1: \documentstyle[11pt]{article}
2: %
3: \def\one{1\hskip -.37em 1}
4: \def\proj{E\hskip -.69em I}
5: %
6: \setlength{\textwidth}{155mm}
7: \setlength{\textheight}{220mm}
8: \setlength{\topmargin}{-5mm}
9: \setlength{\oddsidemargin}{3mm}
10: \setlength{\evensidemargin}{3mm}
11: %
12: \setlength{\parskip}{1mm}
13: \setlength{\parindent}{10mm}
14: \setlength{\baselineskip}{7mm}
15: %
16: \begin{document}
17: %
18: \begin{titlepage}
19: \begin{flushright}
20: YITP-00-27\\
21: %cond-mat/0006070\\
22: June 2000
23: \end{flushright}
24: %
25: \begin{centering}
26:
27: {\ }\vspace{1cm}
28:
29: {\Large\bf On Electric Fields in Low Temperature Superconductors}\\
30:
31: \vspace{2.0cm}
32:
33: Jan Govaerts\footnote{On leave from the Institute of Nuclear
34: Physics, Catholic University of Louvain, Louvain-la-Neuve, Belgium\\
35: \indent{\ }\hspace{5pt}E-mail: {\tt govaerts@fynu.ucl.ac.be}}\\
36:
37: \vspace{0.6cm}
38:
39: {\em C.N. Yang Institute for Theoretical Physics}\\
40: {\em State University of New York at Stony Brook}\\
41: {\em Stony Brook NY 11794-3840, USA}\\
42:
43: \vspace{0.6cm}
44:
45: Damien Bertrand and Geoffrey Stenuit\footnote{
46: E-mail: {\tt bertrand, stenuit@fynu.ucl.ac.be}}\\
47:
48: \vspace{0.6cm}
49:
50: {\em Institut de Physique Nucl\'eaire,
51: Universit\'e catholique de Louvain}\\
52: {\em 2, Chemin du Cyclotron, B-1348 Louvain-la-Neuve, Belgium}
53:
54: \vspace{2cm}
55: \begin{abstract}
56:
57: \noindent The manifestly Lorentz covariant Landau-Ginzburg equations
58: coupled to Maxwell's equations are considered as a possible framework
59: for the effective description of the interactions between low temperature
60: superconductors and magnetic as well as electric fields. A specific
61: experimental set-up, involving a nanoscopic superconductor and only
62: static applied fields whose geometry is crucial however, is described,
63: which should allow to confirm or invalidate the covariant model through the
64: determination of the temperature dependency of the critical
65: magnetic-electric field phase diagram and the identification
66: of some distinctive features it should display.
67:
68: \end{abstract}
69:
70: \vspace{10pt}
71:
72: \end{centering}
73:
74: \vspace{5pt}
75:
76: \noindent PACS numbers: 74., 74.20.De, 74.25.Dw
77:
78: \vspace{25pt}
79:
80: \end{titlepage}
81:
82: \setcounter{footnote}{0}
83:
84: \noindent{\bf 1. Introduction.}
85: It is a widely held belief that (low $T_c$) superconductors cannot
86: sustain electric fields in static configurations. The argument\cite{Tinkham}
87: is directly based on the first of the London equations,
88: \begin{equation}
89: \frac{\partial}{\partial t}\left(\Lambda\vec{J}\right)=\vec{E}\ \ \ ,\ \ \
90: \vec{\partial}\times\left(\Lambda\vec{J}\right)=-\vec{B}\ \ \ ,
91: \label{eq:London}
92: \end{equation}
93: where $\Lambda$ is a phenomenological parameter proper to the superconducting
94: material, $\vec{J}$ is the supercurrent density, and $\vec{E}$, $\vec{B}$ are
95: the electric and magnetic fields, respectively. Indeed, given that relation
96: as well as the property of infinite conductivity, any nonvanishing electric
97: field within the sample must set into motion a dissipationless supercurrent,
98: hence a displacement of charges which very rapidly leads to an exact screening
99: of any such electric field, certainly for time independent configurations
100: (the case of stationary configurations with externally sustained supercurrents
101: in the presence of magnetic vortices in type II superconductors is a different
102: matter, see for example Refs.\cite{Josephson,Bok, Kolacek}).
103:
104: However, such a situation raises a series of puzzles of varying degrees of
105: concern. First, it is physically inconceivable that an applied
106: electric field would discontinuously drop to zero from its external value
107: when crossing the surface of a superconducting sample. There
108: ought to exist some skin effect with a characteristic nonvanishing
109: penetration depth however small. Nevertheless, none of the parameters
110: appearing in the London equations provides for such an electric
111: penetration length. Moreover, in the above picture for the screening of
112: electric fields, mention is made of an electromagnetic
113: charge density which also is not accounted for in the London equations,
114: nor more generally in the Landau-Ginzburg (LG) equations (see below).
115:
116: Another concern at a more formal level is the fact that the coupling
117: of the London and LG equations to the electromagnetic fields is not
118: spacetime covariant. Indeed, under Lorentz boosts, the
119: supercurrent density $\vec{J}$ ought to transform as the space
120: components of a 4-vector whose time component would then play
121: the role of the aforementioned missing charge density, while the electric
122: $\vec{E}$ and magnetic $\vec{B}$ fields transform as components of the
123: two index antisymmetric field strength tensor $F_{\mu\nu}$.
124: To illustrate this point more vividly perhaps, consider
125: a flat infinite superconducting slab submitted to an external homogeneous
126: magnetic field lying parallel to its surface. In such a case,
127: the magnetic field will only partially penetrate the sample, with
128: a characteristic penetration length related to the parameter $\Lambda$
129: above\cite{Tinkham}. Imagine now performing a Lorentz boost in a
130: direction both parallel to the surface of the slab and perpendicular
131: to the applied magnetic field. According to the Lorentz covariance of
132: Maxwell's equations, in the boosted frame there appears now an electric
133: field perpendicular to the surface of the sample, {\sl also within
134: the volume of the superconductor where the magnetic field in the
135: initial frame is nonvanishing\/}. Thus Lorentz covariance requires
136: the possibility of electric fields on the same footing as magnetic
137: ones within superconductors, with an electric penetration length
138: equal to the familiar magnetic one. Nevertheless, the existence of
139: such electric fields within superconductors is incompatible with
140: the London equations.
141:
142: One may take issue with the above covariance argument,
143: since the superconducting sample itself defines a preferred frame,
144: thereby insisting that the physics should be described only with respect
145: to that specific rest-frame. Even though that frame is obviously
146: distinguished, the coupling of the London and LG equations
147: to the electromagnetic fields should be consistent with the covariance
148: properties of Maxwell's equations, even within that frame.
149: In addition, one would also like to have available
150: a manifestly covariant framework in which to study the interactions
151: of electromagnetic fields and moving superconducting samples,
152: an issue which the usual London and LG equations are unable to address,
153: and which is certainly accounted for to anyone's satisfaction in the
154: description of electromagnetic interactions with ordinary conductors. It
155: remains true nevertheless that some physical characterizations of
156: superconductors can be defined with respect only to the rest-frame, such
157: as for example the frame dependent notion of the free energy whose value
158: determines the occurrence of the superconducting-normal phase transition
159: only when evaluated in the rest-frame.
160:
161: Note that any spacetime covariant extension of the usual London and
162: LG equations entails a time dependent LG (TDLG) equation in which space and
163: time variations are on the same footing. Namely, a covariant
164: TDLG equation is necessarily of second-order in time derivatives
165: as it is in space derivatives, in sharp contrast with the usual
166: non covariant first-order TDLG equations encountered in the
167: literature\cite{Tinkham}. In particular, in a covariant setting, the
168: time scale associated to time dependent fluctuations is then naturally
169: set by the time it takes light to travel the distance of some mean value
170: of the penetration and coherence lengths. In contradistinction for
171: first-order TDLG equations, this time scale is specified
172: in terms of an additional parameter, the relaxation constant\cite{Tinkham}.
173: As a matter of fact, the latter quantity involves Boltzmann's constant,
174: showing that the usual TDLG equations apply rather to the time
175: dependency encured through thermodynamic fluctuations in the
176: superconducting order parameter. As such, these TDLG equations do not
177: provide a framework in which to study the intrinsically genuine time
178: dependent dynamics of superconductors coupled to time varying electromagnetic
179: fields, in the absence of thermodynamic fluctuations. A covariant
180: extension of the London and LG equations would also provide such
181: a dynamic framework, of relevance for instance to the
182: dynamics of ensembles of magnetic vortices interacting among one another
183: and with their electromagnetic environment.
184:
185: A manifestly covariant extension of the LG equation which immediately
186: comes to one's mind is of course the so-called U(1) Higgs model of
187: particle physics, whose construction itself was motivated by the
188: BCS and LG theories in the late 50's. Indeed, as an effective theory
189: for superconductivity, this model coincides with the original LG
190: formulation for stationnary configurations, and readily provides\cite{Weinberg}
191: a description of all the remarkable quantum phenomena of superconductivity.
192: It may thus appear somewhat surprising that the covariant formulation
193: has not been used to further explore superconducting phenomena
194: possibly lying beyond the boundaries of the usual London and LG equations.
195: The purpose of this Letter is to suggest examples of
196: investigations along such lines.
197:
198: After some considerations on the covariant London and LG equations presented
199: in the next section, section 3 identifies a specific set-up which should
200: enable to establish experimentally whether the covariant, rather than
201: the usual noncovariant approach is relevant to superconducting phenomena
202: in the presence of electric fields. What appears to us to be quite a
203: remarkable circumstance is that this experimental confirmation of the
204: covariant model should prove to be possible already using only a static,
205: time independent configuration of external fields, through the observation
206: of the superconducting-normal phase transition and the determination
207: of the phase diagram in the $(B,E)$ plane for a specific geometry of the
208: applied fields.
209:
210: \vspace{10pt}
211:
212: \noindent{\bf 2. The covariant LG equations.}
213: As mentioned already, the considered model, of application only to low
214: temperature superconductors, is that of a U(1) gauge invariant coupling to
215: the electromagnetic interactions of a complex scalar field $\psi$ of charge
216: $q=-2e<0$ (the Cooper pair charge) and with self-interactions determined
217: through the usual LG potential $(|\psi|^2-1)^2$ properly normalized.
218: The order parameter $\psi$ is normalized to the square-root of the
219: Cooper pair density in the bulk in the absence
220: of any electromagnetic field (see Ref.\cite{Gov1} for some further
221: details). Rather than listing all the relevant equations
222: in terms of the physical quantities, let us already use the following
223: choice of units. Space and time coordinates, namely $\vec{x}$ and
224: the combination $x^0=ct$ with $c$ being the speed of light in vacuum,
225: are measured in units of the penetration length $\lambda(T)$. Similarly,
226: magnetic $\vec{B}$ and electric $\vec{E}/c$ fields are measured in units
227: of $\Phi_0/(2\pi\lambda^2(T))$, where $\Phi_0=2\pi\hbar/|q|$ is the usual
228: quantum of flux. Note that these units are temperature dependent,
229: since the penetration length $\lambda(T)$ is, with a dependency
230: we shall model\cite{Tinkham} through
231: $\lambda(T)=\lambda(0)\left(1-(T/T_c)^4\right)^{-1/2}$,
232: $T_c$ being of course the critical temperature. Finally, the order
233: parameter $\psi$ is parametrized according to $\psi=f e^{i\theta}$,
234: with $f$ real and $f^2$ thus measuring the relative Cooper pair density.
235: In terms of these units, space and time coordinates are denoted
236: $\vec{u}$ and $\tau$, and the magnetic and electric fields $\vec{b}$
237: and $\vec{e}$, respectively. Finally, let us also introduce the
238: quantities,
239: \begin{equation}
240: j^0=\frac{q}{\hbar}\frac{\lambda^3(T)}{f^2}\,\mu_0\,c\rho_{\rm em}\ \ \ ,\ \ \
241: \vec{j}=\frac{q}{\hbar}\frac{\lambda^3(T)}{f^2}\,\mu_0\,\vec{J}_{\rm em}\ ,
242: \end{equation}
243: where $\mu_0$ is the usual vacuum magnetic permitivity, and
244: $(c\rho_{\rm em},\vec{J}_{\rm em})$ are the superconducting electromagnetic
245: charge and current densities (constructed in terms of $\psi$), indeed
246: defining a 4-vector under Lorentz transformations.
247: Note that these relations show that $(f^2j^0,f^2\vec{j})$ is in fact
248: proportional to this electromagnetic 4-supercurrent, which must be
249: locally conserved.
250:
251: The latter remark is also confirmed by the inhomogeneous Maxwell
252: equations (all space derivatives are of course with respect
253: to $\vec{u}$),
254: \begin{equation}
255: \vec{\partial}\cdot\vec{e}=-f^2\,j^0\ \ \ ,\ \ \
256: \vec{\partial}\times\vec{b}-\partial_\tau\vec{e}=-f^2\vec{j}\ ,
257: \label{eq:Maxi}
258: \end{equation}
259: which indeed, as is usual, imply the local conservation property
260: $\partial_\tau\left(f^2j^0\right)+\vec{\partial}\cdot
261: \left(f^2\vec{j}\right)=0$.
262: The remaining electromagnetic equations of motion are given by,
263: \begin{equation}
264: \partial_\tau\vec{j}\,+\,\vec{\partial}j^0=-\vec{e}\ \ \ ,\ \ \
265: \vec{\partial}\times\vec{j}=\vec{b}\ ,
266: \label{eq:London2}
267: \end{equation}
268: which are recognized as the appropriate covariant extension of the
269: London equations in (\ref{eq:London}). Note that only the first London
270: equation is modified by the inclusion of a contribution of the supercharge
271: density, as was indeed required, while the second London equation,
272: essential to the Meissner effect, retains its original form.
273: The homogeneous Maxwell equations,
274: \begin{equation}
275: \vec{\partial}\times\vec{e}+\partial_\tau\vec{b}=\vec{0}\ \ \ ,\ \ \
276: \vec{\partial}\cdot\vec{b}=0\ ,
277: \end{equation}
278: follow from the covariant London equations (\ref{eq:London2}) (as they
279: do also from the noncovariant ones in (\ref{eq:London})).
280:
281: The equations of motion for the order parameter $\psi$ are given,
282: on the one hand, by the covariant LG equation
283: \begin{equation}
284: \left[\vec{\partial}^2-\partial^2_\tau\right]f=
285: \left[\vec{j}^2-{j^0}^2\right]f-\kappa^2(1-f^2)f\ ,
286: \label{eq:LG}
287: \end{equation}
288: where the LG parameter $\kappa=\lambda(T)/\xi(T)$---$\xi(T)$ being
289: the coherence length---is essentially temperature independent\cite{Tinkham},
290: and on the other hand, by the following conditions for the quantum phase
291: $\theta$,
292: \begin{equation}
293: \partial_\tau\theta=-j^0+\varphi\ \ \ ,\ \ \
294: \vec{\partial}\theta=\vec{j}-\vec{a}\ .
295: \label{eq:theta}
296: \end{equation}
297: In these latter relations, $\varphi$ and $\vec{a}$ are the scalar and
298: vector gauge potentials defined such that
299: \begin{equation}
300: \vec{e}=-\vec{\partial}\varphi-\partial_\tau\vec{a}\ \ \ ,\ \ \
301: \vec{b}=\vec{\partial}\times\vec{a}\ .
302: \end{equation}
303: Finally, these equations are subject to boundary conditions requiring
304: vanishing values for those components of the vectors $\vec{\partial}f$
305: and $\vec{j}$ which are perpendicular to the surfaces of the
306: superconducting sample in contact with an insulating material.
307:
308: The advantage of using this representation of the system is that the
309: number of gauge dependent variables is kept to a minimum\cite{Gov1}. Only the
310: quantities $\theta$, $\varphi$ and $\vec{a}$ are defined up to the following
311: gauge transformations,
312: \begin{equation}
313: \theta'=\theta+\chi\ \ ,\ \
314: \varphi'=\varphi+\partial_\tau\chi\ \ ,\ \
315: \vec{a}'=\vec{a}-\vec{\partial}\chi\ ,
316: \end{equation}
317: $\chi(\vec{u},\tau)$ being an arbitrary function, while the absolute
318: sign of $f$ may also be subject to gauge transformations\cite{Gov2}.
319:
320: In fact, this decoupling of gauge variant and gauge invariant quantities
321: may be rendered complete when substituting the covariant London equations
322: (\ref{eq:London2}) into the inhomogeneous Maxwell equations (\ref{eq:Maxi}).
323: In addition to the covariant LG equation (\ref{eq:LG}), one then finds
324: for the $(j^0,\vec{j})$ 4-supercurrent,
325: \begin{equation}
326: \left[\vec{\partial}^2-\partial^2_\tau\right]j^0=
327: f^2 j^0-\partial_\tau
328: \left[\partial_\tau j^0+\vec{\partial}\cdot\vec{j}\right]\ \ \ ,\ \ \
329: \left[\vec{\partial}^2-\partial^2_\tau\right]\vec{j}=
330: f^2\vec{j}+\vec{\partial}
331: \left[\partial_\tau j^0+\vec{\partial}\cdot\vec{j}\right]\ .
332: \label{eq:current}
333: \end{equation}
334: Any solution to this set of three coupled equations for $j^0$, $\vec{j}$
335: and $f$ then leads to specific values for $\vec{b}$ and $\vec{e}$
336: through the covariant London equations (\ref{eq:London2}), and in turn,
337: once the gauge potentials $\varphi$ and $\vec{a}$ related to these fields
338: determined up to the gauge transformations parametrized by $\chi$,
339: the corresponding solution for the quantum phase $\theta$ is also
340: finally obtained from (\ref{eq:theta})\cite{Gov2}.
341:
342: Although nonlinear, the equations (\ref{eq:LG}) and (\ref{eq:current})
343: also establish that these covariant LG equation admit progressive
344: wave solutions with covariant dispersion relations in a linear
345: regime. As a matter of fact, the phase and group velocities for
346: variations in the 4-supercurrent and the order parameter $f$ are different,
347: unless the LG parameter $\kappa$ takes the same critical value
348: $\kappa_c=1/\sqrt{2}$ as the one which is so crucial to the understanding
349: of the stability and interaction properties of magnetic vortices.
350: Indeed, for a fluctuation of wave number $k$ (in units of $1/|\vec{u}|$)
351: around the vacuum solution
352: $(j^0=0,\vec{j}=\vec{0},f=1)$, the group velocities are,
353: respectively,
354: \begin{equation}
355: v_j=\frac{k}{\sqrt{k^2+1}}\ \ \ ,\ \ \
356: v_f=\frac{k}{\sqrt{k^2+2\kappa^2}}\ ,
357: \end{equation}
358: thus showing that when $\kappa>\kappa_c$ (resp. $\kappa<\kappa_c$),
359: waves in the 4-supercurrent will overtake (resp. be overtaken by)
360: those in the order parameter $f$. In other words, within the superconductor,
361: fluctuations in the electromagnetic fields will propagate more rapidly
362: (resp. slowly) than those in the order parameter, in accordance with
363: the relative magnetic and superconducting rigidities that the parameters
364: $\lambda(T)$ and $\xi(T)$ characterize.
365:
366: Clearly, such properties are totally different in the case of the
367: usual noncovariant first-order TDLG equation, in which time scales are then
368: normalized with respect to the relaxation parameter, which itself is
369: temperature dependent. The ensuing dispersion relations are then
370: linear in frequency, implying that the group velocities of fluctuations
371: in the supercurrent $\vec{j}$ and in the order parameter $f$ are
372: then also identical, independently of the value for $\kappa$.
373: Such differences between the covariant and noncovariant frameworks
374: must lead to distinct physical properties in the case of time dependent
375: configurations in ultra-high frequency regimes,
376: $\nu\sim c/\lambda(T),c/\xi(T)$, an issue which, however, is beyond
377: the scope of this work.
378:
379: To conclude this general discussion, let us
380: also give the expression for the free energy $E$ of the system in
381: the covariant form,
382: \begin{displaymath}
383: \left(\frac{\lambda^3(T)}{2\mu_0}
384: \left(\frac{\Phi_0}{2\pi\lambda^2(T)}\right)^2\right)^{-1}\,E=
385: \int_{(\infty)}d^3\vec{u}\left\{
386: \left[\vec{e}-\vec{e}_{\rm ext}\right]^2+
387: \left[\vec{b}-\vec{b}_{\rm ext}\right]^2\right\}\ +\
388: \end{displaymath}
389: \begin{equation}
390: +\ \int_\Omega d^3\vec{u}\left\{
391: \left(\partial_\tau f\right)^2+\left(\vec{\partial}f\right)^2+
392: f^2\left({j^0}^2+\vec{j}^2\right)+\frac{1}{2}\kappa^2\left(1-f^2\right)^2
393: -\frac{1}{2}\kappa^2\right\}\ ,
394: \label{eq:E1}
395: \end{equation}
396: where the normalization factor related to our choice of units is displayed
397: together with $E$ in the l.h.s., $\vec{e}_{\rm ext}$ and $\vec{b}_{\rm ext}$
398: are externally applied electric and magnetic fields, respectively, and
399: $\Omega$ stands for the volume of the superconducting sample.
400:
401: The same expression is also of application to the noncovariant model,
402: in which case one has $j^0=0$ and $\vec{e}=\vec{0}$ within the superconductor,
403: and the quadratic term in $\partial_\tau f$ is to be replaced by
404: a linear term while the time coordinate is then also measured in units of
405: the relaxation parameter for the TDLG equation. The term in
406: $\left[\vec{e}-\vec{e}_{\rm ext}\right]^2$ measures the energy required
407: to expulse the electric field from the superconductor. In the
408: noncovariant case, and in accordance with the first London equation,
409: we shall thus assume that the associated penetration depth
410: is essentially vanishing for all practical purposes. For physics
411: reasons, such an approximation cannot be very reliable when it
412: comes to nanoscopic superconductors, but we shall use it as
413: a working hypothesis anyway. Note that the free energy $E$
414: is defined here in such a way that it vanishes at the superconducting-normal
415: phase transition. And as a last remark, clearly, in the case of
416: stationary configurations and in the absence of any electric fields,
417: the equations of both approaches coincide
418: with the usual noncovariant time independent LG equations.
419:
420: \vspace{10pt}
421:
422: \noindent{\bf 3. Characterizing the phase transition.}
423: In order to identify a specific geometry of applied fields which could
424: help discriminate experimentally between the two approaches already in a
425: static configuration, consider again the situation of the flat infinite
426: slab of the Introduction, this time subjected not only to the homogeneous
427: magnetic field parallel to its surface, but also to an homogeneous
428: electric field applied perpendicularly to its surface (an electric
429: field parallel to the slab does not induce a supercharge
430: distribution $j^0$, hence neither a feature distinctive from the
431: noncovariant model). The slab is
432: taken to be of thickness $2a$, while the external electric field
433: $\vec{e}_{\rm ext}$ is aligned along the $x$ axis, and the external magnetic
434: field $\vec{b}_{\rm ext}$ along the $y$ axis, with components
435: $e_{\rm ext}$ and $b_{\rm ext}$, respectively (the origin of this
436: coordinate system is of course positioned in the center of the slab).
437: For this specific geometry, the expulsion of the magnetic field is
438: achieved through an induced supercurrent $\vec{j}$ circulating along
439: the $z$ axis, also parallel to the slab, while that of the electric field
440: is achieved through the appearance of a nonvanishing supercharge density
441: $j^0$, an occurrence which simply cannot arise in the noncovariant approach.
442: Both these effects imply a deviation from its canonical value of unity
443: for the order parameter $f$. In view of the symmetries of the problem, both
444: $j^0(u)$ and $j^z(u)$ are odd functions of the normalized $u=x/\lambda(T)$
445: coordinate along the $x$ axis, while $f(u)$ is even.
446:
447: Given the equations and the different conditions imposed on these quantities
448: at the boundaries $u=\pm u_a\equiv\pm a/\lambda(T)$, it proves possible as
449: well as useful to express both $j^0(u)$ and $j^z(u)$ in terms of a single
450: function $j(u)$
451: \begin{equation}
452: j^0(u)=-e_{\rm ext} j(u)\ \ \ ,\ \ \
453: j^z(u)=-b_{\rm ext} j(u)\ ,
454: \end{equation}
455: so that one has for the electric and magnetic fields within the
456: sample,
457: \begin{equation}
458: e(u)=e_{\rm ext}\frac{d}{du}j(u)\ \ \ ,\ \ \
459: b(u)=b_{\rm ext}\frac{d}{du}j(u)\ ,
460: \end{equation}
461: thus showing once again that Lorentz covariance implies that the
462: penetration lengths for both types of fields are identical.
463: The set of equations to be considered then reduces to,
464: \begin{equation}
465: \frac{d^2}{du^2}j(u)=f^2(u)j(u)\ \ \ ,\ \ \
466: \frac{d^2}{du^2}f(u)=\left(b^2_{\rm ext}-e^2_{\rm ext}\right)j^2(u)f(u)-
467: \kappa^2\left(1-f^2(u)\right)f(u)\ ,
468: \label{eq:slab1}
469: \end{equation}
470: subject to the boundary conditions
471: \begin{equation}
472: \frac{d}{du}j(u)_{|u=\pm u_a}=1\ \ \ ,\ \ \
473: \frac{d}{du}f(u)_{|u=\pm u_a}=0\ .
474: \end{equation}
475: Note already the subtle interplay between the magnetic and electric
476: field contributions to the LG equation for $f(u)$, which leads to
477: values larger than unity for $f(u)$ in the electric regime
478: $e^2_{\rm ext}>b^2_{\rm ext}$, while, as is usual, $f(u)$ remains less
479: than unity in the magnetic regime $b^2_{\rm ext}>e^2_{\rm ext}$.
480: The existence of these two regimes is a direct and distinctive
481: consequence of manifest Lorentz covariance; only the magnetic one arises
482: in the noncovariant approach (see below). Note also that in view
483: of these equations, the solutions for $j(u)$ and $f(u)$ are necessarily
484: functions of the specific combination $(b^2_{\rm ext}-e^2_{\rm ext})$ only,
485: indeed justifying this notion of electric or magnetic regimes.
486:
487: Up to the normalisation factor displayed in (\ref{eq:E1}) as well
488: as the infinite surface of the slab, the free energy ${\cal E}$
489: of configurations obeying these equations is simply given by
490: \begin{equation}
491: {\cal E}=2u_a\left\{\left[1-\frac{1}{u_a}j(u_a)\right]
492: \left(b^2_{\rm ext}+e^2_{\rm ext}\right)\ -\
493: \frac{1}{u_a}\int_0^{u_a}du\left[
494: \left(b^2_{\rm ext}-e^2_{\rm ext}\right)j^2f^2+
495: \frac{1}{2}\kappa^2f^4\right]\right\}\ .
496: \end{equation}
497: Consequently, the curve in the $(b,e)$ phase diagram which
498: characterizes the superconducting-normal phase transition obeys
499: the following equation in the covariant approach,
500: \begin{equation}
501: b^2+e^2=\frac{1}{\left[1-\frac{1}{u_a}j(u_a)\right]}\,
502: \frac{1}{u_a}\int_0^{u_a}du\left[
503: \left(b^2-e^2\right)j^2f^2+
504: \frac{1}{2}\kappa^2f^4\right]\ .
505: \label{eq:critical1}
506: \end{equation}
507: The noteworthy property of this relation is that the l.h.s. involves
508: only the combination $(b^2+e^2)$, while the r.h.s. is only a function of
509: the combination $(b^2-e^2)$ of the external fields (since the solutions
510: $j(u)$ and $f(u)$ also share that property).
511:
512: Before addressing the specific consequences of this equation for the
513: $(b,e)$ phase diagram, let us consider the corresponding expressions in
514: the noncovariant approach. In that case, one has of course $j^0(u)=0$
515: (which also implies $e(u)=0$ within the superconductor, as follows
516: from the first London equation) as well as $j^z(u)=-b_{\rm ext}j(u)$.
517: The equations then remain as given in (\ref{eq:slab1}), including the
518: boundary conditions, with the only but important difference that the factor
519: $(b^2_{\rm ext}-e^2_{\rm ext})$ appearing in the LG equation is
520: of course replaced by $b^2_{\rm ext}$ only. Hence, the noncovariant
521: LG equations only admit the magnetic regime of solutions.
522: Note that in this case, the solutions for $j(u)$ and $f(u)$ are
523: then also functions of the $b^2_{\rm ext}$ external field only.
524: Consideration of the expression for the free energy then leads to
525: the following condition of criticality in the $(b,e)$ phase diagram
526: in the noncovariant case,
527: \begin{equation}
528: b^2+\frac{1}{\left[1-\frac{1}{u_a}j(u_a)\right]}e^2=
529: \frac{1}{\left[1-\frac{1}{u_a}j(u_a)\right]}\,
530: \frac{1}{u_a}\int_0^{u_a}du\left[
531: b^2j^2f^2+\frac{1}{2}\kappa^2f^4\right]\ .
532: \label{eq:critical2}
533: \end{equation}
534: In spite of the apparent similarity with (\ref{eq:critical1}), recall
535: however that the r.h.s. of this expression is a function of $b^2$ only,
536: while the l.h.s. is no longer the simple combination $b^2+e^2$
537: characteristic of a circle since the coefficient multiplying the term
538: in $e^2$ is also a function of $b^2$.
539:
540: Note that the conditions (\ref{eq:critical1}) and (\ref{eq:critical2})
541: coincide in the limit that no electric field is applied, $e=0$, as they
542: should of course. Moreover in the absence of any magnetic field, $b=0$,
543: (\ref{eq:critical2}) implies the existence of a nonvanishing critical
544: electric field, $e_0=\kappa/\sqrt{2}$, or in physical units
545: $E_0(T)/c=\left(\lambda(0)/\lambda(T)\right)^2B^\infty_c(0)$,
546: $B^\infty_c(0)=\Phi_0/(2\sqrt{2}\pi\lambda(0)\xi(0))$ being the usual
547: thermodynamic critical magnetic field in the bulk at zero temperature.
548: Clearly, the existence of such a critical electric field even in the
549: noncovariant approach is consequence of our definition for the free
550: energy in (\ref{eq:E1}) which accounts for the expulsed electric
551: energy density through the term in $\left[\vec{e}-\vec{e}_{\rm ext}\right]^2$.
552: In particular, this critical electric field $E_0(T)$ vanishes
553: at the critical temperature $T_c$, as does the critical magnetic field
554: $B_0(T)$ in the absence of any electric field, $e=0$.
555:
556: \vspace{10pt}
557:
558: \noindent{\bf 4. The $(B,E)$ phase diagram.}
559: A complete unravelling of the consequences of the criticality conditions
560: (\ref{eq:critical1}) and (\ref{eq:critical2}) requires of course
561: a numerical approach. Nevertheless, an analysis in some limiting
562: situations already suffices to gain insight into the differences
563: implied by the two models. An obvious such situation is obtained in
564: the macroscopic limit, namely when the slab half-thickness $a$ is
565: much larger than both the penetration and coherence lengths. For all
566: practical purposes, the function $j(u)$ then essentially vanishes
567: whereas the order parameter retains its canonical value of unity
568: within most of the volume of the sample, except for a small region close
569: to the surface. Hence in the above expressions of criticality,
570: in the limit that $a\rightarrow\infty$, only the contribution in
571: $\kappa^2f^4/2$ tends to dominate, leading in both cases to the
572: condition,
573: \begin{equation}
574: b^2+e^2\simeq \frac{1}{2}\kappa^2\ \ \ ,\ \ \ a\gg\lambda(T),\xi(T)\ .
575: \label{eq:macro1}
576: \end{equation}
577: Since this will prove to be useful, let us normalize the measurement
578: of these fields to the value $b_0$ of the critical magnetic
579: field in the absence of any electric field (in the macroscopic limit,
580: we thus have $b_0=\kappa/\sqrt{2}$). In terms of the physical
581: quantities, one then obtains the following approximation to the
582: criticality condition in the $(B,E)$ phase diagram
583: \begin{equation}
584: \left(\frac{B}{B_0}\right)^2+\left(\frac{E/c}{B_0}\right)^2\simeq 1\ \ \ ,
585: \ \ \ a\gg\lambda(T),\xi(T)\ .
586: \label{eq:macro2}
587: \end{equation}
588: Hence in the macroscopic limit, the two models are not distinguished
589: in their $(B,E)$ phase diagrams. In particular, both their critical
590: magnetic, $B_0$, and electric, $E_0$, fields (in the absence each time
591: of the other field) reach a vanishing value at the critical
592: temperature $T_c$.
593:
594: Consider now the nanoscopic limit, namely when $a\ll\lambda(T),\xi(T)$.
595: In practice, this situation may be encountered indeed for nanoscopic
596: samples close to the critical temperature $T_c$. In such a case,
597: one may develop series expansion solutions in $u$ for the functions $j(u)$
598: and $f(u)$ in order to evaluate the criticality conditions
599: (\ref{eq:critical1}) and (\ref{eq:critical2}). However, since
600: contributions of order $b^2$ and $e^2$ appear on both sides of these
601: equations, in order to be of sufficient accuracy,
602: the expansion in $u$ must at the same time include at least
603: the first order corrections in $b^2$ and $e^2$ in the r.h.s. of
604: (\ref{eq:critical1}) and (\ref{eq:critical2}) as well.
605: For this reason, it is more relevant to consider a weak field
606: expansion for the solutions independently of whether $u_a$ is small or not,
607: to be used to compute to first order in $b^2$ and $e^2$ the r.h.s. of
608: the criticality conditions above, and then
609: eventually take the nanoscopic limit. Note that given the result
610: (\ref{eq:macro1}), critical fields are at least of the order of
611: $\kappa/\sqrt{2}$, so that such a weak field expansion should be
612: warranted for small values of the LG parameter $\kappa$, namely
613: for type I superconductors.
614:
615: After some work, one then finds that the criticality conditions
616: (\ref{eq:critical1}) and (\ref{eq:critical2}),
617: evaluated to first order in $b^2$ and $e^2$, imply the following
618: constraint on the physical fields in the $(B,E)$ phase diagram,
619: \begin{equation}
620: \left(\frac{B}{B_0}\right)^2+C\left(\frac{E/c}{B_0}\right)^2\simeq 1\ ,
621: \label{eq:nano1}
622: \end{equation}
623: where as before $B_0$ stands for the critical magnetic field value
624: in the absence of any electric field, $E=0$, which is in general a function
625: of temperature and of $a$ of course, while $C$ is a factor
626: given by the following expressions,
627: \begin{equation}
628: \begin{array}{r c l}
629: {\rm covariant\ model}&:&\ \ C=\left(\frac{1+\beta}{1-\beta}\right)\ ,\\
630: & & \\
631: {\rm noncovariant\ model}&:&\ \
632: C=\left(\frac{u_a}{u_a-\tanh u_a}\right)
633: \left(\frac{1}{1-\beta}\right)\ ,
634: \end{array}
635: \end{equation}
636: where
637: \begin{displaymath}
638: \beta=\frac{u_a}{16(u_a-\tanh u_a)^2}\frac{1}{(\kappa^2-2)^2}\left\{
639: 8\kappa\sqrt{2}\frac{\tanh^2u_a}{\tanh(\kappa\sqrt{2}u_a)}-
640: (3\kappa^4-10\kappa^2+16)\tanh u_a+\right.
641: \end{displaymath}
642: \begin{equation}
643: \left.+(5\kappa^4-22\kappa^2+16)\tanh^3u_a
644: +(\kappa^2-2)(3\kappa^2-4)\frac{u_a}{\cosh^4u_a}\right\}\ .
645: \label{eq:B}
646: \end{equation}
647:
648: These expressions are valid in the weak field approximation to
649: first order whatever the value for $a$. Taking now the nanoscopic limit
650: as well, one finds $\left(1-\tanh(u_a)/u_a\right)^{-1}=
651: 3/u^2_a\left[1+{\cal O}(u^2_a)\right]$
652: and $\beta=1/2\left[1+{\cal O}(u^2_a)\right]$, leading finally to the
653: following criticality conditions in the $(B,E)$ phase diagram
654: in the weak field limit,
655: \begin{equation}
656: \begin{array}{r c l c l}
657: {\rm covariant\ model}&:&\ \
658: \left(\frac{B}{B_0}\right)^2+3\left(\frac{E/c}{B_0}\right)^2\simeq 1\ &,&
659: \ a\ll\lambda(T),\xi(T)\ ,\\
660: & & & & \\
661: {\rm noncovariant\ model}&:&\ \
662: \left(\frac{B}{B_0}\right)^2+6\left(\frac{\lambda(0)}{a}\right)^2
663: \frac{1}{1-\left(\frac{T}{T_c}\right)^4}\left(\frac{E/c}{B_0}\right)^2\simeq
664: 1\ &,& \ a\ll\lambda(T),\xi(T)\ .
665: \label{eq:nano2}
666: \end{array}
667: \end{equation}
668: Since the critical magnetic field $B_0$ does vanish towards the
669: critical temperature $T=T_c$, so do all the critical fields $B$ and $E$
670: which are defined by either of these relations, and thus in particular also
671: the critical electric field $E_0$ in the absence of any magnetic field,
672: $B=0$, as was already remarked previously in the noncovariant case.
673: However, by having chosen to normalize the measurements of these fields
674: to $B_0$, a very distinctive feature appears for the covariant model when
675: compared to the noncovariant one. Indeed, the ratio $E/(cB_0)$ always retains
676: a finite and nonvanishing value, whatever the critical values for $B$ and
677: $E$ within the intervals $[0,B_0]$ and $[0,E_0]$, {\sl even in the limit
678: of the critical temperature $T_c$\/}, whereas in the noncovariant model,
679: that same ratio $E/(cB_0)$ must vanish like $\sqrt{1-(T/T_c)^4}$ (given our
680: chosen model for $\lambda(T)$). In particular, in the weak field
681: approximation and including the result (\ref{eq:macro1}) valid for
682: macroscopic samples, one thus derives in the covariant case the following
683: bounds for the critical electric field $E_0$,
684: \begin{equation}
685: \sqrt{\frac{1-\beta}{1+\beta}}<\frac{E_0/c}{B_0}<1\ ,
686: \label{eq:boundE0cov1}
687: \end{equation}
688: with $E_0/(cB_0)$ moving towards lower values within that interval
689: when the critical temperature $T_c$ is approached (recall that $\beta$
690: is also temperature dependent through $u_a$). In the nanoscopic limit
691: $a\ll\lambda(T),\xi(T)$, these same bounds reduce to
692: \begin{equation}
693: \frac{1}{\sqrt{3}}<\frac{E_0/c}{B_0}<1\ .
694: \label{eq:boundE0cov2}
695: \end{equation}
696: In contradistinction in the noncovariant case, the lower bound
697: on $E_0/(cB_0)$ always vanishes, since one then finds,
698: \begin{equation}
699: \sqrt{\left(1-\frac{1}{u_a}\tanh u_a\right)(1-\beta)}<\frac{E_0/c}{B_0}<1\ ,
700: \label{eq:boundE0non1}
701: \end{equation}
702: reducing in the nanoscopic limit $a\ll\lambda(T),\xi(T)$ to
703: \begin{equation}
704: \frac{1}{\sqrt{6}}\left(\frac{a}{\lambda(0)}\right)
705: \sqrt{1-\left(\frac{T}{T_c}\right)^4}<\frac{E_0/c}{B_0}<1\ .
706: \label{eq:boundE0non2}
707: \end{equation}
708:
709: As a matter of fact, this type of consideration may be refined further
710: still in the covariant case. Indeed, an obvious solution to the covariant LG
711: equations is $j(u)=\sinh u/\cosh u_a$, $f(u)=1$ in the case that
712: $e_{\rm ext}=b_{\rm ext}$, a fact which, as was remarked previously, is
713: a distinctive feature of the covariant approach, since this solution
714: defines precisely the boundary between the magnetic and electric regimes
715: of superconductivity, and as such its existence is a direct consequence
716: of Lorentz covariance. Hence, the critical condition
717: (\ref{eq:critical1}) for the corresponding fields $b_1$ and $e_1$
718: simplifies in this specific instance to the exact result, valid under all
719: circumstances,
720: \begin{equation}
721: b_1^2+e_1^2=2b_1^2=2e_1^2=\frac{1}{2}\kappa^2
722: \frac{1}{1-\frac{1}{u_a}\tanh u_a}\ .
723: \end{equation}
724: When the weak field approximation is also warranted for the evaluation
725: of the critical magnetic field $B_0$, this result combines
726: with those above to lead to the following bounds,
727: \begin{equation}
728: \frac{1}{\sqrt{2}}\sqrt{1-\beta}<\frac{B_1}{B_0}=\frac{E_1/c}{B_0}<
729: \frac{1}{\sqrt{2}}\ ,
730: \label{eq:boundE1cov1}
731: \end{equation}
732: and in the nanoscopic limit,
733: \begin{equation}
734: \frac{1}{2}<\frac{B_1}{B_0}=\frac{E_1/c}{B_0}<\frac{1}{\sqrt{2}}\ \ \ ,
735: \ \ \ a\ll\lambda(T),\xi(T)\ ,
736: \label{eq:boundE1cov2}
737: \end{equation}
738: whereas in the noncovariant case, one finds similarly in the
739: nanoscopic limit
740: \begin{equation}
741: \frac{1}{\sqrt{1+6\left(\frac{\lambda(0)}{a}\right)^2
742: \frac{1}{1-\left(\frac{T}{T_c}\right)^4}}}<
743: \frac{B_1}{B_0}=\frac{E_1/c}{B_0}<\frac{1}{\sqrt{2}}\ \ \ ,\ \ \
744: \ \ \ a\ll\lambda(T),\xi(T)\ .
745: \label{eq:boundE1non2}
746: \end{equation}
747: Hence here again for those specific configurations such that
748: $B_{\rm ext}=E_{\rm ext}/c$, the lower bound on $B_1/B_0$
749: reaches a vanishing value at the critical temperature in the
750: noncovariant case, whereas that lower bound remains finite and is only
751: mildly temperature dependent in the covariant case.
752:
753: The existence of such finite bounds on the values for $E_1/(cB_1)$
754: as a function of temperature in the covariant case, translates into
755: the following nice characterization in terms of the $(B/B_0,E/(cB_0)$ phase
756: diagram. Indeed, the limits (\ref{eq:boundE1cov2}) (which are more
757: refined in (\ref{eq:boundE1cov1})) imply that the phase boundary
758: curve in that diagram must always cross the diagonal line $B=E/c$ within the
759: interval of $B/B_0$ or $E/(cB_0)$ values defined by these bounds in
760: the covariant model, whatever the value for the temperature (see Fig.1).
761: Such a property is simply not met in the noncovariant model (see Fig.2).
762: Similarly, the lower bounds
763: (\ref{eq:boundE0cov1}) or (\ref{eq:boundE0cov2}) on $E_0/(cB_0)$ imply that,
764: when approaching the critical temperature, the same phase boundary curve
765: at $B/B_0=0$ cannot move below a specific finite value in the covariant model,
766: while it must necessarily do so in the noncovariant one.
767:
768: As a conclusion thus, which should remain valid beyond the
769: specific limits con\-si\-de\-red here, it appears that by choosing
770: to normalize the measurement of critical electric and magnetic fields
771: to the critical magnetic field in the absence of any electric field,
772: for a given nanoscopic sample with this specific geometry of applied
773: fields and by approaching the critical temperature,
774: the $(B,E)$ phase diagram provides the necessary distinctive features
775: which should enable to discriminate experimentally between the covariant
776: and noncovariant mo\-dels, and in any case confirm or invalidate the
777: description offered by the Lorentz covariant LG equations. Indeed, as
778: was remarked previously, the ordinary noncovariant framework is not
779: physically realistic when it comes to nanoscopic samples in the presence
780: of electric fields, since it ignores the partial penetration, albeit
781: small, of the electric field into the sample's surface.
782: The present analysis has concentrated on the weak field approximation,
783: essentially in the nanoscopic limit. Similar distinctive differences
784: between the two models should also exist for larger values of $\kappa$,
785: in ways still to be investigated requiring then a detailed numerical
786: study which is not pursued in this Letter.
787:
788:
789: \vspace{10pt}
790:
791: \noindent{\bf 5. Numerical solutions.}
792: Here, we present the results of the numerical resolution of the
793: LG equations and of the criticality conditions
794: (\ref{eq:critical1}) and (\ref{eq:critical2}) for only
795: one situation, which is close enough both to the discussion of the
796: previous section and to an experimentally realistic situation.
797: Namely, we take the following parameter values
798: \begin{equation}
799: \frac{a}{\lambda(0)}=5\ \ \ ,\ \ \ \kappa=0.02\ .
800: \label{eq:values}
801: \end{equation}
802: Indeed, this value for $\kappa$ is typical for aluminium (Al),
803: while tabulated values of $\lambda(0)$
804: for Al---$\lambda(0)=16-50$ nm with $T_c=1.18$ K---would imply that
805: the slab is then a few hundred nanometers thick,
806: within reach of present lithographic techniques for Al on
807: a SiO$_2$ substrate. Moreover, the critical magnetic field $B^\infty_c(0)$
808: for Al is also on the order of 100 Gauss, so that the required
809: electric field values for a measurement of the $(B,E)$ phase
810: diagram would reach into 3 MV/m, namely 3 V/$\mu$m, certainly
811: also a reasonable range of values for such a nanoscopic device.
812: Of course, compared to the infinite slab model, such a device will
813: be subjected to finite size corrections. Presumably, such corrections
814: would imply that the role played by $\lambda(T)$ and $\kappa$ in our
815: analysis would be replaced by some effective quantities whose values
816: would not differ to a great extent from those of Al in the bulk.
817: Such corrections may be assessed only once a specific device is designed.
818:
819: In Fig.1 (resp. Fig.2), we present the
820: $(B/B_0,E/(cB_0))$ phase diagram for the covariant (resp. noncovariant)
821: model, given the values in (\ref{eq:values}), for a series of temperatures
822: in the range from $T=0$ to $T=T_c$. The general behaviour
823: of the phase diagram as a function of temperature is indeed the one described
824: in the previous section. In particular in the covariant model,
825: and as a function of temperature, the critical electric field values
826: $E_0/(cB_0)$ and $E_1/(cB_0)$ obey the different
827: finite lower (and upper bounds) derived from the analytical discussion,
828: including those given in (\ref{eq:boundE0cov1}) and (\ref{eq:boundE1cov1})
829: when considering the associated values for $\beta$. In contradistinction,
830: in the noncovariant case, the ratio $E/(cB_0)$ reaches a
831: vanishing value when approaching the critical temperature, while in this
832: case as well it way be checked that the different lower bounds
833: (\ref{eq:boundE0non1}), (\ref{eq:boundE0non2}) and
834: (\ref{eq:boundE1non2}) are indeed also obeyed.
835:
836: Such results, as well as the other considerations of this Letter
837: show that it should be possible to experimentally
838: discriminate between the ordinary noncovariant LG equations and
839: the covariant ones advocated here, by determining the critical
840: $(B,E)$ phase diagram of a nanoscopic superconducting sample for
841: temperatures approaching its critical temperature.
842: The geometry of the applied fields is crucial
843: for this purpose, with the external magnetic field parallel to the
844: sample's surface and the external electric field perpendicular to it.
845: By normalizing the measurement of fields to that of the critical magnetic
846: field in the absence of any electric field, distinctive differences
847: between the two approaches are best brought to the fore, and should
848: enable to confirm or invalidate the covariant approach. Moreover,
849: if the experiment should also allow for an absolute calibration
850: of the applied fields, the comparison between the two models may be
851: refined still further by considering the temperature dependency of the
852: critical value for applied magnetic $B$ and electric $E/c$ fields of
853: equal magnitude, this temperature dependency being
854: constrained to lie within a specific interval
855: whose existence is a direct consequence of the manifest Lorentz covariance
856: of the covariant model. We hope to be able to report on such measurements in
857: the future, but lithographic problems have hindered any progress until now.
858:
859: \noindent{\bf Acknowledgements.}
860: J.G. wishes to thank Peter van Nieuwenhuizen and the members of the
861: C.N. Yang Institute for Theoretical Physics at the State University of
862: New York at Stony Brook for their kind hospitality. The work
863: of G.S. is financially supported as a Scientific Collaborator of the
864: ``Fonds National de la Recherche Scientifique" (FNRS, Belgium).
865:
866: \clearpage
867:
868: \newpage
869:
870: \begin{thebibliography}{99}
871:
872: \bibitem{Tinkham} See for example,\\
873: M. Tinkham, {\sl Introduction to Superconductivity\/}, 2nd Edition
874: (McGraw Hill, New York, 1996);\\
875: J.R. Waldram, {\sl Superconductivity of Metals and Cuprates\/}
876: (Institute of Physics, Bristol, 1996).
877:
878: \bibitem{Josephson} B.D. Josephson, {\sl Phys. Lett.\/} {\bf 16} (1965) 242.
879:
880: \bibitem{Bok} J. Bok and J. Klein, {\sl Phys. Rev. Lett.\/}
881: {\bf 20} (1968) 660.
882:
883: \bibitem{Kolacek} J. Kol\'a\u{c}ek and E. Kawate, {\sl Phys. Lett.\/}
884: {\bf A 260} (1999) 300;\\
885: J. Kol\'a\u{c}ek and P. Va\u{s}ek, {\sl Hall voltage
886: sign reversal in type II superconductors\/}, {\tt cond-mat/9911222};\\
887: J. Kol\'a\u{c}ek, P. Lipavsk\'y and V. \u{S}pi\u{c}ka,
888: {\sl Electric field in type II superconductors\/}, {\tt cond-mat/9911283}.
889:
890: \bibitem{Weinberg} S. Weinberg, {\sl The Quantum Theory of Fields\/}
891: (Cambridge University Press, Cambridge, 1996), Vol. II, pp. 332-352.
892:
893: \bibitem{Gov1} J. Govaerts, G. Stenuit, D. Bertrand and O. van der Aa,
894: {\sl Phys. Lett.\/} {\bf A 267 } (2000) 56 ({\tt cond-mat/9908451}).
895:
896: \bibitem{Gov2} J. Govaerts, in preparation.
897:
898: \end{thebibliography}
899:
900: \clearpage
901:
902: \noindent{\bf Figure Captions}
903:
904: \vspace{10pt}
905:
906: \noindent Figure 1: The phase diagram $(B/B_0,E/(cB_0))$ for the
907: covariant LG equations with the va\-lues (\ref{eq:values}). Shown
908: from top to bottom are the curves associated to the following
909: increasing temperature values, $T/T_c=0,0.8766,0.9659,0.9935,0.9996$.
910: The diagonal line determines those configurations such that
911: $B_{\rm ext}=E_{\rm ext}/c$, the two vertical dot-dashed lines
912: at $B/B_0=1/2$ and $B/B_0=1/\sqrt{2}$ correspond to the lower and
913: upper bounds (\ref{eq:boundE1cov2}) obeyed by the critical electric $E_1/c$
914: and magnetic $B_1$ fields of equal strength in the nanoscopic limit
915: of the weak field ap\-pro\-xi\-mation for the covariant LG equations,
916: while the horizontal dashed line at $E/(cB_0)=1/\sqrt{3}$ corresponds to
917: the lower bound (\ref{eq:boundE0cov2}) on the critical electric field
918: $E_0/(cB_0)$ in the same approximation.
919: The existence of these finite bounds is the distinctive prediction
920: of the covariant model and a direct consequence of its manifest Lorentz
921: covariance.
922:
923: \vspace{20pt}
924:
925: \noindent Figure 2: The same as in Fig.1 for the noncovariant
926: model. In this case, the horizontal and two vertical lines are displayed
927: only for the purpose of comparison with the covariant model.
928:
929:
930: \end{document}
931: