cond-mat0601461/p3.tex
1: \documentclass[twocolumn,showpacs,amsmath,amssymb,prl]{revtex4}
2: %\documentclass[preprint,showpacs,amsmath,amssymb,prl]{revtex4}
3: 
4: \usepackage{epsfig}
5: 
6: \begin{document}
7: \title{Pairing symmetry in the anisotropic Fermi superfluid
8: under $p$-wave Feshbach Resonance}
9: \author{Chi-Ho Cheng and Sung-Kit Yip}
10: \affiliation{ Institute of Physics, Academia Sinica, Taipei,
11: Taiwan}
12: 
13: \date{\today}
14: 
15: 
16: \begin{abstract}
17: The anisotropic Fermi superfluid of ultra-cold Fermi atoms under
18: the $p$-wave Feshbach resonance is studied theoretically. The
19: pairing symmetry of the ground state is determined by the strength of
20: the atom-atom magnetic dipole interaction. It is $k_z$
21: for a strong dipole interaction; while it becomes $k_z - i
22: \beta k_y$, up to a rotation about $\hat z$, for a weak one
23: (Here $\beta < 1$ is a numerical coefficient).
24: By changing the external magnetic field
25: or the atomic gas density, a phase transition between these two
26: states can be driven. We discuss how the pairing symmetry of
27: the ground state can be determined in the time-of-flight
28: experiments.
29: \end{abstract}
30: 
31: \pacs{03.75.Ss, 05.30.Fk, 34.90.+q}
32: 
33: \maketitle
34: 
35: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
36: 
37: \section{I. Introduction}
38: 
39: The trapped Bose or Fermi atomic gases
40: \cite{Dalfovo99,Castin01} are usually weakly
41: interacting since the inter-particle distances between the atoms are
42: typically
43: much larger than their scattering lengths.
44: However, it has been recognized that the
45: inter-particle interaction can in fact be tuned via Feshbach
46: resonances \cite{Stwalley76,Tiesinga92}. The interaction changes from weakly
47: to strongly attractive by varying the external magnetic field
48: across the Feshbach resonance. For a two-component Fermi gas,
49: the ground state evolves from a Bardeen-Cooper-Schrieffer (BCS)
50:  superfluid with
51: long-ranged (compared with inter-particle distances) Cooper pairing
52: to a Bose-Einstein condensate (BEC) of tightly bound pairs
53: \cite{cross0,nozieres,randeria90}.
54: This cross-over has recently been a subject of
55: intensive theoretical
56: \cite{ohashi,falco04,perali,java,numeric}
57: and experimental
58: \cite{regal,bourdel,zwierlein,bartenstein,kinast}
59: investigations.
60: 
61: Most previous investigations deal with $s$-wave Feshbach
62: resonance. Therefore, both the Cooper pairs and the bound states
63: between two fermions have $s$-wave symmetries. Recent experiments
64: demonstrated $p$-wave Feshbach resonances
65: \cite{Regal03,Ticknor04,Zhang04,Gunter05}. It thus raises the
66: possibility of $p$-wave Fermi superfluid and BEC of $p$-wave
67: ``molecules" \cite{klinkhamer,ohashi05,ho05} (see also
68: \cite{alt}).
69: 
70: To begin, we first recall the well-known superfluid
71: $^3$He \cite{Leggett75}. $^3$He has (nuclear) spin $1/2$, in
72: general {\it not} spin-polarized and basically spatially
73: isotropic, that is, the interaction between two atoms
74: is dependent of the orientation of their relative
75: distance.
76:  Back in the 60's, Anderson and Morel \cite{Anderson61}
77: investigated theoretically the superfluid state for this system by
78: assuming that the pairing exists only between the same (say
79: ${\uparrow}$) species. They showed that the ground state corresponds to Cooper
80: pairing in the $l = 1$, $m = 1$ channel, that is, the pairing
81: wavefunction has the symmetry of the spherical harmonic
82: $Y_{11}(\hat k) \propto (\hat k_x + i \hat k_y)$
83:  (or its spatial rotation, such as $\hat k_z - i \hat k_y$).
84:  This orbital symmetry is
85: realized in the $^3$He A-phase and is known as the ``axial" state
86: \cite{Leggett75}. One can thus suspect that for a
87: spin-polarized but otherwise spatially isotropic system, the
88: pairing is again in the $Y_{11}(\hat k)$ state.
89: This has been confirmed in \cite{ho05}.
90: 
91: However, in the atomic gas system, as demonstrated and explained
92: by Ticknor {\it et al.} \cite{Ticknor04} in the context of
93: $^{40}{\rm K}$,
94:  the magnetic dipole interaction, which breaks
95: rotational symmetry, may be important.
96: (An Alkali atom has one additional electron beyond
97: closed electronic shell(s), hence it
98: possesses a magnetic moment mainly due to
99: this extra electron, though the precise value of
100: this moment depends on the hyperfine
101: interaction and the external magnetic field).
102: They show that, due to this
103: magnetic dipole  interaction,
104:  for magnetic field along $\hat z$, the $l=1$, $m=0$
105: resonance occurs at a higher magnetic field than the $m = {\pm} 1$
106: ones (see also \cite{Gunter05}) For a given magnetic field, the
107: induced effective interaction is anisotropic. In our previous
108: paper \cite{Cheng05} (see also \cite{gurarie}), we discussed
109: theoretically the expected ground state order parameter under this
110: circumstance. Since the interaction of the $m = 0$ channel is more
111: attractive than that of the $m = {\pm} 1$, in the case that the $m
112: = 0$ and $m = {\pm} 1$ resonances are sufficiently far apart, we
113: showed that the pairing can only occur in the $m=0$ channel, and
114: the ground state becomes $Y_{10}(\hat k) \propto k_z$. The system
115: is either the BCS state with $k_z$ Cooper pairing or the BEC state
116: of bosonic ``molecules" in the $m=0$ channel, depending on the
117: detuning. Fermion pairing or bosons in $m=\pm 1$ channel do not
118: exist. The orbital symmetry of the pairing is the same as the
119: ``polar" phase in the $^3$He literature \cite{Leggett75}.
120: 
121: For the case that two resonances in $m=0$ and $m=\pm 1$ channels
122: are closed to each other, we showed  that the ground state
123: symmetry is $k_z - i \beta k_y$ up to rotation about $\hat z$,
124: where $\hat z$ is along the magnetic field direction and $\beta <
125: 1$. This state is thus intermediate between the axial $(\beta =
126: 1)$ the the polar $(\beta = 0)$ phases.
127: 
128: 
129: This paper provides some details of our earlier investigation \cite{Cheng05},
130: as well as some additional information.  It
131: is organized as follows. We first study the Cooper
132: pairing symmetry of a weak-coupling BCS model,
133: where the p-wave pairing interaction is anisotropic.
134: We show in that case how the analogous results cited above arise.
135: Then we return to the slightly more involved case of Feshbach resonance
136: and determine the general phase diagram for this case.
137: We first provide the results for the case where the system can be
138: described by an effective single channel model,
139: i.e., purely in terms of atoms interacting with each other
140: via a two-body instantaneous interaction.
141: Then we study the necessary modifications to these results
142: when this approximation is relaxed.
143: Finally a probe of the pairing symmetry based on the time-of-flight
144: experiment is discussed.
145: 
146: 
147: 
148: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
149: \section{II. Lessons from Anisotropic BCS Model}
150: 
151: In the spirit of the work of Anderson and Morel \cite{Anderson61},
152: we first consider a weak-coupling BCS model of
153:  a Fermi system with one spin species under the
154: interaction $-V_{\vec k - \vec k'}$.
155: (We shall however study the case where this $V$ is anisotropic,
156: see below).
157: The Hamiltonian of our system
158: is $H=H_f + H_V$, where
159: \begin{eqnarray}
160: H_f = \sum_{\vec k} (\epsilon_k-\mu )
161:  a_{\vec k}^{\dagger} a_{\vec k}
162: \end{eqnarray}
163: \begin{eqnarray}
164: H_V = - \sum_{\vec k,\vec k',\vec q} V_{\vec k-\vec k'}
165: a_{\vec q+\vec k'}^\dagger a_{\vec q-\vec
166: k'}^\dagger
167: a_{\vec q-\vec k} a_{\vec q+\vec k} \nonumber \\
168: \end{eqnarray}
169: Here $\epsilon_k = \hbar^2 k^2/2M$, and $\mu$ is the chemical potential,
170: and $a_{\vec k}$ is the annihilation operator for
171: a Fermion with wave-vector $\vec k$.
172:  Standard BCS theory proposes the
173: following trial ground state wavefunction
174: \begin{eqnarray}
175: |G\rangle = \prod_k (u_{\vec k} + v_{\vec k} a_{\vec
176: k}^\dagger a_{-\vec k}^\dagger) |0\rangle
177: \end{eqnarray}
178: with normalization condition $|u_{\vec k}|^2 + |v_{\vec k}|^2 =1$.
179: Minimizing the ground state energy $\langle G|H|G\rangle$ with
180: respect to $u_{\vec k}$ and $v_{\vec k}$ gives the excitation
181: spectrum
182: \begin{eqnarray}
183: E_{\vec k} = \sqrt{(\epsilon_k-\mu)^2+|\Delta_{\vec k}|^2}
184: \end{eqnarray}
185: where the energy gap $\Delta_{\vec k}$ satisfies
186: \begin{eqnarray} \label{gap1}
187: \Delta_{\vec k} = \sum_{\vec k'}V_{\vec k-\vec
188: k'}\frac{\Delta_{\vec k'}}{2E_{\vec k'}}
189: \end{eqnarray}
190: In case the interaction is $p$-wave, only the $l=1$ component
191: survive in $V_{\vec k-\vec k'}$. We write
192: \begin{eqnarray} \label{vk}
193: V_{\vec k-\vec k'} = 2\pi  \sum_{m=-1}^{m=1} V_{m}(k,k')
194: Y_{1m}(\hat k) Y_{1m}^*(\hat k')
195: \end{eqnarray}
196: We shall study the situation where
197: $V_1 = V_{-1} \le V_0$.
198: We also decompose the energy gap in spherical harmonics of $l=1$
199: partial waves,
200: \begin{eqnarray} \label{dk}
201: \Delta_{\vec k} &=& \Delta f(\hat k) \nonumber \\ &=& \Delta
202: \sum_{m=-1}^{m=1} f_{m} Y_{1m}(\hat k)
203: \end{eqnarray}
204: with normalization $\sum_{m} |f_{m}|^2 =1$. Substituting
205: Eqs.(\ref{vk})-(\ref{dk}) into Eq.(\ref{gap1}), we have
206: %\begin{widetext}
207: \begin{eqnarray} \label{gap2}
208: f(\hat k) = \pi \sum_{\vec k'} \sum_{m} V_{m}(k,k')Y_{1m}(\hat k)
209: \frac{Y_{1m}^*(\hat k')f(\hat k')}{[(\epsilon_{k'}-\mu)^2+\Delta^2
210: |f(\hat k')|^2]^\frac{1}{2}}
211:  \nonumber \\
212: \end{eqnarray}
213: %\end{widetext}
214: Suppose further that $V_{m}(k,k')=V_{m}$ is independent of $k$ and
215: $k'$, Eq.(\ref{gap2}) becomes
216: \begin{eqnarray}
217: f_{m} = \pi V_{m} \sum_{\vec k} \frac{Y_{1m}^*(\hat
218: k)f(\hat k)}{[(\epsilon_{k}-\mu)^2+\Delta^2 |f(\hat
219: k)|^2]^\frac{1}{2}}
220: \end{eqnarray}
221: We perform the integration in $\vec k$ in the following way,
222: \begin{eqnarray}
223: \sum_{\vec k} \rightarrow \rho_0 \int_{-\Lambda}^\Lambda d\epsilon
224: \int d\Omega
225: \end{eqnarray}
226: where $\rho_0$ is the density of state around the Fermi surface,
227: $\Lambda$ is the energy cutoff, and $d\Omega$ is solid angle. Then
228: we get totally 3 nonlinear equations,
229: %\begin{widetext}
230: \begin{eqnarray} \label{gm}
231: f_{m} = 2\pi \rho_0 V_{m} \left[
232:  f_{m} \log(\frac{2\Lambda}{\Delta})
233:  - \int d\Omega Y_{1m}^*(\hat k)f(\hat k)\log|f(\hat k)|
234:  \right] \nonumber \\
235: \end{eqnarray}
236: %\end{widetext}
237: for $m=-1,0,1$.
238: 
239:  We would like to find the solution
240: to Eq. (\ref{gm}) with the anisotropy
241:  characterized by the ratio $V_1/V_0$.
242: Particular solutions to
243: Eq.(\ref{gm}) can be easily found at two limits.
244: When $V_1/V_0=0$,
245:  $\Delta_{\vec k} \propto Y_{10}(\hat
246: k)\propto k_z$ because no other pairing can be possible.
247:  At the
248: other limit, $V_1/V_0=1$,
249:  the system is again isotropic the ground
250: state pairing is $\Delta_{\vec k} \propto Y_{11}(\hat k)$,
251: or its rotations.
252: In terms of the language by
253: $f_m$'s, for $V_1/V_0=0$,
254: $f_0=1, f_1=f_{-1}=0$ (up to a gauge transformation).
255: For $V_1/V_0=1$,  a particular solution is
256: $f_0=1/\sqrt{2}, f_1=f_{-1}=1/2$ such that
257: \begin{eqnarray}
258: \Delta_{\vec k} &\propto& \frac{1}{\sqrt{2}}Y_{10}(\hat k)
259:  + \frac{1}{2}(Y_{11}(\hat k) + Y_{1,-1}(\hat k)) \nonumber \\
260:  &\propto& k_z - i k_y
261: \end{eqnarray}
262: Note that it does not contradict with the pairing $Y_{11}(\hat
263: k)\propto k_x + i k_y$ since these states are degenerate in an isotropic
264: limit $V_1=V_{-1}=V_0$.
265: With
266: gradually tuning the parameter $0< V_1/V_0 < 1$, we will show
267: below that $\Delta(\hat k) \propto k_z - i\beta k_y$ up to a
268: $z$-axis rotation with $\beta < 1$.
269: 
270: 
271: Before solving the gap equation
272: Eq.(\ref{gm}) for the general intermediate values
273: of the pairing $0< V_1/V_0 < 1$, it is helpful to first
274: identify the critical $V_1^*/V_0$ such that $f_1$ and $f_{-1}$
275: start to deviate from zero (i.e., when the pairing starts to deviate from
276: $Y_{10}$).  Let us take the gauge where $f_0$ is real.
277: Linearizing Eq.(\ref{gm}) in $f_1$ and $f_{-1}$
278: around the point $(f_0,f_1,f_{-1})=(1,0,0)$, we have
279: \begin{eqnarray}
280: \frac{1}{2\pi\rho_0 V_1^*} \left( \begin{array}{c}
281:  f_1 \\ f_{-1}^* \end{array} \right) =
282: \left( \begin{array}{cc}
283: A_1 & A_2 \\
284: A_2 & A_1 \end{array} \right)
285:  \left( \begin{array}{c}
286:  f_1 \\ f_{-1}^* \end{array} \right)
287: \end{eqnarray}
288: where $A_1= \log(2\Lambda/\Delta) -\int d\Omega |Y_{11}|^2
289: \log|Y_{10}|-1/2$, $A_2=1/2$. Thus when we increase $V_1$, the
290: first non-trivial solution for $f_1$ and $f_{-1}$ is obtained when
291: $(f_1, f_{-1}^*)$ gives the  largest eigenvalue of the matrix.  It
292: can be easily seen that this corresponds to the solution $f_1 =
293: f_{-1}^*$, with eigenvalue $A_1 + A_2$.  The
294: critical value $V_1^{*}$ thus satisfies
295: \begin{equation}
296: \frac{1}{ 2\pi\rho_0 V_1^*} =  \left[ \log(\frac{2\Lambda}{\Delta})
297:  - \int d\Omega |Y_{11}|^2 \log|Y_{10}| \right]
298:  \\ \label{gm1}
299: \end{equation}
300: On the other hand, at this point (from Eq. (\ref{gm}) with $m=0$),
301: \begin{equation}
302: \frac{1}{ 2\pi\rho_0 V_0} = \left[ \log(\frac{2\Lambda}{\Delta})
303:  - \int d\Omega |Y_{10}|^2 \log|Y_{10}| \right]
304:  \label{gm2}
305: \end{equation}
306: Eliminating $\log(2\Lambda/\Delta)$ in Eqs.(\ref{gm1})-(\ref{gm2})
307: gives
308: \begin{eqnarray}
309: V_1^*/V_0 = \frac{1}{1+2\pi\rho_0 V_0}
310: \end{eqnarray}
311: 
312: For $V_1^*/V_0 < V_1/V_0 < 1$, $f_{\pm 1}$ becomes finite.
313: Provided no additional phase transitions occur,
314: we expect then the relation $f_1 = f_{-1}^*$ remains valid.
315: Furthermore,
316: our system is rotationally invariant around $\hat z$. Under this
317: rotation, $f_1 \to f_1 e^{i \phi}$ but $f_{-1} \to f_{-1} e^{- i
318: \phi}$ where $\phi$ is the rotational angle.  By suitable choice
319: of $\phi$ we can thus make both $f_1$ and $f_{-1} = f_1^*$ real
320: and positive. We shall confine ourselves to this case without loss
321: of generality.  Under this choice the order parameter has
322: the form $k_z - i \beta k_y$.
323: 
324: Hence, when $V_1^*/V_0 < V_1/V_0 < 1$, we search for the solution $f_0>0,
325: f_1=f_{-1}>0$ under the normalization condition $f_0^2 + 2 f_1^2 =
326: 1$. The solution $f_0$ (where $f_1=f_{-1}=\sqrt{(1-f_0^2)/2}$ )
327: for different $V_1/V_0$ and  $\rho_0 V_0$ is
328: presented in Fig.\ref{f0}. For given $\rho_0 V_0$,
329: the pairing is $k_z$ ($f_0=0$) at small $V_1/V_0$ until the latter
330: exceeds the critical $V_1^*/V_0$. Beyond this critical $V_1^*/V_0$,
331: $f_0$ decreases gradually with increasing $V_1/V_0$. At
332: $V_1/V_0=1$, $f_0=1/\sqrt{2}$ for all $\rho_0 V_0$'s meaning that
333: the pairing becomes $k_z - i k_y$ (corresponding to $\beta = 1$).
334: 
335: 
336: 
337: 
338: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
339: %\vspace{15pt}
340: \begin{figure}[tbh]
341: \begin{center}
342: \includegraphics[width=3in]{f0.eps}
343: \end{center}
344: \vspace{-5pt}
345:  \caption{$f_0$ as a function of $V_1/V_0$ for $\rho_0 V_0$'s
346: given in the legend.}
347:  \label{f0}
348:  \vspace{-5pt}
349: \end{figure}
350: 
351: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
352: 
353: 
354: 
355: 
356: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
357: \section{III. P-WAVE FESHBACH PAIRING}
358: 
359: We return to the case of p-wave Feshbach resonance.
360: We begin with the total Hamiltonian $H = H_f + H_b + H_{\alpha}$
361: where again $H_f$ is for the free fermions,
362: \begin{eqnarray}  \label{Hf}
363: H_f &=& \sum_{\vec k} (\epsilon_k  - \mu)
364:  a_{\vec k}^{\dagger} a_{\vec k} \ .
365: \end{eqnarray}
366: $H_b$ is for the molecules in the ``closed" channels, with $b_{\vec q, m}$
367: the annihilation operator for a Boson with angular momentum $l =
368: 1$ and $z$-axis projection $m$ with momentum $\vec q$,
369: \begin{eqnarray}
370: H_b &=& \sum_{\vec q} \left(
371:     \frac{\epsilon_q}{2} + \delta_m  - 2 \mu \right)
372:    b_{\vec q, m}^{\dagger} b_{\vec q, m}
373: \label{Hb}
374: \end{eqnarray}
375: Here $\delta_m$'s are the (bare) detuning of the $l=1$, $m$ resonance.
376: $H_\alpha$ represents the Feshbach coupling, which is a
377: generalization of the ones already commonly employed
378: \cite{ohashi,Fesh} for $s$-wave Feshbach
379: resonances to the case of several $l =1$, $m = 0, \pm 1$ closed
380: channels \cite{ohashi05,ho05}:
381: %\begin{widetext}
382: \begin{eqnarray} \label{Hal}
383: H_{\alpha} =
384:    \sum_{m,\vec q, \vec k}
385:    \varphi^*_m(\vec k)
386:       b_{\vec q, m}^\dagger a_{- \vec k + \vec q/2} a_{\vec k + \vec q/2}
387:    + {\rm h.c.}
388: \end{eqnarray}
389: %\end{widetext}
390: where $\varphi_m(\vec k)=-\frac{i \sqrt{ 4 \pi}}{L^{3/2}}  k Y_{1
391: m}(\hat k) \alpha_m$. The factor $Y_{1 m}(\hat
392: k)$ reflects the symmetry of the $l, m$ bound state and the linear
393: factor in $k$ arises from the small momentum approximation for the
394: coupling. $\alpha_m$ is the corresponding coupling constant,
395: it has dimension $[E][L]^{5/2}$.  More precisely $\alpha_m$ can be
396: an analytic function of $k$.  For low energy (density) phenomena,
397: we shall however only need its value $\alpha_m \equiv \alpha_m(k=0)$.
398: 
399: We shall apply mean-field theory to the Hamiltonian $H$.
400: In this approximation, we discard all terms that involve
401: finite momentum Bosons $b_{\vec q, m}$ with $\vec q \ne 0$,
402: and regard $b_{\vec q = 0, m}$ as c-numbers.
403: It is then convenient to introduce
404: \begin{eqnarray}
405: D_m &=& - \frac{i\sqrt{4\pi}}{L^\frac{3}{2}} \alpha_m b_{0,m}  \label{defDm}
406: \\
407: \Delta_{\vec k} &=& \sum_m D_m k Y_{1 m}(\hat k) \label{defDelta}
408: \end{eqnarray}
409: Hence the mean-field Hamiltonian for the Fermions becomes
410: \begin{eqnarray} \label{hmf}
411: H^{\rm mf} = \sum_{\vec k} (\epsilon_k -\mu)a_{\vec k}^\dagger
412: a_{\vec k} + (\Delta_{\vec k}^* a_{-\vec k}a_{\vec k} + {\rm
413: h.c.})
414: \end{eqnarray}
415: This can be solved via the familiar
416: Bogoliubov transformation, which gives
417: \begin{eqnarray}
418: \langle a_{-\vec k}a_{\vec k} \rangle =
419:  - \frac{\Delta_{\vec k}}{2[(\epsilon_k-\mu)^2+|\Delta_{\vec
420: k}|^2]^\frac{1}{2}}
421: \end{eqnarray}
422: Minimization with respect to the c-numbers $b_{\vec q = 0, m}$ give
423: \begin{eqnarray}
424: b_{0,m} (\delta_m - 2\mu) = - \sum_{\vec k}
425:  \varphi^*_m(\vec k) \langle a_{-\vec k}a_{\vec k} \rangle
426: \end{eqnarray}
427: %The superfluid phase is then characterized by Cooper pairing
428: %parameters $\langle a_{-\vec k}a_{\vec k} \rangle$ and three
429: %different fermion-pair of bound states $b_{0,m}$ under
430: %Bose-Einstein condensation.
431: which can be written as,
432: \begin{eqnarray} \label{Dm}
433: (\delta_m -2 \mu ) D_m
434:   = \frac { 2 \pi} {L^3}  \sum_{\vec k}
435:    \frac {|\alpha_m|^2
436:     k   Y_{1 m}^*(\hat k) \Delta_{\vec k}} { \left[
437: (\epsilon_k-\mu)^2+|\Delta_{\vec k}|^2
438: \right]^\frac{1}{2}  }
439: \label{gap-un}
440: \end{eqnarray}
441: 
442: The R.H.S. of this equation is formally divergent
443: (This divergence is only formal:  the sum may actually
444: converge if the $k$ dependences of $\alpha_m(k)$
445: are included)
446: Nevertheless,
447: it can be "renormalized" by re-expressing this equation
448: in terms of parameters that enter the
449: two-body scattering amplitude
450: (c.f. \cite{ho05} and also the theoretical references
451: in the s-wave case \cite{cross0,falco04,perali,java,numeric}.)
452: Let us then consider the scattering of two particles in vacuum.
453: We denote $f_{\vec k}(\vec k')$ as the
454: scattering amplitude for
455: two incident particles with relative momentum $\vec k$ and
456: out-going momentum $\vec k'$.
457: We can define
458:  $f_m(k)$ (do not confuse
459: this $f_m(k)$ with those in Sec II)
460: via
461: \begin{equation}
462: f_{\vec k} (\vec k') = \sum_m  f_m(k)
463: (4 \pi) Y_{1m}(\hat k') Y_{1m}^{*} (\hat k)
464: \label{fmdef}
465: \end{equation}
466: This equation is the generalization of the corresponding
467: one in scattering theory to the case of anisotropic
468: interaction (but still with rotational invariance around $\hat z$).
469:   It reduces to those in standard
470: quantum mechanics textbooks \cite{Merzbacher}
471: if $f_m(k)$ is independent of $m$.
472: $f_{\vec k}(\vec k')$ is related
473: to the $T$-matrix by
474: \begin{eqnarray} \label{flm}
475: f_{\vec k}(\vec k') = -\frac{M L^3}{4\pi \hbar^2}
476: T_{\vec k',\vec k}(E=2\epsilon_k + i0^+)
477: \end{eqnarray}
478: The $T$-matrix can be evaluated in the standard way:
479: \begin{eqnarray} \label{tmatrix}
480: T_{\vec k',\vec k}(E) &=& \sum_m \big[ \varphi_m(\vec k'){\cal
481: G}^{(m)}_{b_0}(E)\varphi^*_m(\vec k) \nonumber \\
482:  &&+ \varphi_m(\vec k'){\cal G}^{(m)}_{b_0}(E)\Pi^{(mm)}(E) {\cal
483: G}^{(m)}_{b_0}(E)\varphi_m^*(\vec k) + \cdots \big] \nonumber \\
484:  %&=& \sum_m \left[ \frac{\varphi_m(\vec k')\varphi_m^*(\vec k)}{[{\cal
485: %G}^{(m)}_{b_0}(E)]^{-1}-\Pi^{(mm)}(E)} \right]
486:  %\nonumber \\
487:  &=& \sum_m \frac{\varphi_m^*(\vec k')\varphi_m(\vec k)}
488:               {E-\delta_m-\Pi^{(m)}(E)}
489: \end{eqnarray}
490: where
491: \begin{eqnarray} \label{pi}
492: \Pi^{(mm')}(E) &=& \sum_{\vec p,\omega} \varphi_m^*(\vec p) {\cal
493: G}_{f}(\vec p,\omega) {\cal G}_{f}(-\vec p, E - \omega)
494: \varphi_{m'}(\vec p) \nonumber \\
495:  &=& \sum_{\vec p} \frac{\varphi_m^*(\vec p)\varphi_{m'}(\vec p)}
496:     {E-2\epsilon_p}   \ .
497: \end{eqnarray}
498: ${\cal G}_{f}(\vec p, \omega)$
499: is the Greens function for free Fermions
500: at wavevector $\vec p$ ($p$ has dimension of $[L]^{-1}$),
501:  frequency $\omega$,
502: and ${\cal G}^{(m)}_{b_0}(E)$ is the Greens
503: functions of free  $m$-Boson
504:   at zero momentum $\vec q = 0$ and energy $E$.
505: Here we have already made use of a simplification due to the fact
506: that $\Pi^{(m m')}$ vanishes for $m \ne m'$ (see Eq. (\ref{pi})),
507: hence there are no cross-terms between different $m$'s in Eq.
508: (\ref{tmatrix}).
509: 
510: 
511: Substituting Eqs.(\ref{tmatrix})-(\ref{pi}) into Eq.(\ref{flm}),
512: and using the definition (\ref{fmdef}), we get
513: \begin{eqnarray} \label{high1}
514: -\frac{M}{4\pi \hbar^2}\frac{k^2}{f_m(k)} =
515: \frac{2\epsilon_k-\delta_m}{|\alpha_m|^2} -
516: \frac{1}{L^3}\sum_{\vec p} \frac{p^2}{2\epsilon_k-2\epsilon_p +
517: i0^+} \nonumber \\
518: \end{eqnarray}
519: The sum over $\vec p$ in the R.H.S. formally diverges.
520: However, we are interested only on its dependence
521: on $k$ at small $k$'s.  We thus expand the
522: R.H.S. in powers of $k$:
523: \begin{eqnarray} \label{high2}
524: && \frac{2\epsilon_k-\delta_m}{|\alpha_m|^2} -
525: \frac{1}{L^3}\sum_{\vec p} \frac{p^2}{2\epsilon_k-2\epsilon_p +
526: i0^+} \nonumber \\
527:  &=& -( \frac{\delta_m}{|\alpha_m|^2}
528:    -  \frac{M}{\hbar^2} \frac{1}{L^3} \sum_{\vec p} 1  ) \nonumber \\
529:  &&  + (  \frac{\hbar^2}{ M |\alpha_m|^2}
530:    +  \frac{M}{\hbar^2} \frac{1}{L^3} \sum_{\vec p} \frac{1}{p^2} )k^2 -
531:    i\frac{M}{4\pi \hbar^2}k^3
532: \end{eqnarray}
533: 
534: 
535: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
536: \vspace{15pt}
537: %\begin{figure}[tbh]
538: \begin{figure}
539: \begin{center}
540: \includegraphics[width=3in]{mag.eps}
541: \end{center}
542: %\vspace{-5pt}
543:  \caption{Plot of the fit to the experimental results
544: of Ref.\cite{Ticknor04} for $-1/v_m$ and $-c_m$ for $^{40}K$ atoms in the
545: $ |f, m_f\rangle = | 9/2, -7/2\rangle$ hyperfine state.
546: $v_1 = v_{-1}$ and $c_1 = c_{-1}$.
547:  $-1/v_m$'s vanish at the resonant fields $B_m^*$, with
548: $B_0^* > B_{\pm 1}^*$.}
549:  \label{mag}
550:  %\vspace{-5pt}
551: \end{figure}
552: 
553: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
554: 
555: 
556: 
557: 
558: 
559: In the low-energy limit, Eq.(\ref{high1})-({\ref{high2}) should
560: reduce to the general result
561:  for the low-energy scattering between a pair
562: of particles.  Using the same notation adopted in
563: Ref.\cite{Ticknor04}, the scattering amplitude is parametrized as
564: \begin{eqnarray} \label{high3}
565: f_m(k) = \frac{k^2}{ -v_m^{-1}+ c_m k^2 - i k^3} \label{fm}
566: \end{eqnarray}
567: where the dimensions of $v_m$ and $c_m$ are $\rm [L]^3$ and $\rm
568: [L]^{-1}$, respectively.
569: The magnetic field dependent parameters
570: $v_m$, $c_m$ are in principle available experimentally.
571: An example is as shown in  Fig.\ref{mag}.
572:  Our renormalization scheme is to absorb the bare
573: parameters $\delta_m$ and $\alpha_m$ into the physical
574: (renormalized) parameters $v_m$, $c_m$ by identifying the
575: equivalence between Eqs.(\ref{high1})-(\ref{high2}) and
576: Eq.(\ref{high3}). Hence
577: \begin{eqnarray}
578: - \frac{1}{v_m} &=& \frac{ 4 \pi \hbar^2}{M} (
579:   \frac{\delta_m}{|\alpha_m|^2}
580:    - \frac{M}{\hbar^2} \frac{1}{L^3} \sum_{\vec p} 1 )
581: \label{vm} \\
582: c_m &=& - \frac{ 4 \pi \hbar^2}{M} (
583:   \frac{\hbar^2}{ M |\alpha_m|^2}
584:    +  \frac{M}{\hbar^2} \frac{1}{L^3} \sum_{\vec p} \frac{1}{p^2} )
585: \label{cm}
586: \end{eqnarray}
587: %The divergent sums over $\vec p$ on the R.H.S. of the above two
588: %equations can be regulated either by introducing a cut-off or
589: %invoking the fact that the coupling $\tilde \alpha_m$ must
590: %We shall express physical quantities  in terms of
591: %$v_m$ and $c_m$.
592:  In Eq. (\ref{vm}) and (\ref{cm}), if the $p$ dependence
593: of $\alpha_m$'s had been included, the sums involving $\vec p$
594: should have the extra factors $ |\alpha_m(p)|^2/ |\alpha_m(0)|^2$.
595: When treating the many-body problem, as we shall see below, the
596: only "divergent" sums are again $ \sum_{\vec p} |\alpha_m(p)|^2/
597: |\alpha_m(0)|^2$ and $ \sum_{\vec p} |\alpha_m(p)|^2/
598: |\alpha_m(0)|^2 p^2$. We shall apply Eq. (\ref{vm}) and (\ref{cm})
599: to eliminate these sums in favor of $v_m$ and $c_m$. After this
600: procedure, low energy phenomena (involving only small $p$'s) can
601: be parameterized entirely by $v_m$, $c_m$ and $\alpha_m(0)$. To
602: simplify our presentation, we shall not include
603:  explicitly the factors
604: $ |\alpha_m(p)|^2/ |\alpha_m(0)|^2$
605: in our calculations below.
606: 
607: Returning to Eq.(\ref{Dm}), we rewrite its R.H.S as
608: \begin{eqnarray}
609: && \frac{2\pi}{L^3}\sum_{\vec k}
610:  \frac{|\alpha_m|^2 k Y_{1 m}^*(\hat k) \Delta_{\vec k}}
611:  {\left[ (\epsilon_k-\mu)^2+|\Delta_{\vec k}|^2
612:  \right]^\frac{1}{2}} \nonumber \\
613:  &=& \frac{2\pi|\alpha_m|^2}{L^3}\sum_{\vec k,m'}
614: \frac{D_{m'} Y_{1 m}^*(\hat k) Y_{1 m'}(\hat k) k^2}
615:  {\left[ (\epsilon_k-\mu)^2+|\Delta_{\vec k}|^2
616:  \right]^\frac{1}{2}} \nonumber \\
617: &=& \frac{2\pi|\alpha_m|^2}{L^3}\sum_{\vec k,m'}D_{m'} Y_{1
618: m}^*(\hat k) Y_{1 m'}(\hat k)
619:   \nonumber \\
620:  && \times \left\{ h(\vec k) +
621: \frac{k^2}{\epsilon_k}(1+\frac{\mu}{\epsilon_k}-\frac{|\Delta_{\vec
622: k}|^2}{2\epsilon_k^2}) \right\}
623: \label{rngap}
624: \end{eqnarray}
625: where
626: \begin{eqnarray}
627: h(\vec k) \equiv \frac{k^2}{\left[ (\epsilon_k-\mu)^2 +
628: |\Delta_{\vec k}|^2 \right]^{1/2} }
629:  - \frac{k^2}{\epsilon_k}(1+\frac{\mu}{\epsilon_k}-\frac{|\Delta_{\vec
630: k}|^2}{2\epsilon_k^2}) \nonumber \\
631: \label{hk}
632: \end{eqnarray}
633: The sum involving $h(\vec k)$ in Eq. (\ref{rngap}) is convergent.
634: For the rest of the terms, their divergences are the same as in
635: either Eqs. (\ref{vm}) and (\ref{cm}). These sums can thus be expressed
636: in terms of the physical parameters $v_m$ and $c_m$.
637: Using Eq.(\ref{rngap}), the "gap equation" Eq.(\ref{Dm})
638:  can therefore be rewritten as
639: \begin{widetext}
640: \begin{eqnarray}
641:  \frac{M}{4\pi}D_0(-\frac{1}{v_0}+\frac{2Mc_0\mu}{\hbar^2})
642:  - \frac{3M}{10\pi}(\frac{M^2 c_0}{4\pi\hbar^4}+\frac{1}{|\alpha_0|^2})
643:  \left[D_0(2|D_1|^2+3|D_0|^2+2|D_{-1}|^2)-2D_{0}^*D_1 D_{-1}
644:    \right] \nonumber \\
645:  = \frac{2\pi}{L^3}\sum_{\vec k, m}D_m Y_{10}^*(\hat k) Y_{1
646: m}(\hat
647:  k) h(\vec k)
648:  \label{D0}
649: \end{eqnarray}
650: \begin{eqnarray}
651:  \frac{M}{4\pi}D_1(-\frac{1}{v_1}+\frac{2Mc_1\mu}{\hbar^2})
652:  - \frac{3M}{10\pi}(\frac{M^2 c_1}{4\pi\hbar^4}+\frac{1}{|\alpha_1|^2})
653:  \left[ D_1
654: (3|D_1|^2+2|D_0|^2+6|D_{-1}|^2)-D_{-1}^*D_0^2 \right] \nonumber \\
655:  = \frac{2\pi}{L^3}\sum_{\vec k, m}D_m Y_{11}^*(\hat k) Y_{1 m}(\hat
656:  k) h(\vec k)
657:  \label{D1}
658: \end{eqnarray}
659: \end{widetext}
660: and a corresponding equation with $m= 1 \leftrightarrow m=-1$.
661: Eq.(\ref{D0})-(\ref{D1}) are to be solved under the constraint from the
662: number equation
663: \begin{eqnarray}
664: n = \frac{1}{L^3} ( \sum_{\vec k} \langle a_{\vec
665: k}^{\dagger} a_{\vec k} \rangle
666:    + 2 \sum_m |b_{0,m}|^2 )
667: \label{n1}
668: \end{eqnarray}
669: The first term is from open-channel atoms whereas
670: the second term is from the closed-channel molecules.
671: From BCS theory,
672: \begin{displaymath}
673: \langle a_{\vec k}^{\dagger} a_{\vec k} \rangle
674: = \frac{1}{2} \left( 1-\frac{\epsilon_k - \mu}{[
675: (\epsilon_k-\mu)^2+|\Delta_{\vec k}|^2]^\frac{1}{2}} \right)
676: \end{displaymath}
677:  The first term in Eq. (\ref{n1}) is thus again formally divergent:
678:  $ \langle a_{\vec k}^{\dagger} a_{\vec k} \rangle$ $\sim$
679: $ \frac{ |\Delta_{\vec k} |^2} { 4 \epsilon_k^2 }$ at large $k$.
680:  However, this sum
681: it can be treated in analogous manner as Eq. (\ref{rngap}) by
682: employing again Eq. (\ref{cm}).
683:  The result is
684: \begin{eqnarray}
685: n &=& \frac{1}{2L^3} \sum_{\vec k} ( 1-\frac{\epsilon_k - \mu}{[
686: (\epsilon_k-\mu)^2+|\Delta_{\vec k}|^2]^\frac{1}{2}}
687:            - \frac{ |\Delta_{\vec k} |^2} { 2 \epsilon_k^2 } )
688:            \nonumber \\
689:   && - \frac{1}{4\pi}\sum_m(\frac{M^2
690: c_m}{4\pi\hbar^4}+\frac{1}{|\alpha_m|^2})|D_m|^2
691:    + \frac{1}{2\pi}\sum_m
692:   \frac{|D_m|^2}{|\alpha_m|^2}  \nonumber \\
693: \label{n}
694: \end{eqnarray}
695: 
696: Eqs.(\ref{D0}), (\ref{D1}), and (\ref{n}) are our principal
697: equations, with parameters characterizing the Feshbach resonances
698: entirely in terms of $v_m$, $c_m$ and $1/|\alpha_m|^2$.
699:  These equations determine
700: the order parameters $D_m$ and chemical potential $\mu$ for given
701: density $n$ and ``interaction parameters" $v_m$, $c_m$ and
702: $1/|\alpha_m|^2$.
703: 
704: 
705: 
706: 
707: \section{IV. BEC-BCS crossover and quantum phase transitions}
708: 
709: 
710: 
711: Generally with Feshbach resonance for the sub-channel $m$, $1/v_m$ is field
712: dependent, vanishing at the resonant field $B_m^{*}$.  In contrast, $c_m$ has
713: a definite sign ({\it c.f.} Fig. \ref{mag}).
714:  For the ease of discussions, we shall assume that
715: $c_m < 0$ and field independent, $-1/v_m$  is an increasing
716: function of field
717:   $- 1/v_m > (<) 0$
718: for $B > (<) B_m^{*}$, as in the case of $^{40}$K.
719:  (This corresponds to the case where
720: $\delta_m$ is an increasing function of field and $\alpha_m$
721: weakly field dependent, {\it c.f.} Eqs.(\ref{vm}) and (\ref{cm})).
722: For $B < B_m^{*}$, a bound state appears.  The energy of this
723: bound state is given by $- \epsilon_{b,m} = - \hbar^2
724: {\kappa_m}^2/M$ with $k = i \kappa_m$ being a pole for $f_m (k)$,
725: that is, $ - 1/v_m - c_m \kappa_m^2 - \kappa_m^3 = 0$.
726: For given $m$ and small detuning below that resonance,
727: $\kappa_m$ is small and is given by $\kappa_m^2 = 1/ [
728: (-c_m)(v_m) ]$.  Since $1/v_m$ should be roughly linear in $B$
729: near the resonance, $\epsilon_{b,m}$ increases linearly with $
730: (B_m^{*} - B)$ (in contrast to $s$-wave, where it is quadratic;
731: see also \cite{ho05})
732: These results for small detuning apply provided $\kappa_m << (-c_m)$,
733: or equivalently $1/v_m \ll (-c_m)^3$.
734: We shall always confine ourselves to this regime for
735: negative detunings (even when we write $ B \ll B_m^*$).
736: For larger negative detunings we need to take into account
737: higher order terms in $k^2$ in denominator of $f_m(k)$.
738: 
739: 
740: 
741: 
742: Moreover, as explained in Ref.\cite{Ticknor04}, due to the dipole
743: interaction, $B_0^{*} > B_{1}^{*} = B_{-1}^{*}$,
744: as can be seen again in Fig \ref{mag}.
745: Thus in the field
746: range of interest, $ - 1/v_1 = -1/v_{-1} > -1/v_0$. We can say
747: that, at a given field, the effective interaction between the
748: Fermions is less attractive for relative angular momentum
749: projections $m = \pm 1$ than $m = 0$.
750: 
751: 
752: Since the interaction is less attractive for angular momentum
753: projections $m = \pm 1$, for sufficiently large difference between
754: $-1/v_{0}$ and $-1/v_{\pm 1}$ we expect (and verify below) that
755: the pairing is entirely in the $m = 0$ partial wave.  We thus
756: first begin our analysis by assuming that only $D_0$ is
757: non-vanishing.
758: 
759: 
760: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
761: \vspace{25pt}
762: \begin{figure}[tbh]
763: %\begin{figure}
764: \begin{center}
765: \includegraphics[width=3in]{gap_c100.eps}
766: \end{center}
767: %\vspace{-5pt}
768:  \caption{The dimensionless parameters $\tilde D_0$ and $\tilde
769:  \mu$ as functions of $-1/\tilde v_0$.
770: $\tilde c_0= \tilde c_1=-100$ and $\tilde D_{\pm 1} =0$ in this
771: case. The dashed line represents $ - \epsilon_b / 2 \epsilon_F$.
772:   }
773:  \label{gap_c100}
774:  %\vspace{-5pt}
775: \end{figure}
776: %\vspace{10pt}
777: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
778: 
779: 
780: 
781: 
782: 
783: 
784: For simplicity, we shall first drop the terms with explicit $1 /
785: |\alpha_m|^2$ factors. If this is valid, then our equations
786: (\ref{D0}), (\ref{D1}), (\ref{n}) are same as a corresponding
787: model consisting only of Fermions interacting among each other
788: with an interaction such that the corresponding
789:  two-body scattering amplitude parameters are given
790: by the same $c_m$ and $v_m$.
791: In analogy with the s-wave literature, we shall call this the
792: single-channel approximation
793: (note there are still three  ($m = -1, 0, 1$) Feshbach resonances).
794:  The effect from finite $1/|\alpha_m|^2$ will be discussed at the end
795: of this Section.
796: 
797: With these simplifications, Eq. (\ref{D0}) and (\ref{n}) become
798: \begin{equation}
799:  \frac{M}{4\pi}(-\frac{1}{v_0}+\frac{2Mc_0\mu}{\hbar^2})
800:  + \frac{9M}{40\pi^2\hbar^4}(-c_0)|D_0|^2
801:  = \frac{2\pi}{L^3}\sum_{\vec k, m} |Y_{10}(\hat k)|^2 h(\vec k)
802:  \label{D00}
803: \end{equation}
804: and
805: \begin{eqnarray}
806: n &=& \frac{1}{2L^3} \sum_{\vec k} ( 1-\frac{\epsilon_k - \mu}{[
807: (\epsilon_k-\mu)^2+|\Delta_{\vec k}|^2]^\frac{1}{2}}
808:            - \frac{ |\Delta_{\vec k} |^2} { 2 \epsilon_k^2 } )
809:            \nonumber \\
810:   && + \frac{M^2}{(4\pi)^2\hbar^4} (-c_0) |D_0|^2
811: \label{n00}
812: \end{eqnarray}
813: respectively.
814:  Eqs.(\ref{D0}) and (\ref{n}) can be solved
815: simultaneously similar to the $s$-wave case.
816: By gauge invariance, we shall choose $D_0$ to be real
817: without loss of generality.
818: It is convenient to express the results in
819: dimensionless form. We thus define $\tilde \mu\equiv \mu/\epsilon_{\rm
820: F}$, $\tilde D_m \equiv D_m/ \hbar v_{\rm F}$, $\tilde c_m \equiv
821: n^{-1/3}c_m$, and $\tilde v_m \equiv n v_m$ where $\epsilon_{\rm
822: F} \equiv \hbar^2 k^2_{\rm F}/2M$, $v_{\rm F} \equiv \hbar k_{\rm F}/M$,
823: and $k_{\rm F}^3 \equiv 6\pi^2 n$.
824: The results are as shown in
825: Fig. \ref{gap_c100} (for the case $\tilde c_0 = \tilde c_1 = -
826: 100$, see below for the reason of this choice).
827: 
828: In the BCS regime, corresponding to ``large" external magnetic
829: field $B\gg B^*_0$ thus $-{\tilde v_0}^{-1} \to \infty$, we have
830: $\tilde \mu \simeq 1$, and $\tilde D_0\simeq 0$.
831: We can ignore the term $\propto |D_0|^2$
832:  on the left hand side of Eq.(\ref{D00}).
833:  This equation then
834: reduces to the corresponding one in Sec II with the effective
835: pairing interaction $V_0$ in the $m=0$ channel there
836: proportional to $ (-1/v_0 + 2M c_0 \mu/\hbar^2 )^{-1}$ $=$
837: $(-1/v_0 + c_0 k_F^2)^{-1}$.
838: This combination follows from the factor
839:  $(\delta_0 - 2 \mu)$ on the left hand side of Eq. (\ref{gap-un}).
840: 
841: 
842: In the BEC regime,  corresponding to $B\ll B^*_0$ or $-{\tilde
843: v}_0^{-1}\rightarrow -\infty$, $\mu \simeq -\epsilon_{b,0}/2$, and
844: $\tilde D_0$ approaches to a constant. In Eq. (\ref{D00}) we can
845: ignore all terms explicit in $\Delta_{\vec k}$ or $D_0$.  Writing
846: $\kappa = (2 M |\mu|)^{1/2}$ and performing the sum over $\vec k$,
847: we find that $\kappa$ obeys $-1/v_0 -c_0 \kappa^2 = \kappa^3$.
848: Comparing this with the equation for the two body bound state, we
849: find $|\mu| = \epsilon_{b 0}/2$ as claimed. In Eq. (\ref{n00}), we
850: can expand the square-root $[(\epsilon_k - \mu) + |\Delta_{\vec
851: k}|^2]^{-1/2}$ $\approx$ $ (\epsilon_k + |\mu|) - |\Delta_{\vec
852: k}|^2/ 2 (\epsilon_k + |\mu|)$. We then get
853: \begin{equation}
854: n = \frac{M^2}{(4\pi)^2 \hbar^4} |D_0|^2 \left[ (-c_0) - \frac{3}{2}\kappa
855: \right]
856: \label{n0bec}
857: \end{equation}
858: where $\kappa$ was defined above.  For small detuning
859: $\kappa \ll (-c_0)$, {\it i.e.} $1/\tilde v_0 \ll |\tilde c_0|^3$,
860: a relation well-satisfied in the range in Fig \ref{gap_c100}
861: and others below, we then obtain
862:  $\tilde D_0 \simeq (32 \pi/3)^{1/3}
863: (- \tilde c_0)^{-1/2}$.
864: 
865: The ``cross-over" behavior in Fig.\ref{gap_c100} is analogous to
866: the $s$-wave case, where the corresponding $x$-axis is $ x = - 1/
867: (n^{1/3} a_s)$ where $a$ is the $s$-wave scattering length.
868:  Note here
869: $D_0$ has the dimension of $\rm [E][L]^{-1}$ and
870: behaves differently from the $s$-wave $\Delta$ in the BEC limit:
871: it saturates rather than continue to increase.
872: Rigorously speaking, we actually expect a (quantum) phase transition
873: at the point where $\mu$ crosses zero \cite{randeria90,klinkhamer}:
874: the system is gapless for $\mu > 0$ but gapful if $\mu < 0$.
875: However, we found no evident changes of slope
876: in $\mu$ and $\Delta$ when this transition is crossed,
877: and thus expect this phase transition is very weak thermodynamically.
878: Similar results were found in Ref \cite{ho05} for the axial
879: ($D_1 \ne 0$, $D_0 = D_{-1} = 0$ state).
880: 
881: We have also performed calculations for other values of $\tilde
882: c_0$. The size of the crossover region is roughly proportional to
883: the value of $\tilde c_0$ (which is in turn inversely proportional
884: to the density).  For example, for $\tilde c_0 = -200$ (not shown), the
885: corresponding results can be captured well by replacing the
886: $x$-axis by $ - 1/ 2 \tilde v_0$ and dividing $\tilde D_0$ by $1 /
887: \sqrt{2}$ in Fig.\ref{gap_c100}.
888: 
889: 
890: 
891: 
892: 
893: 
894: 
895: 
896: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
897: %Under the assumption that $\tilde D_1=\tilde D_{-1}^*$ being real,
898: %and $\tilde c_0=\tilde c_1=\tilde c_{-1}$, the principal equations
899: %in dimensionless form are
900: %\begin{widetext}
901: %\begin{eqnarray}
902: %1-\frac{(6\pi^2)^\frac{2}{3}}{16\pi}(-\tilde c_0)(\tilde
903: %D_0^2+2\tilde D_1^2) &=& \frac{3}{8\pi}f(\tilde\mu,\tilde
904: %D_0,\tilde D_1) \label{func} \\
905: %\tilde D_0 \left\{ -\frac{1}{\tilde v_0} -
906: %(6\pi^2)^\frac{2}{3}(-\tilde
907: %c_0)\left[\tilde\mu-\frac{3}{10\pi}(3\tilde D_0^2+2\tilde
908: %D_1^2)\right] \right\} &=& 9 \tilde D_0 g(\tilde\mu,\tilde D_0,\tilde D_1)
909: %\label{gunc} \\
910: %\tilde D_1 \left\{ -\frac{1}{\tilde v_1} -
911: %(6\pi^2)^\frac{2}{3}(-\tilde
912: %c_0)\left[\tilde\mu-\frac{3}{10\pi}(\tilde D_0^2+9\tilde
913: %D_1^2)\right] \right\} &=& 9 \tilde D_1 h(\tilde\mu,\tilde
914: %D_0,\tilde D_1) \label{hunc}
915: %\end{eqnarray}
916: %\end{widetext}
917: %where
918: %\begin{widetext}
919: %\begin{eqnarray}
920: %f(\tilde\mu,\tilde D_0,\tilde D_1) &=& \int_0^\infty dx \int
921: %d\Omega\ x^2 \left( 1 -
922: %\frac{x^2-\tilde\mu}{[(x^2-\tilde\mu)^2+\frac{3}{\pi}x^2\tilde D^2
923: %]^\frac{1}{2}}-\frac{3}{2\pi x^2}\tilde D^2 \right) \\
924: %g(\tilde\mu,\tilde D_0,\tilde D_1) &=& \int_0^\infty dx \int
925: %d\Omega\ x^2\cos^2\theta\left(
926: %\frac{x^2}{[(x^2-\tilde\mu)^2+\frac{3}{\pi}x^2\tilde D^2
927: %]^\frac{1}{2}} - (1+\frac{\tilde\mu}{x^2}-\frac{3}{2\pi x^2}\tilde
928: %D^2 ) \right) \\
929: %h(\tilde\mu,\tilde D_0,\tilde D_1) &=& \int_0^\infty dx \int
930: %d\Omega\ x^2\sin^2\theta\sin^2\varphi \left(
931: %\frac{x^2}{[(x^2-\tilde\mu)^2+\frac{3}{\pi}x^2\tilde D^2
932: %]^\frac{1}{2}} - (1+\frac{\tilde\mu}{x^2}-\frac{3}{2\pi x^2}\tilde
933: %D^2 ) \right)
934: %\end{eqnarray}
935: %\end{widetext}
936: %with $\tilde D^2 \equiv \tilde D_0^2 \cos^2\theta+2\tilde
937: %D_1^2\sin^2\theta\sin^2\varphi$.
938: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
939: 
940: 
941: 
942: The above behavior applies only to sufficiently large $ -\tilde
943: v_{\pm 1}^{-1} - ( -\tilde v_0^{-1}) > 0$.  When this difference
944: is sufficiently small, $\tilde D_{\pm 1}$ will become finite.
945: Similar to the treatment in Section II for the
946: anisotropic weak-coupling BCS model,
947: we first identify the critical $\tilde
948: v_{\pm 1}^{* -1}$ such that $D_1$ and/or $D_{-1}$ first start to deviate from
949: zero.
950: For simplicity, in the results below we shall take $c_0 = c_{\pm 1}$.
951: This is satisfied to a good approximation at least for
952: $^{40}$K near the resonances (see Fig \ref{mag}).
953: We linearize
954: Eqs.(\ref{D0})-(\ref{D1}) (and the latter with
955: $m=1 \leftrightarrow m=-1$) in $D_1$ and $D_{-1}$.
956: We obtain
957: \begin{eqnarray}
958: -\frac{M}{4\pi v_{1}^*} \left( \begin{array}{c}
959:  D_1 \\ D_{-1}^* \end{array} \right) =
960: \left( \begin{array}{cc}
961: A_1 & A_2 \\
962: A_2 & A_1 \end{array} \right)
963:  \left( \begin{array}{c}
964:  D_1 \\ D_{-1}^* \end{array} \right)
965: \end{eqnarray}
966: where
967: \begin{eqnarray}
968: A_1 &=& \frac{2\pi}{L^3}\sum_{\vec k} |Y_{11}|^2 h(\vec k)
969:  -\frac{2\pi}{L^3}\sum_{\vec k} D_0^2 |Y_{10}|^2 |Y_{11}|^2 g(\vec
970:  k) \nonumber \\
971: && - \frac{M^2 c_1}{2\pi \hbar^2}(\mu-\frac{3M}{10\pi \hbar^2}D_0^2) \\
972:  A_2 &=& \frac{2\pi}{L^3}\sum_{\vec k} D_0^2 |Y_{10}|^2 |Y_{11}|^2
973:  g(\vec k) - \frac{3M^3 c_1}{40\pi^2 \hbar^4}D_0^2 \\
974:  g(\vec k) &\equiv& \frac{k^4}{2}(
975:  \frac{1}{\left[(\epsilon_k -\mu)^2+ D_0^2 |Y_{10}|^2 k_z^2 \right]^{3/2}}
976:  - \frac{1}{\epsilon_k^3} )
977: \end{eqnarray}
978: Here $h(\vec k)$ is as defined in Eq. (\ref{hk}) but with only
979: $D_0 \ne 0$. We can verify that $A_2 > 0$.  Therefore the largest
980: eigenvalue for the A-matrix, hence the smallest value for
981: $v_{1}^*$, belongs to the class $D_1 = D_{-1}^*$. The
982: corresponding critical value is then given by $ -\frac{M}{4\pi
983: v_{1}^*} = A_1 + A_2$. On the other hand, from Eq.(\ref{D0}), we
984: have
985: \begin{eqnarray}
986: -\frac{M}{4\pi v_0}
987:  = \frac{2\pi}{L^3}\sum_{\vec k} |Y_{10}|^2 h(\vec k)
988:   - \frac{M^2 c_0}{2\pi\hbar^2}(\mu-\frac{9M}{20\pi\hbar^2}D_0^2)
989:   \nonumber \\
990: \end{eqnarray} Combining these two equations
991: and re-writing it in dimensionless form, we have
992: \begin{eqnarray}
993: &&-\frac{1}{\tilde v_1^*}+\frac{1}{\tilde v_0} =
994: \frac{3(6\pi^2)^\frac{2}{3}}{5\pi}{\tilde D_0}^2(-\tilde c_0)
995: \nonumber \\
996: &&+ 9\pi \int_0^\infty dx \int_{-1}^1 dy
997:  \frac{x^4(1-3 y^2)}{[(x^2-\tilde \mu)^2+\frac{3}{\pi}\tilde D_0^2 x^2
998: y^2]^{1/2}} \nonumber \\
999:  \label{v*}
1000: \end{eqnarray}
1001: 
1002: In the BCS limit, $\tilde D_0\rightarrow 0$, the first term in the
1003: R.H.S. of Eq.(\ref{v*}) is negligible whereas the second,
1004:  which we shall call $K_1$, is finite.
1005: The dominant contribution to $K_1$ occurs near $x \simeq \sqrt{\tilde
1006: \mu}$, where the integrand diverges if $\tilde D_0$
1007: were zero exactly. To evaluate $K_1$ in the limit $\tilde D_0 \to
1008: 0$, we add and subtract the integral
1009: \begin{equation}
1010: K_2 \equiv 9\pi \int_{-\xi_0}^{\xi_0} d\xi \int_{-1}^1 dy
1011:  \frac{(1-3y^2)}{2[\xi^2+\frac{3}{\pi}\tilde
1012: D_0^2 y^2 ]^\frac{1}{2}}
1013: \label{K2}
1014: \end{equation}
1015: where $\xi \equiv x^2 - \tilde \mu$, and $\xi_0 \gg \tilde D_0$
1016:  is an arbitrary cut-off.  The integrand for
1017: $K_1 - K_2$ now has no divergence near $ x^2 - \tilde \mu \approx 0$,
1018: and their difference vanishes in the $\tilde D_0 \to 0$ limit due
1019: to the $y$ integral.
1020: $K_2$ can be evaluated by first integrating with respect to $\xi$,
1021: we get
1022: \begin{eqnarray}
1023: && 9\pi  \int_{-1}^1 dy \int_{-\xi_0}^{\xi_0} d\xi
1024:  \frac{(1-3y^2)}{2[\xi^2+\frac{3}{\pi}\tilde
1025: D_0^2 y^2 ]^\frac{1}{2}} \nonumber \\
1026: &=& -18\pi  \int_{-1}^1 dy\ (1-3y^2)\log y
1027: \nonumber \\
1028: &=& 12\pi
1029: \end{eqnarray}
1030: In going from the first to the second line, we have
1031: used the fact that
1032: $ \int_{-1}^1 dy\ (1-3y^2) = 0$, hence the terms involving
1033: $\xi_0$ and $\tilde D_0$ drop out in the  $\xi_0 \gg \tilde D_0$ limit.
1034: Therefore in the BCS limit we obtain $-\tilde
1035: v_1^{* -1} + \tilde v_0^{-1} \rightarrow 12\pi = 37.7$.
1036: 
1037: In the BEC limit, recalling that $\tilde D_0$ approaches
1038: a constant  $\simeq (32 \pi/3)^{1/3} (- \tilde c_0)^{-1/2}$
1039: whereas $\tilde \mu$ is large and negative,
1040:  we find that the second term in the RHS of Eq. (\ref{v*}) is
1041: negligible compared with the first (provided $\kappa \ll (-c_0)$).
1042: Using the value for $\tilde D_0$, Eq. (\ref{v*}) yields $-\tilde
1043: v_1^{* -1} + \tilde v_0^{-1}\rightarrow 48\pi/5 = 30.2$. In the
1044: intermediate regime, the numerical results for $-\tilde v_1^{* -1}
1045: + \tilde v_0^{-1}$ as a function of $-\tilde v_0^{-1}$ is shown as
1046: the thick black line ($\tilde D_0 = 0$) in Fig.\ref{beta_c}.
1047: 
1048: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1049: %\vspace{25pt}
1050: %\begin{figure}[tbh]
1051: %\begin{figure}
1052: %\begin{center}
1053: %\includegraphics[width=3in]{mixed4d.eps}
1054: %\end{center}
1055: %\vspace{-5pt}
1056: % \caption{Contour plot of $\tilde D_1$ as a function of
1057: %$-1/\tilde v_1+ 1/\tilde v_0$ and $-1/\tilde v_0$ for
1058: % $\tilde c_0= \tilde c_1= -100$. The line for $\tilde
1059: %D_{\pm 1} =0$ corresponds to the critical value $-1/\tilde v_1^*+
1060: %1/\tilde v_0$.} \label{mixed4d}
1061:  %\vspace{-5pt}
1062: %\end{figure}
1063: 
1064: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1065: 
1066: 
1067: We now solve for the order parameters $D_{m}$ when
1068: $-\tilde v_1^{-1} + \tilde v_0^{-1}$ is less than the critical
1069: value.  As already mentioned, we have already made use of
1070:  gauge invariance to  choose
1071: $D_0$ to be real. Under this choice, the solutions we found belong to the
1072: class $D_{1} = D_{-1}^{*}$. Writing $D_{1} = |D_{1}| e^{i \chi}$,
1073: $\Delta_{\vec k}$ then has the angular dependence $\propto$ $ D_0
1074: \hat k_z + i \sqrt{2} |D_1| \hat k \cdot \hat a$ $\propto (\hat
1075: k_z - i \beta \hat k \cdot \hat a)$ where $\hat a = (\cos \chi)
1076: \hat y + (\sin \chi) \hat x$ is a unit vector perpendicular to
1077: $\hat z$ and $\beta = \sqrt{2} |D_{1} / D_{0}|$. A particular
1078: solution is given by the case where $D_1$ and $D_{-1}$ are both
1079: real where $\hat a = \hat y$. Other solutions are simply
1080: related to this one by a rotation about $\hat z$.  Without loss of
1081: generality, we shall therefore assume that $D_{m}$'s are all
1082: real below.
1083: 
1084: 
1085: The contour plot of $\beta$ (by solving Eq. (\ref{D0})-(\ref{n})
1086: without the $1/|\alpha_m|^2$ terms) is shown in Fig.\ref{beta_c}.
1087: (see Fig. 2 of \cite{Cheng05} for a plot of $D_{\pm 1}$) $\beta$
1088: monotonically increases downwards or towards the right in this
1089: diagram. In the $\beta = 0$ ($D_{\pm 1} = 0$) phase, the state is
1090: rotationally invariant about $\hat z$, whereas this symmetry is
1091: broken in the $\beta \ne 0$ phase. There is a (quantum) phase
1092: transition between these two phases when one crosses the critical
1093: line $-\tilde v_1^{* -1} + \tilde v_0^{-1}$.
1094: 
1095: 
1096: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1097: \vspace{45pt}
1098: %\begin{figure}[tbh]
1099: \begin{figure}
1100: \begin{center}
1101: \includegraphics[width=3in]{beta_c.eps}
1102: \end{center}
1103: %\vspace{-5pt}
1104:  \caption{Contour plot of $\beta$ as a function of
1105: $-1/\tilde v_1+ 1/\tilde v_0$ and $-1/\tilde v_0$ for
1106:  $\tilde c_0= \tilde c_{\pm 1}= -100$.
1107: The line for $\beta =0$ corresponds to the critical value
1108: $-1/\tilde v_1^*+ 1/\tilde v_0$.
1109: } \label{beta_c}
1110:  \vspace{15pt}
1111: \end{figure}
1112: 
1113: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1114: 
1115: For $\tilde
1116: c_0=-200$, the results are qualitatively the same
1117: provided we double the
1118: value of $-1/\tilde v_0$. The corresponding plot is shown in
1119: Fig.\ref{beta_c200}.
1120: 
1121: In the intermediate splitting regime, we thus predict the state is
1122: $k_z - i\beta k_y$ on the BCS side, whereas it is
1123: $ k_z$ on the BEC side. For large positive detuning, the splitting
1124: should be less relevant and the pairing state should resemble more
1125: that of the isotropic system. On the BEC side, the system should
1126: be closer to a Bose condensate of lowest energy molecules ($\hat
1127: k_z$).
1128: 
1129: 
1130: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1131: \vspace{75pt}
1132: %\begin{figure}[tbh]
1133: \begin{figure}
1134: \begin{center}
1135: \includegraphics[width=3in]{beta_c200.eps}
1136: \end{center}
1137: %\vspace{-5pt}
1138:  \caption{Contour plot of $\beta$ as a function of
1139: $-1/\tilde v_1+ 1/\tilde v_0$ and $-1/\tilde v_0$ for
1140:  $\tilde c_0= \tilde c_{\pm 1}= -200$.
1141: %The line for $\tilde D_{\pm 1} =0$ corresponds to the critical value
1142: %$-1/\tilde v_1^*+ 1/\tilde v_0$.
1143: } \label{beta_c200}
1144:  \vspace{5pt}
1145: \end{figure}
1146: 
1147: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1148: 
1149: 
1150: 
1151: For the $^{40}$K case studied in Ref.\cite{Ticknor04}, the
1152: Feshbach resonances are at $B^*_0 \approx 198.8 {\rm G}$ and
1153: $B^*_1 \approx 198.4 {\rm G}$.  There, $c_1$ and $c_0$ are both
1154: only weakly field dependent and are approximately given by $-0.02
1155: a_0^{-1}$ (see Fig \ref{mag}).
1156:  Our choice of $\tilde c = - 100$ above corresponds to a
1157: density of roughly $10^{-11} a_0^{-3}$ = $6.7 \times 10^{13} {\rm
1158: cm}^{-3}$.
1159:  The range of $-1/{\tilde v_0}$ for $\tilde \mu$ to go
1160: from $0$ to $1$
1161: corresponds to roughly $0.5$G
1162:   if we take the same gas density.
1163: (This field range is proportional to $n^{2/3}$ ).
1164:  Near the resonant fields, $ -v_1^{-1} + v_0^{-1}
1165: \approx 2.1 \times 10^{-8} a_0^{-3}$ and is roughly field
1166: independent. Thus the density determines the values for both
1167: $\tilde c_{0, 1}$ and $ -\tilde v_1^{-1} +\tilde v_0^{-1}$ while
1168: varying the magnetic field corresponds roughly to moving along a
1169: horizontal line on our phase diagram of Fig.\ref{beta_c} (with
1170: increasing field towards the right and the distance of the line
1171: from the horizontal axis proportional to $n^{-1}$).
1172: For the density cited above, $ -\tilde v_1^{-1} +\tilde v_0^{-1}
1173: \approx 2000$, hence we expect only the $k_z$ phase to be observed.
1174: To observe the phase transition, we need a $^{40}K$ gas of
1175: a much higher density, or another gas species with much smaller splitting
1176: between the $m= 0$ and $m = {\pm 1}$ resonances.
1177: 
1178: 
1179: 
1180: 
1181: In the above treatment, we have assumed that the terms with
1182: explicit $1/|\alpha_m|^2$ factors in Eqs.(\ref{D0})-(\ref{n}) can
1183: be dropped. The validity of this assumption has to
1184:  be investigated by
1185: first-principle calculations of two-atom collisions.
1186: Following Ref \cite{chevy}, we
1187: define the quantity
1188: \begin{equation}
1189: \eta_m \equiv \frac{M^2 |\alpha_m|^2}{\hbar^4 L^3}\sum_{\vec p}\frac{1}{p^2}
1190: \   .
1191: \end{equation}
1192: $\eta_m$ quantifies the relative strength of coupling between the
1193: $m$ closed-channel and the continuum.  With this notation, Eq.
1194: (\ref{cm}) becomes
1195: \begin{equation}
1196: c_m = - \frac{4 \pi \hbar^4}{M^2 |\alpha_m|^2} (1 + \eta_m) \ .
1197: \end{equation}
1198: The single channel approximation used above corresponds to
1199: $\eta_m \to \infty$. We now investigate the effect of finite $\eta_m$
1200: on the many-body problem. Expressing the Eqs.(\ref{D0}),
1201: (\ref{D1}), and (\ref{n}) by $\eta_m$ instead of $1/|\alpha_m|^2$,
1202: we get
1203: \begin{widetext}
1204: \begin{eqnarray}
1205:  \frac{M}{4\pi}D_0(-\frac{1}{v_0}+\frac{2Mc_0\mu}{\hbar^2})
1206:  - \frac{\eta_0}{1+\eta_0}\frac{3M^3c_0}{40\pi^2\hbar^4}
1207:  \left[D_0(2|D_1|^2+3|D_0|^2+2|D_{-1}|^2)-2D_{0}^*D_1 D_{-1}
1208:    \right] \nonumber \\
1209:  = \frac{2\pi}{L^3}\sum_{\vec k, m}D_m Y_{10}^*(\hat k) Y_{1
1210: m}(\hat
1211:  k) h(\vec k)
1212: \end{eqnarray}
1213: \begin{eqnarray}
1214:  \frac{M}{4\pi}D_1(-\frac{1}{v_1}+\frac{2Mc_1\mu}{\hbar^2})
1215:  - \frac{\eta_1}{1+\eta_1}\frac{3M^3c_1}{40\pi^2\hbar^4}
1216:  \left[ D_1
1217: (3|D_1|^2+2|D_0|^2+6|D_{-1}|^2)-D_{-1}^*D_0^2 \right] \nonumber \\
1218:  = \frac{2\pi}{L^3}\sum_{\vec k, m}D_m Y_{11}^*(\hat k) Y_{1 m}(\hat
1219:  k) h(\vec k)
1220: \end{eqnarray}
1221: \end{widetext}
1222: \begin{eqnarray}
1223: n &=& \frac{1}{2L^3} \sum_{\vec k} ( 1-\frac{\epsilon_k - \mu}{[
1224: (\epsilon_k-\mu)^2+|\Delta_{\vec k}|^2]^\frac{1}{2}}
1225:            - \frac{ |\Delta_{\vec k} |^2} { 2 \epsilon_k^2 } )
1226:            \nonumber \\
1227:   && - \frac{M^2}{(4\pi)^2\hbar^4}\sum_m
1228:        \frac{2+\eta_m}{1+\eta_m} |D_m|^2 c_m
1229: \label{neta}
1230: \end{eqnarray}
1231: We note here that the last term of Eq.(\ref{neta}) involves a
1232: factor $(2 + \eta_m)/(1 + \eta_m)$.  This term actually arises
1233: from two contributions.  The first one, involving $2/(1 +
1234: \eta_m)$, is due to the closed-channel molecules (the last term in
1235: Eq.(\ref{n})), and a second one,  involving $\eta_m /(1 + \eta_m)$
1236: due to the open-channel atoms.
1237: 
1238: 
1239: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1240: \vspace{50pt}
1241: \begin{figure}[tbh]
1242: %\begin{figure}
1243: \begin{center}
1244: \includegraphics[width=3in]{eta.eps}
1245: \end{center}
1246: %\vspace{-5pt}
1247:  \caption{Plots of (a) $\tilde D_0$ and (b) $\tilde \mu$
1248:  as functions of $-1/\tilde v_0$ at
1249:  $\tilde c_0= \tilde c_{\pm 1}= -100$, $\tilde D_{\pm 1}=0$ for different
1250: $\eta_0$'s.
1251: $\eta_0=\infty$ corresponds that of Fig \ref{gap_c100}.
1252:  The dashed line in (b) represents to the two-body result.}
1253: \label{eta}
1254:  \vspace{5pt}
1255: \end{figure}
1256: 
1257: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1258: 
1259: To show the effect of finite $\eta_0$, we plot the
1260: general results for
1261:  $\tilde D_0$ and $\tilde \mu$ in the state $\tilde
1262: D_{\pm 1} = 0$  in Fig.\ref{eta}. The value of $\eta_0$ mainly
1263: affects the results on the BEC side. In fact, deep in the BCS
1264: limit the results are independent of $\eta_0$, as expected since
1265: all atoms are basically in the open-channels. $\tilde \mu$ is
1266: still basically linear in $ 1/ \tilde v_0$ with the same  slope,
1267: and thus the width of the "crossover" region remains to be
1268: determined by $\tilde c_0$ only (but not $\eta_0$). In the BEC
1269: limit, when $\eta_0$ decreases from $\infty$, the contribution
1270: from bound state molecules becomes more important (see in
1271: particular Eq. (\ref{neta}) and the discussions below it).
1272:  The chemical potential $\tilde
1273: \mu$ becomes closer to the two-body value for decreasing $\eta_0$
1274: as shown in Fig. \ref{eta}(b).
1275: $\tilde D_0$ has the limiting value
1276: $\simeq (32 \pi/3)^{1/3}
1277: (- \tilde c_0)^{-1/2} [ (1 + \eta_0)/(2 + \eta_0) ]^{1/2}$,
1278: and thus decreases with decreasing $\eta_0$.
1279: 
1280: The phase transition between the $k_z$ state and
1281: the $k_z - i \beta k_y$ state can be investigated as before.
1282: The value for $ -\tilde v_1^{* -1} +\tilde v_0^{-1}$
1283: is not affected by the value of $\eta_m$'s in the
1284: BCS limit.
1285:  In the BEC limit,
1286: if $\eta_0 = \eta_{\pm 1} \equiv \eta$,
1287: we obtain $ -\tilde v_1^{* -1} +\tilde v_0^{-1}
1288: \rightarrow (30.2) \eta/(2+\eta)$.
1289: This critical value thus decreases with decreasing $\eta$.
1290: Its behavior is as shown in Fig \ref{veta}.
1291: The $\eta$ dependence of this transition line has also
1292: been investigated in Ref \cite{gurarie}.
1293: 
1294: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1295: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1296: \vspace{70pt}
1297: \begin{figure}[tbh]
1298: %\begin{figure}
1299: \begin{center}
1300: \includegraphics[width=3in]{veta.eps}
1301: \end{center}
1302: %\vspace{-5pt}
1303:  \caption{The transition line $ -\tilde v_1^{* -1} +\tilde v_0^{-1}$
1304: between the $k_z$ phase and the $k_z - i \beta k_y$ phase
1305: for $\eta$'s in the legend. $\tilde c_0 = \tilde c_{\pm 1} = -100$.}
1306: \label{veta}
1307:  \vspace{-5pt}
1308: \end{figure}
1309: 
1310: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1311: 
1312: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1313: 
1314: 
1315: We conclude with a discussion on the condition $\eta \gg 1$, or
1316: equivalently $\hbar^4/(\alpha_m M)^2 \ll |c_m| $, for the validity
1317: of the effective single-channel approximation. It is worth noting
1318: that the corresponding condition for the s-wave Feshbach resonance
1319: superfluidity is quite different.  For this latter case, the
1320: single-channel approximation near resonanace is valid if
1321: \cite{renorm,diener04,mackie05,szymanska05,partridge05} the
1322: corresponding parameter $\alpha_s$ in Eq.(\ref{Hal}) satisfies
1323: $\hbar^4/(|\alpha_s|^2 M^2) \ll 1/n^{1/3}$. This case is often
1324: also referred to the "wide resonance" regime. Though for both s
1325: and p-wave resonances, the single-channel approximation would be
1326: valid for sufficiently strong coupling (large $\alpha$) between
1327: the closed and open channels, for s-wave the condition for a
1328: single-channel approximation is always satisfied when the gas is
1329: sufficiently dilute. In contrast, for p-wave the condition $\eta
1330: \gg 1$ is independent of the density of the gas. This difference
1331: can be understood by noting that the dimension of $\alpha_s$ is
1332: $\rm [E][L]^{3/2}$, whereas that of $\alpha_m$ here is $\rm
1333: [E][L]^{5/2}$. The effective range $r_s$ for s-wave Feshbach
1334: resonance  \cite{renorm,diener04} is proportional to
1335: $\hbar^4/(|\alpha_s|^2 M^2)$, and the additional term $\propto
1336: \sum (1/p^2)$ in Eq. (\ref{cm}) does not arise.
1337: 
1338: For the s-wave case, the terms "single-channel approximation"
1339: and "wide-resonance" have been used interchangably.
1340: However, we should note that for our p-wave case,
1341: the condition $\eta \gg 1$ is not necessarily
1342: in conflict with the fact that the
1343:  the peak in scattering cross-section discussed in
1344: Fig.1 of Ref.\cite{Ticknor04} is narrow.  This p-wave scattering
1345: cross-section is proportional to $|f_m|^2$, which is given by
1346: \begin{eqnarray}
1347: |f_m|^2 = \frac{k^4}{(\frac{1}{v_m} + |c_m| k^2)^2 + k^6}
1348: \end{eqnarray}
1349: For $B > B_m^*$, where $1/v_m < 0$, the cross-section thus has a
1350: peak at finite energy $E = \hbar^2 k_o^2/M$ where $k_o^2 = 1/ (|c_m v_m|)$.
1351: The denominator
1352:  can be written as $ c_m^2 (k^2 - k_o^2)^2 + k^6$.
1353: For $B \gtrsim B_m^*$, $|f_m|^2$ as a function of $k^2$ then peaks
1354: very near $k^2 = k_o^2$ and has a width approximately given by
1355: $k_o^3/|c_m| = 1/ (|c_m|^{5/2}(-v_m)^{3/2})$ $\ll k_o^2$, and therefore a very
1356: narrow (in energy) resonance. This is entirely a result of the
1357: special form of $f$ for finite angular momentum.
1358:    This narrowness of the peak is therefore not directly
1359: connected with the assumption on the magitude of $\eta$ made
1360: here.
1361: 
1362: In addition, the condition for $\eta \gg 1$ here is also {\em not}
1363: necessarily in conflict with the assumption of a ``narrow" resonance in the
1364: context discussed by Gurarie {\it et al} \cite{gurarie}.  They defined a
1365: parameter $\gamma \propto \alpha_m^2 M^2 n^{1/3}$, and called the
1366: case $\gamma \ll 1$ to be a narrow resonance.  Their condition is
1367: fulfilled always at sufficiently low densities. Thus the resonance
1368: can be narrow in their language whereas the single-channel
1369: approximation can still be valid if $n^{1/3}  \ll \hbar^4/(\alpha_m M)^2
1370: \ll |c_m| $.
1371: 
1372: 
1373: 
1374: 
1375: 
1376: \section{V. experimental probe of pairing symmetry }
1377: 
1378: It was suggested by Altman {\it et al.} \cite{altman} that
1379: the measurement of the atom shot noise in the time-of-flight (TOF)
1380: absorption image can reveal the properties of the many-body
1381: states, including the symmetry of the pairing state.
1382:  Later the proposal was carried out by Greiner {\it et al.}
1383: \cite{greiner} for $^{40}$K atoms under an $s$-wave
1384: Feshbach resonance. As discussed in previous section, the
1385: pairing state induced by $p$-wave Feshbach resonance is either
1386: $k_z$ or $k_z - i \beta k_y$ where $\beta$ can vary under a change
1387: in the magnetic field or the gas density. In this section, we
1388: propose how $\beta$ can be determined in a TOF experiment.
1389: 
1390: In the TOF experiment, one can measure the quantity
1391: \begin{eqnarray}
1392: {\cal G}(\vec r, -\vec r) &\equiv& \langle n(\vec r)n(-\vec
1393: r)\rangle_t - \langle n(\vec r) \rangle_t \langle n(-\vec r)
1394: \rangle_t \nonumber
1395: \\ &\propto& \langle n_{\vec k}n_{-\vec k}\rangle -
1396: \langle n_{\vec k} \rangle \langle n_{-\vec k} \rangle
1397: \end{eqnarray}
1398: where $\vec k \equiv m \vec r/(\hbar t)$. $t$ is the time after
1399: the trapping potential and the interaction between atoms are
1400: suddenly turned off. Using the mean-field
1401: Hamiltonian in Eq.(\ref{hmf}), we have
1402: \begin{eqnarray}
1403: {\cal G}(\vec r, -\vec r) &\propto& \frac{|\Delta_{\vec
1404: k}|^2}{(\epsilon_k - \mu)^2 + |\Delta_{\vec k}|^2} \nonumber \\
1405: &=& \frac{\frac{3}{\pi}\tilde D_0^2 (\tilde k_z^2 + \beta^2 \tilde
1406: k_x^2)}{(\tilde k^2 - \tilde \mu)^2 + \frac{3}{\pi}\tilde D_0^2
1407: (\tilde k_z^2 + \beta^2 \tilde k_x^2)}
1408: \end{eqnarray}
1409: where $\tilde {\vec k} \equiv \vec k/k_{\rm F}$.
1410: The two phases $k_z$ and $k_z - i \beta (\vec k \cdot \hat a)$
1411: can be distinguished by the presence of anisotropy in
1412: the $x-y$ plane.  For example, consider the weighted average
1413: over $\phi$ defined by
1414: \begin{equation}
1415: S_0 (\alpha) \equiv \langle
1416: {\rm cos} \left( 2 (\phi + \alpha) \right)
1417: {\cal G}(\vec r, -\vec r) \rangle
1418: \end{equation}
1419: This average vanished in the $k_z$ phase.
1420: For $\beta \ne 0$, $S_0 (\alpha)$ becomes finite
1421: and can in principle be evaluated.
1422: The presence of the $\tilde k_z^2 + \beta^2 \tilde k_x^2$
1423: term in the denominator
1424: complicates the analysis.  There are nevertheless
1425: two possible simplifications.  In the BEC limit,
1426: the terms involving $\tilde D_0^2$ in the denominator
1427: are negligible compared with the first.
1428: In the BCS limit we can confine ourselves to data where
1429:  $r\gg R_{\rm F}$, equivalently $k\gg
1430: k_{\rm F}$. In both cases the correlation function has
1431: the angular dependence
1432: \begin{eqnarray}
1433: {\cal G}(\vec r, -\vec r)
1434: %_{r\gg R_{\rm F}}
1435: &\propto&
1436: %(\frac{R_{\rm F}}{r})^2
1437: (\tilde k_z^2 + \beta^2 \tilde k_x^2)
1438: \end{eqnarray}
1439: Then, for $\hat a = ({\rm cos} \chi) \hat y + ({\rm sin} \chi) \hat x$
1440: as before, we get
1441: \begin{equation}
1442: S_0(\alpha) \propto - D_0^2 \beta^2 {\rm cos} \left( 2 (\chi - \alpha) \right)
1443: \ .
1444: \end{equation}
1445: Thus $S_0(\alpha)$ will reveal the finiteness of $\beta$ as well
1446: as the value of $\chi$ hence the direction of $\hat a$
1447: (up to a sign).  The values of $\beta$ however are still difficult
1448: to determine in this way, both because the need to evaluate
1449: the proportionality coefficient in $S_0$ as well as the
1450: fact that experiments so far detect a correlation in s-wave
1451: substantially less than the expected theoretical value.
1452: However, we can consider instead
1453: %for $r\gg R_{\rm F}$, says, for example, $r \gtrsim 2 R_{\rm F}$,
1454: \begin{eqnarray}
1455: S_1 &\equiv& \langle \cos^2(\hat r\cdot
1456: \hat z) {\cal G}(\vec r,-\vec r)  \rangle
1457: %|_{r\gg R_{\rm F}}
1458:  \\
1459: S_2 &\equiv& \langle \sin^2(\hat r\cdot \hat z) {\cal G}(\vec
1460: r,-\vec r) \rangle
1461: %|_{r\gg R_{\rm F}}
1462: \end{eqnarray}
1463: These give
1464: \begin{eqnarray}
1465: \frac{S_1}{S_2}= \frac{3+\beta^2}{2+4\beta^2}
1466: \end{eqnarray}
1467: and hence the value of $\beta$ can be determined via
1468: \begin{eqnarray}
1469: \beta^2 = {\frac{3-2(S_1/S_2)}{4 (S_1/S_2) - 1}}
1470: \end{eqnarray}
1471: For the polar state, $\beta = 0$ and
1472: $S_1 = 3 S_2/2 $.  For the axial state,  $\beta=1$ with $S_1=2S_2/3$.
1473: 
1474: 
1475: \section{VI. Conclusion}
1476: 
1477: In conclusion, due to the magnetic dipolar interaction between two
1478: fermionic atoms, the superfluidity in trapped gases induced by a
1479: $p$-wave Feshbach resonance is expected to behave very differently
1480: from the more familiar example of $^3$He. The symmetry of the
1481: ground state can be tuned by varying either the density of the gas
1482: or the applied external magnetic field. The phase can be a polar
1483: state or a state intermediate between the polar and the axial
1484: phase. We also propose how the pairing symmetry can be detected
1485: from an experimental investigation.
1486: 
1487: 
1488: 
1489: \section{Acknowledgement}
1490: 
1491: S.K.Y. is very grateful to T.-L. Ho and A. J. Leggett for
1492: useful discussions.
1493: This research is supported by the National Science Foundation
1494: under Grant No. PHY99-07949 and the National Science Council of
1495: Taiwan under Grant Nos. NSC93-2112-M-001-016 and NSC94-2112-M-001-002.
1496: The postdoctoral position of C.H.C. is supported
1497: by NSC under Grant No.
1498: NSC93-2816-M-001-0007-6 and Academia Sinica.
1499: 
1500: 
1501: 
1502: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1503: 
1504: \begin{thebibliography}{}
1505: \bibitem{Dalfovo99}
1506: F. Dalfovo {\it et al.}, \rmp {\bf 71}, 463 (1999) and references
1507: therein.
1508: \bibitem{Castin01}
1509: Y. Castin, in {\it Coherent atomic matter waves}, Les Houches
1510: Session LXXII, NATO Advanced Study Institute, edited by R. Kaiser,
1511: C. Westbrook, and F. David (EDP Science; Spinger-Verlag, 2001) and
1512: references therein.
1513: \bibitem{Stwalley76} W. C. Stwalley, Phys. Rev. Lett. {\bf 37},
1514:   1628 (1976)
1515: \bibitem{Tiesinga92}
1516: E. Tiesinga {\it et al.}, \pra {\bf 46}, R1167 (1992);
1517:   {\it ibid}, \pra {\bf 47}, 4114 (1993)
1518: \bibitem{cross0}
1519: A.J. Leggett, J. Phys. (Paris), Colloq. {\bf 41}, C7-19 (1980).
1520: \bibitem{nozieres}
1521: P. Nozieres and S. Schmitt-Rink, J. Low Temp. Phys. {\bf 59}, 195
1522: (1985).
1523: \bibitem{randeria90} M. Randeria, J.-M. Duan and L.-Y. Shieh,
1524:   Phys. Rev. B {\bf 41}, 327 (1990)
1525: 
1526: \bibitem{ohashi}
1527: Y. Ohashi and A. Griffin, \pra {\bf 67}, 033603 (2003).
1528: 
1529: \bibitem{falco04}
1530: G. M. Falco and H. T. C. Stoof, Phys. Rev. Lett.
1531: {\bf 92}, 130401 (2004)
1532: 
1533: \bibitem{perali}
1534: A. Perali, \prl {\bf 92}, 220404 (2004); {\it ibid}, {\bf 93}, 100404 (2004)
1535: 
1536: \bibitem{java} J. Javanainen {\it et al},
1537:  Phys. Rev. Lett. {\bf 95}, 110408 (2005)
1538: 
1539: \bibitem{numeric} S.-Y. Chang {\it et al}, Phys. Rev. A,
1540:    {\bf 70}, 043602 (2004); G. Astrakharchik {\it et al}
1541:   Phys. Rev. Lett. {\bf 93}, 200404 (2004);
1542:   A. Bulgac {\it et al} cond-mat/0505374 (2005)
1543: 
1544: \bibitem{regal}
1545: C.A. Regal, \prl {\bf 92}, 040403 (2004).
1546: \bibitem{bourdel}
1547: T. Bourdel {\it et al.}, \prl {\bf 93}, 050401 (2004).
1548: \bibitem{zwierlein}
1549: M.W. Zwierlein {\it et al.}, \prl {\bf 92}, 120403 (2004).
1550: \bibitem{bartenstein}
1551: M. Bartenstein {\it et al.}, \prl {\bf 92}, 120401 (2004).
1552: \bibitem{kinast}
1553: J. Kinast {\it et al.}, Science {\bf 307}, 1296 (2005).
1554: \bibitem{Regal03}
1555: C.A. Regal {\it et al.}, \prl {\bf 90}, 053201 (2003).
1556: \bibitem{Ticknor04}
1557: C. Ticknor {\it et al.}, \pra {\bf 69}, 042712 (2004).
1558: \bibitem{Zhang04}
1559: J. Zhang {\it et al.}, \pra {\bf 70}, 030702(R) (2004).
1560: 
1561: \bibitem{Gunter05}
1562: K. G\"unter {\it et al.}, cond-mat/0507632 (2005)
1563: 
1564: 
1565: \bibitem{klinkhamer}
1566: F.R. Klinkhamer and G.E. Volovik, JETP Lett. {\bf 80}, 389 (2004).
1567: \bibitem{ohashi05}
1568: Y. Ohashi, \prl {\bf 94}, 050403 (2005).
1569: \bibitem{ho05}
1570: T.L. Ho and R.B. Diener, \prl {\bf 94}, 090402 (2005).
1571: 
1572: \bibitem{alt}  For an alternative mechanism for p-wave in
1573: trapped Fermi gases, see
1574: L. You and M. Marinescu, Phys. Rev. A {\bf 60}, 2324 (1999).
1575: 
1576: \bibitem{Leggett75}
1577: A.J. Leggett, \rmp {\bf 75}, 331 (1975).
1578: \bibitem{Anderson61}
1579: P.W. Anderson and P. Morel, Phys. Rev. {\bf 123}, 1911 (1961).
1580: 
1581: \bibitem{Cheng05}
1582: C.H. Cheng and S.-K. Yip, \prl {\bf 95}, 070404 (2005).
1583: 
1584: 
1585: \bibitem{gurarie}
1586: V. Gurarie, L. Radzihovsky, and A.V. Andreev, \prl {\bf 94},
1587: 230403 (2005); see also cond-mat/0410620v3.
1588: 
1589: 
1590: %\bibitem{Duine03}
1591: %R.A. Duine and H.T.C. Stoof, J. Opt. B {\bf 5}, S212 (2003).
1592: \bibitem{Fesh}
1593: E. Timmermanns {\it et al.}, Phys. Rep. {\bf 315}, 199 (1999);
1594: M. Holland {\it et al}, Phys. Rev. Lett. {\bf 87}, 120406 (2001)
1595: 
1596: \bibitem{Merzbacher} E. Merzbacher, Quantum Mechanics,
1597: John Wiley and Sons, New York (1970)
1598: 
1599: \bibitem{renorm}
1600: R. Combescot, \prl {\bf 91}, 120401 (2003); G.M. Bruun and C.J.
1601: Pethick, {\it ibid} {\bf 92}, 140404 (2004).
1602: 
1603: \bibitem{diener04}
1604: R.B. Diener and T.-L. Ho, cond-mat/0405174.
1605: 
1606: \bibitem{mackie05}
1607: M. Mackie and J. Piilo, Phys. Rev. Lett.
1608: {\bf 94}, 060403 (2005)
1609: 
1610: \bibitem{szymanska05}
1611: M.H. Szymanska {\it et al.}, cond-mat/0501728.
1612: \bibitem{partridge05} G. B. Partridge {\it et al},
1613: Phys. Rev. Lett. {\bf 95}, 020404 (2005)
1614: 
1615: 
1616: 
1617: \bibitem{chevy}
1618: F. Chevy {\it et al.}, \pra {\bf 71}, 062710 (2005).
1619: 
1620: \bibitem{altman}
1621: E. Altman, E. Demler, and M.D. Lukin, \pra {\bf 70}, 013603
1622: (2004).
1623: \bibitem{greiner}
1624: M. Greiner, C.A. Regal, J.T. Stewart, and D.S. Jin, \prl {\bf 94},
1625: 110401 (2005).
1626: 
1627: 
1628: 
1629: 
1630: \end{thebibliography}
1631: 
1632: %\newpage
1633: 
1634: 
1635: \end{document}
1636: