cond-mat0503400/sc.tex
1: %\documentclass[aps,preprint]{revtex4}
2: \documentclass[aps,twocolumn,article,showpacs,showkeys]{revtex4}
3: 
4: \usepackage{amsmath,amssymb,mathrsfs}
5: \usepackage{latexsym}
6: \usepackage{graphicx,psfrag} 
7: 
8: % boldsymbol (requires amsmath)
9: \newcommand{\bs}[1]{\boldsymbol{#1}}
10: % commutator and anticommutator
11: \newcommand{\bigc}[2]{\bigl[#1,#2\bigr]}
12: \newcommand{\Bigc}[2]{\Bigl[#1,#2\Bigr]}
13: \newcommand{\biggc}[2]{\biggl[#1,#2\biggr]}
14: \newcommand{\Biggc}[2]{\Biggl[#1,#2\Biggr]}
15: \newcommand{\comm}[2]{\left[#1,#2\right]}
16: \newcommand{\anticomm}[2]{\left\{#1,#2\right\}}
17: % of simply takes its argument and places it in () sized brackets 
18: \newcommand{\of}[1]{\left(#1\right)}
19: % sum on nearest neighbor bonds
20: \newcommand{\bond}{\left\langle i, j \right\rangle}
21: %\newcommand{\bondsum}{\sum_{\left\langle i, j \right\rangle}}
22: \newcommand{\nbond}{\left\langle\left\langle i, j \right\rangle\right\rangle}
23: %\newcommand{\nbondsum}{\sum_{\left\langle\left\langle i, j 
24: %        \right\rangle\right\rangle}}
25: % defines the function absolute value of
26: \newcommand{\abs}[1]{\big|#1\bigr|}
27: % 1/2
28: \newcommand{\half}{$\frac{1}{2}$ }
29: % notation for vacuum, an empty set inside a ket
30: \newcommand{\vac}{\left|\,0\,\right\rangle}
31: % define the bra and ket, and braket
32: \newcommand{\ket}[1]{\left|#1\right\rangle}
33: \newcommand{\bra}[1]{\left\langle#1\right|}
34: \newcommand{\braket}[1]{\bigl\langle#1\bigr\rangle}
35: % simplifies using the up and down arrows to denote spin
36: \newcommand{\up}{\uparrow}
37: \newcommand{\dw}{\downarrow}
38: % standard begin-end abreviations
39: \def\bd{\begin{displaymath}}
40: \def\ed{\end{displaymath}}
41: \def\be{\begin{equation}}
42: \def\ee{\end{equation}}
43: \def\bea{\begin{eqnarray}}
44: \def\eea{\end{eqnarray}}
45: \def\bi{\begin{itemize}}
46: \def\ei{\end{itemize}}
47: \def\bn{\begin{enumerate}}
48: \def\en{\end{enumerate}}
49: 
50: \def\ie{{\it i.e.},\ }
51: \def\eg{{\it e.g.}\ }
52: \def\ea{{\it et.al.}}
53: 
54: \begin{document}
55: \title{Is electromagnetic gauge invariance spontaneously violated 
56: in superconductors?}
57: 
58: \author{Martin Greiter} 
59: \email{greiter@tkm.uni-karlsruhe.de}
60: \affiliation{Institut f\"ur Theorie der Kond.\ Materie,
61:   Universit\"at Karlsruhe, Postfach 6980, D-76128 Karlsruhe}
62: 
63: \begin{abstract}
64:   We aim to give a pedagogical introduction to those elementary
65:   aspects of superconductivity which are not treated in the classic
66:   textbooks.  In particular, we emphasize that global U(1) phase
67:   rotation symmetry, and not gauge symmetry, is spontaneously
68:   violated, and show that the BCS wave function is, contrary to claims
69:   in the literature, fully gauge invariant.  We discuss the nature of
70:   the order parameter, the physical origin of the many degenerate
71:   states, and the relation between formulations of superconductivity
72:   with fixed particle numbers {\it vs.}\ well defined phases.  We
73:   motivate and to some extend derive the effective field theory at low
74:   temperatures, explore symmetries and conservation laws, and justify the
75:   classical nature of the theory.  Most importantly, we show that the
76:   entire phenomenology of superconductivity essentially follows from
77:   the single assumption of a charged order parameter field.  This
78:   phenomenology includes Anderson's characteristic equations of
79:   superfluidity, electric and magnetic screening, the Bernoulli Hall
80:   effect, the balance of the Lorentz force, as well as the quantum
81:   effects, in which Planck's constant manifests itself through the
82:   compactness of the U(1) phase field.  The latter effects include
83:   flux quantization, phase slippage, and the Josephson effect.
84: % \\  PACS numbers: 74.20.-z, 11.15.-q, 11.15.Ex.%~\cite{pacs}
85: \end{abstract}
86: 
87: 
88: 
89: \pacs{74.20.-z, 11.15.-q, 11.15.Ex.}
90: 
91: \keywords{superconductivity; gauge invariance; order parameter;
92:   effective field theory; Higgs mechanism}
93: 
94: \maketitle
95: %\tableofcontents
96: 
97: \section*{CONTENTS}
98: \noindent{\small
99: \begin{tabular}{r@{\hspace{2mm}}p{75mm}r}
100: I. &Introduction &1\\[3pt]
101: II. &Gauge invariance &2\\[3pt]
102: III. &Order parameter considerations &4\\[3pt]
103: IV. &Effective field theory &6\\[3pt]
104: V. &Phenomenology and the Higgs mechanism &11\\[3pt]
105: VI. &Quantum effects &15\\[3pt]
106: \multicolumn{2}{l}{Appendix} &17\\[3pt]
107: \multicolumn{2}{l}{Acknowledgements} &17
108: \end{tabular}}
109: 
110: 
111: \section{INTRODUCTION}
112: %\section{Introduction}
113: \label{sec:int}
114: %
115: Many years ago, Steven Weinberg mentioned to me that he was
116: disconcerted that none of the classic textbooks on
117: superconductivity would explain the phenomenon in terms of the
118: Higgs mechanism~\cite{sid} for the electromagnetic gauge field.  This
119: concern is of course very well justified, and it was most likely with
120: this concern in mind that Weinberg has included a section on
121: superconductivity in his treatment of spontaneous symmetry
122: breaking and the Higgs mechanism in his %by now likewise classic
123: series of volumes entitled {\it The quantum theory of
124:   fields}\hspace{2pt}~\cite{wein}.  When I was asked recently to
125: present some lectures on superconductivity, I opened his book
126: expecting to find a particularly lucid exposition of this in condensed
127: matter physics rarely emphasized perspective.  I found the exposition
128: I was looking for, but to my surprise, build around the
129: following statement: {\it A superconductor is simply a material in
130:   which electromagnetic gauge invariance is spontaneously broken.}
131: What Weinberg means with this statement is just that the
132: electromagnetic gauge field ``acquires a mass'' due to the Higgs
133: mechanism in a superconductor, as particle physicists often speak of
134: {spontaneously broken gauge invariance} interchangeably with
135: {the Higgs mechanism}.
136: 
137: Nonetheless, I am not perfectly at ease with the above statement,
138: which is, by the way, by no means specific to Weinberg's exposition,
139: but widely believed and accepted.  While it is obvious that Weinberg
140: fully understands the matter, the statement may still be misleading to
141: a young student who is learning the subject for the first time.  The
142: problem is that the statement is, if one takes it literally, not
143: correct: gauge invariance cannot spontaneously break down as a matter
144: of principle, and in particular is not broken in a superconductor, as
145: I will explain in the following section.
146: 
147: This paper is organized as follows.  In Section~\ref{sec:gau}, we
148: discuss the statement quoted above including the danger which may result
149: from a literal interpretation of it in depth.  In particular, we show
150: that the BCS ground state is, in contrast to statements made in the
151: literature, fully gauge invariant.  The crucial ingredient often
152: omitted is that gauge transformations involve, in addition to the
153: standard transformation of gauge fields, local phase rotations of {\it
154:   both} creation (and annihilation) operators {\it and} wave
155: functions.
156: %
157: In Section~\ref{sec:ord}, we discuss the nature of the order parameter in
158: superconductors, with a particular emphasis on finite systems, which
159: always possess a unique ground state.  The arising subtlety is
160: explained by drawing an analogy to quantum antiferromagnets, which
161: also possess a unique and rotationally invariant ground state for
162: finite systems.
163: %
164: In Section~\ref{sec:eff}, we motivate and elaborate the effective
165: field theory of a superconductor at low temperatures, which contains
166: the theory of a neutral superfluid as the special case where the
167: charge is set to zero.  In particular, we obtain the particle density
168: and current as well as the energy and momentum density from the
169: physical symmetries of the theory, invariance under global U(1) phase
170: rotations of the order parameter and invariance under translations in
171: time and space.  The quest for a consistent definition of the
172: superfluid velocity yields a relation between current and momentum
173: densities in the superfluid, which in turn requires corrections to the
174: effective Lagrangian.  Since the density of the superfluid is
175: essentially the ``momentum conjugate'' of the order parameter phase,
176: Hamilton's equations yield physical information not contained in the
177: Euler--Lagrange equations; specifically, we obtain a gauge invariant
178: generalization of Anderson's characteristic equations of superfluidity
179: to the case of superconductors.  We conclude this Section with a brief
180: justification of the classical nature of the effective field theory.
181: %
182: In Section~\ref{sec:phe}, we discuss the phenomenology of
183: superconductors as compared to neutral superfluids, or, in general
184: terms, the Higgs mechanism.  To begin with, we briefly address the
185: phenomenology of neutral superfluids including vortex quantization,
186: and
187: %review the formalism of 
188: give a general introduction to the Higgs mechanism in field theories.
189: We then turn to the phenomenology of simply connected superconductors,
190: solve the equations of motion, obtain electric and magnetic screening,
191: London's equation, the Bernoulli Hall effect, and the balance of the
192: Lorentz force.  We demonstrate that the Higgs mechanism never
193: corresponds to a spontaneous violation of a gauge symmetry, and that
194: it is incorrect to interpret it in terms of ``a mass acquired by the
195: electromagnetic gauge field'', as the massive field is no longer a
196: gauge field.  Specifically, we show that the massive vector field,
197: which may alternatively be used to describe a (simply connected)
198: superconductor, is correctly interpreted as a four-vector formed by
199: the chemical potential and the three components of the superfluid
200: velocity.  We conclude this Section with a discussion of the subtle
201: difference between the physical invariance of the theory under global
202: U(1) rotations of the order parameter phase and gauge invariance,
203: which is nothing but a local invariance of our description of the
204: system.
205: %
206: In the last Section, we review a family of ``quantum effects'': the
207: quantization of magnetic flux in superconductors, phase slippage, and
208: the Josephson effect in both neutral superfluids and superconductors.
209: In these effects, Planck's constant manifests itself in the
210: phenomenology through the compactness of the order parameter phase
211: field; these effects require either a non-trivial topology or more
212: than one superfluid.  We derive them from the effective field theory
213: introduced in section~\ref{sec:eff}, and thereby demonstrate that the
214: very few assumptions made in motivating the effective theory are
215: sufficient to account for them.
216: 
217: \section{GAUGE INVARIANCE}
218: %\section{Gauge invariance}
219: \label{sec:gau}
220: 
221: To illustrate how dangerous the statement quoted in the introduction
222: is in the case of superconductivity, where we do not only have a
223: description in terms of an effective field theory but also a
224: microscopic description in terms of model Hamiltonians and trial wave
225: functions, I will at first assume the statement was true and take it
226: literally.  I will pretend to be a student who has just learned that
227: electromagnetic gauge invariance is spontaneously violated in a
228: superconductor.  Well, what does this mean?  A spontaneously broken
229: symmetry means that the Hamiltonian of a given system in the
230: thermodynamic limit is invariant under a given symmetry transformation
231: (\ie commutes with the generator(s) of this symmetry) while the ground
232: state is not invariant.  There are many ground states, which transform
233: into each other under the symmetry transformations.  A classic example
234: is ferromagnetism: The Hamiltonian is rotationally invariant, while
235: any particular ground state, specified by the direction the
236: magnetization vector points to, is not.  So if gauge invariance is
237: broken in a superconductor, this must mean that the ground state of
238: the superconductor does not share the gauge invariance of the
239: Hamiltonian.  Indeed, a glance at the BCS wave
240: function~\cite{bcs,schrief,degenn,tinki}
241: %{BardeenCooperSchrieffer57pr1,BardeenCooperSchrieffer57pr2,Schrieffer1964,deGennes1966,Tinkham1996}
242: \begin{equation}
243:   \label{e:bcs}
244:   \ket{\psi%^{\scriptstyle\rm BCS}
245:     _\phi}=\prod_{\bs{k}}\left(u_{\bs{k}} + v_{\bs{k}} e^{i\phi} 
246:     c_{\bs{k}\up}^\dagger\,c_{-\bs{k}\dw}^\dagger\right)\vac,
247: \end{equation} 
248: where the coefficients $u_{\bs{k}}$ and $v_{\bs{k}}$ are chosen real
249: and $\phi$ is an arbitrary phase, appears to confirm this picture.
250: There are many different ground state wave functions, labeled by
251: $\phi$, which transform into each other under an electromagnetic gauge
252: transformation given by
253: \begin{equation}
254:   \label{e:gaugec}
255:   c_{\bs{k}\sigma}^\dagger\rightarrow 
256:   e^{i{\textstyle\frac{e}{\hbar c}}\Lambda} c_{\bs{k}\sigma}^\dagger,
257: \end{equation} 
258: which is tantamount to taking 
259: \begin{displaymath}
260:   \phi\rightarrow \phi+\frac{2e}{\hbar c}\Lambda .
261: \end{displaymath}
262: For simplicity, we have chosen $\Lambda$ independent of spacetime. 
263: The electron charge throughout this article is $-e$.
264: 
265: Next, I the student ask myself whether these many BCS wave functions
266: for different parameters $\phi$ correspond to physically different
267: states.  I know that gauge transformations are not physical
268: transformations: gauge invariance is an invariance of a description of
269: a system, while other symmetries correspond to invariances under
270: physical transformations, like rotations or translations, which affect
271: the physical state in question.  For example, if the Hamiltonian for
272: given system (like a ferromagnet) is invariant under rotations in
273: space, this implies that if we rotate a given eigenstate, we will
274: obtain another eigenstate.  Depending on whether the original state is
275: rotationally invariant or not, it will transform into itself or into a
276: physically different state.  A gauge transformation, by contrast, will
277: only transform our description of a system from one gauge to another,
278: without ever having any effect on the physical state of the system.
279: Gauge transformations are comparable to rotations or translations of
280: the coordinate system we use to describe a system.  Another way of
281: seeing the difference is by noting that it is possible to rotate or
282: translate a superconductor in the laboratory, but as a matter of
283: principle not possible to gauge transform it.  Returning to the
284: superconductor, I the student conclude that if the many different
285: ground states only differ by a gauge transformation, they cannot be
286: physically different.  The ground state of a superconductor must hence
287: be physically unique.
288: 
289: In fact, there is another way of looking at the problem which appears
290: to confirm this conclusion.  A BCS superconductor can not only be
291: described in the grand-canonical ensemble, where the chemical
292: potential rather than the number of particles is fixed, but also in
293: the canonical ensemble, where the number of particles or pairs is
294: fixed.  Following Anderson~\cite{phil66}, we can project out a (not
295: normalized) state with $N$ pairs from (\ref{e:bcs}) via
296: \begin{equation}
297:   \label{e:project}
298:   \ket{\psi_N} = \int_0^{2\pi}\!\!d\phi\, e^{-iN\phi}\ket{\psi_\phi}
299: \end{equation} 
300: and obtain (see appendix)
301: \begin{eqnarray}
302:   \label{e:bcsn}
303:   \ket{\psi_N}\!\!\! &=&\!\!\!\int\!\! 
304:   d^{3\,}\!\bs{x}_1\!\ldots d^{3\,}\!\bs{x}_{2\!N}\,
305:   \varphi(\bs{x}_1-\bs{x}_2)\cdot\!\ldots\!\cdot
306:   \varphi(\bs{x}_{2\!N-1}\!-\bs{x}_{2\!N})
307:   \,\cdot\cr
308:   && \cdot\,\psi^\dagger_\up(\bs{x}_1)\psi^\dagger_\dw(\bs{x}_2)\ldots
309:   \psi^\dagger_\up(\bs{x}_{2N-1})\psi^\dagger_\dw(\bs{x}_{2N})\vac \!,
310: \end{eqnarray}
311: where the real-space creation operator fields $\psi^\dagger_\sigma(\bs{x})$
312: are simply the Fourier transforms of the momentum-space creation operators
313: $c_{\bs{k}\sigma}^\dagger$,
314: \begin{equation}
315:   \label{e:ft}
316:   \psi^\dagger_\sigma(\bs{x})=\frac{1}{\sqrt{V}}\sum_{\bs{k}} 
317:   e^{-i\bs{k}\bs{x}} c^\dagger_{\bs{k}\sigma }, \
318:   c^\dagger_{\bs{k} \sigma}=\frac{1}{\sqrt{V}}\!\int\!\! d^{3\,}\!\bs{x}\, 
319:   e^{i\bs{k}\bs{x}} \psi^\dagger_\sigma(\bs{x}).
320: \end{equation}
321: % \end{displaymath}
322: The wave function for each of the individual pairs, which only depends
323: on the relative coordinate, is (up to a normalization) given by
324: \begin{equation}
325:   \label{e:varphi}
326:   \varphi(\bs{x})=\frac{1}{V}\sum_{\bs{k}} \frac{v_{\bs{k}}}{u_{\bs{k}}}
327:   e^{i\bs{k}\bs{x}}.  
328: \end{equation}
329: This form nicely illustrates that all the
330: pairs have condensed into the same state, which is the essence of
331: superfluidity.  As $\varphi(\bs{x})$ is uniquely determined for a given
332: Hamiltonian, the ground state of the superconductor (\ref{e:bcsn})
333: once more appears to be unique and non-degenerate.
334: 
335: So far the students train of thought.  The conclusion reached is of
336: course completely wrong:  a superfluid, and in particular a superconductor,
337: is characterized by a spontaneously broken symmetry, and, at least in the
338: thermodynamic limit, there are many degenerate ground states.  There 
339: are several mistakes in the students analysis.  The first
340: is his literal interpretation of the statement quoted in the
341: inroduction.  In fact, {\it a gauge symmetry cannot spontaneously
342:   break down as a matter of principle, since it is not a physical
343:   symmetry of the system to begin with, but merely an invariance of
344:   description}~\cite{elit}. The only way to violate a gauge symmetry is
345: by choosing a gauge, which again has only an effect on our
346: description, but not on the physical system itself.
347: 
348: In particular, the BCS ground state does not violate gauge invariance,
349: even though statements to the contrary have been made in the literature.
350: The apparent contradiction with (\ref{e:bcs}) and 
351: (\ref{e:gaugec}) can be resolved by recalling that a gauge transformation
352: only affects our description of the system, and is analogous to a rotation
353: of the coordinate system we use in the example of a ferromagnet:  if
354: we rotate the coordinate system accordingly, a ground state with 
355: the magnetization vector pointing in the $z$-direction in the original 
356: coordinate system will ``transform'' into a state with the magnetization
357: vector pointing in the $x$-direction in the new coordinate system,
358: while the physical state has not been affected at all.
359: So while the BCS wave function may look different in a different gauge,
360: the state itself will remain the same.
361: 
362: It is worthwhile to rephrase this statement in equations.
363: To begin with, let us consider a (relativistic quantum) field
364: theory.  Electromagnetic gauge invariance is the invariance of a given
365: theory under U(1) rotations of the complex scalar fields which carry
366: the charge:
367: \begin{equation}
368:   \label{e:gaugeprel}
369:   \psi^\dagger(x)\rightarrow 
370:   e^{i{\textstyle\frac{e}{\hbar c}}\Lambda(x)}\psi^\dagger(x),\quad 
371:   \psi(x)\rightarrow e^{-i{\textstyle\frac{e}{\hbar c}}\Lambda(x)}\psi(x),
372: \end{equation}
373: where $x$ denotes spacetime.  We use the conventions
374: $(x^\mu)=(x^0,x^1,x^2,x^3)=(ct,x,y,z)$,
375: $(x_\mu)=g_{\mu\nu}x^\nu$, $1=g_{00}=-g_{11}=-g_{22}=-g_{33}$.
376: If the theory contains gradient terms in these fields (as it usually does),
377: gauge invariance demands that they are minimally coupled to 
378: a U(1) gauge field, \ie the gradient terms must enter the Lagrangian as 
379: \begin{displaymath}
380:   \Bigl(\partial_\mu + i{\frac{e}{\hbar c}} A_\mu (x)\Bigr)\psi^\dagger(x)
381:   \quad \text{or}\quad
382:   \Bigl(\partial_\mu - i{\frac{e}{\hbar c}} A_\mu (x)\Bigr)\psi(x),
383: \end{displaymath}
384: where $(\partial_\mu)\equiv (\partial/\partial x^\mu) =
385: ({\frac{1}{c}}\partial_t, \nabla)$.  The gauge field
386: $(A_\mu)=(\Phi,-\bs{A})$ must transform according to
387: \begin{equation}
388:   \label{e:gaugearel}
389:   A_\mu(x)\rightarrow A_\mu(x)-\partial_\mu \Lambda(x).  
390: \end{equation}
391: The statement that the theory is gauge invariant simply means
392: that the Lagrangian is invariant under the combined transformation
393: (\ref{e:gaugeprel}) and (\ref{e:gaugearel}).  It is not a physical
394: invariance, but an invariance of description, as it only amounts to a
395: reparametrization of fields. 
396: 
397: The concept of gauge invariance is implemented in a very similar way
398: in non-relativistic quantum mechanics, where the gauge field $\bs{A}$ 
399: is no longer considered a dynamical variable, but an externally applied 
400: vector potential, and we usually do not describe a system by
401: a Lagrange density, but by a Hamiltonian operator and its
402: eigenstates.  For pedagogical reasons, let us first assume a formulation 
403: in second quantization.  Electromagnetic gauge invariance means once
404: again that the description is invariant under U(1) rotations of the
405: particle creation and annihilation operator fields~\cite{simp},  
406: \begin{equation}
407:   \label{e:gaugep}
408:   \psi^\dagger_\sigma(\bs{x})\rightarrow 
409:   e^{i{\textstyle\frac{e}{\hbar c}}\Lambda(\bs{x})}
410:   \psi_\sigma^\dagger(\bs{x}), \quad 
411:   \psi_\sigma(\bs{x})\rightarrow
412:   e^{-i{\textstyle\frac{e}{\hbar c}}\Lambda(\bs{x})}\psi_\sigma(\bs{x}).
413: \end{equation}
414: The kinetic part of the Hamiltonian will again contain gradient terms
415: in the operator fields, which once again must be minimally coupled to
416: the electromagnetic gauge field.  For example, the standard kinetic
417: Hamiltonian for a quadratic dispersion 
418: \begin{equation}
419:   \label{e:ham}
420:   H_{\scriptscriptstyle\textrm{kin}}=
421:   \frac{1}{2m}\sum_\sigma\!\int\!d^{3\,}\!\bs{x}\,
422:   \psi^\dagger_\sigma(\bs{x}) 
423:   \left(-i\hbar\nabla +i\frac{e}{\hbar}\bs{A(\bs{x})}\right)^2
424:   \psi_\sigma(\bs{x})
425: \end{equation}
426: is obviously invariant under (\ref{e:gaugep}) provided we transform
427: the gauge field simultaneously according to
428: \begin{equation}
429:   \label{e:gaugea}
430:   \bs{A(\bs{x})}\rightarrow \bs{A(\bs{x})}+\nabla\Lambda(\bs{x}).
431: \end{equation}
432: Let us now turn to the gauge transformation properties of the
433: eigenstates.  Consider a general $N$ electron eigenstate
434: \begin{eqnarray}
435:   \label{e:eigen}
436:   \ket{\varphi}\!\!\! &=&\!\!\!\sum_{\sigma_1\ldots\sigma_N}\!\int\!
437:   d^{3\,}\!\bs{x}_1\ldots d^{3\,}\!\bs{x}_N\,
438:   \varphi(\bs{x}_1\ldots \bs{x}_N;\sigma_1\ldots\sigma_N)\,\cdot\cr
439:   &&\rule{0pt}{12pt}\hspace{15mm}
440:   \psi^\dagger_{\sigma_1}(\bs{x}_1)\ldots\psi^\dagger_{\sigma_N}(\bs{x}_N)
441:   \vac \!.
442: \end{eqnarray}
443: The state is invariant under (\ref{e:gaugep}) provided we transform
444: the wave function according to
445: \begin{eqnarray}
446:   \label{e:gaugepsi}
447:   &&\hspace{-9mm}
448:   \varphi(\bs{x}_1\ldots \bs{x}_N;\sigma_1\ldots\sigma_N)\,\rightarrow\,\cr
449:   &&\rule{0pt}{22pt}\hspace{1mm}
450:   \prod_{j=1}^N e^{-i{\textstyle\frac{e}{\hbar c}}\Lambda(\bs{x_j})}\,
451:   \varphi(\bs{x}_1\ldots \bs{x}_N;\sigma_1\ldots\sigma_N).
452: \end{eqnarray}
453: This already illustrates the statement phrased in words above: A gauge
454: transformation leaves physical states invariant.  This is just not
455: obvious in every formulation.  If we formulate a problem in
456: non-relativistic quantum mechanics in first quantization, a gauge
457: transformation will only amount to (\ref{e:gaugea}) and
458: (\ref{e:gaugepsi}), as we do not even introduce the operator fields
459: $\psi^\dagger$ and $\psi$.  As $\bs{A}(\bs{x})$ implements an
460: externally applied magnetic field, we must choose a gauge in order to
461: obtain explicit expressions for the Hamiltonian and the eigenstates.
462: The vector potential and the wave functions will have different
463: functional forms in different gauges.  The gauge rotations
464: (\ref{e:gaugep}) of the particle creation and annihilation operators,
465: by contrast, only amount to a local change of variables; we could write
466: \begin{displaymath}
467:   \begin{array}{rcl}\psi^\dagger_\sigma(\bs{x}) \!\!&\! \rightarrow \!&\!\!
468:     {\psi^\dagger_\sigma}'(\bs{x})=
469:     e^{i{\textstyle\frac{e}{\hbar c}}\Lambda(\bs{x})}\,
470:     \psi_\sigma^\dagger(\bs{x}),\\\rule{0pt}{18pt}
471:     \quad \psi_\sigma(\bs{x}) \!\!&\! \rightarrow \!&\!\!
472:     {\psi_\sigma}'(\bs{x})= 
473:     e^{-i{\textstyle\frac{e}{\hbar c}}\Lambda(\bs{x})}\,
474:     \psi_\sigma(\bs{x}),
475: \end{array}
476: \end{displaymath}
477: and then simply omit the primes.  This part of the gauge
478: transformation is often omitted as a choice of convention.
479: 
480: In the case of a BCS superconductor, such a convention would be all but
481: propitious, as it would suggests that the ground state is not gauge
482: invariant.  The apparent contradiction in the students train of
483: thought is immediately resolved as one uses the full and correct
484: prescription for a gauge transformation, 
485: \begin{equation}
486:   \label{e:gaugecc}
487:   c_{\bs{k}\sigma}^\dagger\rightarrow 
488:   e^{i{\textstyle\frac{e}{\hbar c}}\Lambda} c_{\bs{k}\sigma}^\dagger,
489:   \qquad \phi\rightarrow\phi-\frac{2e}{\hbar c}\Lambda,
490: \end{equation} 
491: where the transformation of the phase $\phi$ is the equivalent of 
492: (\ref{e:gaugepsi}) above.  Then the BCS ground state 
493: \begin{displaymath}
494:   \prod_{\bs{k}}\left(u_{\bs{k}} + v_{\bs{k}} e^{i\phi} 
495:     c_{\bs{k}\up}^\dagger\,c_{-\bs{k}\dw}^\dagger\right)\vac \!,
496: \end{displaymath}
497: is evidently gauge invariant; it is merely the label $\phi$ in
498: $\ket{\psi_\phi}$ which will be adjusted under a gauge transformation.
499: The transformation $\phi\rightarrow\phi-\frac{2e}{\hbar c}\Lambda$ is
500: also required for the classical (or Ginzburg--Landau) order parameter
501: field $\Psi^*(\bs{x})$, which is given by the expectation value of the
502: operator field
503: \begin{equation}
504:   \label{e:op}
505:   \hat\Psi^\dagger(\bs{x})\equiv 
506:   \psi^\dagger_\up(\bs{x})\,\psi^\dagger_\dw(\bs{x}),
507: \end{equation}
508: to have to the correct gauge transformation properties.  
509: %
510: For the BCS ground state,
511: \begin{equation}
512:   \label{e:bcsop}
513:   \Psi^*(\bs{x})= %\propto %\equiv
514:   \bra{\psi_\phi}\,\psi^\dagger_\up(\bs{x})\,\psi^\dagger_\dw(\bs{x}) 
515:   \ket{\psi_\phi}
516:   =\frac{1}{V}\sum_{\bs{k}} v_{\bs{k}}^* u_{\bs{k}}^{\phantom{*}}e^{-i\phi}
517: \end{equation} 
518: will transform as a field of charge $-2e$ under (\ref{e:gaugecc}),
519: \begin{equation}
520:   \label{e:opgauge}
521:   \Psi^*(\bs{x})\rightarrow 
522:   e^{i{\textstyle\frac{2e}{\hbar c}}\Lambda(\bs{x})}\Psi^*(\bs{x}).
523: \end{equation} 
524: This is the physically correct prescription.  When we couple
525: $\Psi^*(\bs{x})$ minimally to the electromagnetic gauge field, as
526: required by (\ref{e:opgauge}), we obtain the correct effective field
527: theory description of superconductivity.  This theory displays the
528: Higgs mechanism and yields London's equation.  (By contrast, if we
529: were to adhere to (\ref{e:gaugec}), $\Psi^*(\bs{x})$ would be
530: invariant, could not be coupled to the electromagnetic gauge field,
531: and no sensible effective field theory could be formulated.)
532: 
533: \section{ORDER PARAMETER CONSIDERATIONS}
534: %\section{Order parameter considerations}
535: \label{sec:ord}
536: 
537: Before proceeding further with the Higgs mechanism, %in this direction, 
538: I would like to return to the students train of thought and explain
539: what is wrong with his conclusion drawn from the BCS wave function in
540: position space.  The problem here is that while the ground state is
541: indeed unique for a finite system, there are many degenerate states in
542: the thermodynamic limit, which correspond to different numbers of
543: particles.  To understand this issue in depth, it is best to first
544: recall how rotational symmetry is spontaneously violated in
545: ferromagnets and antiferromagnets.  As a minimal model, we consider a
546: three dimensional cubic lattice of spins with spin quantum number $S$
547: and assume the Heisenberg Hamiltonian~\cite{auer}
548: \begin{equation}
549:   \label{e:heisen}
550:   H_J = J \sum_{\bond} \bs{S}_i \bs{S}_j
551: \end{equation} 
552: where the sum extends over all nearest-neighbor bonds $\bond$ and
553: $J<0$ ($J>0$) for a ferromagnet (an antiferromagnet).
554: 
555: In the case of a ferromagnet, the order parameter is given
556: by the total spin operator 
557: \begin{equation}
558:   \label{e:stot}
559:   \bs{S}_{\rm tot}\equiv\sum_i \bs{S}_i,
560: \end{equation} 
561: where the sum extends over all lattice sites.  It commutes with the
562: Hamiltonian,
563: \begin{equation}
564:   \label{e:commstot}
565:   \comm{H_J}{\bs{S}_{\rm tot}}=0,
566: \end{equation}
567: and it is hence possible to choose simultaneous eigenstates of the
568: Hamiltonian and the order parameter.  In other words, the degenerate
569: eigenstates of the order parameter corresponding to all possible
570: directions the magnetization vector
571: \begin{displaymath}
572:   \bs{M}=\braket{\bs{S}_{\rm tot}}%_0
573: \end{displaymath}
574: can point to, are simultaneously degenerate eigenstates of the
575: Hamiltonian.  (For a ferromagnet with $N$ spins, all the spins align
576: and the ground states are just the states with maximal total spin
577: $S_{\rm tot}=NS$.)  The ground state of the Hamiltonian is vastly
578: (\ie $2S_{\rm tot}\!+\!1$ fold) degenerate even if the
579: system is finite.
580: 
581: The situation is different in the case of the antiferromagnet.  The
582: order parameter is given by the N\'eel vector, which in operator form
583: is given by
584: \begin{equation}
585:   \label{e:neel}
586:   \bs{\hat N}\equiv\sum_{i\in {\rm A}}\bs{S}_i-\sum_{j\in {\rm B}}\bs{S}_j,
587: \end{equation} 
588: where A and B denote the two sublattices of the (bipartite) cubic
589: lattice.  It does {\it not} commute with the Hamiltonian:
590: \begin{displaymath}
591:   \comm{H_J}{\bs{\hat N}}\ne 0.
592: \end{displaymath}
593: This implies that we cannot choose simultaneous eigenstates for the
594: Hamiltonian and the order parameter.  In fact, a theorem due to
595: Marshall~\cite{marshall} states that the ground state for $N$ even is
596: unique and a spin singlet, or in other words, rotationally invariant.
597: (It is possible to choose simultaneous eigenstates of $H_J$ and
598: $\bs{S}_{\rm tot}$ as (\ref{e:commstot}) holds independently of the
599: sign of $J$.)  The classical N\'eel order parameter,
600: \begin{displaymath}
601:   \bs{N}=\bigl\langle\bs{\hat N}\bigr\rangle%_0
602: \end{displaymath}
603: will vanish for any finite system.  This is not to say that there is
604: no order for a finite system; it just manifests itself only through
605: long-range correlations in the staggered spin-spin correlation function: 
606: \begin{displaymath}
607:   \braket{\bs{S}_i \bs{S}_j }\rightarrow \pm\ \hbox{const.}\qquad 
608:   \hbox{as}\ i\!-\!j\rightarrow\infty
609: \end{displaymath}
610: where the $+$ sign applies for $i$ and $j$ on the same sublattice, 
611: the $-$ sign for $i$ and $j$ on different sublattices, 
612: and $i\!-\!j\rightarrow\infty$ is understood 
613: to denote a very large separation within (the finite volume of) the 
614: system.  As we approach the thermodynamic limit,
615: the difference in energy between the lowest singlet and lowest
616: eigenstates for %higher total spin 
617: $S_{\rm tot}=1,2,3\ldots$ vanishes, and the ground state becomes
618: degenerate (see Fig.\ \ref{f:spec}a).  These degenerate states can
619: now be classified by the directions of the N\'eel vector $\bs{N}$, and
620: the spontaneous breakdown of rotation symmetry is evident.
621: \begin{figure}[tbhp]
622:   \begin{center}
623:     \vspace{2mm}
624:     \includegraphics[width=\columnwidth]{spec.eps}
625:   \end{center}
626:   \caption{In antiferromagnets (a) and superconductors (b), the 
627:     ground state is unique for finite systems but degenerate in the
628:     thermodynamic limit.}
629:   \label{f:spec}
630: \end{figure}
631: 
632: The situation in superconductors is analogous to the antiferromagnet:
633: The (operator valued) order parameter (\ref{e:op}) does not commute
634: with the BCS Hamiltonian for any finite system even if we work in the
635: grand-canonical ensemble, and the ground state for any finite volume
636: will have a well defined particle number.  The difference in energy
637: between a system with $N$ or $N\pm 1$ or $N\pm 2$ {\it etc.}\ pairs,
638: however, will vanish in the thermodynamic limit (see
639: \mbox{Fig.~\ref{f:spec}b)}, and the many degenerate ground states can
640: be classified by the phase $\phi$ of the (classical) order parameter
641: \begin{equation}
642:   \label{e:bcsop2}
643:   \Psi^*(\bs{x})= \braket{\hat\Psi^\dagger(\bs{x})}=
644:   |\Psi^*(\bs{x})| e^{-i\phi(\bs{x})}.
645: \end{equation} 
646: The broken symmetry is of course also present in a system with a fixed
647: number of particles, but like in the case of the antiferromagnet, only
648: as a long-range correlation of the (operator valued) order parameter
649: field:
650: \begin{equation}
651:   \label{e:odlro}
652:   \braket{\hat\Psi^\dagger(\bs{x})\hat\Psi(\bs{y})}\rightarrow 
653:   \hbox{const.}\qquad \hbox{as}\ |\bs{x}-\bs{y}|\rightarrow\infty,
654: \end{equation}
655: where $\hat\Psi(\bs{x})\equiv \psi_\dw(\bs{x})\psi_\up(\bs{x})$
656: is simply the hermitian conjugate of $\hat\Psi^\dagger(\bs{x})$.  This
657: correlation is referred to as off diagonal long-range order
658: (ODLRO)~\cite{odlro}.  This type of order is characteristic to all
659: superfluids, whether charged (like a superconductor) or neutral (like
660: liquid helium), whether fermionic (like a superconductor or $^3$He) or
661: bosonic (like $^4$He).  For a bosonic superfluid, the (operator
662: valued) order parameter $\hat\Psi^\dagger(\bs{x})$ and its hermitian
663: conjugate $\hat\Psi(\bs{x})$ no longer create or annihilate a pair of
664: fermions, but simply create or annihilate a single boson (like a
665: $^4$He atom).
666: 
667: The ODLRO is already evident from the position space wave function
668: (\ref{e:bcsn}): Since all the pairs have condensed into the same
669: quantum state, which is translationally invariant as it does not
670: depend on the center-of-mass coordinates of the pairs, we expect to
671: obtain a finite overlap with the original ground state if we rather
672: clumsily (\ie via $\hat\Psi(\bs{y})$) remove a pair of particles at some
673: location $\bs{y}$ and equally clumsily (\ie via
674: $\hat\Psi^\dagger(\bs{x})$) recreate it at a distant location
675: $\bs{x}$.  In a superfluid or superconductor with a fixed number of
676: particles, the phase $\phi$ will align over the entire system, like
677: the direction of the staggered magnetization or N\'eel vector will
678: align in an antiferromagnet.
679: 
680: To illustrate the significance of the phase once more, let us consider
681: a large (but finite) superconductor A, and describe it as a
682: combination of two superconductors B and C: %(see \mbox{Fig.~\ref{fig:abc})}.  
683: \begin{center}
684: %\begin{picture}(180,40)(-100,-20)
685: \begin{picture}(180,40)(-90,-20)
686: % horizonal lines
687: \put(-55,-16){\makebox(0,0){\rule{70.pt}{ 0.3pt}}}
688: \put( 55,-16){\makebox(0,0){\rule{70.pt}{ 0.3pt}}}
689: \put(-55, 16){\makebox(0,0){\rule{70.pt}{ 0.3pt}}}
690: \put( 55, 16){\makebox(0,0){\rule{70.pt}{ 0.3pt}}}
691: %vertical lines
692: \put(-90,0){\makebox(0,0){\rule{ 0.3pt}{32.pt}}}
693: \put(-20,0){\makebox(0,0){\rule{ 0.3pt}{32.pt}}}
694: \put( 20,0){\makebox(0,0){\rule{ 0.3pt}{32.pt}}}
695: \put( 62,0){\makebox(0,0){\rule{ 0.3pt}{32.pt}}}
696: \put( 90,0){\makebox(0,0){\rule{ 0.3pt}{32.pt}}}
697: %labels
698: \put(  0,  0){\makebox(0,0){\large $=$}}
699: \put(-55,  0){\makebox(0,0){\large A}}
700: \put( 41,  0){\makebox(0,0){\large B}}
701: \put( 76,  0){\makebox(0,0){\large C}}
702: \end{picture}
703: \end{center}
704: If we label the ground states of each superconductor by its phase, 
705: we can obviously write
706: \begin{displaymath}
707:   \ket{\psi_\phi^{\text{A}}}=
708:   \ket{\psi_\phi^{\text{B}}}\otimes\ket{\psi_\phi^{\text{C}}}
709: \end{displaymath}
710: as the phase $\phi$ of the order parameter will align over the entire system.
711: If we now transform to a description in terms of fixed numbers of
712: pairs $N_a$ for each superconductor,
713: \begin{displaymath}
714:   \ket{\psi^a_N} = \int_0^{2\pi}\!\!d\phi\, 
715:   e^{-iN_a\phi}\ket{\psi^a_\phi},
716: \end{displaymath}
717: where $a$ can be A, B, or C, the ground state of A is no longer
718: a direct product of the ground states of B and C:
719: \begin{displaymath}
720:   \ket{\psi_{N_{\text{A}}}^{\text{A}}}
721:   \ne\ket{\psi_{N_{\text{B}}}^{\text{B}}}
722:   \otimes\ket{\psi_{N_{\text{A}}-N_{\text{B}}}^{\text{C}}},
723: \end{displaymath}
724: no matter how we choose $N_{\text{B}}$, as the phases no longer align.  So
725: while it is possible to describe a superconductor in a canonical
726: ensemble (\ie with fixed particle number), it is highly awkward to do
727: so.  It is comparable to a description of an antiferromagnet with
728: long-range N\'eel order in terms of an overall singlet ground state of
729: the system.
730: 
731: The most significant difference between an antiferromagnet and a
732: superfluid or superconductor with regard to %the concept of 
733: the order parameter is that the broken rotational symmetry in the
734: former case is much more evident to us, as all the macroscopic objects
735: in our daily life experience violate rotational symmetry at one level
736: or another.  In particular, the structure of the material in which
737: antiferromagnetic order occurs provides us already with a reference
738: frame for the direction the N\'eel order parameter may point to.  In
739: the case of the superconductor, we need a second superconductor to
740: have a reference direction for the phase, and an interaction between
741: the order parameter in both superconductors to detect a relative
742: difference in the phases.  (In practice, such an interaction may be
743: accomplished by a pair tunneling or so-called Josephson junction.)
744: The interference experiments will of course only be sensitive to the
745: relative phase, and not the absolute phase in any of the
746: superconductors, as all phases can, as a matter of principle, only be
747: specified relative to some reference phase.  In principle, the same is
748: true for rotational invariance, but in this case the fixed stars
749: provide us with a reference frame we perceive as ``absolute''.
750: 
751: We may conclude at this point that in a superfluid or superconductor,
752: a symmetry is spontaneously violated, but this symmetry is {\em not}
753: gauge invariance, but global U(1) phase rotation symmetry. This is
754: already evident from the fact that the discussion above made no
755: reference to whether the order parameter field
756: $\hat\Psi^\dagger(\bs{x})$ is charged or not, and equally well applies
757: to neutral superfluids, where $\hat\Psi^\dagger(\bs{x})$ carries no
758: charge.
759: 
760: There is, however, a very important difference between these two
761: cases.  If the order parameter field is neutral, the excitation
762: spectrum of the system contains a gapless (or in the language of
763: particle physics ``massless'') mode, a so-called Goldstone
764: boson~\cite{sid}, which physically corresponds to very slow spatial
765: variations in the direction (as for the case of broken rotational
766: invariance) or phase (as for the case of a superfluid) of the
767: classical order parameter field.  If the order parameter field is
768: charged, however, it couples to the electromagnetic gauge field, and
769: the Goldstone boson is absent due to the Higgs mechanism.  The
770: physical principle underlying the mechanism was discovered by
771: Anderson~\cite{phil63} in the context of superconductivity: as the
772: electromagnetic interaction is long-ranged, the mode corresponding to
773: very slow spatial variations in the phase $\phi$ of the
774: superconducting order parameter, which implies currents by the
775: equation of motion and hence also variations in the density of the
776: superfluid by the continuity equation, acquires a gap (or ``mass'')
777: given by the plasma frequency.
778: 
779: \section{EFFECTIVE FIELD THEORY} 
780: %\section{Effective field theory} 
781: \label{sec:eff}
782: 
783: Most of the phenomenology of superfluidity or superconductivity can be
784: derived from a simple effective field theory, which in the latter
785: case displays the Higgs mechanism.
786: %
787: It is probably best to turn directly to the low-energy effective
788: Lagrangian for the superconductor, as it contains the superfluid as
789: the special case where the coupling $e^*$ of the order parameter to
790: the electromagnetic gauge field is set to zero.  To motivate the
791: Lagrangian, recall first the Ginzburg--Landau~\cite{lg} expansion of the
792: free energy density in terms of the order parameter (which is now
793: normalized differently from (\ref{e:bcsop}) above) in the vicinity of
794: the critical temperature $T_{\text{c}}$, where the transition between normal
795: and superconducting phases occurs:
796: \begin{eqnarray}
797:   f(T,\Psi)\!\!&\!=\!&\!\! 
798:   \frac{1}{2m^*} \biggl|\Bigl(
799:   -i\hbar\nabla+\frac{e^*}{c}\bs{A}(\bs{x})\Bigr)\Psi(\bs{x})\biggr|^2
800:   \nonumber \\ \label{e:lg} \rule{0pt}{20pt} 
801:   &&\hspace{-10mm}+\, a(T)|\Psi(\bs{x})|^2+\frac{1}{2}b(T)|\Psi(\bs{x})|^4
802:   + \frac{1}{8\pi}\bs{B}(\bs{x})^2\hspace{5mm}
803: \end{eqnarray}
804: where $m^*$ and $-e^*=-2e$ are the effective mass and charge of the
805: electron pairs, respectively, and
806: $\bs{B}=\nabla\times\bs{A}$ is the magnetic field.
807: The material parameter $a(T)$ changes sign to become negative as we
808: pass through the transition from above, while $b(T)$ has to remain
809: positive.  
810: Minimizing the free energy in the superconducting phase yields that
811: (i) the gradient term must vanish,
812: (ii) $|\Psi(\bs{x})|^2=-{a}/{b}$,
813: %\begin{displaymath}
814: %  |\Psi(\bs{x})|^2=-\frac{a}{b}\, ,
815: %\end{displaymath}
816: and (iii) $\bs{B}=0$.  
817: This means that the amplitude $\Psi_0$ of the order parameter
818: $\Psi(\bs{x})=\Psi_0e^{i\phi}$ has to be fixed while the phase $\phi$,
819: which labels the many degenerate ground state configurations, can be
820: arbitrary as long as the variation over the sample is given by
821: \begin{displaymath}
822:   \nabla\phi=-\frac{e^*}{\hbar c}\bs{A},
823: \end{displaymath}
824: which implies $\phi(\bs{x})=\text{const.}$ if we choose the gauge
825: $\bs{A}(\bs{x})=0$.  In the vicinity of the transition, we may treat
826: $\Psi$ as a small parameter, which implies that the expansion
827: (\ref{e:lg}) provides us with a complete description of the system at
828: the level of thermodynamics.
829: 
830: The Ginzburg--Landau expansion is also helpful in motivating the low
831: energy effective Lagrange density at low temperatures.  To begin with, 
832: we may assume that since the amplitude fluctuations are massive, they
833: do not enter in the low energy description.
834: Taking $|\Psi(\bs{x})|$ to be constant, the free energy density above
835: reduces to a constant, an electromagnetic field contribution, and
836: \begin{equation}
837:   \label{e:fmag}
838:   f_{\text{mag}}=\frac{n_{\text s}}{2m^*}\,
839:   \Big(\hbar\nabla\phi(\bs{x})+\frac{e^*}{c}\bs{A}(\bs{x})\Bigl)^2,
840: \end{equation}
841: where $n_{\text s}=|\Psi_0|^2$ is a phenomenological parameter which
842: depends on the material and the temperature.  It has the dimension of
843: a density and is equal to the density of the superfluid in the absence
844: of currents and inhomogeneities at $T=0$, as we shall see below.  It
845: is usually referred to as the superfluid density, but it would be more
846: appropriate to use the superfluid stiffness ${n_{\text s}}/{m^*}$ as a
847: parameter instead~\cite{phil66}.  We will also see below that Galilean
848: invariance of the superfluid implies that $m^*$ is the bare mass of
849: the superfluid particles, \ie $m^*=2m_{\text{e}}$ for Cooper
850: pairs~\cite{rel}.
851: 
852: We take (\ref{e:fmag}) to be part of the potential energy in the
853: effective Lagrange density for the superfluid.  The remaining
854: contribution arises from the coupling of the charge of the superfluid
855: to the electrostatic potential $\Phi(\bs{x})$, which is in leading
856: order given by
857: \begin{displaymath}
858:   f_{\text{el}}=-n_{\text s} e^*\Phi(\bs{x}).
859: \end{displaymath}
860: This term is usually not included in the free energy of the
861: superconductor, as it is always canceled off by another such term with
862: opposite sign arising from the uniform positive background charge.  It
863: is essential to our effective field theory here, however, as it is
864: part of the Lagrange density for the superfluid, while the uniform
865: background charge is accounted for by another Lagrange density
866: \begin{equation}
867:   \label{e:lb}
868:   \mathcal{L}_{\text b}(x)=-n_{\text{s}} e^*\Phi(x),
869: \end{equation}
870: where $x=(ct,\bs{x})$ denotes spacetime.  Note that $f_{\text{el}}$
871: is not invariant under (time dependent) gauge transformations
872: \begin{eqnarray}
873:   \label{e:gauge}
874:   \phi(x)\!&\!\to \!&\!\phi(x)-\frac{e^*}{\hbar c}\Lambda(x),
875:   \nonumber\\\rule{0pt}{16pt}
876:   \Phi(x)\!&\!\to \!&\!\Phi(x)-\frac{1}{c}\partial_t \Lambda(x),
877:   \\\rule{0pt}{12pt}
878:   \bs{A}(x)\!&\!\to \!&\!\bs{A}(x)+\nabla\Lambda(x).
879:   \nonumber
880: \end{eqnarray}
881: 
882: We now turn to the kinetic energy term in the effective Lagrange density.
883: The simplest gauge invariant Lagrange density containing
884: both potential energy terms above is given by~\cite{feyn}
885: \begin{eqnarray}
886: \label{e:lmin}
887: \mathcal{L}_{\text s}(x)
888: \!&\!=\!&\!-\,n_{\text{s}} 
889: \Big(\hbar\partial_t \phi(x)-e^*\Phi(x)\Bigl)\cr\rule{0pt}{20pt}
890: &&\!-\,\frac{n_{\text{s}}}{2m^*}
891: \Big(\hbar\nabla\phi(x)+\frac{e^*}{c}\bs{A}(x)\Bigl)^2.\hspace{5mm}
892: \end{eqnarray}
893: This Lagrange density, however, cannot be complete.  The only term
894: containing a time derivative, $\partial_t \phi(x)$, appears as a total
895: derivative (here time derivative of $\phi$) and hence does not affect
896: the Euler--Lagrange equations of motion.  It is nonetheless of
897: physical significance, as it both ensures gauge invariance and
898: accounts for the leading contribution to the particle density, as we
899: will see below.
900: 
901: To obtain a second order time derivative term, 
902: recall that the characteristic feature of a neutral (\ie $e^*=0$)
903: superfluid is that the only excitation at low energies is a sound wave
904: with a linear dispersion
905: \begin{equation}
906: \label{e:lindis}
907: \omega (\bs{k}) =v|\bs{k}|,
908: \end{equation}
909: where $\bs{k}$ is the wave number and $v$ is the velocity of sound in
910: the fluid.  As we wish the effective Lagrange density for the
911: superfluid both to be gauge invariant and to yield (\ref{e:lindis}) as
912: an equation of motion for $e^*=0$, we arrive at
913: \begin{eqnarray}
914:   \label{e:ls}
915:   \mathcal{L}_{\text s}(x)
916:   \!&\!=\!&\!-\,
917:   n_{\text{s}}\Big(\hbar\partial_t \phi(x)-e^*\Phi(x)\Bigl)
918:   \\ \nonumber\rule{0pt}{18pt}
919:   &&\hspace{-13mm}\!\!+\frac{n_{\text{s}}}{2m^*}
920:   \biggl\{\frac{1}{v^2}\Big(\hbar\partial_t \phi(x)\!-\!
921:   e^*\Phi(x)\Bigl)^2- 
922:   \Big(\hbar\nabla\phi(x)\!+\!\frac{e^*}{c}\bs{A}(x)\Bigl)^{\!2}
923:   \biggr\}.
924: \end{eqnarray}
925: With 
926: \begin{equation}
927:   \label{e:dmu}
928:   D_\mu\phi\equiv\hbar\:\! \partial_\mu \phi-\frac{e^*}{c}A_\mu\:\!, 
929: \end{equation}
930: where $(\partial_\mu) %\equiv\frac{\partial}{\partial x^\mu}
931: = ({\frac{1}{c}}\partial_t, \nabla)$ and $(A_\mu)=(\Phi,-\bs{A})$,
932: the Lagrange density may also be written
933: \begin{equation}
934:   \label{e:ls2}
935:   \mathcal{L}_{\text s}=-%\:\!\hbar
936:   \:\! c\:\! n_{\text{s}}\;\!D_0\phi\:\!+\:\!
937:   \frac{n_{\text{s}}}{2m^*}\,
938:   \biggl\{\frac{c^2}{v^2}\,(D_0\phi)^2-(D_i\phi)^2\biggr\},
939: \end{equation}
940: where $i=1,2,3$.
941: The total Lagrangian of the system is given by
942: \begin{equation}
943:   \label{e:ltot}
944:   L=\int\! d^{3\,}\!\bs{x} 
945:   \left\{\mathcal{L}_{\text s}(x) + \mathcal{L}_{\text b}(x) + 
946:     \mathcal{L}_{\text{em}}(x)\right\},
947: \end{equation}
948: where 
949: \begin{equation}
950:   \label{e:lmax}
951:   \mathcal{L}_{\text{em}}=-\frac{1}{16\pi}F_{\mu\nu}F^{\mu\nu}
952:   \quad\text{with}\quad
953:   F_{\mu\nu}\equiv\partial_\mu A_\nu-\partial_\nu A_\mu
954: \end{equation}
955: denotes the standard Maxwell Lagrange density for electromagnetism. 
956: 
957: The astonishing feature is now that this simple Lagrangian for the
958: compact U(1) field $\phi(x)$ (compact since the values $\phi$ and
959: $\phi+2\pi$ describe the same physical state and hence must be
960: identified) coupled to the electromagnetic gauge field accounts for
961: all the essential features of superfluidity or superconductivity.
962: There are also important corrections to it, but we will discover
963: them automatically as we proceed.
964: 
965: In order understand the physical content of the Lagrangian, it is
966: highly instructive to study its symmetries, in particular particle
967: number conservation and invariance under translations in space and
968: time.  We wish our analysis to apply both to the case of a neutral and
969: a charged superfluid.  In the former case, the theory is no longer
970: invariant under a local U(1) gauge transformation (as the
971: electromagnetic gauge transformation (\ref{e:gaugeprel}) reduces to
972: the identity transformation for $e^*=0$) but still invariant under a
973: global U(1) rotation
974: \begin{equation}
975:   \label{e:rot}
976:   \Psi(x)\rightarrow 
977:   e^{i\lambda}\Psi(x)\quad\text{or}\quad
978:   \phi(x)\rightarrow \phi(x)+\lambda,
979: \end{equation}
980: where $\lambda$ is independent of spacetime.  Physically, this
981: symmetry corresponds to particle (or Cooper pair) number conservation.
982: According to Noether's theorem~\cite{pes}, if under a given transformation
983: the Lagrange density only changes by a total derivative,
984: \begin{displaymath}
985:   \quad D\mathcal{L}(x)\equiv 
986:   \frac{d\mathcal{L}(x,\lambda)}{d\lambda}\bigg|_{\lambda=0}\biggr.
987:   =\partial_\mu F^\mu(x),
988: \end{displaymath}
989: there is a conserved current associated with this symmetry: 
990: \begin{equation}
991:   \label{e:current}
992:   J^\mu(x)=\text{const.}\cdot\left\{
993:     \frac{\delta L_{\text s}}{\delta(\partial_\mu\phi(x))} D\phi(x)-F^\mu(x),
994:   \right\}
995: \end{equation}
996: where
997: \begin{displaymath}
998:   D\phi(x)\equiv\frac{d\phi(x,\lambda)}{d\lambda}\bigg|_{\lambda=0}\biggr.
999: \end{displaymath}
1000: Current conservation means $\partial_\mu J^\mu=0$.  Since
1001: (\ref{e:rot}) yields $F^\mu(x)=0$ and $D\phi(x)=1$, the particle
1002: four-current $(J^\mu)=(c\rho,\bs{J})$ is given by
1003: \begin{equation}
1004:   \label{e:currents}
1005:   J^\mu(x)
1006:   =-\frac{1}{\hbar}\frac{\delta L_{\text s}}{\delta(\partial_\mu\phi(x))} 
1007:   =-\frac{\delta L_{\text s}}{\delta(D_\mu\phi(x))}, 
1008: \end{equation}
1009: where we have chosen the normalization such that the electric current
1010: equals the charge times the particle current:
1011: \begin{displaymath}
1012:   -c\;\!\frac{\delta L_{\text s}}{\delta(A_\mu(x))}
1013:   =-e^*J^\mu(x). 
1014: \end{displaymath}
1015: The Lagrange density (\ref{e:ls}) yields for the particle density 
1016: \begin{eqnarray}
1017:   \label{e:rhocur}
1018:   \rho(x)\!&\!=\!&\! 
1019:   -\frac{1}{\hbar}\frac{\delta L_{\text s}}{\delta(\partial_t\phi(x))}
1020:   \\ \label{e:rho}\rule{0pt}{16pt} 
1021:   \!&\!=\!&\! 
1022:   n_{\text s}-\frac{n_{\text s}}{m^*}\frac{1}{v^2}
1023:   \Bigl(\hbar\partial_t \phi(x)-e^*\Phi(x)\Bigr)
1024: \end{eqnarray}
1025: % \begin{equation}
1026: %   \label{e:rho}
1027: %   \rho(x)
1028: %   =-\frac{1}{\hbar}\frac{\delta L_{\text s}}{\delta(\partial_t\phi(x))}
1029: %   =n_{\text s}-\frac{n_{\text s}}{m^*}\frac{1}{v^2}
1030: %   \Bigl(\hbar\partial_t \phi(x)-e^*\Phi(x)\Bigr)
1031: % \end{equation}
1032: and for the particle current
1033: \begin{eqnarray}
1034:   \label{e:jcur}
1035:   \bs{J}(x)\!&\!=\!&\!
1036:   -\frac{1}{\hbar}\frac{\delta L_{\text s}}{\delta(\nabla\phi(x))}
1037:   \\ \label{e:j}\rule{0pt}{16pt} 
1038:   \!&\!=\!&\! 
1039:   \frac{n_{\text s}}{m^*}
1040:   \Bigl(\hbar\nabla\phi(x)+\frac{e^*}{c}\bs{A}(x)\Bigr).
1041: \end{eqnarray}
1042: % \begin{equation}
1043: %   \label{e:j}
1044: %   \bs{J}(x)
1045: %   =-\frac{1}{\hbar}\frac{\delta L_{\text s}}{\delta(\nabla\phi(x))}
1046: %   =\frac{n_{\text s}}{m^*}
1047: %   \Bigl(\hbar\nabla\phi(x)+\frac{e^*}{c}\bs{A}(x)\Bigr).
1048: % \end{equation}
1049: The corresponding conservation law is just the continuity equation
1050: \begin{equation}
1051: \label{e:cont}
1052: \partial_t\rho+\nabla\bs{J}=0.
1053: \end{equation}
1054: 
1055: Note that since our Lagrangian (\ref{e:ltot}) does not depend on $\phi(x)$,
1056: but only on derivatives of $\phi(x)$, \ie
1057: \begin{displaymath}
1058:   \frac{\delta L_{\text s}}{\delta(\phi(x))}=0,
1059: \end{displaymath}
1060: (\ref{e:currents}) implies that the current conservation law
1061: \begin{displaymath}
1062:   -\hbar\partial_\mu J^\mu(x)=\partial_\mu 
1063:   \frac{\delta L_{\text s}}{\delta(\partial_\mu\phi(x))}=0
1064: \end{displaymath}
1065: is equivalent the Euler--Lagrange equation for the field $\phi(x)$.
1066: For a neutral superfluid, we obtain 
1067: \begin{equation}
1068:   \left(\frac{1}{v^2}\partial_t^2 -\nabla^2\right)\phi(x)\!=0
1069: \end{equation}
1070: and hence the dispersion (\ref{e:lindis}) by Fourier transformation.
1071: 
1072: The most important implication of (\ref{e:currents}) for the
1073: particle four-current is, however, that the density  
1074: %$-\hbar c \rho(x)$ is equal to
1075: $\rho(x)$ is up to a numerical factor equal to
1076: the momentum field $\pi(x)$ conjugate to $\phi(x)$:
1077: \begin{equation}
1078:   \label{e:mom}
1079:   -\hbar\rho(x)=\pi(x)\equiv
1080:   \frac{\delta L_{\text s}}{\delta(\partial_t\phi(x))}. 
1081: \end{equation}
1082: We may hence go over to an Hamiltonian formulation, and write
1083: the Hamiltonian density
1084: \begin{equation}
1085:   \label{e:h}
1086:   \mathcal{H}_{\text s}(x)\equiv 
1087:   -\hbar\rho(x)\:\! \partial_t\phi(x)-\mathcal{L}_{\text s}(x),
1088: \end{equation}
1089: which is now considered a functional of 
1090: $\rho(x)$, $\partial_i\phi(x)$, $\Phi(x)$, %$A_0(x)$, 
1091: and $A_i(x)$, but not
1092: $\partial_t\phi(x)$.  (In principle, $\mathcal{H}_{\text s}$ could
1093: also depend through $\mathcal{L}_{\text s}$ on $\phi(x)$ and $x$.
1094: Note also that (\ref{e:h}) as the generator of time translations is
1095: not invariant under time dependent gauge transformations, while the
1096: equations of motions below are invariant.)  The Hamiltonian is of
1097: course given by
1098: \begin{equation}
1099:   \label{e:htot}
1100:   H_{\text s}=\int\! d^{3\,}\!\bs{x}\;\! \mathcal{H}_{\text s}(x) .
1101: \end{equation}
1102: Hamilton's equations are in analogy to the familiar equations
1103: \begin{displaymath}
1104:   \dot q=\frac{\partial H(p,q)}{\partial p},\
1105:   \dot p=-\frac{\partial H(p,q)}{\partial q}
1106: \end{displaymath}
1107: from classical mechanics given by
1108: \begin{equation}
1109:   \label{e:h1}
1110:   \partial_t\phi(x)=\frac{\delta H_{\text s}}{\delta (\pi(x))}
1111:   =-\frac{1}{\hbar}\frac{\delta H_{\text s}}{\delta (\rho(x))}
1112: \end{equation}
1113: and
1114: \begin{equation}
1115:   \label{e:h2}
1116:   -\hbar\:\!\partial_t \rho(x)=\partial_t \pi(x)=\partial_i 
1117:   \frac{\delta H_{\text s}}{\delta (\partial_i\phi(x))}
1118:   -\frac{\delta H_{\text s}}{\delta (\phi(x))}.
1119: \end{equation}
1120: With regard to the explicit equations of motion for the fields, these
1121: equations are equivalent to the Euler--Lagrange equation.  They
1122: provide, however, additional information regarding the physical
1123: interpretation.  To extract this information, it is best
1124: to study first the other conservation laws corresponding to
1125: energy and momentum.
1126: 
1127: 
1128: The theory is invariant under spacetime translations $x\rightarrow
1129: x-e\lambda$, where $e$ is an arbitrary unit vector in spacetime (\eg
1130: $e=(1,0,0,0)$ or $e^\nu=\delta_{0\nu}$ for a translation in time).
1131: The infinitesimal translations are equivalent to the field and density
1132: transformations~\cite{pes}
1133: \begin{eqnarray}
1134: \label{e:transl}
1135: \phi(x)\!&\!\rightarrow\!&\!\phi(x+e\lambda)= 
1136: \phi(x)+\lambda e^\nu\partial_\nu\phi(x),
1137: \nonumber\\ \nonumber \rule{0pt}{15pt}
1138: A_\mu(x)\!&\!\rightarrow\!&\! A_\mu(x+e\lambda)= 
1139: A_\mu(x)+\lambda e^\nu\partial_\nu A_\mu(x),
1140: \\ \nonumber \rule{0pt}{15pt}
1141: \mathcal{L}(x)\!&\!\rightarrow\!&\!\mathcal{L}(x+e\lambda)
1142: =\mathcal{L}(x)+\lambda\partial_\nu(e^\nu\mathcal{L}(x)),
1143: \end{eqnarray}
1144: which implies $D\phi=e^\nu\partial_\nu\phi$, 
1145: $DA_\mu=e^\nu\partial_\nu A_\mu$, and $F^\mu(x)=e^\mu\mathcal{L}(x)$.
1146: The conserved current associated  
1147: with this symmetry is according to (\ref{e:current}) given by
1148: \begin{displaymath}
1149: J^\mu
1150: = \frac{\delta L}{\delta(\partial_\mu\phi)} 
1151: e^\nu\partial_\nu\phi
1152: +\frac{\delta L}{\delta(\partial_\mu A_\kappa)} e^\nu\partial_\nu A_\kappa
1153: -e^\mu\mathcal{L}
1154: = e_\nu T_{\scriptscriptstyle\text{can}}^{\mu\nu},
1155: \end{displaymath}
1156: where the canonical energy-momentum tensor
1157: $T_{\scriptscriptstyle\text{can}}^{\mu\nu}$ is the sum of the
1158: contributions from the superfluid, the uniformly charged background,
1159: and the electromagnetic field:
1160: \begin{displaymath}
1161: T_{\scriptscriptstyle\text{can}}^{\mu\nu}=
1162: T_{\text{s},\scriptscriptstyle\text{can}}^{\mu\nu}+
1163: T_{\text{b},\scriptscriptstyle\text{can}}^{\mu\nu}+
1164: T_{\text{em},\scriptscriptstyle\text{can}}^{\mu\nu}
1165: \end{displaymath}
1166: where 
1167: \begin{eqnarray}
1168: \label{e:tcans}\rule{0pt}{15pt}
1169: T_{\text{s},\scriptscriptstyle\text{can}}^{\mu\nu}\!&\!=\!&\!
1170: \frac{\delta L_{\text s}}{\delta(\partial_\mu\phi)} 
1171: \partial^{\:\!\nu\!}\phi
1172: -g^{\mu\nu}\mathcal{L}_{\text s},
1173: \\ \nonumber\rule{0pt}{16pt}
1174: T_{\text{b},\scriptscriptstyle\text{can}}^{\mu\nu}\!&\!=\!&\!
1175: -g^{\mu\nu}\mathcal{L}_{\text b},
1176: \\ \label{e:tcane}\rule{0pt}{20pt}
1177: T_{\text{em},\scriptscriptstyle\text{can}}^{\mu\nu}\!&\!=\!&\!
1178: \frac{\delta L_{\text{em}}}{\delta(\partial_\mu A_\kappa)} 
1179: \partial^{\:\!\nu\!}A_\kappa
1180: -g^{\mu\nu}\mathcal{L}_{\text{em}}.
1181: \end{eqnarray}
1182: % Only the total current is conserved,
1183: % $\partial_\mu T_{\scriptscriptstyle\text{can}}^{\mu\nu}=0$. 
1184: The conservation law $\partial_\mu
1185: T_{\scriptscriptstyle\text{can}}^{\mu\nu}=0$ describes energy
1186: conservation for $\nu=0$ and momentum conservation for $\nu=i$.  The
1187: $\mu=0$ components of $T_{\scriptscriptstyle\text{can}}^{\mu\nu}$
1188: correspond to energy and momentum densities; in particular, $-c\int\!
1189: d^{3\,}\!x\, T_{\scriptscriptstyle\text{can}}^{00}$ generates
1190: translations in time and $\int\! d^{3\,}\!x\,
1191: T_{\scriptscriptstyle\text{can}}^{0i}$ translations in space.
1192: 
1193: In the case of a gauge theory, like the theory of a charged superfluid
1194: we consider here, it is not possible to interpret
1195: $T_{\text{s,}\scriptscriptstyle\text{can}}^{00}$ as the energy or
1196: $\frac{1}{c}T_{\text{s,}\scriptscriptstyle\text{can}}^{0i}$ as the
1197: kinematical momentum density of the superfluid.  The reason is simply
1198: that (\ref{e:tcans}) (and also (\ref{e:tcane})) is not gauge
1199: invariant.  To circumvent this problem, we simply supplement the naive
1200: translations by a suitable gauge transformation, such that the fields
1201: transform covariantly:
1202: \begin{eqnarray}
1203: \phi&\!\rightarrow\!&\! \phi+\lambda 
1204: e^\nu\Bigl(\partial_\nu\phi-\frac{e^*}{\hbar c}A_\nu\Bigr),
1205: \nonumber \\ \nonumber \rule{0pt}{15pt}
1206: A_\kappa\!\!&\!\rightarrow\!&\! A_\kappa+\lambda 
1207: e^\nu (\partial_\nu A_\kappa-\partial_\kappa A_\nu).
1208: \end{eqnarray}
1209: The gauge transformation is hence given by (\ref{e:gauge}) with
1210: $\Lambda(x)=\lambda e^\nu A_\nu(x)$.  This yields the ``kinematical''
1211: energy momentum tensor, with contributions from the superfluid and the
1212: electromagnetic field%~\cite{gww}
1213: \begin{eqnarray}
1214:   \label{e:tensor}
1215:   T_{\text s}^{\mu\nu}\!&\!=\!&\!
1216:   \frac{\delta L_{\text s}}{\delta(D_\mu\phi)} D^{\:\!\nu\!}\phi
1217:   -g^{\mu\nu}_{\phantom{\mu\nu}}\mathcal{L}_{\text s},
1218:   \label{e:ts} \\ \nonumber \rule{0pt}{20pt}
1219:   T_{\text{em}}^{\mu\nu}\!&\!=\!&\!
1220:   \frac{\delta L_{\text{em}}}{\delta(F_{\mu\kappa})} 
1221:   F_{\phantom{\:\!\nu}\kappa}^{\:\!\nu}
1222:   -g^{\mu\nu}_{\phantom{\mu\nu}}\mathcal{L}_{\text{em}}.
1223: \end{eqnarray}
1224: These expressions are manifestly gauge invariant.  Using
1225: (\ref{e:currents}) with $\mu=0$ or (\ref{e:rhocur}), 
1226: we can write the energy density of the superfluid
1227: \begin{equation}
1228:   \label{e:eden}
1229:   T_{\text s}^{00}(x)= -\rho(x)
1230:   \Bigl(\hbar\partial_t\phi(x) -e^* \Phi(x)\Bigr)-\mathcal{L}_{\text s}(x).
1231: \end{equation}
1232: Note that this is numerically equal to
1233: \begin{equation}
1234:   \label{e:eden1}
1235:   T_{\text s}^{00}(x)=\mathcal{H}_{\text s}(x)+\rho(x)e^* \Phi(x).
1236: \end{equation}
1237: Similarly, we can write the momentum density
1238: \begin{equation}
1239: \label{e:mden}
1240: \frac{1}{c}T_{\text s}^{0i}(x)=
1241: \rho(x)\Bigl(\hbar\nabla\phi(x)+\frac{e^*}{c}\bs{A}(x)\Bigr).
1242: \end{equation}
1243: 
1244: 
1245: We can use this expression to introduce the superfluid velocity
1246: $\bs{v}_{\text s}(x)$.  In terms of $\bs{v}_{\text s}(x)$, the momentum
1247: density of the superfluid has to be given by
1248: \begin{equation}
1249:   \label{e:tvs}
1250:   \frac{1}{c}T_{\text s}^{0i}(x)\stackrel{!}{=}
1251:   \rho(x)\, m^*\;\!\!\bs{v}_{\text s}(x),
1252: \end{equation}
1253: which leads us to define
1254: \begin{equation}
1255:   \label{e:vs}
1256:   m^*\bs{v}_{\text s}(x)\equiv
1257:   \hbar\nabla\phi(x)+\frac{e^*}{c}\bs{A}(x)=D_i\phi(x).
1258: \end{equation}
1259: Since $\bs{v}_{\text s}(x)$ is to be interpreted as a physical velocity, it
1260: has to transform like a velocity under a Galilean transformation, 
1261: \begin{equation}
1262:   \label{e:gal1}
1263:   \bs{v}_{\text s}(x)\ \rightarrow\ \bs{v}_{\text s}(x) + \bs{u}.
1264: \end{equation}
1265: The total momentum of the superfluid will hence transform according to
1266: \begin{equation}
1267:   \label{e:gal2}
1268:   \frac{1}{c}\int\! d^{3\,}\!\bs{x}\, T_{\text s}^{0i}(x)
1269:   \ \rightarrow\ \frac{1}{c}\int\! d^{3\,}\!\bs{x}\, T_{\text s}^{0i}(x) 
1270:   +\!\int\! d^{3\,}\!\bs{x}\:\!\rho(x)\  m^*\bs{u},
1271: \end{equation}
1272: which implies directly that in a translationally invariant system,
1273: $m^*$ has to be the bare mass of the superfluid particles or Cooper
1274: pairs~\cite{conn}.
1275: 
1276: It should also be possible to express the particle current in terms 
1277: of the superfluid velocity.  
1278: Since the same particles which carry the momentum also
1279: carry the current, the particle current has to be given by 
1280: \begin{equation}
1281: \label{e:js}
1282: \bs{J}(x)=\rho(x)\:\!\bs{v}_{\text s}(x).
1283: \end{equation}
1284: This is almost, but not quite, equivalent to our earlier expression
1285: (\ref{e:j}), as $n_{\text s}$ is only the leading contribution to
1286: $\rho(x)$.  So either (\ref{e:js}) with (\ref{e:vs}) or (\ref{e:j}) is
1287: not fully correct.  To see which one, recall that we have only used
1288: the general expression for the density %(\ref{e:currents}) with $\mu=0$
1289: (\ref{e:rhocur})
1290: as defined through particle number conservation %(\ref{e:currents}) 
1291: in obtaining (\ref{e:mden}) and hence (\ref{e:vs}) and (\ref{e:js})
1292: from (\ref{e:ts}), while we have used the explicit expression for the
1293: Lagrange density (\ref{e:ls}) in obtaining (\ref{e:j}).  In other
1294: words, only symmetry considerations enter in (\ref{e:mden}), while
1295: (\ref{e:j}) depends explicitly on the Lagrange density.  The
1296: expression for the momentum density (\ref{e:mden}), and hence our
1297: definition of the superfluid velocity (\ref{e:vs}), is therefore
1298: exact, while the expression (\ref{e:j}) for the particle current is
1299: only an approximation~\cite{phila}.
1300: 
1301: The expression for the current, however, will assume the exact and
1302: physically correct form (\ref{e:js}) if we introduce suitable
1303: corrections to the effective Lagrangian.  To obtain these, we simply
1304: require the Lagrangian to satisfy~\cite{gww}
1305: \begin{equation}
1306:   \label{e:gww}
1307:   \frac{1}{c}\:\!T_{\text s}^{0i}(x)= m^*J^i(x)
1308: \end{equation}
1309: or
1310: \begin{equation}
1311: \frac{1}{c}\frac{\delta L_{\text s}}{\delta(D_0\phi)} D^{\:\!i\!}\phi
1312:  = -m^*\frac{\delta L_{\text s}}{\delta(D_i\phi)}.
1313:  \end{equation}
1314: Upon integration of this equation we find that the Lagrange density
1315: must be of the form
1316: \begin{equation}
1317: \mathcal{L}_{\text s}
1318: =P\Bigl(cD_0\phi+\frac{1}{2m^*} (D_i\phi)^2\Bigr),
1319: \end{equation}
1320: where $P$ is an arbitrary polynomial.
1321: Our superfluid Lagrange density (\ref{e:ls2}) will assume this form if we
1322: add third and fourth order corrections in $D_\mu\phi$; the full
1323: superfluid Lagrange density is then given by~\cite{gww,schakel}
1324: \begin{eqnarray}
1325: \label{e:ls3}
1326: \mathcal{L}_{\text s}\!&\!=\!&\!-\,n_{\text{s}}\;\!
1327: \Bigl(cD_0\phi+\frac{1}{2m^*} (D_i\phi)^2\Bigr)\cr\rule{0pt}{18pt}
1328: &&\!+\,\frac{n_{\text{s}}}{2m^*}\frac{1}{v^2}\;\!
1329: \Bigl(cD_0\phi+\frac{1}{2m^*} (D_i\phi)^2\Bigr)^2.
1330: \end{eqnarray}
1331: It yields for the particle density 
1332: \begin{eqnarray}
1333: \label{e:rho3}
1334: \rho(x)\!&\!=\!&\!
1335: -\frac{1}{c}\frac{\delta L_{\text s}}{\delta(D_0\phi)}
1336: \nonumber \\ \rule{0pt}{17pt}
1337: %\!&\!=\!&\!
1338: && \hspace{-10mm}=
1339: n_{\text s}-\frac{n_{\text s}}{m^*}\frac{1}{v^2}
1340: \Bigl(cD_0\phi+\frac{1}{2m^*}(D_i\phi)^2\Bigr)
1341: \\ \nonumber \rule{0pt}{17pt}
1342: %\!&\!=\!&\!
1343: && \hspace{-10mm}=
1344: n_{\text s}-\frac{n_{\text s}}{m^*}\frac{1}{v^2}
1345: \biggl\{\Bigl(\hbar\partial_t \phi-e^*\Phi\Bigr)+\frac{1}{2m^*}
1346: \Bigl(\hbar\nabla\phi+\frac{e^*}{c}\bs{A}\Bigl)^2\biggr\}
1347: \end{eqnarray}
1348: and for the particle current
1349: \begin{equation}
1350: \label{e:j3}
1351: \bs{J}(x)=-\frac{\delta L_{\text s}}{\delta(D_i\phi)}
1352: =\rho(x)\:\!\bs{v}_{\text s}(x),
1353: \end{equation}
1354: where $\rho(x)$ and $\bs{v}_{\text s}(x)$ are given by (\ref{e:rho3}) and
1355: (\ref{e:vs}).
1356: 
1357: Let us now return to Hamilton's equations, and in particular their
1358: physical interpretation.  With (\ref{e:eden1}) we may rewrite
1359: (\ref{e:h1}) as
1360: \begin{eqnarray}
1361: \label{e:h1'}
1362: \hbar\partial_t\phi(x)
1363: \!&\!=\!&\!-\frac{\delta H_{\text s}}{\delta (\rho(x))}
1364: =-\frac{\partial\mathcal{H}_{\text s}(x)}{\partial\rho(x)}
1365: =-\frac{\partial T_{\text s}^{00}(x)}{\partial\rho(x)}+e^*\Phi(x)
1366: \nonumber\\ %\nonumber
1367: \rule{0pt}{14pt}
1368: \!&\!\!=\!\!&\!\!-\mu(x)+e^*\Phi(x),
1369: \end{eqnarray}
1370: where we have used the definition of the chemical potential.  This is one
1371: of two equations Anderson~\cite{phil66} refers to as ``characteristic of
1372: superfluidity''.  In analogy to the definition (\ref{e:vs}) of
1373: $\bs{v}_{\text s}(x)$, we rewrite it for later purposes as
1374: \begin{equation}
1375: \label{e:mu}
1376: -\mu(x)=\hbar\partial_t\phi(x)-e^*\Phi(x)=cD_0\phi(x).
1377: \end{equation}
1378: Taking the gradient and adding $\frac{e^*}{c}\partial_t\bs{A}(x)$ 
1379: on both sides of (\ref{e:h1'}), we obtain 
1380: \begin{eqnarray}
1381:   \label{e:muelch}
1382:   \partial_t\Bigl(\hbar\nabla\phi(x)+\frac{e^*}{c}\bs{A}(x)\Bigr)
1383:   \!&\!=\!&\!-\nabla\mu(x)-e^*\bs{E}(x)\cr \rule{0pt}{16pt}
1384:   \!&\!=\!&\!-\nabla\mu_{\scriptscriptstyle\text{el.chem.}}(x),
1385: \end{eqnarray}
1386: where we used the definitions of the electric field,
1387: \begin{displaymath}
1388: \bs{E}\equiv -\nabla\Phi-\frac{1}{c}\partial_t\bs{A},
1389: \end{displaymath}
1390: and of the electrochemical potential.  With (\ref{e:vs}) we may write  
1391: \begin{equation}
1392: \label{e:acc}
1393: m^*\:\!\partial_t\bs{v}_{\text s}(x)
1394: =-\nabla\mu_{\scriptscriptstyle\text{el.chem.}}(x).
1395: \end{equation}
1396: The gradient of the electrochemical potential (or chemical potential
1397: for a neutral superfluid) is usually~\cite{phil66} identified with
1398: minus the total force on the particles, and (\ref{e:acc}) is referred
1399: to as the ``acceleration equation''.  This is, however, not quite
1400: correct.  $\partial_t \bs{v}_{\text s}$ in (\ref{e:acc}) denotes the
1401: time derivative in the superfluid velocity field at spacetime $x$
1402: (known as ``local acceleration'' in hydrodynamics), while the force on
1403: the particles is given by the time derivative of the velocity of a
1404: given particle in the fluid at $x$ (``substantial acceleration'' in
1405: hydrodynamics):
1406: \begin{equation}
1407: \label{e:force}
1408: \frac{1}{m^*}\bs{F}=\frac{d\bs{v}_{\text s}}{dt}
1409: =\partial_t\bs{v}_{\text s}+(\bs{v}_{\text s}\nabla)\bs{v}_{\text s}.
1410: \end{equation}
1411: Nonetheless, (\ref{e:acc}) is one of the fundamental equations in the
1412: phenomenology of superfluidity.  It states %, for one thing, 
1413: that if
1414: there is a gradient in the electrochemical (or chemical) %for $e^*=0$) 
1415: potential in a superconductor (or superfluid), the superfluid
1416: will be ``accelerated'' without any frictional damping.  On the other
1417: hand, if the superfluid flow is stationary, the electrochemical (or
1418: chemical) potential has to be constant across the superconductor (or
1419: superfluid).  Since a voltmeter measures a difference in the
1420: electrochemical potential, there cannot be a voltage across a
1421: superconductor unless the flow is ``accelerated''.
1422: 
1423: Let us now turn to Hamilton's second equation (\ref{e:h2}).  We first
1424: rewrite it as
1425: \begin{equation}
1426: \label{e:h2'}
1427: -\hbar\:\!\partial_t\rho(x)
1428: =\partial_i \frac{\delta H_{\text s}}{\delta (\partial_i\phi(x))}
1429: -\frac{\partial\mathcal{H}_{\text s}(x)}{\partial\phi(x)}.
1430: \end{equation}
1431: Since the last term in (\ref{e:eden1}), $\rho(x)e^* \Phi(x)$, does not
1432: depend on the phase $\phi(x)$, we may replace $\mathcal{H}_{\text
1433:   s}(x)$ in the last term in (\ref{e:h2'}) by $T_{\text s}^{00}(x)$.
1434: Integrating the resulting equation over the superfluid, discarding a
1435: boundary term, and defining a ``global'' derivative with respect to
1436: the phase,
1437: \begin{displaymath}
1438:   \frac{\partial F[\phi(x)]}{\partial\phi}\equiv
1439: \lim_{\Delta\phi\rightarrow 0}\frac{F[\phi(x)+\Delta\phi]}{\Delta\phi},
1440: \end{displaymath}
1441: where $F[\phi(x)]$ is an arbitrary functional of $\phi(x)$
1442: and $\Delta\phi$ an infinitesimal independent of spacetime, yields
1443: \begin{displaymath}
1444: -\hbar\:\!\partial_t \int\!d^{3\,}\!{\bs x}\,\rho(x)
1445: =-\frac{\partial}{\partial\phi}\,\int\!\! 
1446: d^{3\,}\!{\bs x}\:T_{\text s}^{00}(x)\,
1447: \end{displaymath}
1448: or
1449: \begin{equation}
1450: \label{e:h2''}
1451: \hbar\:\!\partial_t N=\frac{\partial E_{\text s}}{\partial\phi}\:\!,
1452: \end{equation}
1453: where $N$ is the number of particles (or pairs) and $E_{\text s}$ the
1454: energy of the superfluid.  This is the other ``characteristic
1455: equation'' of superfluidity~\cite{phil66}.
1456: 
1457: This concludes our derivation or motivation of the fundamental
1458: equations of superfluidity in the limit of low temperatures and low
1459: energies.  Before turning to the phenomenology these equations imply,
1460: I would like to digress briefly and justify one of the implicit
1461: assumptions made above.  The assumption is that we can describe the
1462: macroscopic quantum phenomena of superfluidity with a classical
1463: effective field theory, or in other words, that we may consider both
1464: the phase $\phi(x)$ and its conjugate field $\pi(x)=-\hbar\rho(x)$ as
1465: thermodynamic variables.  To justify this assumption, let us
1466: canonically quantize the theory by imposing
1467: \begin{equation}
1468:   \label{e:canq}
1469:   \bigc{\hat\phi(\bs{x},t)}{\hat\pi(\bs{y},t)}=i\hbar\delta(\bs{x}-\bs{y}).
1470: \end{equation}
1471: Integration of $\bs{y}$ over the superfluid yields
1472: \begin{displaymath}
1473:   \label{e:canq2}
1474:   \bigc{\hat\phi(\bs{x})}{\hat N}=-i,
1475: \end{displaymath}
1476: which in turn implies the uncertainty relation
1477: \begin{displaymath}
1478:   \Delta\phi(\bs{x})\;\!\Delta N\ge\,\frac{1}{2}.
1479: \end{displaymath}
1480: If we assume that the number of particles in the superfluid takes on a
1481: macroscopic value of order $N\approx 10^{20}$, a $\Delta N$ of the
1482: order of $\sqrt{N}$ implies a relative uncertainty of order $10^{-10}$
1483: in the particle number and the phase.  These numbers are comparable to
1484: the position and momentum uncertainties of a macroscopic object.  The
1485: description of a macroscopic superfluid in terms of a classical field
1486: theory is therefore as appropriate as the classical description of any
1487: other macroscopic object.  This is of course not in contradiction with
1488: the fact that Planck's constant $\hbar$ appears in this effective
1489: field theory.  We will see below that it manifests itself in a family
1490: of ``quantum effects'', which are related to the compactness of the
1491: U(1) field $\phi(x)$.  These effects require either a non-trivial
1492: topology or more than one superfluid, and are very similar for neutral
1493: and for charged superfluids.
1494: 
1495: %\section{PHENOMENOLOGY OF SUPERFLUIDITY AND THE HIGGS MECHANISM}
1496: %\section{Phenomenology of superfluidity and the Higgs mechanism}
1497: \section{PHENOMENOLOGY AND THE HIGGS MECHANISM}
1498: %\section{Phenomenology and the Higgs mechanism}
1499: \label{sec:phe}
1500: 
1501: To begin with, however, let us consider the superfluid flow in a
1502: simply connected superfluid.  The phenomenology depends strikingly on
1503: whether the fluid is charged or not.  For a neutral superfluid,
1504: $e^*=0$, and the gauge field decouples completely.  Even for a fixed
1505: set of boundary conditions, we have an infinite set of solutions for
1506: the superfluid flow, corresponding via
1507: \begin{displaymath}
1508: m^*\bs{v}_{\text s}(x)=\hbar\nabla\phi(x)  
1509: \end{displaymath}
1510: to all possible choices of the phase field $\phi(x)$.  In a simply
1511: connected superfluid, the flow will be vortex-free, \ie
1512: \begin{displaymath}
1513:   \nabla\times\bs{v}_{\text s}(x)=0,
1514: \end{displaymath}
1515: and subject to boundary constraints, but apart from this, it only has
1516: to satisfy the continuity equation as an equation of motion.  
1517: 
1518: 
1519: The simplest example of a multiply connected superfluid is a
1520: superfluid with a line defect, or vortex, along which the magnitude
1521: $|\Psi(x)|$ of the superfluid order parameter vanishes.  The
1522: phase $\phi(x)$ still has to be single valued everywhere in the fluid,
1523: but being a phase, its value may change by a multiple of $2\pi$ as we
1524: circumvent the line defect along a closed curve $\partial S$:
1525: \begin{equation}
1526:   \label{e:vortex}
1527:   \oint_{\partial S}\!\nabla\phi(x)d\bs{l}=2\pi n
1528: \end{equation}
1529: where $n$ is an integer.  The angular momentum of each superfluid
1530: ``particle'' around the vortex is hence quantized in units of
1531: $\hbar$.  With Stokes theorem and the definition
1532: \begin{equation}
1533:   \label{e:vorti}
1534:   \bs{\omega}(x)=\nabla\times\bs{v}_{\text s}(x),
1535: \end{equation}
1536: we may express this alternatively as quantization condition for the 
1537: vorticity
1538: \begin{equation}
1539:   \label{e:vortiq}
1540:   \int_{S}\bs{\omega}(x)\!\cdot\!\bs{n} da =\frac{2\pi\hbar\;\!n}{2m^*},
1541: \end{equation}
1542: where $\bs{n}$ is a unit vector normal to the surface and the area
1543: integral extends over any open surface $S$ which is pierced by the
1544: vortex once.  The quantization of vortices in a superfluid is the
1545: simplest of the ``quantum effects'' alluded to above, where Planck's
1546: constant $\hbar$ enters in the phenomenology through the compactness
1547: of the field $\phi(x)$.  (If $\phi(x)$ was not compact, we could
1548: eliminate $\hbar$ completely from the effective theory by rescaling
1549: $\phi(x)\rightarrow \hbar\phi(x)$.)
1550: 
1551: 
1552: Let us now turn to the phenomenology of a simply connected charged
1553: superfluid or superconductor, which displays the Higgs mechanism.  The
1554: essence of the mechanism is that the phase field $\phi(x)$ looses its
1555: independent significance in the presence of the gauge field.  
1556: There are two ways of seeing this.  The first is on the level of the
1557: equations of motion.  We can simply choose a gauge such that
1558: $\phi(x)=0$ everywhere in the fluid; any other choice of gauge can be
1559: brought into this gauge via %a gauge transformation 
1560: (\ref{e:gauge}) with 
1561: \begin{displaymath}
1562:   \Lambda(x)=\frac{\hbar c}{e^*}\phi(x).
1563: \end{displaymath}
1564: The second is on the level of the effective Lagrangian.  We may
1565: introduce a new vector field
1566: \begin{equation}
1567:   \label{e:newfields}
1568:   -\frac{e^*}{c}A_\mu'\equiv
1569:   D_\mu\phi=\hbar\:\! \partial_\mu \phi-\frac{e^*}{c}A_\mu\:\!. 
1570: \end{equation}
1571: In terms of this field, the superfluid Lagrange density (\ref{e:ls3})
1572: looks the same except that all terms containing $\phi$ have
1573: disappeared.  In particular, the terms quadratic in the derivatives of
1574: $\phi$ in $\mathcal{L}_{\text s}$ have turned into a mass term
1575: \begin{equation}
1576:   \label{e:mass}
1577:   \frac{n_{\text{s}}}{2m^*}\biggl\{\frac{1}{v^2}
1578:   \left(e^*A_0'\right)^2-
1579:   \Big(\frac{e^*}{c}\bs{A}'\Bigl)^2\biggr\}
1580: \end{equation}
1581: for the vector field.  The Maxwell Lagrange density and the Lagrange
1582: density for the uniform neutralizing background charge take the same
1583: form with $F_{\mu\nu}$ and $A_\mu$ replaced by $F_{\mu\nu}'$ and
1584: $A_\mu'$, respectively, except for a total derivative or boundary term
1585: we discard.  Thus the massless gauge field $A_\mu$ is replaced by a
1586: massive vector field $A_\mu'$, while the Goldstone boson $\phi$
1587: disappeared.  The total number of degrees of freedom, however, is
1588: preserved: before, the massless vector field has two (the two helicity
1589: states of the photon) and the Goldstone boson one, while the massive
1590: vector field after the change of variables has three degrees of
1591: freedom.  In Sidney Coleman's words, ``the vector field has eaten the
1592: Goldstone bosons and grown heavy''~\cite{sid}.  We will return to this
1593: issue after studying the phenomenology of the superconductor using the
1594: equations of motion.
1595: 
1596: The Euler--Lagrange equation for $A_\mu$,
1597: \begin{equation}
1598:   \label{e:eulag}\nonumber
1599:   \partial_\mu \frac{\delta L}{\delta(\partial_\mu A_\nu)}
1600:   -\frac{\delta L}{\delta A_\nu}=0,
1601: \end{equation}
1602: yields Maxwell's electrodynamics with electric charge density
1603: $-e^*(\rho-n_{\text s})$ and current $-e^*\bs{J}$,
1604: \begin{eqnarray}
1605:   \nabla\cdot\bs{E}\!&\!=\!&\!-4\pi e^*(\rho-n_{\text s}),
1606:   \label{e:max1} \\ \rule{0pt}{20pt} \label{e:max2}
1607:   \nabla\times\bs{B}-\frac{1}{c}\partial_t\bs{E}
1608:   \!&\!=\!&\!-\frac{4\pi e^*}{c} \bs{J},
1609: \end{eqnarray}
1610: where 
1611: \begin{equation}
1612:   \label{e:eandb}\nonumber
1613:   \bs{E}=-\nabla\Phi-\frac{1}{c}\partial_t\bs{A},\
1614:   \bs{B}=\nabla\times\bs{A},
1615: \end{equation}
1616: and $\rho$ and $\bs{J}$ are given by (\ref{e:rho3}) and (\ref{e:j3}),
1617: respectively.  In principle, we could also obtain the continuity
1618: equation (\ref{e:cont}) as the Euler--Lagrange equation for $\phi$,
1619: but since $L$ depends on derivatives of $A_\mu$ only through
1620: $F_{\mu\nu}$ and $\phi$ is minimally coupled to $A_\mu$,
1621: (\ref{e:cont}) is automatically satisfied by any solution of
1622: (\ref{e:max1}) and (\ref{e:max2}).  This is consistent with the fact
1623: that $\phi$ has lost its independent significance due to the Higgs
1624: mechanism.
1625: 
1626: For convenience, we choose the gauge $\phi(x)=0$.  Then (\ref{e:rho3}), 
1627: (\ref{e:j3}), and (\ref{e:vs}) imply
1628: \begin{equation}
1629:   4\pi e^*(\rho-n_{\text s})=
1630:   \frac{4\pi{e^*}^2 n_{\text s}}{m^*v^2}
1631:   \Bigl(\Phi-\frac{e^*}{2m^*c^2}\bs{A}^2\Bigl),
1632:   \label{e:rho4} 
1633: \end{equation}
1634: \vspace{-2pt}
1635: \begin{equation}
1636:   \label{e:j4}
1637:   \frac{4\pi e^*}{c}\bs{J}  =
1638:   \frac{4\pi{e^*}^2 n_{\text s}}{m^*c^2}\bs{A}
1639:   \biggl\{1+\frac{e^*}{m^*v^2}
1640:   \Bigl(\Phi-\frac{e^*}{2m^*c^2}\bs{A}^2\Bigl)\biggr\}.
1641: \end{equation}
1642: Let us now restrict our attention to quasistatic phenomena, where we
1643: can neglect the time derivative terms.  The analysis given below implies
1644: that this assumption holds for frequencies significantly smaller than
1645: $c/\lambda_{\text L}$, where 
1646: \begin{equation}
1647: \lambda_{\text L} = \sqrt{\frac{m^*c^2}{4\pi{e^*}^2 n_{\text s}}} 
1648: %\frac{1}{\lambda_{\text L}^2}=\frac{4\pi{e^*}^2 n_{\text s}}{m^*c^2} 
1649: \end{equation}
1650: is the London penetration depth.
1651: Then %the equations 
1652: (\ref{e:max1})--(\ref{e:j4}) reduce to
1653: \begin{eqnarray}
1654:   \nabla^2\Phi \!&\!=\!&\! 
1655:   \frac{c^2}{v^2}\frac{1}{\lambda_{\text L}^{\;2}}
1656:   \Bigl(\Phi-\frac{{e^*}}{2m^*c^2}\bs{A}^2\Bigl),
1657:   \label{e:eom1} \\ \rule{0pt}{20pt} \label{e:eom2}
1658:   \nabla^2\bs{A}\!-\!\nabla(\nabla\bs{A})\!&\!=\!&\! 
1659:   \frac{1}{\lambda_{\text L}^{\;2}}
1660:   \bs{A}\biggl\{1\!+\!\frac{e^*}{m^*v^2}
1661:   \Bigl(\Phi\!-\!\frac{{e^*}}{2m^*c^2}\bs{A}^2\Bigl)\!\biggr\}.\qquad
1662: \end{eqnarray}
1663: Let us first look at the linear terms in these equations, \ie the
1664: solution for infinitesimal $\Phi$ and $\bs{A}$.  Under quasistatic
1665: conditions, (\ref{e:max2}) implies $\nabla\bs{J}=0$ and with
1666: (\ref{e:j4}) for infinitesimal fields $\nabla\bs{A}=0$.  The equations
1667: reduce to
1668: \begin{eqnarray}
1669:   \nabla^2\Phi \!&\!=\!&\! 
1670:   \frac{c^2}{v^2}\frac{1}{\lambda_{\text L}^{\;2}}\Phi,
1671:   \label{e:eomi1}\\ \rule{0pt}{20pt} \label{e:eomi2}
1672:   \nabla^2\bs{A}\!&\!=\!&\! 
1673:   \frac{1}{\lambda_{\text L}^{\;2}}
1674:   \bs{A},
1675: \end{eqnarray}
1676: \ie we have electric screening in addition to magnetic screening, but
1677: with a screening length reduced by a factor $v/c$.  This leads us to
1678: conjecture that the dominant energy is the Coulomb interaction, which
1679: effects charge neutrality or $\rho(x)\approx n_{\text s}$.  We now
1680: simply assume that this is a valid approximation, and justify it {\sl
1681:   a posteriori}.  Then (\ref{e:rho4}) implies
1682: \begin{equation}
1683:   \label{e:hall}
1684:    \Phi-\frac{{e^*}}{2m^*c^2}\bs{A}^2=0,
1685: \end{equation}
1686: and (\ref{e:j4}) reduces to
1687: \begin{equation}
1688:   \label{e:j5}
1689:   \bs{J}=\frac{e^* n_{\text s}}{m^*c}\bs{A}.
1690: \end{equation}
1691: Taking the curl of this equation, we obtain London's
1692: equation~\cite{london,london1}
1693: \begin{equation}
1694:   \label{e:lon2}
1695:   \nabla\times\bs{J}=\frac{e^* n_{\text s}}{m^*c}\bs{B}.
1696: \end{equation}
1697: Under quasistatic conditions, we have again $\nabla\bs{J}=0$ and with
1698: (\ref{e:j5}) $\nabla\bs{A}=0$, which implies that (\ref{e:eom2})
1699: reduces to (\ref{e:eomi2}).  The solution of (\ref{e:eomi2}) describes
1700: exponential screening with penetration depth $\lambda_{\text L}$.  If
1701: we have, for example, a superconductor which occupies the half-space
1702: $x>0$ subject to an external magnetic field $\bs{B}=B_0\bs{\hat y}$ at
1703: the boundary $x=0$, we obtain
1704: \begin{equation}
1705:   \label{e:sol}
1706:   \bs{A}=A_0\;\! e^{-x/\lambda_{\text L}}\bs{\hat z},\ 
1707:   \bs{B}=B_0\;\! e^{-x/\lambda_{\text L}}\bs{\hat y},\
1708: %  \ \text{and}\
1709:   \bs{J}=J_0\;\! e^{-x/\lambda_{\text L}}\bs{\hat z},
1710: \end{equation}
1711: where
1712: \begin{displaymath}
1713:   A_0=\lambda_{\text L} B_0,\    
1714:   J_0=\sqrt{\frac{n_{\text s}}{4\pi m^*}}\;\! B_0.
1715: \end{displaymath}
1716: The screening of the magnetic field is known as the Meissner effect.
1717: According to (\ref{e:hall}), the vector potential implies
1718: an electrostatic potential 
1719: \begin{equation}
1720:   \label{e:hall2}
1721:    \Phi%(x)
1722:    =\frac{{B_0}^2}{8\pi e^* n_{\text s}}\;\! e^{-2x/\lambda_{\text L}}.
1723: \end{equation}
1724: This potential allows us to verify the validity of our approximation
1725: $\rho(x)\approx n_{\text s}$.  Substituting (\ref{e:hall2}) into
1726: (\ref{e:eomi1}), we find that the ratio of the neglected term to the
1727: terms kept is
1728: \begin{equation}
1729:   \label{e:approx}
1730:   \frac{\nabla^2\Phi}{\frac{c^2}{v^2}\frac{1}{\lambda_{\text L}^{\;2}}\Phi}
1731:   =\frac{4 v^2}{c^2}%\approx 10^{-6}, 
1732:   \ll 1,
1733: \end{equation}
1734: \ie the approximation is excellent.
1735: 
1736: The electrostatic potential (\ref{e:hall2}) is called the London or
1737: Bernoulli Hall effect~\cite{london,gww}.  To understand its physical
1738: origin, it is best to rewrite (\ref{e:hall}) with (\ref{e:vs}) for
1739: $\phi(x)=0$ in terms of the superfluid velocity:
1740: \begin{equation}
1741:   \label{e:hall4}
1742:   -e^*\Phi+\frac{1}{2}m^*{\bs{v}_{\text s}}^2=0.
1743: \end{equation}
1744: The electrostatic potential simply compensates the kinetic energy
1745: contribution to the chemical potential, as required by (\ref{e:h1'})
1746: with $\phi(x)=0$.  For stationary flow, this condition reduces to the
1747: requirement that the electrochemical potential
1748: $\mu_{\scriptscriptstyle\text{el.chem.}}$ is constant across the
1749: superconductor.  In practice, the London Hall effect can only be
1750: measured with capacitive contacts, as ohmic contacts are sensitive to
1751: the electrochemical rather than the electrostatic potential~\cite{bok}.
1752: The effect furnishes us with an independent meaning of the superfluid
1753: density $n_{\text s}$ or the effective mass $m^*$, while under
1754: quasistatic conditions all other effects~\cite{rot} depend only on the
1755: superfluid stiffness $n_{\text s}/m^*$.  The underlying theoretical
1756: reason is that the London Hall effect is a consequence of the
1757: corrections incorporated in the effective Lagrange density
1758: (\ref{e:ls3}).  In these terms, the parameter $m^*$ enters by itself,
1759: while (apart from a total time derivative term irrelevant to the
1760: equations of motion) only the combination $n_{\text s}/m^*$ entered in
1761: the previous approximative Lagrangian (\ref{e:ltot}) with
1762: (\ref{e:ls}), (\ref{e:lb}) and (\ref{e:lmax}).
1763: 
1764: Since we have given a precise definition of the superfluid velocity,
1765: it is legitimate to ask whether the London Hall effect (\ref{e:hall4})
1766: balances the Lorentz force, or, if not, what other forces balance it.
1767: The total electromagnetic force on a given
1768: particle with charge $-e^*$ in the fluid is given by
1769: \begin{eqnarray}
1770:    \bs{F}_{\text{em}}\!&\!=\!&\! 
1771:    -e^*\bigl(\bs{E}
1772:    +\frac{\bs{v}_{\text s}}{c}\times\bs{B}\bigr)
1773:    \label{e:lorentz} \\ \rule{0pt}{20pt} \nonumber
1774:    \!&\!=\!&\!
1775:    -e^*\Bigl(-\frac{1}{c}\partial_t\bs{A}-\nabla\Phi
1776:    +\frac{\bs{v}_{\text s}}{c}\times (\nabla\times\bs{A})\Bigr).
1777: \end{eqnarray}
1778: With (\ref{e:hall4}) and (\ref{e:vs}) we obtain 
1779: \begin{eqnarray}
1780:   \bs{F}_{\text{em}}\!&\!=\!&\! 
1781:   m^*\Bigl(\partial_t\bs{v}_{\text s}
1782:   +\frac{1}{2}\nabla (\bs{v}_{\text s})^2
1783:   -\bs{v}_{\text s}\times (\nabla\times\bs{v}_{\text s})\Bigr)
1784:   \label{e:lorentz2}\nonumber \\ \rule{0pt}{20pt} \nonumber
1785:   \!&\!=\!&\!
1786:   m^*\bigl(\partial_t\bs{v}_{\text s}
1787:   +(\bs{v}_{\text s}\nabla )\bs{v}_{\text s}\bigr)
1788:   =m^*\frac{d\bs{v}_{\text s}}{dt}.
1789: \end{eqnarray}
1790: Thus the gradient $\nabla\Phi$ of the electrostatic potential
1791: (\ref{e:hall4}) does not balance the Lorentz force
1792: $\frac{1}{c}\bs{v}_{\text s}\times\bs{B}$, but both terms together
1793: account for the difference between local and substantial acceleration
1794: (\ref{e:force}) in the superfluid.  This difference is significant
1795: when, for example, the flow is stationary but does not follow a
1796: straight line.
1797: 
1798: Let us summarize how the Higgs mechanism manifests itself in the
1799: equations of motion.  For a neutral superfluid with $e^*=0$, there are
1800: many solutions to
1801: \begin{displaymath}
1802:   \bs{J}(x)=\rho\bs{v}_{\text s}
1803:   =\frac{\rho}{m^*}\Bigl(\hbar\nabla\phi+\frac{e^*}{c}\bs{A}\Bigr)
1804: \end{displaymath}
1805: for fixed boundary conditions, corresponding to all possible
1806: configurations for the phase field $\phi(x)$.  These solutions reflect
1807: the existence of the massless sound mode described by the field
1808: $\phi(x)$.  For the charged superfluid, however, all the different
1809: configurations for $\phi(x)$ merely correspond to different choices of
1810: gauge; as far as the current or magnetic field distributions are
1811: concerned, all these solutions are always equivalent to one for which
1812: we have $\phi(x)=0$.  Thus the field $\phi$ does no longer describe an
1813: excitation.  For a simply connected superconductor, it is only
1814: meaningful in that it assures gauge invariance, both on the level of
1815: the Lagrange density and on the level of the equations of motion.  The
1816: solution of these equations is physically (\ie apart from the freedom
1817: to choose the gauge) unique for a given set of boundary conditions.
1818: 
1819: We now return to the manifestation of the Higgs mechanism on the level
1820: of the Lagrangian.  For a simply connected superconductor, we have
1821: already seen that we may eliminate the phase field $\phi$ if we
1822: introduce a new vector field $A_\mu'$ according to
1823: (\ref{e:newfields}).  The mass term (\ref{e:mass}) we find for
1824: $A_\mu'$ may appear to violate gauge invariance, as mass terms
1825: generally do, and may, at first sight, to be taken as a signature of a
1826: spontaneously broken gauge invariance.  For one thing, however, gauge
1827: invariance is not violated.  From the definition (\ref{e:newfields})
1828: it is clear that the new field simply transforms as
1829: \begin{displaymath}
1830:   A_\mu'(x)\rightarrow A_\mu'(x)
1831: \end{displaymath}
1832: under a gauge transformation (\ref{e:gauge}).  The Lagrangian hence
1833: remains manifestly gauge invariant, and has to remain gauge invariant,
1834: as it is the same Lagrangian as before expressed in terms of different
1835: fields.  Furthermore, if a symmetry is spontaneously broken, it is never
1836: violated on the level of the Lagrangian or the Hamiltonian, but only
1837: on the level of the ground state.
1838: 
1839: In the literature, one sometimes finds the statement that ``the gauge
1840: field acquires a mass'' due to the Higgs mechanism.  This is not
1841: exactly to the point, as it suggests that the massive vector field
1842: $A_\mu'$ is still a gauge field, while we have just seen that it is
1843: gauge invariant.  In the case of a superconductor, we even know how to
1844: interpret the individual components of $A_\mu'$ physically.
1845: According to (\ref{e:newfields}), (\ref{e:mu}), and (\ref{e:vs}),
1846: \begin{equation}
1847:   \label{eq:a'}
1848:   -\frac{e^*}{c}\left(A_\mu'\right)
1849:   =\bigl(-\frac{\mu}{c},m^*\bs{v}_{\text s}\bigr).
1850: \end{equation}
1851: The Higgs mechanism hence does not imply that ``the electromagnetic
1852: gauge field acquires a mass'', but only that we can describe the
1853: superconductor in terms of gauge invariant fields, that is, in
1854: terms of the chemical potential $\mu(x)$ and the superfluid velocity
1855: $\bs{v}_{\text s}(x)$.  If we do this, we also have to express the
1856: Maxwell Lagrange density (\ref{e:lmax}) in terms of $\mu$ and
1857: $\bs{v}_{\text s}$.  With
1858: \begin{displaymath}
1859:   -\frac{1}{2}F_{\mu\nu}F^{\mu\nu}=\bs{E}^2-\bs{B}^2
1860: \end{displaymath}
1861: we obtain for the total Lagrange density 
1862: \begin{eqnarray}
1863:   \mathcal{L}%_{\text{tot}}
1864:   \!&\!=\!&\!
1865:   \frac{1}{8\pi {e^*}^2}
1866:   \Bigl\{(\nabla\mu+m^*\partial_t\bs{v}_{\text s})^2-
1867:   c^2{m^*}^2 (\nabla\times\bs{v}_{\text s})^2\Bigr\}
1868:   \nonumber \\ \nonumber
1869:    \rule{0pt}{20pt}
1870:   &&\hspace{-7mm} -n_{\text{s}}\biggl\{
1871:   \mu+\Bigl(-\mu+\frac{1}{2}m^*\bs{v}_{\text s}^2\Bigr)
1872:   +\frac{1}{2m^*}\frac{1}{v^2}
1873:   \Bigl(-\mu+\frac{1}{2}m^*\bs{v}_{\text s}^2\Bigr)^2\biggr\}.\\
1874:   \label{e:l2}
1875: \end{eqnarray}
1876: The Euler--Lagrange equations we obtain from (\ref{e:l2}) for $\mu$
1877: and $\bs{v}_{\text s}$ are equivalent to (\ref{e:max1})--(\ref{e:j4}),
1878: and yield exactly the same solution as above.
1879: %
1880: Writing the Lagrangian in terms of $\mu$ and $\bs{v}_{\text s}$ does
1881: not yield any practical advantage, but clearly illustrates that gauge
1882: invariance has become irrelevant---it is not broken, but has simply
1883: left the stage.  Since all the fields are gauge invariant,
1884: (\ref{e:l2}) does not even provide a framework to think of a spontaneous
1885: violation of a gauge invariance.
1886: 
1887: These considerations apply to every field theory which displays the
1888: Higgs mechanism.  In any such theory, the Lagrange density is invariant
1889: under a global physical symmetry for a matter field, and invariant
1890: under a local gauge symmetry, which affects both the matter field and
1891: the gauge field.  The global symmetry is ``physical'' as we can
1892: classify the states of matter according to their transformation
1893: properties, while the gauge symmetry is ``unphysical'' as gauge
1894: transformations have no effect on the states of matter, but only on
1895: our description of these states.  In our example of a superfluid,
1896: charged or neutral, the global symmetry transformation is 
1897: \begin{equation}
1898:   \label{e:global}
1899:   \phi(x)\rightarrow\phi(x)+\lambda,
1900: \end{equation}
1901: where $\lambda$ is independent of spacetime.  This symmetry is
1902: spontaneously violated, which means that there are many degenerate
1903: ground states which map into each other under (\ref{e:global}).  For a
1904: neutral superfluid, we obtain a massless mode according to Goldstone's
1905: theorem.  The situation is more subtle for a superconductor, as the
1906: matter field is coupled to a gauge field and the Lagrange density is
1907: also invariant under the gauge transformation (\ref{e:gauge}).  This
1908: ``unphysical'' symmetry, however, seems to contain the physical
1909: symmetry as the special case
1910: \begin{equation}
1911:   \label{e:globalgauge}
1912:   \Lambda(x)=-\frac{\hbar c}{e^*}\lambda.  
1913: \end{equation}
1914: {The formal equivalence of the transformation (\ref{e:global}) and
1915: (\ref{e:gauge}) with (\ref{e:globalgauge}) is 
1916: at the root of the widely established but incorrect interpretation of
1917: (\ref{e:global}) as a gauge transformation, and in particular of the
1918: spontaneous violation of (\ref{e:global}) as a spontaneous violation
1919: of a gauge symmetry.}  (This is presumably the reason why particle
1920: physicists like Steven Weinberg speak of ``spontaneously broken gauge
1921: symmetries'' interchangeably with ``the Higgs mechanism''.)  The
1922: problem here is that the equivalence is only formal.  The gauge
1923: transformation (\ref{e:globalgauge}) represents a transformation of
1924: our description, similar to a rotation of a coordinate system we use
1925: to describe a physical state, while the transformation
1926: (\ref{e:global}) corresponds to a transformation of our physical
1927: state, like a rotation of a physical system.  Clearly, a
1928: (counterclockwise) rotation of the coordinate system has the same
1929: effect on our equations as a (clockwise) rotation of the physical
1930: system we describe with these equations, but the transformations are
1931: all but equivalent.  It is hence incorrect to refer to the spontaneous
1932: violation of (\ref{e:global}) as a spontaneous violation of gauge
1933: symmetry.  A gauge symmetry cannot be spontaneously violated as a
1934: matter of principle.  
1935: 
1936: The difference between the ``physical'' symmetry (\ref{e:global}) and
1937: the gauge symmetry (\ref{e:gauge}) can also be appreciated at the
1938: level of conservation laws.  The former yields particle number (or
1939: charge) as a conserved quantity, according to (\ref{e:cont}), while
1940: there is no conservation law associated with the latter.  In the
1941: literature, (\ref{e:global}) is often referred to as a global gauge
1942: transformation, and the conservation of charge attributed to gauge
1943: invariance.  This view, however, is not consistent.  If one speaks of
1944: a global gauge symmetry, this symmetry has to be a proper subgroup of
1945: the local gauge symmetry group.  The alleged global gauge symmetry
1946: hence cannot be a ``physical'' symmetry while the local gauge symmetry
1947: is an invariance of description, or be spontaneously violated while
1948: the local symmetry is fully intact.  The difference between the global
1949: phase rotation (\ref{e:global}) and a global gauge rotation
1950: (\ref{e:globalgauge}) is even more at evident at the level of quantum
1951: states.  The BCS ground state (\ref{e:bcs}) is, for example, not
1952: invariant under (\ref{e:global}), while it is fully gauge invariant,
1953: as we have seen in Section \ref{sec:gau}
1954: 
1955: The conclusions regarding the physical significance (or maybe better
1956: insignificance) of gauge transformations we reached here for
1957: superconductors hold for any field theory which displays the Higgs
1958: mechanism.
1959: 
1960: \section{QUANTUM EFFECTS}
1961: %\section{Quantum effects}
1962: \label{sec:qua}
1963: 
1964: This discussion of the Higgs mechanism applies only to simply
1965: connected superconductors.  If we have a nontrivial topology or more
1966: than one superfluid, the phase field $\phi$ reassumes physical
1967: significance through its compactness, that is, the fact that its value
1968: is only defined modulo $2\pi$.  In these situations, we are not
1969: allowed to set $\phi(x)=0$ in the equations of motion or eliminate it
1970: from the effective Lagrange density via (\ref{e:newfields}) or
1971: (\ref{eq:a'}), as we would loose the information regarding the
1972: compactness.  Since the phase field $\phi$ is multiplied by Planck's
1973: constant $\hbar$ whenever it enters in the Lagrange density, any
1974: effect due to the compactness of $\phi$ will depend on $\hbar$, and
1975: only exist for $\hbar\ne 0$.  Therefore we refer to them as ``quantum
1976: effects''.
1977: 
1978: The simplest of these effects in superconductors is the quantization
1979: of magnetic flux, which is analogous to the quantization of vorticity
1980: in neutral superfluids.  The effect was predicted by London in a
1981: footnote in his first book~\cite{london,london2} almost a decade
1982: before BCS proposed their microscopic theory.  Consider a macroscopic
1983: superconductor with a hole in it, which may either be a hole in the
1984: superconducting material or a line defect or vortex in the
1985: superconducting order parameter.  Like in the case of a vortex in a neutral
1986: superfluid, the phase field $\phi$ has to be single valued everywhere 
1987: in the superconductor, but may change by a multiple of $2\pi$ as we
1988: circumvent the hole or defect along a closed curve $\partial S$:
1989: \begin{equation}
1990:   \label{e:vortex2}
1991:   \oint_{\partial S}\!\nabla\phi(x)d\bs{l}=2\pi n,
1992: \end{equation}
1993: where $n$ is an integer.  We now take $\partial S$ well inside the
1994: superconductor, that is, separated at each point by a distance much
1995: larger than the penetration depth $\lambda_{\text L}$ from the hole or
1996: defect.  Then, according to the Meissner effect or our derivation of
1997: London's equation above, which still applies locally, the superfluid
1998: velocity
1999:  \begin{displaymath}
2000:   \bs{v}_{\text s}
2001:   =\frac{1}{m^*}\Bigl(\hbar\nabla\phi+\frac{e^*}{c}\bs{A}\Bigr)
2002: \end{displaymath}
2003: has to vanish along $\partial S$, and (\ref{e:vortex2}) implies 
2004: \begin{equation}
2005:   \label{e:vortex3}
2006:   \oint_{\partial S}\!\bs{A}(x)d\bs{l}
2007:   =\int_{S}\!\bs{B}(x)\!\cdot\!\bs{n} da
2008: %  =\frac{2\pi\hbar c}{e^*} n,
2009:   =\frac{h c}{e^*}\cdot n,
2010: \end{equation}
2011: where we have used Stokes theorem once more.  The magnetic flux through the
2012: hole or vortex is hence quantized in units of $hc/e^*$, which for
2013: $e^*=2e$ is half of the Dirac flux quantum.  Note that the vorticity 
2014: (\ref{e:vorti}) is not quantized in a superconductor.
2015: 
2016: 
2017: \begin{figure} %[tb]
2018:   \begin{center}
2019:     \vspace{2mm}
2020:     \psfrag{vortex}{vortex}
2021:     \psfrag{(a)}{(a)}
2022:     \psfrag{(b)}{(b)}
2023:     \includegraphics[width=\columnwidth]{slip.eps}
2024:   \end{center}
2025:   \caption{Phase slippage: A vortex moving in a superfluid induces a 
2026:     transverse gradient in the (electro) chemical potential by
2027:     dragging a branch cut in the phase of the order parameter with
2028:     it.}
2029:   \label{f:slip}
2030: \end{figure}
2031: We now review two further quantum effects, which are similar in
2032: neutral and charged superfluids; as in our derivation of the effective
2033: theory above, the equations for the latter case contain the former as
2034: the special case $e^*=0$.  One of the effects is phase
2035: slippage~\cite{phil66}.  Consider two points $1$ and $2$ in a
2036: superfluid, which are connected by a vertical path (see
2037: \mbox{Fig.~\ref{f:slip}a).}  Now imagine we adiabatically move a vortex
2038: from the very far left across the path to the very far right.  This
2039: process yields a difference in the electrochemical potential between
2040: the two points, which is according to (\ref{e:muelch}) given by
2041: \begin{equation}
2042:   \label{e:deltamu}
2043:   \Delta\mu_{\scriptscriptstyle\text{el.chem.}}=
2044:   -\partial_t \int_{1}^{2}\! 
2045:   \Bigl(\hbar\nabla\phi(x)+\frac{e^*}{c}\bs{A}(x)\Bigr) d\bs{l}.
2046: \end{equation}
2047: where the line integral is taken along the path between the points.
2048: The time integrated difference in the electrochemical potential is
2049: hence given by the difference between the line integral at the end of
2050: the process and the line integral at the beginning. %,
2051: Let us first consider the case of a neutral superfluid, \ie $e^*=0$.
2052: The line integral of $\nabla\phi$ will have changed by $2\pi$, as the
2053: difference in the paths is topologically equivalent to encircling the
2054: vortex once (see \mbox{Fig.~\ref{f:slip}b)}. 
2055: Alternatively, we may say the vortex has dragged a branch cut
2056: across the path.  (We assume that at the beginning and the end,
2057: the vortex is so far away from points $1$ and $2$ that 
2058: we can neglect its influence on the line integral.)
2059: If we now have a continuous flow of vortices across the path, 
2060: the line integral will pick up a contribution of $2\pi\hbar$
2061: from each of them, and we obtain a chemical potential difference
2062: \begin{equation}
2063:   \label{e:phslip}
2064:   \Delta\mu = h\langle\partial_t 
2065:   N_{\text{v}}\rangle_{\scriptscriptstyle\text{av}}, 
2066: \end{equation}
2067: where $\langle\partial_t N_{\text{v}}\rangle_{\scriptscriptstyle\text{av}}$ is
2068: the average rate of vortices crossing the path.
2069: 
2070: Let us now turn to the case of a superconductor, where we assume that
2071: during the entire process the distance between the vortex and either
2072: of the points $1$ and $2$ is much larger than the penetration
2073: depth.  Since the line integral we obtain when encircling a
2074: superconducting vortex along a circle well inside the superconductor
2075: is zero,
2076: \begin{equation}
2077:   \label{e:encircle}
2078:   \oint_{R\gg\lambda_{\text L}}\!\! \, 
2079:   \Bigl(\hbar\nabla\phi(x)+\frac{e^*}{c}\bs{A}(x)\Bigr)d\bs{l}=0,
2080: \end{equation}
2081: we do not obtain a difference in the electrochemical potential as we
2082: move an isolated vortex carrying a magnetic flux quantum across the
2083: path.  So, at first sight, it may appear as there is no phase slippage
2084: effect in superconductors.  The situation just described, however, is
2085: not the general one, as we dragged a unit of magnetic flux with the
2086: vortex from the very far left to the very far right.  This produced a
2087: Hall effect which exactly canceled the phase slippage effect.  If we
2088: consider a situation where we have a large, almost uniform magnetic
2089: field and an Abrikosov vortex lattice or liquid in which the distance
2090: between the vortices is much smaller than the penetration depth, and
2091: we have a flow of vortices across the path, the magnetic field will
2092: remain to a reasonable approximation unaffected by the flow and we
2093: recover (\ref{e:phslip}) for the electrochemical potential difference.
2094: The voltage we measure between the two points is then given by
2095: \begin{equation}
2096:   \label{e:nenrst}
2097:   U=\frac{h}{2e}\langle\partial_t 
2098:   N_{\text{v}}\rangle_{\scriptscriptstyle\text{av}}.
2099: \end{equation}
2100: This voltage is known as the Nernst effect in superconductors.
2101: 
2102: 
2103: The last and possibly most striking quantum effect we review is the
2104: Josephson effect~\cite{jos}.
2105: %
2106: Consider two superfluids or superconductors S$_1$ and S$_2$, which are
2107: weakly coupled, say by a narrow constriction for superfluid particles
2108: or a tunneling barrier for Cooper pairs. % (see Fig.\ \ref{fig:jos}).
2109: The only requirement for the effect is that there is an energy
2110: associated with the weak link, which depends on the (gauge invariant)
2111: phase difference $\Delta\phi$ between two points 2 and 1 in
2112: superfluids S$_2$ and S$_1$:
2113: \begin{equation}
2114:   \label{e:josen}
2115:   E_{\text{junction}}=f(\Delta\phi) 
2116: \end{equation}
2117: with 
2118: \begin{equation}
2119:   \label{e:josph}
2120:   \Delta\phi \equiv
2121:   \phi(2)-\phi(1)+\frac{e^*}{\hbar c}\int_{1}^{2}\! \bs{A}(x) d\bs{l},
2122: \end{equation}
2123: where the line integral is taken along the path the superfluid
2124: particles take when they move from one superfluid to the other.  Note
2125: that $\Delta\phi$ divided by the distance between the points 2 and 1
2126: is just the discrete version of the gradient term
2127: \begin{displaymath}
2128:   \hbar\nabla\phi(x)+\frac{e^*}{c}\bs{A}(x)
2129: \end{displaymath}
2130: we already encountered in the Ginzburg--Landau free energy, where the
2131: magnetic energy (\ref{e:fmag}) was essentially given by its square.
2132: We assume that $E_{\text{junction}}$ is likewise minimal for
2133: $\Delta\phi=0$, which implies that the first term in a Taylor
2134: expansion around this minimum is quadratic in $\Delta\phi$.  In the
2135: case of the junction, however, this term is not sufficient.  Since
2136: $\phi$ is only defined modulo $2\pi$, $f(\Delta\phi)$ has to be a
2137: periodic function of $\Delta\phi$.  Josephson has shown that to a
2138: reasonable approximation, it is given by
2139: \begin{equation}
2140:   \label{e:josf}
2141:   f(\Delta\phi)=-E_0 \cos(\Delta\phi).
2142: \end{equation}
2143: Let us now assume a situation where both macroscopic superfluids are
2144: in a state of equilibrium, but the phases are not necessarily aligned
2145: relative to each other.  Then only the energy stored in the junction
2146: depends on the phases of the superfluids, and the ``characteristic
2147: equation'' (\ref{e:h2''}) becomes for superfluid S$_2$
2148: \begin{equation}
2149:   \label{e:josh2}
2150:   \hbar\:\!\partial_t N_2
2151:   =\frac{\partial E_{\text{junction}}(\Delta\phi)}{\partial\phi(2)}\:\!,
2152: \end{equation}
2153: where $N_2$ is the number of superfluid particles or Cooper pairs in
2154: S$_2$.  (We would also obtain a similar equation for S$_1$, but since we
2155: assume $N_1+N_2=\text{const.}$ and $E_{\text{junction}}$ only depends
2156: on $\phi(2)-\phi(1)$, it does not provide any additional information.)  
2157: The particle current from superfluid S$_1$ to S$_2$ is hence given by 
2158: \begin{equation}
2159:   \label{e:josj}
2160:   J_{1\rightarrow 2}=\frac{1}{\hbar}
2161:   \frac{\partial E_{\text{junction}}(\Delta\phi)}{\partial\phi(2)}
2162:   =\frac{1}{\hbar}E_0 \sin(\Delta\phi).
2163: \end{equation}
2164: On the other hand, since the other ``characteristic
2165: equation'' (\ref{e:h1'}) holds for each superfluid,
2166: \begin{equation}
2167:   \label{e:h1''}
2168:   \hbar \partial_t \bigl(\phi(2)-\phi(1)\bigr)
2169:   =-\bigl(\mu(2)-\mu(1)\bigr)+e^* \bigl(\Phi(2)-\Phi(1)\bigr).
2170: \end{equation}
2171: If we add  
2172: \begin{displaymath}
2173:   \partial_t \frac{e^*}{c}\int_{1}^{2}\! \bs{A}(x)d\bs{l}
2174: \end{displaymath}
2175: to both sides of (\ref{e:h1''}), we obtain
2176: \begin{equation}
2177:   \label{e:josdt}
2178:   \hbar \partial_t \Delta\phi
2179:   =-\bigl(\mu(2)-\mu(1)\bigr)-e^*\int_{1}^{2}\! \bs{E}(x)d\bs{l}
2180:   =-\Delta\mu_{\scriptscriptstyle\text{el.chem.}}, 
2181: \end{equation}
2182: % On the other hand, if we integrate the other ``characteristic
2183: % equation'' (\ref{e:muelch}) across the junction, we obtain
2184: % \begin{equation}
2185: %   \label{e:josdt}
2186: %   \hbar \partial_t \Delta\phi
2187: %   =-\Delta\mu_{\scriptscriptstyle\text{el.chem.}}, 
2188: % \end{equation}
2189: or, if we take $\Delta\mu_{\scriptscriptstyle\text{el.chem.}}$ time
2190: independent,
2191: \begin{displaymath}
2192:   \Delta\phi(t)= 
2193:   -\frac{\Delta\mu_{\scriptscriptstyle\text{el.chem.}}}{\hbar}\cdot t
2194: %   + \text{const.}
2195: %   +\Delta\phi(t=0)
2196:    +\Delta\phi_0
2197: \end{displaymath}
2198: Substitution into (\ref{e:josj}) yields
2199: \begin{equation}
2200:   \label{e:josj'}
2201:   J_{1\rightarrow 2}=\frac{E_0}{\hbar}
2202: %  \sin(2\pi\nu t+\text{const.}),
2203: %  \sin\bigl(2\pi\nu t+\Delta\phi(t=0)\bigr),
2204:   \sin\bigl(2\pi\nu t+\Delta\phi_0\bigr),
2205: \end{equation}
2206: where 
2207: \begin{equation}
2208:   \label{e:josfreq}
2209:   \nu\equiv-\frac{\Delta\mu_{\scriptscriptstyle\text{el.chem.}}}{h}
2210: \end{equation}
2211: is the Josephson frequency.  This implies that if the electrochemical
2212: potential is equal for both superfluids, we find a DC particle current
2213: depending on the initial alignment of the phases.  If there is a
2214: difference in the potential, however, the current will oscillate with
2215: frequency $\nu$.  This is called the AC Josephson effect.  The effect
2216: exists for both neutral and charged superfluids, but it is much easier
2217: to measure in a superconductor, as we can realize a difference in the
2218: electrochemical potential by applying a voltage $U$ across the
2219: junction, $\Delta\mu_{\scriptscriptstyle\text{el.chem.}}=-2eU$, and
2220: easily measure oscillations in the electrical current.
2221: 
2222: Note that the Josephson effect, so astonishing its phenomenology may
2223: be, follows through the ``characteristic'' equations of superfluidity
2224: directly from the fact that there is a broken symmetry in
2225: superfluids and that the compact phase field which labels the
2226: different degenerate ground states is the field conjugate to the
2227: density in the superfluid.  
2228: % The other assumptions we made in this article, that the order
2229: % parameter carries charge in the case of a superconductor, that there
2230: % exists a linearly dispersing sound mode in neutral superfluids, which
2231: % disappears due to the Higgs mechanism in charged superfluids, and that
2232: % both current and momentum are carried by the same species of
2233: % particles, was not required to explain any of the quantum effects.
2234: The other assumption we made in this article, the assumption that both
2235: current and momentum are carried by the same species of particles in a
2236: superfluid, was not required to explain any of the quantum effects.
2237: 
2238: 
2239: %\newpage
2240: \section*{APPENDIX}
2241: 
2242: In this appendix, we 
2243: %We now 
2244: %
2245: %explicitly show that in a BCS state all the electron pairs have
2246: %condensed into one and the same two-particle state.
2247: %
2248: derive the position space wave function (\ref{e:bcsn}) by
2249: projecting the BCS state (\ref{e:bcs}) onto a fixed number of pairs $N$.
2250: %We omit factors of normalization. The BCS state may be written 
2251: %
2252: The (unnormalized) BCS state may be written
2253: \begin{eqnarray}
2254:   \nonumber \ket{\psi_\phi} \!&\!=\!&\!
2255:   \prod_{\bs{k}}\left(1 +  e^{i\phi} \frac{v_{\bs{k}}}{u_{\bs{k}}} 
2256:     c_{\bs{k}\up}^\dagger\,c_{-\bs{k}\dw}^\dagger\right)\vac
2257: %  \\ \nonumber\rule{0pt}{18pt}  \!&\!=\!&\!
2258: \end{eqnarray}
2259: \pagebreak
2260: \begin{eqnarray}
2261:   \nonumber \phantom{\ket{\psi_\phi}} \!&\!=\!&\!
2262:   \prod_{\bs{k}}\exp\Bigl(e^{i\phi}\frac{v_{\bs{k}}}{u_{\bs{k}}} 
2263:     c_{\bs{k}\up}^\dagger\,c_{-\bs{k}\dw}^\dagger\Bigr)\vac 
2264:   \\ \nonumber\rule{0pt}{18pt}  \!&\!=\!&\!
2265:   \exp\Bigl(e^{i\phi} \sum_{\bs{k}} \frac{v_{\bs{k}}}{u_{\bs{k}}} 
2266:     c_{\bs{k}\up}^\dagger\,c_{-\bs{k}\dw}^\dagger\Bigr)\vac
2267:   \\ \nonumber\rule{0pt}{18pt}  \!&\!=\!&\!
2268:   \exp\bigl(e^{i\phi} b^\dagger\bigr)\vac \!.
2269: \end{eqnarray}
2270: The pair creation operator $b^\dagger$ is given by
2271: \begin{eqnarray}
2272:   \nonumber
2273:   b^\dagger\!&\equiv&\!\sum_{\bs{k}} \frac{v_{\bs{k}}}{u_{\bs{k}}} 
2274:   c_{\bs{k}\up}^\dagger\,c_{-\bs{k}\dw}^\dagger
2275:   \\ \nonumber\rule{0pt}{18pt}  \!&=&\!
2276:   \int\! d^{3\,}\!\bs{x}_1 d^{3\,}\!\bs{x}_{2}\,\varphi(\bs{x}_1-\bs{x}_2)\,
2277:   \psi^\dagger_\up(\bs{x}_1)\psi^\dagger_\dw(\bs{x}_2)\vac \!,
2278: \end{eqnarray}
2279: where $\varphi(\bs{x})$ is given by (\ref{e:varphi}).  If we now project
2280: out a state with $N$ pairs according to (\ref{e:project}), we obtain
2281: \begin{displaymath}
2282:   \ket{\psi_N}=
2283: %  \frac{1}{2\pi}
2284:   \int_0^{2\pi}\!\!d\phi\,e^{-iN\phi}\exp\bigl(e^{i\phi}b^\dagger\bigr)\vac=
2285:   \frac{2\pi}{N!}\;\!
2286:   \bigl(b^\dagger\bigr)^N \vac \!,
2287: \end{displaymath}
2288: % \begin{eqnarray}
2289: %   \nonumber
2290: %   \ket{\psi_N} 
2291: %   \!&\!\!=\!\!&\! 
2292: %   \int_0^{2\pi}\!\!d\phi\,e^{-iN\phi}\exp\bigl(e^{i\phi}b^\dagger\bigr)\vac 
2293: %   \\ \nonumber\rule{0pt}{18pt}  \!&\!\!=\!\!&\!
2294: % %  \frac{1}{N!} 
2295: %   \bigl(b^\dagger\bigr)^N \vac
2296: % \end{eqnarray}
2297: which is (up to a normalization) equivalent to (\ref{e:bcsn}).
2298: 
2299: \section*{ACKNOWLEDGMENTS}
2300: 
2301: I wish to thank T.\ Kopp, 
2302: D.\ Schuricht, M.\ Vojta, and P.\ W\"olfle for discussions of various 
2303: aspects of this work.
2304: %fundamental aspects of superconductivity.
2305: 
2306: %\newpage
2307: %\bibliographystyle{apsrmp}
2308: %\bibliography{../bib/book,../bib/paper,../bib/schuricht,../bib/martin}
2309: %\end{document}
2310: 
2311: \begin{thebibliography}{99}
2312:   
2313: %\bibitem{pacs} The PACS number {\it 11.15.Ex} stands for {\it
2314: %    Spontaneous breaking of gauge symmetries}, a classification which
2315: %  should, according to this review, not even exist.
2316: 
2317: \bibitem{sid} For an excellent review and references, see %for example
2318:   S.\ Coleman's 1973 Erice Lectures {\it Secret Symmetry}, reprinted in 
2319:   S.\ Coleman, {\it Aspects of Symmetry} (Cambridge University Press, 1985).
2320:   
2321: \bibitem{wein} S.\ Weinberg, {\it The Quantum Theory of Fields}, Vol.\ 
2322:   II (Cambridge University Press, 1996).
2323:   
2324: \bibitem{bcs} J.\ Bardeen, L.N.\ Cooper, and J.R.\ Schrieffer, {\it
2325:     Phys.\ Rev.\ }{\bf 108}, 1175 (1957).
2326: 
2327: \bibitem{schrief} J.R.\ Schrieffer, {\it Theory of Superconductivity}
2328: %  (Benjamin/Addison Wesley, 1964).
2329:   (Benjamin, 1964).
2330:   
2331: \bibitem{degenn} P.G.\ de Gennes, {\it Superconductivity of Metals and
2332:     Alloys} 
2333: %  (Benjamin/Addison Wesley, 1966).
2334:   (Benjamin, 1966).
2335: 
2336: \bibitem{tinki} Michael Tinkham, {\it Introduction to Superconductivity}
2337:   (McGraw Hill, 1996).
2338: 
2339: \bibitem{phil66} %See for example 
2340:   P.W.\ Anderson, {\it Rev.\ of Mod.\ Phys.\ }{\bf 38}, 298 (1966).
2341:   
2342: \bibitem{elit} Note also that gauge symmetry is a local symmetry, and
2343:   a local cannot break down in principle as a consequence of Elitzur's
2344:   theorem; S.~Elitzur, {\it Phys.\ Rev.\ D\ }{\bf 12}, 3978 (1975).
2345: 
2346: %\bibitem{conv} $x^\mu=(x^0,x^1,x^2,x^3)=(t,x,y,z)$,
2347: %  $x_\mu=g_{\mu\nu}x^\nu$, $1=g_{00}=-g_{11}=-g_{22}=-g_{33}$.
2348:   
2349: \bibitem{simp} For simplicity, we consider only time independent gauge
2350:   fields and transformations.
2351: 
2352: \bibitem{auer} A.\ Auerbach, {\it Interacting Electrons and Quantum
2353:     Magnetism} (Springer, 1994).
2354:   
2355: \bibitem{marshall} W.\ Marshall, {\it Proc.\ R.\ Soc.\ London}\ {\bf
2356:   A232}, 48 (1955).
2357: 
2358: \bibitem{odlro} O.\ Penrose and L.\ Onsager, {\it Phys.\ Rev.\ }{\bf
2359:     104}, 576 (1956);  C.N.\ Yang, {\it Rev.\ of Mod.\ Phys.\ }{\bf
2360:     34}, 694 (1962).
2361:   
2362: \bibitem{phil63} P.W.\ Anderson, {\it Phys.\ Rev.\ }{\bf 130}, 439 (1963).
2363:   
2364: \bibitem{lg} V.L.\ Ginzburg and L.D.\ Landau, {\it J.\ exp.\ theo.\ 
2365:     Phys.\ }{\bf 20}, 1064 (1950).
2366:   
2367: \bibitem{rel} There are relativistic corrections to this, arising \eg
2368:   from the binding energy of the Cooper pairs.  The theory of those is
2369:   still controversial; see for example M.\ Liu, {\it Phys.\ Rev.\ 
2370:     Lett.\ }{\bf 81}, 3223 (1998) and references therein.
2371: 
2372: \bibitem{feyn} R.P.\ Feynman, R.B.\ Leighton, and M.\ Sands, {\it The
2373:     Feynman Lectures of Physics} (Addison Wesley, 1965).  Feynman
2374:   takes the point of view that since $\Psi(x)$ may be interpreted as a
2375:   condensate wave function, it should obey a Schr\"odinger equation.
2376:   The corresponding Lagrange density is given by
2377: %  $$
2378: %  \mathcal{L}_{\text s}(x)=\\
2379: %  \Psi^*(x)\biggl\{\bigl( i\hbar\partial_t+e^*\Phi(x)\bigr)
2380: %  -\frac{1}{2m^*}
2381: %  \Bigl(-i\hbar\nabla+\frac{e^*}{c}\bs{A}(x)\Bigr)^2\biggr\}\Psi(x).
2382: %  $$
2383: %  \vspace{-4pt}
2384:   \begin{displaymath}
2385:   \begin{array}{r@{}l}\displaystyle
2386:   \mathcal{L}_{\text s}(x)=
2387:   \Psi^*(x)\biggl\{&\displaystyle
2388:   \bigl(i\hbar\partial_t+e^*\Phi(x)\bigr)
2389:   \\\displaystyle
2390:   -\frac{1}{2m^*}
2391:   &\displaystyle
2392:   \Bigl(-i\hbar\nabla+\frac{e^*}{c}\bs{A}(x)\Bigr)^2\biggr\}\Psi(x).
2393:   \end{array}
2394: %  \vspace{-4pt}
2395:   \end{displaymath}
2396:   If one takes the amplitude $|\Psi(x)|$ constant, this reduces
2397:   to the Lagrange density given in the text.
2398:   
2399: \bibitem{pes} See for example M.E.\ Peskin and D.V.\ Schroeder, {\it
2400:     An Introduction to Quantum Field Theory} (Addison Wesley, 1995),
2401:   page 15.
2402:   
2403: \bibitem{conn} Note, however, that there is no connection as of yet 
2404:   between the effective mass $m^*$ in (\ref{e:tvs})--(\ref{e:gal2})
2405:   and $m^*$ in the effective Lagrange density.  The equivalence is
2406:   only established with (\ref{e:js}) and (\ref{e:gww}) below.
2407:   
2408: \bibitem{phila} Anderson~\cite{phil66} refers to our definition
2409:   (\ref{e:vs}) of $\bs{v}_{\text s}(x)$ as a ``pseudo-identity'', as
2410:   the framework of his analysis does not provide the means to
2411:   recognize the approximative nature of (\ref{e:j}).
2412:   
2413: \bibitem{gww} M.\ Greiter, F.\ Wilczek, and E.\ Witten, {\it Mod.\ Phys.\ 
2414:     Lett.\ B\ }{\bf 3}, 903 (1989).
2415: 
2416: \bibitem{schakel} A.M.J.\ Schakel, {\it Mod.\ Phys.\ 
2417:     Lett.\ B\ }{\bf 14}, 927 (1990).
2418:   
2419: \bibitem{london} \label{london} F.\ London, {\it Superfluids}, Vol.~I (Wiley,
2420:   1950).
2421:   
2422: \bibitem{london1} Strictly speaking, this is London's second and more
2423:   famous equation.  The first equation,
2424:   \begin{displaymath}
2425: %    \partial_t \bs{J}_{\text{el.}}=\frac{{e^*}^2 n_{\text s}}{m^*}\bs{E},
2426:     \partial_t \bs{J}=-\frac{e^* n_{\text s}}{m^*}\bs{E},
2427:   \end{displaymath}
2428:   is obtained from (\ref{e:muelch}) with (\ref{e:j3}) if we neglect the
2429:   $\nabla\mu$ term, which is a rather crude approximation.
2430:   
2431: \bibitem{bok} J.\ Bok and J.\ Klein, {\it Phys.\ Rev.\ Lett.\ }{\bf
2432:     20}, 660 (1968).
2433:   
2434: \bibitem{rot} There is yet another effect which depends on the Cooper
2435:   pair mass directly.  If one rotates a superconductor, the rotation
2436:   of the uniformly charged background induces a uniform magnetic field
2437:   of strength
2438:   \begin{displaymath}
2439: %  \begin{equation}
2440: %    \bs{B}=\frac{n_{\text{s}}}{\rho}\frac{m^*c}{e^*}2\bs{\omega},
2441:     \bs{B}=\frac{2m^*c}{e^*}\;\!\bs{\Omega}
2442: %  \end{equation}
2443:   \end{displaymath}
2444:   well inside the superconductor, where $\bs{\Omega}$ is the angular
2445:   velocity of rotation.  To derive this effect, one simply has to add the
2446:   electric current $e^* n_{\text{s}}(\bs{\Omega}\times\bs{r})$ induced
2447:   by the rotating background charge to the right of (\ref{e:max2}):  
2448:   \begin{displaymath}
2449:     \nabla\times\bs{B}-\frac{1}{c}\partial_t\bs{E}=
2450:     -\frac{4\pi e^*}{c}\bigl(\bs{J}-n_{\text{s}}(\bs{\Omega}\times\bs{r})\bigr).
2451:   \end{displaymath}
2452:   Discarding the time derivative term and taking the curl, one obtains with
2453:   (\ref{e:j5}) and $\nabla\times(\bs{\Omega}\times\bs{r})=2\bs{\Omega}$
2454:   \begin{displaymath}
2455:     \nabla\times\nabla\times\bs{B}=-\frac{4\pi e^*}{c}
2456:     \Bigl(\frac{e^* n_{\text s}}{m^*c}\bs{B}-n_{\text{s}}\,2\bs{\Omega}\Bigr).
2457:   \end{displaymath}
2458:   Solving this equation yields $\nabla\times\nabla\times\bs{B}=0$ and
2459:   therefore a uniform magnetic field %$\bs{B}$ as given above 
2460:   deep inside the superconductor.  See Reference \ref{london}, \S 12;
2461:   J.\ Tate, B.\ Cabrera, S.B.\ Felch, and J.T.\ Anderson, {\it Phys.\ 
2462:     Rev.\ Lett.\ }{\bf 162}, 845 (1989).
2463: 
2464: \bibitem{london2} See page 152.  Note that this prediction shows that
2465:   London had fully understood the significance of vector potentials in
2466:   quantum mechanics almost a decade before the acclaimed work by
2467:   Y.\ Aharonov and D.\ Bohm.
2468:   
2469: \bibitem{jos} B.D.\ Josephson, {\it Phys.\ Lett.\ }{\bf 1}, 251
2470:   (1962).
2471: 
2472: 
2473: %\bibitem{lie} We have excluded the total derivative term $-\hbar
2474: %  n_{\text s}$ from our considerations, as it is irrelevant to the
2475: %  equations of motion.
2476: 
2477: \end{thebibliography}
2478: 
2479: 
2480: \end{document}
2481: 
2482: \bibitem{} 
2483: 
2484: \bibitem{} 
2485: 
2486: 
2487: 
2488: 
2489: ************************************************************************
2490: % \begin{itemize}
2491: % \item[I.] Introduction
2492: % %\item[I.] INTRODUCTION  
2493: % % 
2494: % \item[II.] Gauge invariance  
2495: % %\item[II.] GAUGE INVARIANCE
2496: %   
2497: %   Why gauge invariance cannot break down spontaneously as a matter of
2498: %   principle.  Gauge transformations in (quantum) field theory and
2499: %   quantum mechanics.  Gauge invariance of the microscopic theory of
2500: %   superconductivity.
2501: % %
2502: % \item[III.] Order parameter considerations.  
2503: %   
2504: %   Relation between formulations of superconductivity with fixed
2505: %   particle numbers {\it vs}.\ well defined phases.  Physical origin of
2506: %   the many degenerate ground states.
2507: % %
2508: % \item[IV.]  Effective field theory at low temperatures.  
2509: %   
2510: %   Motivation of the effective Lagrangian for neutral superfluids and
2511: %   superconductors.  Particle density and current from order parameter
2512: %   phase rotation symmetry.  Hamiltonian formulation.  Energy and
2513: %   momentum density from time and space translation symmetry.
2514: %   Definition of a superfluid velocity and corrections to the effective
2515: %   Lagrangian.  Derivation of a generalized version of Anderson's
2516: %   characteristic equations of superfluidity from Hamilton's equations.
2517: %   Local {\it vs.}\ substantial acceleration.  Justification of the
2518: %   classical nature of the field theory.
2519: % %
2520: % \item[V.] Phenomenology and the Higgs mechanism.  
2521: %   
2522: %   Phenomenology of neutral superfluids, vortex quantization.  General
2523: %   introduction to the Higgs mechanism.  Phenomenology of
2524: %   superconductors, electric and magnetic screening, London's equation,
2525: %   Bernoulli Hall effect, balance of the Lorentz force.  Higgs
2526: %   mechanism exemplified with superconductors, effective Lagrangian in
2527: %   terms of the chemical potential and the superfluid velocity.
2528: %   Distinction between spontaneously broken global U(1) symmetry and
2529: %   intact gauge symmetry.
2530: % %
2531: % \item[VI.] {Quantum effects}  
2532: % %\item[\bf VI.] {\bf Quantum effects}  
2533: %   
2534: %   Planck's constant manifests itself through the compactness of the
2535: %   phase field in flux quantization, phase slippage, and the Josephson
2536: %   effect.
2537: % \end{itemize}
2538: