\magnification 1200 \vsize 23 truecm \hsize 15truecm \baselineskip 18truept \centerline {\bf Off-diagonal Long Range Order} \vskip 0.5cm \centerline {\bf and Superconductive Electrodynamics} \vskip 1cm \centerline {{by Geoffrey L. Sewell}\footnote{$^{(a)}$}{Electronic mail: G.L.Sewell@qmw.ac.uk}} \vskip 1cm \centerline {Department of Physics, Queen Mary and Westfield College} \vskip 0.5cm \centerline {Mile End Road, London E1 4NS} \vskip 1.5cm \centerline {\bf Abstract} \vskip 0.5cm\noindent \vskip 1.5cm We present a general, model independent, quantum statistical derivation of superconductive electrodynamics from the assumptions of off-diagonal long range order (ODLRO), local gauge covariance and thermodynamic stability. On this basis, we obtain the Meissner and Josephson effects, the quantisation of trapped magnetic flux, and the metastability of supercurrents. A key to these results is that the macroscopic wave function, specified by the ODLRO condition, enjoys the rigidity property that London$^{1,2}$ envisaged for the microstate of a superconductor. \vskip 2cm PACS Numbers 74.20.-z; 64.60.My; 05.30.-d \vfill\eject \centerline {\bf I.Introduction} \vskip 0.2cm The object of this article is to provide a model-independent derivation of the electromagnetic properties of superconductors from their order structure, the essential imput being the assumptions of off-diagonal long range order, gauge covariance and thermodynamic stability. Thus, our approach to the subject is centred on very general principles, and is therefore at the opposite pole from that provided by the computational techniques of many-body theory. \vskip 0.2cm In order to explain the need for such an approach, let us first recall that, in the case of metallic superconductivity, the electron pairing hypothesis$^{3,4}$ has been amply substantiated, at an empirical level, by the the measured values of the magnetic flux quantisation$^5$ and of the Josephson tunnelling frequency$^6$. Furthermore, the widely accepted microscopic theory of superconductivity, devised by Bardeen, Cooper and Schrieffer (BCS)$^7$ on the basis of this hypothesis, provides an accurate picture of the thermodynamics$^{8-10}$ of superconductors, especially of their second order phase transitions. On the other hand, the electrodynamics of this theory is seriously flawed in that, firstly, it violates the principle of local gauge covariance$^{11,12},$ and consequently does not even admit a precise definition of the local current density; and, secondly, there is no indication that its ansatz for current-carrying states have the metastability properties$^{13,14}$ of supercurrents. In fact, the violation of the gauge principle by the BCS theory arises from its {\it truncation} of a fully gauge invariant model (Fr\"ohlich's electron-phonon system$^{15}$), that retains only the interactions that give rise to the electron pairing. Attempts$^{16,17}$ to remedy the situation by taking some account of the residual interactions have led to derivations of the Meissner effect that are only {\it approximately} gauge covariant. Since exact gauge covariance is required for a consistent electrodynamics, this is no solution of the problem. As regards ceramic, i.e. high $T_{c},$ superconductors, the microscopic theory is less developed, and has certainly not led to an electrodynamics, even though various interesting ideas concerning its underlying quantum mechanisms have been proposed, e.g. in Refs. 18-23. \vskip 0.2cm Thus, there is a need for a quantum-based gauge covariant electrodynamics of the superconducting phase. Since, at an empirical level, this electrodynamics has such sharply defined {\it qualitative} characteristics, common to the vast variety of superconducting materials, a corresponding quantum theory should surely isolate their origins in general, qualitative terms. We remark here that the traditional techniques of many-body theory$^{24,25}$ are unsuited to this purpose, since these are designed for essentially approximative calculations rather than precise classifications. \vskip 0.2cm In view of these considerations, we have adopted a different approach$^{26-28}$ to superconductive electrodynamics, based on the hypothesis of {\it off-diagonal long range order} (ODLRO), which encapsulates the qualitative features of the electron pairing in a gauge covariant way. This hypothesis was proposed by Yang$^{29}$ as a characterisation of the structure of the superconducting phase, following a similar proposal by O.Penrose$^{30,31}$ in connection with the theory of superfluid Helium. Formally, the ODLRO condition may be expressed in terms of the relevant quantised matter field ${\psi}$ by the formulae $$\lim_{{\vert}y{\vert}\to{\infty}}{\lbrack}{\langle} {\psi}(x){\psi}^{\star}(x+y){\rangle}-{\Phi}(x) {\Phi}^{\star}(x+y){\rbrack}=0 \ for \ bosons, \eqno(1.1a)$$ and $$\lim_{{\vert}y{\vert}\to{\infty}}\bigl[{\langle} {\psi}_{\uparrow}(x_{1}) {\psi}_{\downarrow}(x_{2}) {\psi}_{\downarrow}^{\star}(x_{2}^{\prime}+y) {\psi}_{\uparrow}^{\star}(x_{1}^{\prime}+y){\rangle} -{\Phi}(x_{1},x_{2}){\Phi}^{\star}(x_{1}+y,x_{2}+y)\bigr]$$ $$=0 \ for \ fermions, \eqno(1.1b)$$ where, in both cases, ${\Phi}$ is a complex-valued function, often termed the {\it macroscopic wave function}, such that ${\Phi}(x+y),$ or ${\Phi}(x_{1}+y,x_{2}+y),$ does not tend to zero as ${\vert}y{\vert}{\rightarrow}{\infty}.$ In the bosonic case, ODLRO is simply a generalisation of Bose- Einstein condensation$^{30,31}$ to interacting systems. \vskip 0.2cm We note here that the ODLRO condition is fulfilled not only by the BCS ansatz, but also some of those for ceramic superconductivity$^{19,21}$, as well as that of Feynman$^{32}$ for superfluid Helium (cf Ref. 31). It is also satisfied in the low temperature phases of a number of tractable, gauge covariant models, namely the ideal Bose gas$^{33}$, the hard sphere Bose fluid on a lattice$^{34}$, and, at least at zero temperature, the Hubbard model with {\it attractive} interaction between electrons on the same site$^{35}$. \vskip 0.2cm Our project, then, is to derive the principal electrodynamic properties of superconductors from precisely specified assumptions of ODLRO, gauge covariance and thermodynamic stability. In fact, progress towards this objective has already been achieved in Refs. 26-28 and 36, which have provided derivations of the Meissner effect$^{26},$ the quantisation of trapped magnetic flux$^{27,36},$ and the Josephson effect$^{27},$ as well as a sketched approach to the theory of persistent currents$^{28}$. \vskip 0.2cm The present article will be devoted to a review and further development of this work. Our essential aim will be to provide a coherent account of the way in which the electrodynamics of superconductors, including the metastabilty of the supercurrents, stems from their order structure. We shall keep the mathematics quite simple, formulating the theory within the standard second quantisation framework of condensed matter physics. We remark here that one may put the whole theory onto a completely rigorous basis by recasting it, as in Ref. 27, within the framework of operator algebraic statistical mechanics. \vskip 0.2cm We shall organise the presentation of the theory as follows. We start, in Sec. II, with a general formulation of our gauge-covariant model of matter in interaction with a classical magnetic field. This formulation covers the cases of both continuous and lattice systems. A key result here is that the dynamics of the model, in the presence of a uniform magnetic induction $B,$ is covariant with respect to the {\it regauged space translations}$^{37,26}$ $${\psi}(x){\rightarrow}{\psi}(x+a){\exp}\bigl({-ie(B{\times}a).x \over 2{\hbar}c}\bigr),\eqno(1.2)$$ where $a$ is an arbitrary displacement and the sinusoidal factor arises from the corresponding regauging of the magnetic vector potential. \vskip 0.2cm In Sec. III, we employ this result to derive the Meissner effect, as in Ref. 26. The essential point is that the formula (1.2) for regauged space translations, together with the assumption of ODLRO, leads to the conclusion that {\it either} ${\Phi}$ {\it or} $B$ must vanish. The requirement of thermodynamic stability of the ordered phase then implies that the latter condition prevails, for sufficiently weak applied fields. Thus, we have a Meissner effect, since normally diamagnetic systems can accommodate non-zero uniform induction. The key to this result is that the macroscopic wave function exhibits a rigidity, in the face of the applied field, of the kind envisaged by London$^1$ at a more microscopic level. It is worth remarking here that the principle of local gauge covariance, which was an obstacle to the previous theories, is an essential ingredient of our argument leading to the Meissner effect! \vskip 0.2cm In Sec. IV, we extend that argument, along the lines of Ref. 36, to derive the quantisation of trapped magnetic flux in multiply-connected systems. \vskip 0.2cm In Sec. V, we formulate the theory of persistent currents in a multiply-connected body, such as a ring. Here, the supercurrents are none other than the currents that implement the Meissner effect by screening the trapped magnetic flux from the interior of the body$^2,$ and the essential problem is that of the stabilisation of both the trapped flux and its screening current. In fact, this is a problem of {\it metastability}, since the current-carrying state has higher free energy than that of true thermal equilibrium at the same temperature. Thus, adopting our earlier characterisation$^{14}$ of metastable states by thermodynamic stability against strictly local, rather than global, disturbances, we formulate the condition for the persistence of currents in terms of a variational principle, that is subject to the constraint of the flux quantisation. In this way, we characterise the phenomenon of superconductivity itself, i.e. the persistence of currents, by a relatively simple thermodynamic assumption, together with that of ODLRO. Furthermore, we show that this phenomenon is intimately related to a superselection rule, that forbids locally induced transitions between states with different flux quantum numbers. \vskip 0.2cm In Sec. VI, we formulate the Josephson effect and derive its tunnelling frequency form the properties of the macroscopic wave function. \vskip 0.2cm We conclude, in Sec. VII, with a brief discussion of open problems of the theory. \vskip 0.5cm \centerline {\bf II. The General Model} \vskip 0.2cm Our model, ${\Sigma},$ is an infinitely extended system, consisting of electrons and possibly another species of particles, e.g. phonons or ions, in a space, $X,$ which may be either a Euclidean continuum or a lattice. Points in $X$ will usually be denoted by $x,$ sometimes by $y,a$ or $b.$ We shall denote the electronic component of ${\Sigma}$ by ${\Sigma}_{el},$ and the other component, if any, by ${\Sigma}^{\prime}.$ We shall assume that the dynamics of the model is covariant w.r.t. gauge transformations of both first and second kind, space translations and time reversals. These assumptions represent general demands of quantum mechanics and electromagnetism, and are fulfilled by particular models, such as those of Fr\"ohlich$^{15}$ and Hubbard$^{38}$, on which theories of metallic and high $T_{c}$ superconductivity, respectively, are based. \vskip 0.2cm For simplicity, we shall generally employ a notation appropriate to the case where $X$ is a continuum. This may easily be translated into the corresponding one for lattice case by standard procedures employed in gauge theories$^{39}$. \vskip 0.2cm We shall describe the electronic subsystem ${\Sigma}_{el}$ in terms of a quantised field, $${\psi}=\pmatrix{{\psi}_{1}\cr {\psi}_{-1}\cr}{\equiv} \pmatrix{{\psi}_{\uparrow}\cr {\psi}_{\downarrow}\cr}, \eqno(2.1)$$ which satisfies the canonical anticommutation relations $$[{\psi}_{s}(x),{\psi}_{s^{\prime}}^{\star}(x^{\prime})]_{+}= {\delta}_{ss^{\prime}}{\delta}(x-x^{\prime}); \ [{\psi}_{s}(x),{\psi}_{s^{\prime}}(x^{\prime})]_{+}=0\eqno(2.2)$$ We shall sometimes use the symbol ${\psi}^{\#}$ to denote either ${\psi}$ or ${\psi}^{\star}.$ \vskip 0.2cm The observables$^{40}$ of ${\Sigma}_{el}$ are formed from the polynomials in ${\psi}$ and ${\psi}^{\star}$ that are invariant under gauge transformations of the first kind, ${\psi}{\rightarrow}{\psi}e^{i{\alpha}},$ with ${\alpha}$ constant. Thus, they are generated algebraically by operators of the form $${\psi}_{s_{1}}^{\star}(x_{1}).. \ .{\psi}_{s_{n}}^{\star}(x_{n}){\psi}_{s_{n+1}}(x_{n+1}).. \ .{\psi}_{s_{2n}}(x_{2n}).$$ \vskip 0.2cm We shall be concerned with the properties of the the system in the presence of a classical electromagnetic field $(E,B),$ represented by a scalar potential ${\phi}$ and a vector potential, $A:$ thus, $E=-{\nabla}{\phi}-c^{- 1}{\partial}A/{\partial}t$ and $B=curlA.$ We assume that the dynamics of the model is covariant w.r.t. gauge transformations of the second kind, as given by the formula $$A{\rightarrow}A+{\nabla}{\chi}, \ {\phi}{\rightarrow} {\phi}-c^{-1}{{\partial}{\chi}\over {\partial}t}, \ {\psi}{\rightarrow}{\psi}{\exp}({ie{\chi}\over {\hbar}c}) \eqno(2.3),$$ where ${\chi}$ is an arbitrary function of position and time. We assume that the observables representing the position-dependent densities of electronic charge, current and magnetic polarisation in the presence of this field are given by the standard formulae $${\rho}=-e{\psi}^{\star}{\psi}; \ j(x)= {ie\over 2}({\psi}^{\star}{\nabla}{\psi}- ({\nabla}{\psi}^{\star}){\psi})+{e\over c}A{\psi}^{\star}{\psi}; \ and \ m={e{\hbar}\over mc}{\psi}^{\star}{\sigma}{\psi}, \eqno(2.4)$$ where $-e$ is the electronic charge and ${\sigma}$ is the spin vector, whose components $({\sigma}_{1},{\sigma}_{2},{\sigma}_{3})$ are the Pauli matrices. We note here that ${\rho}, \ j$ and $m$ are invariant w.r.t. all gauge transformations. \vskip 0.2cm We assume that the formal Hamiltonian of the model is of the form $$H={\int}({\hbar}{\nabla}{\psi}^{\star}+ {ie\over c}A{\psi}^{\star}).({\hbar}{\nabla}{\psi}-{ie\over c}A)dx- {\int}(B.m+{\rho}{\phi})dx+V_{el}+H_{int}+H^{\prime},\eqno(2.5)$$ where $V_{el}$ is the potential energy of the interelectronic interactions, $H^{\prime}$ is the Hamiltonian for ${\Sigma}^{\prime}$ and $H_{int}$ is the energy of interaction between ${\Sigma}_{el}$ and ${\Sigma}^{\prime}.$ We assume that these three contributions to $H$ are all independent of the potentials ${\phi}$ and $A,$ and thus that $H$ is gauge invariant. Further, we stipulate that the magnetic interactions between the electrons, that stem from the sources $j(x)$ and $m(x)$ according to Maxwell's equations, are not incorporated into either $V_{el}$ or $H_{int},$ but are represented by the dependence of $B$ on these sources (cf. Comment at the end of this Section). In the theory that follows, we shall confine our considerations to situations where $B$ is static, $E=0,$ and, except in Sec. 6, we shall take it that ${\phi}=0$ and ${\chi}$ is time-independent. \vskip 0.2cm We shall assume that the dynamics is covariant w.r.t. space translations and time reversals. The transformations of the fields ${\psi}, \ A,$ corresponding to space translations $a({\in}X),$ are given by $$A(x){\rightarrow}(x+a), \ {\psi}(x){\rightarrow}{\psi}(x+a).\eqno(2.6)$$ The operation of time reversal serves to transform ${\psi}_{s}, \ {\psi}_{s}^{\star}$ and $A$ to ${\psi}_{-s}^{\star}, \ {\psi}_{-s}$ and $-A,$ respectively, and to invert the order of the terms in the operator products. Thus, its effective action on the electronic observables and vector potential is given by $${\psi}_{s_{1}}^{\#}(x_{1}). \ .{\psi}_{s_{n}}^{\#}(x_{n}) {\rightarrow}{\psi}_{-s_{n}}^{{\#}{\star}}(x_{n}). \ . {\psi}_{-s_{1}}^{{\#}{\star}}(x_{1}); \ A{\rightarrow} -A.\eqno(2.7)$$ \vskip 0.2cm Specialising now to the case where the magnetic induction $B$ is uniform and so may be represented by the vector potential $A(x)={1\over 2}B{\times}x,$ and choosing ${\chi}(x)=- (B{\times}x).a,$ we have the relation $A(x)+{\nabla}{\chi}(x)={\chi}(x-a).$ Hence, by (2.3) and (2.6), the dynamics of ${\Sigma}$ is covariant w.r.t. the transformation$^{37,26}$ $${\psi}(x){\rightarrow}{\psi}_{a}(x){\equiv} {\psi}(x+a){\exp}({-ie(B{\times}x).a)\over e{\hbar}c}), \ A(x){\rightarrow}A(x),\eqno(2.8)$$ together with the transformation of the ${\Sigma}^{\prime}-$ observables corresponding to space translations. Since (2.6) consists of a space translation, compensated by a gauge transformation in such a way as to leave $A$ unchanged, we term it a {\it regauged space translation}. \vskip 0.2cm We denote by $Q_{t}$ the time-translate, in Heisenberg representation, of an arbitrary observable, $Q,$ of ${\Sigma}.$ A dynamical characterisation of thermal equilibrium states of the system at inverse temperature ${\beta}$ is given by the Kubo- Martin-Schwinger (KMS) condition$^{41,42}$, i.e., $${\langle}Q_{t}Q^{\prime}{\rangle}_{A}= {\langle}Q^{\prime}Q_{t+i{\hbar}{\beta}}{\rangle}_{A}, \eqno(2.9)$$ for arbitrary observables $Q$ and $Q^{\prime},$ where the angular brackets denote expectation value and the suffix $A$ indicates its dependence on $A$. Most importantly, this condition is valid even for infinite systems$^{43}$, where the traditional Gibbs canonical formulation is not directly applicable. In the case where $B$ is uniform, the covariance of the dynamics w.r.t regauged space translations ensures that the model supports equilibrium states that are invariant under these translations. \vskip 0.2cm Turning now to the thermodynamics of the model, we define $f(B,T)$ to be its global free energy density at temperature $T,$ under the action of the uniform induction, $B.$ This function can be formulated by standard statistical mechanical methods$^{44}$ in terms of the microscopic description of ${\Sigma},$ with $B$ taken to be a given control variable. \vskip 0.2cm On the other hand, this induction is {\it not} given, a priori, in the physical situation where ${\Sigma}$ is subjected to a magnetic field, $H_{ex},$ due to a fixed source of current density $J_{ex}.$ For, in this situation, the induction, $B,$ and the equilibrium state of ${\Sigma}$ co-determine one another. \vskip 0.2cm To formulate the thermodynamics of ${\Sigma}$ under these conditions, we have to incorporate the energy of the field, $B,$ and of its interaction with the source, $J_{ex},$ into the picture. Thus, we have to consider the system under consideration to comprise the composite, ${\tilde {\Sigma}},$ of ${\Sigma}$ and the field $B.$ The contribution to the internal energy of ${\tilde {\Sigma}}$ due to the induction, $B,$ and its coupling to the source is then $${\int}({1\over 2}B^{2}-c^{-1}A.J_{ex})dx,$$ and, in view of the Maxwell equation $curlH_{ex}=c^{-1}J_{ex},$ this reduces to $${\int}({1\over 2}B^{2}-H_{ex}.B)dx.\eqno(2.10)$$ Thus, in the case where $B$ and $H_{ex}$ are uniform, the free energy density of ${\tilde {\Sigma}}$ is $${\phi}(B,T,H_{ex})=f(B,T)+{1\over 2}B^{2}-B.H_{ex}.\eqno(2.11)$$ The equilibrium value of $B,$ corresponding to given $H_{ex}$ and $T,$ is then obtained by minimising ${\phi},$ and the resultant Gibbs free energy density is $${\tilde f}(H_{ex},T)=min_{B}{\phi}(B,T,H_{ex}).\eqno(2.12)$$ \vskip 0.2cm {\bf Comment.} This formulation of the thermodynamics of ${\tilde {\Sigma}}$ is, of course, semi-classical, since the field $B$ is treated as classical. In this picture, the field energy (2.10) stems from the magnetic interactions of the currents and spins in ${\Sigma}$ both with one another and with the external source. For a discussion of the problem of a fully quantum formulation, see Ref. 27, Sec.4. \vfill\eject \centerline {\bf III. ODLRO and the Meissner Effect.} \vskip 0.2cm In order to formulate ODLRO, we introduce the {\it pair field} $${\Psi}(x_{1},x_{2})={\psi}_{\uparrow}(x_{1}) {\psi}_{\downarrow}(x_{2}).\eqno(3.1)$$ We then say that a state of ${\Sigma}$ possesses the property of ODLRO if there is a {\it classical} two-point field ${\Phi}(x_{1},x_{2})$ such that$^{29}$ $$\lim_{{\vert}y{\vert}\to{\infty}}\bigl[{\langle} {\Psi}(x_{1},x_{2}) {\Psi}^{\star}(x_{1}^{\prime}+y,x_{2}^{\prime}+y){\rangle}_{A}- {\Phi}(x_{1},x_{2}) {\Phi}^{\star}(x_{1}^{\prime}+y,x_{2}^{\prime}+y)\bigr]=0,\eqno(3.2)$$ and, further, ${\Phi}(x_{1}+y,x_{2}+y),$ does not tend to zero as ${\vert}y{\vert}{\rightarrow}{\infty}.$ In this case, ${\Phi}$ is termed the {\it macroscopic wave function} of the state. \vskip 0.2cm {\bf Note.} Although ${\Psi}$ is not an an observable, in the sense specified in Sec. II, the quantity in angular brackets in the ODLRO condition (3.2) is. \vskip 0.2cm In order to relate ODLRO to superconductive electrodynamics, we note that the essential distinction between normal diamagnetism and the Meissner effect is that the former can support a uniform, static, non-zero magnetic induction and the latter cannot. Thus, we base our derivation of the Meissner effect on considerations of the response of a state possessing ODLRO to the action of a uniform magnetic field. \vskip 0.2cm The following two Propositions, which we shall prove at the end of this Section, provide us with the key to the relationship between ODLRO and the electromagnetic properties of our model. \vskip 0.2cm {\bf Proposition 3.1.} {\it (1) The ODLRO condition (3.2) defines the macroscopic wave function ${\Phi},$ up to a phase factor, i.e., if ${\Phi}_{1}$ and ${\Phi}_{2}$ are two such functions which satisfy this condition for the same state, then ${\Phi}_{1}=e^{i{\alpha}}{\Phi}_{2},$ where ${\alpha}$ is some real constant. \vskip 0.2cm (2) In the case where $A=0$ and where the state is invariant with respect to time reversals, the function ${\Phi}$ is real-valued, up to a constant phase factor.} \vskip 0.2cm {\bf Proposition 3.2.} {\it If the system is in a translationally invariant state possessing the property of ODLRO, then \vskip 0.2cm (1) in the case where $X$ is a continuum, the magnetic induction $B$ vanishes, and \vskip 0.2cm (2) in the case where $X$ is a lattice, $B$ is restricted to the discrete set of values ${\lbrace}B^{(n)}{\rbrace},$ given by the condition that $B^{(n)}.(a{\times}b)$ is an integral multiple of ${\pi}e/{\hbar}c,$ for any $a, \ b$ in $X.$ Thus, the vectors $B^{(n)}$ are the sites of an associated lattice, ${\cal B},$ defined by this condition.} \vskip 0.2cm {\bf Comments.} (1) Prop. 3.2 demonstrates that, in the continuum case, translationally invariant (including equilibrium) states with ODLRO do not admit uniform magnetic fields, i.e., they exhibit the Meissner effect. \vskip 0.2cm (2) In the case where $X$ is a lattice, this conclusion must be modified by the possibility that ODLRO might be compatible with a non-zero quantised induction $B^{(n)}.$ \vskip 0.2cm (3) The question of whether ODLRO, with or without the quantised magnetic induction $B^{(n)},$ prevails is a thermodynamic one. \vskip 0.2cm (4) The removal of ODLRO by even infinitessimal changes in $B$ from the value $0$ or, in the lattice case, $B^{(n)}$ suggests the advent of a phase transition. \vskip 0.2cm (5) In the case of lattice systems, the non-zero $B^{(n)}$'s are of the order of ${\hbar}c/el^{2},$ where $l$ is the spacing of the lattice $X.$ Thus, for typical values of $l,$ e.g. $10^{-8}$cm., they are of the order of $10^{9}$G, which is not only many orders of magnitude larger than any known critical fields for superconductors, but also much larger than the internal fields in ferromagnets. \vskip 0.2cm We base our treatment of the thermodynamics and phase structure of the system, which is evidently needed in view of Comment (3), on the following assumptions. \vskip 0.2cm {\bf (III.1)} {\it (a) The equilibrium state of ${\Sigma}$ corresponding to any given $(A,T)$ is unique and therefore translationally invariant. \vskip 0.2cm (b) Similarly, the equilibrium state of ${\tilde {\Sigma}}$ corresponding to given $(H_{ex},T)$ is unique and translationally invariant.} \vskip 0.2cm {\bf Note.} This assumption precludes the applicability of our treatment to the mixed phase of type II superconductors, since this breaks the space translational symmetry of the system. \vskip 0.2cm The next assumption, prompted by the above Comment (5), excludes ODLRO states with non-zero quantised induction $B^{(n)}$ from the theory, and thus puts the results of Prop. 3.2 for continuous and lattice systems onto the same footing. \vskip 0.2cm {\bf (III.2)} {\it Even in the case of lattice systems, the equilibrium states of ${\tilde {\Sigma}}$ that possess the property of ODLRO carry no magnetic induction.} \vskip 0.2cm The next assumption follows the line suggested by the above Comment (4). \vskip 0.2cm {\bf (III.3)} {\it (a) For $B=0,$ the system ${\Sigma}$ undergoes a phase transition at a temperature $T_{c},$ such that, for $T