cond-mat0510672/bec.tex
1:  %% ****** Start of file template.aps ****** %
2: %%
3: %%
4: %%   This file is part of the APS files in the REVTeX 4 distribution.
5: %%   Version 4.0 of REVTeX, August 2001
6: %%
7: %%
8: %%   Copyright (c) 2001 The American Physical Society.
9: %%
10: %%   See the REVTeX 4 README file for restrictions and more information.
11: %%
12: %
13: % This is a template for producing manuscripts for use with REVTEX 4.0
14: % Copy this file to another name and then work on that file.
15: % That way, you always have this original template file to use.
16: %
17: % Group addresses by affiliation; use superscriptaddress for long
18: % author lists, or if there are many overlapping affiliations.
19: % For Phys. Rev. appearance, change preprint to twocolumn.
20: % Choose pra, prb, prc, prd, pre, prl, prstab, or rmp for journal
21: %  Add 'draft' option to mark overfull boxes with black boxes
22: %  Add 'showpacs' option to make PACS codes appear
23: %  Add 'showkeys' option to make keywords appear
24: %\documentclass[aps,prl,preprint,groupedaddress]{revtex4}
25: %\documentclass[aps,prl,preprint,superscriptaddress]{revtex4}
26: \documentclass[aps,pra,twocolumn,groupedaddress,showpacs]{revtex4}
27: \usepackage{epsfig}
28: \usepackage{graphicx}
29: \usepackage{amsfonts}
30: \usepackage{latexsym}
31: \usepackage{bm}
32: %\usepackage{pdfsync}
33: %\usepackage{pdfsync}
34: 
35: 
36: % You should use BibTeX and apsrev.bst for references
37: % Choosing a journal automatically selects the correct APS
38: % BibTeX style file (bst file), so only uncomment the line
39: % below if necessary.
40: %\bibliographystyle{apsrev}
41: 
42: \def\kx{\left( {p k_x \over 2} \right)}
43: \def\ky{\left( {p k_y \over 2} \right)}
44: \def\bk{{\bf k}}
45: \def\bx{{\bf x}}
46: \def\by{{\bf y}}
47: \newcommand{\boldgreek}[1]{\mbox{\boldmath$\mathbf{#1}$\unboldmath}}
48: \def\bpartial{\boldgreek{\partial}}
49: \def\8{\infty}
50: \def\ohm{\frac{1}{2m}}
51: \def\oh{\frac{1}{2}}
52: \def\ot{\frac{1}{3}}
53: \def\oq{\frac{1}{4}}
54: \def\tt{\frac{2}{3}}
55: \def\ft{\frac{4}{3}}
56: \def\tq{\frac{3}{4}}
57: \def\d{\partial}
58: \def\i{\imath\,}
59: \def\ih{\frac{\imath}{2}\,}
60: \def\undertext#1{\vtop{\hbox{#1}\kern 1pt \hrule}}
61: \def\ra{\rightarrow}
62: \def\lfa{\leftarrow}
63: \def\Ra{\Rightarrow}
64: \def\lra{\longrightarrow}
65: \def\ler{\leftrightarrow}
66: \def\lrb#1{\left(#1\right)}
67: \def\O#1{O\left(#1\right)}
68: \def\VEV#1{\left\langle\,#1\,\right\rangle}
69: \def\tr{\hbox{tr}\,}
70: \def\trb#1{\tr\lrb{#1}}
71: \def\dd#1{\frac{d}{d#1}}
72: \def\dbyd#1#2{\frac{d#1}{d#2}}
73: \def\pp#1{\frac{\partial}{\partial#1}}
74: \def\pbyp#1#2{\frac{\partial#1}{\partial#2}}
75: \def\ff#1{\frac{\delta}{\delta#1}}
76: \def\fbyf#1#2{\frac{\delta#1}{\delta#2}}
77: \def\pd#1{\partial_{#1}}
78: \def\br{\\ \nonumber & &}
79: \def\brr{\right. \\ \nonumber & &\left.}
80: \def\inv#1{\frac{1}{#1}}
81: \def\be{\begin{equation}}
82: \def\ee{\end{equation}}
83: \def\bea{\begin{eqnarray} & &}
84: \def\eea{\end{eqnarray}}
85: \def\ct#1{\cite{#1}}
86: \def\rf#1{(\ref{#1})}
87: \def\EXP#1{\exp\left(#1\right)}
88: \def\TEXP#1{\hat{T}\exp\left(#1\right)}
89: \def\INT#1#2{\int_{#1}^{#2}}
90: \def\MAT{{\it Mathematica }}
91: \def\LHS{left-hand side }
92: \def\RHS{right-hand side }
93: \def\COM#1#2{\left\lbrack #1\,,\,#2\right\rbrack}
94: \def\AC#1#2{\left\lbrace #1\,,\,#2\right\rbrace}
95: 
96: \def\grsq{{\bf \nabla}^2}
97: 
98: \def\cH{{\cal H}}
99: \def\rf#1{(\ref{#1})}
100: \def\t{\tilde}
101: \def\cE{{\cal E}}
102: \def\cD{{\cal D}}
103: \def\sign{{\rm sign}}
104: 
105: \def\rfs#1{Eq.~\rf{#1}}
106: 
107: \def\nn{\nonumber}
108: 
109: \begin{document}
110: 
111: % Use the \preprint command to place your local institutional report
112: % number in the upper righthand corner of the title page in preprint mode.
113: % Multiple \preprint commands are allowed.
114: % Use the 'preprintnumbers' class option to override journal defaults
115: % to display numbers if necessary
116: %\preprint{}
117: 
118: %Title of paper
119: %\title{Properties of the BCS-BEC condensate in the BEC regime}
120: \title{Properties of the strongly paired fermionic condensates}
121: 
122: % repeat the \author .. \affiliation  etc. as needed
123: % \email, \thanks, \homepage, \altaffiliation all apply to the current
124: % author. Explanatory text should go in the []'s, actual e-mail
125: % address or url should go in the {}'s for \email and \homepage.
126: % Please use the appropriate macro foreach each type of information
127: 
128: % \affiliation command applies to all authors since the last
129: % \affiliation command. The \affiliation command should follow the
130: % other information
131: % \affiliation can be followed by \email, \homepage, \thanks as well.
132: \author{J. Levinsen}
133: \author{V. Gurarie}
134: 
135: %\email[]{Your e-mail address}
136: %\homepage[]{Your web page}
137: %\thanks{}
138: %\altaffiliation{}
139: \affiliation{Department of Physics, CB390, University of Colorado,
140: Boulder CO 80309}
141: 
142: 
143: %Collaboration name if desired (requires use of superscriptaddress
144: %option in \documentclass). \noaffiliation is required (may also be
145: %used with the \author command).
146: %\collaboration can be followed by \email, \homepage, \thanks as well.
147: %\collaboration{}
148: %\noaffiliation
149: 
150: \date{\today}
151: 
152: \begin{abstract}
153: We study a gas of fermions undergoing a wide resonance $s$-wave
154: BCS-BEC crossover, in the BEC regime at zero temperature. We
155: calculate the chemical potential and the speed of sound of this
156: Bose-Einstein-condensed gas, as well as the condensate depletion, in the low
157: density approximation. We discuss how higher order terms in the low
158: density expansion can be constructed. We demonstrate that the
159: standard BCS-BEC gap equation is invalid in the BEC regime and is
160: inconsistent with the results obtained here. The low
161: density approximation we employ breaks down in the intermediate
162: BCS-BEC crossover region. Hence our theory is unable to predict how
163: the chemical potential and the speed of sound evolve once the
164: interactions are tuned towards the BCS regime. As a part of our
165: theory, we derive the well known result for the bosonic scattering
166: length diagrammatically and check that there are no bound states of
167: two bosons.
168: \end{abstract}
169: 
170: % insert suggested PACS numbers in braces on next line
171: \pacs{03.75.Hh, 03.75.Ss}
172: % insert suggested keywords - APS authors don't need to do this
173: %\keywords{}
174: 
175: %\maketitle must follow title, authors, abstract, \pacs, and \keywords
176: \maketitle
177: 
178: % body of paper here - Use proper section commands
179: 
180: \section{Introduction}
181: 
182: 
183: 
184: In their seminal papers A. Leggett \cite{Leggett1980} and P.
185: Nozi\`eres and R. Schmitt-Rink \cite{Nozieres1985} studied a system
186: of spin-1/2 fermions with attractive short-ranged interactions in
187: the singlet channel at low temperature. If the interactions are
188: weak, the fermions form a Bardeen-Cooper-Schrieffer (BCS)
189: superconductor. If the interactions are made progressively stronger,
190: at some critical interaction strength a bound state of two fermions
191: becomes possible. These bound states are then bosons, which undergo
192: Bose-Einstein condensation (BEC). Studies by these and many other
193: authors demonstrated that as the interaction strength is increasing,
194: the BCS superfluid smoothly evolves into the  BEC superfluid,
195: without undergoing any phase transition in between.
196: 
197: Quite remarkably, Leggett, and Nozi\`eres and Schmitt-Rink afterwards,
198: found that the BCS gap equation, the mean-field equation which
199: describes the properties of the BCS superconductor, evolves under the
200: strengthening of the potential into the Schr\"odinger equation of a
201: bound pair of fermions. This happens despite the fact that the gap
202: equation is derived on the assumption of weak interactions between the
203: fermions and should in principle break down as the interaction
204: strength is increased. Interestingly, this observation was first made
205: as far back as in 1966 by V. N. Popov, \cite{Popov1966}, but his
206: contribution, at least within the context of the BCS-BEC crossover,
207: was not noticed until later. Similarly, the BCS-BEC crossover was
208: first studied in 1969 by D. M. Eagles \cite{Eagles1969} while the two
209: channel model \cite{Timmermans1999} was first studied by Yu. B. Rumer
210: \cite{Rumer1959}.
211: %Popov also
212: %observed the breakdown of the gap equation described in this paper
213: %(see below).
214: 
215: The fact that the gap equation describes the BCS superconductor
216: well, combined with the fact that the gap equation is also well
217: suited to describe the bound pairs of fermions on the opposite side
218: of the crossover, and taking into account that there is no phase
219: transition in between, allowed Nozi\`eres and Schmitt-Rink to
220: conclude that it may be used to interpolate between these two
221: regimes, when the potential is neither too weak nor too strong and
222: the fermions form an intermediate crossover BCS-BEC superfluid.
223: 
224: Interest in this subject was recently revived when the BCS-BEC
225: superfluid was obtained experimentally in the cold atomic systems
226: where the interactions can be tuned with the help of Feshbach
227: resonances
228: \cite{Timmermans1999,Holland2001,Timmermans2001,Hulet2003,Jin2004,Ketterle2004}.
229: Thus for the first time theoretical predictions regarding the BCS-BEC
230: crossover can be tested experimentally.
231: 
232: Given the lack of justification for the gap equation in the BEC
233: regime beyond predicting the binding energy of a pair of fermions,
234: the question remained whether it can be used to predict other
235: properties of the BEC condensate, such as the condensate depletion,
236: excitation spectrum and so on.
237: 
238: %Many subsequent publications of other authors went further than
239: %that, claiming that the gap equation provides not just a qualitative
240: %interpolation between the BCS and the BEC regimes, but also a good
241: %approximation to the actual physics in the intermediate BCS-BEC
242: %regime. These claims, although theoretically unsubstantiated, were
243: %nevertheless adopted in many publications and checked against
244: %numerics with varying success.
245: 
246: In this paper, we show that the gap equation is actually not valid
247: on the BEC side of the crossover. Although correctly predicting the
248: binding energy of the fermionic bound pairs, it fails to describe
249: the interactions of these dimers properly. As a result, if taken
250: literally, it incorrectly predicts the physical properties of the
251: BEC superfluid.
252: 
253: Indeed, it is generally believed in the literature
254: \cite{Leggett1980,Nozieres1985} that the transition between the BCS
255: and the BEC phases occurs without a phase transition making these
256: phases qualitatively the same phase.
257: %Indeed, the lack of phase transition between the BCS and BEC phases
258: %makes them qualitatively the same phase.
259: Thus on a qualitative level,
260: it is possible to use the gap equation to obtain order of magnitude
261: estimates of the parameters of the BEC phase. However, if one is
262: interested in quantitative calculations, one needs to go beyond the
263: gap equation. Barring the exact solution of the problem, which most
264: likely is not possible here, the next best thing is to identify a
265: small parameter and do the calculations as an expansion in powers of
266: that parameter. The gap equation, when written on the BEC side of the
267: crossover, implicitly assumes that the small parameter is the
268: interaction strength between the bosons. This assumption is invalid;
269: it has been known for some time now that the Born approximation fails
270: to describe the interactions between the bosons properly
271: \cite{Petrov2005}.
272: 
273: In this paper, we identify as a small parameter the so-called gas
274: parameter of the Bose gas, and do all calculations as an
275: expansion in powers of that parameter.
276: 
277: The gap equation should be replaced by another equation on the BEC
278: side of the crossover. That equation, described below, cannot be
279: derived exactly, but can only be obtained as an expansion in powers of
280: the gas parameter. In the lowest order approximation, it allows us to
281: compute the chemical potential $\mu_b$, the speed of sound $u$ in the
282: superfluid, and the condensate depletion to be, at zero temperature,
283: \begin{eqnarray} \label{eq:results}
284: &&\mu_b = \frac{4 \pi n_b}{m_b} a_b,\,\,\, u^2=\frac{4 \pi n_b}{m_b^2}
285: a_b, \nn \\
286: &&n_{0,b} = n_b \left[1-\frac{8}{3}
287: \sqrt{\frac{n_b \, a_b^3}{\pi}} \right].
288: \end{eqnarray}
289: Here
290: \begin{equation} \label{eq:shlyap} a_b\approx 0.60 \,a
291: \end{equation}
292: is the scattering length of the bosonic dimers, whose approximate
293: relationship to the scattering length of a pair of fermions $a$ has
294: first been derived by Petrov, Salomon, and Shlyapnikov
295: \cite{Petrov2005}. The coefficient $0.60$ is approximate in the sense
296: of being determined only numerically, although the procedure to find
297: it is in principle exact. $n_b$ is the density of bosons, which in the
298: lowest order approximation in density can be replaced by $n/2$, where
299: $n$ is the density of the fermions. $n_{0,b}$ is the condensate
300: density. $m_b=2m$ is the mass of the bosons, where $m$ is the fermion
301: mass. $\mu_b$ is the chemical potential of the bosons, which can be
302: related to the chemical potential $\mu$ of the fermions by demanding
303: that $\mu_b-2\mu$ coincides with the binding energy released when the
304: boson is formed. Throughout this paper we adopt units in which $\hbar
305: = 1$.
306: 
307: Our results \rfs{eq:results} coincide with the behavior of the
308: dilute interacting Bose gas \cite{AbrikosovBook}. We would like to
309: emphasize that they are not exact, but are obtained as the lowest
310: order approximation in powers of the gas parameter $a
311: n^{\frac{1}{3}}$. It is possible to use the techniques described in
312: this paper and calculate the next order corrections to
313: \rfs{eq:results}. They will no longer necessarily coincide with the
314: next order corrections in the standard dilute Bose gas. We will not
315: calculate them here, however, and limit our discussion with merely
316: indicating how these higher order corrections can be obtained.
317: 
318: 
319: As the BEC condensate is tuned towards the crossover BCS-BEC regime,
320: $a$ increases reaching infinity at the unitary point which lies
321: somewhere in the intermediate BCS-BEC region. Thus the approximation
322: used in this paper breaks down as the intermediate regime is
323: approached. Our technique is unable to tell us anything about the
324: crossover regime, and we will not attempt to study it in this paper.
325: 
326: All calculations in this paper are done at zero temperature. We
327: think that our techniques can also be adopted to finite
328: temperatures, including studies of the critical superfluid transition
329: temperature, either in the large $N$ approximation or using some
330: other technique. See also Ref.~\cite{Andersen2004} for more
331: references on this subject. We will not attempt to do this in this
332: paper.
333: 
334: It is important to note that the interaction potential between
335: fermions is characterized not only by the scattering length $a$, but
336: also by the effective range $r_0$. Unlike $a$ which changes as the
337: interaction potential is tuned through the crossover, $r_0$ remains
338: roughly the same. If $\left| r_0 \right| \gg n^{-\frac{1}{3}}$, then
339: the fermions are said to be in the narrow resonance regime. In that
340: regime, the gap equation is a good approximation to the actual
341: physics of the BCS-BEC crossover everywhere \cite{Gurarie2005}.
342: However, the wide resonance crossover where $\left| r_0 \right| \ll
343: n^{-\frac{1}{3}}$, $\left|r_0 \right| \ll \left|a\right|$ is more
344: relevant for the current experiments, and it is the wide resonance
345: crossover which received most attention in the literature.
346: Everything we described above this paragraph pertains to the wide
347: resonance regime. In order to be certain that we work in the wide
348: resonance regime, we employ the one channel model which guarantees
349: that $r_0$ is vanishingly small.
350: %Alternatively, we could work with
351: %he two channel model, but then to insure small $r_0$ we would have
352: %o work in the limit of large coupling.
353: 
354: An important part of the theory presented in this paper is the
355: diagrammatic derivation of \rfs{eq:shlyap}. It has come to our
356: attention that while our work was in progress, this was derived
357: diagrammatically, independently of us, by I.~V.~Brodsky, M.~Y.~Kagan,
358: A.~V.~Klaptsov, R.~Combescot, and X.~Leyronas, in
359: Ref.~\cite{Brodsky2005}. The techniques these authors employed to
360: arrive at \rfs{eq:shlyap} coincide with the ones discussed here.
361: 
362: Finally, we note that a problem similar to the one studied in this
363: paper was examined by L. V. Keldysh and A. N. Kozlov in the context
364: of excitons in Ref.~\cite{Keldysh1968}. Those authors, however,
365: concentrated on the case of Coulomb interactions of interest for the
366: physics of excitons, while we are interested in short ranged
367: interactions, which have quite  a different physics. In particular,
368: the universal results \rfs{eq:results} do not hold for their system.
369: Additionally, they, as well as Ref.~\cite{Popov1966}, used an
370: operator version of perturbation theory which is hard to generalize
371: beyond the lowest order. In contrast, our theory is based on a
372: functional integral and is easily generalizable to arbitrary order.
373: It is still remarkable that these authors understood the breakdown
374: of the gap equation in the BEC regime long before this became an
375: important issue in the BCS-BEC crossover literature.
376: 
377: The rest of the paper is organized as follows. In section
378: \ref{sec:one} we introduce the formalism and discuss how the
379: calculations of interest to us can be performed. In section
380: \ref{sec:two} we set up a perturbative expansion in powers of the gas
381: parameter. In the next section \ref{sec:three} we perform the actual
382: calculations and derive \rfs{eq:results}.  Section \ref{sec:BCS}
383: discusses an alternative derivation of the results of section
384: \ref{sec:three} and the origin of the breakdown of the BCS-BEC gap
385: equation. Finally, in section \ref{sec:scat}, which is followed by
386: conclusions, we compute the scattering amplitude between bosons and
387: the bosonic scattering length \rfs{eq:shlyap}, a necessary step in the
388: calculations performed in this paper. Some of the technical details of
389: the calculations are discussed in the appendices.
390: 
391: 
392: 
393: \section{Formulation of the problem}
394: \label{sec:one}
395: 
396: Consider a gas of spin-1/2 fermions, interacting with some short
397: ranged interaction in the $s$-wave channel. The most straightforward
398: way to study this gas is by means of a functional integral
399: \begin{equation} \label{eq:partst}
400: Z=\int \cD \bar \psi_\uparrow \cD \psi_\uparrow \cD \bar
401: \psi_\downarrow \cD \psi_\downarrow ~e^{iS_f},
402: \end{equation}
403: where $S_f=S_0+S_{\rm int}$ is the action consisting of the free
404: part
405: \begin{equation} \label{eq:S}
406: S_0=\sum_{\sigma=\uparrow,\downarrow} \int d^3x~ dt~ \bar
407: \psi_{\sigma} \left( i \pp{t}+\frac{1}{2m}\frac{\d^2}{\d {\bf
408: x}^2}+\mu \right) \psi_\sigma
409: \end{equation}
410: and the interaction part \begin{equation} \label{eq:Sint} S_{\rm
411: int} = \lambda \int d^3 x~dt~ \bar \psi_{\uparrow} \bar \psi_{
412: \downarrow} \psi_{\downarrow} \psi_{\uparrow}.
413: \end{equation}
414: This term describes interactions which happen at one point in space.
415: We need to remember that by itself such an interaction is unphysical
416: and has to be supplemented by a momentum cutoff $\Lambda$.
417: $\lambda>0$ to reflect that we choose an attractive interaction
418: potential.
419: 
420: As is standard in the treatment of superconductivity, we introduce
421: the Hubbard-Stratonovich field $\Delta$, which allows us to rewrite
422: \rfs{eq:partst} in the equivalent way
423: \begin{equation} \label{eq:part1}
424: Z=\int \cD \bar \psi_\uparrow \cD \psi_\uparrow \cD \bar
425: \psi_\downarrow \cD \psi_\downarrow \cD \Delta \cD \bar
426: \Delta~e^{iS},
427: \end{equation}
428: where the action $S$ is now given by $S=S_0+S_{\rm HS}$, and
429: \begin{equation} \label{eq:HS}
430: S_{\rm HS}= -\int d^3x~ dt~\left[ \frac{1}{\lambda} \bar \Delta
431:  \Delta +
432: \left( \Delta \bar \psi_{\uparrow} \bar \psi_{\downarrow} + \bar
433: \Delta \psi_{\downarrow} \psi_{\uparrow} \right)
434: \right].\end{equation}
435: The total action $S$ is now quadratic in
436: fermions, so the fermions can be integrated out. This results in the
437: following partition function
438: \begin{equation} \label{eq:part}
439: Z=\int \cD \Delta \cD \bar \Delta~e^{iS_\Delta},
440: \end{equation}
441: where the effective action $S_{\Delta}$ is given by
442: \begin{eqnarray}
443: \label{eq:Seff1} S_{\Delta}&=&-i~\tr \log \left( \matrix { \omega^+
444: -\frac{{\bf p}^2}{2m}+\mu & -\Delta \cr - \bar \Delta & \omega^+ +
445: \frac{{\bf p}^2}{2m}-\mu } \right) \cr & & -\frac{1}{\lambda}
446:  \int d t d^3 x~\bar \Delta \Delta,
447: \end{eqnarray}
448: where $\omega^+=\omega+i0\, \sign~\omega$. The action $S_\Delta$ and
449: the functional integral \rfs{eq:part} represent the starting point
450: in our theory, although sometimes it is also advantageous to keep
451: the fermions explicitly, as in \rfs{eq:part1} and \rfs{eq:HS}.
452: 
453: In the modern literature on Feshbach resonances, the system described
454: in Eqs.\rf{eq:partst}, \rf{eq:S}, and \rf{eq:Sint} is often referred
455: to as the one channel model. The interactions between fermions in this
456: model are such that their effective range $r_0$ is vanishingly small
457: (in the limit of a large momentum cutoff $\Lambda$). And indeed, the
458: results described in this paper are applicable to the limit $|r_0| \ll
459: n^{-{\frac{1}{3}}}$, $\left| r_0 \right| \ll \left|a \right|$ only,
460: thus the one channel model is a natural starting point for our
461: study. Alternatively, one could also study the two channel model
462: \cite{Timmermans1999}. The two channel model involves an additional
463: parameter which allows control of the value of $r_0$ independently of
464: $a$. However, if we chose to work with the two channel model, to
465: ensure the smallness of $r_0$ we would need to study the limit of
466: infinitely strong coupling within that model. In that limit, the
467: distinction between the two channel and one channel models
468: disappear. This is explained in more detail in Appendix
469: \ref{AppendixTC}. Because of this equivalence, we will not study the
470: two channel model any further in this paper.
471: 
472: The standard approach at this stage is to find the extremum of the
473: action $S_\Delta$ with respect to the field $\Delta$. Assuming that
474: the extremum of the action is when the field $\Delta$ takes some
475: constant value $\Delta_0$, we find the BCS-BEC gap equation
476: \begin{equation}
477: \frac{1}{\lambda} =\frac{1}{2} \int \frac{d^3 p}{(2 \pi)^3}
478: \frac{1}{\sqrt{ \left(\frac{ p^2}{2 m} -  \mu \right)^2 + \Delta_0^2
479: }}.
480: \end{equation}
481: The integral in the gap equation is up to momenta of the order of
482: the cutoff $\Lambda$. It is advantageous to trade the interaction
483: strength $\lambda$ for the fermionic scattering length $a$ using the
484: relation
485: \begin{equation}\label{eq:aforlambda}
486: a=\left(-\frac{4 \pi}{m \lambda} + \frac{2
487: \Lambda}{\pi}\right)^{-1}.
488: \end{equation}
489: This relation is fairly standard. For completeness, we include its
490: derivation in Appendix \ref{AppendixB}. Then the BCS-BEC gap
491: equation acquires the form
492: \begin{equation} \label{eq:BCSBEC}
493: -\frac{m}{4 \pi a} =\frac{1}{2} \int \frac{d^3 p}{(2 \pi)^3} \left [
494: \frac{1}{\sqrt{ \left(\frac{ p^2}{2 m} - \mu \right)^2 + \Delta_0^2 }}
495: - \frac{2 m}{p^2} \right].
496: \end{equation} The integral over momentum $p$ is now convergent and
497: can be extended to infinity.
498: 
499: If $a$ is negative and $|a| n^{\frac{1}{3}} \ll 1$ (this corresponds
500: to the weak attraction), then we say that the gas of fermions is in
501: the BCS state. In this regime, the gap equation has been
502: investigated in detail in the literature devoted to superconductors.
503: It was shown that the fluctuations of $\Delta$ about its mean field
504: value $\Delta_0$, which solves \rfs{eq:BCSBEC}, are indeed small,
505: and the physics of the superconductors can be captured by solving
506: \rfs{eq:BCSBEC}.
507: 
508: As the interaction strength is increased, $a$ becomes more and more
509: negative, reaching negative infinity at the interaction strength
510: which corresponds to the threshold of bound state creation. For even
511: stronger potentials, $a$ becomes positive, first large and then it
512: gets smaller. In this regime there is no justification for using
513: \rfs{eq:BCSBEC}.
514: 
515: Yet Ref.~\cite{Nozieres1985} argued that  in the ``deep BEC" regime
516: where $a n^{\frac{1}{3}} \ll 1$, \rfs{eq:BCSBEC} can be
517: reinterpreted as the Schr\"odinger equation of a pair of fermions,
518: with chemical potential playing the role of half of the energy. So
519: the gap equation then correctly predicts the formation of bosonic
520: dimers. Indeed, solving the gap equation in this regime by
521: anticipating that $\mu \ll -\Delta_0$,  we neglect $\Delta_0$  and
522: find within this approximation
523: \begin{equation} \label{eq:bind} \mu = -\frac{1}{2 m
524: a^2}.
525: \end{equation}
526: $1/ma^2$ is the binding energy of two fermions, as expected.
527: 
528: Nevertheless there is no reason to expect that the gap equation
529: holds beyond \rfs{eq:bind}. For example, one could expect that
530: attempting to use it to compute the deviation of $\mu$ from
531: \rfs{eq:bind} due to nonzero $\Delta_0$ in \rfs{eq:BCSBEC} would
532: lead to an incorrect relation between the boson chemical potential
533: $\mu_b=2\mu+1/(ma^2)$ and $\Delta_0$. On the other hand knowing the
534: proper relation between $\mu_b$ and $\Delta_0$ would allow us to
535: compute the physical properties of the BEC condensate, such as the
536: speed of sound. So there is a need to develop a reliable technique
537: which allows computation of these quantities.
538: 
539: 
540: 
541: 
542: In this paper we  investigate the properties of the system in the
543: low density ``deep BEC" regime, that is, in the regime where $a$ is
544: positive and small, so that $a n^{\frac{1}{3}} \ll 1$.  To this end,
545: we will calculate the normal and anomalous propagators of the
546: bosonic $\Delta$ field using a diagrammatic expansion. In doing so,
547: we will keep only the diagrams lowest in the density of the system.
548: However, we will not assume that the interactions are weak, or in
549: other words, that the interactions can be treated in the Born
550: approximation.
551: 
552: 
553: The fact that the Born approximation is not applicable to the
554: bosonic dimers has been known for quite a while. Indeed, the dimer's
555: scattering length in the Born approximation is $a_b=2 a$
556: \cite{Melo1993}, while the correct scattering length is given by
557: \rfs{eq:shlyap}. Although first derived in Ref.~\cite{Petrov2005} by
558: solving the four-body Schr\"odinger equation, it can also be derived
559: by  summing all the diagrams contributing to the scattering exactly.
560: In this paper we
561:  perform this summation and derive \rfs{eq:shlyap}
562: diagrammatically in section \ref{sec:scat}. This allows us to
563: recognize $a_b$ when the diagrams contributing to it appear within
564: the diagrams for the normal and anomalous bosonic propagators in our
565: theory.
566: 
567: Once the bosonic propagators are calculated, we will impose the
568: condition that they have a pole at zero frequency and momentum,
569: representing the sound mode of the condensate. This will give us a
570: relation between the chemical potential $\mu$, the scattering length
571: $a$, and the expectation value $\Delta_0$ of the field $\Delta$,
572: which will replace the gap equation \rfs{eq:BCSBEC}. In the dilute
573: Bose gas literature, this  is usually referred to as the
574: Hugenholtz-Pines relation \cite{Hugenholtz1959}. This equation
575: replaces the gap equation in our theory.
576: 
577: If instead of this proper procedure we had used the gap equation, we
578: would have arrived at \rfs{eq:results} with $a_b=2a$, which is the
579: Born approximation to the dimer scattering length. Thus the gap
580: equation fails because it cannot treat the interaction between
581: dimers beyond Born approximation.
582: 
583: In order to be able to determine the unknown quantities $\mu$ and
584: $\Delta_0$, one also needs another equation. Its role is usually
585: played by the particle number equation, which states that the total
586: particle density is equal to $n$,
587: \begin{equation} \label{eq:pne}
588: n=\frac{1}{V} \int d^3 x~\left[\VEV{\bar \psi_\uparrow
589: \psi_\uparrow}+\VEV{\bar \psi_\downarrow \psi_\downarrow} \right],
590: \end{equation}
591: where $V$ is the volume of the system. In its most naive
592: incarnation, we calculate the particle density by assuming that the
593: field $\Delta$ does not fluctuate and is equal to its mean value
594: $\Delta_0$. Then we find
595: \begin{equation} \label{eq:pne1}
596: n=\int \frac{d^3p}{(2\pi)^3}
597: \left[1-\frac{\frac{p^2}{2m}-\mu}{\sqrt{\left(\frac{p^2}{2m}-\mu\right)^2+\Delta_0^2}}
598: \right].
599: \end{equation}
600: We are going to show that this equation is indeed correct in the
601: lowest order approximation in density. However, if one wants to go
602: beyond the lowest order approximation, \rfs{eq:pne1} acquires
603: nontrivial corrections.
604: 
605: 
606: \section{Diagrammatic Expansion}
607: \label{sec:two}
608: 
609: To set up a diagrammatic expansion, we first write down free
610: propagators which follow from \rfs{eq:HS} or \rfs{eq:Seff1}, taking
611: into account that we work in the BEC regime where $\mu<0$. The
612: propagator of the fermions is given by
613: \begin{equation}
614: G_0(p,\omega)=\frac{1}{\omega-\frac{p^2}{2m}+\mu+i0}.
615: \end{equation}
616: Note that since  $\mu<0$, the propagator is retarded.
617: 
618: The propagator of bosons is, naively, equal to $-\lambda$. However,
619: we also need to include in it the self energy correction due to
620: fermionic loops. In other words, we need to expand the effective
621: action $S_\Delta$ up to quadratic order in $\Delta$ and the
622: corresponding term in the expansion corrects the propagator. This
623: expansion is performed in Appendix \ref{AppendixB}. We find, again
624: taking into account that $\mu<0$,
625: \begin{equation} \label{eq:D0}
626: D_0(p,\omega)=\frac{4 \pi}{m} \frac{1}{a^{-1}-\sqrt{m} \sqrt{-
627: \omega+\frac{p^2}{4m}-2 \mu-i0}}.
628: \label{eq:bosonprop}
629: \end{equation}
630: The bosonic propagator is also retarded, as any propagator of free
631: bosons must be at zero temperature \cite{AbrikosovBook}.
632: 
633: The fact that all the free propagators are retarded crucially
634: simplifies the diagrammatic expansion. Note that in the BCS regime
635: where $\mu>0$ this simplification would not have taken place.
636: Similar effect happens in the more standard dilute Bose gas
637: \cite{AbrikosovBook}.
638: 
639: The diagrammatic technique must also involve the lines beginning or
640: ending in the condensate. These are assigned the value $\Delta_0$.
641: Although $\Delta_0^2$ does not coincide with the density of
642: particles in the condensate, it is proportional to it. In turn, the
643: condensate density is always lower than the total density $n$. Thus
644: the expansion in powers of the density $n$ can be replaced by the
645: expansion in powers of $\Delta_0$. Since each line beginning or
646: ending in the condensate is assigned the value $\Delta_0$,  the
647: expansion in the density can also be understood as the expansion in
648: numbers of lines  beginning or ending in the condensate.
649: 
650: Now we would like to construct self energy corrections to the
651: propagators $G_0$ and $D_0$, to find the low density approximation
652: to the exact fermionic normal and anomalous propagators $G_n$ and
653: $G_a$ and exact bosonic normal and anomalous propagators $D_n$ and
654: $D_a$. We need to know $G_{n,a}$ to calculate the particle density
655: \rfs{eq:pne}. We need to know $D_{n,a}$ to relate the chemical
656: potential $\mu$ to other parameters in the theory.
657: 
658: We denote $\Sigma_{n,a}$ the normal and anomalous self energy
659: corrections to $G_{n,a}$. Keeping with tradition
660: \cite{AbrikosovBook}, we reserve the notation $\Sigma_{11}$ and
661: $\Sigma_{20}$ for the normal and anomalous self energy terms in
662: bosonic propagators $D_{n,a}$. We write
663: $\Sigma=\Sigma^{(1)}+\Sigma^{(2)}+\dots$, where $\Sigma^{(j)}$
664: denotes the contribution to $\Sigma$ from the diagrams with $j$ lines
665: beginning or ending in the condensate.
666: 
667: \begin{figure}[bt]
668: \includegraphics[height=1 in]{sigmaa}
669: \caption{The lowest order contribution in the low density
670:   approximation to the anomalous fermion self energy, $\Sigma_a$. The
671:   dashed line is a condensate line.}
672: \label{sigmaa}
673: \end{figure}
674: The only diagram with one condensate line which contributes to
675: $\Sigma_a$ is shown on Fig \ref{sigmaa}. We find
676: \begin{equation}
677: \Sigma_a^{(1)}=\Delta_0.
678: \end{equation}
679: The reason for the absence of any other diagrams at this order is
680: the absence of vertex corrections in our theory at $\mu<0$ when all
681: the propagators are retarded. Thus it is a feature of the BEC regime
682: only.
683: 
684: There are no normal self energy terms with one external line, thus
685: $\Sigma_n^{(1)}=0$. Computing the normal Green's function $G_n$ with
686: this one self energy correction is equivalent to computing the
687: correlation functions in \rfs{eq:pne} by assuming that $\Delta$ does
688: not fluctuate and is equal to $\Delta_0$. This gives \rfs{eq:pne1}.
689: 
690: \begin{figure}[bt]
691: \includegraphics[height=.7 in]{sigmaan}
692: \caption{The next to leading order contribution in the low density
693:   approximation to the normal fermion self energy.}
694: \label{sigmaan}
695: \end{figure}
696: No diagrams with two condensate lines contribute to the
697: anomalous self energy, and thus $\Sigma_a^{(2)}=0$. The
698: diagrams contributing to the normal self energy at this order
699: are shown in Fig. \ref{sigmaan}. The square block shown on this figure
700: is equal to the sum of all diagrams which contribute to boson-fermion
701: scattering. There is an infinite number of such diagrams, and they can
702: all be summed by solving the Lippmann-Schwinger integral
703: equation. This will be discussed in more details in section
704: \ref{sec:scat}. For now we denote the result of summation as $t^{\rm
705: bf}$. Then the contribution to the normal self energy
706: at this order is given by
707: \begin{equation}
708: \Sigma_{n}^{(2)}=\Delta_0^2 t^{\rm bf}.
709: \end{equation}
710: 
711: %There are substantially more diagrams with two external lines
712: %contributing to $\Sigma_{a,n}$, and they are shown on Fig
713: %\ref{sigmaan}. The square block shown on this figure is equal to the
714: %sum of all diagrams which contribute to boson-fermion
715: %scattering. There is an infinite number of such diagrams, and they can
716: %all be summed by solving the Lippmann-Schwinger integral
717: %equation. This will be discussed in more details in section
718: %\ref{sec:scat}. For now we denote the result of summation as $t^{\rm
719: %bf}$. Then the contribution to the anomalous and normal self energies
720: %at this order is given by
721: %\begin{equation}
722: %\Sigma_{n,a}^{(2)}=\Delta_0^2 t^{\rm bf}.
723: %\end{equation}
724: 
725: 
726: 
727: If needed, diagrams with three or higher number of external lines
728: can also be constructed.
729: 
730: 
731: \begin{figure}[bt]
732: \includegraphics[height=1.5 in]{sigma}
733: \caption{The normal (top) and anomalous (bottom) self energy
734:   contributions to the bosonic propagator. The wavy lines are bosonic
735:   propagators.}
736: \label{sigma}
737: \end{figure}
738: Turning to the bosonic propagator, we find that there are no diagrams
739: contributing to $\Sigma_{11}$ and $\Sigma_{20}$ at the order of
740: $\Delta_0$. At the order of $\Delta_0^2$, the self energy diagrams are
741: shown on Fig \ref{sigma}.  The square box denotes the sum of all the
742: diagrams which contribute to the $T$-matrix of scattering of a boson
743: off another boson.  Just as those for $t^{\rm bf}$, these diagrams can
744: also be summed up using the appropriate Lippmann-Schwinger
745: equation. We denote the result of summation $t^{\rm bb}$. $t^{\rm bb}$
746: is calculated in section \ref{sec:scat}. Thus we find
747: \begin{equation} \label{eq:set}
748: \Sigma_{11}^{(2)} = 2 \Delta_0^2 t^{\rm bb}, \ \Sigma_{20}^{(2)}=
749: \Delta_0^2 t^{\rm bb}.
750: \end{equation}
751: The coefficient $2$ appears as the appropriate combinatorial factor.
752: 
753: 
754: It is also possible to consider diagrams higher order in $\Delta_0$
755: which contribute to the bosonic normal and anomalous self energy.
756: One such diagram, which contributes to $\Sigma_{11}^{(4)}$, is shown
757:  on Fig \ref{highorder}. There are infinitely many similar diagrams
758:  contributing to $\Sigma^{(4)}_{11}$ and $\Sigma_{20}^{(4)}$. We
759:  will not discuss these any further.
760: \begin{figure}[bt]
761: \includegraphics[height=1.2 in]{highorder}
762: \caption{A possible higher order contribution to the bosonic self energy.}
763: \label{highorder}
764: \end{figure}
765: 
766: 
767: 
768: 
769: \section{Calculation of the speed of sound and the condensate depletion}
770: \label{sec:three}
771: 
772: Once the self energy terms are known, the procedure for calculating
773: the chemical potential, the speed of sound, and the condensate
774: depletion is fairly standard.
775: 
776: First of all, we calculate the normal and anomalous bosonic Green's
777: functions $D_{n,a}$, given $\Sigma_{11}$ and $\Sigma_{20}$. The
778: calculation involves solving the appropriate Dyson equation
779: \cite{AbrikosovBook} and the result is
780: \begin{eqnarray}
781: D_n(p,\omega)&=&
782: \frac{D_0^{-1}(-p,-\omega)-\Sigma_{11}(-p,-\omega)}{D(p,\omega)},
783: \cr D_a(p,\omega)&=&\frac{\Sigma_{20}(p,\omega)}{D(p,\omega)},
784: \end{eqnarray}
785: where
786: \begin{eqnarray}
787: D(p,\omega)&=&-\left[\Sigma_{20}(p,\omega) \right]^2+ \left(
788: D_0^{-1}(p,\omega) -\Sigma_{11}(p,\omega)\right)\times \cr && \times
789: \left( D_0^{-1}(-p,-\omega) - \Sigma_{11}(-p,-\omega) \right). \cr
790: &&
791: \end{eqnarray}
792: The Hugenholtz-Pines relation then takes the form
793: \begin{equation}
794: D(0,0)=0.
795: \end{equation}
796: In the lowest order in density, the self energy terms $\Sigma_{11}$
797: and $\Sigma_{20}$ vanish. Then the Hugenholtz-Pines relation reads
798: \begin{equation}
799: D_0^{-1}(0,0)=0.
800: \end{equation}
801: Recalling the definition of $D_0$, \rfs{eq:D0}, we immediately find
802: $\mu=-1/(2 m a^2)$, the same relation as the one which follows from
803: the gap equation, \rfs{eq:bind}.
804: 
805: In the first nonvanishing order  in density, the self energy terms
806: are given by \rfs{eq:set}. We are interested in the self energy at
807: zero momentum and frequency, thus $t^{\rm bb}$ also needs to be
808: computed at zero momentum and frequency. In section
809: \ref{sec:scat} we show how to calculate $t^{\rm bb}$ in vacuum by
810: solving the appropriate Lippmann-Schwinger equation. We, however,
811: need $t^{\rm bb}$ at a finite chemical potential $\mu$. Fortunately,
812: if
813: \begin{equation} \label{eq:impcond} \left| \mu+1/(2 m a^2) \right| \ll \left| \mu \right|,
814: \end{equation} then $t^{\rm bb}$ at a finite $\mu$ is approximately
815: given by its expression computed in vacuum at zero momentum and
816: frequency. In other words, it is proportional to $a_b$, where $a_b$
817: is the bosonic scattering length in vacuum, \rfs{eq:shlyap}. The
818: Lippmann-Schwinger equation can only be solved numerically, so we
819: only know the numerical value of $a_b$.
820: 
821: More precisely, $t^{\rm bb}$ at zero momentum and frequency, and at
822: chemical potential satisfying \rfs{eq:impcond} is related to the
823: scattering length $a_b$ as follows
824: \begin{equation} \label{eq:tbbfrac}
825: t^{\rm bb} = \frac{2 \pi a_b}{m Z^2},
826: \end{equation}
827: where $Z$ is the residue of the bosonic propagator $D_0$  at its
828: pole, or
829: \begin{equation} \label{eq:Z}
830: Z= \frac{8 \pi }{m^2 a}.
831: \end{equation}
832: We expect \rfs{eq:impcond} to hold as we expect $\mu$ to deviate
833: only slightly from its zero density value \rfs{eq:bind} in the low
834: density approximation employed here.
835: 
836: Armed with these relations, we introduce a bosonic chemical potential
837: $$
838: \mu_b=2\mu+\frac{1}{ m a^2} $$ and solve the Hugenholtz-Pines
839: relation in the next order in density
840: \begin{equation} \label{eq:ph}
841: \mu_b=\frac{ a \, a_b \, m \Delta_0^2}{4 }.
842: \end{equation}
843: There are actually two solutions of the Hugenholtz-Pines relation,
844: but one of them is well known to be unphysical \cite{AbrikosovBook}.
845: 
846: We remark that if we had instead decided to solve the gap equation
847: \rfs{eq:BCSBEC} by expanding its square root in powers of
848: $\Delta_0^2$, we would have obtained $\mu_b=\Delta_0^2 \,a^2 m/2$,
849: as if the bosonic scattering length were $a_b=2a$. This is
850: manifestly incorrect. Thus we see that the gap equation fails in the
851: BEC regime. In fact, it is easy to see exactly where the gap
852: equation breaks down, and what comes in its place. This analysis is
853: performed in section~\ref{sec:BCS}.
854: 
855: Now we use the particle number equation \rfs{eq:pne1} to relate
856: $\Delta_0$ to the particle density $n$. Within the lowest
857: approximation in density, we can substitute $\mu=-1/(2 m a^2)$ to
858: find
859: \begin{equation} \label{eq:pne2}
860: n=\frac{a m^2 \Delta_0^2}{4 \pi}.
861: \end{equation}
862: Using the next order approximation for the chemical potential would
863: only contribute to the further expansion of \rfs{eq:pne2} in powers
864: of $\Delta_0$, where, however, we would also need to take into
865: account the terms $\Sigma_{a,n}^{(2)}$. Fortunately this is not
866: needed for the calculations in the lowest order presented here.
867: 
868: Combining \rfs{eq:ph} and \rfs{eq:pne2} we find
869: \begin{equation}
870: \mu_b = \frac{\pi a_b  n}{m} =\frac{4 \pi a_b n_b}{m_b} ,
871: \end{equation}
872: where $m_b=2 m$ is the mass of the bosonic dimers and $n_b=n/2$ is
873: their density. This is the first of the results advertised in the
874: beginning of the paper in \rfs{eq:results}.
875: 
876: 
877: 
878: To find the speed of sound, we impose the condition
879: \begin{equation}
880: D(p,\omega)=0.
881: \end{equation}
882: This gives us a relation between $\omega$ and $p$. For small
883: $\omega$ and $p$, with the help of \rfs{eq:pne2} it reduces to
884: \begin{equation}
885: \omega^2 = p^2 \frac{4 \pi n_b a_b}{m_b^2}.
886: \end{equation}
887: Comparing with $\omega^2=u^2p^2$, where $u$ is the speed of sound,
888: gives the second result in \rfs{eq:results}.
889: 
890: Finally, we calculate the condensate depletion. The density of bosons
891: not in the condensate can be calculated as
892: \begin{equation} \label{eq:cd1}
893: \delta n_b = \frac{i}{Z} \lim_{t \rightarrow - 0} \int \frac{d
894: \omega \, d^3 p}{(2\pi)^4}~ D_n(p,\omega)\,  e^{-i \omega t}.
895: \end{equation}
896: The calculation of the integral in \rfs{eq:cd1} demands knowing the
897: full frequency and momentum dependence of the self energy terms
898: $\Sigma_{11}$ and $\Sigma_{20}$. This in principle can only be found
899: numerically, since the knowledge of full frequency and momentum
900: dependence of $t^{\rm bb}$ is required. However, in the limit of
901: small density where self energy terms are small, the relevant
902: frequency and momentum which contribute to the integral in
903: \rfs{eq:cd1} are in the range $\omega \sim \Delta_0^2$ and $p^2/(4m)
904: \sim \Delta_0^2$. Thus we expand $D_0^{-1}$ in powers of $\Delta_0^2
905: \sim |\omega+p^2/(2m_b)-\mu_b| \ll 1/(m a^2)$ to find the
906: approximate expression for the free bosonic propagator \rfs{eq:D0}
907: \begin{equation}
908: \frac{1}{Z} D_0(p,\omega) \approx \frac{1} {\omega-\frac{p^2}{4m} +
909: \mu_b+i0}.
910: \end{equation}
911: At the same time $Z \Sigma_{20}$ coincides with its expression for
912: the dilute interacting Bose gas. With this substitution,
913: $D_n(p,\omega)$ is equal to that of the dilute Bose gas, with the
914: scattering length $a_b$. Thus the standard formula for the
915: condensate depletion \cite{AbrikosovBook} holds, and it reads
916: \begin{equation} \label{eq:cd3}
917: n_{0,b}=n_b-\delta n_b \approx n_b \left[1-\frac{8}{3}
918: \sqrt{\frac{n_b \, a_b^3}{\pi}} \right],
919: \end{equation}
920: which is the final result in \rfs{eq:results}. We emphasize that this
921: is also an expansion in powers of $na^3$, and that next order terms
922: will likely depend on the functional dependence of the self energy
923: terms on the frequency and momentum.
924: 
925: Let us summarize what one needs to do to go beyond the lowest order
926: approximation discussed above.  The diagrams for the higher order
927: corrections to $\Sigma_{11}$ and $\Sigma_{20}$, such as the one shown
928: on Fig.~\ref{highorder}, have to be evaluated. There will be an
929: infinite number of such diagrams, and they can possibly be summed up
930: using a numerical technique similar to the one employed here to
931: calculate $t^{\rm bb}$. One also needs to calculate higher order
932: corrections to $\Sigma_a$ and $\Sigma_n$. In particular, at the order
933: of $\Delta_0^2$, $t^{\rm bf}$ has to be evaluated. Because of the
934: nonzero chemical potential present in the external lines of the
935: diagrams shown on Fig.~\ref{sigmaan}, $t^{\rm bf}$ will not reduce
936: simply to the scattering length $a_{bf}$, and so extra work will be
937: needed to find the value of
938: $\Sigma_n^{(2)}$. As pointed out by Hugenholtz and Pines in Ref.
939: \cite{Hugenholtz1959}, at each order one also needs to correct the
940: propagators such that a gapless sound mode is produced at this order.
941: 
942: Once the self energy terms are found in the desired order, one needs
943: to solve the Hugenholtz-Pines relation in the next order of
944: $\Delta_0^2$. This also involves taking into account the dependence
945: of $t^{\rm bb}$ on $\mu_b$, which was neglected here due to $\mu_b$
946: being zero up to terms of the order of $\Delta_0^2$. This allows to
947: find the chemical potential. Given the chemical potential, one can
948: also find corrections to the spectrum by looking at the poles of the
949: propagator. For that, the expansion of $t^{\rm bb}$ in powers of
950: momentum will also be needed. This expansion is also required in
951: order to find corrections to the condensate depletion, given at the
952: lowest order in \rfs{eq:cd3}.
953: 
954: %Let us expand the effective action $S_\Delta$ in powers of $\Delta$.
955: %It is straightforward to see that this expansion involves only the
956: %even powers of $\Delta$ and $\bar \Delta$. The first term of this
957: %expansion is proportional to $\bar \Delta \Delta$, and it corrects
958: %the propagator of $\Delta$. The subsequent terms of the fourth,
959: %sixth, etc,  order give effective interactions of $\Delta$ field.
960: 
961: \section{The origin of the failure of the BCS-BEC gap equation}
962: \label{sec:BCS}
963: 
964: An alternative method of deriving the results of the previous section
965: is presented here. A convenient way to derive the BCS-BEC gap equation
966: is by shifting $\Delta$ by $\Delta_0$ in \rfs{eq:part1} and
967: differentiating the resulting $Z$ with respect to $\bar\Delta_0$,
968: assuming that $\Delta_0$ gives an extremum of $\log Z$. This gives
969: \begin{equation} \label{eq:opb}
970: -\frac{\Delta_0}{\lambda} = \VEV{\psi_{\downarrow} \psi_{\uparrow}}.
971: \end{equation}
972: To derive the naive gap equation \rfs{eq:BCSBEC} one usually evaluates
973: the right hand side of \rfs{eq:opb} at constant
974: $\Delta=\Delta_0$. While this procedure works in the BCS regime, it
975: actually breaks down in the opposite BEC regime.
976: 
977: Indeed, on the BEC side of the crossover, it is more appropriate to
978: expand everything in powers of $\Delta_0$. So we instead compute the
979: right hand side of \rfs{eq:opb} in terms of diagrams with a fixed
980: number of condensate lines.
981: 
982: The simplest diagram contributing to \rfs{eq:opb} is shown on
983: Fig.~\ref{fig:cond1}. There are two condensate lines shown on that
984: diagram. The line on the left is not an actual line. Rather it is
985: drawn simply for convenience, and the vertex where it is attached to
986: the diagram represents $\psi_{\downarrow} \psi_{\uparrow}$ whose
987: correlation function we are computing. The line on the right,
988: however, is the actual condensate line. Thus this diagram has one
989: condensate line and is proportional to $\Delta_0$.
990: 
991: \begin{figure}[hbt]
992: \includegraphics[height=.5 in]{cond1}
993: \caption{The simplest diagram contributing to the gap equation
994: \rfs{eq:opb}.} \label{fig:cond1}
995: \end{figure}
996: 
997: If we use this diagram for the right hand side of \rfs{eq:opb}, we
998: derive \rfs{eq:BCSBEC} with the substitution $\Delta_0=0$. Thus this
999: reproduces the lowest order gap equation, and gives \rfs{eq:bind}
1000: for $\mu$.
1001: 
1002: This is the only possible diagram with one condensate line. There are no
1003: diagrams with two condensate lines either. Once we allow for three
1004: external condensate lines, we find one possible diagram shown on
1005: Fig.~\ref{fig:cond2}.
1006: 
1007: \begin{figure}[hbt]
1008: \includegraphics[height=.9 in]{cond2}
1009: \caption{The simplest diagram contributing to the gap equation at
1010:   order $\Delta_0^3$.}
1011: \label{fig:cond2}
1012: \end{figure}
1013: 
1014: This diagram, when taken into account, gives the usual gap equation
1015: \rfs{eq:BCSBEC} where the square root has been expanded up to linear
1016: order with respect to $\Delta_0^2$. We know that solving
1017: \rfs{eq:BCSBEC} at this order gives $a_b=2a$ \cite{Melo1993}. And
1018: indeed, the diagram shown on Fig.~\ref{fig:cond2} is nothing but the
1019: simplest process contributing to the boson-boson $T$-matrix, which is
1020: the very first diagram on Fig.~\ref{fig:diagrams}. In other words, it
1021: is the Born approximation for the boson-boson scattering.
1022: 
1023: We know that the Born approximation breaks down in the BEC regime.
1024: Instead, at this order we need to take all the diagrams with three
1025: external lines, which are equal to $t^{\rm bb}$, as shown on
1026: Fig.~\ref{fig:cond3}.
1027: \begin{figure}[hbt]
1028: \includegraphics[height=.9 in]{cond3}
1029: \caption{The sum of all diagrams contributing to the gap equation at
1030:   order $\Delta_0^3$.}
1031: \label{fig:cond3}
1032: \end{figure}
1033: This gives the correct gap equation, with $2a$ being replaced by
1034: $0.60\,a$.
1035: 
1036: To summarize, instead of using the naive gap equation \rfs{eq:BCSBEC}
1037: which does not take into account fluctuations around $\Delta_0$, in
1038: the BEC regime we use \rfs{eq:opb}, expanding the right hand side in
1039: powers of $\Delta_0$. This procedure goes beyond the Born
1040: approximation and reproduces the correct results in the BEC regime.
1041: 
1042: %If we would like to go to higher orders in $\Delta_0$, we would
1043: %have to consider six boson processes and sum all the diagrams which
1044: %correspond to them. This will deviate even further from the naive
1045: %gap equation \rfs{eq:BCSBEC}.
1046: 
1047: 
1048: 
1049: \section{Scattering of particles and bound states}
1050: \label{sec:scat}
1051: \begin{figure}[ht]
1052: \includegraphics[height=0.36 in]{firstfb}
1053: \caption{Diagrams contributing to the $t$-matrix of fermion-boson
1054:   scattering.}
1055: \label{fig:firstfb}
1056: \end{figure}
1057: In this section we consider the scattering of a fermion by a bound
1058: state of fermions and scattering between bound states of fermions.
1059: All calculations proceed in vacuum so we set $\mu=0$ for this
1060: entire section. As argued in section \ref{sec:three}, these vacuum
1061: scattering processes are precisely those needed to calculate the
1062: normal and anomalous fermion self energies at next to leading order
1063: in the low density expansion and the lowest order boson normal and
1064: anomalous self energies, respectively.
1065: 
1066: The scattering $t$-matrix consists of all diagrams with incoming and
1067: outgoing lines corresponding to the scattering particles. Since we
1068: work in vacuum, there is no hole propagation, or all the propagators
1069: are retarded. These observations greatly simplify the possible
1070: diagrams. In fact, had any of the propagators been advanced, there
1071: would be no hope of summing the diagrams contributing to the
1072: scattering processes exactly. Fortunately, when all the propagators
1073: are retarded, we are able to sum all the diagrams using
1074: Lippmann-Schwinger type integral equations. We emphasize that all the
1075: calculations proceed in the regime of vanishing effective range $r_0$,
1076: so the results of this sections are valid only for potentials where
1077: $|a| \gg |r_0|$.
1078: 
1079: We first consider scattering of a fermion and a bound state of two
1080: fermions, a bosonic dimer. This process was first solved in the zero
1081: effective range approximation by Skorniakov and Ter-Martirosian
1082: \cite{Skorniakov1956} in the related problem of neutron-deuteron
1083: scattering. We include here a somewhat detailed description of this
1084: scattering process in part because it is a useful exercise before
1085: considering the more complicated case of scattering between two
1086: bosons.
1087: \begin{figure}[ht]
1088: \includegraphics[height=1.85 in]{fb}
1089: \caption{The integral equation for the $t$-matrix of fermion-boson
1090:   scattering.} \label{fig:fb}
1091: \end{figure}
1092: 
1093: 
1094: The first few diagrams contributing to the $t$-matrix are shown in
1095: Fig. \ref{fig:firstfb}, where the external lines are not part of the
1096: $t$-matrix. Even though all the propagators are retarded,  there is
1097: still an infinite number of diagrams containing retarded propagators
1098: only, and they all contribute at the same order. Indeed, from these
1099: first examples it is clear that any such contribution to the
1100: $t$-matrix will have $n\geq0$ boson lines of order $a/m$, see Eq.
1101: (\ref{eq:bosonprop}), $2n+1$ fermion lines of order $1/E\sim ma^2$
1102: and $n$ integrations of order $Ep^3\sim m^{-1}a^{-5}$. Thus each
1103: diagram contributing to the $t$-matrix is of order $ma^2$.
1104: 
1105: There is no hope of computing these diagrams one by one and summing
1106: them up. Instead, we derive an integral equation which their sum
1107: satisfies.  Let the incoming boson and fermion have on-shell
1108: 4-momenta $(\vec k, k^2/4m-b)$ and $(-\vec k,k^2/2m)$, respectively,
1109: and let the outgoing boson and fermion have 4-momenta $(\vec
1110: p,p_0+k^2/4m-b)$ and $(-\vec p,-p_0+k^2/2m)$, respectively. Here $b$
1111: is the binding energy of the dimer, $-1/(ma^2)\equiv -b$. The
1112: $t$-matrix with these kinematics is denoted $t^{\rm bf}_{\vec
1113: k}(\vec p,p_0)$. Then the $t$-matrix satisfies the integral equation
1114: \begin{eqnarray}
1115: t^{\rm bf}_{\vec k}(\vec p,p_0) & = & -G_0(\vec k+\vec p,-k^2/4m-b+p_0)
1116: \nn \\ && \hspace{-2.2cm}
1117: -i\int\frac{d^4q}{(2\pi)^4}t^{\rm bf}_{\vec k}(\vec
1118: q,q_0) D_0(\vec q,k^2/4m-b+q_0)\nn \\ && \hspace{-2.2cm} \times
1119: G_0(-\vec q,\frac{k^2}{2m}-q_0)G_0(\vec p+\vec
1120: q,p_0+q_0-\frac{k^2}{4m}-b),
1121: \label{eq:threebody}
1122: \end{eqnarray}
1123: which is depicted in Fig. \ref{fig:fb}. The minus sign in front of
1124: the right hand side of the equation is due to fermionic
1125: anticommutations. The reason for letting outgoing momenta be
1126: off-shell is that this makes it possible to solve the integral
1127: equation. The on-shell point has $|\vec k|=|\vec p|$ and $p_0=0$. It
1128: is possible to integrate out the loop energy on the right hand side
1129: in Eq. (\ref{eq:threebody}) by noting that $t^{\rm bf}_{\vec k}(\vec
1130: q,q_0)$ is analytic in the upper half plane in $q_0$ which can be
1131: easily seen by looking at the diagrams contributing to the
1132: $t$-matrix. This integration sets $q_0=k^2/2m-q^2/2m$ and we thus
1133: set $p_0=k^2/2m-p^2/2m$, to have the same dependence in the
1134: $t$-matrix on both sides in the equation. Define $t^{\rm bf}_{\vec
1135: k}(\vec p,p_0=k^2/2m-p^2/2m)\equiv t^{\rm bf}_{\vec k}(\vec p)$.
1136: 
1137: At low energies scattering is dominated by $s$-wave scattering, thus
1138: Eq. (\ref{eq:threebody}) can be averaged first over directions of
1139: $\vec k$ and then over directions of $\vec p$. The angular average of
1140: $t^{\rm bf}_{\vec k}(\vec p)$ is denoted $t^{\rm bf}_{k}(p)$.  Then
1141: the integral equation for the $t$-matrix of fermion-boson scattering
1142: becomes
1143: \begin{eqnarray}
1144: t^{\rm bf}_k(p) & = & \frac m{2pk}\ln\frac{p^2+pk+k^2-mE}{p^2-pk+k^2-mE}
1145: \nn \\ &&
1146: +\frac1{\pi}\int^\infty_0dq\,\frac qp\ln\frac{q^2+qp+p^2-mE}
1147: {q^2-qp+p^2-mE}
1148: \nn \\ &&
1149: \times \frac{t^{\rm bf}_k(q)}{a^{-1}-\sqrt m\sqrt{-E+3q^2/4m}}.
1150: \label{eq:inteqbf}
1151: \end{eqnarray}
1152: Here, $E=3k^2/4m-b$ is the total energy which is assumed to be less
1153: than zero. To calculate the scattering amplitude each external bosonic
1154: leg has to be renormalized by the square root of the residue of the
1155: pole of the bosonic propagator at the energy of the bound state,
1156: $\sqrt Z=\sqrt{8\pi/(m^2a)}$, compare with \rfs{eq:Z}. The external
1157: fermionic legs are free propagators and do not have to be
1158: renormalized. The scattering amplitude is evaluated on shell and is
1159: \begin{equation}
1160: T^{\rm bf}(k) = Zt^{\rm bf}_{k}(k).
1161: \end{equation}
1162: The relationship between the scattering length and the scattering
1163: amplitude is
1164: \begin{equation}
1165: T^{\rm bf}(0) = \frac{3\pi}ma_{bf}.
1166: \end{equation}
1167: Solving Eq. (\ref{eq:inteqbf}) it is found that $a_{bf}\approx1.18a$.
1168: 
1169: \begin{figure}[ht]
1170: \includegraphics[height=1.85 in]{tinteq}
1171: \caption{The integral equation for the $t$-matrix of boson-boson
1172:   scattering. $\Gamma$ is the sum of all two boson irreducible
1173:   diagrams.} \label{fig:inteqgamma}
1174: \end{figure}
1175: 
1176: We now turn to the scattering between two bosons where the bosons are
1177: bound states of two distinguishable fermions. This process was first
1178: solved by Petrov {\it et al} \cite{Petrov2005} by solving the quantum
1179: mechanical 4-body problem. As we already mentioned, recently, while
1180: the work reported here was in progress, the problem was solved in
1181: Ref.~\cite{Brodsky2005} by using the same diagrammatic approach as we
1182: employ below.
1183: 
1184: As in the case of fermion-boson scattering there are no condensate
1185: lines internally in the diagrams and no hole propagation.
1186: 
1187: \begin{figure*}[hbt]
1188: \includegraphics[height=0.6 in]{diagrams}
1189: \caption{The simplest diagrams contributing to the two boson
1190:   irreducible diagram, $\Gamma$. The first diagram on the right hand
1191:   side corresponds to $\Gamma^{(0)}$ defined in Eq. (\ref{gamma0}).}
1192:   \label{fig:diagrams}
1193: \end{figure*}
1194: 
1195: Consider the different diagrams which contribute to the $t$-matrix of
1196: boson-boson scattering. The $t$-matrix consists of all possible
1197: diagrams with two incoming and two outgoing bosons, the external legs
1198: not included. The crucial point is that all of these diagrams are of
1199: the same order, namely they are all proportional to $m^3a^3$ which we
1200: will show below. This means that the diagrams contributing to the
1201: $t$-matrix do not form a perturbation series and it is not adequate to
1202: keep only the Born approximation to the $t$-matrix. Instead an
1203: infinite number of diagrams must be taken into account. The summation
1204: of all diagrams contributing to the $t$-matrix may be performed by
1205: using an integral equation. In particular, the $t$-matrix may be built
1206: from a series of two boson irreducible diagrams, which results in an
1207: integral equation as shown in Fig. \ref{fig:inteqgamma}. Two boson
1208: irreducible diagrams are understood to be diagrams which cannot be cut
1209: in half by cutting two boson lines only.
1210: 
1211: Let the kinematics be as depicted on the figure with the incoming
1212: energies and momenta chosen on-shell and the outgoing energies and
1213: momenta off-shell to allow for the solution of the integral
1214: equation. The on-shell condition is $|\vec p|=|\vec k|$ and
1215: $p_0=0$. The $t$-matrix with kinematics as shown on the left hand side
1216: of Fig. \ref{fig:inteqgamma} is denoted by $t^{\rm bb}_{\vec k}(\vec
1217: p,p_0)$. The two boson irreducible diagram with $(\pm \vec q,E/2\pm
1218: q_0)$ incoming 4-momenta and $(\pm \vec p,E/2\pm p_0)$ outgoing
1219: 4-momenta is given by $\tilde\Gamma_k(\vec q,q_0;\vec p,p_0)$. $E$ is
1220: the total energy, $E=-2b+k^2/2$. The integral equation corresponding
1221: to these kinematics is
1222: \begin{eqnarray}
1223: t^{\rm bb}_{\vec k}(\vec p,p_0) & = & \tilde\Gamma_k(\vec k,0;\vec p,p_0) \nn
1224: \\ && \hspace{-1cm}+i\int\frac{d^4q}{(2\pi)^4}D_0(\vec q,E/2+q_0)D_0(-\vec q,
1225: E/2-q_0)\nn \\ && \hspace{-1cm}\times\tilde\Gamma_k(\vec q,q_0,\vec p,p_0)t^{\rm
1226: bb}_{\vec k}(\vec q,q_0).
1227: \label{inteqgamma}
1228: \end{eqnarray}
1229: Thus far the integral equation does not depend on whether the
1230: constituents of the dimers are bosons or fermions. Since the
1231: scattering is dominated by $s$-wave scattering we average
1232: Eq. (\ref{inteqgamma}) over directions of $\vec k$, then directions of
1233: $\vec p$ and finally the integration over directions of $\vec q$ can
1234: be trivially performed. Let the corresponding angular averages be
1235: $t^{\rm bb}_k(p,p_0)$ and $\Gamma_k(q,q_0;p,p_0)$ and let $k\equiv |\vec k|$,
1236: $p\equiv |\vec p|$, and $q\equiv |\vec q|$. Then the integral equation
1237: becomes
1238: \begin{widetext}
1239: \begin{equation}
1240: t^{\rm bb}_k(p,p_0) = \Gamma_k(k,0;p, p_0)+\frac{4i}\pi
1241: \int
1242: \frac{\Gamma_k(q,q_0;p,p_0)t^{\rm bb}_k(q,q_0) q^2dq\,dq_0}{1+
1243: \sqrt{(q^2/4-E/2)^2-q_0^2}-\sqrt2\sqrt{q^2/4-E/2+
1244: \sqrt{(q^2/4-E/2)^2-q_0^2}}},
1245: \label{inteqgammas}
1246: \end{equation}
1247: \end{widetext}
1248: which contains no remaining angular dependence and where the product
1249: of bosonic propagators has been written explicitly using
1250: Eq. (\ref{eq:bosonprop}). Since all diagrams are of the same order of
1251: magnitude, it is convenient to let all momenta be measured in units of
1252: $a^{-1}$ and energies in units of $b=1/(ma^2)$. $t^{\rm bb}$ and
1253: $\Gamma$ are then measured in units of $m^3a^3$. For simplicity this
1254: is the case in Eq. (\ref{inteqgammas}) and will be the case in the
1255: remainder of this section.
1256: 
1257: \begin{figure*}[hbt]
1258: \includegraphics[height=.69 in]{xiinteq}
1259: \caption{Integral equation for the two boson irreducible diagram with
1260:   two bosons coming in and one boson and two fermions going out.}
1261: \label{fig:inteqxi}
1262: \end{figure*}
1263: 
1264: The simplest two boson irreducible diagrams are shown in
1265: Fig. \ref{fig:diagrams} where the external lines are not a part of
1266: $\Gamma$. It is clear that there is an infinite number of such
1267: diagrams. Letting $n$ count the number of bosonic propagators in the
1268: diagram contributing to $\Gamma$, any such diagram has $4+2n$
1269: fermionic propagators, each of order $ma^2$, $n$ bosonic propagators
1270: of order $a/m$ and $n+1$ integrations of order $m^{-1}a^{-5}$. Thus
1271: each of these diagrams is of order $m^3a^3$. To see that $t^{\rm bb}$
1272: is also of order $m^3a^3$ note that any diagram contributing to
1273: $t^{\rm bb}$ will contain $n+1$ factors of $\Gamma$, $2n$ bosonic
1274: propagators, and $n$ integrations, resulting in any diagram
1275: contributing to $t^{\rm bb}$ being proportional to $m^3a^3$.
1276: 
1277: Letting the first diagram on the left in Fig. \ref{fig:diagrams} equal
1278: the $t$-matrix corresponds to the Born approximation which gives
1279: $a_b=2a$. Using this same diagram for $\Gamma$ reproduces the result
1280: of Ref. \cite{Pieri2000}, $a_b=0.78a$. However in both of these methods
1281: an infinite number of diagrams of the same order as those considered
1282: is ignored.
1283: 
1284: The sum of two boson irreducible diagrams may again be obtained
1285: through solving an integral equation. The integral equation to be
1286: solved is depicted in Fig. \ref{fig:inteqxi} where $\Xi$ denotes all
1287: two boson irreducible diagrams with two incoming bosons, one outgoing
1288: boson and two outgoing fermions. The two boson irreducible diagram
1289: $\Gamma$ may then be obtained by tying together the two external
1290: fermion lines.  Furthermore, since the two boson irreducible diagram
1291: $\Gamma$ does not depend on angles, we choose to let $\Xi$ be averaged
1292: over the direction of the incoming momentum.  Define a four vector
1293: $K\equiv (\vec 0,E/2)$ which describes the incoming energy carried by
1294: a boson. Then the integral equation corresponding to
1295: the kinematics shown on the figure is
1296: \begin{widetext}
1297: \begin{eqnarray}
1298: \Xi_k(q,p_1,p_2) & = & -\int\frac{d\Omega_{\vec q}}{4\pi}
1299: \left[G_0 \left(\frac{K}{2}+q-p_1 \right)G_0\left( \frac K2 -q-p_2 \right)+G_0\left(\frac K2+q-p_2 \right)
1300: G_0\left(\frac K2-q-p_1 \right)\right] \nn \\
1301: && \hspace{-3.1cm}-i\int\frac{d^4Q}{(2\pi)^4}G_0\left(\frac{K}2+Q
1302: \right)G_0\left(\frac{K}2-Q-p_1-p_2\right)
1303: \left[\Xi_k(q,p_1,Q)D_0(K-Q-p_1)+\Xi_k(q,Q,p_2)D_0(K-Q-p_2)\right].
1304: \label{inteqxi}
1305: \end{eqnarray}
1306: \end{widetext}
1307: The minus signs in front of both the first order terms and the higher
1308: order terms are a result of anti-commuting fermions. The integral over
1309: $d\Omega_{\vec q}$ is the averaging over angles of incoming momentum.
1310: $\Gamma$ is then found from $\Xi$ by connecting the outgoing fermion
1311: lines and integrating over the loop momentum. The precise relation
1312: between $\Gamma$ and $\Xi$ is
1313: \begin{eqnarray}
1314: \Gamma_k(q;p) & = & \frac i2\int\frac{d^4q'}
1315: {(2\pi)^4}G_0(K/2+q')G_0(K/2+p-q') \nn \\
1316: && \times \Xi_k(q,q',p-q').
1317: \label{gammaxi}
1318: \end{eqnarray}
1319: It is not essential to include both lowest order contributions to
1320: $\Xi$ in Eq. (\ref{inteqxi}), the only difference if only one of these
1321: were included is that then the factor $1/2$ in Eq. (\ref{gammaxi})
1322: should be removed. However, the symmetric structure of
1323: Eq. (\ref{inteqxi}) means that
1324: \begin{equation}
1325: \Xi_k(q,p_1,p_2) = \Xi_k(q,p_2,p_1)
1326: \label{eq:symmetry}
1327: \end{equation}
1328: which will be useful in the solution of these equations. How we solve
1329: the set of equations (\ref{inteqgammas}),(\ref{inteqxi}), and
1330: (\ref{gammaxi}) is described in detail in Appendix \ref{app:scat}.
1331: 
1332: As in the case of fermion-boson scattering, to calculate the
1333: scattering amplitude each external bosonic leg has to be renormalized
1334: by $\sqrt Z=\sqrt{8\pi/(m^2a)}$. The scattering amplitude is then
1335: evaluated on shell
1336: \begin{equation}
1337: T^{\rm bb}(k) = Z^2 t^{\rm bb}_k(k,0).
1338: \end{equation}
1339: with $k\equiv |\vec k|$. Here units are restored to the
1340: $t$-matrix. The scattering amplitude is related to the boson-boson
1341: scattering length by
1342: \begin{equation}
1343: T^{\rm bb}(0) = \frac{2\pi}ma_b.
1344: \end{equation}
1345: By solving Eq. (\ref{inteqgammas}) using the method described above
1346: and in Appendix \ref{app:scat} we find $a_b\approx0.60a$ in complete
1347: agreement with Refs. \cite{Petrov2005,Brodsky2005}.
1348: 
1349: Using the above formalism it may also be checked that there are no
1350: bound states of a pair of bosons. If such bound states exist, the
1351: interaction between the bosons in this theory can become
1352: effectively attractive, rendering the bosonic gas unstable. A bound
1353: state of two bosons would correspond to a pole in the scattering
1354: amplitude, that is, to a solution of the homogenous version of the
1355: integral equation (\ref{inteqgamma}) for the $t$-matrix. Varying the
1356: total energy, that is $k^2$ for $k^2\leq0$, we do not find any
1357: solutions to the homogenous equation and thus do not find any bound
1358: states.
1359: 
1360: It should be mentioned that the interaction between bosons does not
1361: have to be renormalized, in contrast to the interaction between bound
1362: states of bosons \cite{Efimov1971}. In the latter case, the
1363: interaction between a boson and a dimer of bosons depends on the
1364: ultraviolet cutoff and a possible solution to this difficulty is to
1365: add a three body force counterterm to the theory and use an additional
1366: input, such as an experimentally measured three body scattering length
1367: \cite{Bedaque1998}. However, this difficulty is absent in the case of
1368: scattering between a fermion and a dimer of fermions and also in the
1369: problem of scattering between two dimers of fermions.
1370: 
1371: 
1372: 
1373: 
1374: \section{Conclusion}
1375: 
1376: In this paper we discussed the low density expansion of the BCS-BEC
1377: condensate in the BEC regime. We found the dispersion of its
1378: Bogoliubov modes (the speed of sound), its chemical potential, and
1379: the condensate depletion in the lowest order approximation in powers
1380: of the gas parameter $a n^{\frac{1}{3}}$. Notice that the gas
1381: parameter increases as the system is tuned towards the BCS-BEC
1382: crossover regime, eventually reaching infinity at the so-called
1383: unitary point lying in between the BCS and BEC condensate. So the
1384: theory developed here works in the BEC regime only and breaks down
1385: in the crossover area.
1386: 
1387: Throughout the paper we emphasized that the BCS-BEC gap equation
1388: actually breaks down in the BEC regime. We would like to remark
1389: further on the origin of this breakdown. The gap equation is derived
1390: by minimizing the effective action of the condensate, given by
1391: \rfs{eq:Seff1}. This is correct only if the fluctuations about this
1392: minimum are small. If they are not small, the fluctuations must be
1393: taken into account. In that case, minimization of the effective action
1394: should be replaced by minimization of the effective potential, which
1395: takes into account fluctuations, as discussed in any textbook on field
1396: theory. In practice, it is possible to replace the evaluation of the
1397: minimum of the effective potential by the Hugenholtz-Pines relation,
1398: as is done here. The advantage of this second procedure is in the fact
1399: that it automatically gives the excitation spectrum in addition to the
1400: chemical potential. The two techniques are however equivalent. Most
1401: importantly, the calculations performed in this paper demonstrate that
1402: in including the fluctuations into the effective potential, one needs
1403: to go well beyond the Gaussian fluctuations approximation. In fact,
1404: terms up to infinite order (in terms of the number of loops) have to
1405: be summed up, all of them being of the same order in the gas
1406: parameter.  Fortunately, this is possible to do, and this is what has
1407: been accomplished in this paper. Alternatively,
1408: the gap equation corrected by fluctuations was calculated in
1409: section~\ref{sec:BCS}.
1410: %%change
1411: We demonstrated how the naive gap equation breaks down due to
1412: fluctuations, and how the fluctuations effectively replace the Born
1413: approximation with the full boson-boson scattering amplitude. If we
1414: wanted, we could follow the procedure outlined in
1415: section~\ref{sec:BCS} as an alternative to the Hugenholtz-Pines
1416: relation.
1417: 
1418: The same argument also demonstrates that the Gross-Pitaevskii equation
1419: of the condensate should follow not from the effective action, but
1420: from the effective potential. In other words, the true bosonic
1421: scattering length \rfs{eq:shlyap} should be used in its quartic term,
1422: as opposed to its Born approximation value $2 a_b$.  The
1423: Gross-Pitaevskii equation will of course be valid only at length
1424: scales much bigger than $a$.
1425: 
1426: All calculations in this paper are done in the lowest approximation
1427: in density, which significantly simplifies the work needed to be
1428: done. Finding higher order corrections to \rfs{eq:results} is one
1429: possible direction of further work along the lines discussed in this
1430: paper. Another possible direction is to study the condensate at
1431: finite temperature. This should probably be done in the large $N$
1432: approximation such as the one used in Ref.~\cite{Baym2000}.
1433: Generalizing the techniques of our paper to finite temperature
1434: should be a promising direction of further research.
1435: 
1436: On the other hand, we do not expect that these techniques will help
1437: to shed light on the BCS-BEC crossover regime, especially at the
1438: unitary point. The small parameter utilized here becomes infinity
1439: there. The unitary regime can only be understood numerically
1440: \cite{Bulgac2005}, barring an invention of an exact solution.
1441: 
1442: 
1443: 
1444: \begin{acknowledgments}
1445: We thank J. Shepard, L. Radzihovsky, D. Sheehy, and A. Lamacraft for
1446: useful discussions. This work was supported by the NSF grant
1447: DMR-0449521.  J.~L. also wishes to thank the Danish Research Agency
1448: for support.
1449: \end{acknowledgments}
1450: 
1451: \appendix
1452: 
1453: \section{The Two channel model}
1454: \label{AppendixTC}  The two channel model is defined by its
1455: functional integral
1456: \begin{equation} \label{eq:tc1}
1457: Z=\int \cD \bar \psi_\uparrow \cD \psi_\uparrow \cD \bar
1458: \psi_\downarrow \cD \psi_\downarrow \cD b \cD \bar b~e^{iS_{\rm
1459: tc}},
1460: \end{equation}
1461: where the action $S_{\rm tc}$ is  given by $S_{\rm
1462: tc}=S_0+S_{0b}+S_{ab}$. Here $S_0$ is the free fermion action given
1463: in \rfs{eq:S}. $S_{0b}$ is the free boson action given by
1464: \begin{equation} \label{eq:Sb}
1465: S_{0b}= \int d^3x~ dt~ \bar b \left( i
1466: \pp{t}+\frac{1}{4m}\frac{\d^2}{\d {\bf x}^2}+2\mu -\epsilon_0\right)
1467: b,
1468: \end{equation}
1469: where $\epsilon_0$ is the detuning, the parameter controlled by the
1470: magnetic field in the Feshbach resonance setup. $S_{ab}$ is the
1471: interaction term
1472: \begin{equation}
1473: S_{ab}= -g \int d^3 x~dt ~ \left( b \, \bar \psi_{\uparrow} \bar
1474: \psi_{\downarrow} + \bar b \, \psi_{\downarrow} \psi_{\uparrow}
1475: \right).\end{equation} The scattering of fermions  in the two
1476: channel model is characterized by the scattering length $a$ and
1477: effective range $r_0$ given by \cite{Andreev2004}
1478: \begin{equation}
1479: a=-\frac{m g^2}{4 \pi \left(\epsilon_0-\frac{g^2 m \Lambda}{2 \pi^2}
1480: \right)}, \ r_0=-\frac{8\pi}{m^2 g^2}.
1481: \end{equation}
1482: In order to ensure that we work in the wide resonance regime $|r_0|
1483: \ll n^{-{\frac{1}{3}}}$, $|r_0| \ll |a|$, in this paper we need to
1484: take the limit $g \rightarrow \infty$, while simultaneously
1485: adjusting $\epsilon_0$ so that $a$ remains finite. Introducing the
1486: notation $\Delta=g b$, $\bar \Delta=g \bar b$, we find that $g$
1487: disappears from $S_{ab}$, while $S_{0b}$, in the large $g$ limit,
1488: becomes
1489: \begin{equation}
1490: S_{0b} = \left(\frac{m}{4 \pi a} - \frac{m \Lambda}{2 \pi^2}
1491: \right)\int d^3 x dt \, \bar \Delta \Delta.
1492: \end{equation}
1493: We recognize the combination of parameters in brackets as
1494: $-1/\lambda$, thanks to \rfs{eq:aforlambda}. Thus the two channel
1495: model reduces to the one channel model as given by \rfs{eq:S} and
1496: \rfs{eq:HS}.
1497: 
1498: One important lesson which follows from this discussion is that in the
1499: wide resonance regime, it is not justified to use $g$ as a small
1500: parameter and construct perturbative expansion in its powers.  Indeed,
1501: $g$ is not only not small, it should actually be taken to infinity.
1502: 
1503: \section{Fermionic Scattering and renormalized bosonic propagator}
1504: \label{AppendixB}
1505: \begin{figure}[hbt]
1506: \includegraphics[height=.35 in]{ff}
1507: \caption{The bubble diagrams of fermion scattering.}
1508: \label{fig:ff}
1509: \end{figure}
1510: The scattering of fermions governed by the action Eq. (\ref{eq:S})
1511: and interacting via the short range interaction Eq. (\ref{eq:Sint})
1512: can be calculated by summing up the diagrams depicted in Fig.
1513: \ref{fig:ff}. The calculation proceeds in vacuum, so the chemical
1514: potential $\mu$ is set to zero everywhere. This sum gives the
1515: fermion $T$-matrix, $T^{\rm ff}$, and the scattering length is
1516: proportional to the $T$-matrix at zero momentum,
1517: \begin{equation}
1518: a = \frac m{4\pi}T^{\rm ff}(0).
1519: \end{equation}
1520: The bubble diagrams form a geometric series which can be summed to
1521: give
1522: \begin{equation}
1523: T^{\rm ff}(0) = \frac{-\lambda}{1+\lambda\Pi(0)}
1524: \end{equation}
1525: where $\Pi$ is a bubble as shown in Fig. \ref{fig:ff} evaluated at
1526: zero momentum. This bubble is given by
1527: \begin{equation}
1528: \Pi(0)  =  i\int\frac{d^4p}{(2\pi)^4}G_0(p)G_0(-p) \nn \\
1529: =  -\frac{m\Lambda}{2\pi^2}.
1530: \end{equation}
1531: Here, $\Lambda$ is the momentum cut-off. This shows
1532: Eq. (\ref{eq:aforlambda}), namely that
1533: \begin{equation}
1534: a = \left(-\frac{4\pi}{m\lambda}+\frac{2\Lambda}\pi\right)^{-1}.
1535: \end{equation}
1536: 
1537: \begin{figure}[hbt]
1538: \includegraphics[height=.9 in]{bosonprop}
1539: \caption{The bosonic propagator renormalized by fermion loops. The
1540:   thin wavy lines are unrenormalized boson propagators.}
1541: \label{fig:bosonprop}
1542: \end{figure}
1543: The calculation of the bosonic propagator corrected by fermionic
1544: loops proceeds similarly. Having in mind its applications in this
1545: paper, we do this calculation at a finite chemical potential $\mu
1546: \le 0$. Expanding the action $S_\Delta$ in Eq. (\ref{eq:Seff1})
1547: results in a geometric series as shown in Fig. \ref{fig:bosonprop},
1548: which can be summed to give
1549: \begin{equation}
1550: D_0(p) = \frac{-\lambda}{1+\lambda\Pi(p)}.
1551: \label{eq:renormbos}
1552: \end{equation}
1553: The fermion loop can be calculated to give
1554: \begin{eqnarray} \label{eq:polarization}
1555: \Pi(p) & = & i\int\frac{d^4q}{(2\pi)^4}G_0(p+q)G_0(-q) \nn \\
1556: && \hspace{-1cm}=-\frac{m\Lambda}{2\pi^2}+\frac{m^{3/2}}{4\pi}\sqrt{-\omega+p^2/4m-2\mu-i0}.
1557: \end{eqnarray}
1558: Inserting this result in Eq. (\ref{eq:renormbos}), using Eq.
1559: (\ref{eq:aforlambda}) to write the momentum cut-off in terms of $a$,
1560: finally gives
1561: \begin{equation} \label{eq:bosonicpropagator}
1562: D_0(p) = \frac{4\pi}m\frac1{a^{-1}-\sqrt m\sqrt{-\omega+\frac{p^2}{4m}-2\mu-i0}}.
1563: \end{equation}
1564: 
1565: Note that it is crucial that $\mu \le 0$ for the integral in
1566: \rfs{eq:polarization} to be calculated the way it is. Otherwise, the
1567: fermionic propagator is no longer retarded due to hole propagation.
1568: Thus \rfs{eq:bosonicpropagator} applies only in vacuum or in the BEC
1569: regime where $\mu \le 0$.
1570: 
1571: \section{The integral equation for boson-boson scattering}
1572: \label{app:scat}
1573: In this appendix we describe the solution of
1574: Eqs. (\ref{inteqgammas}-\ref{gammaxi}), describing scattering of two
1575: bound states, each consisting of two fermions. We demonstrate how to
1576: integrate out the loop energies in constructing the sum of two boson
1577: irreducible diagrams, $\Gamma$, and we write the corresponding
1578: equations in a more convenient way for numerical studies. We also
1579: describe the numerical methods used. For simplicity, in this appendix
1580: all momenta are measured in units of $a^{-1}$, all energies in units
1581: of $1/(ma^2)$, and $t^{\rm bb}$ and $\Gamma$ in units of $m^3a^3$.
1582: 
1583: \begin{figure}[hbt]
1584: \includegraphics[height=1.8 in]{contour}
1585: \caption{The original and rotated contours of frequency integration used
1586:   in Eq. (\ref{inteqgammas}). The crosses correspond to poles of the
1587:   bosonic propagators which approach the imaginary axis as $k,q\to0$.
1588:   The dots and black lines towards $\pm\infty$ are branch cuts of the
1589:   bosonic propagators and the grey boxes are possible non-analytic
1590:   structure of $\Gamma(q,q_0;p,p_0)$. These do not come closer to the
1591:   imaginary axis than $\pm b$.}
1592: \label{fig:contour}
1593: \end{figure}
1594: To construct $\Gamma_k(q,q_0;p,p_0)$ it is convenient first to rotate
1595: the external energies $p_0$ and $q_0$ onto the imaginary axis which
1596: can be done without crossing any poles. Both applications we have in
1597: mind, calculation of the scattering length and search for bound
1598: states, require the $t$-matrix $t^{\rm bb}_k(p,p_0)$ evaluated on
1599: shell and thus we are only interested in the $t$-matrix evaluated at
1600: $p_0=0$ which is still on the contour.  That it is possible to rotate
1601: the external energies is not immediately obvious. Consider
1602: Eq. (\ref{inteqgammas}) which relates the $t$-matrix to $\Gamma$. The
1603: two bosonic propagators have branch cuts in $q_0$ starting at
1604: $q_0=\pm(1+q^2/4-k^2/4)$ and going towards $\pm\infty$,
1605: respectively. $1+q^2/4-k^2/4>0$ since to calculate the scattering
1606: length we let $k=0$ and to search for bound states we let
1607: $k^2\leq0$. The bosonic propagators also have poles at
1608: $q_0=\pm(q^2/4-k^2/4-i0)$ where the infinitesimal is determined from
1609: the requirement that Eq. (\ref{eq:bosonprop}) describes a retarded
1610: propagator. It will be shown below that also $\Gamma(q,q_0;p,p_0)$ can
1611: only be non-analytic in quadrant $II$ above the branch cut and $IV$
1612: below the branch cut in $p_0$ and $q_0$. We conclude that not only is
1613: it possible to rotate the external energies, but we also move the
1614: contour of integration away from any singularities. This is
1615: illustrated in Fig. \ref{fig:contour}. The only remaining
1616: singularities close to the integration contour are from the poles of
1617: the bosonic propagator as $q\to0$ and $k\to0$. This singularity is
1618: integrable and we will explicitly treat this below by a change of
1619: variables in Eq. (\ref{eq:change}).
1620: 
1621: In the integral equation (\ref{inteqxi}) it is possible
1622: to integrate over the loop energy since $\Xi_k(q,p_1,p_2)$ is analytic
1623: in the lower half planes of both fermion energies. To see this, note
1624: that for any diagram contributing to $\Xi$ the two outgoing fermionic
1625: propagators do not originate from the same bosonic propagator since
1626: then $\Xi$ would be two boson reducible. Instead each of the fermions
1627: originate from a boson which in turn decays into one other fermion
1628: which through a series of propagations forward in time contribute to
1629: the creation of the outgoing boson. Thus $\Xi$ contains only retarded
1630: propagators of $-p_1$ and $-p_2$ and must be analytic in the lower
1631: half planes of the corresponding energies.
1632: 
1633: It is not as simple to integrate out the loop energy in
1634: Eq. (\ref{gammaxi}) since $\Xi_k(q,q',p-q')$ is not analytic in either
1635: $q'_0$ half planes. However, if one considers the integral equation
1636: satisfied by $\Xi_k(q,q',p-q')$, Eq. (\ref{inteqxi}) with $p_1=q'$ and
1637: $p_2=p-q'$, the first product of propagators can be split into a term
1638: analytic in the upper $q'_0$ half plane and a term analytic in the
1639: lower $q'_0$ half plane by noting that
1640: \begin{eqnarray}
1641: G_0(K/2+q-q')G_0(K/2-q-p+q') \nn \\
1642: = \frac{G_0(K/2+q-q')+G_0(K/2-q-p+q')}
1643: {\frac E2-p_0-(\vec q-\vec q\,')^2/2-(\vec q+\vec p-\vec q\,')^2/2+i0}.
1644: \end{eqnarray}
1645: The second product of fermionic propagators is treated similarly.
1646: The remaining part of the integral equation for $\Xi_k(q,q',p-q')$
1647: consists of a term analytic in the upper $q'_0$ half plane and a term
1648: analytic in the lower $q'_0$ half plane. Thus effectively we have
1649: split $\Xi_k(q,q',p-q')$ into
1650: \begin{equation}
1651: \Xi_k(q,q',p-q') = \Xi^+_k(q,q',p-q') + \Xi^-_k(q,q',p-q')
1652: \end{equation}
1653: with $\Xi^+_k(q,q',p-q')$ ($\Xi^-_k(q,q',p-q')$) analytic in the upper
1654: (lower) $q_0'$ half plane. Note that this splitting means that in
1655: complete generality
1656: \begin{equation}
1657: \Xi_k(q,p_1,p_2) = \Xi^+_k(q,p_1,p_2) + \Xi^-_k(q,p_1,p_2).
1658: \label{xisplit}
1659: \end{equation}
1660: 
1661: $\Xi^+_k(q,q',p-q')$ and $\Xi^-_k(q,q',p-q')$ satisfy a set of coupled
1662: integral equations. The equation satisfied by $\Xi^-_k(q,q',p-q')$ is
1663: \begin{widetext}
1664: \begin{eqnarray}
1665: && \Xi^-_k(q,q',p-q') =
1666: %& = &
1667: \nn \\ &&
1668: -\int\frac{d\Omega_{\vec q}}{4\pi}
1669: \left[\frac{G_0(K/2+q-q')}
1670: {\frac E2-p_0-(\vec q-\vec q\,')^2/2-(\vec q+\vec p-\vec q\,')^2/2+i0}
1671: %\right. \nn \\ && \left.
1672: +\frac{G_0(K/2-q-q')}
1673: {\frac E2-p_0-(\vec q+\vec q\,')^2/2-(\vec q-\vec p+\vec q\,')^2/2+i0}
1674: \right] \nn \\ &&
1675: -\left.\int\frac{d^3Q}{(2\pi)^3}G_0(K/2-Q-p)D_0(K-Q-q')\Xi_k(q,q',Q)\right|_{Q_0=\vec
1676:   Q^2/2-E/4}.
1677: \label{ximinusoff}
1678: \end{eqnarray}
1679: \end{widetext}
1680: The right hand side of this equation is related to $\Xi^+$ through
1681: Eq. (\ref{xisplit}). The crucial point is that the right hand side of
1682: Eq. (\ref{ximinusoff}) equals $\Xi^+_k(q,p-q',q')$ which can be seen
1683: by writing the corresponding equation for $\Xi^+_k(q,q',p-q')$ and
1684: using the change of variables $p-q'\leftrightarrow q'$ along with
1685: Eq. (\ref{eq:symmetry}). Thus
1686: $\Xi^+_k(q,p-q',q')=\Xi^-_k(q,q',p-q')$. Consequently, using
1687: Eq. (\ref{gammaxi}), we conclude that $\Xi^+_k(q,q',p-q')$ and
1688: $\Xi^-_k(q,q',p-q')$ has identical contributions to $\Gamma$. We also
1689: note that
1690: \begin{equation}
1691: \Xi_k(q,p_1,p_2) = \Xi^-_k(q,p_1,p_2) + \Xi^-_k(q,p_2,p_1),
1692: \label{xisplitminus}
1693: \end{equation}
1694: which may be inserted in the right hand side of
1695: Eq. (\ref{ximinusoff}). The result is an un-coupled integral equation
1696: for $\Xi^-_k(q,q',p-q')$.
1697: 
1698: We now perform the $q'_0$-integration in the lower half plane in
1699: Eq. (\ref{gammaxi}) using Eq. (\ref{xisplitminus}). This integration
1700: sets $q'_0=\vec q\,'^2/2-E/4>0$. At this value of the total outgoing
1701: fermion energy Eq. (\ref{ximinusoff}) reduces to
1702: \begin{widetext}
1703: \begin{eqnarray}
1704: && \Xi_k^-(q;\vec q\,',\vec q\,'^2/2-E/4;\vec p-\vec q\,',p_0-\vec
1705: q\,'^2/2+E/4)
1706: \nn \\ &=&
1707: %=
1708: \int\frac{d\Omega_{\vec q}}{4\pi}
1709: \left[
1710: \frac{G_0(K/2+q-q')}
1711: {p_0+(\vec q-\vec q\,')^2/2+(\vec q+\vec p-\vec q\,')^2/2-E/2-i0}+
1712: \frac{G_0(K/2-q-q')}
1713: {p_0+(\vec q+\vec q\,')^2/2+(-\vec q+\vec p-\vec q\,')^2/2-E/2-i0}
1714: \right]
1715: \nn \\
1716: && +\frac1{2\pi^2}\int d^3Q\,\frac
1717: {\Xi^-_k(q;\vec q\,',\vec q\,'^2/2-E/4;\vec Q,\vec Q/2-E/4)
1718: +\Xi^-_k(q;\vec Q,\vec Q/2-E/4;\vec q\,',\vec q\,'^2/2-E/4)}
1719: {\left(p_0+\vec Q^2/2+(\vec p+\vec Q)^2/2-E/2-i0\right)\left(1-\sqrt{-E+3\vec
1720:     Q^2/4+3\vec q\,'^2/4+\vec Q\cdot \vec q\,'/2}\right)}.
1721: \label{ximinus}
1722: \end{eqnarray}
1723: \end{widetext}
1724: The right hand side of this equation only contains $\Xi^-$ evaluated
1725: at fermion energies related to the corresponding momentum by $Q_0=\vec
1726: Q^2-E/4$ and $q'_0=\vec q\,'^2/2-E/4$, which we will call ``on
1727: shell''. We will solve the integral equation by first letting
1728: \begin{equation}
1729: p_0\to \vec p\,^2/2+\vec q\,'^2-\vec p\cdot\vec q\,'-E/2,
1730: \end{equation}
1731: which means that the left hand side is also evaluated ``on shell''.
1732: The resulting integral equation is an integral equation in $|\vec
1733: q'|$, $|\vec p-\vec q\,'|$, and the angle between these, using $|\vec
1734: k|$, $|\vec p|$, $|\vec q|$, and $q_0$ as input. Subsequently the ``on
1735: shell'' $\Xi^-$ can be inserted in the right hand side of
1736: Eq. (\ref{ximinus}) to evaluate $\Xi^-$ at any values of the outgoing
1737: energy.
1738: 
1739: It is now possible to conclude that $\Gamma(q,q_0;p,p_0)$ is analytic
1740: in quadrant $I$ and $III$ in $p_0$ and $q_0$ and that furthermore any
1741: non-analyticity is far from the rotated contour of integration. The
1742: remaining pole of a fermionic propagator in Eq. (\ref{gammaxi}) after
1743: the $q_0'$ integration in the lower half plane is at $p_0=1-k^2/4+\vec
1744: q\,'^2/2+(\vec p-\vec q\,')^2/2-i0$ where the real part is greater
1745: than $1-k^2/4\geq1$. In Eq. (\ref{ximinus}) the poles in $p_0$ are in
1746: quadrant $II$ and also has real part less than
1747: $-1+k^2/4\leq-1$. Finally, also in Eq. (\ref{ximinus}), the pole in
1748: $q_0$ is in quadrant $IV$ below the branch cut.
1749: 
1750: We now define the ``on shell'' function
1751: \begin{eqnarray}
1752: \tilde\Xi_k(q;p_1,p_2,\cos\theta) & \equiv &
1753: \nn\\ &&
1754: \hspace{-2.2cm}
1755: \Xi^-_k(q;\vec p_1,\vec
1756: p_1\,^2/2-E/4; \vec p_2,\vec p_2\,^2-E/4).
1757: \label{eq:defineonshell}
1758: \end{eqnarray}
1759: Here, $p_{1,2} \equiv |\vec p_{1,2}|$ and $\theta$ is the angle
1760: between these vectors. We then find the final set of equations to be
1761: solved to construct $\Gamma$:
1762: \begin{widetext}
1763: \begin{eqnarray}
1764: \tilde\Xi_k(q;p_1,p_2,\cos\theta) & = & -\int\frac{d\Omega_{\vec
1765:     q}}{4\pi}
1766: \left[
1767: \frac1{q^2+p_1^2+p^2_2-\vec q\cdot \vec p_1+\vec q\cdot\vec p_2-E}
1768: \frac1{-q_0+p_1^2/2+(\vec q-\vec p_1)^2/2-E/2} \right.
1769: \nn \\
1770: &&
1771: \hspace{-1.8cm}
1772: \left.+
1773: \frac1{q^2+p_1^2+p^2_2+\vec q\cdot \vec p_1-\vec q\cdot\vec p_2-E}
1774: \frac1{q_0+p_1^2/2+(\vec q+\vec p_1)^2/2-E/2}
1775: \right] \nn \\
1776: &&
1777: \hspace{-1.8cm}
1778: +\frac1\pi\int_{-1}^1d(\cos\phi)\int_0^\infty Q^2dQ\frac{1}
1779: {1-\sqrt{-E+3Q^2/4+3p_1^2/4+Qp_1\cos\phi/2}} \nn \\
1780: &&
1781: \hspace{-1.8cm}
1782: \times\frac{\tilde\Xi_k(q,p_1,Q,\cos\phi)+\tilde\Xi_k(q;Q,p_1,\cos\phi)}
1783: {\sqrt{(Q^2+p_1^2+p_2^2+Qp_1\cos\phi+p_1p_2\cos\theta+Qp_2\cos\theta\cos\phi-E)^2
1784: -Q^2p_2^2(1-\cos^2\theta)(1-\cos^2\phi)}}.
1785: \label{onshellxi}
1786: \end{eqnarray}
1787: We now insert Eq. (\ref{onshellxi}) in the right hand side of
1788: Eq. (\ref{ximinus}) using Eq. (\ref{eq:defineonshell}), and
1789: subsequently Eq. (\ref{ximinus}) in Eq. (\ref{gammaxi}). Performing
1790: the integral over the azimuthal angle in the $\vec Q$ integral we find
1791: \begin{eqnarray}
1792: \Gamma_k(q,q_0;p,p_0) & = & \Gamma_k^{(0)}(q,q_0;p,p_0)+
1793: \frac1{4\pi^3}\int_{-1}^1 d(\cos\theta)\int_0^\infty
1794: \frac{q'^2dq'}{p_0-q'^2-p^2/2+p q'
1795: \cos\theta+E/2} \nn \\ && \times \int_{-1}^1d(\cos\phi) \int_0^\infty
1796: \frac{Q^2dQ}{1-\sqrt{-E+3Q^2/4+3q'^2/4+Q q'\cos\phi/2}}
1797: \nn \\ && \times
1798: \frac{\tilde\Xi_k(q;q',Q,\cos\phi)+
1799: \tilde\Xi_k(q;Q,q',\cos\phi)}{
1800: \sqrt{(Q^2+p^2/2+Qp\cos\theta\cos\phi+p_0-E/2)^2-Q^2p^2
1801:     (1-\cos^2\theta)(1-\cos^2\phi)}},
1802: \label{eq:finalgamma}
1803: \end{eqnarray}
1804: where $\Gamma^{(0)}$ is the contribution from the first diagram on the
1805: right hand side in Fig. \ref{fig:diagrams}, which can be found to be
1806: \begin{eqnarray}
1807: \Gamma_k^{(0)}(q,q_0;p,p_0) & = & \frac1{16\pi^2}\int_0^\infty
1808: \frac{dq'}{pq}\frac1{-E/2+q'^2+p^2/4+q^2/4}
1809: \ln\frac{-q_0^2+(-E/2+q'^2+p^2/4+q^2/4-q'p)^2}{-q_0^2+(-E/2+q'^2+p^2/4+q^2/4+q'p)^2}
1810: \nn \\ &&
1811: \times\ln\frac{-p_0^2+(-E/2+q'^2+p^2/4+q^2/4-q'q)^2}{-p_0^2+(-E/2+q'^2+p^2/4+q^2/4+q'q)^2}.
1812: \label{gamma0}
1813: \end{eqnarray}
1814: \end{widetext}
1815: With the energies rotated to the imaginary axis all equations used to
1816: find the $t$-matrix, Eq. (\ref{inteqgammas}) at $dq_0\to i\,dq_0$, and
1817: Eqs. (\ref{onshellxi}-\ref{gamma0}) are real. This is not immediately
1818: obvious but may be easily checked numerically.
1819: 
1820: To treat the poles in Eq. (\ref{inteqgammas}) of the
1821: bosonic propagators at $q_0=\pm(q^2/4-k^2/4-i0)$ as $k,q\to0$ we
1822: perform the following change of variables on the external energies and
1823: momenta to ``polar'' coordinates
1824: \begin{eqnarray}
1825: q^2/4 & = & R^2_1\cos\theta_1, \hspace{1cm} q_0 = R_1^2\sin\theta_1,
1826: \nn \\
1827: p^2/4 & = & R^2_2\cos\theta_2, \hspace{1cm} p_0 = R_2^2\sin\theta_2,
1828: \label{eq:change}
1829: \end{eqnarray}
1830: where $R_i\in[0,\infty[$ and $\theta_i\in[0,\pi/2]$. Here we restrict
1831: the integration over $q_0$ to the upper imaginary axis, using the
1832: symmetry of the integrand in Eq. (\ref{inteqgammas}) as
1833: $q_0\to-q_0$. This symmetry follows from the symmetry of the
1834: problem. For convergence reasons it is advantageous to integrate over
1835: finite intervals, and to this end we use the change of variables
1836: \begin{equation}
1837: R_i = \frac2{z_i+1}-1.
1838: \end{equation}
1839: We employ the same change of variables on the internal loop momenta
1840: $Q$ and $q'$ in Eqs. (\ref{onshellxi}) and (\ref{eq:finalgamma}),
1841: \begin{equation}
1842: Q = \frac2{z_Q+1}-1, \,\,\, q' = \frac2{z'+1}-1.
1843: \end{equation}
1844: To solve the integral equations (\ref{inteqgammas}) and
1845: (\ref{onshellxi}) we use the Nystrom method, writing the integral
1846: equations as matrix equations and inverting these to find the
1847: solution. Evaluation of the integrals in these integral equations is
1848: performed using Gauss-Legendre quadrature \cite{numrecipes}.
1849: 
1850: 
1851: %The high frequency asymptotics can be found if we note that in
1852: %this regime $g(p)$ can effectively be replaced by $M (2 \pi)^3
1853: %\delta(p)$, where $M$ sets the characteristic scale of the
1854: %correlation function $g(p)$. In other words, all the length scales
1855: %become much smaller than $l$, the disorder correlation length.
1856: %Then $Q_{1,2} = M \lambda_{1,2}$ and we find
1857: %\begin{equation}
1858: %\lambda_1 = - {p^2 \over 2 M} - {i \over 2} \sqrt{  4 {p^2 \over
1859: %M} \omega^2 - {p^4 \over M^2 }}.
1860: %\end{equation}
1861: %Therefore, the density of states is given by
1862: %\begin{equation}
1863: %\rho(\omega) = {1 \over \pi \omega^3} \int_0^{2 \omega \sqrt{M}}
1864: %{p^2 dp \over 2 \pi^2} \sqrt{  4 {p^2 \over M} \omega^2 - {p^4
1865: %\over M^2 }}
1866: %\end{equation}
1867: 
1868: 
1869: 
1870: 
1871: 
1872: 
1873: 
1874: % Recently low frequency bosonic excitations in random media were
1875: %discussed in a very general framework in
1876: %Ref.~\cite{Gurarie2002,Gurarie2003a}. The discussion in these
1877: %publications showed that the behavior of those excitations depend
1878: %crucially on whether or not they are Goldstone modes. In turn, the
1879: %Goldstone modes can be further subdivided into the phonon-like and
1880: %magnon-like modes, depending on the structure of their
1881: %Hamiltonians.
1882: 
1883: %The phonon-like modes in random elastic media were first
1884: %thoroughly discussed two decades ago in Ref.~\cite{John1982}. One
1885: %of the basic results of that publication was that in an elastic
1886: %media with short range correlated random elastic parameters the
1887: %low frequency density of states goes as
1888: %\begin{equation}
1889: %\label{eq:DoS} \rho(\omega) \propto \omega^{d-1}, \, \omega \ll
1890: %\omega_{D},
1891: %\end{equation}
1892: %where $d$ is the dimensionality of the media, and $\omega_D$ is
1893: %the Debye frequency. This result can be anticipated without any
1894: %calculations. The long wavelength nature of low frequency phonons
1895: %(and indeed of any Goldstone modes) tend to average local disorder
1896: %over distances of the order of wavelength and therefore the
1897: %fluctuations in the elastic properties of the media can be
1898: %completely neglected. On the other hand, it is well known that the
1899: %density of states in any clean elastic media follows \rfs{eq:DoS}.
1900: 
1901: 
1902: 
1903: 
1904: 
1905: 
1906: 
1907: 
1908: 
1909: 
1910: 
1911: 
1912: 
1913: % References should be done using the \cite, \ref, and \label commands
1914: %\section{}
1915: % Put \label in argument of \section for cross-referencing
1916: %\section{\label{}}
1917: %\subsection{}
1918: %\subsubsection{}
1919: 
1920: % If in two-column mode, this environment will change to single-column
1921: % format so that long equations can be displayed. Use
1922: % sparingly.
1923: %\begin{widetext}
1924: % put long equation here
1925: %\end{widetext}
1926: 
1927: % figures should be put into the text as floats.
1928: % Use the graphics or graphicx packages (distributed with LaTeX2e)
1929: % and the \includegraphics macro defined in those packages.
1930: % See the LaTeX Graphics Companion by Michel Goosens, Sebastian Rahtz,
1931: % and Frank Mittelbach for instance.
1932: %
1933: % Here is an example of the general form of a figure:
1934: % Fill in the caption in the braces of the \caption{} command. Put the label
1935: % that you will use with \ref{} command in the braces of the \label{} command.
1936: % Use the figure* environment if the figure should span across the
1937: % entire page. There is no need to do explicit centering.
1938: 
1939: % \begin{figure}
1940: % \includegraphics{}%
1941: % \caption{\label{}}
1942: % \end{figure}
1943: 
1944: % Surround figure environment with turnpage environment for landscape
1945: % figure
1946: % \begin{turnpage}
1947: % \begin{figure}
1948: % \includegraphics{}%
1949: % \caption{\label{}}
1950: % \end{figure}
1951: % \end{turnpage}
1952: 
1953: % tables should appear as floats within the text
1954: %
1955: % Here is an example of the general form of a table:
1956: % Fill in the caption in the braces of the \caption{} command. Put the label
1957: % that you will use with \ref{} command in the braces of the \label{} command.
1958: % Insert the column specifiers (l, r, c, d, etc.) in the empty braces of the
1959: % \begin{tabular}{} command.
1960: % The ruledtabular enviroment adds doubled rules to table and sets a
1961: % reasonable default table settings.
1962: % Use the table* environment to get a full-width table in two-column
1963: % Add \usepackage{longtable} and the longtable (or longtable*}
1964: % environment for nicely formatted long tables. Or use the the [H]
1965: % placement option to break a long table (with less control than
1966: % in longtable).
1967: % \begin{table}%[H] add [H] placement to break table across pages
1968: % \caption{\label{}}
1969: % \begin{ruledtabular}
1970: % \begin{tabular}{}
1971: % Lines of table here ending with \\
1972: % \end{tabular}
1973: % \end{ruledtabular}
1974: % \end{table}
1975: 
1976: % Surround table environment with turnpage environment for landscape
1977: % table
1978: % \begin{turnpage}
1979: % \begin{table}
1980: % \caption{\label{}}
1981: % \begin{ruledtabular}
1982: % \begin{tabular}{}
1983: % \end{tabular}
1984: % \end{ruledtabular}
1985: % \end{table}
1986: % \end{turnpage}
1987: 
1988: % Specify following sections are appendices. Use \appendix* if there
1989: % only one appendix.
1990: %\appendix
1991: %\section{}
1992: 
1993: % If you have acknowledgments, this puts in the proper section head.
1994: %\begin{acknowledgments}
1995: % put your acknowledgments here.
1996: %\end{acknowledgments}
1997: 
1998: % Create the reference section using BibTeX:
1999: \bibliography{bec}
2000: 
2001: \end{document}
2002: %
2003: % ****** End of file template.aps ******
2004: