1:
2: \documentclass[eqsecnum,showkeys,secnumarabic,showpacs,twocolumn,hyperref,pra]{revtex4}
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: \usepackage{amsfonts}
5: \usepackage{amsmath}
6: \usepackage{amssymb}
7: \usepackage{graphicx}
8: \usepackage{hyperref}
9:
10: \setcounter{MaxMatrixCols}{10}
11:
12: \newtheorem{theorem}{Theorem}
13: \newtheorem{acknowledgement}[theorem]{Acknowledgement}
14: \newtheorem{algorithm}[theorem]{Algorithm}
15: \newtheorem{axiom}[theorem]{Axiom}
16: \newtheorem{claim}[theorem]{Claim}
17: \newtheorem{conclusion}[theorem]{Conclusion}
18: \newtheorem{condition}[theorem]{Condition}
19: \newtheorem{conjecture}[theorem]{Conjecture}
20: \newtheorem{corollary}[theorem]{Corollary}
21: \newtheorem{criterion}[theorem]{Criterion}
22: \newtheorem{definition}[theorem]{Definition}
23: \newtheorem{example}[theorem]{Example}
24: \newtheorem{exercise}[theorem]{Exercise}
25: \newtheorem{lemma}[theorem]{Lemma}
26: \newtheorem{notation}[theorem]{Notation}
27: \newtheorem{problem}[theorem]{Problem}
28: \newtheorem{proposition}[theorem]{Proposition}
29: \newtheorem{remark}[theorem]{Remark}
30: \newtheorem{solution}[theorem]{Solution}
31: \newtheorem{summary}[theorem]{Summary}
32: \newenvironment{proof}[1][Proof]{\noindent\textbf{#1.} }{\ \rule{0.5em}{0.5em}}
33:
34: \begin{document}
35:
36: \preprint{}
37: \title{Extended Ensemble Theory, Spontaneous Symmetry Breaking, and Phase
38: Transitions}
39: \author{Ming-wen Xiao}
40: \affiliation{Department of Physics, Nanjing University, Nanjing 210093, China}
41: \keywords{Quantum ensemble theory, Quantum statistical mechanics, General
42: studies of phase transitions}
43: \pacs{05.30.Ch, 05.30.-d, 64.60.-i}
44:
45: \begin{abstract}
46: In this paper, as a personal review, we suppose a possible extension of
47: Gibbs ensemble theory so that it can provide a reasonable description to
48: phase transitions and spontaneous symmetry breaking. The extension is
49: founded on three hypotheses, and can be regarded as a microscopic edition of
50: the Landau phenomenological theory of phase transitions. Within its
51: framework, the stable state of a system is determined by the evolution of
52: order parameter with temperature according to such a principle that the
53: entropy of the system will reach its minimum in this state. The evolution of
54: order parameter can cause change in representation of the system
55: Hamiltonian; different phases will realize different representations,
56: respectively; a phase transition amounts to a representation transformation.
57: Physically, it turns out that phase transitions originate from the automatic
58: interference among matter waves as temperature is cooled down. Typical
59: quantum many-body systems are studied with this extended ensemble theory. We
60: regain the Bardeen-Cooper-Schrieffer solution for the weak-coupling
61: superconductivity, and prove that it is stable. We find that
62: negative-temperature and laser phases arise from the same mechanism as phase
63: transitions, and that they are instable. For the ideal Bose gas, we
64: demonstrate that it will produce Bose-Einstein condensation (BEC) in the
65: thermodynamic limit, which confirms exactly Einstein's deep physical
66: insight. In contrast, there is no BEC either within the phonon gas in a
67: black body or within the ideal photon gas in a solid body. We prove that it
68: is not admissible to quantize Dirac field by using Bose-Einstein statistics.
69: We show that a structural phase transition belongs physically to the BEC
70: happening in configuration space, and that a double-well anharmonic system
71: will undergo a structural phase transition at a finite temperature. For the $%
72: O(N)$-symmetric vector model, we demonstrate that it will yield spontaneous
73: symmetry breaking and produce Goldstone bosons; and if it is coupled with a
74: gauge field, the gauge field will obtain a mass (Higgs mechanism). Also, we
75: show that an interacting Bose gas is stable only if the interaction is
76: repulsive. For the weak interaction case, we find that the BEC is a
77: \textquotedblleft $\lambda $\textquotedblright -transition and its
78: transition temperature can be lowered by the repulsive interaction. In
79: connection with liquid $^{4}\mathrm{He}$, it is found that the specific heat
80: at constant pressure $C_{P}$ will show a $T^{3}$ law at low temperatures,
81: which is in agreement with the experiment. If the system is further cooled
82: down, the theory predicts that $C_{P}$ will vanish linearly as $T\rightarrow
83: 0$, which is anticipating experimental verifications.
84: \end{abstract}
85:
86: \received{October 25, 2005}
87: \revised{\today}
88: \startpage{1}
89: \endpage{}
90: \maketitle
91: \tableofcontents
92:
93: \section{Introduction}
94:
95: \subsection{The Problem of Gibbs Ensemble Theory \label{PGET}}
96:
97: Since Gibbs \cite{Gibbs} established the ensemble theory of statistical
98: mechanics, there has arisen the problem as to whether it has the ability to
99: describe phase transitions \cite{Yang1}.
100:
101: Let us investigate the problem through the famous example, i.e., the
102: conventional low-$T_{c}$ superconductivity, which can be described by the
103: so-called Bardeen-Cooper-Schrieffer (BCS) Hamiltonian \cite{BCS1,BCS2,BCS3},
104: \begin{align}
105: H(c)& =\sum_{\mathbf{k}}\varepsilon (\mathbf{k})(c_{\mathbf{k}\uparrow
106: }^{\dag }c_{\mathbf{k}\uparrow }+c_{-\mathbf{k}\downarrow }^{\dag }c_{-%
107: \mathbf{k}\downarrow }) \notag \\
108: & -g\sum_{\mathbf{k},\mathbf{k}^{\prime }}c_{\mathbf{k}^{\prime }\uparrow
109: }^{\dag }c_{-\mathbf{k}^{\prime }\downarrow }^{\dag }c_{-\mathbf{k}%
110: \downarrow }c_{\mathbf{k}\uparrow }, \label{BCS}
111: \end{align}%
112: where $\varepsilon (\mathbf{k})$ denotes the energy relative to Fermi level,
113: $g>0$ the coupling strength, $c_{\mathbf{k}\uparrow }$ and $c_{\mathbf{k}%
114: \downarrow }$ ($c_{\mathbf{k}\uparrow }^{\dag }$ and $c_{\mathbf{k}%
115: \downarrow }^{\dag }$) the destruction (creation) operators for the
116: electrons with up and down spins, respectively. Here, the notation $H(c)$
117: expresses that the Hamiltonian is a function of $c_{\mathbf{k}\uparrow }$, $%
118: c_{\mathbf{k}\downarrow }$, $c_{\mathbf{k}\uparrow }^{\dag }$, and $c_{%
119: \mathbf{k}\downarrow }^{\dag }$.
120:
121: For BCS superconductivity, what is the most important physically is that the
122: Hamiltonian $H(c)$ will remain invariant under the transformation,%
123: \begin{equation}
124: G(\vartheta ,c)H(c)G^{\dagger }(\vartheta ,c)=H(c),
125: \end{equation}%
126: where
127: \begin{equation}
128: G(\vartheta ,c)=e^{-i\vartheta \sum_{\mathbf{k}}(c_{\mathbf{k}\uparrow
129: }^{\dag }c_{\mathbf{k}\uparrow }+c_{\mathbf{k}\downarrow }^{\dag }c_{\mathbf{%
130: k}\downarrow })},\text{ }\vartheta \in \lbrack 0,2\pi ). \label{Gauge}
131: \end{equation}%
132: This invariance is known as gauge symmetry. As a direct consequence of this
133: symmetry, one has%
134: \begin{equation}
135: \left\langle c_{-\mathbf{k}\downarrow }c_{\mathbf{k}\uparrow }\right\rangle =%
136: \mathrm{Tr}\big(c_{-\mathbf{k}\downarrow }c_{\mathbf{k}\uparrow }\rho (H(c))%
137: \big)=0, \label{Zero}
138: \end{equation}%
139: where $\langle \cdots \rangle $ denotes the Gibbs ensemble average with
140: respect to the Hamiltonian $H(c)$, and
141: \begin{equation}
142: \rho (H)=\frac{e^{-\beta H}}{\mathrm{Tr}\!\left( e^{-\beta H}\right) }
143: \end{equation}%
144: is the statistical density operator where $\beta =1/\left( k_{B}T\right) $
145: with $k_{B}$ and $T$ being the Boltzmann constant and temperature,
146: respectively. To prove Eq. (\ref{Zero}), it is sufficient to heed that
147: \begin{equation}
148: G\left( \frac{\pi }{2},c\right) c_{-\mathbf{k}\downarrow }c_{\mathbf{k}%
149: \uparrow }G^{\dag }\left( \frac{\pi }{2},c\right) =-c_{-\mathbf{k}\downarrow
150: }c_{\mathbf{k}\uparrow }. \label{Negative}
151: \end{equation}
152:
153: Eq. (\ref{Zero}) indicates that the electron-pair amplitude $\left\langle
154: c_{-\mathbf{k}\downarrow }c_{\mathbf{k}\uparrow }\right\rangle $ can never
155: become nonzero according to Gibbs ensemble theory, in other words, there can
156: not happen superconductivity for the BCS Hamiltonian of Eq. (\ref{BCS})
157: within Gibbs ensemble theory no matter how low the temperature is. This
158: proves definitely that Gibbs ensemble theory has no ability to describe
159: phase transitions.
160:
161: Even so, it is still hoped to do, as least as possible, some modification to
162: Gibbs theory so that it could describe phase transitions. Up to now, there
163: are two kinds of modification, they both arise from the theoretical studies
164: on the so-called Ising model \cite{Ising}, and become the popular and
165: predominant believes about phase transitions in the current world of
166: statistical physics.
167:
168: The first is to introduce the thermodynamic limit in the end of calculation
169: \cite{Onsager,Kaufmann,Yang2,Yang3,Georgii,GEmch,Emch,Bratteli}, that is,%
170: \begin{equation}
171: \left\langle c_{-\mathbf{k}\downarrow }c_{\mathbf{k}\uparrow }\right\rangle
172: =\lim_{V\rightarrow +\infty }\mathrm{Tr}\big(c_{-\mathbf{k}\downarrow }c_{%
173: \mathbf{k}\uparrow }\rho (H(c))\big), \label{Zero1}
174: \end{equation}%
175: where the $\lim_{V\rightarrow +\infty }$ means to take the thermodynamic
176: limit ($V\rightarrow +\infty $, $N\rightarrow +\infty $, and $N/V=$
177: constant) after performing the trace. Because%
178: \begin{equation}
179: \mathrm{Tr}\big(c_{-\mathbf{k}\downarrow }c_{\mathbf{k}\uparrow }\rho (H(c))%
180: \big)=0,
181: \end{equation}%
182: irrespective of the magnitudes of $V$, $N$, and $N/V$ ($H(c)$ is invariant
183: under gauge transformation, and that is true independently of $V$, $N$, and $%
184: N/V$, namely, the gauge symmetry has nothing to do with $V$, $N$, and $N/V$%
185: .), one has%
186: \begin{equation}
187: \lim_{V\rightarrow +\infty }\mathrm{Tr}\big(c_{-\mathbf{k}\downarrow }c_{%
188: \mathbf{k}\uparrow }\rho (H(c))\big)=0. \label{Zero2}
189: \end{equation}%
190: Therefore, this kind of modification can not make Gibbs ensemble theory have
191: the ability to describe phase transitions. Indeed, the thermodynamic limit
192: is important for calculating the statistical average of observable, but
193: itself alone is insufficient to act as the physical criterion for phase
194: transitions.
195:
196: The second is to introduce both the thermodynamic limit and an auxiliary
197: external field \cite{Ising,Kramers,Yang4,Huang},
198: %TCIMACRO{\TeXButton{Bwidetxt}{\begin{widetext}} }%
199: %BeginExpansion
200: \begin{widetext}
201: %EndExpansion
202: \begin{equation}
203: \left\langle c_{-\mathbf{k}\downarrow }c_{\mathbf{k}\uparrow }\right\rangle
204: =\lim_{\phi \rightarrow 0}\lim_{V\rightarrow +\infty }\mathrm{Tr\!}\left(
205: c_{-\mathbf{k}\downarrow }c_{\mathbf{k}\uparrow }\rho \Big(H(c)+\sum_{%
206: \mathbf{k}}(\phi c_{-\mathbf{k}\downarrow }c_{\mathbf{k}\uparrow }+\phi
207: ^{\dag }c_{\mathbf{k}\uparrow }^{\dag }c_{-\mathbf{k}\downarrow }^{\dag })%
208: \Big)\right) , \label{Pair}
209: \end{equation}%
210: %TCIMACRO{\TeXButton{Ewidetxt}{\end{widetext}} }%
211: %BeginExpansion
212: \end{widetext}
213: %EndExpansion
214: where, as pointed out by Huang \cite{Huang}, the limit $\phi \rightarrow 0$
215: must be taken after the thermodynamic limit $V\rightarrow +\infty $. The
216: purpose of introducing an auxiliary external field $\phi $ is to break the
217: gauge symmetry of the system, as can be easily seen from the term,%
218: \begin{equation}
219: \sum_{\mathbf{k}}(\phi c_{-\mathbf{k}\downarrow }c_{\mathbf{k}\uparrow
220: }+\phi ^{\dag }c_{\mathbf{k}\uparrow }^{\dag }c_{-\mathbf{k}\downarrow
221: }^{\dag }). \label{Broken}
222: \end{equation}%
223: Since the gauge symmetry has been broken with regard to the entire
224: Hamiltonian,
225: \begin{equation}
226: H(c)+\sum_{\mathbf{k}}(\phi c_{-\mathbf{k}\downarrow }c_{\mathbf{k}\uparrow
227: }+\phi ^{\dag }c_{\mathbf{k}\uparrow }^{\dag }c_{-\mathbf{k}\downarrow
228: }^{\dag }), \label{ExField}
229: \end{equation}%
230: in so far as the auxiliary external field $\phi $ is infinitesimal but
231: nonzero, the trace and the limit $V\rightarrow +\infty $ of Eq.(\ref{Pair})
232: become nonzero. Further, if the limit $\phi \rightarrow 0$ does not tend to
233: zero, the electron-pair amplitude $\left\langle c_{-\mathbf{k}\downarrow }c_{%
234: \mathbf{k}\uparrow }\right\rangle $ will get nonzero. One thus concludes
235: that there appear Cooper pairs in the system, and that the system has gone
236: into the superconducting phase. However, this conclusion can not hold in
237: physics, for the modification suffers the serious problem: The limit
238: procedure employed by Eq.(\ref{Pair}) is physically equivalent to the
239: scenario that an infinitesimal auxiliary external field is first added to
240: the system so as to induce the symmetry of the system to break down, and
241: taken off from the system finally. But, in the actual situation, the
242: inducement of such an external field is unnecessary, the symmetry itself
243: breaks down spontaneously and does not need any help of external force.
244: Anyhow, this kind of modification contains unphysical operations, and thus
245: can not be accepted in principle.
246:
247: Here, it is also significant and worth while to give a brief discussion to
248: the two-dimensional Ising model because it can be solved exactly within the
249: two modifications,%
250: \begin{equation}
251: H_{\mathrm{I}}=-J\sum_{\langle ij\rangle}s_{i}s_{j},
252: \end{equation}
253: where $s_{i}=\pm1$ denotes the spin variable on site $i$, the symbol $%
254: \langle ij\rangle$ means that the sites $i$ and $j$ \ are the nearest
255: neighbors, and $J>0$ is the exchange coupling between a nearest-neighbor
256: pair of spins.
257:
258: As is well known, Onsager \cite{Onsager} proved rigorously that the specific
259: heat of the Ising Hamiltonian $H_{\mathrm{I}}$ is singular at $T=T_{c}>0$
260: within the first modification of Gibbs ensemble theory. It hints that the
261: system might undergo a paramagnetic-ferromagnetic phase transition at $T_{c}$
262: in the thermodynamic limit. However, just as Eqs. (\ref{Zero1}--\ref{Zero2}%
263: ), the magnetization $m$,
264: \begin{equation}
265: m=\lim_{N\rightarrow \infty }\mathrm{Tr}\big(s_{i}\rho (H_{\mathrm{I}})\big)%
266: =0, \label{mIsing}
267: \end{equation}%
268: is always equal to zero due to the fact that the parity symmetry of $H_{%
269: \mathrm{I}}$ with respect to $s_{i}=1$ and $s_{i}=-1$ has nothing to do with
270: the magnitude of $N$ where $N$ represents the total number of sites. Eq. (%
271: \ref{mIsing}) proves definitely that there is no paramagnetic-ferromagnetic
272: phase transition at any temperature within the first modification of Gibbs
273: ensemble theory even if the specific heat is singular.
274:
275: The above proof also shows that it can not be solved within the framework of
276: the first modification whether or not the onset of the singularity at $%
277: T=T_{c}$ manifests a phase transition. In order to justify that the
278: phenomenon occurring at $T=T_{c}$ is a phase transition, Yang \cite{Yang4}
279: went beyond the first modification and turned to the second one, which was,
280: in fact, suggested originally by Ising himself \cite{Ising} in studying the
281: one-dimensional Ising model. Wonderfully, Yang succeeded in proving
282: rigorously that%
283: \begin{eqnarray}
284: m &=&\lim_{B\rightarrow 0}\lim_{N\rightarrow \infty }\mathrm{Tr}\Big(%
285: s_{i}\rho \big(H_{\mathrm{I}}-B\sum_{i}s_{i}\big)\Big) \notag \\
286: &=&\left\{
287: \begin{array}{ll}
288: \text{zero,} & T\geq T_{c} \\
289: \text{nonzero, } & T<T_{c},%
290: \end{array}%
291: \right. \label{mnzero}
292: \end{eqnarray}%
293: where $B$ stands for an external magnetic field. Mathematically, it seems
294: reasonable to say from Eq. (\ref{mnzero}) that a paramagnetic-ferromagnetic
295: phase transition will occur at $T=T_{c}$. Nevertheless, as pointed out
296: above, Eq. (\ref{mnzero}) imports unphysical operations from the
297: introduction of the external field $B$, and this leads to the result that
298: the phase transition will not occur automatically but has to be driven and
299: decided from outside the system, which directly contradicts the basic
300: experimental fact that any phase transition occurs itself spontaneously.
301: Therefore, the second modification, or rather Ising's criterion, can not be
302: used as a physical criterion to justify whether there exists a
303: paramagnetic-ferromagnetic phase transition at $T=T_{c}$.
304:
305: It should be stressed again that it is the two rigorous works of Refs. \cite%
306: {Onsager} and \cite{Yang4} that establish the two kinds of modification to
307: the Gibbs ensemble theory.
308:
309: On all accounts, the two kinds of modification must both be discarded, and
310: it is necessary to modify or extend Gibbs ensemble theory anew. That is just
311: the main purpose of this paper.
312:
313: Recently, Gibbs ensemble theory has been extended by C. Tsallis and his
314: followers with the conception of nonextensive entropy \cite%
315: {Tsallis1,Tsallis2,Tsallis3}, which is being under controversy. In this
316: paper, as a personal review, we would like to suppose another possible
317: extension of the Gibbs ensemble theory so that it can provide a reasonable
318: description to phase transitions and spontaneous symmetry breaking (SSB).
319:
320: \subsection{Landau Phenomenological Theory}
321:
322: Also, there is a macroscopic theory for phase transitions, i.e., Landau
323: phenomenological theory \cite{Landau}.
324:
325: Phenomenological as it is, Landau theory succeeds in providing a unified
326: picture for all the second-kind phase transitions, e.g., superconductivity,
327: superfluidity, magnetism, and structural phase transitions. Physically,
328: Landau picture consists of two basic notions: order parameter and
329: variational principle. Order parameter describes the degree of order of the
330: system, it is zero in the disordered phase, and nonzero in the ordered one.
331: Variational principle yields the equation of motion of order parameter and
332: controls the evolution of order parameter with temperature, a system must
333: arrive at the minimum of Helmholtz or Gibbs free energy in its stable state.
334: According to this picture, a system will evolve with the evolution of its
335: order parameters, and produce phase transitions spontaneously, i.e., without
336: any help or drive from outside the system. In short, Order parameter and
337: variational principle are two characteristics of Landau theory.
338:
339: On the other hand, Landau pointed out that a phase transition of the second
340: kind reflects physically the change in symmetry of the system: the
341: disordered phase has a higher symmetry than the ordered one. He found that
342: this change in symmetry can be described through the representations of the
343: symmetry group, a phase transition corresponds to a representation
344: transformation, different phases will realize different representations,
345: respectively. Accordingly, Landau established the relation between the order
346: parameter and the representation. Using this relation, it can be easily
347: determined from the variational principle which representation will be
348: realized at a certain temperature. Representation transformation and
349: spontaneous symmetry breaking are the other two characteristics of Landau
350: theory.
351:
352: To sum up, a system will select and realize automatically different
353: representations at different temperatures according to the variational
354: principle of order parameter. That is the mechanism for phase transitions
355: discovered by Landau theory, it crystallizes Landau's ideas about phase
356: transitions: order parameter, variational principle, representation
357: transformation, and spontaneous symmetry breaking. Theoretically, this
358: mechanism gives a reasonable answer to the problem why a phase transition or
359: SSB can happen at a certain temperature. In applications, it also agrees
360: quite well with various phase transitions, its effectiveness and
361: universality are well known to condensed matter physicists. In a word,
362: Landau's ideas on phase transitions are successful and of great value in
363: physics!
364:
365: Regrettably, it is impossible to deduce Landau theory from Gibbs ensemble
366: theory because there is no variational principle of order parameter within
367: the latter. This demonstrates again the deficiency of Gibbs ensemble theory.
368: In fact, this deficiency was pointed out by Born and Fucks \cite{Born} early
369: in 1938. In 1937, Mayer \cite{Mayer} published his famous paper on
370: gas-liquid transition, which is based directly on Gibbs ensemble theory.
371: This work was seriously doubted and questioned by Born and Fucks\ \cite{Born}%
372: : \textquotedblleft How can the gas molecules `know' when they have to
373: coagulate to form a liquid and solid?\textquotedblright\ Obviously, owing to
374: lack of Landau mechanism, it is hard for Gibbs ensemble theory to answer the
375: question posed by Born and Fucks.
376:
377: Gibbs ensemble theory should be extended to incorporate the Landau's ideas
378: so as to describe phase transitions and answer the question posed by Born
379: and Fucks. That is another intention of this paper.
380:
381: \section{The Extended Ensemble Theory \label{EET}}
382:
383: In order to describe phase transitions and spontaneous symmetry breaking, we
384: shall, as a personal review, extend Gibbs ensemble theory with three
385: hypotheses, which can be stated, taking the BCS superconductivity as an
386: instance, as follows.
387:
388: (1). The system Hamiltonian is represented by $H^{\prime}(\phi,c)$,
389: \begin{equation}
390: H^{\prime}(\phi,c)=e^{iD(\phi,c)}H(c)e^{-iD(\phi,c)}, \label{HPrime}
391: \end{equation}
392: where
393: \begin{equation}
394: D(\phi,c)=\sum_{\mathbf{k}}(\phi_{\mathbf{k}}c_{-\mathbf{k}\downarrow }c_{%
395: \mathbf{k}\uparrow}+\phi_{\mathbf{k}}^{\dag}c_{\mathbf{k}\uparrow}^{\dag
396: }c_{-\mathbf{k}\downarrow}^{\dag}). \label{Dfactor}
397: \end{equation}
398: Here $\phi$ is an internal field of the system, it is also the order
399: parameter for BCS superconductivity. Evidently, $H^{\prime}(\phi,c)$ is not
400: invariant under the gauge transformation of $G(\vartheta,c)$ as long as $%
401: \phi\neq0$, namely, the gauge symmetry will be broken for $%
402: H^{\prime}(\phi,c) $ if $\phi\neq0$. From now on, we shall call $D(\phi,c)$
403: the phase-transition operator, for sake of convenience.
404:
405: (2). The statistical average of an observable $F(c)$ is defined as
406: \begin{equation}
407: \mathcal{F}(\phi,\beta)=\langle F(c)\rangle=\mathrm{Tr}\big(F(c)\rho
408: (H^{\prime}(\phi,c))\big). \label{Average}
409: \end{equation}
410: We remark that the average is now a function of both temperature and order
411: parameter, which is expressed explicitly by the two arguments of $\mathcal{F}%
412: (\phi,\beta)$.
413:
414: (3). The order parameter $\phi $ is determined by the minimum of the entropy
415: of the system,
416: \begin{eqnarray}
417: \delta S &=&0, \label{Variation} \\
418: \Delta S &\geq &0, \label{Increment}
419: \end{eqnarray}%
420: where $S$ denotes the entropy,
421: \begin{eqnarray}
422: S(\phi ,\beta ) &=&\langle -\ln \!\left( \rho (H(c))\right) \rangle \notag
423: \\
424: &=&-\mathrm{Tr}\big(\!\ln \!\left( \rho (H(c))\right) \rho (H^{\prime }(\phi
425: ,c))\big). \label{Entropy}
426: \end{eqnarray}%
427: As a function of the order parameter, the entropy controls the evolution of
428: state of the system with temperature.
429:
430: The first hypothesis means that the system Hamiltonian can take different
431: representations at different temperatures, e.g., it can take the symmetric
432: representation ($\phi=0$) at a high temperature, and the asymmetric
433: representation ($\phi\neq0$) at a low temperature. Which representation it
434: will take is determined by the third hypothesis. After the representation is
435: so determined, the statistical average can be calculated with respect to
436: this representation, as stated in the second hypothesis.
437:
438: Obviously, the extended theory will reduce to the original one if $\phi =0$,
439: that is to say, the original theory holds only for the normal phase of the
440: system. The broken-symmetry phase will be described by the extended theory.
441:
442: Within the framework of the extended ensemble theory, a system will realize
443: different representations of the same system Hamiltonian at different
444: temperatures according to the principle of least entropy with respect to
445: order parameter, this mechanism for phase transitions is, in spirit, the
446: same as that given by Landau phenomenological theory; hence, it can be said
447: that Landau mechanism has been incorporated into the extended ensemble
448: theory. Also, because the order parameter is an internal field of the system
449: itself, symmetry breaking will occur in a completely spontaneous way rather
450: than forced by an external field, there is no unphysical operation within
451: the extended theory. In a word, the extended ensemble theory is
452: conceptionally in accordance with Landau's ideas on phase transitions: order
453: parameter, variational principle, representation transformation, and
454: spontaneous symmetry breaking.
455:
456: In Sec. \ref{POPT}, we shall show further that phase transitions originate
457: physically from the wave nature of matter.
458:
459: Now, let us establish the relationship between the extended ensemble theory
460: and thermodynamics, it can be implemented as follows.
461:
462: First, the internal energy $U$ can be obtained from Eq. (\ref{Average}),
463: replacing $F(c)$ by the Hamiltonian $H(c)$. It is a function of temperature $%
464: T$ and volume $V$, that is,
465: \begin{equation}
466: U=U(T,V).
467: \end{equation}
468: From the internal energy $U(T,V)$, the specific heat at constant volume, $%
469: C_{V}$, can be calculated through%
470: \begin{equation}
471: C_{V}=C_{V}(T,V)=\left( \frac{\partial U}{\partial T}\right) _{V}.
472: \end{equation}
473:
474: Then, we can obtain the thermodynamical entropy $S_{th}$,%
475: \begin{equation}
476: S_{th}=S_{th}(T,V)=\int_{0}^{T}C_{V}(T^{\prime},V)\frac{\mathrm{d}T^{\prime}%
477: }{T^{\prime}}.
478: \end{equation}
479: It is worth paying attention to the significant difference between the
480: thermodynamical entropy $S_{th}$ and the statistical entropy $S$ defining in
481: Eq. (\ref{Entropy}): $S$ measures the degree of order of the system, whereas
482: $S_{th}$ measures the amount of heat absorbed or rejected by the system,
483: i.e., $dQ=TdS_{th}$; besides, $S$ contains the basic information of the
484: system, it controls the evolution of state of the system, as a consequence, $%
485: S_{th}$ derives itself from $S$.
486:
487: At last, we arrive at%
488: \begin{equation}
489: F=F(T,V)=U(T,V)-TS_{th}(T,V),
490: \end{equation}%
491: where $F$ is the Helmholtz free energy of the system, with $T$ and $V$ as
492: its natural variables. As a thermodynamic potential of the system, $F$ can
493: generate other thermodynamic quantities. By such a way, the extended
494: ensemble theory connects with thermodynamics.
495:
496: In the end of this section, let us explain why we choose the BCS model
497: rather than the Ising model as the starting instance to develop the theory
498: for phase transitions and spontaneous symmetry breaking.
499:
500: As is well known, Ising model $H_{\mathrm{I}}$ is a classical discrete
501: Hamiltonian. Simple as it is mathematically, it is, however, ill defined
502: from the point of view of physics. On one hand, a classical Hamiltonian, $%
503: H=H(q,p)$ where $q$ and $p$ denote the generalized coordinates and momenta
504: of the system, must be continuous\ in phase space, that is to say, $H(q,p)$
505: is a continuous function of $q$ and $p$; there does not exist any classical
506: system which corresponds to a discrete Hamiltonian; therefore, $H_{\mathrm{I}%
507: }$ does not correspond to a classical system. On the other hand, the
508: discrete quantity $s_{i}=\pm 1$ is not a quantum operator but a pure integer
509: variable, therefore, $H_{\mathrm{I}}$ does not correspond to a quantum
510: system, either. In a word, Ising model can not correspond to any physical
511: system; its solution, whether rigorous or not, has no real physical
512: relevance \cite{Yang1}, which was, in fact, realized and pointed out as the
513: "Ising disease" by people early in 1940's and 1950's \cite{Yang1}.
514:
515: To understand that point further, let us first suppose that Ising model
516: could represent a real physical system, and then analyze the statistical
517: properties of this system in detail. First, if the system is not placed in
518: an external magnetic field, then, according to the Onsager's solution \cite%
519: {Ising,Huang}, its specific heat will diverge logarithmically at $T=T_{c}>0$%
520: ; however, as indicated by Eq. (\ref{mIsing}), there appears no spontaneous
521: magnetization bellow $T_{c}$, namely, the state of the system will remain
522: nonmagnetic at any temperature, whether $T>T_{c}$ or $T<T_{c}$. There is no
523: paramagnetic-ferromagnetic phase transition in the Ising system if it is not
524: placed in an external magnetic field, though its specific heat is singular.
525: That is the first characteristic behavior of the Ising system. Secondly, if
526: the system is placed in an external magnetic field, then, according to the
527: Yang's solution \cite{Yang4}, there will appear an induced magnetization
528: bellow $T_{c}$. That is to say, the Ising system will transform from a
529: paramagnetic state at $T>T_{c}$ into a ferromagnetic state at $T<T_{c}$ when
530: it is placed in an external magnetic field, here, the transformation is an
531: induced transformation other than a spontaneous phase transition. That is
532: the second characteristic behavior of the Ising system. In sum, an Ising
533: system will exhibit two distinct behaviors, respectively, depending on
534: whether the system is placed in an external magnetic field or not. It will
535: exhibit the first characteristic behavior in the case without the
536: application of any external magnetic field and the second one in the case
537: with the application of an external magnetic field. The two behaviors should
538: be checked experimentally with respect to the same physical system. However,
539: such a physical system that can simultaneously show both the two
540: characteristic behaviors of the Ising model has never been observed and
541: reported experimentally. In other words, there does not exist the supposed
542: Ising system in nature at all! That is not surprising, it just reflects the
543: ill definition of the Ising model. Here, it should be pointed out that the
544: solutions of Onsager and Yang are not identical but essentially different
545: from each other, whether in the sense of mathematics or in the sense of
546: physics: In the limit $B\rightarrow 0$, the Yang's solution can not reduce
547: to the Onsager's solution, as can be easily seen by comparing Eq. (\ref%
548: {mIsing}) and Eq. (\ref{mnzero}); they represent the statistical properties
549: of the Ising system with and without the interaction of an external magnetic
550: field, respectively.
551:
552: Needless to say, an Ising system is completely distinguishable from a real
553: physical system of phase transition. For the latter, the transition happens
554: itself spontaneously, needing no inducement of a corresponding external
555: field. Evidently, this nature of phase transitions is neither the same as
556: the first characteristic behavior of the Ising system, i.e., a singular
557: specific heat but no phase transition, nor the same as the second
558: characteristic behavior of the Ising system, i.e., a transformation induced
559: but not spontaneous. There are clear and definite discrepancies between the
560: statistical behavior of a real physical system of phase transition and those
561: of the Ising system. Any phase transition can not be described by Ising
562: model. Of course, if one disregards physically the ill definition of Ising
563: model, the distinction between the solutions of Onsager and Yang, and the
564: discrepancies between the statistical behavior of a real physical system of
565: phase transition and those of the Ising system, he can simulate some phase
566: transitions and critical phenomena quite well with Ising model \cite%
567: {Kadanoff1,Kadanoff2,Kadanoff3,Wilson1,Wilson2,Wilson3,Wilson4,Wilson5,Fisher}%
568: .
569:
570: Theoretically, Ising model is ill defined; experimentally, Ising system does
571: not exist in nature at all. That is the reason why we do not choose it as
572: the starting instance to develop the theory of phase transitions and SSB. As
573: for the BCS model \cite{BCS1,BCS2,BCS3}, quite the contrary, it is of high
574: value in theory because it contributes a new quantum concept, i.e., Cooper
575: pairs, which is profound and rather hard to understand theoretically but yet
576: verified exactly by the quantized flux experiments. Actually, it is the BCS
577: theory that began our microscopic understanding of phase transitions.
578:
579: Although Ising's criterion for phase transitions is problematic in physics,
580: it is much suggestive in mathematics, indeed. To some degree, it can be said
581: that the present modification is just equivalent to substituting the
582: external field proposed by Ising with an internal field determined by the
583: system itself, as can be seen from expanding the right-hand side of Eq. (\ref%
584: {HPrime}) to the linear term of $\phi $ and then comparing it with Eq. (\ref%
585: {ExField}).
586:
587: \section{Application to the Ideal Fermi Gas \label{IFG}}
588:
589: Let us first apply the extended ensemble theory to the ideal Fermi gas. It
590: is the simplest case, and we shall see that this case is exactly solvable.
591:
592: For the ideal Fermi gas, there is no interaction ($g=0$), Eq. (\ref{BCS})
593: reduces to
594: \begin{equation}
595: H(c)=\sum_{\mathbf{k}}\varepsilon (\mathbf{k})(c_{\mathbf{k}\uparrow
596: }^{\dagger }c_{\mathbf{k}\uparrow }+c_{-\mathbf{k}\downarrow }^{\dag }c_{-%
597: \mathbf{k}\downarrow }),
598: \end{equation}%
599: where $\varepsilon (\mathbf{k})=\hbar ^{2}\mathbf{k}^{2}/\left( 2m\right)
600: -\mu $ with $m$ and $\mu $ being the mass of the fermions and the chemical
601: potential of the system, respectively.
602:
603: To facilitate the calculation of the entropy, we reformulated Eq. (\ref%
604: {Entropy}) as
605: \begin{equation}
606: S(\phi ,\beta )=-\mathrm{Tr}\!\left( \ln \!\Big(\rho \big(H(e^{-iD(\phi
607: ,c)}ce^{iD(\phi ,c)})\big)\Big)\rho (H(c))\right) . \label{Sphi}
608: \end{equation}%
609: By use of Eq. (\ref{Dfactor}), we find
610: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}} }%
611: %BeginExpansion
612: \begin{widetext}
613: %EndExpansion
614: \begin{subequations}
615: \label{Ctrans}
616: \begin{eqnarray}
617: e^{-iD(\phi ,c)}c_{\mathbf{k}\uparrow }e^{iD(\phi ,c)} &=&\cos \left( \theta
618: _{\mathbf{k}}\right) c_{\mathbf{k}\uparrow }+i\sin \left( \theta _{\mathbf{k}%
619: }\right) e^{-i\phi _{\mathbf{k}}}c_{-\mathbf{k}\downarrow }^{\dagger } \\
620: e^{-iD(\phi ,c)}c_{-\mathbf{k}\downarrow }e^{iD(\phi ,c)} &=&\cos \left(
621: \theta _{\mathbf{k}}\right) c_{-\mathbf{k}\downarrow }-i\sin \left( \theta _{%
622: \mathbf{k}}\right) e^{-i\phi _{\mathbf{k}}}c_{\mathbf{k}\uparrow }^{\dagger
623: },
624: \end{eqnarray}%
625: where $\theta _{\mathbf{k}}=|\phi _{\mathbf{k}}|$ and $\varphi _{\mathbf{k}%
626: }=\arg (\phi _{\mathbf{k}})$. Substituting them into Eq. (\ref{Sphi}), we
627: have
628: \end{subequations}
629: \begin{equation}
630: S(\phi ,\beta )=\ln \!\big(\mathrm{Tr}(e^{-\beta H(c)})\big)+2\beta \sum_{%
631: \mathbf{k}}\varepsilon (\mathbf{k})\left[ \cos ^{2}(\theta _{\mathbf{k}%
632: })f(\varepsilon (\mathbf{k}))+\sin ^{2}(\theta _{\mathbf{k}})f(-\varepsilon (%
633: \mathbf{k}))\right] , \label{Sphi1}
634: \end{equation}%
635: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
636: %BeginExpansion
637: \end{widetext}
638: %EndExpansion
639: where%
640: \begin{equation}
641: f(\varepsilon )=\frac{1}{e^{\beta \varepsilon }+1}
642: \end{equation}%
643: is the Fermi distribution function.
644:
645: From Eqs. (\ref{Sphi1}) and (\ref{Variation}), it follows that
646: \begin{equation}
647: \sin(2\theta_{\mathbf{k}})=0, \label{IFGOrder}
648: \end{equation}
649: which is the equation of order parameter. Obviously, this equation has the
650: solutions,%
651: \begin{equation}
652: \theta_{\mathbf{k}}=0,\text{ }\pi/2\text{ },\pi,\text{ }3\pi/2.
653: \label{Theta}
654: \end{equation}
655:
656: In addition, one can easily deduce from Eq. (\ref{Sphi1}) that%
657: \begin{equation}
658: \Delta S=\sum_{\mathbf{k}}\Delta s_{\mathbf{k}},
659: \end{equation}%
660: where
661: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}}}%
662: %BeginExpansion
663: \begin{widetext}%
664: %EndExpansion
665: \begin{equation}
666: \Delta s_{\mathbf{k}}=\left\{
667: \begin{array}{rlll}
668: \beta \varepsilon (\mathbf{k})\tanh \!\left( \frac{\beta \varepsilon (%
669: \mathbf{k})}{2}\right) \left[ 1-\cos \left( 2\delta \theta _{\mathbf{k}%
670: }\right) \right] & \geq & 0,\text{ } & \theta _{\mathbf{k}}=0,\text{ or }\pi
671: \\
672: -\beta \varepsilon (\mathbf{k})\tanh \!\left( \frac{\beta \varepsilon (%
673: \mathbf{k})}{2}\right) \left[ 1-\cos \left( 2\delta \theta _{\mathbf{k}%
674: }\right) \right] & \leq & 0,\text{ } & \theta _{\mathbf{k}}=\pi /2,\text{ or
675: }3\pi /2,%
676: \end{array}%
677: \right.
678: \end{equation}%
679: with $\delta \theta _{\mathbf{k}}$ being the deviation of $\theta _{\mathbf{k%
680: }}$ from the corresponding solution.
681:
682: For the solution that all $\theta _{\mathbf{k}}=0$ or $\pi $, we have $%
683: \Delta S\geq 0$, which meets the requirement of Eq. (\ref{Increment}),
684: thereby, this solution corresponds to a stable phase. Observe that the order
685: parameter $\phi =0$ when all $\theta _{\mathbf{k}}=0$, one recognizes
686: immediately that this phase is just the normal phase of the system, which
687: can also be seen from the following equation,
688: \begin{equation}
689: \mathcal{F}(\phi ,\beta )=\mathrm{Tr}\big(F(c)\rho (H^{\prime }(\phi ,c))%
690: \big)=\mathrm{Tr}\big(F(c)\rho (H(c))\big),\text{ }\forall \theta _{\mathbf{k%
691: }}=0,\text{ or }\pi .
692: \end{equation}%
693: We are familiar with the properties of this normal phase from the standard
694: books on quantum statistical mechanics. What it is worth stressing here is
695: that the normal phase is proved to be stable at any temperature within the
696: framework of the extended ensemble theory. Of course, that is reasonable,
697: and just what we expect.
698:
699: As to the other solutions, they also have physical senses though they
700: correspond to instable phases because the entropies for them are not
701: minimal. Let us first discuss the phase with all $\theta _{\mathbf{k}}=\pi
702: /2 $ or $3\pi /2$. We find
703: \begin{equation}
704: \mathcal{F}(\phi ,\beta )=\mathrm{Tr}\big(F(c)\rho (H^{\prime }(\phi ,c))%
705: \big)=\mathrm{Tr}\!\left( F(c)\frac{e^{-\beta ^{\prime }H(c)}}{\mathrm{Tr}%
706: \!\left( e^{-\beta ^{\prime }H(c)}\right) }\right) ,\text{ }\forall \theta _{%
707: \mathbf{k}}=\frac{\pi }{2},\text{ or }\frac{3\pi }{2},
708: \end{equation}%
709: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
710: %BeginExpansion
711: \end{widetext}
712: %EndExpansion
713: where $\beta ^{\prime }=1/(k_{B}T^{\prime })$ with $T^{\prime }=-T<0$. That
714: is to say, the temperature of the system goes negative. This shows that
715: negative temperatures root from the same microscopic mechanism as that for
716: phase transitions, so we call this instable phase the negative-temperature
717: phase. In comparison, the normal phase is said to be the
718: positive-temperature phase ($T>0$). This negative-temperature phase can not
719: be realized physically because its internal energy is unbounded above.
720:
721: Apart from the negative-temperature phase, all the other instable phases are
722: partially negative-temperature phases, which means that the particles of the
723: system are distributed partially among the negative-temperature
724: (single-particle) states, and partially among the positive-temperature
725: (single-particle) states,
726: \begin{equation}
727: \langle c_{\mathbf{k}\sigma }^{\dag }c_{\mathbf{k}\sigma }\rangle =\left\{
728: \begin{array}{ll}
729: \frac{1}{e^{\beta \varepsilon (\mathbf{k})}+1},\text{ } & \theta _{\mathbf{k}%
730: }=0,\text{ or }\pi \\
731: \frac{1}{e^{\beta ^{\prime }\varepsilon (\mathbf{k})}+1},\text{ } & \theta _{%
732: \mathbf{k}}=\pi /2,\text{ or }3\pi /2.%
733: \end{array}%
734: \right.
735: \end{equation}%
736: That is to say, some of the states take $0$ or $\pi $, the others take $\pi
737: /2$ or $3\pi /2$. Each $\mathbf{k}$ with $\theta _{\mathbf{k}}=0$ or $\pi $
738: represents a positive-temperature state, and each $\mathbf{k}$ with $\theta
739: _{\mathbf{k}}=\pi /2$ or $3\pi /2$ represents a negative-temperature state.
740: Because the partially negative-temperature phase is an instable phase, it
741: can emit photons when particles transfer from a negative-temperature state
742: into a positive-temperature one. Therefore, a partially negative-temperature
743: system can constitute a laser.
744:
745: To make that point more specified, let us consider a semiconductor that is
746: described by the following Hamiltonian,
747: \begin{equation}
748: H(c)=\sum_{n,\mathbf{k}}\varepsilon _{n}(\mathbf{k})(c_{n,\mathbf{k}\uparrow
749: }^{\dag }c_{n,\mathbf{k}\uparrow }+c_{n,-\mathbf{k}\downarrow }^{\dag }c_{n,-%
750: \mathbf{k}\downarrow }),
751: \end{equation}%
752: where $n$ denotes the energy-band index, $\varepsilon _{n}(\mathbf{k})$ the
753: energy of the electrons in the $n$th band relative to the Fermi level of the
754: system, and $c_{n,\mathbf{k}\sigma }$ the annihilation operator of the
755: electrons with momentum $\mathbf{k}$ and spin $\sigma $ in the $n$th band.
756: Replacing the $D(\phi ,c)$ of Eq. (\ref{Dfactor}) with
757: \begin{equation}
758: D(\phi ,c)=\sum_{n,\mathbf{k}}(\phi _{n,\mathbf{k}}c_{n,-\mathbf{k}%
759: \downarrow }c_{n,\mathbf{k}\uparrow }+\phi _{n,\mathbf{k}}^{\dag }c_{n,%
760: \mathbf{k}\uparrow }^{\dag }c_{n,-\mathbf{k}\downarrow }^{\dag }),
761: \end{equation}%
762: and following the same procedure as for the ideal Fermi gas, one can easily
763: get
764: \begin{equation}
765: \langle c_{n,\mathbf{k}\sigma }^{\dag }c_{n,\mathbf{k}\sigma }\rangle
766: =\left\{
767: \begin{array}{ll}
768: \frac{1}{e^{\beta \varepsilon _{n}(\mathbf{k})}+1},\text{ } & \theta _{n,%
769: \mathbf{k}}=0,\text{ or }\pi \\
770: \frac{1}{e^{\beta ^{\prime }\varepsilon _{n}(\mathbf{k})}+1},\text{ } &
771: \theta _{n,\mathbf{k}}=\pi /2,\text{ or }3\pi /2,%
772: \end{array}%
773: \right.
774: \end{equation}%
775: where $\theta _{n,\mathbf{k}}=|\phi _{n,\mathbf{k}}|$.
776:
777: Suppose that there are, for example, two electrons per site. If the system
778: is at zero temperature and stays in the normal phase where all $\theta _{n,%
779: \mathbf{k}}=0$ or $\pi $, the valence band, i.e., the $0$th band, is fully
780: filled, the conduction band, i.e. the $1$st band, and all the other higher
781: bands are empty. This is the well-known picture for semiconductors (or
782: insulators), which is depicted in Fig. 1.
783: %TCIMACRO{%
784: %\TeXButton{Fig1}{\begin{figure}[htbp]
785: %\includegraphics[scale=0.35,angle=-90]{Fig1.eps}
786: %\caption
787: %{Energy bands for the normal (positive-temperature) phase of a semiconductor
788: %where $\mu$ denotes the chemical potential.}
789: %\end{figure}}}%
790: %BeginExpansion
791: \begin{figure}[htbp]
792: \includegraphics[scale=0.35,angle=-90]{Fig1.eps}
793: \caption
794: {Energy bands for the normal (positive-temperature) phase of a semiconductor
795: where $\mu$ denotes the chemical potential.}
796: \end{figure}%
797: %EndExpansion
798:
799: Now, let us consider such a partially negative-temperature phase,
800: \begin{equation}
801: \theta _{n,\mathbf{k}}=\left\{
802: \begin{array}{ll}
803: \pi /2,\text{ or }3\pi /2,\text{ } & n=1 \\
804: 0,\text{ or }\pi ,\text{ } & n\neq 1.%
805: \end{array}%
806: \right.
807: \end{equation}%
808: If this phase is at zero temperature, the conduction band is fully filled,
809: all the other bands are empty. It implies that the population is inverted
810: between the valence and conduction bands, which is depicted in Fig. 2.
811: %TCIMACRO{%
812: %\TeXButton{Fig2}{\begin{figure}[htbp]
813: %\includegraphics[scale=0.35,angle=-90]{Fig2.eps}
814: %\caption
815: %{Energy bands for a partially negative-temperature phase of a semiconductor
816: %where $\mu
817: %$ denotes the chemical potential. Here, the population is inverted between the
818: %valance and conduction bands.}
819: %\end{figure}} }%
820: %BeginExpansion
821: \begin{figure}[htbp]
822: \includegraphics[scale=0.35,angle=-90]{Fig2.eps}
823: \caption
824: {Energy bands for a partially negative-temperature phase of a semiconductor
825: where $\mu
826: $ denotes the chemical potential. Here, the population is inverted between the
827: valance and conduction bands.}
828: \end{figure}
829: %EndExpansion
830: As mentioned above, such a population-inversion phase is instable, it will
831: finally turn into the stable phase, i.e., the normal phase, with the
832: electrons in the conduction band hopping into the valence band and
833: simultaneously releasing energies to the surroundings. It will make a laser
834: if the energies are released by means of emitting photons. That is somewhat
835: an ideal semiconductor laser, it is a \textquotedblleft
836: two-level\textquotedblright\ system, a more realistic laser can be achieved
837: in the so-called \textquotedblleft three-level\textquotedblright\ system,%
838: \begin{equation}
839: \theta _{n,\mathbf{k}}=\left\{
840: \begin{array}{ll}
841: 0,\text{ or }\pi ,\text{ } & n=0 \\
842: \pi /2,\text{ or }3\pi /2,\text{ } & \exists \mathbf{k}\text{ for }n=2 \\
843: 0,\text{ or }\pi ,\text{ } & \text{otherwise,}%
844: \end{array}%
845: \right.
846: \end{equation}%
847: where the $0$th and $2$nd bands are partially filled, the $1$st band is
848: empty, which is depicted in Fig. 3. Here, the population is not inverted
849: between the $0$th and $2$nd bands, i.e., the population of the $2$nd band is
850: less than that of the $0$th band, but it is inverted between the $1$st and $%
851: 2 $nd bands. This kind of \textquotedblleft three-level\textquotedblright\
852: semiconductor laser has been realized in experiments \cite{Laud}. For this
853: reason, we call the partially negative-temperature phase the laser phase.
854: Obviously, a laser phase cannot be described by the original Gibbs ensemble
855: theory \cite{Gibbs}.
856: %TCIMACRO{%
857: %\TeXButton{Fig3}{\begin{figure}[htbp]
858: %\includegraphics[scale=0.35,angle=-90]{Fig3.eps}
859: %\caption{Energy bands for a laser phase of a semiconductor where $\mu
860: %$ denotes the
861: %chemical potential. Here, both the valance and 2th bands are partially filled, the
862: %population is inverted between the 2th and conduction bands.}
863: %\end{figure}}}%
864: %BeginExpansion
865: \begin{figure}[htbp]
866: \includegraphics[scale=0.35,angle=-90]{Fig3.eps}
867: \caption{Energy bands for a laser phase of a semiconductor where $\mu
868: $ denotes the
869: chemical potential. Here, both the valance and 2th bands are partially filled, the
870: population is inverted between the 2th and conduction bands.}
871: \end{figure}%
872: %EndExpansion
873:
874: Traditionally, negative temperatures are something obscure, and hard to
875: understand why they are higher or hotter than positive temperatures \cite%
876: {Purcell,Ramsey,Pathria}. From the viewpoint of the extended ensemble
877: theory, the reason is that both the negative-temperature and laser phases
878: are instable, they will trend, by giving up parts of their internal energy
879: to the surroundings, towards the stable phase, viz., the
880: positive-temperature phase. That is the microscopic interpretation about
881: negative temperatures.
882:
883: Finally, we note that the ideal Fermi gas cannot produce superconductivity
884: because there does not exist any Cooper pair in any case of the solutions of
885: Eq. (\ref{Theta}),
886: \begin{equation}
887: \left\langle c_{-\mathbf{k}\downarrow }c_{\mathbf{k}\uparrow }\right\rangle =%
888: \mathrm{Tr}\big(c_{-\mathbf{k}\downarrow }c_{\mathbf{k}\uparrow }\rho
889: (H^{\prime }(\phi ,c))\big)=0,
890: \end{equation}%
891: this result is just as expected, and right in accordance with the BCS theory
892: \cite{BCS1,BCS2,BCS3}.
893:
894: In summary, it is proved, within the framework of the extended ensemble
895: theory, that the ideal Fermi gas can not produce superconductivity, its
896: normal phase is stable at any temperature. Besides, the extended ensemble
897: theory gives a microscopic interpretation to the negative-temperature and
898: laser phases: they originate from the same microscopic mechanism as that for
899: phase transitions, and most importantly they are instable.
900:
901: \section{Application to BCS Superconductivity}
902:
903: \subsection{The BCS\ Mean-Field Theory}
904:
905: To study BCS superconductivity with the extended ensemble theory, one needs
906: first to solve Eq. (\ref{Variation}) to obtain the order parameter, and then
907: to prove that it can satisfy the requirement of Eq. (\ref{Increment}).
908: Unfortunately, that is rather difficult because it is impossible to
909: calculate $S(\phi ,\beta )$ rigorously when $H(c)$ contains an interaction.
910: As a result, we have to seek approximations. Before doing so, it is
911: worthwhile to give a brief survey to the self-consistent mean-field theory
912: due to BCS \cite{BCS2,BCS3}, which can be summarized as follows,
913: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}} }%
914: %BeginExpansion
915: \begin{widetext}
916: %EndExpansion
917: \begin{gather}
918: H(c)\Longrightarrow H_{\mathrm{MF}}(c)=\sum_{\mathbf{k}}\varepsilon (\mathbf{%
919: k})(c_{\mathbf{k}\uparrow }^{\dag }c_{\mathbf{k}\uparrow }+c_{-\mathbf{k}%
920: \downarrow }^{\dag }c_{-\mathbf{k}\downarrow })-\Delta \sum_{\mathbf{k}}(c_{-%
921: \mathbf{k}\downarrow }c_{\mathbf{k}\uparrow }+c_{\mathbf{k}\uparrow }^{\dag
922: }c_{-\mathbf{k}\downarrow }^{\dag })+\frac{1}{g}\Delta ^{2}, \label{BCSMF1}
923: \\
924: \left\langle F(c)\right\rangle =\mathrm{Tr}\big(F(c)\rho \left( H(c)\right) %
925: \big)\Longrightarrow \left\langle F(c)\right\rangle =\mathrm{Tr}\big(%
926: F(c)\rho \left( H_{\mathrm{MF}}(c)\right) \big), \label{BCSMF2}
927: \end{gather}%
928: where%
929: \begin{equation}
930: \Delta =g\sum_{\mathbf{k}}\left\langle c_{-\mathbf{k}\downarrow }c_{\mathbf{k%
931: }\uparrow }\right\rangle =g\sum_{\mathbf{k}}\mathrm{Tr}\big(c_{-\mathbf{k}%
932: \downarrow }c_{\mathbf{k}\uparrow }\rho \left( H_{\mathrm{MF}}(c)\right) %
933: \big). \label{BCSMF3}
934: \end{equation}%
935: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
936: %BeginExpansion
937: \end{widetext}
938: %EndExpansion
939: Eq. (\ref{BCSMF1}) and Eq. (\ref{BCSMF3}) constitute a pair of
940: self-consistent equations to determine both the mean-field Hamiltonian $H_{%
941: \mathrm{MF}}(c)$ and the energy gap $\Delta $. After $H_{\mathrm{MF}}(c)$ is
942: so determined, the average of observable $F(c)$ can be evaluated according
943: to Eq. (\ref{BCSMF2}).
944:
945: Obviously, if $\Delta \neq 0$, the mean-field Hamiltonian $H_{\mathrm{MF}%
946: }(c) $ is not invariant under the gauge transformation of Eq. (\ref{Gauge}),
947: i.e., the gauge symmetry is broken with respect to $H_{\mathrm{MF}}(c)$. It
948: should be noted that, as shown in Eqs. (\ref{BCSMF2}) and (\ref{BCSMF3}), $%
949: H_{\mathrm{MF}}(c)$ must be used in place of $H(c)$ when calculating
950: statistical averages. Otherwise, $\Delta $ becomes zero, there is no Cooper
951: pair and superconductivity, as has been demonstrated in Eq. (\ref{Zero}).
952: That is the key point of the BCS mean-field theory. This kind of mean-field
953: theory is also widely used in studying other phase transitions and SSB,
954: Bogoliubov theory of a weakly interacting Bose gas being a famous example
955: \cite{Bogoliubov,Zagrebnov}. Nevertheless, as pointed out by Emch in Sec.
956: 1.1.f of Ref. \cite{Emch}, it contains an inescapable paradox. That is, the
957: spectrum of the system Hamiltonian $H_{\mathrm{MF}}(c)$ depends on
958: temperature, which contradicts the fact that the spectrum of an operator is
959: an invariant property of this operator itself and should not depend on
960: temperature. Now, for the extended ensemble theory, the normal and
961: superconducting phases correspond, as stated in Sec. \ref{EET}, to the
962: symmetric and asymmetric representations of the same system Hamiltonian,
963: respectively, it excludes such paradox absolutely.
964:
965: Although the approximation of Eqs. (\ref{BCSMF1}--\ref{BCSMF3}) contains the
966: paradox, its results are astonishingly in good agreement with the
967: experiments. Therefore, we need to search a formally identical but
968: essentially different approximation within the extended ensemble theory so
969: as to interpret the superconductivity anew.
970:
971: In language of Green's function (GF), the above approximation can be
972: translated into the following formalism,
973: \begin{gather}
974: \langle \langle c_{\mathbf{k}\uparrow }|c_{\mathbf{k}\uparrow }^{\dag
975: }\rangle \rangle _{\omega }^{\mathrm{MF}}=\frac{\omega +\varepsilon (\mathbf{%
976: k})}{\omega ^{2}-\left[ \varepsilon ^{2}(\mathbf{k})+\Delta ^{2}\right] },
977: \label{GMF1} \\
978: \langle \langle c_{\mathbf{k}\uparrow }|c_{-\mathbf{k}\downarrow }\rangle
979: \rangle _{\omega }^{\mathrm{MF}}=\frac{\Delta }{\omega ^{2}-\left[
980: \varepsilon ^{2}(\mathbf{k})+\Delta ^{2}\right] }, \label{GMF2} \\
981: \Delta =-g\sum_{\mathbf{k}}\int_{-\infty }^{+\infty }\frac{\mathrm{d}\omega
982: }{\pi }f(\omega )\mathrm{Im}\langle \langle c_{\mathbf{k}\uparrow }|c_{-%
983: \mathbf{k}\downarrow }\rangle \rangle _{\omega }^{\mathrm{MF}}, \label{GMF3}
984: \end{gather}%
985: where $\langle \langle A|B\rangle \rangle _{\omega }^{\mathrm{MF}}$ denotes
986: the retarded Green's function defined with respect to $H_{\mathrm{MF}}(c)$.
987: Now, Eqs. (\ref{GMF2}) and (\ref{GMF3}) constitute a pair of self-consistent
988: equations to determine the two complex-valued functions, $\langle \langle c_{%
989: \mathbf{k}\uparrow }|c_{-\mathbf{k}\downarrow }\rangle \rangle _{\omega }^{%
990: \mathrm{MF}}$ and $\Delta $, in contrast to Eqs. (\ref{BCSMF1}) and (\ref%
991: {BCSMF3}) where an operator is involved. Therefore, Eqs. (\ref{GMF1}--\ref%
992: {GMF3}) are formally adaptable to transplanting into the extended ensemble
993: theory. As is well known, BCS superconductivity can be explained in terms of
994: these Green's functions \cite{BCS2,BCS3}. We shall reestablish them within
995: the framework of the extended ensemble theory. Of course, they must be
996: redefined, with respect to $H^{\prime }(\phi ,c)$ instead of $H_{\mathrm{MF}%
997: }(c)$.
998:
999: \subsection{Interpretation of BCS Superconductivity}
1000:
1001: Now, we devote ourselves to reestablishing a formalism similar to Eqs. (\ref%
1002: {GMF1}--\ref{GMF3}), and interpreting the superconductivity with the
1003: extended ensemble theory.
1004:
1005: Following Eq. (\ref{Sphi}), the entropy of a BCS superconductor can be
1006: written as%
1007: \begin{align}
1008: S(\phi ,\beta )& =\ln \!\Big(\mathrm{Tr}\big(e^{-\beta H(c)}\big)\Big)
1009: \notag \\
1010: & +\beta \sum_{\mathbf{k}}\varepsilon (\mathbf{k})(\overline{d_{\mathbf{k}%
1011: \uparrow }^{\dag }d_{\mathbf{k}\uparrow }}+\overline{d_{-\mathbf{k}%
1012: \downarrow }^{\dag }d_{-\mathbf{k}\downarrow }}) \notag \\
1013: & -\beta g\sum_{\mathbf{k}}\sum_{\mathbf{k}^{\prime }}\overline{d_{\mathbf{k}%
1014: ^{\prime }\uparrow }^{\dag }d_{-\mathbf{k}^{\prime }\downarrow }^{\dag }d_{-%
1015: \mathbf{k}\downarrow }d_{\mathbf{k}\uparrow }}\,, \label{BCSEtr}
1016: \end{align}%
1017: where
1018: \begin{subequations}
1019: \label{dOpr}
1020: \begin{eqnarray}
1021: d_{\mathbf{k\uparrow }} &=&e^{-iD(\phi ,c)}c_{\mathbf{k\uparrow }}e^{iD(\phi
1022: ,c)} \\
1023: d_{-\mathbf{k\downarrow }} &=&e^{-iD(\phi ,c)}c_{-\mathbf{k\downarrow }%
1024: }e^{iD(\phi ,c)},
1025: \end{eqnarray}%
1026: and
1027: \end{subequations}
1028: \begin{equation}
1029: \overline{A(c)}\equiv \mathrm{Tr}\big(A(c)\rho \left( H(c)\right) \big).
1030: \end{equation}%
1031: As the usual mean-field approximation, we decouple the last term on the
1032: right-hand side of Eq. (\ref{BCSEtr}) as follows,
1033: \begin{equation}
1034: \overline{d_{\mathbf{k}^{\prime }\uparrow }^{\dag }d_{-\mathbf{k}^{\prime
1035: }\downarrow }^{\dag }\times d_{-\mathbf{k}\downarrow }d_{\mathbf{k}\uparrow }%
1036: }=\overline{d_{\mathbf{k}^{\prime }\uparrow }^{\dag }d_{-\mathbf{k}^{\prime
1037: }\downarrow }^{\dag }}\times \overline{d_{-\mathbf{k}\downarrow }d_{\mathbf{k%
1038: }\uparrow }}\text{\thinspace }. \label{BCSFact}
1039: \end{equation}%
1040: It results in
1041: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}}}%
1042: %BeginExpansion
1043: \begin{widetext}%
1044: %EndExpansion
1045: \begin{align}
1046: S(\phi ,\beta )& =\ln \!\Big(\mathrm{Tr}\big(e^{-\beta H(c)}\big)\Big)+\beta
1047: \sum_{\mathbf{k}}\varepsilon (\mathbf{k})\left[ \cos ^{2}(\theta _{\mathbf{k}%
1048: })(\overline{c_{\mathbf{k}\uparrow }^{\dag }c_{\mathbf{k}\uparrow }}+%
1049: \overline{c_{-\mathbf{k}\downarrow }^{\dag }c_{-\mathbf{k}\downarrow }}%
1050: )+\sin ^{2}(\theta _{\mathbf{k}})(\overline{c_{\mathbf{k}\uparrow }c_{%
1051: \mathbf{k}\uparrow }^{\dag }}+\overline{c_{-\mathbf{k}\downarrow }c_{-%
1052: \mathbf{k}\downarrow }^{\dag }})\right] \notag \\
1053: & -\frac{1}{4}\beta g\sum_{\mathbf{k}^{\prime }}\sin (2\theta _{\mathbf{k}%
1054: ^{\prime }})e^{i\varphi _{\mathbf{k}^{\prime }}}(\overline{c_{\mathbf{k}%
1055: ^{\prime }\uparrow }^{\dag }c_{\mathbf{k}^{\prime }\uparrow }}-\overline{c_{-%
1056: \mathbf{k}^{\prime }\downarrow }c_{-\mathbf{k}^{\prime }\downarrow }^{\dag }}%
1057: )\times \sum_{\mathbf{k}}\sin (2\theta _{\mathbf{k}})e^{-i\varphi _{\mathbf{k%
1058: }}}(\overline{c_{\mathbf{k}\uparrow }^{\dag }c_{\mathbf{k}\uparrow }}-%
1059: \overline{c_{-\mathbf{k}\downarrow }c_{-\mathbf{k}\downarrow }^{\dag }})%
1060: \text{{\Large ,}} \label{SupS}
1061: \end{align}%
1062: where Eqs. (\ref{dOpr}), (\ref{Ctrans}), and (\ref{Zero}) have been used.
1063:
1064: Substitution of Eq. (\ref{SupS}) into Eq. (\ref{Variation}) gives the
1065: equations of order parameter,%
1066: \begin{gather}
1067: 2\varepsilon (\mathbf{k})\sin (2\theta _{\mathbf{k}})+g\cos (2\theta _{%
1068: \mathbf{k}})\sum_{\mathbf{k}^{\prime }}\cos (\varphi _{\mathbf{k}}-\varphi _{%
1069: \mathbf{k}^{\prime }})\sin (2\theta _{\mathbf{k}^{\prime }})(\overline{c_{%
1070: \mathbf{k}^{\prime }\uparrow }^{\dag }c_{\mathbf{k}^{\prime }\uparrow }}-%
1071: \overline{c_{-\mathbf{k}^{\prime }\downarrow }c_{-\mathbf{k}^{\prime
1072: }\downarrow }^{\dag }})=0, \\
1073: \sin (2\theta _{\mathbf{k}})\sum_{\mathbf{k}^{\prime }}\sin (\varphi _{%
1074: \mathbf{k}}-\varphi _{\mathbf{k}^{\prime }})\sin (2\theta _{\mathbf{k}%
1075: ^{\prime }})(\overline{c_{\mathbf{k}^{\prime }\uparrow }^{\dag }c_{\mathbf{k}%
1076: ^{\prime }\uparrow }}-\overline{c_{-\mathbf{k}^{\prime }\downarrow }c_{-%
1077: \mathbf{k}^{\prime }\downarrow }^{\dag }})=0.
1078: \end{gather}%
1079: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
1080: %BeginExpansion
1081: \end{widetext}
1082: %EndExpansion
1083: They have a trivial solution,%
1084: \begin{equation}
1085: \phi =0, \label{STrv}
1086: \end{equation}%
1087: and a nontrivial solution,%
1088: \begin{equation}
1089: \left\{
1090: \begin{array}{l}
1091: \varphi _{\mathbf{k}}=\varphi _{0}=\mathrm{constant} \\
1092: \varepsilon (\mathbf{k})\sin (2\theta _{\mathbf{k}})-\Lambda \cos (2\theta _{%
1093: \mathbf{k}})=0,%
1094: \end{array}%
1095: \right. \label{SNTrv}
1096: \end{equation}%
1097: where%
1098: \begin{equation}
1099: \Lambda =-ie^{i\varphi _{0}}g\sum_{\mathbf{k}}\left\langle c_{-\mathbf{k}%
1100: \downarrow }c_{\mathbf{k}\uparrow }\right\rangle .
1101: \end{equation}%
1102: The trivial solution always remains zero, it does not change with
1103: temperature. In contrast, the nontrivial one will depend on temperature, it
1104: may be zero for some temperatures, and nonzero for the others. Without loss
1105: of generality, we shall take $\varphi _{0}=\pi /2$, the nontrivial solution
1106: is then simplified as%
1107: \begin{gather}
1108: \varepsilon (\mathbf{k})\sin (2\theta _{\mathbf{k}})-\Lambda \cos (2\theta _{%
1109: \mathbf{k}})=0, \label{SupTheta} \\
1110: \Lambda =g\sum_{\mathbf{k}}\left\langle c_{-\mathbf{k}\downarrow }c_{\mathbf{%
1111: k}\uparrow }\right\rangle , \label{SupLambda}
1112: \end{gather}%
1113: they are a pair of coupled equations to determine both $\theta _{\mathbf{k}}$
1114: and $\Lambda $.
1115:
1116: If $g=0$, Eq. (\ref{SupLambda}) shows that $\Lambda=0$, then Eq. (\ref%
1117: {SupTheta}) comes back to the equation (\ref{IFGOrder}) for the ideal Fermi
1118: gas. Physically, the BCS\ Hamiltonian of Eq. (\ref{BCS}) can describe only
1119: the weak-coupling superconductivity. For the weak-coupling
1120: superconductivity, the coupling strength $g$ is quite small, from Eqs. (\ref%
1121: {SupTheta}) and (\ref{SupLambda}) it follows that $\Lambda$ and $\theta_{%
1122: \mathbf{k}}$ are also small. This makes it easy for us to do further
1123: approximations, as will be seen from below.
1124:
1125: In order to solve Eqs. (\ref{SupTheta}) and (\ref{SupLambda}), one needs $%
1126: \left\langle c_{-\mathbf{k}\downarrow }c_{\mathbf{k}\uparrow }\right\rangle $%
1127: , it can be obtained by the corresponding retarded Green's function $\langle
1128: \langle c_{\mathbf{k}\uparrow }|c_{-\mathbf{k}\downarrow }\rangle \rangle
1129: _{\omega }$ defined with respect to $H^{\prime }(\phi ,c)$ \cite{Zubarev}.
1130: As is well known, $\langle \langle c_{\mathbf{k}\uparrow }|c_{-\mathbf{k}%
1131: \downarrow }\rangle \rangle _{\omega }$ satisfies the equation of motion,%
1132: \begin{align}
1133: \omega \langle \langle c_{\mathbf{k}\uparrow }|c_{-\mathbf{k}\downarrow
1134: }\rangle \rangle _{\omega }& =\left\langle \left\{ c_{\mathbf{k}\uparrow
1135: },c_{-\mathbf{k}\downarrow }\right\} \right\rangle \notag \\
1136: & -\langle \langle c_{\mathbf{k}\uparrow }|\left[ c_{-\mathbf{k}\downarrow
1137: },H^{\prime }(\phi ,c)\right] \rangle \rangle _{\omega }, \label{GFMotion}
1138: \end{align}%
1139: where $\left\{ A,B\right\} $ denotes the anticommutator of $A$ and $B$.
1140: Expanding $H^{\prime }(\phi ,c)$ into the power series of $\phi $,
1141: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}} }%
1142: %BeginExpansion
1143: \begin{widetext}
1144: %EndExpansion
1145: \begin{equation}
1146: H^{\prime }(\phi ,c)=H(c)+i\left[ D(\phi ,c),H(c)\right] +\frac{i^{2}}{2!}%
1147: \left[ D(\phi ,c),\left[ D(\phi ,c),H(c)\right] \right] +\cdots ,
1148: \end{equation}%
1149: and substituting it into Eq. (\ref{GFMotion}), we have%
1150: \begin{align}
1151: \omega \langle \langle c_{\mathbf{k}\uparrow }|c_{-\mathbf{k}\downarrow
1152: }\rangle \rangle _{\omega }& =\left\langle \left\{ c_{\mathbf{k}\uparrow
1153: },c_{-\mathbf{k}\downarrow }\right\} \right\rangle -\langle \langle c_{%
1154: \mathbf{k}\uparrow }|\left[ c_{-\mathbf{k}\downarrow },H(c)\right] \rangle
1155: \rangle _{\omega }-i\langle \langle c_{\mathbf{k}\uparrow }|\left[ c_{-%
1156: \mathbf{k}\downarrow },\left[ D(\phi ,c),H(c)\right] \right] \rangle \rangle
1157: _{\omega } \notag \\
1158: & -\frac{i^{2}}{2!}\langle \langle c_{\mathbf{k}\uparrow }|\left[ c_{-%
1159: \mathbf{k}\downarrow },\left[ D(\phi ,c),[D(\phi ,c),H(c)]\right] \right]
1160: \rangle \rangle _{\omega }+\cdots .
1161: \end{align}%
1162: As stated above, $\phi $ is small for the weak-coupling superconductivity,
1163: it is thus rational to maintain only the zeroth-order term,%
1164: \begin{equation}
1165: \omega \langle \langle c_{\mathbf{k}\uparrow }|c_{-\mathbf{k}\downarrow
1166: }\rangle \rangle _{\omega }=-\varepsilon (\mathbf{k})\langle \langle c_{%
1167: \mathbf{k}\uparrow }|c_{-\mathbf{k}\downarrow }\rangle \rangle _{\omega
1168: }+g\langle \langle c_{\mathbf{k}\uparrow }|c_{\mathbf{k}\uparrow }^{\dag
1169: }\sum_{\mathbf{k}^{\prime }}c_{-\mathbf{k}^{\prime }\downarrow }c_{\mathbf{k}%
1170: ^{\prime }\uparrow }\rangle \rangle _{\omega }.
1171: \end{equation}%
1172: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
1173: %BeginExpansion
1174: \end{widetext}
1175: %EndExpansion
1176: With regard to the second GF on the right-hand side, we do the factorization
1177: as in Eq. (\ref{BCSFact}),
1178: \begin{equation}
1179: g\sum_{\mathbf{k}^{\prime }}\langle c_{-\mathbf{k}^{\prime }\downarrow }c_{%
1180: \mathbf{k}^{\prime }\uparrow }\rangle =g\sum_{\mathbf{k}^{\prime }}\overline{%
1181: d_{-\mathbf{k}^{\prime }\downarrow }d_{\mathbf{k}^{\prime }\uparrow }}%
1182: =\Lambda ,
1183: \end{equation}%
1184: this leads to%
1185: \begin{equation}
1186: \left[ \omega +\varepsilon (\mathbf{k})\right] \langle \langle c_{\mathbf{k}%
1187: \uparrow }|c_{-\mathbf{k}\downarrow }\rangle \rangle _{\omega }=\Lambda
1188: \langle \langle c_{\mathbf{k}\uparrow }|c_{\mathbf{k}\uparrow }^{\dag
1189: }\rangle \rangle _{\omega }. \label{WGFMotion}
1190: \end{equation}%
1191: The GF on the right-hand side can be obtained by the same procedure as for $%
1192: \langle \langle c_{\mathbf{k}\uparrow }|c_{-\mathbf{k}\downarrow }\rangle
1193: \rangle _{\omega }$,%
1194: \begin{equation}
1195: \left[ \omega -\varepsilon (\mathbf{k})\right] \langle \langle c_{\mathbf{k}%
1196: \uparrow }|c_{\mathbf{k}\uparrow }^{\dag }\rangle \rangle _{\omega
1197: }=1+\Lambda ^{\dagger }\langle \langle c_{\mathbf{k}\uparrow }|c_{-\mathbf{k}%
1198: \downarrow }\rangle \rangle _{\omega },
1199: \end{equation}%
1200: where
1201: \begin{equation}
1202: \Lambda ^{\dagger }=g\sum_{\mathbf{k}}\langle c_{\mathbf{k}\uparrow
1203: }^{\dagger }c_{-\mathbf{k}\downarrow }^{\dagger }\rangle =\Lambda .
1204: \label{DLambda}
1205: \end{equation}%
1206: From Eqs. (\ref{WGFMotion}--\ref{DLambda}), it follows that%
1207: \begin{gather}
1208: \langle \langle c_{\mathbf{k}\uparrow }|c_{\mathbf{k}\uparrow }^{\dag
1209: }\rangle \rangle _{\omega }=\frac{\omega +\varepsilon (\mathbf{k})}{\omega
1210: ^{2}-\left[ \varepsilon ^{2}(\mathbf{k})+\Lambda ^{2}\right] },
1211: \label{WGFMotion1} \\
1212: \langle \langle c_{\mathbf{k}\uparrow }|c_{-\mathbf{k}\downarrow }\rangle
1213: \rangle _{\omega }=\frac{\Lambda }{\omega ^{2}-\left[ \varepsilon ^{2}(%
1214: \mathbf{k})+\Lambda ^{2}\right] }, \label{WGFMotion2} \\
1215: \Lambda =-g\sum_{\mathbf{k}}\int_{-\infty }^{+\infty }\frac{\mathrm{d}\omega
1216: }{\pi }f(\omega )\text{\textrm{Im}}\langle \langle c_{\mathbf{k}\uparrow
1217: }|c_{-\mathbf{k}\downarrow }\rangle \rangle _{\omega }. \label{DLambda1}
1218: \end{gather}%
1219: They form a self-consistent mean-field solution to the Green's functions
1220: responsible for BCS superconductivity within the framework of the extended
1221: ensemble theory. Comparing them with Eqs. (\ref{GMF1}--\ref{GMF3}), one sees
1222: that the present solution is identical to the BCS solution except that the
1223: statistical average is now defined with respect to $H^{\prime }(\phi ,c)$.
1224: Thus far, the BCS mean-field results have been transplanted into the
1225: extended ensemble theory, the paradox removed.
1226:
1227: With the help of Eq. (\ref{WGFMotion2}), Eqs. (\ref{DLambda1}) and (\ref%
1228: {SupTheta}) can be simplified as%
1229: \begin{eqnarray}
1230: \Lambda &=&\frac{1}{2}g\Lambda \sum_{\mathbf{k}}\frac{\tanh \left( \frac{1}{2%
1231: }\beta \xi _{\mathbf{k}}\right) }{\xi _{\mathbf{k}}}, \label{GapEq} \\
1232: \theta _{\mathbf{k}} &=&\left\{
1233: \begin{array}{ll}
1234: \frac{1}{2}\arcsin \left( \frac{\Lambda }{\xi _{\mathbf{k}}}\right) ,\text{ }
1235: & \varepsilon (\mathbf{k})\geq 0 \\
1236: -\frac{1}{2}\arcsin \left( \frac{\Lambda }{\xi _{\mathbf{k}}}\right) ,\text{
1237: } & \varepsilon (\mathbf{k})<0,%
1238: \end{array}%
1239: \right. \label{OrderEq}
1240: \end{eqnarray}%
1241: where%
1242: \begin{equation}
1243: \xi _{\mathbf{k}}=\sqrt{\varepsilon ^{2}(\mathbf{k})+\Lambda ^{2}}.
1244: \end{equation}%
1245: Eq. (\ref{OrderEq}) together with Eq. (\ref{GapEq}) gives the mean-field
1246: solution to $\phi $ ($\phi _{\mathbf{k}}=i\theta _{\mathbf{k}}$), the order
1247: parameter for BCS superconductivity.
1248:
1249: Eq. (\ref{GapEq}) is familiar in the BCS mean-field theory \cite{BCS2,BCS3},
1250: it has the nontrivial solution,
1251: \begin{equation}
1252: \Lambda =\left\{
1253: \begin{array}{ll}
1254: \mathrm{zero,}\text{ } & T\geq T_{c} \\
1255: \text{\textrm{nonzero}}\mathrm{,}\text{ } & T<T_{c},%
1256: \end{array}%
1257: \right. \label{PseudoOrder}
1258: \end{equation}%
1259: where $T_{c}$ is determined by the equation,
1260: \begin{equation}
1261: 1=\frac{1}{2}g\sum_{\mathbf{k}}\frac{\tanh \left( \frac{1}{2}\beta
1262: _{c}\varepsilon (\mathbf{k})\right) }{\varepsilon (\mathbf{k})},
1263: \end{equation}%
1264: with $\beta _{c}=1/k_{B}T_{c}$. From Eqs. (\ref{PseudoOrder}) and (\ref%
1265: {OrderEq}), we find
1266: \begin{equation}
1267: \phi _{\mathbf{k}}=\left\{
1268: \begin{array}{ll}
1269: \mathrm{zero,}\text{ } & T\geq T_{c} \\
1270: \text{\textrm{nonzero}}\mathrm{,}\text{ } & T<T_{c}.%
1271: \end{array}%
1272: \right. \label{RealOrder}
1273: \end{equation}%
1274: Since $\phi $ is the order parameter of the system, Eq. (\ref{RealOrder})
1275: indicates that $T_{c}$ is the transition temperature for BCS
1276: superconductivity.
1277:
1278: Compared with Eq. (\ref{RealOrder}), Eq. (\ref{PseudoOrder}) shows that the
1279: energy gap $\Lambda $ behaviors with temperature just like an order
1280: parameter. However, this behavior of the energy gap is not a property of
1281: itself, but derives from the property of the order parameter, as can be seen
1282: from the equation,%
1283: \begin{equation}
1284: \Lambda =g\sum_{\mathbf{k}}\mathrm{Tr}\big(c_{-\mathbf{k}\downarrow }c_{%
1285: \mathbf{k}\uparrow }\rho \left( H^{\prime }(\phi ,c)\right) \big)=\left\{
1286: \begin{array}{ll}
1287: \mathrm{zero,}\text{ } & \phi =0 \\
1288: \text{\textrm{nonzero}}\mathrm{,}\text{ } & \phi \neq 0,%
1289: \end{array}%
1290: \right.
1291: \end{equation}%
1292: which demonstrates that the forming of Cooper pairs is a direct consequence
1293: of the gauge-symmetry breaking caused by the internal spontaneous field $%
1294: \phi $. In the extended ensemble theory, the order parameter is a physical
1295: quantity more fundamental than the energy gap.\ This picture is
1296: significantly distinct from the BCS mean-field theory where the energy gap
1297: is the most fundamental quantity that is responsible for the
1298: superconductivity.
1299:
1300: In sum, Eqs. (\ref{WGFMotion1}), (\ref{WGFMotion2}), (\ref{GapEq}), and (\ref%
1301: {OrderEq}) constitute a self-consistent mean-field theory for BCS
1302: superconductivity within the framework of the extended ensemble theory, its
1303: nontrivial solution reproduces the BCS mean-field results on the
1304: conventional low-$T_{c}$ superconductivity.
1305:
1306: We are now confronted with the important task to prove the stability of this
1307: nontrivial solution. By expanding the $S(\phi ,\beta )$ of Eq. (\ref{SupS})
1308: into a Taylor or Volterra series in the neighborhood of the nontrivial
1309: solution $\phi _{0}$, one finds
1310: \begin{equation}
1311: \Delta S=S(\phi ,\beta )-S(\phi _{0},\beta )=\delta ^{2}S+\delta
1312: ^{3}S+\delta ^{4}S+\cdots , \label{DeltaSBCS}
1313: \end{equation}%
1314: where $\delta ^{n}S$ represents the $n$th power term of the expansion of $%
1315: S(\phi ,\beta )$, or the $n$th variation of $S$. Specifically, the second
1316: variation $\delta ^{2}S$ has the form,
1317: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}} }%
1318: %BeginExpansion
1319: \begin{widetext}
1320: %EndExpansion
1321: \begin{align}
1322: \delta ^{2}S& =\frac{1}{8}\beta g\Lambda ^{2}\sum_{\mathbf{k}}\sum_{\mathbf{k%
1323: }^{\prime }}\frac{\tanh (\frac{1}{2}\beta \xi _{\mathbf{k}})}{\xi _{\mathbf{k%
1324: }}}\frac{\tanh (\frac{1}{2}\beta \xi _{\mathbf{k}^{\prime }})}{\xi _{\mathbf{%
1325: k}^{\prime }}}\left( \delta \varphi _{\mathbf{k}}-\delta \varphi _{\mathbf{k}%
1326: ^{\prime }}\right) ^{2}+2\beta \sum_{\mathbf{k}_{1}}\xi _{\mathbf{k}%
1327: _{1}}\tanh (\frac{1}{2}\beta \xi _{\mathbf{k}_{1}})\left( \delta \theta _{%
1328: \mathbf{k}_{1}}\right) ^{2} \notag \\
1329: & +\beta \left\{ \sum_{\mathbf{k}_{2}}2\xi _{\mathbf{k}_{2}}\tanh (\frac{1}{2%
1330: }\beta \xi _{\mathbf{k}_{2}})\left( \delta \theta _{\mathbf{k}_{2}}\right)
1331: ^{2}-g\sum_{\mathbf{k}_{2}}\sum_{\mathbf{k}_{2}^{\prime }}\varepsilon (%
1332: \mathbf{k}_{2})\frac{\tanh (\frac{1}{2}\beta \xi _{\mathbf{k}_{2}})}{\xi _{%
1333: \mathbf{k}_{2}}}\varepsilon (\mathbf{k}_{2}^{\prime })\frac{\tanh (\frac{1}{2%
1334: }\beta \xi _{\mathbf{k}_{2}^{\prime }})}{\xi _{\mathbf{k}_{2}^{\prime }}}%
1335: \delta \theta _{\mathbf{k}_{2}}\delta \theta _{\mathbf{k}_{2}^{\prime
1336: }}\right\} , \label{DeltaS2}
1337: \end{align}%
1338: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
1339: %BeginExpansion
1340: \end{widetext}
1341: %EndExpansion
1342: where $\delta \varphi _{\mathbf{k}}$ and $\delta \theta _{\mathbf{k}}$
1343: represent the variations of $\varphi _{\mathbf{k}}$ and $\theta _{\mathbf{k}%
1344: } $ from the nontrivial solution $\phi _{0}$, respectively. Here, the set of
1345: $\mathbf{k}$ has been separated into two subsets: the subset of $\mathbf{k}%
1346: _{1}$ and the subset of $\mathbf{k}_{2}$, where $\mathbf{k}_{1}$ and $%
1347: \mathbf{k}_{2}$ satisfy $\varepsilon (\mathbf{k}_{1})=0$ and $\varepsilon (%
1348: \mathbf{k}_{2})\neq 0$, respectively. All the contributions from the subset
1349: of $\mathbf{k}_{2}$ are included by the term within the curly brackets.
1350:
1351: Let us prove that this term is definitely positive, it can be done by
1352: proving that the eigenvalues of the matrix $M$,%
1353: \begin{align}
1354: M_{\mathbf{k}_{2}\mathbf{k}_{2}^{^{\prime }}}& =2\xi _{\mathbf{k}_{2}}\tanh (%
1355: \frac{1}{2}\beta \xi _{\mathbf{k}_{2}})\delta _{\mathbf{k_{2}k}%
1356: _{2}^{^{\prime }}}-g\varepsilon (\mathbf{k}_{2})\varepsilon (\mathbf{k}%
1357: _{2}^{\prime }) \notag \\
1358: & \times \frac{\tanh (\frac{1}{2}\beta \xi _{\mathbf{k}_{2}})}{\xi _{\mathbf{%
1359: k}_{2}}}\frac{\tanh (\frac{1}{2}\beta \xi _{\mathbf{k}_{2}^{\prime }})}{\xi
1360: _{\mathbf{k}_{2}^{\prime }}}, \label{Mkk}
1361: \end{align}%
1362: are all positive. The eigenvalue function for $M$ is,
1363: \begin{equation}
1364: |\Omega (\lambda )|=0, \label{EigenValueEq}
1365: \end{equation}%
1366: where%
1367: \begin{equation}
1368: \Omega _{\mathbf{k}_{2}\mathbf{k}_{2}^{^{\prime }}}(\lambda )=\lambda \delta
1369: _{\mathbf{k_{2}k}_{2}^{^{\prime }}}-M_{\mathbf{k}_{2}\mathbf{k}%
1370: _{2}^{^{\prime }}}.
1371: \end{equation}%
1372: Observe that the second term on the right-hand side of Eq. (\ref{Mkk}) is a
1373: product of the two factors corresponding to the row $\mathbf{k}_{2}$ and the
1374: column $\mathbf{k}_{2}^{^{\prime }}$ respectively, we can reduce Eq. (\ref%
1375: {EigenValueEq}) into
1376: \begin{equation}
1377: |\widetilde{\Omega }(\lambda )|=0, \label{EigenValueEq1}
1378: \end{equation}%
1379: where
1380: \begin{equation}
1381: \widetilde{\Omega }_{\mathbf{k}_{2}\mathbf{k}_{2}^{^{\prime }}}(\lambda )=%
1382: \frac{\lambda -2\xi _{\mathbf{k}_{2}}\tanh (\frac{1}{2}\beta \xi _{\mathbf{k}%
1383: _{2}})}{g\left[ \varepsilon (\mathbf{k}_{2})\tanh (\frac{1}{2}\beta \xi _{%
1384: \mathbf{k}_{2}})/\xi _{\mathbf{k}_{2}}\right] ^{2}}\delta _{\mathbf{k_{2}k}%
1385: _{2}^{^{\prime }}}+1.
1386: \end{equation}%
1387: Obviously,%
1388: \begin{equation}
1389: \lambda =2\xi _{\mathbf{k}_{2}}\tanh (\frac{1}{2}\beta \xi _{\mathbf{k}%
1390: _{2}})>0 \label{LambdaRoot}
1391: \end{equation}%
1392: is an eigenvalue of Eq. (\ref{EigenValueEq1}) because at least two rows ($%
1393: \pm \mathbf{k}_{2}$) of the determinant $|\widetilde{\Omega }(\lambda )|$
1394: are identical. Apart from those eigenvalues, there are possibly other ones,
1395: they satisfy the equation,
1396: \begin{equation}
1397: 1=g\sum_{\mathbf{k}_{2}}\frac{\left[ \varepsilon (\mathbf{k}_{2})\tanh (%
1398: \frac{1}{2}\beta \xi _{\mathbf{k}_{2}})/\xi _{\mathbf{k}_{2}}\right] ^{2}}{%
1399: 2\xi _{\mathbf{k}_{2}}\tanh (\frac{1}{2}\beta \xi _{\mathbf{k}_{2}})-\lambda
1400: }. \label{EigenValueEq2}
1401: \end{equation}%
1402: To derive this equation from Eq. (\ref{EigenValueEq1}), it is sufficient to
1403: heed that every element of $\widetilde{\Omega }(\lambda )$ contains the
1404: number $1$. If $\lambda \leq 0$, one has%
1405: \begin{equation}
1406: g\sum_{\mathbf{k}_{2}}\frac{\left[ \varepsilon (\mathbf{k}_{2})\tanh (\frac{1%
1407: }{2}\beta \xi _{\mathbf{k}_{2}})/\xi _{\mathbf{k}_{2}}\right] ^{2}}{2\xi _{%
1408: \mathbf{k}_{2}}\tanh (\frac{1}{2}\beta \xi _{\mathbf{k}_{2}})-\lambda }<%
1409: \frac{1}{2}g\sum\limits_{\mathbf{k}}\frac{\tanh (\frac{1}{2}\beta \xi _{%
1410: \mathbf{k}})}{\xi _{\mathbf{k}}}. \label{Inequ}
1411: \end{equation}%
1412: Eqs. (\ref{EigenValueEq2}) and (\ref{Inequ}) lead us to%
1413: \begin{equation}
1414: \frac{1}{2}g\sum\limits_{\mathbf{k}}\frac{\tanh \left( \frac{1}{2}\beta \xi
1415: _{\mathbf{k}}\right) }{\xi _{\mathbf{k}}}>1.
1416: \end{equation}%
1417: This inequality contradicts the result of Eq. (\ref{GapEq}),%
1418: \begin{equation}
1419: \frac{1}{2}g\sum\limits_{\mathbf{k}}\frac{\tanh (\frac{1}{2}\beta \xi _{%
1420: \mathbf{k}})}{\xi _{\mathbf{k}}}\leq 1.
1421: \end{equation}%
1422: That is to say, the root of Eq. (\ref{EigenValueEq2}) must be greater than
1423: zero if it has any. Combination of this result with Eq. (\ref{LambdaRoot})
1424: demonstrates that all the eigenvalues of the matrix $M$ are definitely
1425: positive, which ends our proof for the positivity of the term within the
1426: curly brackets of Eq. (\ref{DeltaS2}).
1427:
1428: This positivity implies that%
1429: \begin{equation}
1430: \delta ^{2}S\geq 0.
1431: \end{equation}%
1432: When $\delta ^{2}S>0$, we know from Eq. (\ref{DeltaSBCS}) that $\Delta S>0$,
1433: it meets the requirement of Eq. (\ref{Increment}). However, if $\delta
1434: ^{2}S=0$, we must generally examine $\delta ^{3}S$, $\delta ^{4}S$, or even
1435: more higher variations of $S$ to check whether $\Delta S\geq 0$, which is
1436: obviously complicated to handle. Fortunately, we need not do that in the
1437: case considered now. When $\delta ^{2}S=0$, Eq. (\ref{DeltaS2}) shows that%
1438: \begin{equation}
1439: \delta \theta _{\mathbf{k}_{2}}=0,\text{ }\mathrm{if}\text{ }T\geq T_{c};%
1440: \text{ }\mathrm{and}\text{ }\left\{
1441: \begin{array}{l}
1442: \delta \varphi _{\mathbf{k}}=\delta \varphi _{\mathbf{k}^{\prime }} \\
1443: \delta \theta _{\mathbf{k}_{1}}=0 \\
1444: \delta \theta _{\mathbf{k}_{2}}=0,%
1445: \end{array}%
1446: \right. \text{ }\mathrm{if}\text{ }T<T_{c}.
1447: \end{equation}%
1448: With them, one can easily verify that $S(\phi ,\beta )=S(\phi _{0},\beta )$.
1449: Namely, $\Delta S=0$ if $\delta ^{2}S=0$.
1450:
1451: To sum up, $\Delta S>0$ if $\delta ^{2}S>0$, and $\Delta S=0$ if $\delta
1452: ^{2}S=0$. Since $\delta ^{2}S\geq 0$, we conclude that the requirement $%
1453: \Delta S\geq 0$ is satisfied, the nontrivial solution is stable.%
1454: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}}}%
1455: %BeginExpansion
1456: \begin{widetext}%
1457: %EndExpansion
1458:
1459: As regards the trivial solution, $\phi =0$ ($\Lambda =0$), it is identical
1460: to the nontrivial one when $T\geq T_{c}$, and thus is stable at $T\geq T_{c}$%
1461: . If $T<T_{c}$, we find
1462: \begin{equation}
1463: \delta ^{2}S=\beta \left\{ \sum_{\mathbf{k}_{2}}2\varepsilon (\mathbf{k}%
1464: _{2})\tanh \left( \frac{1}{2}\beta \varepsilon (\mathbf{k}_{2})\right)
1465: \left( \delta \theta _{\mathbf{k}_{2}}\right) ^{2}-g\sum_{\mathbf{k}%
1466: _{2}}\sum_{\mathbf{k}_{2}^{\prime }}\tanh \left( \frac{1}{2}\beta
1467: \varepsilon (\mathbf{k}_{2})\right) \tanh \left( \frac{1}{2}\beta
1468: \varepsilon (\mathbf{k}_{2}^{\prime })\right) \delta \theta _{\mathbf{k}%
1469: _{2}}\delta \theta _{\mathbf{k}_{2}^{\prime }}\right\} . \label{NDS2}
1470: \end{equation}%
1471: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
1472: %BeginExpansion
1473: \end{widetext}
1474: %EndExpansion
1475: Accordingly, it has positive eigenvalues,%
1476: \begin{equation}
1477: \lambda =2\varepsilon (\mathbf{k}_{2})\tanh \left( \frac{1}{2}\beta
1478: \varepsilon (\mathbf{k}_{2})\right) >0,
1479: \end{equation}%
1480: In addition, we also have
1481: \begin{equation}
1482: 1=g\sum_{\mathbf{k}_{2}}\frac{\left[ \tanh \left( \frac{1}{2}\beta
1483: \varepsilon (\mathbf{k}_{2})\right) \right] ^{2}}{2\varepsilon (\mathbf{k}%
1484: _{2})\tanh \left( \frac{1}{2}\beta \varepsilon (\mathbf{k}_{2})\right)
1485: -\lambda }.
1486: \end{equation}%
1487: In contrast to Eq. (\ref{EigenValueEq2}), this equation has a negative root
1488: of $\lambda $ when $T<T_{c}$. To see it clearly, let us consider, for
1489: instance, the zero temperature case,%
1490: \begin{equation}
1491: 1=2g\sum_{\varepsilon (\mathbf{k})>0}\frac{1}{2\varepsilon (\mathbf{k}%
1492: )-\lambda }=2g\mathcal{N}(0)\int_{0}^{\hbar \omega _{D}}\mathrm{d}%
1493: \varepsilon \frac{1}{2\varepsilon -\lambda },
1494: \end{equation}%
1495: where, as usual, $\mathcal{N}(0)$ denotes the density of states at the Fermi
1496: level, and $\omega _{D}$ the Debye frequency. Apparently, it has the
1497: solution,%
1498: \begin{equation}
1499: \lambda =-\frac{2\hbar \omega _{D}}{\exp (\frac{1}{g\mathcal{N}(0)})-1}<0.
1500: \end{equation}%
1501: The existence of both the positive and negative eigenvalues implies that the
1502: trivial solution is a saddle point at $T<T_{c}$. Therefore, the term within
1503: the curly brackets of Eq. (\ref{NDS2}) is indefinite, that is, it can change
1504: in sign. As a consequence, the trivial solution, or rather the normal phase,
1505: is unstable at $T<T_{c}$. That is just the Cooper instability \cite%
1506: {BCS1,BCS2,BCS3}.
1507:
1508: Of the two solutions, the nontrivial is the only one that is stable within
1509: the whole temperature range.
1510:
1511: With the nontrivial solution obtained, other physical quantities can be
1512: easily expressed and calculated in terms of the two Green's functions $%
1513: \langle\langle c_{\mathbf{k}\uparrow}|c_{\mathbf{k}\uparrow}^{\dag}\rangle%
1514: \rangle_{\omega}$ and $\langle\langle c_{\mathbf{k}\uparrow}|c_{-\mathbf{k}%
1515: \downarrow}\rangle\rangle_{\omega}$, as was done in Refs. \cite{BCS3}, \cite%
1516: {Rickayzen} and \cite{Ambegaokar}, with the results unchanged and the same
1517: as the BCS theory.
1518:
1519: So far, BCS superconductivity has been interpreted within the extended
1520: ensemble theory. In this interpretation, the order parameter for BCS
1521: superconductivity evolves with temperature according to the principle of
1522: least entropy. At high temperatures ($T\geq T_{c}$), it is zero and stable,
1523: the system Hamiltonian realizes the representation with perfect gauge
1524: symmetry, the resulting phase is normal and disordered. As temperature
1525: decreases and goes below the critical temperature ($T<T_{c}$), the zero
1526: solution becomes unstable, instead, there arises a new stable solution of
1527: the order parameter, it is nonzero. At the same time, the system Hamiltonian
1528: transforms into a new representation with gauge symmetry broken, the
1529: electrons are hence formed into Cooper pairs, and the resulting phase gets
1530: superconducting. Thus and so, the extended ensemble theory answers why, when
1531: and how the electrons can coagulate to form Cooper pairs and
1532: superconductivity, the problem posed by Born and Fucks \cite{Born}.
1533:
1534: Lastly, it should be noted that the BCS and Landau theories are unified into
1535: a single formalism within the framework of the extended ensemble theory.
1536:
1537: \section{Bose-Einstein Condensation \label{BECIdeal}}
1538:
1539: After the discussion of BCS superconductivity, we proceed now to study
1540: Bose-Einstein condensation (BEC). We shall concentrate on three systems: the
1541: ideal Bose gas, the photon gas in a black body, and the ideal phonon gas in
1542: a solid body. They are the most important cases of Bose systems: the first
1543: is a system with the conservation of particles, but the other two are not;
1544: the second belongs to gauge fields whereas the third does not. In addition,
1545: we shall also discuss the quantization of Dirac field from a standpoint of
1546: the extended ensemble theory.
1547:
1548: \subsection{The Ideal Bose Gas}
1549:
1550: Let us begin with the ideal Bose gas. Its Hamiltonian reads as follows,
1551: \begin{equation}
1552: H(b)=\sum_{\mathbf{k}}(\epsilon _{\mathbf{k}}-\mu )b_{\mathbf{k}}^{\dag }b_{%
1553: \mathbf{k}}, \label{Hibg}
1554: \end{equation}%
1555: where $b_{\mathbf{k}}$ ($b_{\mathbf{k}}^{\dag }$) is the annihilation
1556: (creation) operator for the bosons with a momentum $\hbar \mathbf{k}$, and $%
1557: \epsilon _{\mathbf{k}}=\hbar ^{2}\mathbf{k}^{2}/\left( 2m\right) $ the
1558: single-particle energy, and $\mu $ the chemical potential. This Hamiltonian
1559: is gauge invariant,%
1560: \begin{equation}
1561: G(\vartheta ,b)H(b)G^{\dag }(\vartheta ,b)=H(b),
1562: \end{equation}%
1563: where
1564: \begin{equation}
1565: G(\vartheta ,b)=e^{-i\vartheta \sum_{\mathbf{k}}b_{\mathbf{k}}^{\dag }b_{%
1566: \mathbf{k}}},\text{ }\vartheta \in \lbrack 0,2\pi ).
1567: \end{equation}%
1568: The invariance has an important consequence,
1569: \begin{equation}
1570: \left\langle b_{\mathbf{k}}\right\rangle =\mathrm{Tr}\big(b_{\mathbf{k}}\rho
1571: (H(b))\big)=0. \label{BECGauge}
1572: \end{equation}%
1573: It signifies that the condensation amplitude will be zero forever if the
1574: gauge symmetry does not break down.
1575:
1576: In order to examine whether the gauge symmetry can break down spontaneously
1577: or not, one needs, following the hypotheses in Sec. \ref{EET}, to consider
1578: the phase-transition operator,
1579: \begin{equation}
1580: D(\eta,b)=\sum_{\mathbf{k}}(\eta_{\mathbf{k}}^{\dag}b_{\mathbf{k}}+\eta_{%
1581: \mathbf{k}}b_{\mathbf{k}}^{\dag}), \label{BECPTO}
1582: \end{equation}
1583: where the internal field $\eta$ is the order parameter for BEC. The reason
1584: for introducing Eq. (\ref{BECPTO}) can be seen from comparing Eqs. (\ref%
1585: {BECGauge}) and (\ref{BECPTO}) with Eqs. (\ref{Zero}) and (\ref{Dfactor})
1586: respectively.
1587:
1588: We shall consider first the finite volume case where $V$ and $N$ are both
1589: finite, and then the thermodynamic limit case where $V\rightarrow+\infty$,
1590: and $N\rightarrow+\infty$, but $N/V$ is finite. As will be seen, the two
1591: cases are significantly different.
1592:
1593: \subsubsection{The Finite Volume Case}
1594:
1595: In this case, the wave vector $\mathbf{k}$ is discrete. The entropy of the
1596: system can be written as
1597: \begin{equation}
1598: S(\eta,\beta)=S(0,\beta)+\beta\sum_{\mathbf{k}}(\epsilon_{\mathbf{k}}-\mu
1599: )\eta_{\mathbf{k}}^{\dag}\eta_{\mathbf{k}}, \label{BECS}
1600: \end{equation}
1601: where
1602: \begin{equation}
1603: S(0,\beta)=-\mathrm{Tr}\big(\ln\left( \rho(H(b))\right) \rho(H(b))\big)\!
1604: \end{equation}
1605: is independent of the order parameter $\eta$.
1606:
1607: According to Eqs. (\ref{Variation}) and (\ref{Increment}), we have
1608: \begin{gather}
1609: \frac{\partial S}{\partial \eta _{\mathbf{k}}}=0, \label{OrderParam0} \\
1610: \Delta S=\beta \sum_{\mathbf{k}}(\epsilon _{\mathbf{k}}-\mu )\delta \eta _{%
1611: \mathbf{k}}^{\dag }\delta \eta _{\mathbf{k}}\geq 0, \label{BECDelta}
1612: \end{gather}%
1613: where $\delta \eta _{\mathbf{k}}$ represents the variation of $\eta _{%
1614: \mathbf{k}}$ from the solution given by Eq. (\ref{OrderParam0}). Eq. (\ref%
1615: {BECDelta}) shows that the chemical potential $\mu $ must satisfy the
1616: following condition,%
1617: \begin{equation}
1618: \mu \leq \epsilon _{\mathbf{k}}.
1619: \end{equation}%
1620: As $\epsilon _{\mathbf{k}}\geq 0$, this condition is equivalent to
1621: \begin{equation}
1622: \mu \leq 0. \label{StableCond}
1623: \end{equation}%
1624: This inequality is the physical condition for the stability of the system.
1625:
1626: In combination with Eq. (\ref{BECS}), Eq. (\ref{OrderParam0}) gives us the
1627: equation of order parameter,
1628: \begin{equation}
1629: (\epsilon _{\mathbf{k}}-\mu )\eta _{\mathbf{k}}=0.
1630: \end{equation}%
1631: It has two solutions: the trivial one,%
1632: \begin{equation}
1633: \eta _{\mathbf{k}}=0,
1634: \end{equation}%
1635: and the nontrivial one,%
1636: \begin{equation}
1637: \eta _{\mathbf{k}}=\eta _{0}\delta _{\mathbf{k,}0},\text{ }\mathrm{only}%
1638: \text{ }\mathrm{if}\text{ }\mu =0.
1639: \end{equation}
1640:
1641: For the nontrivial solution, $\mu =0$, it causes a contradiction,%
1642: \begin{eqnarray}
1643: \eta _{0}^{+}\eta _{0} &=&N-\sum_{\mathbf{k}}\mathrm{Tr}\big(b_{\mathbf{k}%
1644: }^{\dag }b_{\mathbf{k}}\rho (H(b))\big) \notag \\
1645: &=&N-\sum_{\mathbf{k}}\frac{1}{e^{\beta \epsilon _{\mathbf{k}}}-1},
1646: \label{Lesb0}
1647: \end{eqnarray}%
1648: where the left-hand side can not be negative, but the right-hand side is
1649: negative because the term of $\mathbf{k}=0$ is infinite. Therefore, the
1650: nontrivial solution must be discarded.
1651:
1652: With regard to the trivial solution, it corresponds to the normal phase of
1653: the system. It does not cause any contradiction,%
1654: \begin{equation}
1655: N=\sum_{\mathbf{k}}\mathrm{Tr}\big(b_{\mathbf{k}}^{\dag }b_{\mathbf{k}}\rho
1656: (H(b))\big)=\sum_{\mathbf{k}}\frac{1}{e^{\beta \left( \epsilon _{\mathbf{k}%
1657: }-\mu \right) }-1},
1658: \end{equation}%
1659: this equation can be satisfied at any temperature. In fact, it is just the
1660: equation for the chemical potential $\mu $, its solution will give $\mu =\mu
1661: (T)$. Because $N$ is finite, $\mu $ will be less than zero irrespective of
1662: temperature, i.e.,%
1663: \begin{equation}
1664: \mu (T)<0,\text{ for }T>0.
1665: \end{equation}%
1666: It implies that the normal phase of the system is always stable,%
1667: \begin{equation}
1668: \Delta S>0.
1669: \end{equation}%
1670: This proves that there is no phase transition in the finite volume case, the
1671: system will stay in its normal phase forever.
1672:
1673: \subsubsection{The Thermodynamic Limit Case \label{TLC}}
1674:
1675: In this case, $V\rightarrow +\infty $, we should use the density of entropy
1676: instead of the entropy itself,
1677: \begin{eqnarray}
1678: \hspace{-0.3in}s(\eta ,\beta ) &=&\lim_{V\rightarrow +\infty }\frac{S(\eta
1679: ,\beta )}{V} \notag \\
1680: &=&s(0,\beta )+\beta \lim_{V\rightarrow +\infty }\frac{1}{V}\sum_{\mathbf{k}%
1681: }(\epsilon _{\mathbf{k}}-\mu )\eta _{\mathbf{k}}^{\dagger }\eta _{\mathbf{k}%
1682: }, \label{sdensity}
1683: \end{eqnarray}%
1684: where
1685: \begin{equation}
1686: s(0,\beta )=-\lim_{V\rightarrow +\infty }\frac{1}{V}\mathrm{Tr}\big(\ln
1687: \left( \rho (H(b))\right) \rho (H(b))\big)
1688: \end{equation}%
1689: is a term irrelevant to $\eta $. The limit $V\rightarrow +\infty $ can be
1690: achieved as follows.
1691:
1692: As usual, suppose that the system is box normalized, i.e.,
1693: \begin{equation}
1694: k_{x(y,z)}=n_{x(y,z)}\frac{2\pi }{l_{x(y,z)}},\text{\ }n_{x(y,z)}\in
1695: %TCIMACRO{\U{2124} }%
1696: %BeginExpansion
1697: \mathbb{Z}
1698: %EndExpansion
1699: , \label{Quantum}
1700: \end{equation}%
1701: where $%
1702: %TCIMACRO{\U{2124} }%
1703: %BeginExpansion
1704: \mathbb{Z}
1705: %EndExpansion
1706: $ denotes the set of integers, and $l_{x}$, $l_{y}$, and $l_{z}$ the
1707: dimensions of the box along the $x$, $y$ and $z$ directions respectively ($%
1708: V=l_{x}l_{y}l_{z}$). As is well known, each $\mathbf{k}$ corresponds to an
1709: eigenfunction, all the eigenfunctions constitute a Hilbert space. Here, the
1710: Hilbert space is a complete space of integrable functions, with the inner
1711: product defined as a Lebesgue integral. It should be pointed out that the
1712: inner product must be defined as a Lebesgue rather than Riemann integral
1713: because the former can ensure the completeness of an integrable function
1714: space whereas the latter can not \cite{Hewitt}. This implies that all the
1715: problems relevant to a Hilbert space of integrable functions must be treated
1716: with Lebesgue theory. On the other hand, Lebesgue measure and integration is
1717: itself the mathematical foundation of probability theory \cite%
1718: {Kolmogorov,Feller} and statistical physics \cite{Georgii}. We shall
1719: therefore perform the thermodynamic limit according to the Lebesgue theory
1720: of integration.
1721:
1722: Formally, Lebesgue integration includes both the sum over discrete numbers
1723: and the integral on a continuous region. When $V$ is finite, a sum over $%
1724: \mathbf{k}$ is a Lebesgue integral of discrete form. It will transform into
1725: a Lebesgue integral of continuous form as $V\rightarrow +\infty $. To show
1726: the transformation, let us consider, e.g., the sum over $\mathbf{k}$ in Eq. (%
1727: \ref{sdensity}). In the first place, we write it into an explicit form of
1728: Lebesgue integral,
1729: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}} }%
1730: %BeginExpansion
1731: \begin{widetext}
1732: %EndExpansion
1733: \begin{equation}
1734: \frac{1}{V}\sum_{\mathbf{k}}(\epsilon _{\mathbf{k}}-\mu )\eta _{\mathbf{k}%
1735: }^{\dagger }\eta _{\mathbf{k}}=\frac{1}{\left( 2\pi \right) ^{3}}\mathbf{%
1736: \sum_{\mathbf{k}}}(\epsilon _{\mathbf{k}}-\mu )\eta _{\mathbf{k}}^{\dag
1737: }\eta _{\mathbf{k}}m(A_{\mathbf{k}}\cap
1738: %TCIMACRO{\U{211d} }%
1739: %BeginExpansion
1740: \mathbb{R}
1741: %EndExpansion
1742: ^{3})\mathbf{=}\frac{1}{\left( 2\pi \right) ^{3}}\int\limits_{E}h(\mathbf{p})%
1743: \mathrm{d}\mathbf{p}, \label{LI}
1744: \end{equation}%
1745: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
1746: %BeginExpansion
1747: \end{widetext}
1748: %EndExpansion
1749: where%
1750: \begin{gather}
1751: %TCIMACRO{\U{211d} }%
1752: %BeginExpansion
1753: \mathbb{R}
1754: %EndExpansion
1755: ^{3}=\left( -\infty ,+\infty \right) \times \left( -\infty ,+\infty \right)
1756: \times \left( -\infty ,+\infty \right) , \\
1757: m(A_{\mathbf{k}}\cap
1758: %TCIMACRO{\U{211d} }%
1759: %BeginExpansion
1760: \mathbb{R}
1761: %EndExpansion
1762: ^{3})=\frac{\left( 2\pi \right) ^{3}}{V}, \\
1763: A_{\mathbf{k}}=[k_{x}\mathbf{,}k_{x}\mathbf{+}\frac{2\pi }{l_{x}})\times
1764: \lbrack k_{y}\mathbf{,}k_{y}\mathbf{+}\frac{2\pi }{l_{y}})\times \lbrack
1765: k_{z}\mathbf{,}k_{z}\mathbf{+}\frac{2\pi }{l_{z}}), \label{ASET} \\
1766: h(\mathbf{p})=\mathbf{\sum_{\mathbf{k}}}(\epsilon _{\mathbf{k}}-\mu )\eta _{%
1767: \mathbf{k}}^{\dag }\eta _{\mathbf{k}}\chi _{A_{\mathbf{k}}}(\mathbf{p}),
1768: \label{Stepf} \\
1769: \chi _{A_{\mathbf{k}}}(\mathbf{p})=\left\{
1770: \begin{array}{ll}
1771: 1, & \mathbf{p}\in A_{\mathbf{k}} \\
1772: 0, & \mathbf{p}\notin A_{\mathbf{k}}.%
1773: \end{array}%
1774: \right. \label{Characterf}
1775: \end{gather}%
1776: As usual, $\chi _{A_{\mathbf{k}}}(\mathbf{p})$ represents the characteristic
1777: function of the measurable set $A_{\mathbf{k}}$, and $m(A_{\mathbf{k}}\cap
1778: %TCIMACRO{\U{211d} }%
1779: %BeginExpansion
1780: \mathbb{R}
1781: %EndExpansion
1782: ^{3})$ the Lebesgue measure of the set $A_{\mathbf{k}}$. From Eqs. (\ref%
1783: {ASET}--\ref{Characterf}), it follows that
1784: \begin{eqnarray}
1785: \lim_{V\rightarrow +\infty }h(\mathbf{p}) &=&(\epsilon _{\mathbf{p}}-\mu
1786: )\eta _{\mathbf{p}}^{\dag }\eta _{\mathbf{p}} \notag \\
1787: &\equiv &[\epsilon (\mathbf{p})-\mu ]\eta ^{\dag }(\mathbf{p})\eta (\mathbf{p%
1788: }). \label{VLimit}
1789: \end{eqnarray}%
1790: Now, in the second place, let us take the thermodynamic limit on both the
1791: sides of Eq. (\ref{LI}),
1792: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}}}%
1793: %BeginExpansion
1794: \begin{widetext}%
1795: %EndExpansion
1796: \begin{equation}
1797: \lim_{V\rightarrow +\infty }\frac{1}{V}\sum_{\mathbf{k}}(\epsilon _{\mathbf{k%
1798: }}-\mu )\eta _{\mathbf{k}}^{\dagger }\eta _{\mathbf{k}}=\frac{1}{\left( 2\pi
1799: \right) ^{3}}\lim_{V\rightarrow +\infty }\int\limits_{%
1800: %TCIMACRO{\U{211d} }%
1801: %BeginExpansion
1802: \mathbb{R}
1803: %EndExpansion
1804: ^{3}}h(\mathbf{p})\mathrm{d}\mathbf{p}=\frac{1}{\left( 2\pi \right) ^{3}}%
1805: \int\limits_{%
1806: %TCIMACRO{\U{211d} }%
1807: %BeginExpansion
1808: \mathbb{R}
1809: %EndExpansion
1810: ^{3}}\lim_{V\rightarrow +\infty }h(\mathbf{p})\mathrm{d}\mathbf{p}=\frac{1}{%
1811: \left( 2\pi \right) ^{3}}\int\limits_{%
1812: %TCIMACRO{\U{211d} }%
1813: %BeginExpansion
1814: \mathbb{R}
1815: %EndExpansion
1816: ^{3}}[\epsilon (\mathbf{k})-\mu ]\eta ^{\dag }(\mathbf{k})\eta (\mathbf{k})%
1817: \mathrm{d}\mathbf{k}.
1818: \end{equation}%
1819: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
1820: %BeginExpansion
1821: \end{widetext}
1822: %EndExpansion
1823: In such a natural way, a discrete sum over $\mathbf{k}$ will transform into
1824: a continuous Lebesgue integral in the thermodynamic limit.
1825:
1826: With the help of the above equation, one arrives at%
1827: \begin{gather}
1828: \left[ \epsilon (\mathbf{k})-\mu \right] \eta (\mathbf{k})=0,
1829: \label{BECCOrder} \\
1830: \Delta s=\beta \frac{1}{\left( 2\pi \right) ^{3}}\int\limits_{%
1831: %TCIMACRO{\U{211d} }%
1832: %BeginExpansion
1833: \mathbb{R}
1834: %EndExpansion
1835: ^{3}}[\epsilon (\mathbf{k})-\mu ]\delta \eta ^{\dag }(\mathbf{k})\delta \eta
1836: (\mathbf{k})\mathrm{d}\mathbf{k}\geq 0. \label{BECCDelta}
1837: \end{gather}%
1838: The second equation means that the chemical potential must be less than or
1839: equal to zero,%
1840: \begin{equation}
1841: \mu \leq 0, \label{BECMu}
1842: \end{equation}%
1843: it is the condition for the stability of the system. The first equation is
1844: equivalent to
1845: \begin{equation}
1846: \lbrack \epsilon (\mathbf{k})-\mu ]\eta ^{\dag }(\mathbf{k})\eta (\mathbf{k}%
1847: )=0, \label{Eta}
1848: \end{equation}%
1849: which can be easily verified with respect to the two cases: $\eta ^{\dag }(%
1850: \mathbf{k})=0$ and $\eta ^{\dag }(\mathbf{k})\neq 0$. Eq. (\ref{Eta}) has
1851: two solutions: the trivial one,%
1852: \begin{equation}
1853: \eta (\mathbf{k})=0,
1854: \end{equation}%
1855: and the nontrivial one,%
1856: \begin{equation}
1857: \eta (\mathbf{k})=\xi \sqrt{\delta (\mathbf{k})},\text{ only if }\mu =0,
1858: \end{equation}%
1859: where $\xi $ is a complex which does not depend on $\mathbf{k}$, and $\delta
1860: (\mathbf{k})$ the Dirac $\delta $ function.
1861:
1862: For the trivial solution, the system is in its normal phase, this phase can
1863: exist only at high temperatures, which can be easily deduced from the
1864: equation of chemical potential,%
1865: \begin{eqnarray}
1866: n &\equiv &\lim_{V\rightarrow +\infty }\frac{N}{V} \notag \\
1867: &=&\lim_{V\rightarrow +\infty }\frac{1}{V}\sum_{\mathbf{k}}\mathrm{Tr}\big(%
1868: b_{\mathbf{k}}^{\dag }b_{\mathbf{k}}\rho (H(b))\big) \notag \\
1869: &=&\frac{1}{\left( 2\pi \right) ^{3}}\int\limits_{%
1870: %TCIMACRO{\U{211d} }%
1871: %BeginExpansion
1872: \mathbb{R}
1873: %EndExpansion
1874: ^{3}}\frac{1}{e^{\beta (\epsilon \left( \mathbf{k}\right) -\mu )}-1}\mathrm{d%
1875: }\mathbf{k} \notag \\
1876: &=&\frac{(2m)^{3/2}}{4\pi ^{2}\hbar ^{3}}\int_{0^{+}}^{+\infty }\mathrm{d}%
1877: \epsilon \frac{\epsilon ^{1/2}}{e^{\beta (\epsilon -\mu )}-1}, \label{Lesb1}
1878: \end{eqnarray}%
1879: where we have reduced the Lebesgue integral into an improper Riemann
1880: integral in the end. Because, as shown by Eq. (\ref{BECMu}), $\mu $ must be
1881: less than or equal to zero, the above equation can not hold if $T<T_{c}$
1882: where $T_{c}$ is determined by
1883: \begin{equation}
1884: n=\frac{(2m)^{3/2}}{4\pi ^{2}\hbar ^{3}}\int_{0^{+}}^{+\infty }\mathrm{d}%
1885: \epsilon \frac{\epsilon ^{1/2}}{e^{\beta _{c}\epsilon }-1}. \label{Lesb2}
1886: \end{equation}%
1887: As is well known, this gives%
1888: \begin{equation}
1889: T_{c}=\frac{2\pi \hbar ^{2}}{mk_{B}}\left( \frac{n}{\zeta (3/2}\right)
1890: ^{2/3},
1891: \end{equation}%
1892: where $\zeta (x)$ denotes the Riemann zeta function.
1893:
1894: For the nontrivial solution, $\mu $ must be zero, the order parameter is
1895: determined by%
1896: \begin{eqnarray}
1897: n &=&\frac{1}{\left( 2\pi \right) ^{3}}\int\limits_{%
1898: %TCIMACRO{\U{211d} }%
1899: %BeginExpansion
1900: \mathbb{R}
1901: %EndExpansion
1902: ^{3}}\eta ^{\dagger }(\mathbf{k})\eta (\mathbf{k})\mathrm{d}\mathbf{k+}\frac{%
1903: 1}{\left( 2\pi \right) ^{3}}\int\limits_{%
1904: %TCIMACRO{\U{211d} }%
1905: %BeginExpansion
1906: \mathbb{R}
1907: %EndExpansion
1908: ^{3}}\frac{1}{e^{\beta \epsilon \left( \mathbf{k}\right) }-1}\mathrm{d}%
1909: \mathbf{k} \notag \\
1910: &=&\frac{1}{\left( 2\pi \right) ^{3}}\xi ^{\dagger }\xi \mathbf{+}\frac{%
1911: (2m)^{3/2}}{4\pi ^{2}\hbar ^{3}}\int_{0^{+}}^{+\infty }\mathrm{d}\epsilon
1912: \frac{\epsilon ^{1/2}}{e^{\beta \epsilon }-1}, \label{Lesb3}
1913: \end{eqnarray}%
1914: where we have used the fact that the contribution from the state of $\mathbf{%
1915: k}=0$ is zero when reducing the Lebesgue integral into the improper Riemann
1916: integral in the end. Because $\xi ^{\dagger }\xi \geq 0$, the above equation
1917: can not hold if $T>T_{c}$.
1918:
1919: From those discussions, it follows that%
1920: \begin{equation}
1921: \eta =\left\{
1922: \begin{array}{ll}
1923: \mathrm{zero,}\text{ } & T\geq T_{c} \\
1924: \text{\textrm{nonzero}}\mathrm{,}\text{ } & T<T_{c}.%
1925: \end{array}%
1926: \right.
1927: \end{equation}%
1928: That is to say, a phase transition will happen at $T=T_{c}$. Because $\Delta
1929: s\geq 0$ when $\mu \leq 0$, the normal phase ($\eta =0$) exists and is
1930: stable at $T\geq T_{c}$, the ordered phase ($\eta \neq 0$) exists and is
1931: stable at $T<T_{c}$. Here, we note that it is a natural consequence of the
1932: extended ensemble theory that the chemical potential of the ideal Bose gas
1933: is fixed at zero ($\mu =0$) in the ordered phase ($T<T_{c}$), in contrast to
1934: the Einstein's treatment where it is fixed by physical insight \cite%
1935: {Einstein1,Einstein2}.
1936:
1937: Obviously, the above results are the same as the BEC given by Einstein \cite%
1938: {Einstein1,Einstein2} if $\xi ^{\dagger }\xi /(2\pi )^{3}$ were regarded as
1939: the density of the bosons condensed onto the state of $\mathbf{k}=0$.
1940: Therefore, the transition happening at $T=T_{c}$ must be identified
1941: physically as the Bose-Einstein condensation. It should be pointed out that,
1942: generally, the quantity $\xi ^{\dagger }\xi /(2\pi )^{3}$ can not be
1943: interpreted as the density of condensed bosons, as will be clarified in Sec. %
1944: \ref{WIBG}.
1945:
1946: So far, we have presented a new description to the Bose-Einstein
1947: condensation of the ideal Bose gas. In comparison with the Einstein's
1948: description \cite{Einstein1,Einstein2}, we now describe the BEC with two new
1949: conceptions: spontaneous symmetry breaking and Lebesgue integration. Within
1950: this new description, the BEC happens simultaneously with the spontaneous
1951: breaking of the gauge symmetry; the integration is defined and performed in
1952: Lebesgue way. That is fundamentally different from Einstein's description
1953: \cite{Einstein1,Einstein2} where the BEC happens with no symmetry breaking,
1954: and the integration is done in Riemann way.
1955:
1956: Here, it is worth emphasizing the important role played by Lebesgue
1957: integration in the study of the BEC of the ideal Bose gas. As has been seen,
1958: there arises the Bose distribution function,
1959: \begin{equation}
1960: f(\epsilon \left( \mathbf{k}\right) )=\frac{1}{e^{\beta \epsilon \left(
1961: \mathbf{k}\right) }-1}, \label{FSingular}
1962: \end{equation}%
1963: which is unbounded above on the set $E$. Because the integrand of a Riemann
1964: integral must be bounded everywhere, such a function is not Riemann
1965: integrable and can not be handled with Riemann integration. Fortunately,
1966: that makes no trouble. As pointed above, it is, in fact, unnecessary for us
1967: to handle the Bose distribution function with Riemann integration; what
1968: really confronts us in statistical physics is Lebesgue other than Riemann
1969: integration. For Lebesgue integration, it permits more general functions as
1970: integrands, and treats bounded and unbounded functions on an equal footing.
1971: In the sense of Lebesgue integration, the Bose distribution function is
1972: integrable, which can be shown as follows,
1973: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}}}%
1974: %BeginExpansion
1975: \begin{widetext}%
1976: %EndExpansion
1977: \begin{eqnarray}
1978: \lim_{V\rightarrow +\infty }\frac{1}{V}\sum_{\mathbf{k}}\frac{1}{e^{\beta
1979: \epsilon _{\mathbf{k}}}-1} &=&\frac{1}{\left( 2\pi \right) ^{3}}\int\limits_{%
1980: %TCIMACRO{\U{211d} }%
1981: %BeginExpansion
1982: \mathbb{R}
1983: %EndExpansion
1984: ^{3}}\frac{1}{e^{\beta \epsilon \left( \mathbf{k}\right) }-1}\mathrm{d}%
1985: \mathbf{k=}\frac{1}{\left( 2\pi \right) ^{3}}\int\limits_{E_{0}}\frac{1}{%
1986: e^{\beta \epsilon \left( \mathbf{k}\right) }-1}\mathrm{d}\mathbf{k+}\frac{1}{%
1987: \left( 2\pi \right) ^{3}}\int\limits_{%
1988: %TCIMACRO{\U{211d} }%
1989: %BeginExpansion
1990: \mathbb{R}
1991: %EndExpansion
1992: ^{3}\backslash E_{0}}\frac{1}{e^{\beta \epsilon \left( \mathbf{k}\right) }-1}%
1993: \mathrm{d}\mathbf{k} \notag \\
1994: &=&\frac{1}{\left( 2\pi \right) ^{3}}\int\limits_{%
1995: %TCIMACRO{\U{211d} }%
1996: %BeginExpansion
1997: \mathbb{R}
1998: %EndExpansion
1999: ^{3}\backslash E_{0}}\frac{1}{e^{\beta \epsilon \left( \mathbf{k}\right) }-1}%
2000: \mathrm{d}\mathbf{k}=\frac{(2m)^{3/2}}{4\pi ^{2}\hbar ^{3}}%
2001: \int_{0^{+}}^{+\infty }\mathrm{d}\epsilon \frac{\epsilon ^{1/2}}{e^{\beta
2002: \epsilon }-1}, \label{BOSE1}
2003: \end{eqnarray}%
2004: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
2005: %BeginExpansion
2006: \end{widetext}
2007: %EndExpansion
2008: where the set $E_{0}\equiv \{\mathbf{k}=0\}$ includes one point only. $E_{0}$
2009: is a null set, i.e., its measure is zero, thus, the integral on it vanishes.
2010: The rest part, i.e., the integral on $%
2011: %TCIMACRO{\U{211d} }%
2012: %BeginExpansion
2013: \mathbb{R}
2014: %EndExpansion
2015: ^{3}\backslash E_{0}$, reduces finally into an improper Riemann integral. As
2016: is well known, the integral on a null set always vanishes irrespective of
2017: the integrand values on this set, even if they are infinite. Thereby, the
2018: integral on any single state equals zero, the ground state ($\mathbf{k}=0$)
2019: is not more advantageous than the other states ($\mathbf{k}\neq 0$). In
2020: other words, the singularity of $f(\epsilon \left( \mathbf{k}\right) )$ on
2021: the point $\mathbf{k}=0$ has no special meaning, both in mathematics and in
2022: physics.
2023:
2024: This aspect makes the present treatment quite distinct from that of
2025: Einstein, where the integration is handled in Riemann way, with a special
2026: treatment given to the ground state of $\mathbf{k}=0$. Intuitively, Einstein
2027: thought that the bosons would gather in the ground state if $T<T_{c}$, and
2028: thus the contribution from this state is nonzero \cite{Einstein1,Einstein2},
2029: that is,
2030: \begin{eqnarray}
2031: n &=&\lim_{V\rightarrow +\infty }\frac{1}{V}\sum_{\mathbf{k}}\frac{1}{%
2032: e^{\beta \epsilon _{\mathbf{k}}}-1} \notag \\
2033: &=&n_{0}+\frac{(2m)^{3/2}}{4\pi ^{2}\hbar ^{3}}\int_{0^{+}}^{+\infty }%
2034: \mathrm{d}\epsilon \frac{\epsilon ^{1/2}}{e^{\beta \epsilon }-1},
2035: \label{BOSE2}
2036: \end{eqnarray}%
2037: where $n_{0}>0$ is the particle density gathering in the ground state.
2038: Comparing Eqs. (\ref{Lesb3}), (\ref{BOSE1}) and (\ref{BOSE2}) shows more
2039: clearly the difference between the present treatment and Einstein's. From
2040: this difference it follows that the quantity $\xi ^{\dagger }\xi /(2\pi
2041: )^{3} $ needs not to be interpreted as the density of the bosons condensed
2042: onto the state of $\mathbf{k}=0$. In fact, Eq. (\ref{Lesb3}) is physically
2043: just an equation of order parameter. Here, it should be noted that the
2044: Bogoliubov theory of weakly imperfect Bose gas \cite{Bogoliubov,Zagrebnov}
2045: is based on Eq. (\ref{BOSE2}) rather than Eq. (\ref{BOSE1}). At last, why
2046: did Einstein not use Lebesgue integration? there is a historical reason:
2047: Until 1925 when Einstein \cite{Einstein1,Einstein2} studied the BEC,
2048: Lebesgue integration had not yet been introduced to probability theory by
2049: Kolmogorov, who did it in 1933 \cite{Kolmogorov}.
2050:
2051: Also, it is significant to compare the two Lebesgue integrals represented in
2052: Eq. (\ref{Lesb0}) and Eq. (\ref{Lesb3}): Both integrands are unbounded on $%
2053: %TCIMACRO{\U{211d} }%
2054: %BeginExpansion
2055: \mathbb{R}
2056: %EndExpansion
2057: ^{3}$; however, the former is infinite on a set with a nonzero measure ($%
2058: m(A_{\mathbf{k=0}}\cap
2059: %TCIMACRO{\U{211d} }%
2060: %BeginExpansion
2061: \mathbb{R}
2062: %EndExpansion
2063: ^{3})=\left( 2\pi \right) ^{3}/V$), the latter is finite almost everywhere,
2064: or infinite just on a null set ($m\left( E_{0}\cap
2065: %TCIMACRO{\U{211d} }%
2066: %BeginExpansion
2067: \mathbb{R}
2068: %EndExpansion
2069: ^{3}\right) =0$). This difference leads to the result that the nontrivial
2070: solution must be discarded in the finite volume case whereas it has physical
2071: sense in the thermodynamic limit. In a word, the ideal Bose gas can produce
2072: BEC in the thermodynamic limit, but it can not in the finite volume case.
2073: That is a rigorous conclusion deduced from the extended ensemble theory.
2074:
2075: As another instance, the ideal Bose gas shows unambiguously that Gibbs
2076: ensemble theory needs some kind of extension. Otherwise, the conservation of
2077: particles will be broken when $T<T_{c}$, as can be seen by comparing Eqs. (%
2078: \ref{BOSE1}) and (\ref{BOSE2}).
2079:
2080: As the only one of exactly solvable quantum phase transition, the BEC in the
2081: ideal Bose gas demonstrates that the extended ensemble theory postulated in
2082: Sec. \ref{EET} is feasible in physics for us to describe phase transitions.
2083:
2084: Lastly, we would like to stress that the condition of Eq. (\ref{BECMu}),
2085: i.e., $\mu \leq 0$, is not a mathematical limit but just a physical
2086: requirement. Like the case of $\mu \leq 0$, the extended ensemble theory is
2087: also well defined mathematically for $\mu >0$, as is shown in Appendix \ref%
2088: {FSum}. Removing the limitation on the chemical potential is physically
2089: reasonable, it enables the conservation of particles to be handled freely,
2090: and makes it easy to describe the BEC in the interacting Bose gas, as can be
2091: seen in Sec. \ref{WIBG}.
2092:
2093: \textbf{Remark 1}: The ideal Bose gas is a touchstone for the theory of
2094: phase transitions because it is the only one of the many-body systems that
2095: can show a quantum phase transition and can be solved rigorously. Any theory
2096: of phase transitions must be examined by the ideal Bose gas. It can not be
2097: accepted in physics if it can not give a rigorous solution to the ideal Bose
2098: gas; and it can be further applied to the interacting Bose gas only after it
2099: has given a rigorous solution to the ideal Bose gas.
2100:
2101: \textbf{Remark 2}: In 1926, Uhlenbeck \cite{Uhlenbeck} raised an objection
2102: to Einstein's viewpoint on the condensation, he argued that the ideal Bose
2103: gas can not produce BEC if the volume of the system remains finite. This
2104: objection caused an exciting debate during the van der Waals Century
2105: Conference in November 1937 \cite{Pais}. Evidently, the present theory has
2106: given the exact answer to the debate: Uhlenbeck and Einstein are both right,
2107: which is also in accordance with the agreement reached finally in the
2108: conference.
2109:
2110: \textbf{Remark 3}: In the momentum space used here, the condensed phase is
2111: shown to be homogeneous. In Sec. \ref{WIBG}, using the real space, we shall
2112: prove further that the condensed phase can only be homogeneous. There can
2113: not exist any supercurrent or quantized vortex in the ideal Bose gas.
2114:
2115: \textbf{Remark 4}: The rigorous solution for the ideal Bose gas shows that
2116: the extended ensemble theory holds in the critical region.
2117:
2118: \subsection{The Photon Gas in a Black Body}
2119:
2120: For the photon gas in a black body, its Hamiltonian reads,
2121: \begin{equation}
2122: H(a)=\sum_{\mathbf{k}\nu }\omega _{\mathbf{k}}\left( a_{\mathbf{k}\nu
2123: }^{\dag }a_{\mathbf{k}\nu }+\frac{1}{2}\right) , \label{Hphoton}
2124: \end{equation}%
2125: where Coulomb gauge is used, $a_{\mathbf{k}\nu }$ ($a_{\mathbf{k}\nu }^{\dag
2126: }$) is the destruction (construction) operator for the photons with momentum
2127: $\mathbf{k}$ and polarization $\nu $, and $\omega _{\mathbf{k}}=\hbar c%
2128: \mathbf{k}$ represents the photon energy with $c$ the speed of light.
2129:
2130: Analogous to the ideal Bose gas, this system has the phase-transition
2131: operator,
2132: \begin{equation}
2133: D(\xi,a)=\sum_{\mathbf{k}\nu}(\xi_{\mathbf{k}\nu}^{\dag}a_{\mathbf{k}%
2134: \nu}+\xi_{\mathbf{k}\nu}a_{\mathbf{k}\nu}^{\dag}),
2135: \end{equation}
2136: where $\xi$ is the order parameter for BEC.
2137:
2138: From $H(a)$\ and $D(\xi ,a)$, it follows that%
2139: \begin{gather}
2140: \omega _{\mathbf{k}}\xi _{\mathbf{k}\nu }=0, \\
2141: \Delta S=\beta \sum_{\mathbf{k,}\sigma }\omega _{\mathbf{k}}\delta \xi _{%
2142: \mathbf{k},\sigma }^{\dagger }\delta \xi _{\mathbf{k},\sigma }.
2143: \end{gather}%
2144: The first is the equation of motion of order parameter, it has the solution,%
2145: \begin{equation}
2146: \xi _{\mathbf{k}\nu }=\lambda _{\nu }\delta _{\mathbf{k,}0},
2147: \end{equation}%
2148: where $\lambda _{\nu }$ is a complex constant which does not depends on $T$.
2149: Owing to the fact: $\omega _{\mathbf{k}}\geq 0$, this solution is stable,%
2150: \begin{equation}
2151: \Delta S\geq 0. \label{EDS}
2152: \end{equation}%
2153: To determine $\lambda _{\nu }$, it is sufficient to consider the initial
2154: state at $\beta =0$ ($T\rightarrow +\infty $). Physically, any system must
2155: stay in its disordered phase at $\beta =0$. Therefore, $\lambda _{\nu }$
2156: must be zero at $\beta =0$. Since $\lambda _{\nu }$ does not depends on $%
2157: \beta $, it is always zero. That is to say,
2158: \begin{equation}
2159: \xi \left( \beta \right) =0,\text{ for }\beta \geq 0.
2160: \end{equation}%
2161: This equation shows that the photon gas in a black body can not produce BEC,
2162: and must always stay in its normal phase. Eq. (\ref{EDS}) demonstrates that
2163: the normal phase is stable. This result is also valid for the thermodynamic
2164: limit case.
2165:
2166: In short, there can not arise BEC in the photon gas in a black body, the
2167: system stays in its normal phase forever, Planck's radiation law holds at
2168: any temperature. Obviously, this conclusion agrees completely with the
2169: experiment.
2170:
2171: \subsection{The Ideal Phonon Gas in a Solid Body \label{IPGSB}}
2172:
2173: As the third system, let us consider the ideal phonon gas in a solid body,
2174: whose Hamiltonian reads,
2175: \begin{equation}
2176: H(d)=\sum_{\mathbf{k}\sigma }\omega _{\mathbf{k\sigma }}\left( d_{\mathbf{k}%
2177: \sigma }^{\dag }d_{\mathbf{k}\sigma }+\frac{1}{2}\right) , \label{Hphonon}
2178: \end{equation}%
2179: where $d_{\mathbf{k}\sigma }$ ($d_{\mathbf{k}\sigma }^{\dag }$) is the
2180: destruction (construction) operator for the phonons with momentum $\mathbf{k}
2181: $ and polarization $\sigma $, $\omega _{\mathbf{k\sigma }}=\hbar c_{\sigma }%
2182: \mathbf{k}$ represents the phonon energy with $c_{\sigma }$ the velocity of
2183: sound.
2184:
2185: By comparing Eq. (\ref{Hphonon}) with Eq. (\ref{Hphoton}), one concludes at
2186: once that the ideal phonon gas in a solid body can not produce BEC either,
2187: and must also stay in its normal phase forever.
2188:
2189: \subsection{Quantization of Dirac Field}
2190:
2191: At last, let us consider the quantization of Dirac field from the standpoint
2192: of the extended ensemble theory hypothesized in Sec. \ref{EET}. If one
2193: quantizes Dirac field using Bose-Einstein statistics, he has the following
2194: Hamiltonian \cite{Greiner},%
2195: \begin{equation}
2196: H(b;d)=\sum_{\mathbf{p,}\sigma}\omega_{\mathbf{p}}\left( b_{\mathbf{p}%
2197: ,\sigma}^{\dagger}b_{\mathbf{p},\sigma}-d_{\mathbf{p},\sigma}^{\dagger }d_{%
2198: \mathbf{p},\sigma}\right) ,
2199: \end{equation}
2200: where $\mathbf{p}$ and $\sigma$ denote momentum and spin, respectively; $b_{%
2201: \mathbf{p},\sigma}$ ($b_{\mathbf{p},\sigma}^{\dagger}$) and $d_{\mathbf{p}%
2202: \sigma}$ ($d_{\mathbf{p}\sigma}^{\dag}$) are bosonic destruction
2203: (construction) operators; and
2204: \begin{equation}
2205: \omega(\mathbf{p})=\sqrt{\left( c\mathbf{p}\right) ^{2}+\left( mc^{2}\right)
2206: ^{2}}
2207: \end{equation}
2208: with $m$ and $c$ being the electron mass and the speed of light,
2209: respectively.
2210:
2211: Evidently, this system has the phase-transition operator,%
2212: \begin{align}
2213: D(\eta ,b;\xi ,d)& =\sum_{\mathbf{p},\sigma }(\eta _{\mathbf{p}\sigma
2214: }^{\dag }b_{\mathbf{p}\sigma }+\xi _{\mathbf{p}\sigma }^{\dag }d_{\mathbf{p}%
2215: \sigma } \notag \\
2216: & +\eta _{\mathbf{p}\sigma }b_{\mathbf{p}\sigma }^{\dag }+\xi _{\mathbf{p}%
2217: \sigma }d_{\mathbf{p}\sigma }^{\dag }),
2218: \end{align}%
2219: where $\eta $ and $\xi $ are the order parameters for BEC.
2220:
2221: Eqs. (\ref{Variation}) and (\ref{Increment}) imply that%
2222: \begin{gather}
2223: \eta _{\mathbf{p}\sigma }=0, \label{Dirac1} \\
2224: \xi _{\mathbf{p}\sigma }=0, \label{Dirac2} \\
2225: \Delta S=\beta \sum_{\mathbf{p,}\sigma }\omega _{\mathbf{p}}(\delta \eta _{%
2226: \mathbf{p},\sigma }^{\dagger }\delta \eta _{\mathbf{p},\sigma }-\delta \xi _{%
2227: \mathbf{p},\sigma }^{\dagger }\delta \xi _{\mathbf{p},\sigma }).
2228: \label{Dirac3}
2229: \end{gather}%
2230: Eqs. (\ref{Dirac1}) and (\ref{Dirac2}) show that the system should always
2231: stay in its normal phase, but Eq. (\ref{Dirac3}) demonstrates that this
2232: normal phase corresponds to a saddle point, and is thus instable. Therefore,
2233: if Dirac field were quantized using Bose-Einstein statistics, the system
2234: would be instable and collapse. In other words, it is not admissible to
2235: quantize Dirac field using Bose-Einstein statistics. This conclusion is in
2236: accordance with the spin-statistics theorem in quantum field theory \cite%
2237: {Greiner}.
2238:
2239: On the other hand, if Dirac field is quantized using Fermi-Dirac statistics,
2240: its Hamiltonian will read as follows \cite{Greiner},%
2241: \begin{equation}
2242: H(b;d)=\sum_{\mathbf{p,}\sigma }\omega _{\mathbf{p}}(b_{\mathbf{p},\sigma
2243: }^{\dagger }b_{\mathbf{p},\sigma }+d_{\mathbf{p},\sigma }^{\dagger }d_{%
2244: \mathbf{p},\sigma }), \label{EPRP}
2245: \end{equation}%
2246: now $b_{\mathbf{p},\sigma }$ ($b_{\mathbf{p},\sigma }^{\dagger }$) and $d_{%
2247: \mathbf{p}\sigma }$ ($d_{\mathbf{p}\sigma }^{\dag }$) are fermionic
2248: destruction (construction) operators for electrons and positrons. According
2249: to Sec. \ref{EET}, its phase-transition operator is%
2250: \begin{align}
2251: D(\phi ,b;\zeta ,d)& =\sum_{\mathbf{k}}[(\phi _{\mathbf{k}}b_{-\mathbf{k}%
2252: \downarrow }b_{\mathbf{k}\uparrow }+\phi _{\mathbf{k}}^{\dag }b_{\mathbf{k}%
2253: \uparrow }^{\dag }b_{-\mathbf{k}\downarrow }^{\dag }) \notag \\
2254: & +(\zeta _{\mathbf{k}}d_{-\mathbf{k}\downarrow }d_{\mathbf{k}\uparrow
2255: }+\zeta _{\mathbf{k}}^{\dag }d_{\mathbf{k}\uparrow }^{\dag }d_{-\mathbf{k}%
2256: \downarrow }^{\dag })],
2257: \end{align}%
2258: where $\phi $ and $\zeta $ are order parameters. Following the arguments of
2259: Sec. \ref{IFG}, one recognizes immediately that the system is always stable
2260: in its normal phase. This holds even in the original representation of the
2261: Hamiltonian \cite{Greiner},
2262: \begin{equation}
2263: H(c)=\sum_{\mathbf{p}}\omega _{\mathbf{p}}\left( \sum_{s=1}^{2}c_{\mathbf{p}%
2264: ,s}^{\dagger }c_{\mathbf{p},s}-\sum_{s=3}^{4}c_{\mathbf{p},s}^{\dagger }c_{%
2265: \mathbf{p},s}\right) , \label{EHPR}
2266: \end{equation}%
2267: where
2268: \begin{equation}
2269: c_{\mathbf{p},s}=\left\{
2270: \begin{array}{ll}
2271: b_{\mathbf{p},s}, & s=1,2 \\
2272: d_{\mathbf{p},s-2}^{\dagger },\text{ } & s=3,4,%
2273: \end{array}%
2274: \right.
2275: \end{equation}%
2276: \vspace*{-0.3cm}%
2277: \begin{align}
2278: D(\phi ,\zeta ,c)& =\sum_{\mathbf{k}}[(\phi _{\mathbf{k}}c_{-\mathbf{k,}2}c_{%
2279: \mathbf{k,}1}+\phi _{\mathbf{k}}^{\dag }c_{\mathbf{k,}1}^{\dag }c_{-\mathbf{k%
2280: },2}^{\dag }) \notag \\
2281: & +(\zeta _{\mathbf{k}}c_{-\mathbf{k},4}c_{\mathbf{k},3}+\zeta _{\mathbf{k}%
2282: }^{\dag }c_{\mathbf{k},3}^{\dag }c_{-\mathbf{k},4}^{\dag })].
2283: \end{align}%
2284: In contrast to the electron-positron representation of Eq. (\ref{EPRP}), a
2285: minus sign appears in the original representation of Eq. (\ref{EHPR}), but
2286: it can not influence the stability of the system. Anyway, there is no
2287: problem when Dirac field is quantized using Fermi-Dirac statistics.
2288:
2289: In this section, the systems concerned are free boson fields, they can all
2290: be solved rigorously. We shall proceed to the interacting systems in Sec. %
2291: \ref{SPT}. Generally, they can not be solved rigorously, approximations are
2292: needed.\vspace*{0.2in}
2293:
2294: \section{Physical Origination for Phase Transitions \label{POPT}}
2295:
2296: Before proceeding forward to the interacting Bose systems, it is worth
2297: making a research into the physical origination for phase transitions.
2298:
2299: Again, let us take BCS superconductivity as an instance. According to Eq. (%
2300: \ref{Average}), the statistical average of a one-body operator can be
2301: reformulated as
2302: \begin{eqnarray}
2303: \int d\mathbf{r}\langle \psi ^{\dagger }(\mathbf{r})f(\mathbf{r})\psi (%
2304: \mathbf{r})\rangle &=&\int \mathrm{d}\mathbf{r}\mathrm{Tr}\big(\widetilde{%
2305: \psi }^{\dagger }(\mathbf{r})f(\mathbf{r})\widetilde{\psi }(\mathbf{r})\rho
2306: (H(c))\big) \notag \\
2307: &\equiv &\int \mathrm{d}\mathbf{r}\overline{\widetilde{\psi }^{\dagger }(%
2308: \mathbf{r})f(\mathbf{r})\widetilde{\psi }(\mathbf{r})}, \label{Single}
2309: \end{eqnarray}%
2310: where $\psi (\mathbf{r})$ is the electron field, $f(\mathbf{r})$ stands for
2311: the one-body operator in Schr\"{o}dinger picture, and
2312: \begin{equation}
2313: \widetilde{\psi }(\mathbf{r})=e^{-iD(\phi ,c)}\psi (\mathbf{r})e^{iD(\phi
2314: ,c)}.
2315: \end{equation}%
2316: After the phase transition ($\phi \neq 0$), the electron field $\widetilde{%
2317: \psi }(\mathbf{r})$ is separated into two fields,
2318: \begin{equation}
2319: \widetilde{\psi }(\mathbf{r})=\psi (\mathbf{r})+\varphi (\mathbf{r}),
2320: \label{PsiPrime}
2321: \end{equation}%
2322: where
2323: \begin{equation}
2324: \varphi (\mathbf{r})=e^{-iD(\phi ,c)}\psi (\mathbf{r})e^{iD(\phi ,c)}-\psi (%
2325: \mathbf{r}).
2326: \end{equation}%
2327: Substituting Eq. (\ref{PsiPrime}) into Eq. (\ref{Single}), we obtain
2328: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}} }%
2329: %BeginExpansion
2330: \begin{widetext}
2331: %EndExpansion
2332: \begin{equation}
2333: \int \mathrm{d}\mathbf{r}\langle \psi ^{\dagger }(\mathbf{r})f(\mathbf{r}%
2334: )\psi (\mathbf{r})\rangle =\int \mathrm{d}\mathbf{r}\overline{[\psi
2335: ^{\dagger }(\mathbf{r})+\varphi ^{\dagger }(\mathbf{r})]f(\mathbf{r})[\psi (%
2336: \mathbf{r})+\varphi (\mathbf{r})]}\mathbf{.} \label{Pattern1}
2337: \end{equation}%
2338: The right-hand side shows that there exists interference, as usual, it is
2339: described by the cross terms,
2340: \begin{equation}
2341: \overline{\psi ^{\dagger }(\mathbf{r})f(\mathbf{r})\varphi (\mathbf{r})}+%
2342: \overline{\varphi ^{\dagger }(\mathbf{r})f(\mathbf{r})\psi (\mathbf{r})}.
2343: \label{CrossTerms}
2344: \end{equation}%
2345: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
2346: %BeginExpansion
2347: \end{widetext}
2348: %EndExpansion
2349: In terminology of quantum optics, it represents the single-particle
2350: interference. This single-particle interference will vanish if the system
2351: goes into the normal phase where $\varphi (\mathbf{r})=0$. A special case is
2352: the interference appearing in the particle density where $f(\mathbf{r})=1$,
2353: \begin{equation}
2354: \overline{\psi ^{\dagger }(\mathbf{r})\varphi (\mathbf{r})}+\overline{%
2355: \varphi ^{\dagger }(\mathbf{r})\psi (\mathbf{r})}, \label{CrossTerms1}
2356: \end{equation}%
2357: this form of interference is very familiar in optics, Eq. (\ref{CrossTerms})
2358: being the general case.
2359:
2360: In particular, an abnormal form of single-particle interference can appear
2361: in the electron-pair amplitude,
2362: \begin{equation}
2363: \langle \psi _{\downarrow }(\mathbf{r})\psi _{\uparrow }(\mathbf{r})\rangle =%
2364: \overline{[\psi _{\downarrow }(\mathbf{r})+\varphi _{\downarrow }(\mathbf{r}%
2365: )][\psi _{\uparrow }(\mathbf{r})+\varphi _{\uparrow }(\mathbf{r})]},
2366: \end{equation}%
2367: it is described by the abnormal cross terms%
2368: \begin{equation}
2369: \overline{\psi _{\downarrow }(\mathbf{r})\varphi _{\uparrow }(\mathbf{r})}+%
2370: \overline{\varphi _{\downarrow }(\mathbf{r})\psi _{\uparrow }(\mathbf{r})}+%
2371: \overline{\varphi _{\downarrow }(\mathbf{r})\varphi _{\uparrow }(\mathbf{r})}%
2372: ,
2373: \end{equation}%
2374: which have no counterparts in optics, in contrast to Eq. (\ref{CrossTerms1}%
2375: ). Observe Eq. (\ref{Ctrans}), one recognizes that the field $\varphi (%
2376: \mathbf{r})$ contains two parts,%
2377: \begin{equation}
2378: \varphi (\mathbf{r})=\varphi ^{\left( +\right) }(\mathbf{r})+\varphi
2379: ^{\left( -\right) }(\mathbf{r}),
2380: \end{equation}%
2381: where $\varphi ^{\left( +\right) }(\mathbf{r})$ represents the creation
2382: part,
2383: \begin{equation}
2384: \varphi ^{\left( +\right) }(\mathbf{r})=\left[
2385: \begin{array}{l}
2386: \phi _{\downarrow }^{\dagger }(\mathbf{r}) \\
2387: \phi _{\uparrow }^{\dagger }(\mathbf{r})%
2388: \end{array}%
2389: \right] ,
2390: \end{equation}%
2391: with $\phi _{\uparrow }^{\dagger }(\mathbf{r})$ and $\phi _{\downarrow
2392: }^{\dagger }(\mathbf{r})$ being connected with the operators $c_{\mathbf{k}%
2393: \uparrow }^{\dagger }$ and $c_{-\mathbf{k}\downarrow }^{\dagger }$,
2394: respectively; and $\varphi ^{\left( -\right) }(\mathbf{r})$ the annihilation
2395: part, which is connected with the operators $c_{\mathbf{k}\uparrow }$ and $%
2396: c_{-\mathbf{k}\downarrow }$. This separation of two parts leads us to
2397: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}} }%
2398: %BeginExpansion
2399: \begin{widetext}
2400: %EndExpansion
2401: \begin{equation}
2402: \overline{\psi _{\downarrow }(\mathbf{r})\varphi _{\uparrow }(\mathbf{r})}+%
2403: \overline{\varphi _{\downarrow }(\mathbf{r})\psi _{\uparrow }(\mathbf{r})}+%
2404: \overline{\varphi _{\downarrow }(\mathbf{r})\varphi _{\uparrow }(\mathbf{r})}%
2405: =\overline{\psi _{\downarrow }(\mathbf{r})\phi _{\downarrow }^{\dagger }(%
2406: \mathbf{r})}+\overline{\phi _{\uparrow }^{\dagger }(\mathbf{r})\psi
2407: _{\uparrow }(\mathbf{r})}+\overline{\varphi _{\downarrow }^{\left( -\right)
2408: }(\mathbf{r})\phi _{\downarrow }^{\dagger }(\mathbf{r})}+\overline{\phi
2409: _{\uparrow }^{\dagger }(\mathbf{r})\varphi _{\uparrow }^{\left( -\right) }(%
2410: \mathbf{r})}.
2411: \end{equation}%
2412: Clearly, there exists interference on the right-hand side, this interference
2413: induce the abnormal interference on the left-hand side. The abnormal
2414: interference has an important consequence, that is, the electrons are formed
2415: into Cooper pairs in the superconducting phase. Evidently, the abnormal
2416: interference vanishes in the normal phase.
2417:
2418: Also, there exists the two-particle interference in the superconducting
2419: phase, which arises from the statistical average of a two-body operator,%
2420: \begin{equation}
2421: \frac{1}{2}\iint \mathrm{d}\mathbf{r}\mathrm{d}\mathbf{r}^{\prime }\langle
2422: \psi ^{\dagger }(\mathbf{r})\psi ^{\dagger }(\mathbf{r}^{\prime })v(\mathbf{r%
2423: }-\mathbf{r}^{\prime })\psi (\mathbf{r}^{\prime })\psi (\mathbf{r})\rangle =%
2424: \frac{1}{2}\iint \mathrm{d}\mathbf{r\mathrm{d}\mathbf{r}^{\prime }}v(\mathbf{%
2425: r}-\mathbf{r}^{\prime })\overline{\widetilde{\psi }^{\dagger }(\mathbf{r})%
2426: \widetilde{\psi }^{\dagger }(\mathbf{r}^{\prime })\widetilde{\psi }(\mathbf{r%
2427: }^{\prime })\widetilde{\psi }(\mathbf{r})}\mathbf{,} \label{Pattern2}
2428: \end{equation}%
2429: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
2430: %BeginExpansion
2431: \end{widetext}
2432: %EndExpansion
2433: where $v(\mathbf{r}-\mathbf{r}^{\prime })$ stands for the two-body operator
2434: in Schr\"{o}dinger picture.
2435:
2436: In the sense of Eqs. (\ref{Pattern1}) and (\ref{Pattern2}), the
2437: superconducting phase can be identified as a form of interference pattern.
2438: Physically, the onset of an interference pattern will make a system more
2439: structured, and thus increases its degree of order and decreases its
2440: entropy. That is why the superconducting phase has a higher degree of order
2441: and a less amount of entropy than the normal phase.
2442:
2443: From the above discussions, it follows that there will appear interference
2444: in the ordered phase whereas there will not in the normal (disordered)
2445: phase. Obviously, the same conclusion holds for other phase transitions.
2446: Therefore, a phase transition amounts to the onset of a form of
2447: interference. As well-known, interference is a characteristic property of
2448: wave, we thus conclude that phase transitions originate physically from the
2449: wave nature of matter.
2450:
2451: Physically, any system will go in its disordered phase at sufficiently high
2452: temperatures, or as $T\rightarrow+\infty$. Because order parameter equals
2453: zero in disordered phase, the extended ensemble theory should ensure that
2454: there exists a zero solution of order parameter, at least at high
2455: temperatures. So it does, indeed.
2456:
2457: Upon a transformation, Eq. (\ref{Entropy}) can be expressed as
2458: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}} }%
2459: %BeginExpansion
2460: \begin{widetext}
2461: %EndExpansion
2462: \begin{equation}
2463: S(\phi ,\beta )=-\mathrm{Tr}\Big(\!\ln \!\!\big(\rho (e^{-iD(\phi
2464: ,c)}H\left( c\right) e^{iD(\phi ,c)})\big)\rho (H(c))\Big).
2465: \end{equation}%
2466: Considering the expansion,%
2467: \begin{equation}
2468: e^{-iD(\phi ,c)}H\left( c\right) e^{iD(\phi ,c)}=H\left( c\right) -i[D(\phi
2469: ,c),H\left( c\right) ]+\frac{\left( -i\right) ^{2}}{2}[D(\phi ,c),[D(\phi
2470: ,c),H\left( c\right) ]]+\cdots ,
2471: \end{equation}%
2472: we obtain%
2473: \begin{equation}
2474: S(\phi ,\beta )=-\mathrm{Tr}\Big(\ln \!\big(\rho \left( H\left( c\right)
2475: \right) \big)\rho (H(c))\Big)+\beta \frac{\left( -i\right) ^{2}}{2}\mathrm{Tr%
2476: }\Big([D(\phi ,c),[D(\phi ,c),H\left( c\right) ]]\rho (H(c))\Big)+\cdots ,
2477: \end{equation}%
2478: the linear term of $\phi $ vanishes.
2479:
2480: Apart from an irrelevant term, this equation shows that the powers of $\phi $
2481: are equal to or higher than the $2$nd in the expansion of $S(\phi ,\beta )$.
2482: As a result of $\delta S=0$, the powers of $\phi $ are equal to or higher
2483: than the $1$st in the equation of order parameter. It implies that there
2484: always exists a zero solution of $\phi $ at any temperature, i.e., the
2485: trivial solution. In fact, this solution is a stable solution as $%
2486: T\rightarrow +\infty $,%
2487: \begin{equation}
2488: \Delta S=S(\delta \phi ,\beta )-S(0,\beta )=-\mathrm{Tr}\Big(\ln \!\big(\rho
2489: (H(c))\big)\left[ \rho (H^{\prime }(\delta \phi ,c))-\rho (H(c))\right] \Big)%
2490: =0,\text{ }T\rightarrow +\infty .
2491: \end{equation}%
2492: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
2493: %BeginExpansion
2494: \end{widetext}
2495: %EndExpansion
2496: which satisfies the criterion of Eq. (\ref{Increment}). Of course, this
2497: conclusion is also valid for other phase transitions.
2498:
2499: To summarize, there exists a zero solution of order parameter at any
2500: temperature in the extended ensemble theory, this solution becomes stable as
2501: $T\rightarrow +\infty $. Therefore, a system always stays in its normal
2502: phase at sufficiently high temperatures, this high-temperature phase is
2503: disordered and structureless. As temperature decreases, there can arise
2504: spontaneously the quantum interference of matter waves, which makes the
2505: system structured and transform into its ordered phase. That is the physical
2506: picture for phase transitions within the framework of the extended ensemble
2507: theory.
2508:
2509: \textbf{Remark}: In Ref. \cite{Yang5}, Yang suggested an important concept,
2510: i.e., off-diagonal long-range order (ODLRO), to describe quantum phase
2511: transitions. Physically, ODLRO is a generalization of the concept of Copper
2512: pairing. With this generalized concept, he developed many beautiful
2513: theorems, which are characteristics of quantum phases. Regrettably, in that
2514: paper, there is no discussion of the properties of the Hamiltonian that is
2515: needed to ensure the existence of ODLRO at low temperatures. Evidently,
2516: there exists no ODLRO if a system Hamiltonian is in its symmetric
2517: representation. The requirement can be satisfied if the system Hamiltonian
2518: takes its asymmetric, or broken-symmetry, representation. Therefore, those
2519: theorems in Ref. \cite{Yang5} still hold in the extended ensemble theory. As
2520: a matter of fact, the ODLRO's of the reduced density matrices $\rho _{1}$
2521: and $\rho _{2}$ \cite{Yang5}, e.g., $\mathrm{Tr}(c_{i\sigma }\rho (H^{\prime
2522: }(\phi ,c))c_{j\sigma }^{\dagger })$ and $\mathrm{Tr}(c_{k\sigma }c_{l\sigma
2523: }\rho (H^{\prime }(\phi ,c))c_{j\sigma }^{\dagger }c_{i\sigma }^{\dagger })$%
2524: , result physically from the single-particle and two-particle interferences,
2525: respectively. That is to say, off-diagonal long-range order is a physical
2526: manifestation of the quantum interference of matter waves.
2527:
2528: \section{\textbf{Interacting Bose Systems} \label{SPT}}
2529:
2530: In this section, we focus our attention on the interacting Bose systems. We
2531: shall first study the weakly interacting Bose gas, and then the double-well
2532: potential systems where structural phase transitions, Goldstone bosons, and
2533: Higgs mechanism are concerned.\vspace*{0.2in}
2534:
2535: \subsection{Weakly Interacting Bose Gas \label{WIBG}}
2536:
2537: As usual, the Hamiltonian for an interacting Bose gas reads,
2538: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}} }%
2539: %BeginExpansion
2540: \begin{widetext}
2541: %EndExpansion
2542: \begin{equation}
2543: H[\psi ]=\int \mathrm{d}\mathbf{r}\left[ \frac{\hbar ^{2}}{2m}\nabla \psi
2544: ^{\dagger }(\mathbf{r})\cdot \nabla \psi (\mathbf{r})-\mu \psi ^{\dagger }(%
2545: \mathbf{r})\psi (\mathbf{r})\right] +\frac{1}{2}\iint \mathrm{d}\mathbf{r}%
2546: \mathrm{d}\mathbf{r}^{\prime }\psi ^{\dagger }(\mathbf{r})\psi ^{\dagger }(%
2547: \mathbf{r}^{\prime })v(\mathbf{r}-\mathbf{r}^{\prime })\psi (\mathbf{r}%
2548: ^{\prime })\psi (\mathbf{r}),
2549: \end{equation}%
2550: where $\psi (\mathbf{r})$ represents the boson-field operator, and $v(%
2551: \mathbf{r}-\mathbf{r}^{\prime })$ the interaction. Expressed in terms of $%
2552: \psi (\mathbf{r})$ and $\psi ^{\dagger }(\mathbf{r})$, the phase-transition
2553: operator for BEC becomes,
2554: \begin{equation}
2555: D\left[ \eta ,\psi \right] =\int \mathrm{d}\mathbf{r}\left[ \eta ^{\dagger }(%
2556: \mathbf{r})\psi (\mathbf{r})+\eta (\mathbf{r})\psi ^{\dagger }(\mathbf{r})%
2557: \right] .
2558: \end{equation}
2559:
2560: With $H[\psi ]$ and $D\left[ \eta ,\psi \right] $, the entropy of the system
2561: can be obtained,
2562: \begin{align}
2563: S\left[ \eta ,\beta \right] & =-\mathrm{Tr}\big(\ln \left( \rho (H\left[
2564: \psi \right] )\right) \!\rho (H\left[ \psi \right] )\big) \notag \\
2565: & +\beta \bigg\{\int \mathrm{d}\mathbf{r}\left[ \frac{\hbar ^{2}}{2m}\nabla
2566: \eta ^{\dagger }(\mathbf{r})\cdot \nabla \eta (\mathbf{r})-\mu \eta
2567: ^{\dagger }(\mathbf{r})\eta (\mathbf{r})\right] +\iint \mathrm{d}\mathbf{r}%
2568: \mathrm{d}\mathbf{r}^{\prime }\eta ^{\dagger }(\mathbf{r})\eta (\mathbf{r})v(%
2569: \mathbf{r}-\mathbf{r}^{\prime })\overline{\psi ^{\dagger }(\mathbf{r}%
2570: ^{\prime })\psi (\mathbf{r}^{\prime })} \notag \\
2571: & +\iint \mathrm{d}\mathbf{r}\mathrm{d}\mathbf{r}^{\prime }\eta ^{\dagger }(%
2572: \mathbf{r})\eta (\mathbf{r}^{\prime })v(\mathbf{r}-\mathbf{r}^{\prime })%
2573: \overline{\psi ^{\dagger }(\mathbf{r}^{\prime })\psi (\mathbf{r})}+\frac{1}{2%
2574: }\iint \mathrm{d}\mathbf{r}\mathrm{d}\mathbf{r}^{\prime }\eta ^{\dagger }(%
2575: \mathbf{r})\eta (\mathbf{r})v(\mathbf{r}-\mathbf{r}^{\prime })\eta ^{\dagger
2576: }(\mathbf{r}^{\prime })\eta (\mathbf{r}^{\prime })\bigg\}, \label{IBECS}
2577: \end{align}%
2578: where $\overline{F}$ denotes the statistical average of the operator $F$
2579: with respect to $H[\psi ]$. Its variation yields the equation of order
2580: parameter,%
2581: \begin{equation}
2582: \left( -\frac{\hbar ^{2}}{2m}\nabla ^{2}-\mu \right) \eta (\mathbf{r})+\int d%
2583: \mathbf{r}^{\prime }\eta (\mathbf{r})v(\mathbf{r}-\mathbf{r}^{\prime })%
2584: \overline{\psi ^{\dagger }(\mathbf{r}^{\prime })\psi (\mathbf{r}^{\prime })}%
2585: +\int \mathrm{d}\mathbf{r}^{\prime }\eta (\mathbf{r}^{\prime })v(\mathbf{r}-%
2586: \mathbf{r}^{\prime })\overline{\psi ^{\dagger }(\mathbf{r}^{\prime })\psi (%
2587: \mathbf{r})}+\int \mathrm{d}\mathbf{r}^{\prime }\eta (\mathbf{r})v(\mathbf{r}%
2588: -\mathbf{r}^{\prime })\eta ^{\dagger }(\mathbf{r}^{\prime })\eta (\mathbf{r}%
2589: ^{\prime })=0. \label{IOP}
2590: \end{equation}%
2591: That is a generalized Ginzburg-Landau or Gross-Pitaevskii equation. In the
2592: simple case where $v(\mathbf{r}-\mathbf{r}^{\prime })=g\delta (\mathbf{r}-%
2593: \mathbf{r}^{\prime })$, it reduces to%
2594: \begin{equation}
2595: -\frac{\hbar ^{2}}{2m}\nabla ^{2}\eta (\mathbf{r})+\left[ 2g\overline{\psi
2596: ^{\dagger }(\mathbf{r})\psi (\mathbf{r})}-\mu \right] \eta (\mathbf{r}%
2597: )+g\left\vert \eta (\mathbf{r})\right\vert ^{2}\eta (\mathbf{r})=0,
2598: \end{equation}%
2599: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
2600: %BeginExpansion
2601: \end{widetext}
2602: %EndExpansion
2603: which is the standard Ginzburg-Landau \cite{GL} or Gross-Pitaevskii \cite%
2604: {GP1,GP2} equation. In addition to Eq. (\ref{IOP}), one also needs%
2605: \begin{equation}
2606: \int \mathrm{d}\mathbf{r}\eta ^{\dagger }(\mathbf{r})\eta (\mathbf{r})+\int
2607: \mathrm{d}\mathbf{r}\overline{\psi ^{\dagger }(\mathbf{r})\psi (\mathbf{r})}%
2608: =N, \label{ICP}
2609: \end{equation}%
2610: which is the equation of chemical potential. Eq. (\ref{IOP}) and Eq. (\ref%
2611: {ICP}) constitute the two basic equations for an interacting Bose system.
2612:
2613: For simplicity, let us first consider the homogeneous solution,%
2614: \begin{equation}
2615: \eta (\mathbf{r})=\xi , \label{HomoKsi}
2616: \end{equation}%
2617: where $\xi $ is independent of $\mathbf{r}$. Accordingly, Eqs. (\ref{IBECS})
2618: and (\ref{IOP}) get simplified as%
2619: \begin{align}
2620: S\left[ \eta ,\beta \right] & =-\mathrm{Tr}\big(\ln \left( \rho (H\left[
2621: \psi \right] )\right) \!\rho (H\left[ \psi \right] )\big) \notag \\
2622: & +\beta \int \mathrm{d}\mathbf{r}[-\widetilde{\mu }\xi ^{\dagger }\xi +%
2623: \frac{1}{2}v(0)(\xi ^{\dagger }\xi )^{2}], \label{IBECS1}
2624: \end{align}%
2625: \vspace*{-0.3in}%
2626: \begin{equation}
2627: -\widetilde{\mu }\xi +v(0)\left\vert \xi \right\vert ^{2}\xi =0,
2628: \label{IOP1}
2629: \end{equation}%
2630: where%
2631: \begin{align}
2632: \widetilde{\mu }& =\mu -\int \frac{\mathrm{d}\mathbf{k}}{\left( 2\pi \right)
2633: ^{3}}\left[ v(0)+v(-\mathbf{k})\right] \notag \\
2634: & \times \left( -\frac{1}{\beta \hbar }\right) \sum_{n}e^{i\omega _{n}\eta }%
2635: \mathcal{G}(\mathbf{k},i\omega _{n}), \label{RMuW} \\
2636: v(\mathbf{k})& =\int \mathrm{d}\mathbf{r}v(\mathbf{r})e^{-i\mathbf{k}\cdot
2637: \mathbf{r}},
2638: \end{align}%
2639: $\mathcal{G}(\mathbf{k},i\omega _{n})$ being the single-particle Green's
2640: function defined with respect to $H[\psi ]$.
2641:
2642: Eq. (\ref{IBECS1}) has the typical form of landau free energy \cite{Landau}.
2643: From Eqs. (\ref{IBECS1}), (\ref{IOP1}) and (\ref{Increment}), it follows
2644: that the interaction between two particles must be repulsive for an
2645: interacting Bose system to be stable, i.e., $v(\mathbf{r}-\mathbf{r}^{\prime
2646: })>0$. Otherwise, $v(0)<0$, the system will collapse.
2647:
2648: Suppose that the repulsive interaction is weak, we can then treat it as a
2649: perturbation. As a technique of perturbation, we shall employ GF, and adopt
2650: the self-consistent Hartree-Fock approximation as in Ref. \cite{Fetter},\
2651: \begin{align}
2652: \hspace{-1cm}\mathcal{G}(\mathbf{k},i\omega_{n}) & =\frac{1}{i\omega
2653: _{n}-\hbar^{-1}\left[ \varepsilon\left( \mathbf{k}\right) -\mu-\hbar
2654: \Sigma^{\star}\left( \mathbf{k}\right) \right] }, \label{IBECHF1} \\
2655: \hspace{-1cm}\hbar\Sigma^{\star}\left( \mathbf{k}\right) & =\int \frac{%
2656: \mathrm{d}\mathbf{k}^{\prime}}{\left( 2\pi\right) ^{3}}\left[ v(0)+v(\mathbf{%
2657: k}-\mathbf{k}^{\prime})\right] \notag \\
2658: & \times\left( -\frac{1}{\beta\hbar}\right) \sum_{n}e^{i\omega_{n}\eta }%
2659: \mathcal{G}(\mathbf{k}^{\prime},i\omega_{n}), \label{IBECHF2}
2660: \end{align}
2661: where $\hbar\Sigma^{\star}\left( \mathbf{k}\right) $ represents the proper
2662: self-energy.\ By use of Eq. (\ref{RMuW}), $\mathcal{G}(\mathbf{k},i\omega
2663: _{n})$ can be expressed as%
2664: \begin{equation}
2665: \mathcal{G}(\mathbf{k},i\omega_{n})=\frac{1}{i\omega_{n}-\hbar^{-1}\left[
2666: \widetilde{\varepsilon}\left( \mathbf{k}\right) -\widetilde{\mu}\right] },
2667: \label{IBECGF}
2668: \end{equation}
2669: where%
2670: \begin{align}
2671: \widetilde{\varepsilon}\left( \mathbf{k}\right) & =\varepsilon\left( \mathbf{%
2672: k}\right) +\int\frac{\mathrm{d}\mathbf{k}^{\prime}}{\left( 2\pi\right) ^{3}}%
2673: \left[ v(\mathbf{k}-\mathbf{k}^{\prime})-v(-\mathbf{k}^{\prime})\right]
2674: \notag \\
2675: & \times\left( -\frac{1}{\beta\hbar}\right) \sum_{n}e^{i\omega_{n}\eta }%
2676: \mathcal{G}(\mathbf{k}^{\prime},i\omega_{n}). \label{REnergy}
2677: \end{align}
2678: Eq. (\ref{IBECGF}) shows that $\widetilde{\varepsilon}\left( \mathbf{k}%
2679: \right) $ and $\widetilde{\mu}$ are the renormalized energy and chemical
2680: potential, respectively.
2681:
2682: With the help of Eqs. (\ref{HomoKsi}) and (\ref{IBECGF}), Eq. (\ref{ICP})
2683: can be reduced as%
2684: \begin{equation}
2685: \xi^{\dagger}\xi+\int\frac{\mathrm{d}\mathbf{k}}{\left( 2\pi\right) ^{3}}%
2686: \left( -\frac{1}{\beta\hbar}\right) \sum_{n}e^{i\omega_{n}\eta}\mathcal{G}(%
2687: \mathbf{k},i\omega_{n})=n, \label{ICP1}
2688: \end{equation}
2689: where $n=N/V$ is the particle density of the system.
2690:
2691: When $\widetilde{\mu}<0$, Eq. (\ref{IOP1}) has the only solution $\xi=0$. As
2692: shown by Eq. (\ref{IBECS1}), it is a stable solution. In this case, the
2693: system is in its normal phase. The renormalized chemical potential $%
2694: \widetilde{\mu}$ is given by Eq. (\ref{ICP1}),
2695: \begin{equation}
2696: \mathcal{P}\int_{-\infty}^{+\infty}\mathrm{d}\omega\mathcal{N}(\omega ,%
2697: \widetilde{\mu})\frac{1}{e^{\beta\omega}-1}=n, \label{CPGTc}
2698: \end{equation}
2699: where $\mathcal{P}$ denotes the principal value introduced in Appendix \ref%
2700: {FSum}, and $\mathcal{N}(\omega,\widetilde{\mu})$ the density of states,
2701: \begin{equation}
2702: \mathcal{N}(\omega,\widetilde{\mu})=-\int\frac{\mathrm{d}\mathbf{k}}{\left(
2703: 2\pi\right) ^{3}}\mathrm{Im}\mathcal{G}(\mathbf{k},\omega+i0^{+}).
2704: \label{IBECDOS}
2705: \end{equation}
2706: If $\widetilde{\mu}>0$, Eqs. (\ref{IBECS1}) and (\ref{IOP1}) show that the
2707: system will transform into a condensation phase where
2708: \begin{equation}
2709: \left\vert \xi\right\vert ^{2}=\frac{\widetilde{\mu}}{v(0)}.
2710: \end{equation}
2711: Substitution of it into Eq. (\ref{ICP1}) yields
2712: \begin{equation}
2713: \frac{\widetilde{\mu}}{v(0)}+\mathcal{P}\int_{-\infty}^{+\infty}\mathrm{d}%
2714: \omega\mathcal{N}(\omega,\widetilde{\mu})\frac{1}{e^{\beta\omega}-1}=n,
2715: \label{CPLTc}
2716: \end{equation}
2717: which gives the renormalized chemical potential $\widetilde{\mu}$ at $%
2718: T<T_{c} $, where $T_{c}$ is the critical temperature for BEC,%
2719: \begin{equation}
2720: \mathcal{P}\int_{-\infty}^{+\infty}\mathrm{d}\omega\mathcal{N}(\omega ,0)%
2721: \frac{1}{e^{\beta_{c}\omega}-1}=n, \label{ITc}
2722: \end{equation}
2723: that is to say, $\widetilde{\mu}(T_{c})=0$. In sum, Eqs. (\ref{CPGTc}) and (%
2724: \ref{CPLTc}) yield together the renormalized chemical potential of the
2725: system at $T>T_{c}$ and $T<T_{c}$, respectively.
2726:
2727: In order to deduce $\widetilde{\mu }$ and $\xi $ in more detail, let us
2728: suppose further that $v(\mathbf{r}-\mathbf{r}^{\prime })$ is of short range.
2729: Following Ref. \cite{Fetter}, we can approximate $v(\mathbf{k})$ as%
2730: \begin{equation}
2731: v(\mathbf{k})=v(0)\left[ 1-\frac{1}{6}(ka)^{2}\right] , \label{IBECSR}
2732: \end{equation}%
2733: where%
2734: \begin{equation}
2735: a^{2}=\frac{\int d\mathbf{r\,r}^{2}v(\mathbf{r})}{\int d\mathbf{r\,}v(%
2736: \mathbf{r})}.
2737: \end{equation}
2738:
2739: Inserting Eq. (\ref{IBECSR}) into Eq. (\ref{REnergy}) and Eq. (\ref{RMuW})
2740: leads us to%
2741: \begin{align}
2742: \widetilde{\varepsilon }\left( \mathbf{k}\right) & =\frac{\hbar ^{2}\mathbf{k%
2743: }^{2}}{2m^{\star }}, \\
2744: \widetilde{\mu }& =\mu -\left[ 2v(0)-\frac{v(0)nma^{2}}{3\hbar ^{2}}\frac{%
2745: m^{\star }}{m}\right] \notag \\
2746: & \times \mathcal{P}\int_{-\infty }^{+\infty }\mathrm{d}\omega \mathcal{N}%
2747: (\omega ,\widetilde{\mu })\frac{\left( \omega +\widetilde{\mu }\right) }{%
2748: e^{\beta \omega }-1}, \label{IApprox2}
2749: \end{align}%
2750: where $m^{\star }$ is the renormalized mass,%
2751: \begin{equation}
2752: \frac{1}{m^{\star }}=\frac{1}{m}\left[ 1-\frac{v(0)nma^{2}}{3\hbar ^{2}}%
2753: \left( 1-\frac{\left\vert \xi \right\vert ^{2}}{n}\right) \right] ,
2754: \label{IApprox3}
2755: \end{equation}%
2756: and $\mathcal{N}(\omega ,\widetilde{\mu })$ the density of states defined by
2757: Eq. (\ref{IBECDOS}),
2758: \begin{equation}
2759: \mathcal{N}(\omega ,\widetilde{\mu })=\left\{
2760: \begin{array}{ll}
2761: \frac{1}{4\pi ^{2}\hbar ^{3}}(2m^{\star })^{3/2}\left( \omega +\widetilde{%
2762: \mu }\right) ^{1/2},\text{ } & \omega \geq -\widetilde{\mu } \\
2763: 0,\text{ } & \omega <-\widetilde{\mu }.%
2764: \end{array}%
2765: \right. \label{IApprox4}
2766: \end{equation}
2767:
2768: From Eqs. (\ref{ITc}), (\ref{IApprox3}) and (\ref{IApprox4}), it follows that%
2769: \begin{equation}
2770: \frac{T_{c}}{T_{0}}=1-\frac{v(0)nma^{2}}{3\hbar^{2}}, \label{LowTc}
2771: \end{equation}
2772: where $T_{0}$ stands for the critical temperature of the BEC happening in
2773: the corresponding ideal Bose gas,
2774: \begin{equation}
2775: T_{0}=\frac{2\pi\hbar^{2}}{mk_{B}}\left( \frac{n}{\zeta\left( 3/2\right) }%
2776: \right) ^{2/3}.
2777: \end{equation}
2778: Eq. (\ref{LowTc}) manifests that a repulsive interaction will lower the
2779: transition temperature of BEC,%
2780: \begin{equation}
2781: \gamma\equiv\frac{\Delta T}{T_{0}}=\frac{T_{c}-T_{0}}{T_{0}}=-\frac {%
2782: v(0)nma^{2}}{3\hbar^{2}}<0.
2783: \end{equation}
2784:
2785: Also, we have by Eq. (\ref{CPGTc}),%
2786: \begin{equation}
2787: \frac{2}{\sqrt{\pi }\zeta \left( 3/2\right) }t^{3/2}\int_{0}^{+\infty }%
2788: \mathrm{d}x\frac{x^{1/2}}{e^{x-\overline{\mu }/t}-1}=1, \label{MuGTc}
2789: \end{equation}%
2790: where $t=T/T_{c}$ and $\overline{\mu }=\widetilde{\mu }/\left(
2791: k_{B}T_{c}\right) $. This equation gives the $\widetilde{\mu }$ at $T>T_{c}$%
2792: . If $T<T_{c}$, Eq. (\ref{CPLTc}) must be used in place of Eq. (\ref{CPGTc}),%
2793: \begin{align}
2794: 1& =y+\frac{2}{\sqrt{\pi }\zeta \left( 3/2\right) }\left( \frac{1+\gamma }{%
2795: 1+\gamma -\gamma y}\right) ^{3/2} \notag \\
2796: & \times \mathcal{P}\int_{0}^{+\infty }\mathrm{d}x\frac{x^{1/2}}{e^{\left(
2797: x-\nu y\right) /t}-1}, \label{MuLTc}
2798: \end{align}%
2799: where $y=\widetilde{\mu }/\left( v(0)n\right) $, and $\nu =v(0)n/\left(
2800: k_{B}T_{c}\right) $. This equation will yield both the $\widetilde{\mu }$
2801: and $\left\vert \xi \right\vert ^{2}=\widetilde{\mu }/v(0)$ at $T<T_{c}$.
2802:
2803: %TCIMACRO{%
2804: %\TeXButton{Fig4}{\begin{figure*}[hbtp]
2805: %\includegraphics[scale=0.6,angle=-90]{Fig4.eps}
2806: %\caption{
2807: %The chemical potential of the system versus
2808: %temperature, where $\overline{\mu}=\widetilde{\mu}/\left(
2809: %k_{B}T_{c}\right) $, $t=T/T_{c}$, $\nu=v(0)n/\left( k_{B}T_{c}\right) $,
2810: %and $\gamma=\Delta T/T_{0}=-v(0)nma^{2}/\left( 3\hbar^{2}\right) $.
2811: %For the sake of contrast, the chemical potential of the ideal Bose gas
2812: %($\nu=0$, $\gamma=0$) is also plotted.
2813: %}
2814: %\end{figure*}}}%
2815: %BeginExpansion
2816: \begin{figure*}[hbtp]
2817: \includegraphics[scale=0.6,angle=-90]{Fig4.eps}
2818: \caption{
2819: The chemical potential of the system versus
2820: temperature, where $\overline{\mu}=\widetilde{\mu}/\left(
2821: k_{B}T_{c}\right) $, $t=T/T_{c}$, $\nu=v(0)n/\left( k_{B}T_{c}\right) $,
2822: and $\gamma=\Delta T/T_{0}=-v(0)nma^{2}/\left( 3\hbar^{2}\right) $.
2823: For the sake of contrast, the chemical potential of the ideal Bose gas
2824: ($\nu=0$, $\gamma=0$) is also plotted.
2825: }
2826: \end{figure*}%
2827: %EndExpansion
2828:
2829: The numerical results of $\widetilde{\mu }$ and $\left\vert \xi \right\vert
2830: ^{2}$ are shown in Figs. 4 and 5, respectively. We observe that the
2831: renormalized chemical potential $\widetilde{\mu }$ is a monotonically
2832: decreasing function of temperature, at $T=T_{c}$ itself continuous but its
2833: derivative not. This agrees with the $T\geq T_{c}$ result given by Ref. \cite%
2834: {Fetter} and the references therein. When $T<T_{c}$, the chemical potential $%
2835: \widetilde{\mu }$ is set zero in Ref. \cite{Fetter}, as Einstein \cite%
2836: {Einstein1,Einstein2} did in the ideal Bose gas. In the extended ensemble
2837: theory, there is no mathematical limit on $\widetilde{\mu }$, it will change
2838: with temperature even if $T<T_{c}$ so as to ensure the conservation of
2839: particles, as pointed out in Appendix \ref{FSum}. The behavior of $%
2840: \left\vert \xi \right\vert ^{2}$ shown in Fig. 5 is just expected for an
2841: order parameter.
2842:
2843: %TCIMACRO{%
2844: %\TeXButton{Fig5}{\begin{figure*}[htbp]
2845: %\includegraphics[scale=0.6,angle=-90]{Fig5.eps}
2846: %\caption{
2847: %The order parameter for BEC versus temperature,
2848: %where $t=T/T_{c}$, $\nu=v(0)n/\left( k_{B}T_{c}\right) $, and $\gamma=\Delta
2849: %T/T_{0}=-v(0)nma^{2}/\left( 3\hbar^{2}\right) $.
2850: %For the sake of contrast, the order parameter for the ideal Bose gas
2851: %($\nu=0$, $\gamma=0$) is also plotted.
2852: %}
2853: %\end{figure*}}}%
2854: %BeginExpansion
2855: \begin{figure*}[htbp]
2856: \includegraphics[scale=0.6,angle=-90]{Fig5.eps}
2857: \caption{
2858: The order parameter for BEC versus temperature,
2859: where $t=T/T_{c}$, $\nu=v(0)n/\left( k_{B}T_{c}\right) $, and $\gamma=\Delta
2860: T/T_{0}=-v(0)nma^{2}/\left( 3\hbar^{2}\right) $.
2861: For the sake of contrast, the order parameter for the ideal Bose gas
2862: ($\nu=0$, $\gamma=0$) is also plotted.
2863: }
2864: \end{figure*}%
2865: %EndExpansion
2866:
2867: The internal energy of the system, $E$, can be derived conveniently from $%
2868: E=\langle H\rangle +\mu N$,
2869: \begin{align}
2870: \varepsilon & =\nu +\frac{2}{\sqrt{\pi }\zeta \left( 3/2\right) } \notag \\
2871: & \times \int_{0}^{+\infty }\mathrm{d}x\frac{x^{3/2}}{e^{\left( x-\overline{%
2872: \mu }\right) /t}-1},\text{ }T\geq T_{c}, \label{EGTc} \\
2873: \varepsilon & =\nu -\frac{\overline{\mu }^{2}}{2\nu }+\frac{2}{\sqrt{\pi }%
2874: \zeta \left( 3/2\right) }\left( \frac{1+\gamma }{1+\gamma -\gamma y}\right)
2875: ^{5/2} \notag \\
2876: & \times \mathcal{P}\int_{0}^{+\infty }\mathrm{d}x\frac{x^{3/2}}{e^{\left( x-%
2877: \overline{\mu }\right) /t}-1},\text{ }T\leq T_{c}, \label{ELTc}
2878: \end{align}%
2879: where%
2880: \begin{equation}
2881: \varepsilon =\frac{1}{k_{B}T_{c}}\frac{E}{N}.
2882: \end{equation}%
2883: Combination of Eqs. (\ref{MuGTc}) and (\ref{EGTc}) gives $C_{V}$, the
2884: specific heat at constant volume,
2885: \begin{equation}
2886: \frac{C_{V}}{Nk_{B}}=\frac{15}{4}\frac{g_{5/2}(z)}{g_{3/2}(z)}-\frac{9}{4}%
2887: \frac{g_{3/2}(z)}{g_{1/2}(z)},\text{ }T\geq T_{c},
2888: \end{equation}%
2889: where%
2890: \begin{gather}
2891: z=e^{\overline{\mu }/t}, \\
2892: g_{n}(z)=\frac{1}{\Gamma (n)}\int_{0}^{+\infty }\mathrm{d}x\frac{x^{n-1}}{%
2893: z^{-1}e^{x}-1}.
2894: \end{gather}%
2895: This result is the same in form as that for the ideal Bose gas, which is
2896: familiar in standard books on quantum statistical mechanics. If $T\leq T_{c}$%
2897: , $C_{V}$ must be calculated from Eqs. (\ref{MuLTc}) and (\ref{ELTc}).
2898:
2899: %TCIMACRO{%
2900: %\TeXButton{Fig6}{\begin{figure*}[htbp]
2901: %\hspace*{-0.6cm}\includegraphics[scale=0.6,angle=-90]{Fig6.eps}
2902: %\caption{
2903: %The specific heat of a weakly interacting Bose gas
2904: %versus temperature, where $t=T/T_{c}$, $\nu=v(0)n/\left( k_{B}T_{c}\right) $,
2905: %and $\gamma=\Delta T/T_{0}=-v(0)nma^{2}/\left( 3\hbar^{2}\right) $.
2906: %For the sake of contrast, the specific heat of the ideal Bose gas
2907: %($\nu=0$, $\gamma=0$) is also plotted.
2908: %}
2909: %\end{figure*}}}%
2910: %BeginExpansion
2911: \begin{figure*}[htbp]
2912: \hspace*{-0.6cm}\includegraphics[scale=0.6,angle=-90]{Fig6.eps}
2913: \caption{
2914: The specific heat of a weakly interacting Bose gas
2915: versus temperature, where $t=T/T_{c}$, $\nu=v(0)n/\left( k_{B}T_{c}\right) $,
2916: and $\gamma=\Delta T/T_{0}=-v(0)nma^{2}/\left( 3\hbar^{2}\right) $.
2917: For the sake of contrast, the specific heat of the ideal Bose gas
2918: ($\nu=0$, $\gamma=0$) is also plotted.
2919: }
2920: \end{figure*}%
2921: %EndExpansion
2922:
2923: The numerical results of $C_{V}$ are shown in Fig. 6. The shape of $C_{V}$
2924: shows that the BEC occurring in the interacting Bose gas is a
2925: \textquotedblleft $\lambda $\textquotedblright -transition, as is naturally
2926: expected. In contrast to the ideal Bose gas, the weakly interacting Bose gas
2927: manifests itself with two new features: First, $C_{V}$ is discontinuous at $%
2928: T=T_{c}$; and second, $C_{V}$ is linear in temperature when $T\ll T_{c}$.
2929: Evidently, this theoretical result for $C_{V}$ meets the requirement of the
2930: third law of thermodynamics: $C_{V}\rightarrow 0$ as $T\rightarrow 0$. The
2931: discontinuity at $T=T_{c}$ is obviously due to the discontinuity of the
2932: derivatives of $\widetilde{\mu }$ and $\left\vert \xi \right\vert ^{2}$ with
2933: respect to $T$. As to the linearity at low temperatures, it can be explained
2934: as follows.
2935:
2936: Observing that
2937: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}}}%
2938: %BeginExpansion
2939: \begin{widetext}%
2940: %EndExpansion
2941: \begin{equation}
2942: \mathcal{P}\int_{0}^{+\infty }\mathrm{d}x\frac{x^{1/2}}{e^{\left( x-\nu
2943: y\right) /t}-1}=-\int_{0}^{\nu y}x^{1/2}\mathrm{d}x+t\int_{0}^{+\infty }%
2944: \mathrm{d}x\frac{\left( 2\nu y+tx\right) ^{1/2}}{e^{x+\nu y/t}-1}%
2945: +t\int_{0}^{\nu y/t}\mathrm{d}x\frac{\left( \nu y+tx\right) ^{1/2}-\left(
2946: \nu y-tx\right) ^{1/2}}{e^{x}-1},
2947: \end{equation}%
2948: we obtain%
2949: \begin{equation}
2950: \mathcal{P}\int_{0}^{+\infty }\mathrm{d}x\frac{x^{1/2}}{e^{\left( x-\nu
2951: y\right) /t}-1}=-\frac{2}{3}\left( \nu y\right) ^{3/2}+\zeta (2)\left( \nu
2952: y\right) ^{-1/2}t^{2}, \label{X12LTc}
2953: \end{equation}%
2954: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
2955: %BeginExpansion
2956: \end{widetext}
2957: %EndExpansion
2958: to the second power of $t$ as $t\ll 1$. Inserting it into Eq. (\ref{MuLTc}),
2959: we have
2960: \begin{align}
2961: y& =1-\frac{2}{\sqrt{\pi }\zeta \left( 3/2\right) }\left( \frac{1+\gamma }{%
2962: 1+\gamma -\gamma y}\right) ^{3/2} \notag \\
2963: & \times \left[ -\frac{2}{3}\left( \nu y\right) ^{3/2}+\zeta (2)\left( \nu
2964: y\right) ^{-1/2}t^{2}\right] , \label{YLTc}
2965: \end{align}%
2966: it is an iterative equation of $y$. At absolute zero ($t=0$), it reduces to%
2967: \begin{equation}
2968: y_{0}=1+\frac{4}{3\sqrt{\pi }\zeta \left( 3/2\right) }\left( \frac{1+\gamma
2969: }{1+\gamma -\gamma y_{0}}\right) ^{3/2}\left( \nu y_{0}\right) ^{3/2}.
2970: \end{equation}%
2971: This gives the zeroth order solution of $y$, which we denoted by $y_{0}$.
2972: Iterating of Eq. (\ref{YLTc}) to the second power of $t$ leads to%
2973: \begin{equation}
2974: y=y_{0}-\frac{2\zeta (2)}{\sqrt{\pi }\zeta \left( 3/2\right) }\left( \frac{%
2975: 1+\gamma }{1+\gamma -\gamma y_{0}}\right) ^{3/2}\left( \nu y_{0}\right)
2976: ^{-1/2}t^{2}, \label{YLTc1}
2977: \end{equation}%
2978: which shows that $y$, i.e., $\widetilde{\mu }$ and $\left\vert \xi
2979: \right\vert ^{2}$, decreases as $T^{2}$ in the low temperature region ($T\ll
2980: T_{c}$), as can also be seen directly from Figs. 4 and 5. With the same
2981: procedure, one finds%
2982: \begin{equation}
2983: \mathcal{P}\int_{0}^{+\infty }\mathrm{d}x\frac{x^{3/2}}{e^{\left( x-\nu
2984: y\right) /t}-1}=-\frac{2}{5}\left( \nu y\right) ^{5/2}+3\zeta (2)\left( \nu
2985: y\right) ^{1/2}t^{2}. \label{X32LTc}
2986: \end{equation}%
2987: From Eqs. (\ref{YLTc1}), (\ref{X32LTc}), and (\ref{ELTc}), it follows that $%
2988: \varepsilon $ increases as $T^{2}$. Therefore, the specific heat $C_{V}$
2989: will be linear in temperature $T$ when $T\ll T_{c}$.
2990:
2991: If one expands Eqs. (\ref{X12LTc}) and (\ref{X32LTc}) further to the fourth
2992: power of $t$, he arrives at%
2993: \begin{equation}
2994: C_{V}=a(V)T+b(V)T^{3}, \label{CvTV}
2995: \end{equation}
2996: where the expansion coefficients $a(V)$ and $b(V)$ are functions of $V$.
2997: Appendix \ref{TAWIBSLT} shows that this expansion does not depend on the
2998: approximations used here, it is valid provided that the interaction is weak.
2999: Therefore, it is a fundamental property of the weakly interacting Bose gas
3000: that the specific heat $C_{V}$ vanishes linearly as $T\rightarrow0$.
3001:
3002: In connection with the $\lambda$-transition occurring in liquid $^{4}\mathrm{%
3003: He}$, obviously, the shape of $C_{V}$ agrees well with that of the specific
3004: heat observed experimentally \cite{Keesom}. However, the latter is $C_{P}$,
3005: the specific heat at constant pressure rather than at constant volume. As is
3006: well known from thermodynamics, they differ from each other by a
3007: temperature-dependent term,%
3008: \begin{equation}
3009: C_{P}=C_{V}+T\left( \frac{\partial p}{\partial T}\right) _{V}\left( \frac{%
3010: \partial V}{\partial T}\right) _{P}. \label{CpCvDiffer}
3011: \end{equation}
3012: For solids at low temperatures,%
3013: \begin{equation}
3014: C_{P}\approx C_{V},
3015: \end{equation}
3016: because
3017: \begin{equation}
3018: \left( \frac{\partial V}{\partial T}\right) _{P}\approx0.
3019: \end{equation}
3020: That is why the Debye $T^{3}$ law for the phonons, the $T$ law for the
3021: electron gases in normal metals, and the exponential law for BCS
3022: superconductors are observed experimentally in $C_{P}$ albeit they are the
3023: laws of $C_{V}$ at low temperatures. In other words, the theoretical result
3024: of $C_{V}$ for a solid can be verified straightforwardly by the experiments.
3025: However, for a liquid, the situation is different, both the volume and
3026: pressure are much more sensitive to temperature. Therefore, the second term
3027: on the right-hand side of Eq. (\ref{CpCvDiffer}) can not be neglected any
3028: longer, the difference between $C_{P}$ and $C_{V}$ becomes important and
3029: should be taken into account. As a result, to compare the theory with the
3030: experiment for a liquid, the experimental data for $C_{V}$ or the
3031: theoretical result for $C_{P}$ are needed. Regrettably, $C_{V}$ is difficult
3032: to measure experimentally, and $C_{P}$ is difficult to calculate
3033: theoretically. It is thus hard to compare the theoretical result with the
3034: experimental data. Fortunately, $C_{V}$ tends to zero linearly as $%
3035: T\rightarrow0$ for a weakly interacting Bose gas, we can compare the
3036: theoretical result with the experimental data in the limit $T\rightarrow0$.
3037: That can be seen clearly from the following analyses.
3038:
3039: According to the third law of thermodynamics, we have%
3040: \begin{equation}
3041: S_{th}=S_{th}(T,V)=\int_{0}^{T}C_{V}(T^{\prime},V)\frac{\mathrm{d}T^{\prime}%
3042: }{T^{\prime}},
3043: \end{equation}
3044: where, as pointed in Sec. \ref{EET}, $S_{th}$ represents the thermodynamical
3045: entropy. Substituting Eq. (\ref{CvTV}) into this equation, we obtain%
3046: \begin{equation}
3047: S_{th}=a(V)T+\frac{1}{3}b(V)T^{3},
3048: \end{equation}
3049: at low temperatures. Making use of the Maxwell relations,
3050: \begin{gather}
3051: \left( \frac{\partial P}{\partial T}\right) _{V}=\left( \frac{\partial S_{th}%
3052: }{\partial V}\right) _{T}, \\
3053: \left( \frac{\partial V}{\partial T}\right) _{P}=-\left( \frac{\partial
3054: S_{th}}{\partial P}\right) _{T},
3055: \end{gather}
3056: one has%
3057: \begin{align}
3058: \hspace{-1cm}\left( \frac{\partial P}{\partial T}\right) _{V} & =a^{\prime
3059: }(V)T+\frac{1}{3}b^{\prime}(V)T^{3}, \\
3060: \hspace{-1cm}\left( \frac{\partial V}{\partial T}\right) _{P} & =-\left(
3061: \frac{\partial V}{\partial P}\right) _{T}\left[ a^{\prime}(V)T+\frac{1}{3}%
3062: b^{\prime}(V)T^{3}\right] .
3063: \end{align}
3064: Substituting them into Eq. (\ref{CpCvDiffer}) leads us to%
3065: \begin{equation}
3066: C_{P}=a(V)T+\widetilde{b}(V)T^{3}, \label{CpLLTc}
3067: \end{equation}
3068: where the terms with power higher than cube are omitted, and
3069: \begin{equation}
3070: \widetilde{b}(V)=b(V)-\left[ a^{\prime}(V)\right] ^{2}\left( \frac{\partial V%
3071: }{\partial P}\right) _{T}. \label{SoundContr}
3072: \end{equation}
3073: Eq. (\ref{CpLLTc}) presents the form of $C_{P}$ at low temperatures within
3074: the present theory. It contains a linear and a cubic term of $T$, the former
3075: includes only the contribution from $C_{V}$, and the latter includes both
3076: the contributions from $C_{V}$ and the second term on the right-hand side of
3077: Eq. (\ref{CpCvDiffer}). Heeding that $S_{th}=0$ at $T=0$, we have \cite%
3078: {Lifshitz}%
3079: \begin{equation}
3080: u^{2}=-\frac{V^{2}}{mN}\left( \frac{\partial P}{\partial V}\right) _{T=0},
3081: \end{equation}
3082: where $u$ denotes the velocity of sound. This implies that at low
3083: temperatures the velocity of sound is approximate to%
3084: \begin{equation}
3085: u^{2}\approx-\frac{V^{2}}{mN}\left( \frac{\partial P}{\partial V}\right)
3086: _{T}.
3087: \end{equation}
3088: That is to say, the factor $(\partial V/\partial P)_{T}$ on the right-hand
3089: side of Eq. (\ref{SoundContr}) represents physically the effect of sound.
3090: Therefore, the cubic term of $C_{P}$ includes the contribution from the
3091: sound, which is in accordance with the viewpoint of Landau on quantum Bose
3092: liquid $\mathrm{He}$ II \cite{Landau1,Landau2}. This $T^{3}$ law has already
3093: been observed experimentally in liquid $^{4}\mathrm{He}$ at low temperatures
3094: \cite{Keesom}, but it is not an intrinsic property of the BEC because sound
3095: exists in all fluids. The intrinsic property of the BEC is the linear term
3096: of $C_{P}$, it comes purely from $C_{V}$, and is characteristic of a weakly
3097: interacting Bose gas at $T\ll T_{c}$, as mentioned above. There is no such
3098: linear term in the Landau theory of quantum Bose liquid \cite%
3099: {Landau1,Landau2}. To our knowledge, this linear behavior of $C_{P}$ has not
3100: yet been reported experimentally for liquid $^{4}\mathrm{He}$, it is thus a
3101: prediction of the present theory. Eq. (\ref{CpLLTc}) indicates that the
3102: temperature $T$ has to go much lower for $C_{P}$ to show the linear behavior
3103: than to show the $T^{3}$ law. We believe that it would be observed if the
3104: measuring temperature is lowered enough, and if liquid $^{4}\mathrm{He}$
3105: could be regarded as a weakly interacting Bose gas and the $\lambda$%
3106: -transition were a Bose-Einstein condensation.
3107:
3108: \textbf{Remark}: In general, $\left\vert \xi \right\vert ^{2}$ can not be
3109: interpreted as the density of condensed particles. This can be seen from Eq.
3110: (\ref{MuLTc}), it shows that $y>1$ at $T=0\mathrm{K}$, that is,
3111: \begin{equation}
3112: \left\vert \xi \right\vert ^{2}>n
3113: \end{equation}%
3114: at absolute zero. Only in the case of the ideal Bose gas, $\left\vert \xi
3115: \right\vert ^{2}$ can not be greater than the density of particles, i.e.,
3116: \begin{equation}
3117: \left\vert \xi \right\vert ^{2}\leq n
3118: \end{equation}%
3119: at any temperature. Physically, $\xi $ is the order parameter and internal
3120: spontaneous field of the system, there is no \textit{a priori} reason why it
3121: must be interpreted using the the particle density. In fact, any statistical
3122: average of observable, including the particle density, will depend
3123: explicitly on $\xi $ when $T<T_{c}$.
3124:
3125: Now, let us discuss the correlation of the BEC to the superfluidity. As
3126: usual, we set%
3127: \begin{equation}
3128: \eta (\mathbf{r})=\sqrt{\rho (\mathbf{r})}e^{i\varphi (\mathbf{r})},
3129: \end{equation}%
3130: where $\rho (\mathbf{r})=\left\vert \eta (\mathbf{r})\right\vert ^{2}$ and $%
3131: \varphi (\mathbf{r})=\arg (\eta (\mathbf{r}))$. With this, we have%
3132: \begin{eqnarray}
3133: \mathbf{j}(\mathbf{r}) &=&-\frac{i\hbar }{2m}\langle (\nabla -\nabla
3134: ^{\prime })\psi ^{\dagger }(\mathbf{r})\psi (\mathbf{r}^{\prime })\rangle |_{%
3135: \mathbf{r}^{\prime }=\mathbf{r}} \notag \\
3136: &=&\frac{\hbar }{m}\rho (\mathbf{r})\nabla \varphi (\mathbf{r}),
3137: \end{eqnarray}%
3138: obviously, $\mathbf{j}(\mathbf{r})$ is the supercurrent density. If $\varphi
3139: (\mathbf{r})$, the phase of the BEC order parameter, is independent of $%
3140: \mathbf{r}$, there is no supercurrent, $\mathbf{j}(\mathbf{r})=0$;
3141: otherwise, $\mathbf{j}(\mathbf{r})\neq 0$. The supercurrent velocity $%
3142: \mathbf{v}$ is given by
3143: \begin{equation}
3144: \mathbf{v}(\mathbf{r})=\frac{\hbar }{m}\nabla \varphi (\mathbf{r}).
3145: \end{equation}%
3146: Since the order parameter $\eta (\mathbf{r})$ is a single-valued function of
3147: the position $\mathbf{r}$, we have%
3148: \begin{equation}
3149: \oint \mathbf{v}(\mathbf{r})\cdot \mathrm{d}\mathbf{r}=l\frac{h}{m},\text{ }%
3150: l\in
3151: %TCIMACRO{\U{2124} }%
3152: %BeginExpansion
3153: \mathbb{Z}
3154: %EndExpansion
3155: .
3156: \end{equation}%
3157: That is to say, there can exist quantized vortices in the condensed phase if
3158: $\varphi (\mathbf{r})$ depends on $\mathbf{r}$. We remark that there exists
3159: a critical velocity $v_{c}$, the velocity of the supercurrent can not be
3160: greater than it, i.e., $v\leq v_{c}$. That is because the increase of the
3161: supercurrent velocity is unfavorable to the decrease of the entropy of the
3162: system; the larger the $v$ is, the greater the entropy will be, which can be
3163: easily seen from the first integral over $\mathbf{r}$ on the right-hand side
3164: of Eq. (\ref{IBECS}) with, e.g., $\eta (\mathbf{r})=\sqrt{\rho }e^{i\mathbf{k%
3165: }\cdot \mathbf{r}}$ where $\rho =\mathrm{constant}$.
3166:
3167: Finally, we shall prove that there can not exist any supercurrent or
3168: quantized vortex in the ideal Bose gas.
3169:
3170: For the ideal Bose gas, $v(\mathbf{r}-\mathbf{r}^{\prime })=0$, Eq. (\ref%
3171: {IOP}) reduces to%
3172: \begin{equation}
3173: -\frac{\hbar ^{2}}{2m}\nabla ^{2}\eta (\mathbf{r})=\mu \eta (\mathbf{r}),
3174: \label{Keig}
3175: \end{equation}%
3176: obviously, it is the eigenvalue equation of the kinetic energy operator.
3177: This equation can be rewritten as%
3178: \begin{equation}
3179: \frac{1}{2m}\int \mathrm{d}\mathbf{r\,}[-i\hbar \nabla \eta (\mathbf{r}%
3180: )]^{\dagger }\cdot \lbrack -i\hbar \nabla \eta (\mathbf{r})]=\mu \int
3181: \mathrm{d}\mathbf{r\,}\eta ^{\dagger }(\mathbf{r})\eta (\mathbf{r}).
3182: \end{equation}%
3183: It indicates that, for any nontrivial eigenfunction, the chemical potential $%
3184: \mu $ can not be negative, i.e., $\mu \geq 0$. Therefore, there can only
3185: exist the trivial solution, $\eta (\mathbf{r})=0$, when $\mu <0$. The
3186: nontrivial solution, $\eta (\mathbf{r})\neq 0$, can exist only when $\mu
3187: \geq 0$ for the ideal Bose gas.
3188:
3189: To discuss the stabilities of those solutions of Eq. (\ref{Keig}), one needs
3190: to investigate whether $\Delta S\geq 0$. From Eq. (\ref{IBECS}), we find
3191: \begin{equation}
3192: \Delta S=\beta \int \mathrm{d}\mathbf{r}\left[ \frac{\hbar ^{2}}{2m}\nabla
3193: \delta \eta ^{\dagger }(\mathbf{r})\cdot \nabla \delta \eta (\mathbf{r})-\mu
3194: \delta \eta ^{\dagger }(\mathbf{r})\delta \eta (\mathbf{r})\right] ,
3195: \end{equation}%
3196: where $\delta \eta (\mathbf{r})$ represents the variation of the order
3197: parameter $\eta (\mathbf{r})$ from any one of the solutions of Eq. (\ref%
3198: {Keig}). Evidently, $\Delta S\geq 0$ when $\mu \leq 0$, this means that the
3199: system is stable when $\mu \leq 0$. Since $\eta (\mathbf{r})=0$ when $\mu <0$
3200: and $\eta (\mathbf{r})=\mathrm{constant}$ when $\mu =0$, the normal phase
3201: and the homogeneous condensed phase are both stable, which is in accordance
3202: with the results of Sec. \ref{TLC}. As to the case of $\mu >0$, the solution
3203: of Eq. (\ref{Keig}) is inhomogeneous, which is just the case we are now
3204: interested in because an inhomogeneous solution can produce supercurrent and
3205: quantized vortices. To determine whether there can exist any supercurrent
3206: and quantized vortex in the condensed phase, we first reformulate the above
3207: equation as follows,%
3208: \begin{equation}
3209: \Delta S=\beta \int \mathrm{d}\mathbf{r}\left[ \delta \eta ^{\dagger }(%
3210: \mathbf{r})\left( -\frac{\hbar ^{2}}{2m}\nabla ^{2}\right) \delta \eta (%
3211: \mathbf{r})-\mu \delta \eta ^{\dagger }(\mathbf{r})\delta \eta (\mathbf{r})%
3212: \right] ,
3213: \end{equation}%
3214: and then examine the eigenvalue problem,
3215: \begin{eqnarray}
3216: -\frac{\hbar ^{2}}{2m}\nabla ^{2}\delta \eta (\mathbf{r}) &=&E\delta \eta (%
3217: \mathbf{r}), \label{Bessel} \\
3218: \delta \eta (\mathbf{r})|_{r\rightarrow +\infty } &=&0. \label{boundary}
3219: \end{eqnarray}%
3220: Eq. (\ref{boundary}) is the boundary condition obeyed by the variation $%
3221: \delta \eta (\mathbf{r})$. Obviously, this eigenvalue problem always has
3222: solutions when $E>0$, e.g.,%
3223: \begin{equation}
3224: \delta \eta (\mathbf{r})=j_{l}(kr),\text{ }k=\sqrt{\frac{2m}{\hbar ^{2}}E},
3225: \end{equation}%
3226: where $j_{l}(z)$ is the spherical Bessel function. Accompanying those
3227: solutions, $\Delta S$ can be reexpressed as%
3228: \begin{equation}
3229: \Delta S=\beta \int \mathrm{d}\mathbf{r}(E-\mu )\delta \eta ^{\dagger }(%
3230: \mathbf{r})\delta \eta (\mathbf{r}),
3231: \end{equation}%
3232: clearly,%
3233: \begin{equation}
3234: \left\{
3235: \begin{array}{ll}
3236: \Delta S>0,\text{ } & E>\mu \\
3237: \Delta S=0,\text{ } & E=\mu \\
3238: \Delta S<0,\text{ } & E<\mu .%
3239: \end{array}%
3240: \right.
3241: \end{equation}%
3242: It shows that any solution of Eq. (\ref{Keig}) for $\mu >0$ belongs to the
3243: saddle points of the system entropy. In other words, any inhomogeneous
3244: condensed phase is physically instable, needless to say, the supercurrent
3245: and quantized vortex. This proves that there can not exist any supercurrent
3246: or quantized vortex in the ideal Bose gas.
3247:
3248: By the way, the above analyses demonstrate that the condensed phase of the
3249: ideal Bose gas must be homogeneous. There is no inhomogeneous condensation
3250: in the ideal Bose gas. This result is very natural, it confirms the
3251: Einstein's deep physical insight on BEC, again.
3252:
3253: \subsection{Double-well Potential and the BEC in Configuration Space}
3254:
3255: From Sec. \ref{IPGSB}, it is learned that an ideal phonon gas cannot produce
3256: BEC. However, the ideal phonon gas is simply a harmonic approximation to an
3257: actual solid. Generally, the interatomic interaction is anharmonic. It thus
3258: raises the question as to whether an anharmonic system can produce BEC. We
3259: intend to discuss the question in this subsection.
3260:
3261: To that end, it is helpful to make a representation transformation to the
3262: formulation used in Sec. \ref{IPGSB}. There, what we used is the Fock space
3263: where all the observables are expressed in terms of creation and
3264: annihilation operators. In the following, we prefer to use the phase space
3265: where all the observables are expressed in terms of generalized coordinates
3266: and momenta. As is well known, the two spaces are equivalent and can be
3267: transformed into each other according to such a rule \cite{Berezin},
3268: \begin{equation}
3269: \left\{
3270: \begin{array}{l}
3271: q_{i}=\sqrt{\frac{\hbar }{2m_{i}\omega _{i}}}(a_{i}+a_{i}^{\dagger }) \\
3272: p_{i}=-i\sqrt{\frac{m_{i}\hbar \omega _{i}}{2}}(a_{i}-a_{i}^{\dagger }),%
3273: \end{array}%
3274: \right. \label{Transform1}
3275: \end{equation}%
3276: and
3277: \begin{equation}
3278: \left\{
3279: \begin{array}{l}
3280: a_{i}=\sqrt{\frac{m_{i}\omega _{i}}{2\hbar }}q_{i}+i\frac{1}{\sqrt{%
3281: 2m_{i}\hbar \omega _{i}}}p_{i} \\
3282: a_{i}^{\dagger }=\sqrt{\frac{m_{i}\omega _{i}}{2\hbar }}q_{i}-i\frac{1}{%
3283: \sqrt{2m_{i}\hbar \omega _{i}}}p_{i},%
3284: \end{array}%
3285: \right. \label{Transform2}
3286: \end{equation}%
3287: where $q_{i}$ and $p_{i}$ are the $i$th pair of coordinate and momentum; $%
3288: a_{i}$ and $a_{i}^{\dagger }$ the corresponding annihilation and creation
3289: operators; $m_{i}$ the effective mass; and $\omega _{i}>0$ the $i$th
3290: parameter, which can generally take any positive value, and particularly the
3291: natural frequency if $q_{i}$ and $p_{i}$ constitute a harmonic oscillator.
3292: Along with Eqs. (\ref{Transform1}) and (\ref{Transform2}), the main
3293: quantities concerned are transformed as follows,%
3294: \begin{align}
3295: F(a)& \Longleftrightarrow F(q,p), \\
3296: H(a)& \Longleftrightarrow H(q,p), \\
3297: D(\xi ,a)& \Longleftrightarrow D(\eta ,\zeta ,q,p), \label{PTD} \\
3298: S(\xi ,\beta )& \Longleftrightarrow S(\eta ,\zeta ,\beta ),
3299: \label{PSEntropy}
3300: \end{align}%
3301: where%
3302: \begin{gather}
3303: D(\xi ,a)=\sum\limits_{i}(\xi _{i}^{\dag }a_{i}+\xi _{i}a_{i}^{\dag }), \\
3304: D(\eta ,\zeta ,q,p)=\frac{1}{\hbar }\sum\limits_{i}(\eta _{i}q_{i}+\zeta
3305: _{i}p_{i}), \\
3306: \left\{
3307: \begin{array}{l}
3308: \eta _{i}=\sqrt{\frac{m_{i}\hbar \omega _{i}}{2}}(\xi _{i}^{\dag }+\xi _{i})
3309: \\
3310: \zeta _{i}=i\sqrt{\frac{\hbar }{2m_{i}\omega _{i}}}(\xi _{i}^{\dag }-\xi
3311: _{i}).%
3312: \end{array}%
3313: \right.
3314: \end{gather}%
3315: Here, $\xi $ represents the order parameter for BEC in Fock space, $\eta
3316: _{i} $ and $\zeta _{i}$ the corresponding order parameters in phase space.
3317:
3318: Under the action of $D(\eta,\zeta,q,p)$, the statistical averages of $q_{i} $
3319: and $p_{i}$ will change with temperature as
3320: \begin{equation}
3321: \left\{
3322: \begin{array}{l}
3323: \left\langle q_{i}\right\rangle =-\zeta_{i}+\mathrm{Tr}\big(q_{i}\rho\left(
3324: H(q,p)\right) \big) \\
3325: \left\langle p_{i}\right\rangle =\eta_{i}+\mathrm{Tr}\big(p_{i}\rho\left(
3326: H(q,p)\right) \big).%
3327: \end{array}
3328: \right. \label{Distortions}
3329: \end{equation}
3330: If the system produces a BEC, i.e., $\eta_{i}$ and/or $\zeta_{i}$ change
3331: from zero to nonzero with the decreasing of temperature, the average
3332: positions of the particles will redistribute in phase space. In this sense,
3333: the BEC is said to be the phase-space BEC, it changes the distribution of
3334: the system in phase space. If, however, the BEC only causes $\eta_{i}$ (or $%
3335: \zeta_{i}$) to change from zero to nonzero, the particles will redistribute
3336: merely in momentum space (or configuration space), the subspace of phase
3337: space. In such a case, the BEC is said to be the momentum-space BEC (or the
3338: configuration-space BEC), it changes only the distribution of the system in
3339: momentum space (or configuration space). Obviously, once a BEC changes the
3340: distribution of a system in configuration space ($\zeta_{i}\neq0$, $%
3341: \eta_{i}=0$ or not), it induces a structural phase transition (SPT) \cite%
3342: {Blinc}. In this connection, it can be said that a SPT is just an instance
3343: of BEC.
3344:
3345: Commonly, a system Hamiltonian has the form,
3346: \begin{equation}
3347: H(q,p)=\sum_{i}\frac{p_{i}^{2}}{2m_{i}}+\sum_{i}u\left( q_{i}\right) +\frac{1%
3348: }{2}\sum_{i,j}v(q_{i}-q_{j}), \label{SPHamiltonian}
3349: \end{equation}%
3350: where $u\left( q_{i}\right) $ and $v(q_{i}-q_{j})$ stand for the single- and
3351: two-particle potentials, respectively. One can easily verify, using Eqs. (%
3352: \ref{Variation}), (\ref{Increment}), (\ref{PTD}) and (\ref{PSEntropy}), that
3353: $\eta _{i}$ must be zero at any temperature, i.e., there cannot arise the
3354: momentum-space BEC. In consequence, the system with such a form of
3355: Hamiltonian can, at most, produce a configuration-space BEC. Therefore, a
3356: SPT is commonly a configuration-space BEC. In the following, we shall
3357: confine our attention within this kind of SPT. The phase-transition operator
3358: $D(\eta ,\zeta ,q,p)$ can now be simplified as
3359: \begin{equation}
3360: D(\zeta ,p)=\frac{1}{\hbar }\sum\limits_{i}\zeta _{i}p_{i}, \label{PSD}
3361: \end{equation}%
3362: for $\eta _{i}$ vanishes forever.
3363:
3364: In the field of SPT, the Hamiltonian of Eq. (\ref{SPHamiltonian}) is made
3365: concrete as \cite{Blinc},
3366: \begin{align}
3367: H(q,p)& =\sum_{i}\frac{p_{i}^{2}}{2m}+\sum_{i}\left( \frac{1}{2}\mu
3368: ^{2}q_{i}^{2}+\frac{1}{4}gq_{i}^{4}\right) \notag \\
3369: & +\sum_{\langle ij\rangle }\frac{1}{2}\lambda (q_{i}-q_{j})^{2},
3370: \label{PSHamiltonian}
3371: \end{align}%
3372: where $\mu ^{2}$, $g$, and $\lambda $ are constants, and $\langle ij\rangle $
3373: denotes the nearest neighbor sites. The system gets anharmonic if $\mu
3374: ^{2}<0 $ or $g\neq 0$. The terms included by the second sum represent the
3375: single-particle potential that arises from an underlying sublattice of
3376: atoms, which do not participate actively in the phase transition, and the
3377: one included by the third sum represents the elastic coupling between
3378: nearest neighbors, which is described by a harmonic potential with the
3379: coupling constant $\lambda \geq 0$. Obviously, $H(q,p)$ has the discrete
3380: symmetry,
3381: \begin{equation}
3382: q_{i}\Longrightarrow q_{i}^{\prime }=e^{i\frac{\pi }{\hbar }%
3383: \sum\limits_{i}\left( \frac{1}{2}p_{i}^{2}+\frac{1}{2}q_{i}^{2}\right)
3384: }q_{i}e^{-i\frac{\pi }{\hbar }\sum\limits_{i}\left( \frac{1}{2}p_{i}^{2}+%
3385: \frac{1}{2}q_{i}^{2}\right) }=-q_{i},
3386: \end{equation}%
3387: it leads us to
3388: \begin{equation}
3389: \overline{q_{i}}=0, \label{QBar}
3390: \end{equation}%
3391: where
3392: \begin{equation}
3393: \overline{F(q,p)}=\mathrm{Tr}\big(F(q,p)\rho \left( H(q,p)\right) \big).
3394: \end{equation}%
3395: This discrete symmetry will break down if a SPT takes place.
3396:
3397: Let us begin with the simple case of $\lambda=0$. Making use of Eq. (\ref%
3398: {PSD}), we have
3399: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}}}%
3400: %BeginExpansion
3401: \begin{widetext}%
3402: %EndExpansion
3403: \begin{equation}
3404: S(\zeta,\beta)=-\overline{\ln\left( \rho\left( H(q,p)\right) \right) }%
3405: +\beta\sum_{i}\left\{ \frac{1}{2}\left[ \mu^{2}+3g\left( \overline {q_{i}^{2}%
3406: }-\overline{q_{i}}^{2}\right) \right] \zeta_{i}^{2}+\frac{1}{4}%
3407: g\zeta_{i}^{4}\right\} . \label{PSS}
3408: \end{equation}%
3409: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
3410: %BeginExpansion
3411: \end{widetext}
3412: %EndExpansion
3413: It shows that the entropy $S(\zeta,\beta)$ has the typical form of Landau
3414: free energy \cite{Landau}. From Eqs. (\ref{PSS}), (\ref{Variation}), and (%
3415: \ref{Increment}), one can draw three conclusions.
3416:
3417: \begin{enumerate}
3418: \item The case of $g<0$. The system is instable and will collapse because
3419: the entropy is unbounded from below with respect to the variation of $%
3420: \zeta_{i}$.
3421:
3422: \item The case of $g=0$. (a) If $\mu ^{2}>0$, the trivial solution $\zeta
3423: _{i}=0$ is the only minimizer of $S(\zeta ,\beta )$. That is to say, a
3424: system consisting of harmonic oscillators, or rather an ideal phonon gas,
3425: can not produce SPT, it will stay in the normal phase forever, which is in
3426: accordance with the result of Sec. \ref{IPGSB}. (b) If $\mu ^{2}=0$, $\zeta
3427: _{i}$ can take a temperature-independent value, it represents, as indicated
3428: by Eqs. (\ref{Distortions}) and (\ref{QBar}), the equilibrium position of $%
3429: q_{i}$, i.e., $\left\langle q_{i}\right\rangle =-\zeta _{i}$. As the initial
3430: equilibrium position of $q_{i}$ at $\beta =0$, $-\zeta _{i}$ can be chosen
3431: as zero. Therefore, a system composed of free distinguishable particles can
3432: not produce SPT, as is just expected. (c) If $\mu ^{2}<0$, $\zeta _{i}=0$ is
3433: the only solution, but it is a maximizer of $S(\zeta ,\beta )$, the system
3434: is instable and will collapse. This result is quite natural since all
3435: particles go in a reversed harmonic potential.
3436:
3437: \item The case of $g>0$. (a) If $\mu^{2}\geq0$, $\zeta_{i}=0$ is the only
3438: solution, and it is a minimizer of $S(\zeta,\beta)$. There is no phase
3439: transition in the system, the potential is a single well. (b) If $\mu^{2}<0$%
3440: , the potential gets double welled, and it must be determined by the
3441: position fluctuation $\overline{q_{i}^{2}}-\overline{q_{i}}^{2}$ whether the
3442: system can produce a phase transition. Evidently, a transition can occur if
3443: the fluctuation is such a decreasing function of temperature that $\mu
3444: ^{2}+3g(\overline{q_{i}^{2}}-\overline{q_{i}}^{2})$ can change in sign, from
3445: positive to negative. If the fluctuation is so strong that $\mu^{2}+3g[%
3446: \overline{(q_{i}^{2})}-\left( \overline{q_{i}}\right) ^{2}]>0$ up to zero
3447: temperature, the system has to stay in the normal phase forever.
3448: \end{enumerate}
3449:
3450: From the above discussions, it follows that a double-well potential where $%
3451: \mu^{2}<0$ and $g>0$ is the only possible case for the system to produce a
3452: structural phase transition under the condition of $\lambda=0$. One can
3453: easily verify that this conclusion is still valid for $\lambda>0$.
3454:
3455: We are now confronted with performing the evaluation of the position
3456: fluctuation $\overline{q_{i}^{2}}-\overline{q_{i}}^{2}$ so as to ascertain
3457: whether a phase transition can occur or not if $\mu ^{2}<0$ and $g>0$. To
3458: this end, we need to perform a representation transformation to the
3459: Hamiltonian of Eq. (\ref{PSHamiltonian}) because, as well-known, this form
3460: of Hamiltonian has no good basic part and is thus unfavorable for doing
3461: approximations. Complying with quantum field theory, the transformation can
3462: be achieved as follows,
3463: \begin{equation}
3464: \left\{
3465: \begin{array}{l}
3466: q_{i}\implies \phi _{i}=e^{\frac{i}{\hbar }\sum\limits_{j}\nu p_{j}}q_{i}e^{-%
3467: \frac{i}{\hbar }\sum\limits_{j}\nu p_{j}}=q_{i}+\nu \\
3468: p_{i}\implies \pi _{i}=e^{\frac{i}{\hbar }\sum\limits_{j}\nu p_{j}}p_{i}e^{-%
3469: \frac{i}{\hbar }\sum\limits_{j}\nu p_{j}}=p_{i},%
3470: \end{array}%
3471: \right. \label{qpt}
3472: \end{equation}%
3473: where
3474: \begin{equation}
3475: \nu =\sqrt{-\mu ^{2}/g}.
3476: \end{equation}%
3477: In the sense of classical mechanics, the above transformation amounts to
3478: performing a translation so that the new coordinate $\phi _{i}$ is
3479: referenced from $q_{i}=-\nu $, one of the two minimizers of the double-well
3480: potential. Along with Eq. (\ref{qpt}), the new representation $H(\phi ,\pi )$
3481: of the Hamiltonian becomes%
3482: \begin{align}
3483: H(\phi ,\pi )& =e^{\frac{i}{\hbar }\sum\limits_{j}\nu p_{j}}H(q,p)e^{-\frac{i%
3484: }{\hbar }\sum\limits_{j}\nu p_{j}} \notag \\
3485: & =\sum_{i}\left( \frac{p_{i}^{2}}{2m}-\mu ^{2}q_{i}^{2}\right)
3486: +\sum_{i}(g\nu q_{i}^{3}+\frac{1}{4}gq_{i}^{4}) \notag \\
3487: & -\sum_{i}\frac{1}{4}g\nu ^{4}.
3488: \end{align}%
3489: The first sum on the right-hand side represents harmonic oscillators with
3490: natural frequencies $\omega _{i}=\sqrt{-2\mu ^{2}/m}$, it constitutes a
3491: basic Hamiltonian which is favorable for approximation handling, the second
3492: the interaction, and the third an unimportant constant.
3493:
3494: With $H(\phi,\pi)$, one finds
3495: \begin{equation}
3496: \overline{q_{i}^{2}}-\overline{q_{i}}^{2}=\widetilde{\phi_{i}^{2}}-%
3497: \widetilde{\phi_{i}}^{2}=\widetilde{q_{i}^{2}}-\widetilde{q_{i}}^{2},
3498: \label{qfluctuation}
3499: \end{equation}
3500: where $\widetilde{F}$ stands for the statistical average with respect to $%
3501: H(\phi,\pi)$,%
3502: \begin{equation}
3503: \widetilde{F}=\mathrm{Tr}\big(F(q,p)\rho(\widetilde{H}(q,p))\big),
3504: \end{equation}
3505: where $\widetilde{H}(q,p)\equiv H(\phi,\pi)$. According to Eq. (\ref%
3506: {qfluctuation}), we now just need to evaluate the fluctuation $\widetilde{%
3507: q_{i}^{2}}-\widetilde{q_{i}}^{2}$, it can be worked out by the retarded
3508: Green's function $\langle\langle q_{i}|q_{i}\rangle\rangle_{\omega}$ which
3509: is defined with respect to $\widetilde{H}(q,p)$ \cite{Zubarev}. The Green's
3510: function satisfies the following equation of motion,
3511: \begin{align}
3512: \hbar\omega\langle\langle q_{i}|q_{i}\rangle\rangle_{\omega} & =\langle
3513: \lbrack q_{i},q_{i}]\rangle+\langle\langle\lbrack q_{i},\widetilde {H}%
3514: (q,p)]|q_{i}\rangle\rangle_{\omega} \notag \\
3515: & =\frac{i\hbar}{m}\langle\langle p_{i}|q_{i}\rangle\rangle_{\omega}.
3516: \end{align}
3517: Also,%
3518: \begin{align}
3519: \hbar\omega\langle\langle p_{i}|q_{i}\rangle\rangle_{\omega} & =\langle
3520: \lbrack p_{i},q_{i}]\rangle+\langle\langle\lbrack p_{i},\widetilde {H}%
3521: (q,p)]|q_{i}\rangle\rangle_{\omega} \notag \\
3522: & =-i\hbar\big[[1-2\mu^{2}\langle\langle q_{i}|q_{i}\rangle\rangle_{\omega
3523: }+3g\nu\langle\langle q_{i}^{2}|q_{i}\rangle\rangle_{\omega} \notag \\
3524: & +g\langle\langle q_{i}^{3}|q_{i}\rangle\rangle_{\omega}\big].
3525: \end{align}
3526: We would like to truncate the GF chain with the following decoupling
3527: approximation,%
3528: \begin{gather}
3529: \langle\langle q_{i}^{2}|q_{i}\rangle\rangle_{\omega}=\widetilde{q_{i}}%
3530: \langle\langle q_{i}|q_{i}\rangle\rangle_{\omega}=-\nu\langle\langle
3531: q_{i}|q_{i}\rangle\rangle_{\omega}, \\
3532: \langle\langle q_{i}^{3}|q_{i}\rangle\rangle_{\omega}=\widetilde{q_{i}^{2}}%
3533: \langle\langle q_{i}|q_{i}\rangle\rangle_{\omega}=[(\overline{q_{i}^{2}}-%
3534: \overline{q_{i}}^{2})+\nu^{2}]\langle\langle q_{i}|q_{i}\rangle
3535: \rangle_{\omega},
3536: \end{gather}
3537: where we have used the result $\widetilde{\phi_{i}}=\widetilde{q_{i}}+\nu=0 $%
3538: , which follows from Eq. (\ref{QBar}).\ This approximation leads us to%
3539: \begin{equation}
3540: \langle\langle q_{i}|q_{i}\rangle\rangle_{\omega}=\frac{1}{m}\frac{1}{%
3541: \omega^{2}-\omega_{s}^{2}},
3542: \end{equation}
3543: where
3544: \begin{equation}
3545: \omega_{s}=\left[ \frac{g}{m}(\overline{q_{i}^{2}}-\overline{q_{i}}^{2})%
3546: \right] ^{1/2} \label{Omegas}
3547: \end{equation}
3548: represents the renormalized frequency. According to the
3549: fluctuation-dissipation theorem \cite{Zubarev}, the fluctuation $\overline {%
3550: q_{i}^{2}}-\overline{q_{i}}^{2}$ can be obtained by%
3551: \begin{equation}
3552: \overline{q_{i}^{2}}-\overline{q_{i}}^{2}=\mathcal{P}\int_{-\infty}^{+\infty
3553: }d\omega\frac{1}{e^{\beta\hbar\omega}-1}\left[ -\frac{\hbar}{\pi}\mathrm{Im}%
3554: \langle\langle q_{i}|q_{i}\rangle\rangle_{\omega+i0}\right] ,
3555: \end{equation}
3556: it gives rise to%
3557: \begin{equation}
3558: \overline{q_{i}^{2}}-\overline{q_{i}}^{2}=\frac{\hbar}{m\omega_{s}}\left(
3559: \frac{1}{2}+\frac{1}{e^{\beta\hbar\omega_{s}}-1}\right) .
3560: \label{Fluctuation}
3561: \end{equation}
3562: Eq. (\ref{Fluctuation}) and Eq. (\ref{Omegas}) constitute a self-consistent
3563: approximation to the fluctuation $\overline{q_{i}^{2}}-\overline{q_{i}}^{2}$%
3564: . Evidently, the first term on the right-hand side of Eq. (\ref{Fluctuation}%
3565: ) represents the zero-point fluctuation. Eq. (\ref{Fluctuation}) shows that
3566: the fluctuation gets weaker and weaker when temperature goes lower and
3567: lower. At zero temperature, one has%
3568: \begin{equation}
3569: \overline{q_{i}^{2}}-\overline{q_{i}}^{2}=\frac{1}{\sqrt[3]{4mg}}\hbar
3570: ^{2/3}\sim O(\hbar^{2/3}),
3571: \end{equation}
3572: and at high temperatures,
3573: \begin{equation}
3574: \overline{q_{i}^{2}}-\overline{q_{i}}^{2}=\frac{1}{\sqrt{g\beta}}\sim
3575: O(T^{1/2}).
3576: \end{equation}
3577: They imply that there exists a finite temperature $T_{c}$ such that%
3578: \begin{equation}
3579: \left\{
3580: \begin{array}{ll}
3581: \mu^{2}+3g(\overline{q_{i}^{2}}-\overline{q_{i}}^{2})>0,\text{ } & T>T_{c}
3582: \\
3583: \mu^{2}+3g(\overline{q_{i}^{2}}-\overline{q_{i}}^{2})=0,\text{ } & T=T_{c}
3584: \\
3585: \mu^{2}+3g(\overline{q_{i}^{2}}-\overline{q_{i}}^{2})<0,\text{ } & T<T_{c}.%
3586: \end{array}
3587: \right. \label{SignCh}
3588: \end{equation}
3589: From Eq. (\ref{SignCh}) and Eq. (\ref{PSS}), it follows that a SPT will
3590: occur at $T_{c}$,
3591: \begin{equation}
3592: \zeta_{i}=\left\{
3593: \begin{array}{ll}
3594: 0,\text{ } & T\geq T_{c} \\
3595: \pm\left[ -\frac{\mu^{2}}{g}-\frac{3\hbar}{m\omega_{s}}\left( \frac{1}{2}+%
3596: \frac{1}{e^{\beta\hbar\omega_{s}}-1}\right) \right] ^{1/2},\text{ } &
3597: T<T_{c}.%
3598: \end{array}
3599: \right. \label{Zetai}
3600: \end{equation}
3601: Specifically, $\zeta_{i}=\pm\sqrt{-\mu^{2}/g}$ at zero temperature if the
3602: zero-point fluctuation is neglected, i.e., the order parameter will equal
3603: one of the two minimizers of the double-well potential at zero temperature
3604: when the zero-point fluctuation is left out of consideration.
3605:
3606: Combining Eqs. (\ref{Zetai}), (\ref{Distortions}), and (\ref{QBar}), we have
3607: \begin{equation}
3608: \left\langle q_{i}\right\rangle =\left\{
3609: \begin{array}{ll}
3610: 0,\text{ } & T\geq T_{c} \\
3611: \mp \left[ -\frac{\mu ^{2}}{g}-\frac{3\hbar }{m\omega _{s}}\left( \frac{1}{2}%
3612: +\frac{1}{e^{\beta \hbar \omega _{s}}-1}\right) \right] ^{1/2},\text{ } &
3613: T<T_{c}.%
3614: \end{array}%
3615: \right.
3616: \end{equation}%
3617: It shows that for each site the distortion $\left\langle q_{i}\right\rangle $
3618: can take independently either of the two values at $T<T_{c}$, there is no
3619: correlation between the distortions of different sites, therefore, the new
3620: phase below $T_{c}$ is a structural glass. Physically, that is because we
3621: have discarded the coupling between nearest neighbors ($\lambda =0$). If it
3622: is taken into account ($\lambda >0$), one can easily verify that the new
3623: phase below the transition temperature will be ferrodistortive:
3624: \begin{equation}
3625: \left\langle q_{i}\right\rangle =\left\langle q_{j}\right\rangle ,
3626: \end{equation}%
3627: where $i$ and $j$ denote any two different sites. This result is in
3628: agreement with the classical mean-field theory \cite{Blinc}.
3629:
3630: In conclusion, we find that the double-well potential system described by
3631: the Hamiltonian of Eq. (\ref{PSHamiltonian}) can produce a structural phase
3632: transition. Mechanically, that is because the position fluctuations of
3633: particles will decrease with temperature. In the high-temperature regime,
3634: the fluctuations are relatively strong, the system can only stay in its
3635: normal phase. The fluctuations will get weaker and weaker when temperature
3636: goes lower and lower, and finally the system undergoes a transition at a
3637: finite temperature and then turns into a new ordered phase in the
3638: low-temperature regime. We learn from here that the position fluctuations of
3639: particles play an important role in structural phase transitions. Also, we
3640: see that configuration space is very convenient for describing SPT, in
3641: contrast to the Fock space.
3642:
3643: \subsection{Goldstone Bosons, Ginzburg-Landau Equations, and Higgs Mechanism}
3644:
3645: Obviously, the preceding theory for the discrete case can be extended
3646: straightforwardly to the continuum case. Let us consider the so-called $O(N)$%
3647: -symmetric vector model \cite{Justin,Tsvelik},
3648: \begin{align}
3649: H\left[ \psi ,\pi \right] & =\int \mathrm{d}\mathbf{x}\text{ }\Big[\frac{1}{2%
3650: }\pi ^{\dagger }\pi +\frac{1}{2}\nabla \psi ^{\dagger }\cdot \nabla \psi
3651: \notag \\
3652: & +\frac{m^{2}}{2}\psi ^{\dagger }\psi +\frac{g}{4}(\psi ^{\dagger }\psi
3653: )^{2}\Big],
3654: \end{align}%
3655: where $m^{2}<0$ and $g>0$ are the two constant parameters for a double-well
3656: potential, $\psi (\mathbf{x})$ is usually called Higgs field, it represents
3657: a real-valued vector field with $N$ components, and $\pi (\mathbf{x})$ the
3658: corresponding momentum,
3659: \begin{equation}
3660: \lbrack \psi _{i}(\mathbf{x}),\pi _{j}(\mathbf{x}^{\prime })]=i\delta
3661: _{ij}\delta (\mathbf{x}-\mathbf{x}^{\prime }).
3662: \end{equation}%
3663: As is well known, this double-well Hamiltonian is the typical model for SSB
3664: in quantum field theory \cite{Justin,Tsvelik,Muller}. The arguments given in
3665: the preceding subsection are still valid. They reveal that the physical
3666: reason why the $O(N)$ symmetry will break down is that the fluctuation of
3667: the field $\psi (\mathbf{x})$ decreases monotonically with temperature, and
3668: that this SSB belongs essentially to the BEC happening in configuration
3669: space.
3670:
3671: In the condensation phase, the system Hamiltonian takes the form,
3672: \begin{equation}
3673: H^{\prime }\left[ \varphi ,\psi ,\pi \right] =e^{iD\left[ \varphi ,\pi %
3674: \right] }H\left[ \psi ,\pi \right] e^{-iD\left[ \varphi ,\pi \right] },
3675: \label{HPhiPsiPI}
3676: \end{equation}%
3677: where $\varphi (\mathbf{x})$ is the order parameter for Higgs field, and $D%
3678: \left[ \varphi ,\pi \right] $ the phase-transition operator,
3679: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}} }%
3680: %BeginExpansion
3681: \begin{widetext}
3682: %EndExpansion
3683: \begin{equation}
3684: D\left[ \varphi ,\pi \right] =\int \mathrm{d}\mathbf{x}\text{ }\varphi
3685: ^{\dagger }(\mathbf{x})\pi (\mathbf{x}).
3686: \end{equation}%
3687: In more detail, Eq. (\ref{HPhiPsiPI}) can be written as%
3688: \begin{equation}
3689: H^{\prime }\left[ \varphi ,\psi ,\pi \right] =\int \mathrm{d}\mathbf{x}\text{
3690: }\left\{ \frac{1}{2}\pi ^{\dagger }\pi +\frac{1}{2}\nabla (\psi ^{\dagger
3691: }+\varphi ^{\dagger })\cdot \nabla \left( \psi +\varphi \right) +\frac{m^{2}%
3692: }{2}(\psi ^{\dagger }+\varphi ^{\dagger })\left( \psi +\varphi \right) +%
3693: \frac{g}{4}[(\psi ^{\dagger }+\varphi ^{\dagger })\left( \psi +\varphi
3694: \right) ]^{2}\right\} . \label{HPhiPsiPi1}
3695: \end{equation}%
3696: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}}}%
3697: %BeginExpansion
3698: \end{widetext}%
3699: %EndExpansion
3700:
3701: At absolute zero, if one neglects the zero-point fluctuation of the field $%
3702: \psi $, i.e., $\overline{\psi ^{\dagger }\psi }\approx 0$, then he has the
3703: stable solution,
3704: \begin{equation}
3705: \varphi ^{\dagger }\varphi =\nu ^{2},
3706: \end{equation}%
3707: where%
3708: \begin{equation}
3709: \nu =\sqrt{-m^{2}/g}
3710: \end{equation}%
3711: is the minimizer of the double-well potential. Without loss of generality,
3712: we shall take
3713: \begin{equation}
3714: \varphi ^{\dagger }=\left[ \nu ,0,...,0\right] .
3715: \end{equation}%
3716: Substituting it into Eq. (\ref{HPhiPsiPi1}) leads us to
3717: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}}}%
3718: %BeginExpansion
3719: \begin{widetext}%
3720: %EndExpansion
3721: \begin{equation}
3722: H^{\prime }\left[ \varphi ,\psi ,\pi \right] =\int \mathrm{d}\mathbf{x}\text{
3723: }\left[ \frac{1}{2}\pi ^{\dagger }\pi +\frac{1}{2}\nabla \psi ^{\dagger
3724: }\cdot \nabla \psi -m^{2}\psi _{1}^{2}+g\nu \psi _{1}\psi ^{\dagger }\psi +%
3725: \frac{g}{4}(\psi ^{\dagger }\psi )^{2}-\frac{g\nu ^{4}}{4}\right] .
3726: \end{equation}%
3727: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
3728: %BeginExpansion
3729: \end{widetext}
3730: %EndExpansion
3731: That is the zero-temperature representation of the system Hamiltonian, it
3732: shows that the first component field $\psi _{1}(\mathbf{x})$ gets a mass of $%
3733: -2m^{2}$, the others are massless. As well-known, those massless particles
3734: are the so-called Goldstone bosons \cite{Goldstone}. From here, one can see
3735: how Goldstone bosons are produced in a natural way by spontaneous symmetry
3736: breaking within the framework of the extended ensemble theory. Of course, he
3737: should also recognize that the zero-point fluctuation can impose a little
3738: influence on the Goldstone bosons.
3739:
3740: At last, it should be emphasized that, in the $O(N)$-symmetric vector model,
3741: Goldstone bosons originate physically from the Bose-Einstein condensation of
3742: Higgs field.
3743:
3744: If a gauge field is coupled with the Higgs field $\psi(\mathbf{x})$, what
3745: will happen to the gauge field? can it produce BEC as the Higgs field $\psi(%
3746: \mathbf{x})$? To clarify this problem, let us consider, for simplicity, the
3747: coupling of a complex-valued (or two-component) Higgs field $\psi (\mathbf{x}%
3748: )$ and an electromagnetic field $\mathbf{A}(\mathbf{x})$,
3749: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}}}%
3750: %BeginExpansion
3751: \begin{widetext}%
3752: %EndExpansion
3753: \begin{equation}
3754: H\left[ \psi,\pi;\mathbf{A,E}\right] =\int\mathrm{d}\mathbf{x}\text{ }\left[
3755: \pi^{\dagger}\pi+\left( \nabla+ie\mathbf{A}\right) \psi^{\dagger
3756: }\cdot\left( \nabla-ie\mathbf{A}\right) \psi+m^{2}\psi^{\dagger}\psi
3757: +g(\psi^{\dagger}\psi)^{2}+\frac{1}{2}\mathbf{E}^{2}+\frac{1}{2}\left(
3758: \nabla\times\mathbf{A}\right) ^{2}\right] ,
3759: \end{equation}%
3760: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
3761: %BeginExpansion
3762: \end{widetext}
3763: %EndExpansion
3764: where $e$ is the charge of the Higgs field $\psi(\mathbf{x})$, and $\mathbf{E%
3765: }(\mathbf{x})$ the electric field strength which plays the role of the
3766: canonical momentum corresponding to the canonical coordinate $\mathbf{A}(%
3767: \mathbf{x})$,%
3768: \begin{equation}
3769: \lbrack\mathrm{A}_{i}(\mathbf{x}),-\mathrm{E}_{j}(\mathbf{x}^{\prime
3770: })]=i\delta_{ij}\delta(\mathbf{x}-\mathbf{x}^{\prime}).
3771: \end{equation}
3772: Here, the temporal gauge has been used to perform a canonical quantization
3773: to the electromagnetic field, an Abelian gauge field. The reason for
3774: choosing this gauge is that it can be easily applied to quantize non-Abelian
3775: gauge fields, and transformed into other gauges through the path-integral
3776: formalism introduced by Faddeev and Popov \cite{Faddeev}.
3777:
3778: For this coupled system, BEC can be explored through the phase-transition
3779: operator,
3780: \begin{equation}
3781: D\left[ \varphi,\pi\mathbf{;\Lambda},\mathbf{E}\right] =\int\mathrm{d}%
3782: \mathbf{x}\text{ }\left( \varphi\pi+\varphi^{\dagger}\pi^{\dagger }-\mathbf{%
3783: \Lambda}\cdot\mathbf{E}\right) , \label{DPiLambda}
3784: \end{equation}
3785: where $\varphi(\mathbf{x})$ and $\mathbf{\Lambda}(\mathbf{x})$ are the order
3786: parameters for the Higgs field $\psi(\mathbf{x})$ and gauge field $\mathbf{A}%
3787: (\mathbf{x})$, respectively. With $D\left[ \varphi,\pi \mathbf{;\Lambda},%
3788: \mathbf{E}\right] $, one can obtain the entropy of the system,
3789: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}}}%
3790: %BeginExpansion
3791: \begin{widetext}%
3792: %EndExpansion
3793: \begin{align}
3794: S\left[ \varphi,\mathbf{\Lambda,}\beta\right] & =-\mathrm{Tr}\big(\ln (\rho(H%
3795: \left[ \psi,\pi;\mathbf{A,E}\right] ))\rho(H\left[ \psi ,\pi;\mathbf{A,E}%
3796: \right] )\big)+\int\mathrm{d}\mathbf{x}\Big[\left( \nabla+ie\mathbf{\Lambda}%
3797: \right) \varphi^{\dagger}\cdot\left( \nabla -ie\mathbf{\Lambda}\right)
3798: \varphi \notag \\
3799: & +(m^{2}+e^{2}\overline{\mathbf{A}\cdot\mathbf{A}}+4g\overline{%
3800: \psi^{\dagger }\psi})\varphi^{\dagger}\varphi+g\left(
3801: \varphi^{\dagger}\varphi\right) ^{2}+e^{2}\overline{\psi^{\dagger}\psi}%
3802: \mathbf{\Lambda}^{2}+\frac{1}{2}\left( \nabla\times\mathbf{\Lambda}\right)
3803: ^{2}\Big], \label{CoupledPhiLambda}
3804: \end{align}
3805: which shows that the two order parameters $\varphi$ and $\mathbf{\Lambda}$
3806: are coupled together. Eq. (\ref{CoupledPhiLambda}) reminds us of the Landau
3807: free energy for superconductivity \cite{GL,Lifshitz}, indeed, they are both
3808: similar in form.
3809:
3810: Variation of $S$ with respect to $\varphi$ and $\mathbf{\Lambda}$ yields%
3811: \begin{gather}
3812: -\left( \nabla-ie\mathbf{\Lambda}\right) ^{2}\varphi+(m^{2}+e^{2}\overline{%
3813: \mathbf{A}\cdot\mathbf{A}}+4g\overline{\psi^{\dagger}\psi}%
3814: )\varphi+2g\varphi^{\dagger}\varphi\varphi=0, \label{GL1} \\
3815: \nabla\times\nabla\times\mathbf{\Lambda}=-ie\left(
3816: \varphi^{\dagger}\nabla\varphi-\varphi\nabla\varphi^{\dagger}\right)
3817: -2e^{2}(\varphi^{\dagger }\varphi+\overline{\psi^{\dagger}\psi})\mathbf{%
3818: \Lambda.} \label{GL2}
3819: \end{gather}%
3820: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
3821: %BeginExpansion
3822: \end{widetext}
3823: %EndExpansion
3824: This form of two coupled nonlinear equations is known as Ginzburg-Landau
3825: equations \cite{GL,Lifshitz}. Observing that%
3826: \begin{equation}
3827: \langle\mathbf{B}\rangle=-\nabla\times\mathbf{\Lambda,}
3828: \end{equation}
3829: where $\mathbf{B}$ denotes the magnetic field strength, one can specify%
3830: \begin{equation}
3831: \nabla\cdot\mathbf{\Lambda}=0, \label{CoulombLambda}
3832: \end{equation}
3833: which indicates that there are only two independent order parameters for the
3834: electromagnetic field. Eq. (\ref{CoulombLambda}) can be regarded as the
3835: Coulomb gauge to the vector order parameter $\mathbf{\Lambda}$.
3836:
3837: From Eqs. (\ref{CoupledPhiLambda}--\ref{CoulombLambda}), one can easily
3838: verify that there is no BEC for the electromagnetic field $\mathbf{A}$ if it
3839: is not coupled with the Higgs field $\psi $, or it is coupled with $\psi $
3840: but the latter has not condensed yet ($\varphi =0$). This also confirms the
3841: conclusion of Sec. \ref{IPGSB}: an ideal photon gas, or a free
3842: electromagnetic filed, can not produce BEC. Once the Higgs field $\psi $
3843: gets condensed ($\varphi \neq 0$), the electromagnetic field $\mathbf{A}$
3844: must condense down ($\mathbf{\Lambda }\neq 0$) simultaneously; otherwise,
3845: there can arise a linear term of $\delta \mathbf{\Lambda }$ in $\delta S$,
3846: the coupled system will go instable. This shows that the gauge field $%
3847: \mathbf{A}$ will condense together with the Higgs fields $\psi $ when it is
3848: coupled with the latter. In such case, the transition is cooperative, one
3849: usually says that $\varphi $ is the primary order parameter, and $\mathbf{%
3850: \Lambda }$ the secondary one.
3851:
3852: With the help of Eq. (\ref{CoulombLambda}), Eq. (\ref{GL2}) can be reduced as%
3853: \begin{equation}
3854: \nabla^{2}\mathbf{\Lambda}-2e^{2}(\varphi^{\dagger}\varphi+\overline {%
3855: \psi^{\dagger}\psi})\mathbf{\Lambda}=-ie(\varphi^{\dagger}\nabla
3856: \varphi-\varphi\nabla\varphi^{\dagger}). \label{ProcaEq}
3857: \end{equation}
3858: Heeding that $\langle\mathbf{A}\rangle=-\mathbf{\Lambda}$, Eq. (\ref{ProcaEq}%
3859: ) is identical to the static Proca equation \cite{Jackson} except that the
3860: prefactor of $\mathbf{\Lambda}$ may be a function of $\mathbf{r}$, the
3861: counterpart in Proca equation just a constant. In general, the prefactor
3862: cannot be interpreted as the mass of the gauge field. However, once there is
3863: a constant occurring in the prefactor, a mass $\mu$ is generated to the
3864: gauge field, as can be seen more clearly in the following discussion.
3865:
3866: Owing to the nonlinearity, the Ginzburg-Landau equations (\ref{GL1}) and (%
3867: \ref{GL2}) are rather complicated to handle. As usual, let us consider the
3868: zeroth-order approximation of Eq. (\ref{GL1}) at $T=0\mathrm{K}$ \cite%
3869: {GL,Lifshitz},
3870: \begin{equation}
3871: \varphi^{\dagger}(\mathbf{x})\varphi(\mathbf{x})=\nu^{2}/2, \label{NuTZero}
3872: \end{equation}
3873: where the effect of $\mathbf{\Lambda}$\ and\ the zero-point fluctuations of
3874: the fields $\psi$ and $\mathbf{A}$ are omitted. From Eqs. (\ref{ProcaEq})
3875: and (\ref{NuTZero}), it follows that%
3876: \begin{equation}
3877: \nabla^{2}\langle\mathbf{A}\rangle-e^{2}\nu^{2}\langle\mathbf{A}\rangle=0.
3878: \label{GLPicture}
3879: \end{equation}
3880: Since the prefactor $e^{2}\nu^{2}$ is constant now, it can be interpreted as
3881: the mass of the gauge field at zero temperature,\
3882: \begin{equation}
3883: \mu^{2}(T=0)=e^{2}\nu^{2}. \label{MuTZero}
3884: \end{equation}
3885: As well-know, this mass can account for the Meissner effect \cite{Meissner},
3886: i.e., the expulsion of a magnetic field from the interior of a
3887: superconducting material, with a London penetration length $%
3888: \lambda_{L}=\mu^{-1}$ \cite{London}.
3889:
3890: The above formalism is Ginzburg-Landau picture of mass producing, in this
3891: picture, the mass is represented by the equation of motion. There is another
3892: picture, which is due to Higgs \cite{Higgs1,Higgs2,Higgs3}. In Higgs
3893: picture, the mass is represented directly by the system Hamiltonian itself.
3894: Since an equation of motion is in accordance with its Hamiltonian, both
3895: pictures are equivalent physically. To show Higgs picture, we aught to
3896: consider the zero-temperature representation $H_{0}$ of the system
3897: Hamiltonian,%
3898: \begin{align}
3899: H_{0} & =e^{iD\left[ \varphi,\pi\mathbf{;\Lambda},\mathbf{E}\right] }H\left[
3900: \psi,\pi;\mathbf{A,E}\right] e^{-iD\left[ \varphi,\pi \mathbf{;\Lambda},%
3901: \mathbf{E}\right] } \notag \\
3902: & \equiv H[\chi,\pi;\mathbf{W,E}],
3903: \end{align}
3904: where%
3905: \begin{align}
3906: \chi(\mathbf{x}) & =e^{iD\left[ \varphi,\pi\mathbf{;\Lambda},\mathbf{E}%
3907: \right] }\psi(\mathbf{x})e^{-iD\left[ \varphi,\pi\mathbf{;\Lambda },\mathbf{E%
3908: }\right] }, \label{ChiField} \\
3909: \mathbf{W}(\mathbf{x}) & =e^{iD\left[ \varphi,\pi\mathbf{;\Lambda },\mathbf{E%
3910: }\right] }\mathbf{A}(\mathbf{x})e^{-iD\left[ \varphi ,\pi\mathbf{;\Lambda},%
3911: \mathbf{E}\right] }.
3912: \end{align}
3913: As is well known, within the temporal gauge, the electromagnetic field still
3914: has gauge degrees of freedom, the Hamiltonian $H[\psi,\pi;\mathbf{A,E}]$
3915: remains invariant under any time-independent gauge transformation. As
3916: another representation of $H[\psi,\pi;\mathbf{A,E}]$, the Hamiltonian $H%
3917: \left[ \chi,\pi;\mathbf{W,E}\right] $ is also gauge invariant,%
3918: \begin{equation}
3919: H\left[ \chi,\pi;\mathbf{W,E}\right] =e^{iG\left[ \theta,\chi ,\pi,\mathbf{E}%
3920: \right] }H\left[ \chi,\pi;\mathbf{W,E}\right] e^{-iG\left[ \theta,\chi,\pi,%
3921: \mathbf{E}\right] },
3922: \end{equation}
3923: where $\theta\in%
3924: %TCIMACRO{\U{211d} }%
3925: %BeginExpansion
3926: \mathbb{R}
3927: %EndExpansion
3928: $ and%
3929: \begin{equation}
3930: G\left[ \theta,\chi,\pi,\mathbf{E}\right] =\int\mathrm{d}\mathbf{x}\text{ }%
3931: [-ie\theta\left( \pi\chi-\chi^{\dagger}\pi^{\dagger}\right) +\nabla
3932: \theta\cdot\mathbf{E}].
3933: \end{equation}
3934:
3935: As Higgs \cite{Higgs1,Higgs2,Higgs3} did, we\ can re-gauge the fields $\chi(%
3936: \mathbf{x})$ and $\mathbf{W}(\mathbf{x})$\ by letting $\theta (\mathbf{x})$
3937: equal to the phase of $\chi(\mathbf{x})$, that is,%
3938: \begin{equation}
3939: \chi(\mathbf{x})=\rho(\mathbf{x})e^{i\theta(\mathbf{x})},\text{\ }\rho(%
3940: \mathbf{x})\in%
3941: %TCIMACRO{\U{211d} }%
3942: %BeginExpansion
3943: \mathbb{R}
3944: %EndExpansion
3945: ,
3946: \end{equation}
3947: so as to eliminate the Goldstone bosons represented by the phase of $\chi(%
3948: \mathbf{x})$. Upon such a choice of $\theta(\mathbf{x})$, the Hamiltonian $%
3949: H_{0}$ can be expressed as%
3950: \begin{align}
3951: \hspace{-0.2in}H_{0} & =\int\mathrm{d}\mathbf{x}\text{ }\Big[\widetilde{\pi }%
3952: ^{\dagger}\widetilde{\pi}+(\nabla+ie\widetilde{\mathbf{A}})\rho\cdot
3953: (\nabla-ie\widetilde{\mathbf{A}})\rho \notag \\
3954: & +m^{2}\rho^{2}+g\rho^{4}+\frac{1}{2}\mathbf{E}^{2}+\frac{1}{2}(\nabla
3955: \times\widetilde{\mathbf{A}})^{2}\Big], \label{HCoupled}
3956: \end{align}
3957: where
3958: \begin{align}
3959: \widetilde{\pi}(\mathbf{x}) & =e^{iG\left[ \theta,\chi,\pi,\mathbf{E}\right]
3960: }\pi(\mathbf{x})e^{-iG\left[ \theta,\chi,\pi,\mathbf{E}\right] },
3961: \label{PiWave} \\
3962: \widetilde{\mathbf{A}}(\mathbf{x}) & =e^{iG\left[ \theta,\chi,\pi ,\mathbf{E}%
3963: \right] }\mathbf{W}(\mathbf{x})e^{-iG\left[ \theta,\chi ,\pi,\mathbf{E}%
3964: \right] }.
3965: \end{align}
3966:
3967: From Eqs. (\ref{ChiField}) and (\ref{DPiLambda}), it follows that
3968: \begin{equation}
3969: \chi(\mathbf{x})=\psi(\mathbf{x})+\varphi(\mathbf{x}).
3970: \end{equation}
3971: Without loss of generality, we can, from Eq. (\ref{NuTZero}), choose $%
3972: \varphi(\mathbf{x})$ as real,%
3973: \begin{equation}
3974: \varphi(\mathbf{x})=\nu/\sqrt{2}.
3975: \end{equation}
3976: That leads to%
3977: \begin{equation}
3978: \rho(\mathbf{x})=\left( \phi(\mathbf{x})+\nu\right) /\sqrt{2},
3979: \label{RhoPhi}
3980: \end{equation}
3981: where%
3982: \begin{equation}
3983: \phi(\mathbf{x})=\sqrt{2}e^{iG\left[ \theta,\widetilde{\psi},\pi ,\mathbf{E}%
3984: \right] }\psi(\mathbf{x})e^{-iG\left[ \theta,\widetilde{\psi},\pi,\mathbf{E}%
3985: \right] }.
3986: \end{equation}
3987: Eq. (\ref{RhoPhi}) indicates that $\phi(\mathbf{x})$ is merely a real field.
3988:
3989: Substitution of Eq. (\ref{RhoPhi}) into Eq. (\ref{HCoupled}) gives rise to%
3990: \begin{align}
3991: H_{0} & =\int\mathrm{d}\mathbf{x}\text{ }\Big[\frac{1}{2}p^{2}+\frac{1}{2}%
3992: (\nabla+ie\widetilde{\mathbf{A}})\phi\cdot(\nabla-ie\widetilde{\mathbf{A}}%
3993: )\phi \notag \\
3994: & -m^{2}\phi^{2}+g\nu\phi^{3}+\frac{g}{4}\phi^{4}+\frac{1}{2}\mathbf{E}^{2}+%
3995: \frac{1}{2}(\nabla\times\widetilde{\mathbf{A}})^{2} \notag \\
3996: & +\frac{1}{2}e^{2}\nu^{2}\widetilde{\mathbf{A}}\cdot\widetilde{\mathbf{A}}%
3997: +e^{2}\nu\phi\widetilde{\mathbf{A}}\cdot\widetilde{\mathbf{A}} \notag \\
3998: & -\frac{g\nu^{4}}{4}+\frac{1}{2}p_{0}^{2}\Big], \label{HiggsRepre}
3999: \end{align}
4000: where%
4001: \begin{align}
4002: p(\mathbf{x}) & =[\widetilde{\pi}^{\dagger}(\mathbf{x})+\widetilde{\pi }(%
4003: \mathbf{x})]/\sqrt{2}, \\
4004: p_{0}(\mathbf{x}) & =i[\widetilde{\pi}^{\dagger}(\mathbf{x})-\widetilde{\pi }%
4005: (\mathbf{x})]/\sqrt{2}.
4006: \end{align}
4007: Paying attention to%
4008: \begin{equation}
4009: e^{iG\left[ \theta,\chi,\pi,\mathbf{E}\right] }\left( \frac{\chi (\mathbf{x}%
4010: )+\chi^{\dagger}(\mathbf{x})}{\sqrt{2}}\right) e^{-iG\left[ \theta,\chi,\pi,%
4011: \mathbf{E}\right] }=\phi(\mathbf{x})+\nu,
4012: \end{equation}
4013: we have
4014: \begin{gather}
4015: \left[ p(\mathbf{x}),\phi(\mathbf{x}^{\prime})\right] =-i\delta (\mathbf{x}-%
4016: \mathbf{x}^{\prime}), \\
4017: \left[ p_{0}(\mathbf{x}),H_{0}\right] =0.
4018: \end{gather}
4019: That is to say, the $\phi(\mathbf{x})$ and $p(\mathbf{x})$ become a new pair
4020: of canonical field operators whereas $p_{0}(\mathbf{x})$ is just a constant.
4021: Physically, that is because $p_{0}(\mathbf{x})$ corresponds to the
4022: eliminated degree of freedom $\phi_{0}(\mathbf{x})=-i\left( \chi(\mathbf{x}%
4023: )-\chi^{\dagger}(\mathbf{x})\right) /\sqrt{2}$.
4024:
4025: Eq. (\ref{HiggsRepre}) forms Higgs picture of mass producing. Comparing it
4026: with Eq. (\ref{GLPicture}), one sees immediately that the mass of the gauge
4027: field in Higgs picture is identical to that in Ginzburg-Landau picture, as
4028: is expected.
4029:
4030: The above theory for a complex field coupled with an Abelian gauge field can
4031: be easily generalized to the case of an $N$-component field coupled with a
4032: non-Abelian gauge field, with the same conclusion that the gauge field will
4033: obtain a mass through spontaneous symmetry breaking. This manner of mass
4034: producing is the well-known Higgs mechanism, which was successively
4035: contributed by London \cite{London}, Ginzburg and Landau \cite{GL}, Anderson
4036: \cite{Anderson1,Anderson2}, Higgs \cite{Higgs1,Higgs2,Higgs3}, and Kibble
4037: \cite{Kibble}.
4038:
4039: \textbf{Remark}: Spontaneous symmetry breaking can occur without any
4040: fundamental Higgs field in superconductivity, it is thus called "dynamical
4041: symmetry breaking", which implies that Higgs fields are "normally" needed
4042: for symmetry breaking. From the viewpoint of the extended ensemble theory,
4043: that is a prejudice because superconductivity, Goldstone bosons and Higgs
4044: mechanism share the same physical ground. This is in accordance with the
4045: standpoint of Huang expressed in his book \cite{Huang2}.
4046:
4047: \section{Summary and Conclusions}
4048:
4049: So far, a possible extension of Gibbs ensemble theory has been postulated,
4050: as a personal review of the author, to enable a microscopic description to
4051: phase transitions and spontaneous symmetry breaking. The extension is
4052: founded on three hypotheses, which root, in physics, from Landau's ideas on
4053: phase transitions: order parameter, variational principle, representation
4054: transformation, and spontaneous symmetry breaking. In this sense, the
4055: extended Gibbs ensemble theory can be viewed as a microscopic realization of
4056: the Landau phenomenological theory of phase transitions.
4057:
4058: Within the framework of the extended ensemble theory, a phase transition
4059: occurs according to the principle of least entropy, which manifests itself
4060: as an equation of motion of order parameter. This equation determines the
4061: evolution of order parameter with temperature, and thus controls the change
4062: in representation of system Hamiltonian. Different phases correspond to
4063: different representations, and vice versa. A system Hamiltonian will realize
4064: its symmetric representation in the disordered phase, and asymmetric one in
4065: the ordered phase. The change in symmetry results from the change in
4066: representation. That is the Landau mechanism responsible for phase
4067: transitions and spontaneous symmetry breaking in the extended ensemble
4068: theory, it holds in the whole range of temperature, including the critical
4069: region.
4070:
4071: Physically, phase transitions originate from the wave nature of matter. A
4072: system always stays in its normal phase at sufficiently high temperatures,
4073: which is disordered and structureless. As temperature decreases, matter
4074: waves will interfere automatically with one another. This interference makes
4075: the system structured and transform into its ordered phase. That is the
4076: physical picture for phase transitions in the extended ensemble theory.
4077:
4078: Also, the extended ensemble theory has been applied to typical quantum
4079: many-body systems, with the conclusions as follows.
4080:
4081: For the ideal Fermi gas, it can not produce superconductivity, its normal
4082: phase is stable at any temperature.
4083:
4084: Negative-temperature and laser phases arise from the same mechanism as phase
4085: transitions, they are both instable, and will finally turn into the
4086: positive-temperature phase. That is the microscopic interpretation for
4087: negative temperatures and laser phases.
4088:
4089: For the conventional weak-coupling low-$T_{c}$ superconductors, the
4090: mean-field solution due to Bardeen, Cooper and Schrieffer is derived anew,
4091: and proved to be stable within the extended ensemble theory.
4092:
4093: For the ideal Bose gas, it can produce Bose-Einstein condensation only in
4094: the thermodynamic limit, which agrees with Einstein's prediction. But that
4095: holds in the sense of Lebesgue integration rather than Riemann integration.
4096: In general, the order parameter for BEC can not be interpreted as the number
4097: of condensed particles. Besides, there can not exist any supercurrent or
4098: vortex in the ideal Bose gas, the condensation is always homogeneous.
4099:
4100: The ideal phonon gas can not produce Bose-Einstein condensation, neither the
4101: photon gas in a black body.
4102:
4103: It is not admissible to quantize Dirac field using Bose-Einstein statistics.
4104: Otherwise, the system is instable. That is a statistical rather than
4105: mechanical reason why Dirac field has to be quantized using Fermi-Dirac
4106: statistics. This conclusion is in accordance with the spin-statistics
4107: theorem in quantum field theory.
4108:
4109: A structural phase transition belongs physically to the Bose-Einstein
4110: condensation occurring in configuration space. For a double-well anharmonic
4111: system, it is instable at low temperatures, and will undergo a structural
4112: phase transition at a finite temperature. Physically, that is because the
4113: position fluctuation is a monotonically decreasing function of temperature.
4114:
4115: For the $O(N)$-symmetric vector model, the $O(N)$ symmetry will break down
4116: spontaneously, with the presence of Goldstone bosons. In essence, this SSB
4117: is a Bose-Einstein condensation. If the system is coupled with a gauge
4118: field, it can cause the latter to condense together with itself. The
4119: cooperative condensation can be described by the Ginzburg-Landau equations.
4120: After the condensation, the gauge field can obtain a mass, which is the
4121: so-called Higgs mechanism.
4122:
4123: For an interacting Bose gas, it is stable only if the interaction is
4124: repulsive. The BEC present in this system can be described by the
4125: generalized Ginzburg-Landau equation. If the interaction is weak, the BEC is
4126: a \textquotedblleft$\lambda$\textquotedblright-transition, and its
4127: transition temperature can be lowered by the repulsive interaction. As a
4128: characteristic property of this \textquotedblleft$\lambda$%
4129: \textquotedblright-transition, the specific heat at constant volume $C_{V}$
4130: will vanish linearly as $T\rightarrow0$.
4131:
4132: If liquid $^{4}\mathrm{He}$ could be regarded as a weakly interacting Bose
4133: gas, and if the $\lambda $-transition were a Bose-Einstein condensation,
4134: then its specific heat at constant pressure $C_{P}$ would show a $T^{3}$ law
4135: at low temperatures, which is in agreement with the experiment. However, if
4136: temperature goes lower further, it is predicted that $C_{P}$ will exhibit a
4137: linear behavior at relatively lower temperatures, which is in need of
4138: experimental verifications.
4139:
4140: \begin{acknowledgments}
4141: The author would like to thank Prof. Zheng-zhong Li, Prof. Jin-ming Dong,
4142: and Prof. Wei-yi Zhang for their helpful discussions. He is also deeply
4143: grateful to Prof. Vladimir Rittenberg for his valuable advice on how to
4144: revise the paper.
4145: \end{acknowledgments}
4146:
4147: \newpage
4148:
4149: \appendix
4150:
4151: \section{Frequency Summation\label{FSum}}
4152:
4153: In this appendix, we shall first show that, when $\mu>0$, the ideal Bose gas
4154: is still well defined within the extended ensemble theory, and then discuss
4155: the frequency summation, which is very useful when we attempt to solve
4156: problems through Green's functions.
4157:
4158: Upon a transformation as in Eq. (\ref{Sphi}), the statistical average
4159: defined by Eq. (\ref{Average}) can always be transformed into a statistical
4160: average with respect to the Hamiltonian $H(b)$ of Eq. (\ref{Hibg}). It is
4161: thus sufficient for us to show that the statistical average with respect to $%
4162: H(b)$ is properly defined when $\mu >0$. As is well known, every statistical
4163: average can be calculated through a corresponding Green's function, and
4164: every Green's function can be derived from the generating functional of the
4165: system. Therefore, we need only to prove that the generating functional is
4166: properly defined when $\mu >0$.
4167:
4168: According to the path-integral formalism \cite{Negele}, the generating
4169: functional $W[\phi ^{+},\phi ]$ for $H(b)$ can be written as follows,
4170: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}}}%
4171: %BeginExpansion
4172: \begin{widetext}%
4173: %EndExpansion
4174: \begin{equation}
4175: W[\phi ^{+},\phi ]=\mathcal{N}^{-1}\int \mathcal{D}b^{+}\mathcal{D}b\text{%
4176: \thinspace }\exp \!\!\left( \frac{1}{\hbar }\sum_{\mathbf{k},n}\left\{ b_{%
4177: \mathbf{k,}n}^{+}\left[ i\hbar \omega _{n}-\left( \frac{\hbar ^{2}\mathbf{k}%
4178: ^{2}}{2m}-\mu \right) \right] b_{\mathbf{k},n}-\phi _{\mathbf{k},n}b_{%
4179: \mathbf{k,}n}^{+}-\phi _{\mathbf{k,}n}^{+}b_{\mathbf{k},n}\right\} \right) ,
4180: \label{WGenerate}
4181: \end{equation}%
4182: where $\omega _{n}=2n\pi /\beta \hbar $ ($n\in
4183: %TCIMACRO{\U{2124} }%
4184: %BeginExpansion
4185: \mathbb{Z}
4186: %EndExpansion
4187: $) denotes the Mutsubara frequency, and $\mathcal{N}$ the normalization
4188: factor,
4189: \begin{equation}
4190: \mathcal{N}=\int \mathcal{D}b^{+}\mathcal{D}b\text{\thinspace }\exp
4191: \!\!\left( \frac{1}{\hbar }\sum_{\mathbf{k},n}b_{\mathbf{k,}n}^{+}\left[
4192: i\hbar \omega _{n}-\left( \frac{\hbar ^{2}\mathbf{k}^{2}}{2m}-\mu \right) %
4193: \right] b_{\mathbf{k},n}\right) . \label{NGaussian}
4194: \end{equation}
4195:
4196: If $\mu <0$, one gets immediately the familiar result,
4197: \begin{equation}
4198: W[\phi ^{+},\phi ]=\exp \!\!\left( -\frac{1}{\hbar }\sum_{\mathbf{k},n}\phi
4199: _{\mathbf{k},n}\frac{1}{i\hbar \omega _{n}-\left( \frac{\hbar ^{2}\mathbf{k}%
4200: ^{2}}{2m}-\mu \right) }\phi _{\mathbf{k,}n}^{+}\right) . \label{WWG}
4201: \end{equation}%
4202: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
4203: %BeginExpansion
4204: \end{widetext}
4205: %EndExpansion
4206: Since $\hbar ^{2}\mathbf{k}^{2}/\left( 2m\right) -\mu >0$, the normalization
4207: factor $\mathcal{N}$ converges.
4208:
4209: When $\mu \geq 0$, the normalization factor $\mathcal{N}$ will diverge.
4210: However, that does no harm because the normalization factor can be cancelled
4211: by the numerator of Eq. (\ref{WGenerate}), as is frequently encountered in
4212: quantum field theory \cite{Greiner}. After the cancellation, the rest is
4213: still Eq. (\ref{WWG}) except $\mu \geq 0$. As mentioned in Sec. \ref%
4214: {BECIdeal}, an unbounded function as integrand is permissible within
4215: Lebesgue integration. Therefore, the sum in the exponent of Eq. (\ref{WWG})
4216: is proper as a Lebesgue integral. The resulting $W[\phi ^{+},\phi ]$ is also
4217: right because a function is permitted to be unbounded in Lebesgue
4218: integration.
4219:
4220: In a word, the generating functional for $H(b)$ is always properly defined
4221: no matter how large the chemical potential $\mu $ is, which ends our proof.
4222:
4223: Now, there is no mathematical limit on the chemical potential of the ideal
4224: Bose gas, as is the case for the ideal Fermi gas. That is rational and
4225: significant, any limit on chemical potential should be set by the physical
4226: theory itself rather than by the mathematical theory involved. In the
4227: extended ensemble theory, the chemical potential of a system is determined
4228: only by the physical requirement: the conservation of particles, whether the
4229: system obeys Bose-Einstein statistics or Fermi-Dirac statistics; there is no
4230: limit set by the mathematical tool, i.e., Lebesgue integration.
4231:
4232: As a consequence of Eq. (\ref{WWG}), we obtain%
4233: \begin{align}
4234: \mathcal{G}(\mathbf{k},i\omega_{n}) & =-\left( -\frac{1}{\hbar}\right) ^{-2}%
4235: \frac{\delta^{2}W[\phi^{+},\phi]}{\delta\phi_{\mathbf{k,}n}^{+}\delta \phi_{%
4236: \mathbf{k},n}} \notag \\
4237: & =\frac{1}{i\omega_{n}-\hbar^{-1}\left( \frac{\hbar^{2}\mathbf{k}^{2}}{2m}%
4238: -\mu\right) },
4239: \end{align}
4240: where $\mathcal{G}(\mathbf{k},i\omega_{n})$ represents the temperature
4241: Green's function which is defined as \cite{Fetter}%
4242: \begin{equation}
4243: \mathcal{G}(\mathbf{k},\tau-\tau^{\prime})=-\mathrm{Tr}\big(T_{\tau }\{b_{%
4244: \mathbf{k}}(\tau)b_{\mathbf{k}}^{\dagger}(\tau^{\prime})\}\rho (H(b))\big).
4245: \end{equation}
4246: As a function of $\mathbf{k}$, $\mathcal{G}(\mathbf{k},i\omega_{n})$ can be
4247: unbounded if $\mu\geq0$, but that will be all right because $\mathcal{G}(%
4248: \mathbf{k},i\omega_{n})$ can only appear in the integrand of a Lebesgue
4249: integral over $\mathbf{k}$.
4250:
4251: Those discussions indicate that, in the manipulations of $+\infty $ and $%
4252: -\infty $, one can benefit greatly from Lebesgue integration. That is
4253: because Lebesgue integration is designed, \textit{ab initio}, on the set of
4254: extended real numbers: $%
4255: %TCIMACRO{\U{211d} }%
4256: %BeginExpansion
4257: \mathbb{R}
4258: %EndExpansion
4259: ^{\sharp }=%
4260: %TCIMACRO{\U{211d} }%
4261: %BeginExpansion
4262: \mathbb{R}
4263: %EndExpansion
4264: \cup \{+\infty \}\cup \{-\infty \}$ where $%
4265: %TCIMACRO{\U{211d} }%
4266: %BeginExpansion
4267: \mathbb{R}
4268: %EndExpansion
4269: =(-\infty ,+\infty )$ \cite{Hewitt}. Evidently, those benefits can not be
4270: provided by Riemann integration. More benefits of Lebesgue integration will
4271: be seen from the following discussions.
4272:
4273: When one evaluates the statistical average of an observable, he will
4274: encounter the summation over Mutsubara frequencies. As an illustration, let
4275: us consider
4276: \begin{align}
4277: \int_{%
4278: %TCIMACRO{\U{211d} }%
4279: %BeginExpansion
4280: \mathbb{R}
4281: %EndExpansion
4282: ^{3}}\mathrm{d}\mathbf{k\,}\overline{b_{\mathbf{k}}^{\dag }b_{\mathbf{k}}}&
4283: =\int_{%
4284: %TCIMACRO{\U{211d} }%
4285: %BeginExpansion
4286: \mathbb{R}
4287: %EndExpansion
4288: ^{3}}\mathrm{d}\mathbf{k\,}\mathrm{Tr}\big(b_{\mathbf{k}}^{\dag }b_{\mathbf{k%
4289: }}\rho (H(b))\big) \notag \\
4290: & =\int_{%
4291: %TCIMACRO{\U{211d} }%
4292: %BeginExpansion
4293: \mathbb{R}
4294: %EndExpansion
4295: ^{3}}\mathrm{d}\mathbf{k\,}\left( -\frac{1}{\beta \hbar }\right)
4296: \sum_{n}e^{i\omega _{n}\eta }\mathcal{G}(\mathbf{k},i\omega _{n}),
4297: \label{Summation}
4298: \end{align}%
4299: where $\eta =0^{+}$. We shall discuss the thermodynamic limit case. Here, it
4300: is important to note that the frequency summation always goes ahead of the
4301: Lebesgue integral over $\mathbf{k}$ because the thermodynamic limit must be
4302: taken at last.
4303:
4304: To perform the frequency summation, it is helpful to make use of the complex
4305: function,
4306: \begin{equation}
4307: f(z)=\frac{1}{e^{\beta \hbar z}-1},
4308: \end{equation}%
4309: which has simple poles at the Mutsubara frequencies, i.e., $z=i\omega
4310: _{n}=i2n\pi /\beta \hbar $, and transform the summation into a contour
4311: integral \cite{Fetter},%
4312: \begin{equation}
4313: \left( -\frac{1}{\beta \hbar }\right) \sum_{n}e^{i\omega _{n}\eta }\mathcal{G%
4314: }(\mathbf{k},i\omega _{n})=\int_{C}\frac{\mathrm{d}z}{2\pi i}e^{i\eta z}f(z)%
4315: \mathcal{G}(\mathbf{k},z),
4316: \end{equation}%
4317: where the contour $C$ is depicted in Fig. 7. It should be emphasized that
4318: the above procedure requires that the function $\mathcal{G}(\mathbf{k},z)$
4319: must be analytic on the whole imaginary axis lest there be any poles other
4320: than $i\omega _{n}$ on this axis. Sometimes, this requirement can not be met
4321: by every $\mathbf{k}$, e.g.,%
4322: \begin{equation}
4323: \mathcal{G}(\mathbf{k},z)=\frac{1}{z-\hbar ^{-1}(\frac{\hbar ^{2}\mathbf{k}%
4324: ^{2}}{2m}-\mu )},\text{ }\mu \geq 0. \label{Singular}
4325: \end{equation}%
4326: Each such $\mathcal{G}(\mathbf{k},z)$ that satisfies $|\mathbf{k|}=\sqrt{%
4327: 2m\mu }/\hbar $ is singular on $z=0$, and thus can not meet the requirement.
4328: If such singular $\mathbf{k}$'s constitute just a null set $D$, i.e., $%
4329: \mathcal{G}(\mathbf{k},z)$ satisfies the requirement almost everywhere on $%
4330: %TCIMACRO{\U{211d} }%
4331: %BeginExpansion
4332: \mathbb{R}
4333: %EndExpansion
4334: ^{3}$, which is the usual case, then those $\mathbf{k}$'s can be left out of
4335: consideration for they give merely a null contribution to the Lebesgue
4336: integral over $\mathbf{k}$. Therefore, we can rewrite Eq. (\ref{Summation})
4337: as
4338: \begin{align}
4339: & \int_{%
4340: %TCIMACRO{\U{211d} }%
4341: %BeginExpansion
4342: \mathbb{R}
4343: %EndExpansion
4344: ^{3}}\mathrm{d}\mathbf{k\,}\left( -\frac{1}{\beta \hbar }\right)
4345: \sum_{n}e^{i\omega _{n}\eta }\mathcal{G}(\mathbf{k},i\omega _{n}) \notag \\
4346: & =\int_{E}\mathrm{d}\mathbf{k}\int_{C}\frac{\mathrm{d}z}{2\pi i}e^{i\eta
4347: z}f(z)\mathcal{G}(\mathbf{k},z), \label{Sum0}
4348: \end{align}%
4349: where $E=%
4350: %TCIMACRO{\U{211d} }%
4351: %BeginExpansion
4352: \mathbb{R}
4353: %EndExpansion
4354: ^{3}\backslash D$. Here the set $D$ is removed form $%
4355: %TCIMACRO{\U{211d} }%
4356: %BeginExpansion
4357: \mathbb{R}
4358: %EndExpansion
4359: ^{3}$ so that the function $\mathcal{G}(\mathbf{k},z)$ can meet the
4360: requirement everywhere on $E$.
4361: %TCIMACRO{%
4362: %\TeXButton{Fig7}{\begin{figure}[htbp]
4363: %\hspace*{-1.7cm}\includegraphics[scale=0.45,angle=-90]{Fig7.eps}
4364: %\caption{The counters for the frequency summation of a Bose
4365: %system.}
4366: %\end{figure}}}%
4367: %BeginExpansion
4368: \begin{figure}[htbp]
4369: \hspace*{-1.7cm}\includegraphics[scale=0.45,angle=-90]{Fig7.eps}
4370: \caption{The counters for the frequency summation of a Bose
4371: system.}
4372: \end{figure}%
4373: %EndExpansion
4374:
4375: Generally, $\mathcal{G}(\mathbf{k},z)$ are analytic on both the upper and
4376: lower half complex-$z$ planes \cite{Zubarev,Fetter}, one can therefore
4377: deform the contour $C$ into the contours $\Gamma $ and $C^{\prime }$, which
4378: results in
4379: \begin{equation}
4380: \int_{C}\frac{\mathrm{d}z}{2\pi i}e^{i\eta z}f(z)\mathcal{G}(\mathbf{k}%
4381: ,z)=\int_{C^{\prime }}\frac{\mathrm{d}z}{2\pi i}f(z)\mathcal{G}(\mathbf{k}%
4382: ,z). \label{Sum1}
4383: \end{equation}%
4384: The contribution from the contour $\Gamma $ vanishes owing to the
4385: convergence factor $e^{i\eta z}$ \cite{Fetter}. The right-hand side can be
4386: simplified as
4387: \begin{equation}
4388: \int_{C^{\prime }}\frac{\mathrm{d}z}{2\pi i}f(z)\mathcal{G}(\mathbf{k},z)=%
4389: \mathcal{P}\int_{-\infty }^{+\infty }\mathrm{d}\omega \,f(\omega )\mathcal{A}%
4390: (\mathbf{k},\omega ),
4391: \end{equation}%
4392: where $\mathcal{A}(\mathbf{k},\omega )$ is the spectral intensity,
4393: \begin{equation}
4394: \mathcal{A}(\mathbf{k},\omega )=-\frac{1}{2\pi i}\left[ \mathcal{G}(\mathbf{k%
4395: },\omega +i0^{+})-\mathcal{G}(\mathbf{k},\omega -i0^{+})\right] ,
4396: \end{equation}%
4397: and $\mathcal{P}$ represents the principal value,
4398: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}} }%
4399: %BeginExpansion
4400: \begin{widetext}
4401: %EndExpansion
4402: \begin{equation}
4403: \mathcal{P}\int_{-\infty }^{+\infty }\mathrm{d}\omega \,f(\omega )\mathcal{G}%
4404: (\mathbf{k},\omega )=\lim_{\alpha \rightarrow 0^{+}}\left( \int_{-\infty
4405: }^{-\alpha }\mathrm{d}\omega \,f(\omega )\mathcal{A}(\mathbf{k},\omega
4406: )+\int_{\alpha }^{+\infty }\mathrm{d}\omega \,f(\omega )\mathcal{A}(\mathbf{k%
4407: },\omega )\right) . \label{Sum2}
4408: \end{equation}
4409:
4410: Substituting Eqs. (\ref{Sum1}--\ref{Sum2}) into Eq. (\ref{Sum0}), one has
4411: \begin{eqnarray}
4412: \int_{%
4413: %TCIMACRO{\U{211d} }%
4414: %BeginExpansion
4415: \mathbb{R}
4416: %EndExpansion
4417: ^{3}}\mathrm{d}\mathbf{k}\left( -\frac{1}{\beta \hbar }\right)
4418: \sum_{n}e^{i\omega _{n}\eta }\mathcal{G}(\mathbf{k},i\omega _{n}) &=&\int_{E}%
4419: \mathrm{d}\mathbf{k}\lim_{\alpha \rightarrow 0^{+}}\left[ \int_{-\infty
4420: }^{-\alpha }\mathrm{d}\omega \,f(\omega )\mathcal{A}(\mathbf{k},\omega
4421: )+\int_{\alpha }^{+\infty }\mathrm{d}\omega \,f(\omega )\mathcal{A}(\mathbf{k%
4422: },\omega )\right] \notag \\
4423: &=&\lim_{\alpha \rightarrow 0^{+}}\int_{E}\mathrm{d}\mathbf{k}\left[
4424: \int_{-\infty }^{-\alpha }\mathrm{d}\omega \,f(\omega )\mathcal{A}(\mathbf{k}%
4425: ,\omega )+\int_{\alpha }^{+\infty }\mathrm{d}\omega \,f(\omega )\mathcal{A}(%
4426: \mathbf{k},\omega )\right] \notag \\
4427: &=&\lim_{\alpha \rightarrow 0^{+}}\left[ \int_{-\infty }^{-\alpha }\mathrm{d}%
4428: \omega \,f(\omega )\int_{E}\mathrm{d}\mathbf{k}\mathcal{A}(\mathbf{k},\omega
4429: )+\int_{\alpha }^{+\infty }\mathrm{d}\omega \,f(\omega )\int_{E}\mathrm{d}%
4430: \mathbf{k}\mathcal{A}(\mathbf{k},\omega )\right] \notag \\
4431: &=&\mathcal{P}\int_{-\infty }^{+\infty }\mathrm{d}\omega \,f(\omega )\int_{E}%
4432: \mathrm{d}\mathbf{k}\mathcal{A}(\mathbf{k},\omega ).
4433: \end{eqnarray}%
4434: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}}}%
4435: %BeginExpansion
4436: \end{widetext}%
4437: %EndExpansion
4438:
4439: Now, supplementing the set $E$ with the null set $D$, one arrives at%
4440: \begin{align}
4441: & \int_{%
4442: %TCIMACRO{\U{211d} }%
4443: %BeginExpansion
4444: \mathbb{R}
4445: %EndExpansion
4446: ^{3}}\mathrm{d}\mathbf{k}\left( -\frac{1}{\beta \hbar }\right)
4447: \sum_{n}e^{i\omega _{n}\eta }\mathcal{G}(\mathbf{k},i\omega _{n}) \notag \\
4448: & =\mathcal{P}\int_{-\infty }^{+\infty }\mathrm{d}\omega \,f(\omega )\int_{%
4449: %TCIMACRO{\U{211d} }%
4450: %BeginExpansion
4451: \mathbb{R}
4452: %EndExpansion
4453: ^{3}}\mathrm{d}\mathbf{k}\mathcal{A}(\mathbf{k},\omega ). \label{BSum}
4454: \end{align}%
4455: As usual, by introducing the density of states $\mathcal{N}(\omega )$,
4456: \begin{equation}
4457: \mathcal{N}(\omega )=\frac{1}{\left( 2\pi \right) ^{3}}\int_{%
4458: %TCIMACRO{\U{211d} }%
4459: %BeginExpansion
4460: \mathbb{R}
4461: %EndExpansion
4462: ^{3}}\mathrm{d}\mathbf{k}\mathcal{A}(\mathbf{k},\omega ),
4463: \end{equation}%
4464: Eq. (\ref{BSum}) can be reduced into%
4465: \begin{align}
4466: & \frac{1}{\left( 2\pi \right) ^{3}}\int_{%
4467: %TCIMACRO{\U{211d} }%
4468: %BeginExpansion
4469: \mathbb{R}
4470: %EndExpansion
4471: ^{3}}\mathrm{d}\mathbf{k}\left( -\frac{1}{\beta \hbar }\right)
4472: \sum_{n}e^{i\omega _{n}\eta }\mathcal{G}(\mathbf{k},i\omega _{n}) \notag \\
4473: & =\mathcal{P}\int_{-\infty }^{+\infty }\mathrm{d}\omega \frac{\mathcal{N}%
4474: (\omega )}{e^{\beta \hbar \omega }-1}. \label{Principal}
4475: \end{align}%
4476: This is a very useful identity, as its applications, one can easily obtain
4477: the final results of Eqs. (\ref{Lesb1}) and (\ref{Lesb3}). It should be
4478: stressed that Eq. (\ref{Principal}) holds whether the chemical potential $%
4479: \mu $ is less than, equal to, or lager than zero. If $\mu <0$, it reduces to
4480: the common result given in Ref. \cite{Fetter}. It is a generalization of the
4481: common result when $\mu \geq 0$. This generalization removes the
4482: mathematical limit on the chemical potential. One can not reach the result
4483: of Eq. (\ref{Principal}) within Riemann integration.
4484:
4485: Obviously, the above analyses for the ideal Bose gas are also suitable for
4486: other Bose systems, e.g., the interacting Bose gas, which will be studied in
4487: Sec. \ref{WIBG}.
4488:
4489: \textbf{Remark}:\textbf{\ }If the system obeys Fermi-Dirac statistics, the
4490: principal value is unnecessary. Instead of Eq. (\ref{Principal}), one has%
4491: \begin{align}
4492: & \frac{1}{\left( 2\pi \right) ^{3}}\int_{%
4493: %TCIMACRO{\U{211d} }%
4494: %BeginExpansion
4495: \mathbb{R}
4496: %EndExpansion
4497: ^{3}}\mathrm{d}\mathbf{k}\left( \frac{1}{\beta \hbar }\right)
4498: \sum_{n}e^{i\omega _{n}\eta }\mathcal{G}(\mathbf{k},i\omega _{n}) \notag \\
4499: & =\mathcal{P}\int_{-\infty }^{+\infty }\mathrm{d}\omega \frac{\mathcal{N}%
4500: (\omega )}{e^{\beta \hbar \omega }+1} \notag \\
4501: & =\int_{-\infty }^{+\infty }\mathrm{d}\omega \frac{\mathcal{N}(\omega )}{%
4502: e^{\beta \hbar \omega }+1}. \label{Fermi}
4503: \end{align}%
4504: That is simply because the Fermi distribution function is bounded and
4505: continuous at $\omega =0$ for any finite temperature $\beta >0$. A more
4506: direct reason for the unnecessity of the principal value consists in the
4507: fact that, unlike the Bose system, $z=0$ is not the pole of the complex
4508: function,
4509: \begin{equation}
4510: f(z)=\frac{1}{e^{\beta \hbar z}+1},
4511: \end{equation}%
4512: because the Mutsubara frequencies for a Fermi system are all nonzero, i.e., $%
4513: \omega _{n}=\left( 2n+1\right) \pi /\beta \hbar \neq 0$ ($n\in
4514: %TCIMACRO{\U{2124} }%
4515: %BeginExpansion
4516: \mathbb{Z}
4517: %EndExpansion
4518: $). As a consequence, the integration contour can goes as in Fig. 8, which
4519: gives straightforwardly the final result of Eq. (\ref{Fermi}). Evidently,
4520: that makes the Fermi system easier to handle mathematically than the Bose
4521: system.
4522: %TCIMACRO{%
4523: %\TeXButton{Fig8}{\begin{figure}[htbp]
4524: %\hspace*{-1.7cm}\includegraphics[scale=0.45,angle=-90]{Fig8.eps}
4525: %\caption{The counters for the frequency summation of a Fermi system.}
4526: %\end{figure}}}%
4527: %BeginExpansion
4528: \begin{figure}[htbp]
4529: \hspace*{-1.7cm}\includegraphics[scale=0.45,angle=-90]{Fig8.eps}
4530: \caption{The counters for the frequency summation of a Fermi system.}
4531: \end{figure}%
4532: %EndExpansion
4533:
4534: \section{Low-temperature Thermal Activations of a Weakly Interacting Bose
4535: System \label{TAWIBSLT}}
4536:
4537: In Sec. \ref{WIBG}, our discussions were confined within the Hartree-Fock
4538: approximation of Eqs. (\ref{IBECHF1}--\ref{IBECHF2}) and the short-range
4539: approximation of Eq. (\ref{IBECSR}). Within those approximations, one sees
4540: from Eqs. (\ref{X12LTc}) and (\ref{X32LTc}) that the thermal activations at
4541: low temperatures can be expanded into even power series of $T$. We shall
4542: show that, beyond those approximations, this feature still holds as long as
4543: the interaction is weak and perturbative.
4544:
4545: Under the condition assumed, observables can be calculated via GF. The
4546: calculations can be transformed finally into the integrals of the following
4547: form,
4548: \begin{equation}
4549: \mathcal{P}\int_{-\infty }^{+\infty }\mathrm{d}\omega \frac{g\left( \omega ,%
4550: \widetilde{\mu }\right) }{e^{\beta \omega }-1},
4551: \end{equation}%
4552: where $g\left( \omega ,\widetilde{\mu }\right) $ is a real-valued function
4553: of $\omega $ and $\widetilde{\mu }$, which comes from a summation of GF over
4554: $\mathbf{k}$. As a function of $\omega $, $g\left( \omega ,\widetilde{\mu }%
4555: \right) $ should be bounded at bottom or decrease fast with $\omega $ so
4556: that the integral can converge. In both cases, we can cut off the integral
4557: from below at a certain frequency $-\omega _{c}$ ($\omega _{c}>0$), that is,
4558: \begin{equation}
4559: \mathcal{P}\int_{-\infty }^{+\infty }\mathrm{d}\omega \frac{g\left( \omega ,%
4560: \widetilde{\mu }\right) }{e^{\beta \omega }-1}=\mathcal{P}\int_{-\omega
4561: _{c}}^{+\infty }\mathrm{d}\omega \frac{g\left( \omega ,\widetilde{\mu }%
4562: \right) }{e^{\beta \omega }-1}.
4563: \end{equation}%
4564: According to the definition of principal value, the right-hand side can be
4565: written as
4566: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}}}%
4567: %BeginExpansion
4568: \begin{widetext}%
4569: %EndExpansion
4570: \begin{equation}
4571: \mathcal{P}\int_{-\omega _{c}}^{+\infty }\mathrm{d}\omega \frac{g\left(
4572: \omega ,\widetilde{\mu }\right) }{e^{\beta \omega }-1}=\int_{-\omega
4573: _{c}}^{-0^{+}}\mathrm{d}\omega \frac{g\left( \omega ,\widetilde{\mu }\right)
4574: }{e^{\beta \omega }-1}+\int_{0^{+}}^{+\infty }\mathrm{d}\omega \frac{g\left(
4575: \omega ,\widetilde{\mu }\right) }{e^{\beta \omega }-1}.
4576: \end{equation}%
4577: Observe that
4578: \begin{equation}
4579: \frac{1}{e^{\beta \omega }-1}+\frac{1}{e^{-\beta \omega }-1}=-1,
4580: \end{equation}%
4581: we have%
4582: \begin{equation}
4583: \mathcal{P}\int_{-\infty }^{+\infty }\mathrm{d}\omega \frac{g\left( \omega ,%
4584: \widetilde{\mu }\right) }{e^{\beta \omega }-1}=-\int_{-\omega _{c}}^{0}%
4585: \mathrm{d}\omega \,g\left( \omega ,\widetilde{\mu }\right) -\int_{-\omega
4586: _{c}}^{-0^{+}}\mathrm{d}\omega \frac{g\left( \omega ,\widetilde{\mu }\right)
4587: }{e^{-\beta \omega }-1}+\int_{0^{+}}^{+\infty }\mathrm{d}\omega \frac{%
4588: g\left( \omega ,\widetilde{\mu }\right) }{e^{\beta \omega }-1}.
4589: \end{equation}%
4590: Integrated by substitution, it turns into%
4591: \begin{align}
4592: \mathcal{P}\int_{-\infty }^{+\infty }\mathrm{d}\omega \frac{g\left( \omega ,%
4593: \widetilde{\mu }\right) }{e^{\beta \omega }-1}& =-\int_{0}^{\omega _{c}}%
4594: \mathrm{d}\omega \,g\left( \omega -\omega _{c},\widetilde{\mu }\right)
4595: +k_{B}T\int_{0}^{+\infty }\mathrm{d}x\frac{g\left( \omega _{c}+k_{B}Tx,%
4596: \widetilde{\mu }\right) }{e^{x+\beta \omega _{c}}-1} \notag \\
4597: & +k_{B}T\int_{0}^{\beta \omega _{c}}\mathrm{d}x\frac{g\left( k_{B}Tx,%
4598: \widetilde{\mu }\right) -g\left( -k_{B}Tx,\widetilde{\mu }\right) }{e^{x}-1}.
4599: \end{align}%
4600: The second integral on the right-hand side vanishes exponentially as $%
4601: T\rightarrow 0$, it can thus be neglected. The upper limit of the third
4602: integral can be set equal to $+\infty $ at low temperatures for the
4603: integrand decreases as $e^{-x}$ when $x$ becomes very large,
4604: \begin{equation}
4605: \mathcal{P}\int_{-\infty }^{+\infty }d\omega \frac{g\left( \omega ,%
4606: \widetilde{\mu }\right) }{e^{\beta \omega }-1}=-\int_{0}^{\omega
4607: _{c}}d\omega \,g\left( \omega -\omega _{c},\widetilde{\mu }\right)
4608: +k_{B}T\int_{0}^{+\infty }dx\frac{g\left( k_{B}Tx,\widetilde{\mu }\right)
4609: -g\left( -k_{B}Tx,\widetilde{\mu }\right) }{e^{x}-1}. \label{POmega}
4610: \end{equation}%
4611: The second integral on the right-hand side can expanded as%
4612: \begin{equation}
4613: \int_{0}^{+\infty }\mathrm{d}x\frac{g\left( k_{B}Tx,\widetilde{\mu }\right)
4614: -g\left( -k_{B}Tx,\widetilde{\mu }\right) }{e^{x}-1}=2\sum_{n=1}^{+\infty
4615: }\zeta (2n)g^{(2n-1)}\left( 0,\widetilde{\mu }\right) \left( k_{B}T\right)
4616: ^{2n-1}.
4617: \end{equation}%
4618: Substituting it into Eq. (\ref{POmega}) results in%
4619: \begin{equation}
4620: \mathcal{P}\int_{-\infty }^{+\infty }\mathrm{d}\omega \frac{g\left( \omega ,%
4621: \widetilde{\mu }\right) }{e^{\beta \omega }-1}=-\int_{0}^{\omega _{c}}%
4622: \mathrm{d}\omega \,g\left( \omega -\omega _{c},\widetilde{\mu }\right)
4623: +2\sum_{n=1}^{+\infty }\zeta (2n)g^{(2n-1)}\left( 0,\widetilde{\mu }\right)
4624: \left( k_{B}T\right) ^{2n},
4625: \end{equation}%
4626: the sum is an even power series of $T$. To the fourth order, it becomes
4627: \begin{equation}
4628: \mathcal{P}\int_{-\infty }^{+\infty }\mathrm{d}\omega \frac{g\left( \omega ,%
4629: \widetilde{\mu }\right) }{e^{\beta \omega }-1}=-\int_{0}^{\omega _{c}}%
4630: \mathrm{d}\omega \,g\left( \omega -\omega _{c},\widetilde{\mu }\right)
4631: +2\zeta (2)g^{\prime }\left( 0,\widetilde{\mu }\right) \left( k_{B}T\right)
4632: ^{2}+2\zeta (4)g^{(3)}\left( 0,\widetilde{\mu }\right) \left( k_{B}T\right)
4633: ^{4},
4634: \end{equation}%
4635: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
4636: %BeginExpansion
4637: \end{widetext}
4638: %EndExpansion
4639: it is sufficient for the discussion of the low-temperature properties of the
4640: system. This result implies that the renormalized chemical potential $%
4641: \widetilde{\mu }$ and the internal energy $E$ are even functions of $T$ at
4642: low temperatures, which leads to the conclusion of Eq. (\ref{CvTV}): the
4643: specific heat $C_{V}$ of a weakly interacting Bose gas will vanish linearly
4644: as $T\rightarrow 0$.
4645:
4646: The above result reminds us of the specific heat of the electron gas in a
4647: normal metal, which also vanishes linearly as $T\rightarrow 0$. This linear
4648: behavior of the electron gas can be explained analogously. There, what is
4649: concerned is the integral,%
4650: \begin{equation}
4651: \int_{-\infty }^{+\infty }\mathrm{d}\omega \frac{g\left( \omega ,\mu \right)
4652: }{e^{\beta \omega }+1},
4653: \end{equation}%
4654: where $\mu $ is the chemical potential. Along the same way as for the Bose
4655: gas, it can be expressed as
4656: %TCIMACRO{\TeXButton{Bwidetext}{\begin{widetext}}}%
4657: %BeginExpansion
4658: \begin{widetext}%
4659: %EndExpansion
4660: \begin{equation}
4661: \int_{-\infty }^{+\infty }\mathrm{d}\omega \frac{g\left( \omega ,\mu \right)
4662: }{e^{\beta \omega }+1}=\int_{0}^{\omega _{c}}\mathrm{d}\omega \,g\left(
4663: \omega -\omega _{c},\mu \right) +2\sum_{n=1}^{+\infty }\left( 1-\frac{1}{%
4664: 2^{2n-1}}\right) \zeta (2n)g^{(2n-1)}\left( 0,\mu \right) \left(
4665: k_{B}T\right) ^{2n}.
4666: \end{equation}%
4667: To the second order, it is
4668: \begin{equation}
4669: \int_{-\infty }^{+\infty }\mathrm{d}\omega \frac{g\left( \omega ,\mu \right)
4670: }{e^{\beta \omega }+1}=\int_{0}^{\omega _{c}}\mathrm{d}\omega \,g\left(
4671: \omega -\omega _{c},\mu \right) +\zeta (2)g^{\prime }\left( 0,\mu \right)
4672: \left( k_{B}T\right) ^{2}, \label{LowFermi}
4673: \end{equation}%
4674: %TCIMACRO{\TeXButton{Ewidetext}{\end{widetext}} }%
4675: %BeginExpansion
4676: \end{widetext}
4677: %EndExpansion
4678: which is sufficient for the low-temperature properties of the electron gas
4679: in a normal metal because its Fermi temperature is very high, in comparison
4680: to the critical temperature of the BEC happening in a weakly interacting
4681: Bose gas. Eq. (\ref{LowFermi}) also implies that the chemical potential $\mu
4682: $ and the internal energy $E$ are even functions of $T$ at low temperatures,
4683: and that the specific heat at constant volume $C_{V}$ will vanish linearly
4684: as $T\rightarrow 0$. For example, one can easily verify, with the help of
4685: Eq. (\ref{LowFermi}), that the $\mu $, $E$ and $C_{V}$ for the ideal Fermi
4686: gas have the low-temperature forms,%
4687: \begin{eqnarray}
4688: &&\mu (T)=\mu (0)\left[ 1-\frac{\pi ^{2}}{12}\left( \frac{k_{B}T}{\mu (0)}%
4689: \right) ^{2}\right] , \\
4690: &&\frac{E}{N}=\frac{3}{5}\mu (0)\left[ 1+\frac{5\pi ^{2}}{12}\left( \frac{%
4691: k_{B}T}{\mu (0)}\right) ^{2}\right] , \\
4692: &&\frac{C_{V}}{Nk_{B}}=\frac{\pi ^{2}}{2}\frac{k_{B}T}{\mu (0)},
4693: \end{eqnarray}%
4694: which are familiar results. Those discussions show that there exists
4695: similarity even between so different systems!
4696:
4697: \begin{thebibliography}{99}
4698: \bibitem{Gibbs} J. W. Gibbs, \emph{Elementary Principles in Statistical
4699: Mechanics} (Yale University Press, New Haven, 1957).
4700:
4701: \bibitem{Yang1} C. N. Yang, \textquotedblleft Introductory Note on Phase
4702: Transitions and Critical Phenomena\textquotedblright , in \emph{Phase
4703: transitions and Critical Phenomena} (ed. by C. Domb and M. S. Green,
4704: Academic Press, London, 1972), Vol. 1, pp. 1--5.
4705:
4706: \bibitem{BCS1} J. N. Cooper, Phys. Rev. \textbf{104}, 1189 (1956).
4707:
4708: \bibitem{BCS2} J. Bardeen, J. N. Cooper, and J. R. Schrieffer, Phys. Rev.
4709: \textbf{108}, 1175 (1957).
4710:
4711: \bibitem{BCS3} J. R. Schrieffer, \emph{Theory of Superconductivity} (W. A.
4712: Benjamin, New York, 1964).
4713:
4714: \bibitem{Ising} E. Ising, Z. Phys. \textbf{31}, 253 (1925).
4715:
4716: \bibitem{Onsager} L. Onsager, Phys. Rev. \textbf{65}, 117 (1944).
4717:
4718: \bibitem{Kaufmann} B. Kaufmann, Phys. Rev. \textbf{76}, 1232 (1949).
4719:
4720: \bibitem{Yang2} C. N. Yang, and T. D. Lee, Phys. Rev. \textbf{87}, 404
4721: (1952).
4722:
4723: \bibitem{Yang3} C. N. Yang, and T. D. Lee, Phys. Rev. \textbf{87}, 410
4724: (1952).
4725:
4726: \bibitem{Georgii} H.-O. Georgii, \emph{Gibbs Measures and Phase Transitions}
4727: (Walter de Gruyter, Berlin, 1988).
4728:
4729: \bibitem{GEmch} G. G. Emch, \textquotedblleft The $C^{\star }$-Algebraic
4730: Approach to Phase transitions\textquotedblright , in \emph{Phase transitions
4731: and Critical Phenomena} (ed. by C. Domb and M. S. Green, Academic Press,
4732: London, 1972), Vol. 1, pp. 137-175.
4733:
4734: \bibitem{Emch} G. G. Emch, \emph{Algebraic Methods in Statistical Mechanics
4735: and Quantum Field Theory} (John Wiley \& Sons, New York, 1972).
4736:
4737: \bibitem{Bratteli} O. Bratteli and D. W. Robinson, \emph{Operator Algebras
4738: and Quantum Statistical Mechanics} (Springer-Verlag, Berlin, 1996) Vols. 1
4739: and 2.
4740:
4741: \bibitem{Kramers} H. A. Kramers, and G. H. Wannier, Phys. Rev. \textbf{60},
4742: 252 (1941).
4743:
4744: \bibitem{Yang4} C. N. Yang, Phys. Rev. \textbf{85}, 809 (1952).
4745:
4746: \bibitem{Huang} K. Huang, \emph{Statistical Mechanics} (John Wiley \& Sons,
4747: New York, 1987), pp. 300--302, and Chaps. 14 and 15.
4748:
4749: \bibitem{Tsallis1} C. Tsallis, J. Stat. Phys. \textbf{52}, 479 (1988).
4750:
4751: \bibitem{Tsallis2} E. M. F. Curado, and C. Tsallis, J. Phys. A \textbf{24},
4752: L69 (1991).
4753:
4754: \bibitem{Tsallis3} C. Tsallis, R. S. Mendes, and A. P. Plastino, Physica A
4755: \textbf{261}, 534 (1998).
4756:
4757: \bibitem{Landau} L. D. Landau and E. M. Lifshitz, \emph{Statistical Physics}
4758: (Pergamon Press, London, 1958), chap. XIV.
4759:
4760: \bibitem{Born} M. Born, and K. Fucks, Proc. Roy. Soc. \textbf{A166}, 391
4761: (1938).
4762:
4763: \bibitem{Mayer} J. E. Mayer, J. Chem. Phys. \textbf{5}, 67 (1937).
4764:
4765: \bibitem{Kadanoff1} L. P. Kadanoff, Physics \textbf{2}, 263 (1966).
4766:
4767: \bibitem{Kadanoff2} L. P. Kadanoff, W. G\"{o}tze, D. Hamblen, R. Hecht, E.
4768: A. S. Lewis, V. V. Palciauskas, M. Rayl, and J. Swift, Rev. Mod. Phys.
4769: \textbf{39}, 395 (1967).
4770:
4771: \bibitem{Kadanoff3} L. P. Kadanoff, \textquotedblleft Scaling, Universality
4772: and Operator Algebras\textquotedblright , in \emph{Phase transitions and
4773: Critical Phenomena} (ed. by C. Domb and M. S. Green, Academic Press, London,
4774: 1976), Vol. 5a, pp. 1--34.
4775:
4776: \bibitem{Wilson1} K. G. Wilson, Phys. Rev. \textbf{4}, 3174 (1971).
4777:
4778: \bibitem{Wilson2} K. G. Wilson, Phys. Rev. \textbf{4}, 3184 (1971).
4779:
4780: \bibitem{Wilson3} K. G. Wilson, and J. Kogut, Phys. Rep. \textbf{12}, 75
4781: (1974).
4782:
4783: \bibitem{Wilson4} K. G. Wilson, Rev. Mod. Phys. \textbf{47}, 773 (1975).
4784:
4785: \bibitem{Wilson5} K. G. Wilson, Rev. Mod. Phys. \textbf{55}, 583 (1983).
4786:
4787: \bibitem{Fisher} M. E. Fisher, Rev. Mod. Phys. \textbf{70}, 653 (1998).
4788:
4789: \bibitem{Laud} B. B. Laud, \emph{Lasers and Non-linear Optics} (John Wiley
4790: \& Sons, New York, 1991), Chap. 9.
4791:
4792: \bibitem{Purcell} E. M. Purcell and R. V. Pound, Phys. Rev. \textbf{81}, 279
4793: (1951).
4794:
4795: \bibitem{Ramsey} N. F. Ramsey, Phys. Rev. \textbf{103}, 20 (1956).
4796:
4797: \bibitem{Pathria} R. K. Pathria, \emph{Statistical Mechanics} (Pergamon
4798: Press, Oxford, 1972), Sec. 39.
4799:
4800: \bibitem{Bogoliubov} N. N. Bogoliubov, J. Phys. (USSR) \textbf{11}, 23
4801: (1947).
4802:
4803: \bibitem{Zagrebnov} V. A. Zagrebnov, and J. -B. Bru, Phys. Rep. \textbf{350}%
4804: , 291 (2001).
4805:
4806: \bibitem{Zubarev} D. N. Zubarev, \emph{Nonequilibrium Statistical
4807: Thermodynamics} (Consultants Bureau, New York, 1974), Chap. 3.
4808:
4809: \bibitem{Rickayzen} G. Rickayzen, \emph{Theory of Superconductivity} (Wiley,
4810: New York, 1964).
4811:
4812: \bibitem{Ambegaokar} V. Ambegaokar, \textquotedblleft The Green's Function
4813: Method\textquotedblright , in \emph{Superconductivity} (ed. R. D. Parks,
4814: Marcel Dekker, New York, 1969), Vol. 1, Chap. 5.
4815:
4816: \bibitem{Hewitt} E. Hewitt, and K. Stromberg, \emph{Real and Abstract
4817: Analysis} (Springer-Verlag, Berlin, 1965).
4818:
4819: \bibitem{Kolmogorov} A. Kolmogorov, \emph{Grundbegriffe der
4820: Wahrscheinlichkeitsrechnung}, fasc. 3 of Vol. 2 of Ergebnisse der
4821: Mathematic, Berlin, 1933.
4822:
4823: \bibitem{Feller} W. Feller, \emph{An Introduction to Probability Theory and
4824: Its Applications} (John Wiley \& Sons, New York, Vol. 1, 1957, Vol. 2, 1971).
4825:
4826: \bibitem{Einstein1} A. Einstein, Sitzungsber. Klg. Preuss. Akad. \textbf{1924%
4827: }, 261 (1924).
4828:
4829: \bibitem{Einstein2} A. Einstein, Sitzungsber. Klg. Preuss. Akad. \textbf{1925%
4830: }, 3 (1925).
4831:
4832: \bibitem{Uhlenbeck} G. E. Uhlenbeck, \emph{`Over Statistische Methoden in de
4833: Theorie der Quanta'}(Ph.D. thesis, Nyhoff, the Hague, 1927).
4834:
4835: \bibitem{Pais} A. Pais, \emph{`Subtle is the Lord ...' --- The Science and
4836: the Life of Albert Einstein} (Oxford, 1982), Sec. 23d.
4837:
4838: \bibitem{Greiner} W. Greiner, and J. Reinhardt, \emph{Field quantization}
4839: (Springer-Verlag, Berlin, 1996), Chaps. 5 and 12.
4840:
4841: \bibitem{Yang5} C. N. Yang, Rev. Mod. Phys. \textbf{34}, 694 (1962).
4842:
4843: \bibitem{GL} V. L. Ginzburg, and L. D. Landau, Zh. Eksp. Teor. Fiz. \textbf{%
4844: 20}, 1064 (1950).
4845:
4846: \bibitem{GP1} E. P. Gross, Nuovo Cimento \textbf{20}, 454 (1961).
4847:
4848: \bibitem{GP2} L. P. Pitaevskii, Sov. Phys. JETP \textbf{13}, 451 (1961).
4849:
4850: \bibitem{Fetter} A. L. Fetter, and J. D. Walecka, \emph{Quantum Theory of
4851: Many-Particle Systems} (McGraw-Hill, New York, 1971), Chaps. 7--9.
4852:
4853: \bibitem{Keesom} W. H. Keesom, and K. Clausius, Proc, Roy. Acad. (Amsterdam)
4854: \textbf{35}, 307 (1932).
4855:
4856: \bibitem{Lifshitz} E. M. Lifshitz and L. P. Pitaevskii, \emph{Statistical
4857: Physics} (Beijing World Publishing Corporation, Beijing, 1999), Sec. 2 and
4858: Sec. 45.
4859:
4860: \bibitem{Landau1} L. D. Landau, J. Phys. (USSR) \textbf{5}, 71 (1941).
4861:
4862: \bibitem{Landau2} L. D. Landau, J. Phys. (USSR) \textbf{11}, 91 (1947).
4863:
4864: \bibitem{Berezin} F. A. Berezin, \emph{The Method of Second Quantization}
4865: (Academic Press, New York, 1965).
4866:
4867: \bibitem{Blinc} R. Blinc, and B. \v{Z}ek\v{s}, \emph{Soft Modes in
4868: Ferroelectrics and Antiferroelectrics} (North-Holland, Amsterdam, 1974),
4869: Chaps. 1--4.
4870:
4871: \bibitem{Justin} J. Zinn-Justin, \emph{Quantum Field Theory and Critical
4872: Phenomena} (Clarendon Press, Oxford, 1996), Chap. 13.
4873:
4874: \bibitem{Tsvelik} A. M. Tsvelik, \emph{Quantum Field Theory in Condensed
4875: Matter Physics} (World Scientific, Beijing, 2001), Part one.
4876:
4877: \bibitem{Muller} W. Greiner, and B. M\"{u}ller, \emph{Gauge Theory of Weak
4878: Interactions} (Springer-Verlag, Berlin, 2000), Chap. 4.
4879:
4880: \bibitem{Goldstone} J. Goldstone, Nuovo Cimento \textbf{19}, 154 (1961).
4881:
4882: \bibitem{Faddeev} L. D. Faddeev, and V. N. Popov, Phys. Lett. \textbf{25B},
4883: 29 (1967).
4884:
4885: \bibitem{Jackson} J. D. Jackson, \emph{Classical Electrodynamics} (John
4886: Wiley \& Sons, New York, 1999), pp. 600--605.
4887:
4888: \bibitem{Meissner} W. Meissner, and R. Ochsenfeld, Naturwiss. \textbf{21},
4889: 787 (1933).
4890:
4891: \bibitem{London} F. London, and H. London, Proc. Roy. Soc. (London) \textbf{%
4892: A147}, 71 (1935).
4893:
4894: \bibitem{Higgs1} P. W. Higgs, Phys. Rev. Lett. \textbf{12}, 132 (1964).
4895:
4896: \bibitem{Higgs2} P. W. Higgs, Phys. Rev. Lett. \textbf{13}, 508 (1964).
4897:
4898: \bibitem{Higgs3} P. W. Higgs, Phys. Rev. \textbf{145}, 1156 (1966).
4899:
4900: \bibitem{Anderson1} P. W. Anderson, Phys. Rev. \textbf{112}, 1900 (1958).
4901:
4902: \bibitem{Anderson2} P. W. Anderson, Phys. Rev. \textbf{130}, 439 (1963).
4903:
4904: \bibitem{Kibble} T. W. Kibble, Phys. Rev. \textbf{155}, 1554 (1967).
4905:
4906: \bibitem{Huang2} K. Huang, \emph{Quarks, Leptons }\&\emph{\ Gauge Fields}
4907: (World Scientific, Singapore, 1992), Sec. 3.3.
4908:
4909: \bibitem{Negele} J. W. Negele, and H. Orland, \emph{Quantum Many-Particale
4910: Systems} (Addison-Wesley, New York, 1988).
4911: \end{thebibliography}
4912:
4913: \end{document}
4914: