cond-mat0701407/arb.tex
1: \documentstyle[12pt,epsfig]{article}
2: 
3: \topmargin -7mm
4: \evensidemargin 5mm
5: \oddsidemargin 5mm
6: \textwidth=16.5cm
7: \textheight=23.5cm
8: 
9: \newcommand{\be}{\begin{equation}}
10: \newcommand{\ee}{\end{equation}}
11: \newcommand{\Dlt}{\Delta}
12: \newcommand{\dlt}{\delta}
13: \newcommand{\prt}{\partial}
14: \newcommand{\br}{{\bf r}}
15: \newcommand{\bk}{{\bf k}}
16: \newcommand{\bt}{\beta}
17: \newcommand{\vp}{\varphi}
18: \newcommand{\ep}{\varepsilon}
19: \newcommand{\al}{\alpha}
20: \newcommand{\ra}{\rightarrow}
21: \newcommand{\sgm}{\sigma}
22: \newcommand{\gm}{\gamma}
23: \newcommand{\om}{\omega}
24: \newcommand{\Om}{\Omega}
25: \newcommand{\Gm}{\Gamma}
26: \newcommand{\dgr}{\dagger}
27: \newcommand{\lbd}{\lambda}
28: \newcommand{\Lbd}{\Lambda}
29: \newcommand{\cF}{{\cal F}}
30: 
31: \begin{document}
32: 
33: \begin{center}
34: 
35: {\Large{\bf Bose-Einstein-condensed gases with arbitrary strong interactions} \\ [5mm]
36: 
37: V.I. Yukalov$^1$ and E.P. Yukalova$^2$} \\ [3mm]
38: 
39: {\it $^1$Bogolubov Laboratory of Theoretical Physics, \\
40: Joint Institute for Nuclear Research, Dubna 141980, Russia \\ [3mm]
41: $^2$Department of Computational Physics,
42: Laboratory of Information Technologies, \\
43: Joint Institute for Nuclear Research, Dubna 141980, Russia}
44: 
45: \end{center}
46: 
47: \begin{abstract}
48: 
49: Bose-condensed gases are considered with an effective interaction
50: strength varying in the whole range of the values between zero and infinity.
51: The consideration is based on the usage of a representative statistical
52: ensemble for Bose systems with broken global gauge symmetry. Practical
53: calculations are illustrated for a uniform Bose gas at zero temperature,
54: employing a self-consistent mean-field theory, which is both conserving and
55: gapless.
56: 
57: \end{abstract}
58: 
59: \vskip 1cm
60: 
61: {\bf PACS}: 03.75.Hh, 03.75.Kk, 03.75.Nt, 05.30.Ch
62: 
63: \newpage
64: 
65: \section{Introduction}
66: 
67: The properties of systems with Bose-Einstein condensate are currently a topic
68: of great interest, both experimentally and theoretically (see review articles
69: [1--7]). One usually considers weakly interacting Bose gases, whose theory was
70: pioneered by Bogolubov [8,9]. Binary atomic interactions in such gases can be
71: modelled by contact potentials expressed through effective scattering lengths.
72: But the latter can also be made rather large by means of the Feshbach resonance
73: technique, so that effective atomic interactions could become quite strong
74: [7,10,11]. Extension of the Bogolubov theory to Bose systems with strong
75: interactions confronts the well known problem of conserving versus gapless
76: approximations, as was formulated by Hohenberg and Martin [12]. This dilemma
77: has recently been discussed in detail in the review paper by Andersen [5].
78: 
79: The Hohenberg-Martin dilemma of conserving versus gapless theories can be
80: resolved by employing representative statistical ensembles [13]. Using such
81: an ensemble for Bose systems with broken global gauge symmetry makes it
82: straightforward to get a self-consistent theory, both conserving as well as
83: gapless in any given approximation. In particular, the Hartree-Fock-Bogolubov
84: (HFB) approximation, which is by construction conserving, can also be made
85: gapless [14].
86: 
87: In the present paper, we consider an equilibrium Bose system with Bose-Einstein
88: condensate. The main new results are twofold. First, we give a general
89: mathematical foundation for the construction of the grand Hamiltonian for
90: an arbitrary equilibrium system with broken gauge symmetry. The derivation
91: of the grand Hamiltonian and the corresponding equations of motion are valid
92: for any Bose system, whether uniform or nonuniform. Second, in the frame of
93: a self-consistent mean-field theory for a uniform Bose gas, we study, both
94: analytically and numerically, the zero-temperature characteristics as functions
95: of the gas parameter, varying the latter from zero to infinity. Specifically,
96: the condensate fraction, sound velocity, normal and anomalous averages, and
97: the ground-state energy as functions of the gas parameter are investigated.
98: The results are in good agreement with available computer Monter Carlo
99: simulations.
100: 
101: We use the system of units, where $\hbar\equiv 1$ and $k_B\equiv 1$.
102: 
103: \section{Representative Ensemble for Bose-Condensed Systems}
104: 
105: The description of a spinless Bose system at temperature $T>T_c$ above the
106: condensation temperature $T_c$ can be done in terms of the field operators
107: $\psi(\br,t)$ and $\psi^\dgr(\br,t)$ dependeing on the spatial vector $\br$
108: and time $t$. The operators from the algebra of observables and other
109: physical operators are defined in the Fock space $\cF(\psi)$ generated by the
110: field operator $\psi^\dgr$. The related mathematical details of constructing
111: the Fock space $\cF(\psi)$ can be found in books [15,16]. Under the total
112: number of particles $N=<\hat N>$, being the average of the number-of-particle
113: operator $\hat N$, the grand Hamiltonian has the standard form
114: $$
115: H[\psi] = \hat H [\psi] - \mu\hat N \qquad (T > T_c) \; ,
116: $$
117: where $\hat H[\psi]$ is the Hamiltonian energy, which is invariant under the
118: global gauge transformations from the $U(1)$ symmetry group.
119: 
120: At temperatures $T<T_c$, the global gauge symmetry becomes broken. This is
121: achieved by means of the Bogolubov shift [17,18] for the field operators
122: \be
123: \label{1}
124: \psi(\br,t) \; \longrightarrow \; \hat\psi(\br,t)\equiv \eta(\br,t)
125: + \psi_1(\br,t) \; ,
126: \ee
127: in which $\eta(\br,t)$ is the condensate wave function and $\psi_1(\br,t)$
128: is the field operator of uncondensed atoms, enjoying the same Bose
129: commutation relations as $\psi$. The condensate wave function $\eta(\br,t)$
130: is the system order parameter. Now all operators of physical quantities are
131: defined on the Fock space $\cF(\psi_1)$ generated by the field operator
132: $\psi_1^\dgr$. It is important to emphasize that the Fock spaces $\cF(\psi)$
133: and $\cF(\psi_1)$ are mutually orthogonal [13,19].
134: 
135: Thus, below $T_c$, instead of one operator variable $\psi$, there appear two
136: variables, $\eta$ and $\psi_1$. These are linearly independent, being orthogonal
137: to each other,
138: \be
139: \label{2}
140: \int \eta^*(\br,t) \psi_1(\br,t) \; d\br = 0 \; .
141: \ee
142: For two linearly independent variables, there are two normalization conditions.
143: One is the normalization of the condensate function to the number of
144: condensed atoms
145: \be
146: \label{3}
147: N_0 = \int |\eta(\br,t)|^2 \; d\br \; .
148: \ee
149: And another normalization condition is for the operator
150: \be
151: \label{4}
152: \hat N_1 \equiv \int \psi_1^\dgr(\br,t) \psi_1(\br,t)\; d\br \; ,
153: \ee
154: whose average yields the number of uncondensed atoms
155: \be
156: \label{5}
157: N_1 \;  = \; <\hat N_1> \; .
158: \ee
159: The normalization condition (3) can be represented in the same form of the
160: statistical average (5) by using the operator
161: $$
162: \hat N_0 \equiv N_0 \hat 1_\cF \; ,
163: $$
164: where $\hat 1_\cF$ is the unity operator in $\cF(\psi_1)$. Then Eq. (3) is
165: equivalent to the normalization
166: \be
167: \label{6}
168: N_0 \; = \; <\hat N_0> \; .
169: \ee
170: The statistical average of an operator $\hat A$ is defined in the standard
171: way as
172: $$
173: <\hat A> \; \equiv \; {\rm Tr}\; \hat\rho\; \hat A \; ,
174: $$
175: where $\hat\rho$ is a statistical operator and the trace is over $\cF(\psi_1)$.
176: 
177: One more restriction is
178: \be
179: \label{7}
180: <\psi_1(\br,t)> \; = \; 0 \; ,
181: \ee
182: which guarantees the conservation of quantum numbers. This can also be
183: rewritten as the quantum conservation condition
184: \be
185: \label{8}
186: <\hat\Lbd> \; = \; 0
187: \ee
188: for the self-adjoint operator
189: \be
190: \label{9}
191: \hat\Lbd \equiv \int \left [ \lbd(\br,t)\psi_1^\dgr(\br,t) +
192: \lbd^*(\br,t)\psi_1(\br,t) \right ]\; d\br \; ,
193: \ee
194: in which $\lbd(\br,t)$ is a complex function.
195: 
196: Two other common conditions is the normalization of the statistical operator
197: $\hat\rho$,
198: \be
199: \label{10}
200: <\hat 1_\cF> \; = \; 1 \; ,
201: \ee
202: and the definition of the internal energy
203: \be
204: \label{11}
205: E\; = \; <\hat H>
206: \ee
207: as the average of the Hamiltonian energy operator
208: $\hat H\equiv\hat H[\hat\psi]$, which is a functional of the shifted field
209: operator (1).
210: 
211: An equilibrium statistical ensemble for a Bose-condensed system is the pair
212: $\{\cF(\psi_1),\hat\rho\}$ of the space of microstates $\cF(\psi_1)$ and a
213: statistical operator $\hat\rho$. The notion of a representative ensemble stems
214: from the works of Gibbs [20], who emphasized that for the correct description
215: of the given statistical system, in addition to the standard conditions (10)
216: and (11), it is necessary to take into account all other constraints that
217: uniquely define the considered system. The corresponding statistical operator
218: can be found from the maximization of the Gibbs entropy $S\equiv -{\rm Tr}
219: \hat\rho\ln\hat\rho$ under the given statistical conditions. The conditional
220: maximization of the entropy is equivalent to the unconditional minimization
221: of the information functional [16,21]. In the present case, in addition to
222: conditions (10) and (11), we must also take into account the normalization
223: conditions (5) and (6) and the conservation constraint (8). Hence the
224: information functional is
225: $$
226: I[\hat\rho] = {\rm Tr}\; \hat\rho\; \ln\hat\rho + \lbd_0({\rm Tr}\;\hat\rho
227: -1 ) +
228: $$
229: \be
230: \label{12}
231: + \bt ( {\rm Tr}\;\hat\rho \; \hat H -  E ) -
232: \bt\mu_0 ( {\rm Tr}\;\hat\rho \; \hat N_0 - N_0 ) -
233: \bt \mu_1 ( {\rm Tr}\;\hat\rho \; \hat N_1  - N_1 ) -
234: \bt {\rm Tr}\;\hat\rho\; \hat\Lbd \; ,
235: \ee
236: in which $\lbd_0$, $\bt$, $\bt\mu_0$, $\bt\mu_1$, and $\bt\lbd$ are the
237: appropriate Lagrange multipliers guaranteeng the validity of conditions (5),
238: (6), (8), (10), and (11). Minimizing functional (12), we get the statistical
239: operator
240: \be
241: \label{13}
242: \hat\rho = \frac{1}{Z}\; e^{-\bt H} \; ,
243: \ee
244: with the inverse temperature $\bt\equiv 1/T$, the partition function
245: $$
246: Z = \exp(1+\lbd_0) = {\rm Tr}\; e^{-\bt H} \; ,
247: $$
248: and the grand Hamiltonian
249: \be
250: \label{14}
251: H \equiv \hat H - \mu_0 \hat N_0 - \mu_1 \hat N_1 - \hat \Lbd \; .
252: \ee
253: 
254: Let us explain in more detail the role played by the Lagrange multipliers
255: $\mu_0$ and $\mu_1$. Above the Bose-Einstein condensation point, when
256: $N_0=0$ and $\hat N_1=\hat N$, one needs just one Lagrange multiplier
257: $\mu_1=\mu$, coinciding with the system chemical potential, whose role
258: is to preserve the total average number of particles $N=<\hat N>$. However,
259: below the condensation point, when the global gauge symmetry is broken,
260: there are two types of particles, condensed and uncondensed ones. The
261: number of condensed particles $N_0$, according to the Bogolubov theory
262: [8,9,17,18], has to be such that to make the system stable by minimizing
263: the thermodynamic potential. For an equilibrium statistical system with
264: the statistical operator (13), the grand thermodynamic potential is
265: \be
266: \label{15}
267: \Om = - T\ln {\rm Tr} e^{-\bt H} \; ,
268: \ee
269: with the grand Hamiltonian (14). Extremizing the grand potential (15) with
270: respect to the number of condensed particles, from the equation
271: $$
272: \frac{\prt\Om}{\prt N_0}= 0
273: $$
274: one obtains
275: $$
276: \mu_0= \; < \frac{\prt\hat H}{\prt N_0} >\; .
277: $$
278: This means that the Lagrange multiplier $\mu_0$ is responsible for the
279: thermodynamic stability of the system. Another Lagrange multiplier, $\mu_1$,
280: guarantees the normalization condition (5) for the number of uncondensed
281: particles $N_1$. But the latter, since $N_1=N-N_0$, implies that $\mu_1$
282: preserves the total average number of particles $N$. In this way, for a
283: Bose-condensed system, contrary to the system without the Bose-Einstein
284: condensate, there are two conditions on the number of particles. One
285: condition, as earlier, is that the total number of particles be $N$. And
286: another condition is that the number of condensed particles, $N_0$, would
287: be such that to provide the stability of the system by minimizing the
288: thermodynamic potential. This is why one needs two Lagrange multipliers
289: in order to guarantee the validity of these two conditions at each step of
290: any calculational procedure. As is shown below by practical calculations,
291: the use of two Lagrange multipliers makes the theory self-consistent,
292: avoiding the Hohenberg-Martin dilemma and yielding the results that are
293: in agreement with those derived analytically for the weak-coupling limit
294: as well as obtained by Monte-Carlo computer simulations for strong
295: interactions.
296: 
297: It is also important to keep in mind that the introduction of Lagrange
298: multipliers is a technical method allowing us to simplify calculations.
299: The number of the introduced multipliers is connected with the concrete
300: properties of the employed approach. Thus, in the Bogolubov theory
301: [8,9,17,18] one deals with two independent field variables, the condensate
302: wave function $\eta(\br,t)$ and the operator of uncondensed particles
303: $\psi_1(\br,t)$. This is why, as is explained above, it is convenient to
304: introduce two Lagrange multipliers.
305: 
306: One could ask whether we could limit ourselves by introducing a sole
307: Lagrange multiplier. The answer is straightforward: Yes, we could, but
308: the calculational procedure should then be changed. For instance, we could
309: follow the way of Hugenholtz and Pines [22], which was also used by
310: Gavoret and Nozieres [23]. They consider a uniform equilibrium system at
311: zero temperature, defining the number of condensed particles
312: $N_0=N_0(\rho,T)$ as a function of density $\rho$ and temperature $T$ from
313: the extremization of the internal energy $E\equiv<\hat H>$. The found
314: $N_0(\rho,T)$ is substituted explicitly into the Hamiltonian $\hat H$,
315: after which one works with the grand Hamiltonian $H=\hat H-\mu\hat N_1$,
316: where $\hat N_1$ is the operator for the number of  {\it uncondensed}
317: particles. Since the number of condensed atoms has already been defined
318: earlier from the stability condition, one requires now to use the sole
319: Lagrange multiplier aiming at guaranteeing the normalization condition
320: $N_1=<\hat N_1>$ for uncondensed particles, hence, because of the fixed
321: relation $N_1=N-N_0$, preserving the total number of particles $N$. This
322: way of calculations, with details expounded in Refs. [22,23], is
323: mathematically equivalent to the procedure, when $N_0$ has not been fixed
324: in advance but, instead, a Lagrange multiplier $\mu_0$ is introduced to
325: guarantee the stability condition in the process of calculations, after
326: which the number of condensed particles is defined as $N_0=N-N_1$, which
327: specifies the condensate depletion. It is this method of using an additional
328: Lagrange multiplier that is employed in the present paper.
329: 
330: Even more, it could be possible to work introducing no any Lagrange
331: multipliers at all, but which again would necessitate a different
332: calculational procedure. Thus, one could use the Girardeau-Arnowitt
333: approach [24,25] in the frame of the canonical ensemble, which requires
334: to invoke the so-called number-conserving field operators. The weak
335: point of this approach is that, as is well known [24,25], it yields the
336: unphysical gap in the spectrum, when approximate calculations are involved.
337: Girardeau mentioned [26] that an exact theory should not have the gap.
338: Later, Takano [27] demonstrated that, really, the gap should not arise when
339: all terms of the Hamiltonian are taken into account. Unfortunately, there
340: are no exact solutions for a realistic interacting Bose system, so that one
341: always has to resort to some approximations, in the course of which the
342: gap again reappears. This is why it is more convenient to work with the
343: grand canonical ensemble, introducing the Lagrange multipliers that would
344: assure the self-consistency of any calculational scheme.
345: 
346: Thus, the method of Lagrange multipliers has to be treated as a technical
347: procedure allowing us to simplify calculations. The number of the
348: required multipliers is intimately related to the concrete calculational
349: details. In the case of the Bogolubov approach [8,9,17,18], introducing
350: two different field variables, the condensate function $\eta$ and the
351: field operator of uncondensed particles $\psi_1$, it is convenient to
352: define two Lagrange multipliers that guarantee the validity of two
353: conditions, the thermodynamic stability condition and the conservation
354: of the total average number of particles.
355: 
356: When two Lagrange multipliers ate involved, none of them plays the role
357: of the system chemical potential. To define the latter, we may proceed as
358: follows. Keeping in mind that in experiments the total number of particles
359: is usually fixed, we may write for the internal energy (11) the standard
360: relation
361: \be
362: \label{16}
363: E = \; <H> + \;\mu N \; ,
364: \ee
365: connecting $E$ with the average of the grand Hamiltonian $<H>$ and with
366: the system chemical potential $\mu$. At the same time, substituting into
367: Eq. (11) the grand Hamiltonian (14), and taking into account condition
368: (8), we come to the expression
369: \be
370: \label{17}
371: E = \; <H> +\mu_0 N_0 + \mu_1 N_1 \; .
372: \ee
373: 
374: Comparing Eqs. (16) and (17) gives the definition of the system chemical
375: potential
376: \be
377: \label{18}
378: \mu \equiv \mu_0 n_0 + \mu_1 n_1
379: \ee
380: expressed through the Lagrange multipliers $\mu_0$ and $\mu_1$ and the
381: related atomic fractions
382: $$
383: n_0 \equiv \frac{N_0}{N} \; , \qquad n_1 \equiv \frac{N_1}{N} \; .
384: $$
385: 
386: The equations of motion for the variables $\eta$ and $\psi_1$ are given in
387: the usual manner as
388: \be
389: \label{19}
390: i\; \frac{\prt}{\prt t} \; \eta(\br,t) = \frac{\dlt H}{\dlt\eta^*(\br,t)}
391: \ee
392: for the condensate function and, respectively, as
393: \be
394: \label{20}
395: i\; \frac{\prt}{\prt t}\; \psi_1(\br,t) =
396: \frac{\dlt H}{\dlt\psi_1^\dgr(\br,t)}
397: \ee
398: for the field operator of uncondensed atoms.
399: 
400: In this way, the representative statistical ensemble for an arbitrary
401: equilibrium Bose system with broken gauge symmetry is the pair
402: $\{\cF(\psi_1),\hat\rho\}$ of the Fock space of microstates $\cF(\psi_1)$
403: and the statistical operator (13) with the grand Hamiltonian (14). The
404: notion of representative statistical ensembles can be extended to
405: nonequilibrium systems by considering the extremization of an effective
406: action functional [28].
407: 
408: \section{Bose Gas with Contact Interactions}
409: 
410: To specify the consideration, let us take the Hamiltonian energy operator in
411: the usual form
412: \be
413: \label{21}
414: \hat H = \int \hat\psi^\dgr(\br) \left ( -\;
415: \frac{\nabla^2}{2m}\right ) \hat\psi(\br)\; d\br +
416: \frac{1}{2} \; \Phi_0 \int \hat\psi^\dgr(\br) \hat\psi^\dgr(\br)
417: \hat\psi(\br)\hat\psi(\br)\; d\br \; ,
418: \ee
419: corresponding to the contact interaction potential with the strength
420: \be
421: \label{22}
422: \Phi_0 \equiv 4\pi\; \frac{a_s}{m} \; ,
423: \ee
424: where $a_s$ is the scattering length and $m$, mass. Here
425: $\hat\psi(\br)=\hat\psi(\br,t)$ is the shifted field operator (1).
426: 
427: The evolution equation for the condensate function is obtained by averaging
428: Eq. (19). To this end, we need the notation for the local condensate density
429: \be
430: \label{23}
431: \rho_0(\br,t) \equiv |\eta(\br,t)|^2 \; ,
432: \ee
433: normal density of uncondensed atoms
434: \be
435: \label{24}
436: \rho_1(\br,t) \; \equiv \; <\psi_1^\dgr(\br,t)\psi_1(\br,t)> \; ,
437: \ee
438: anomalous density
439: \be
440: \label{25}
441: \sgm_1(\br,t) \; \equiv \; <\psi_1(\br,t)\psi_1(\br,t)> \; ,
442: \ee
443: and the triple correlator
444: \be
445: \label{26}
446: \xi(\br,t) \; \equiv \;
447: <\psi_1^\dgr(\br,t)\psi_1(\br,t)\psi_1(\br,t)> \; .
448: \ee
449: The total local density is
450: \be
451: \label{27}
452: \rho(\br,t) = \rho_0(\br,t) + \rho_1(\br,t) \; .
453: \ee
454: Then, averaging Eq. (19) yields the evolution equation for the condensate
455: function
456: $$
457: i\; \frac{\prt}{\prt t}\; \eta(\br,t)  =\left ( -\;
458: \frac{\nabla^2}{2m} \; - \; \mu_0 \right ) \eta(\br,t) +
459: $$
460: \be
461: \label{28}
462: + \Phi_0 \left [ \rho(\br,t) \eta(\br,t) + \rho_1(\br,t)\eta(\br,t) +
463: \sgm_1(\br,t)\eta^*(\br,t) + \xi(\br,t) \right ] \; .
464: \ee
465: 
466: The equation of motion for the field operator of uncondensed atoms follows
467: from Eq. (20) giving
468: $$
469: i\; \frac{\prt}{\prt t}\; \psi_1(\br,t) = \left ( -\; \frac{\nabla^2}{2m} \;
470: - \; \mu_1\right ) \psi_1(\br,t) +
471: $$
472: \be
473: \label{29}
474: + \Phi_0 \left [ 2\rho_0(\br,t)\psi_1(\br,t) +
475: \eta^2(\br,t)\psi_1^\dgr(\br,t) + \hat X(\br,t) \right ] \; ,
476: \ee
477: where the last term is the correlation operator
478: $$
479: \hat X(\br,t) \equiv \left [ 2\psi_1^\dgr(\br,t) \eta(\br,t) +
480: \eta^*(\br,t)\psi_1(\br,t) + \psi_1^\dgr(\br,t)\psi_1(\br,t)
481: \right ]\psi_1(\br,t) \; .
482: $$
483: 
484: In equilibrium state, one has
485: \be
486: \label{30}
487: \frac{\prt}{\prt t}\; \eta(\br,t) = 0 \; .
488: \ee
489: Also, if the system is uniform, then
490: $$
491: |\eta(\br,t)|^2 = \frac{N_0}{V} \equiv \rho_0 \; ,
492: \qquad \rho_1(\br,t) = \frac{N_1}{V}  =\rho_1 \; ,
493: $$
494: \be
495: \label{31}
496: \sgm_1(\br,t)\equiv \sgm_1 \; , \qquad \xi(\br,t)\equiv\xi \; ,
497: \qquad \rho \equiv \frac{N}{V} = \rho_0 + \rho_1 \; .
498: \ee
499: In that case, equation (28) for the condensate function gives the Lagrange
500: multiplier
501: \be
502: \label{32}
503: \mu_0 =\left ( \rho + \rho_1 +\sgm_1 +
504: \frac{\xi}{\sqrt{\rho_0}} \right ) \Phi_0 \; .
505: \ee
506: 
507: The field operator of uncondensed atoms can be expanded over a complete
508: basis. In general, if the system were nonuniform, in the presence of an
509: external potential, it would be convenient to take a basis formed by natural
510: orbitals [29,30]. For the uniform system under consideration, the natural
511: orbitals are just plane waves $\vp_k(\br)=e^{i\bk\cdot\br}/\sqrt{V}$. The
512: corresponding expansion of $\psi_1$ reads as
513: $$
514: \psi_1(\br,t) = \sum_{k\neq 0} a_k(t) \vp_k(\br) \; .
515: $$
516: With this expansion, the grand Hamiltonian (14) takes the form of a sum
517: \be
518: \label{33}
519: H = \sum_{n=0}^4 H^{(n)}
520: \ee
521: of five terms, classified according to the number of the operators $a_k$
522: or $a_k^\dgr$ in the products. The zero-order term does not contain $a_k$,
523: \be
524: \label{34}
525: H^{(0)} = \left ( \frac{1}{2}\; \rho_0 \Phi_0 - \mu_0 \right ) N_0 \; .
526: \ee
527: 
528: Generally, in order to satisfy condition (8), it is necessary and
529: sufficient [31] that the Hamiltonian (14) would not contain the terms
530: linear in $\psi_1$ or $a_k$. This can be achieved by choosing the
531: corresponding Lagrange multipliers $\lbd(\br,t)$ in Eq. (9). For a
532: uniform system, because of the orthogonality condition (2), one has
533: $\hat\Lbd=0$ and $H^{(1)}=0$ automatically.
534: 
535: The second-order term is
536: \be
537: \label{35}
538: H^{(2)} = \sum_{k\neq 0} \left [ \left ( \frac{k^2}{2m} +
539: 2\rho_0\Phi_0 - \mu_1\right ) a_k^\dgr a_k + \frac{1}{2}\;
540: \rho_0 \Phi_0 \left ( a_k^\dgr a_{-k}^\dgr + a_{-k}a_k
541: \right ) \right ] \; .
542: \ee
543: In the third-order term
544: \be
545: \label{36}
546: H^{(3)} = \sqrt{\frac{\rho_0}{V}} \; \Phi_0 \;
547: {\sum_{p,q}}' \left ( a_q^\dgr a_{q-p} a_p +
548: a_p^\dgr a_{q-p}^\dgr a_q \right ) \; ,
549: \ee
550: the prime on the summation symbol implies that ${\bf p}\neq 0$,
551: ${\bf q}\neq 0$, and ${\bf p}-{\bf q}\neq 0$. The fourth-order term is
552: \be
553: \label{37}
554: H^{(4)} = \frac{\Phi_0}{2V} \; \sum_k {\sum_{p,q}}'
555: a_p^\dgr a_q^\dgr a_{k+p} a_{q-k} \; ,
556: \ee
557: where the prime means that ${\bf p}\neq 0$, ${\bf q}\neq 0$,
558: $\bk+{\bf p}\neq 0$, and $\bk-{\bf q}\neq 0$.
559: 
560: To proceed further, we need to invoke some approximation. A natural
561: mean-field approximation, in the presence of broken gauge symmetry, is the
562: HFB approximation. This is used to simplify the higher-order terms (36) and
563: (37). Employing the designations
564: \be
565: \label{38}
566: \om_k \equiv \frac{k^2}{2m} + 2\rho \Phi_0 - \mu_1
567: \ee
568: and
569: \be
570: \label{39}
571: \Dlt \equiv (\rho_0 +\sgm_1) \Phi_0 \; ,
572: \ee
573: we obtain
574: \be
575: \label{40}
576: H = E_{HFB} + \sum_{k\neq 0} \left [ \om_k a_k^\dgr a_k +
577: \frac{\Dlt}{2}\left ( a_k^\dgr a_{-k}^\dgr + a_{-k} a_k \right )
578: \right ] \; ,
579: \ee
580: where the nonoperator term is
581: \be
582: \label{41}
583: E_{HFB} =  H^{(0)} \; - \; \frac{\Phi_0}{2\rho}
584: \left ( 2\rho_1^2 + \sgm_1^2 \right ) N \; .
585: \ee
586: 
587: Hamiltonian (40) can be diagonalized by means of the Bogolubov canonical
588: transformation $a_k=u_kb_k+v^*_{-k}b^\dgr_{-k}$, in which
589: $$
590: u_k^2 = \frac{\om_k+\ep_k}{2\ep_k} \; , \qquad
591: v_k^2 = \frac{\om_k-\ep_k}{2\ep_k} \; ,
592: $$
593: with the Bogolubov spectrum
594: \be
595: \label{42}
596: \ep_k =\sqrt{\om_k^2 -\Dlt^2} \; .
597: \ee
598: Then one gets
599: \be
600: \label{43}
601: H = E_B + \sum_{k\neq 0} \ep_k b_k^\dgr b_k \; ,
602: \ee
603: where
604: \be
605: \label{44}
606: E_B \equiv E_{HFB} + \frac{1}{2}\; \sum_{k\neq 0} (\ep_k -\om_k)\; .
607: \ee
608: 
609: By the Bogolubov [18] and Hugenholtz-Pines [22] theorems, the spectrum is
610: to be gapless, which implies that
611: \be
612: \label{45}
613: \lim_{k\ra 0} \ep_k = 0 \; , \qquad \ep_k\geq 0\; .
614: \ee
615: From here it follows that
616: \be
617: \label{46}
618: \mu_1 = (\rho +\rho_1 -\sgm_1) \Phi_0 \; .
619: \ee
620: The condensate multiplier (32) in the HFB approximation, when $\xi=0$,
621: becomes
622: \be
623: \label{47}
624: \mu_0 = (\rho + \rho_1 +\sgm_1) \Phi_0 \; .
625: \ee
626: It is important to emphasize that the form of $\mu_1$ in Eq. (46) is
627: necessary and sufficient for making the spectrum gapless. That it is
628: sufficient follows at once after substituting Eq. (46) into the Bogolubov
629: spectrum (42). And the necessity stems from the Bogolubov theorem [18]
630: which can be formulated as the inequality
631: $$
632: \vert \mu_1 - \frac{k^2}{2m} + \Sigma_{12}(\bk,0) -
633: \Sigma_{11}(\bk,0) \vert \leq \frac{k^2}{2m_0} \; ,
634: $$
635: where $\Sigma_{12}$ and $\Sigma_{11}$ are the normal and anomalous
636: self-energies. Setting here $\bk=0$ leads to the Hugenholtz-Pines relation
637: $$
638: \mu_1 = \Sigma_{11}(0,0) - \Sigma_{12}(0,0) \; .
639: $$
640: In the HFB approximation, one has
641: $$
642: \Sigma_{11}(0,0) = 2\rho \Phi_0\; , \qquad \Sigma_{12}(0,0) =
643: (\rho_0 + \sgm_1) \Phi_0\; ,
644: $$
645: from where one immediately obtains Eq. (46).
646: 
647: With multiplier (46), spectrum (42) takes the form
648: \be
649: \label{48}
650: \ep_k = \sqrt{(ck)^2+\left ( \frac{k^2}{2m}\right )^2 } \; ,
651: \ee
652: in which the sound velocity is
653: \be
654: \label{49}
655: c \equiv \sqrt{\frac{\Dlt}{m}} \; .
656: \ee
657: 
658: For the diagonal Hamiltonian (43), it is straightforward to calculate all
659: averages, such as the normal average
660: \be
661: \label{50}
662: n_k \; \equiv \; <a_k^\dgr a_k>
663: \ee
664: and the anomalous average
665: \be
666: \label{51}
667: \sgm_k \; \equiv \; < a_k a_{-k}> \; .
668: \ee
669: Their integration over momenta gives the density of uncondensed particles
670: \be
671: \label{52}
672: \rho_1 = \int n_k \; \frac{d\bk}{(2\pi)^3}
673: \ee
674: and, respectively, the anomalous average
675: \be
676: \label{53}
677: \sgm_1 = \int \sgm_k \; \frac{d\bk}{(2\pi)^3} \; .
678: \ee
679: The quantity $|\sgm_1|$ can be interpreted as the density of pair-correlated
680: atoms [19].
681: 
682: In what follows, we concentrate on the zero-temperature properties of the
683: system. When $T=0$, then $<b_k^\dgr b_k>=0$. The normal average (50) is
684: \be
685: \label{54}
686: n_k = \frac{\om_k-\ep_k}{2\ep_k} \; ,
687: \ee
688: while the anomalous average (51) becomes
689: \be
690: \label{55}
691: \sgm_k = -\; \frac{\Dlt}{2\ep_k} \; .
692: \ee
693: Combining Eqs. (38), (39), (46), and (49), we have
694: $$
695: \om_k = \frac{k^2}{2m} + \Dlt \; , \qquad \Dlt =mc^2 \; .
696: $$
697: The equation for the sound velocity (49) can be represented as
698: \be
699: \label{56}
700: mc^2 = ( \rho_0 + \sgm_1)\Phi_0 \; .
701: \ee
702: Note that from here, in the limit of asymptotically weak interactions, we
703: get the Bogolubov expression
704: $$
705: c \simeq \sqrt{ \frac{\rho_0\Phi_0}{m}} \qquad (\Phi_0\ra 0) \; .
706: $$
707: 
708: For the density of uncondensed atoms (52), we find
709: \be
710: \label{57}
711: \rho_1 = \int \frac{\om_k-\ep_k}{2\ep_k} \; \frac{d\bk}{(2\pi)^3} =
712: \frac{(mc)^3}{3\pi^2} \; .
713: \ee
714: And for the anomalous average (53), we have
715: \be
716: \label{58}
717: \sgm_1 = -\; \frac{\Dlt}{2}\; \int \frac{1}{\ep_k}\;
718: \frac{d\bk}{(2\pi)^3} \; .
719: \ee
720: 
721: The integral $\int d\bk/\ep_k$ in Eq. (58) is ultraviolet divergent.
722: To overcome this, we can resort to the standard procedure of analytic
723: regularization [5,32]. For this purpose, we, first, consider the integral
724: in the limit of asymptotically small $\Phi_0$, when the dimensional
725: regularization is applicable, and then analytically continue the result
726: to arbitrary interactions. The dimensional regularization gives
727: $$
728: \int \frac{1}{\ep_k}\; \frac{d\bk}{(2\pi)^3} = -\;
729: \frac{2}{\pi^2}\; m^{3/2} \; \sqrt{\rho_0\Phi_0} \; .
730: $$
731: Using this in Eq. (58), we obtain
732: \be
733: \label{59}
734: \sgm_1 = \frac{(mc)^2}{\pi^2}\; \sqrt{m\rho_0\Phi_0} \; .
735: \ee
736: 
737: It is important to stress that the anomalous density (59) enjoys the natural
738: limiting property
739: \be
740: \label{60}
741: \sgm_1 \ra 0 \qquad (\rho_0\ra 0) \; .
742: \ee
743: The physics of property (60) is evident. The existence of both the condensate
744: density $\rho_0$ and the anomalous density $\sgm_1$ is caused by the gauge
745: symmetry breaking. Both of them are nonzero as soon as the symmetry is broken,
746: while both become zero if the symmetry is restored. Any of these quantities
747: could be treated as an order parameter for the broken-symmetry phase. So, both
748: these quantities, $\rho_0$ and $\sgm_1$, have to nullify simultaneously, when
749: one of them tends to zero.
750: 
751: We may also note that simplifying $\sgm_1$ by replacing $\sqrt{m\rho_0\Phi_0}$
752: by $mc$, as is done in Ref. [14], is admissible only in the limit of weak
753: interactions, when $\rho_0\sim\rho$, But for strong interactions, when
754: $\rho_0\ra 0$, this replacement does not hold, since then the limiting
755: property (60) is not satisfied. Therefore, the results of Ref. [14] are
756: quantitatively correct in the limit of weak interactions, though for strong
757: interactions, they may give only a qualitative picture. And our aim in the
758: present paper is to give a careful analysis of the system properties for
759: arbitrary strong interactions in the whole range of $\Phi_0\in[0,\infty)$.
760: This requires to employ Eq. (59), which explicitly satisfy the limiting
761: condition (60).
762: 
763: \section{System Characteristics at Varying Interactions}
764: 
765: The effective interaction strength can be characterized by the dimensionless
766: {\it gas parameter}
767: \be
768: \label{61}
769: \gm \equiv \rho^{1/3} a_s \; .
770: \ee
771: One often uses the quantity $\rho a_s^3$ as a parameter quantifying
772: the interaction strength. However, parameter (61), to our mind, is
773: more convenient, since it better distinguishes between weak and
774: strong interactions. Thus, the majority of experiments with ultracold
775: trapped gases [1--7] deals with weakly interacting atoms, so that
776: $\rho a_s^3\sim 10^{-8}-10^{-4}$, which corresponds to
777: $\gm\sim 10^{-3}-10^{-2}$. The effective interactions can be
778: noticeably strengthened by loading atoms in optical lattices [33].
779: Contrary to weakly interacting gases, superfluid $^4$He is a strongly
780: interacting system. The $^4$He atoms can be represented by hard spheres
781: of diameter $a_s$ [34,35]. At saturated vapour pressure, one has [35]
782: $a_s=2.139\AA$ and $\rho\approx 0.022\AA^{-3}$, hence $\rho a_s^3\approx
783: 0.215$, which is yet much less than one. But the corresponding $\gm\approx
784: 0.599$ is close to one. So, weak  interactions are characterized by
785: $\gm\ll 1$, while a strongly interacting system has $\gm\sim 1$.
786: 
787: It is convenient to introduce the dimensionless sound velocity
788: \be
789: \label{62}
790: s \equiv \frac{mc}{\rho^{1/3}} \; ,
791: \ee
792: the fraction of uncondensed atoms
793: \be
794: \label{63}
795: n_1 \equiv \frac{\rho_1}{\rho} = \frac{N_1}{N} \; ,
796: \ee
797: and the anomalous fraction
798: \be
799: \label{64}
800: \sgm \equiv \frac{\sgm_1}{\rho} \; .
801: \ee
802: In these dimensionless quantities, Eq. (56) writes as
803: \be
804: \label{65}
805: s^2 = 4\pi(n_0+\sgm)\gm  \; .
806: \ee
807: Here
808: \be
809: \label{66}
810: n_0 = 1-n_1
811: \ee
812: and
813: \be
814: \label{67}
815: n_1 = \frac{s^3}{3\pi^2} \; ,
816: \ee
817: while the anomalous fraction (64) is
818: \be
819: \label{68}
820: \sgm = \frac{2s^2}{\pi^{3/2}}\; \sqrt{\gm n_0 } \; .
821: \ee
822: Four quantities, $s$, $n_0$, $n_1$, and $\sgm$, are defined by the system
823: of four equations (65) to (68).
824: 
825: In the limit of asymptotically weak interactions, when $\gm\ra 0$ and
826: $\sgm\ra 0$, the condensate fraction and sound velocity tend to the
827: Bogolubov expressions
828: \be
829: \label{69}
830: n_B =  1 \; - \; \frac{8}{3\sqrt{\pi}}\; \gm^{3/2}
831: \ee
832: and, respectively,
833: \be
834: \label{70}
835: s_B = 2\sqrt{\pi\gm} \; .
836: \ee
837: For the higher orders with respect to $\gm$, we find
838: \be
839: \label{71}
840: n_0 \simeq 1 \; - \; \frac{8}{3\sqrt{\pi}}\; \gm^{3/2} \; - \;
841: \frac{64}{3\pi}\; \gm^3 \; - \; \frac{640}{9\pi^{3/2}}\; \gm^{9/2} \; ,
842: \ee
843: \be
844: \label{72}
845: s \simeq 2\sqrt{\pi}\; \gm^{1/2} + \frac{16}{3}\; \gm^2 +
846: \frac{32}{9\sqrt{\pi}}\; \gm^{7/2} \; - \; \frac{3904}{27\pi}\; \gm^5 \; ,
847: \ee
848: \be
849: \label{73}
850: \sgm \simeq \frac{8}{\sqrt{\pi}} \; \gm^{3/2} + \frac{32}{\pi}\;\gm^3 \; -
851: \; \frac{64}{\pi^{3/2}}\; \gm^{9/2} \; .
852: \ee
853: 
854: For asymptotically strong interactions, when $\gm\ra\infty$, we obtain
855: \be
856: \label{74}
857: n_0 \simeq \frac{\pi}{64}\; \gm^{-3} \; - \; \frac{1}{512} \left (
858: \frac{\pi^5}{9}\right )^{1/3} \gm^{-5} \; ,
859: \ee
860: \be
861: \label{75}
862: s \simeq (3\pi^2)^{1/3}\; - \; \frac{1}{64} \left (
863: \frac{\pi^5}{9}\right )^{1/3} \gm^{-3} + \frac{1}{1536} \left (
864: \frac{\pi^7}{3}\right )^{1/3} \gm^{-5} \; ,
865: \ee
866: \be
867: \label{76}
868: \sgm \simeq \frac{(9\pi)^{1/3}}{4} \; \gm^{-1} \; - \;
869: \frac{\pi}{64}\;\gm^{-3} \; - \; \frac{1}{128}\left (
870: \frac{\pi^4}{3}\right )^{1/3} \gm^{-4} + \frac{1}{512} \left (
871: \frac{\pi^5}{9} \right )^{1/3} \gm^{-5} \; .
872: \ee
873: So, the condensate fraction tends to zero, together with the anomalous
874: fraction, as $\gm\ra\infty$, in agreement with condition (60).
875: 
876: For the whole region of $\gm$, we solve numerically the system of Eqs. (65)
877: to (68), showing the results in Figs. 1,2, and 3.
878: 
879: Figure 1 presents the condensate fraction (66) and this fraction (69) in the
880: Bogolubov approximation. At small $\gm$, up to $\gm\approx 0.1$, $n_0$ and
881: $n_B$ practically coincide with each other. For $\gm >0.1$, the Bogolubov
882: approximation overestimates the condensate fraction. But $n_B=0$ at
883: $\gm=0.762$, while $n_0$ is yet finite, though small. The condensate
884: fraction of a homogeneous Bose gas, at zero temperature, as a function of
885: the gas parameter was calculated by means of the Monte Carlo simulation by
886: Giorgini et al. [36] up to $\gm\approx 0.5$. For this region of
887: $\gm$, our $n_0$ is in good agreement with the Monte Carlo calculations.
888: Monte Carlo techniques have also been used for studying the condensate
889: fraction of trapped atoms [37--39]. But the latter results cannot be
890: directly compared with $n_0$ in a homogeneous gas, since in traps, the
891: condensate fraction is a function of spatial variables as well as of the
892: trap shape [37--41]. What is possible and interesting to compare is the
893: condensate fraction in superfluid $^4$He at zero temperature and our
894: $n_0$ at $\gm\approx 0.6$ corresponding to liquid $^4$He with hard-core
895: interactions. For $\gm\approx 0.6$, we have $n_0\approx 0.15$, which is
896: close to the condensate fraction $n_0\approx 0.1$, measured in experiments
897: (as is reviewed in Refs. [3,42]) as well as obtained by Monte Carlo
898: simulations (see review articles [43,44]).
899: 
900: Figure 2 shows the dimensionless sound velocity (62) and its Bogolubov
901: approximation (70). These quantities practically coincide up to $\gm\approx
902: 0.1$. For $\gm>0.1$ the Bogolubov form $s_B$ underestimates $s$ till
903: $\gm\approx 0.7$, after which it overestimates the latter.
904: 
905: In Fig. 3, we compare the fraction of uncondensed atoms (67) with the
906: anomalous fraction (68). As is seen, $\sgm>n_1$ up to $\gm\approx 0.7$. The
907: anomalous fraction $\sgm$ becomes substantially smaller than $n_1$ only for
908: very large $\gm\gg 1$. This is in agreement with other calculations [45]
909: confirming that anomalous averages cannot be neglected at low temperatures.
910: 
911: Let us now analyse the ground-state energy
912: \be
913: \label{77}
914: E\; \equiv \; <H> + \mu N \qquad (T=0) \; .
915: \ee
916: According to Eq. (43),
917: $$
918: <H>\; = \; E_B \qquad (T=0) \; .
919: $$
920: The system chemical potential, defined in Eq. (18), is expressed through the
921: Lagrange multipliers (46) and (47) and the fractions $n_0$ and $n_1$, which
922: gives
923: \be
924: \label{78}
925: \mu = (1+n_1+\sgm-2\sgm n_1)\rho\Phi_0 \; .
926: \ee
927: For the dimensionless chemical potential
928: \be
929: \label{79}
930: \overline\mu \equiv \frac{2m\mu}{\rho^{2/3}} \; ,
931: \ee
932: we get
933: \be
934: \label{80}
935: \overline\mu = 8\pi\gm(1+n_1+\sgm-2\sgm n_1) \; .
936: \ee
937: From Eq. (44), we have
938: \be
939: \label{81}
940: E_B = E_{HFB} + N \int \frac{\ep_k-\om_k}{2\rho}\;
941: \frac{d\bk}{(2\pi)^3} \; ,
942: \ee
943: where the integral is calculated invoking the dimensional regularization
944: [5], giving
945: $$
946: \int \frac{\ep_k-\om_k}{2\rho}\; \frac{d\bk}{(2\pi)^3} =
947: \frac{8(mc)^5}{15\pi^2m\rho} \; .
948: $$
949: Equations (34) and (41) yield
950: \be
951: \label{82}
952: \frac{E_{HFB}}{N} = \frac{\rho\Phi_0}{2}\left ( n_0^2 - 2n_1^2 -\sgm^2
953: \right ) - \mu_0 n_0 \; .
954: \ee
955: Summarizing these formulas, we find
956: \be
957: \label{83}
958: \frac{E}{N} = \frac{\rho\Phi_0}{2}\left ( 1 + n_1^2 - 2\sgm n_1 -
959: \sgm^2 \right ) + \frac{8(mc)^5}{15\pi^2m\rho} \; .
960: \ee
961: It is convenient to define the dimensionless ground-state energy
962: \be
963: \label{84}
964: E_0 \equiv \frac{2mE}{\rho^{2/3}N} \; ,
965: \ee
966: for which we obtain
967: \be
968: \label{85}
969: E_0 =  4\pi\gm \left ( 1 +n_1^2 - 2\sgm n_1 -\sgm^2 +
970: \frac{4s^5}{15\pi^3\gm}  \right ) \; .
971: \ee
972: This can be compared with the Lee-Huang-Yang approximation [46--48]
973: \be
974: \label{86}
975: E_{LHY} = 4\pi\gm \left ( 1 +
976: \frac{128}{15\sqrt{\pi}}\; \gm^{3/2} \right ) \; ,
977: \ee
978: derived for small $\gm$.
979: 
980: In the limits of asymptotically small and large $\gm$, the dimensionless
981: chemical potential (80) tends to
982: \be
983: \label{87}
984: \overline\mu \ra 8\pi\gm + \frac{256}{3}\; \sqrt{\pi}\; \gm^{5/2}
985: \qquad (\gm\ra 0)
986: \ee
987: and, respectively,
988: \be
989: \label{88}
990: \overline\mu \ra 16\pi\gm \qquad (\gm\ra\infty) \; .
991: \ee
992: The dimensionless ground-state energy (85) at small $\gm\ll 1$ possesses the
993: expansion
994: \be
995: \label{89}
996: E_0 \simeq 4\pi\gm + \frac{512}{15}\; \sqrt{\pi}\; \gm^{5/2} +
997: \frac{512}{9}\; \gm^4 \; ,
998: \ee
999: which reproduces the Lee-Huang-Yang approximation (86) for $\gm\ra 0$. And
1000: for large $\gm\gg 1$, we find
1001: \be
1002: \label{90}
1003: E_0 \simeq 8\pi\gm + \frac{6}{5}\left ( 9\pi^4\right )^{1/3} \; - \;
1004: \frac{3}{4}\left ( 3\pi^5\right )^{1/3}\gm^{-1} +
1005: \frac{1}{64}\left ( 3\pi^8\right )^{1/3}\gm^{-4} \; .
1006: \ee
1007: Note that expressions (87) and (88) can also be obtained from Eq. (89)
1008: and (90) using the relation $\mu=\prt E/\prt N$, valid for $T=0$.
1009: 
1010: Figure 4 illustrates the behaviour of the ground-state energy (85) and
1011: the Lee-Huang-Yang approximation (86). The latter is known to practically
1012: coincide with the energy calculated trough the Monte Carlo simulations
1013: [36] up to $\gm\approx 0.4$. As is seen from the figure, our $E_0$ is also
1014: very close to $E_{LHY}$ in the region $0\leq\gm < 0.4$, but is lower than
1015: $E_{LHY}$ for $\gm>0.4$. Hence, $E_0$ well reproduces the available data
1016: of Monte Carlo calculations up to $\gm\approx 0.4$.
1017: 
1018: \section{Conclusion}
1019: 
1020: The notion of representative statistical ensembles is applied to Bose
1021: systems with broken global gauge symmetry. A general procedure is described
1022: for constructing the grand Hamiltonian for the representative ensemble of
1023: an arbitrary equilibrium Bose system. A self-consistent mean-field theory is
1024: developed, which is both conserving and gapless. The properties of a uniform
1025: Bose gas at zero temperature are studied both analytically and numerically
1026: for the gas parameter varying between zero and infinity. Thus, in the frame
1027: of the suggested approach, strongly interacting systems can also be considered.
1028: For instance, as is known [35,43], some of the properties of superfluid
1029: $^4$He can be understood by treating the potential as a hard-core interaction
1030: of diameter $a_s=2.139\AA$, which, at saturated vapor pressure, corresponds
1031: to $\gm\approx 0.6$. For the latter $\gm$, we find the condensate fraction
1032: $n_0$ of order $10\%$, which agrees with the condensate fraction in helium
1033: at zero temperature, measured in experiments [42] and found in Monte Carlo
1034: simulations [43], being also of order $10\%$. Another application of the
1035: developed self-consistent mean-field theory with arbitrary strong
1036: interactions could be the description of Bose-Einstein condensation of
1037: multiquark clusters in nuclear matter [49,50].
1038: 
1039: \vskip 5mm
1040: 
1041: {\bf Acknowledgement}
1042: 
1043: \vskip 3mm
1044: 
1045: One of the authors (V.I.Y.) is grateful for financial support to the
1046: German Research Foundation and for discussions to M. Girardeau, R. Graham,
1047: and H. Kleinert.
1048: 
1049: \newpage
1050: 
1051: \begin{thebibliography}{99}
1052: 
1053: \bibitem{1}
1054: A.S. Parkins and D.F. Walls, Phys. Rep. {\bf 303}, 1 (1998).
1055: 
1056: \bibitem{2}
1057: F. Dalfovo, S. Giorgini, L.P. Pitaevskii, and S. Stringari, Rev. Mod. Phys.
1058: {\bf 71}, 463 (1999).
1059: 
1060: \bibitem{3}
1061: P.W. Courteille, V.S. Bagnato and V.I. Yukalov, Laser Phys. {\bf 11},
1062: 659 (2001).
1063: 
1064: \bibitem{4}
1065: A.L. Fetter, J. Low Temp. Phys. {\bf 129}, 263 (2002).
1066: 
1067: \bibitem{5}
1068: J.O. Andersen, Rev. Mod. Phys. {\bf 76}, 599 (2004).
1069: 
1070: \bibitem{6}
1071: K. Bongs and K. Sengstock, Rep. Prog. Phys. {\bf 67}, 907 (2004).
1072: 
1073: \bibitem{7}
1074: V.I. Yukalov, Laser Phys. Lett. {\bf 1}, 435 (2004).
1075: 
1076: \bibitem{8}
1077: N.N. Bogolubov, J. Phys. (Moscow) {\bf 11}, 23 (1947).
1078: 
1079: \bibitem{9}
1080: N.N. Bogolubov, Moscow Univ. Phys. Bull. {\bf 7}, 43 (1947).
1081: 
1082: \bibitem{10}
1083: E. Timmermans, P. Tommasini, M. Hussein, and A. Kerman, Phys. Rep. {\bf
1084: 315}, 199 (1999).
1085: 
1086: \bibitem{11}
1087: R.A. Duine and H.T.C. Stoof, Phys. Rep. {\bf 396}, 115 (2004).
1088: 
1089: \bibitem{12}
1090: P.C. Hohenberg and P.C. Martin, Ann. Phys. {\bf 34}, 291 (1965).
1091: 
1092: \bibitem{13}
1093: V.I. Yukalov, Phys. Rev. {\bf E 72}, 066119 (2005).
1094: 
1095: \bibitem{14}
1096: V.I. Yukalov and H. Kleinert, Phys. Rev. A {\bf 73}, 063612 (2006).
1097: 
1098: \bibitem{15}
1099: F.A. Berezin, {\it Method of Second Quantization} (Academic, New York,
1100: 1966).
1101: 
1102: \bibitem{16}
1103: V.I. Yukalov, {\it Statistical Green's Functions} (Queen's University,
1104: Kingston, 1998).
1105: 
1106: \bibitem{17}
1107: N.N. Bogolubov, {\it Lectures on Quantum Statistics} (Gordon and
1108: Breach, New York, 1967), Vol. 1.
1109: 
1110: \bibitem{18}
1111: N.N. Bogolubov, {\it Lectures on Quantum Statistics} (Gordon and
1112: Breach, New York, 1970), Vol. 2.
1113: 
1114: \bibitem{19}
1115: V.I. Yukalov, Laser Phys. {\bf 16}, 511 (2006).
1116: 
1117: \bibitem{20}
1118: J.W. Gibbs, {\it Collected Works} (Longmans, New York, 1954), Vol. 2.
1119: 
1120: \bibitem{21}
1121: V.I. Yukalov, Phys. Rep. {\bf 208}, 395 (1991).
1122: 
1123: \bibitem{22}
1124: N.M. Hugenholtz and D. Pines, Phys. Rev. {\bf 116}, 489 (1959).
1125: 
1126: \bibitem{23}
1127: J. Gavoret and P. Nozi\`eres, Ann. Phys. (N.Y.) {\bf 28}, 349 (1964).
1128: 
1129: \bibitem{24}
1130: M. Girardeau and R. Arnowitt, Phys. Rev. {\bf 113}, 755 (1959).
1131: 
1132: \bibitem{25}
1133: M. Girardeau, J. Math. Phys. {\bf 3}, 131 (1962).
1134: 
1135: \bibitem{26}
1136: M. Girardeau,  Phys. Rev. {\bf 115}, 1090 (1959).
1137: 
1138: \bibitem{27}
1139: F. Takano, Phys. Rev. {\bf 123}, 699 (1961).
1140: 
1141: \bibitem{28}
1142: V.I. Yukalov, Laser Phys. Lett. {\bf 3}, 406 (2006).
1143: 
1144: \bibitem{29}
1145: P.O. L\"owdin, Phys. Rev. {\bf 97}, 1474 (1955).
1146: 
1147: \bibitem{30}
1148: A.J. Coleman and V.I. Yukalov, {\it Reduced Density Matrices} (Springer,
1149: Berlin, 2000).
1150: 
1151: \bibitem{31}
1152: N.N. Bogolubov and N.N. Bogolubov Jr., {\it Introduction to Quantum
1153: Statistical Mechanics} (Gordon and Breach, Lausanne, 1994).
1154: 
1155: \bibitem{32}
1156: H. Kleinert, {\it Path Integrals} (World Scientific, Singapore, 2004).
1157: 
1158: \bibitem{33}
1159: K. Xu, Y. Liu, D.E. Miller, J.K. Chim, W. Setiawan, and W. Ketterle,
1160: Phys. Rev. Lett. {\bf 96}, 180405 (2006).
1161: 
1162: \bibitem{34}
1163: M.H. Kalos, Phys. Rev. A {\bf 2}, 250 (1970).
1164: 
1165: \bibitem{35}
1166: M.H. Kalos, D. Levesque, and I. Verlet, Phys. Rev. A {\bf 9}, 2178 (1974).
1167: 
1168: \bibitem{36}
1169: S. Giorgini, J. Boronat, and J. Casulleras, Phys. Rev. A {\bf 60}, 5129
1170: (1999).
1171: 
1172: \bibitem{37}
1173: J.L. DuBois and H.R. Glyde, Phys. Rev. A {\bf 63}, 023602 (2001).
1174: 
1175: \bibitem{38}
1176: J.L. DuBois and H.R. Glyde, Phys. Rev. A {\bf 68}, 033602 (2003).
1177: 
1178: \bibitem{39}
1179: W. Purwanto and S. Zhang, Phys. Rev. A {\bf 72}, 053610 (2005).
1180: 
1181: \bibitem{40}
1182: K. Nho and D. Blume, Phys. Rev. Lett. {\bf 95}, 193601 (2005).
1183: 
1184: \bibitem{41}
1185: K. Nho and D.P. Landau, Phys. Rev. A {\bf 73}, 033606 (2006).
1186: 
1187: \bibitem{42}
1188: F.W. Wirth and R.B. Hallock, Phys. Rev. B {\bf 35}, 34 (1987).
1189: 
1190: \bibitem{343}
1191: D.M. Ceperley, Rev. Mod. Phys. {\bf 67}, 279 (1995).
1192: 
1193: \bibitem{44}
1194: M. Boninsegni, N.V. Prokofiev, and B.V. Svistunov, Phys. Rev. E {\bf 74},
1195: 036701 (2006).
1196: 
1197: \bibitem{45}
1198: V.I. Yukalov and E.P. Yukalova, Laser Phys. Lett. {\bf 2}, 506 (2005).
1199: 
1200: \bibitem{46}
1201: T.D. Lee and C.N. Yang, Phys. Rev. {\bf 105}, 1119 (1957).
1202: 
1203: \bibitem{47}
1204: T.D. Lee, K. Huang, and C.N. Yang, Phys. Rev. {\bf 106}, 1135 (1957).
1205: 
1206: \bibitem{48}
1207: T.D. Lee and C.N. Yang, Phys. Rev. {\bf 112}, 1419 (1958).
1208: 
1209: \bibitem{49}
1210: V.I. Yukalov and E.P. Yukalova, Phys. Part. Nucl. {\bf 28}, 37 (1997).
1211: 
1212: \bibitem{50}
1213: A. Faessler, A.J. Buchmann, M.I. Krivoruchenko, and B.V. Martemyanov,
1214: Phys. Lett. B {\bf 391}, 255 (1997).
1215: 
1216: \end{thebibliography}
1217: 
1218: 
1219: \newpage
1220: 
1221: {\Large{\bf Figure Captions}}
1222: 
1223: \vskip 1cm
1224: 
1225: {\bf Fig. 1}. Condensate fraction $n_0$ (solid line) and its Bogolubov
1226: approximation $n_B$ (dashed line) as functions of the gas parameter $\gm$.
1227: 
1228: \vskip 5mm
1229: 
1230: {\bf Fig. 2}. Dimensionless sound velocity $s$ (solid line) and its
1231: Bogolubov approximation $s_B$ (dashed line) as functions of the gas
1232: parameter $\gm$.
1233: 
1234: \vskip 5mm
1235: 
1236: {\bf Fig. 3}. Fraction of uncondensed atoms $n_1$ (dashed line) and
1237: anomalous fraction $\sgm$ (solid line) as functions of the gas parameter
1238: $\gm$.
1239: 
1240: \vskip 5mm
1241: 
1242: {\bf Fig. 4}. Dimensionless ground-state energy $E_0$ (solid line) and the
1243: Lee-Huang-Yang approximation $E_{LHY}$ (dashed line) as functions of the gas
1244: parameter $\gm$.
1245: 
1246: \newpage
1247: 
1248: \begin{figure}[h]
1249: \centerline{\epsfig{file=Fig1n0B.eps,angle=0,width=16cm}}
1250: \vskip 1cm
1251: \caption{Condensate fraction $n_0$ (solid line) and its Bogolubov
1252: approximation $n_B$ (dashed line) as functions of the gas parameter $\gm$.
1253: }
1254: \label{fig:Fig.1}
1255: \end{figure}
1256: 
1257: \newpage
1258: 
1259: \begin{figure}[h]
1260: \centerline{\epsfig{file=Fig2sB.eps,angle=0,width=16cm}}
1261: \vskip 1cm
1262: \caption{Dimensionless sound velocity $s$ (solid line) and its
1263: Bogolubov approximation $s_B$ (dashed line) as functions of the gas
1264: parameter $\gm$.
1265: }
1266: \label{fig:Fig.2}
1267: \end{figure}
1268: 
1269: \newpage
1270: 
1271: \begin{figure}[h]
1272: \centerline{\epsfig{file=Fig3sgm.eps,angle=0,width=16cm}}
1273: \vskip 1cm
1274: \caption{Fraction of uncondensed atoms $n_1$ (dashed line) and
1275: anomalous fraction $\sgm$ (solid line) as functions of the gas parameter
1276: $\gm$.
1277: }
1278: \label{fig:Fig.3}
1279: \end{figure}
1280: 
1281: \newpage
1282: 
1283: \begin{figure}[h]
1284: \centerline{\epsfig{file=Fig4E.eps,angle=0,width=16cm}}
1285: \vskip 1cm
1286: \caption{Dimensionless ground-state energy $E_0$ (solid line) and the
1287: Lee-Huang-Yang approximation $E_{LHY}$ (dashed line) as functions of the gas
1288: parameter $\gm$.
1289: }
1290: \label{fig:Fig.4}
1291: \end{figure}
1292: 
1293: 
1294: 
1295: \end{document}
1296: