysis also satisfies the Feynman-like relation
386: $\omega_{n_r}(k) \sim k^2/S_{n_r}(k) $.
387: The dynamic structure factors of the three different regimes for $ k a_{r} = 0.5 $ are
388: plotted in Fig. 4. The delta function in Eq. (\ref{delta}) is replaced by the Lorentzian form to
389: plot Fig. 4.
390: \begin{figure}[ht]
391: \includegraphics[width=8.0cm]{wavefcom.eps}
392: \caption{(Color online) Plots of the low-energy density fluctuations in the three different
393: regimes
394: for $ k a_{r} = 0.5 $.}
395: \end{figure}
396:
397:
398: \begin{figure}[ht]
399: \includegraphics[width=8.0cm]{wfcomnew.eps}
400: \caption{(Color online) Plots of the weight factors in the three different regimes.}
401: \end{figure}
402:
403: \begin{figure}[ht]
404: \includegraphics[width=8.0cm]{dsfcom05.eps}
405: \caption{(Color online) Plots of the dynamic structure factors of the three different
406: regimes for $ k a_{r} = 0.5$.}
407: \end{figure}
408:
409: \section{Bragg scattering experiment}
410: The behavior of these multiple peaks in the dynamic structure factor can
411: be resolved in a two-photon Bragg spectroscopy, as shown by Steinhauer
412: {\em et al.} \cite{davidson} for usual BEC. In the two-photon Bragg spectroscopy, the
413: dynamic structure factor can not be measured directly. Actually, the observable
414: in the Bragg scattering experiments is the momentum transferred to the
415: superfluid, which is related to the dynamic structure factor and reflects
416: the behavior of the quasiparticle energy spectrum.
417: The populations in the quasiparticle states can be controlled by using
418: the two-photon Bragg pulse. When the superfluid is irradiated by an external
419: moving optical potential $ V_{\rm op} = V_B(t) \cos(qz-\omega t)$, the excited states
420: are populated by the quasiparticle with energy $ \hbar \omega $ and the momentum
421: $\hbar q$, depending on the value of $q$ and $\omega$ of the optical potential
422: $ V_{\rm op} $. Here, $ V_B $ is the intensity of the Bragg pulse.
423: Suppose the system is subjected to a time-dependent Bragg pulse which is switched
424: on at time $t>0$ and $q$ is also along the $z$-direction.
425: We calculate, similar to the calculation of Refs. \cite{tkg,blak}, the momentum
426: transfer to the superfluid from the moving optical potential
427: and it is given by
428: \bearr
429: P_z(t) & = & \sum_{j,k} \hbar k < \hat \alpha_{j,k}^{\dag} (t) \hat \alpha_{j,k} (t) >
430: = \l (\frac{V_B(t)}{2 \hbar} \r )^2 \nonumber \\
431: & \times & \sum_j \hbar q S_j (\tilde q)
432: \times [F_j (\omega_{-}t) - F_j (\omega_{+}t)],
433: \eearr
434: where $ \hat \alpha_{j,k} (t) $ is the time-evolution of the quasiparticle operator
435: of energy $\hbar \omega_j(k) $ and
436: \be
437: F_j (\omega_{\pm}t) =
438: \l (\frac{\sin[(\omega_j(q) \pm \omega)t/2]}{(\omega_j(q) \pm \omega)/2} \r )^2.
439: \ee
440: For positive $ \omega $ and a given $ \tilde q $ such that $ S_j(\tilde q) $ is maximum,
441: the momentum transferred $ P_z(t)$ is resonant at the frequencies $ \omega = \omega_j(q) $.
442: The width of the each peak goes like $ 2 \pi/t $.
443: For large $t$ and $ \omega_z << \omega_{r} $, one can show that
444: $P_z(t) \sim S(k,\omega)$ \cite{tozo}.
445:
446: In Fig. 5, we plot the net momentum transfer $P_z(t)$ vs the Bragg frequency
447: $ \omega $ for three different choices of the time duration of the Bragg pulses.
448: Figure 5 shows that the shape of the $P_z(t)$ strongly depends on the time duration
449: of the Bragg pulses $ t_B$.
450: When $ t_B = 0.5 T_r $, $P_z(t)$ is a smooth curve with a single peak.
451: Here, we define $ T_{r} = 2 \pi/\omega_{r}$ is radial trapping period.
452: When $t_B = 1.0 T_r $, there is a little evidence of few small peaks start developing in
453: $P_z(t)$.
454: When $ t_B = 2.0 T_r $, the multiple peaks in $P_z(t)$ appears prominently.
455: Therefore, the duration of the Bragg pulses $t_B $ should be greater than the radial
456: trapping period $T_r$ in order to resolve different peaks in the DSF.
457: Figure 5 also shows that the number of quasiparticle modes in unitarity regime is
458: much higher than the other regimes of the crossover. This is due to the fact
459: that $ P_z(t) $ is proportional to the number of quasiparticle modes for a given $k$.
460: \begin{figure}[ht]
461: \includegraphics[width=8.0cm]{timecom.eps}
462: \caption{(Color online) Plots of the momentum transferred
463: in the three different regimes for $k a_{r} = 0.5$ and for different
464: time duration of the Bragg pulses: $t=0.5 T_r$ (solid),
465: $t=1.0 T_r$ (dashed) and $t=2.0 T_r$ (dot-dashed).}
466: \end{figure}
467:
468:
469: \section{summary and conclusions}
470: We have presented the quantized hydrodynamic theory of cigar shaped Fermi
471: superfluid along the BEC-BCS crossover by using the power law form of the
472: equation of state. We have calculated multibranch low energy bosonic modes and
473: the corresponding density fluctuations in three different regimes along the BEC-BCS
474: crossover, namely the weak-coupling BCS, unitarity and molecular BEC states. Then
475: we have presented results of the dynamic structure factor calculation. We have also
476: calculated the momentum transferred to the superfluid by the Bragg pulses and shown
477: that the multibranch nature would be observed when the time duration of the Bragg pulses
478: is greater than the radial trapping period.
479:
480: We have found that the axial propagation of discrete radial modes in the weak-coupling
481: BCS and unitarity regimes are different, although the axial and radial modes
482: are same in both cases. One can identify these two regimes by probing these multibranch
483: modes with the help of the Bragg scattering experiment.
484: We have also seen that the number of quasi particle modes in the unitarity regime is quite
485: large compared to other two regimes for a fixed value of the Bragg momentum and the number
486: of atoms. Therefore, the response function in the unitarity regime will be more prominent
487: than the other two regimes in the crossover.
488: Moreover, one can estimate the sound velocities in different regimes along the crossover by
489: measuring the slope of the phonon modes.
490:
491:
492: \begin{acknowledgments}
493: This work was supported by the Alexander von Humboldt foundation, Germany.
494: \end{acknowledgments}
495:
496: \begin{thebibliography}{2}
497: \bibitem{greiner}
498: M. Greiner, C. A. Regal, and D. S. Jin, Nature {\bf 426}, 537 (2003).
499:
500: \bibitem{jochim}
501: S. Jochim, M. Bartenstein, A. Altmeyer, G. Hendl, S. Riedl,
502: C. Chin, J. H. Denschlag, and R. Grimm, Science {\bf 302}, 2101 (2003).
503:
504: \bibitem{zw}
505: M. W. Zwierlein, C. A. Stan, C. H. Schunck, S. M. F. Raupach, S. Gupta,
506: Z. Hadzibabic, and W. Ketterle, Phys. Rev. Lett. {\bf 91}, 250401 (1993).
507:
508: \bibitem{regal}
509: C. A. Regal, M. Greiner, and D. S. Jin, Phys. Rev. Lett. {\bf 92}, 040403 (2004).
510:
511: \bibitem{barten}
512: M. Bartenstein, A. Altmeyer, S. Riedl, S. Jochim, C. Chin, J. H.
513: Denschlag, and R. Grimm, Phys. Rev. Lett. {\bf 92}, 120401 (2004).
514:
515: \bibitem{bourdel}
516: T. Bourdel, L. Khaykovich, J. Cubizolles, J. Zhang, F. Chevy,
517: M. Teichmann, L. Tarruell, S. J. J. M. F. Kokkelmans, and
518: C. Salomon, Phys. Rev. Lett. {\bf 93}, 050401 (2004).
519:
520: \bibitem{chin}
521: C. Chin, M. Bartenstein, A. Altmeyer, S. Riedl, S. Jochim, J. H. Denschlag,
522: and R. Grimm, Science {\bf 305}, 1128 (2004).
523:
524: \bibitem{hara}
525: K. M. O'Hara, S. L. Hemmer, M. E. Gehm, S. R. Grande,
526: and J. E. Thomas, Science {\bf 298}, 217 (2002).
527:
528: \bibitem{houb}
529: M. Houbiers, H. T. C. Stoof, W. I. McAlexander, and R. G. Hulet,
530: Phys. Rev. A {\bf 57}, R1497 (1998).
531:
532: \bibitem{stwa}
533: W. C. Stwalley, Phys. Rev. Lett. {\bf 37}, 1628 (1976).
534:
535: \bibitem{ties}
536: E. Tiesinga, B. J. Verhaar, and H. T. C. Stoof, Phys. Rev. A {\bf 47},
537: 4114 (1993).
538:
539: \bibitem{freq1}
540: J. Kinast, S. L. Hemmer, M. E. Gehm, A. Turlapov, and J. E. Thomas,
541: Phys. Rev. Lett. {\bf 92}, 150402 (2004).
542:
543: \bibitem{freq2}
544: M. Bartestein, A. Altmeyer, S. Riedl, S. Jochim, C. Chin, J. H. Denschlag, and
545: R. Grimm, Phys. Rev. Lett. {\bf 92}, 203201 (2004).
546:
547: \bibitem{zaremba}
548: E. Zaremba, Phys. Rev. A {\bf 57}, 518 (1998).
549:
550: \bibitem{phonon}
551: J. Stenger, S. Inouye, A. P. Chikkatur, D. M. Stamper-Kurn, D. E. Pritchard, and W.
552: Ketterle,
553: Phys. Rev. Lett. {\bf 82}, 4569 (1999);
554: D. M. Stamper-Kurn, A. P. Chikkatur, A. Gorlitz, S. Inouye, S. Gupta,
555: D. E. Pritchard, and W. Ketterle, Phys. Rev. Lett. {\bf 83}, 2876 (1999).
556:
557: \bibitem{mbs2}
558: J. Steinhauer, N. Katz, R. Orezi, N. Davidson,
559: C. Tozzo and F. Dalfovo, Phys. Rev. Lett. {\bf 90}, 060404 (2003).
560:
561: \bibitem{ric}
562: S. Richard, F. Gerbier, J. H. Thywissen, M. Hugbart, P. Bouyer, and A. Aspect,
563: Phys. Rev. Lett. {\bf 91}, 010405 (2003).
564:
565: \bibitem{gerbier}
566: F. Gerbier, J. H. Thywissen, S. Richard, M. Hugbart, P. Bouyer, and A. Aspect,
567: Phys. Rev. A {\bf 67}, 051602 (2003).
568:
569: \bibitem{deb}
570: B. Deb, J. Phys. B {\bf 39}, 529 (2006).
571:
572: \bibitem{baym}
573: G. M. Bruun and G. Baym, Phys. Rev. A {\bf 74}, 033623 (2006).
574:
575: \bibitem{sound}
576: J. Joseph, B. Clancy, L. Luo, J. Kinast, A. Turlapov, and J. E. Thomas,
577: Phys. Rev. Lett. {\bf 98}, 170401 (2007).
578:
579: \bibitem{tkgsound}
580: T. K. Ghosh and K. Machida, Phys. Rev. A {\bf 73}, 013613 (2006).
581:
582: \bibitem{kim}
583: Y. E. Kim and A. L. Zubarev, Phys. Rev. A {\bf 70}, 033612 (2004).
584:
585: \bibitem{astra}
586: G. E. Astrakharchik, J. Boronat, J. Casulleras, and S. Giorgini, Phys. Rev. Lett. {\bf 93},
587: 200404 (2004).
588:
589: \bibitem{manini}
590: N. Manini and L. Salasnich, Phys. Rev. A {\bf 71}, 033625 (2005).
591:
592: \bibitem{gvs}
593: D. S. Petrov, C. Salomon, and G. V. Shlyapnikov, Phys. Rev. Lett. {\bf 93},
594: 090404 (2004).
595:
596: \bibitem{poly1}
597: H. Heiselberg, Phys. Rev. Lett. {\bf 93}, 040402 (2004).
598:
599: \bibitem{poly2}
600: H. Hu, A. Minguzzi, X. J. Liu, and M. P. Tosi, Phys. Rev. Lett. {\bf 93},
601: 190403 (2004).
602:
603: \bibitem{bulgac}
604: A. Bulgac and G. F. Bertsch, Phys. Rev. Lett. {\bf 94}, 070401 (2005).
605:
606: \bibitem{astra1}
607: G. E. Astrakharchik, R. Combescot, X. Leyronas, and S. Stringari,
608: Phys. Rev. Lett. {\bf 95}, 030404 (2005).
609:
610: \bibitem{davidson}
611: J. Steinhauer, N. Katz, R. Ozeri, N. Davidson, C. Tozzo, and F. Dalfovo,
612: Phys. Rev. Lett. {\bf 90}, 060404 (2003).
613:
614: \bibitem{tkg}
615: T. K. Ghosh, Int. J. Mod. Phys. B {\bf 20}, 5443 (2006).
616:
617: \bibitem{blak}
618: P. B. Blakie, R. J. Ballagh, and C. W. Gardiner, Phys. Rev. A {\bf 65},
619: 033602 (2002).
620:
621: \bibitem{tozo}
622: C. Tozzo and F. Dalfovo, New J. Phys. {\bf 5}, 54 (2003).
623:
624: \end{thebibliography}
625:
626: \end{document}
627:
628: