This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Thermal Stability of Superconductors

Jacob Szeftel1 [email protected]    Nicolas Sandeau2    Michel Abou Ghantous3    Muhammad El-Saba4 1ENS Cachan, LPQM, 61 avenue du Président Wilson, 94230 Cachan, France 2Aix Marseille Univ, CNRS, Centrale Marseille, Institut Fresnel, F-13013 Marseille, France 3American University of Technology, AUT Halat, Highway, Lebanon 4Ain-Shams University, Cairo, Egypt
Abstract

A stability criterion is worked out for the superconducting phase. The validity of a prerequisite, established previously for persistent currents, is thereby confirmed. Temperature dependence is given for the specific heat and concentration of superconducting electrons in the vicinity of the critical temperature TcT_{c}. The isotope effect, mediated by electron-phonon interaction and hyperfine coupling, is analyzed. Several experiments, intended at validating this analysis, are presented, including one giving access to the specific heat of high-TcT_{c} compounds.

pacs:
74.25.Bt

I Introduction

In the mainstream viewpar ; sch ; tin , the thermal properties of superconductors are discussed within the framework of the phenomenological equation by Ginzburg and Landaugin (GL) and the BCS theorybcs . However, since this work is aimed at accounting for the stability of the superconducting state with respect to the normal one, we shall develope an alternative approach, based on thermodynamicslan , the properties of the Fermi gasash and recent resultssz5 ; sz4 , claimed to be valid for all superconductors, including low and high TcT_{c} materials.

The outline is as follows : the specific heat of the superconducting phase is calculated in section 22, which enables us to assess its binding energy and thereby to confirm and refine a necessary condition, established previously for the existence of persistent currentssz4 ; section 33 is concerned with the inter-electron coupling, mediated by the electron-phonon and hyperfine interactions; new experiments, dedicated at validating this analysis, are discussed in section 44 and the results are summarised in the conclusion.

II Binding Energy

As in our previous worksz4 ; sz5 ; sz1 ; sz2 ; sz3 ; sz7 , the present analysis will proceed within the framework of the two-fluid model, for which the conduction electrons comprise bound and independent electrons, in respective temperature dependent concentration cs(T),cn(T)c_{s}(T),c_{n}(T). They are organized, respectively, as a many bound electronsz5 (MBE) state, characterised by its chemical potential μ(cs)\mu(c_{s}), and a Fermi gasash of Fermi energy EF(T,cn)E_{F}(T,c_{n}). The Helmholz free energy of independent electrons per unit volume FnF_{n} and EFE_{F} on the one hand, and the eigenenergy per unit volume s(cs)\mathcal{E}_{s}(c_{s}) of bound electrons and μ\mu on the other hand, are relatedash ; lan , respectively, by EF=FncnE_{F}=\frac{\partial F_{n}}{\partial c_{n}} and μ=scs\mu=\frac{\partial\mathcal{E}_{s}}{\partial c_{s}}. At last, according to Gibbs and Duhem’s lawlan , the two-fluid model fulfilssz4 at thermal equilibrium

EF(T,cn(T))=μ(cs(T)),c0=cs(T)+cn(T),E_{F}(T,c_{n}(T))=\mu(c_{s}(T)),\quad c_{0}=c_{s}(T)+c_{n}(T), (1)

with c0c_{0} being the total concentration of conduction electrons. Solutions of Eq.(1) are given for T=0,TcT=0,T_{c} in Fig.1. Besides, Eq.(1) has been shownsz5 ; sz4 to read for T=TcT=T_{c} (see BB in Fig.1)

EF(Tc,c0)=μ(cs=0)=εc/2,E_{F}(T_{c},c_{0})=\mu(c_{s}=0)=\varepsilon_{c}/2\quad, (2)

with εc\varepsilon_{c} being the energy of a bound electron pairsz5 . Note that Eq.(2) is consistent with the superconducting transition, occuring at TcT_{c}, being of second orderlan , whereas it has been shownsz5 to be of first order at T<TcT<T_{c}, if the sample is flown through by a finite current. Then the binding energy of the superconducting state Eb(T<Tc)E_{b}(T<T_{c}) has been workedsz5 ; gor out as

Eb(T)=TTc(Cs(u)Cn(u))𝑑u,E_{b}(T)=\int_{T}^{T_{c}}\left(C_{s}(u)-C_{n}(u)\right)du\quad, (3)

with Cs(T),Cn(T)=(πkB)23ρ(EF)T,C_{s}(T),C_{n}(T)=\frac{\left(\pi k_{B}\right)^{2}}{3}\rho(E_{F})T, being the electronic specific heat of a superconductor, flown through by a vanishing currentsz5 and that of a degenerate Fermi gasash (kB,ρ(EF)k_{B},\rho(E_{F}) stand for Boltzmann’s constant and the one-electron density of states at the Fermi energy). Due to Eq.(3), a stable superconducting phase Eb>0\Leftrightarrow E_{b}>0 requires Cs(T)>Cn(T)C_{s}(T)>C_{n}(T), which is confirmed experimentallypar ; ash , namely Cs(Tc)3Cn(Tc)C_{s}(T_{c})\approx 3C_{n}(T_{c}).

Refer to caption
Figure 1: Schematic plots of EF(T=0,cn)E_{F}(T=0,c_{n}), EF(Tc,cn)E_{F}(T_{c},c_{n}), EF(T>Tc,cn)E_{F}(T>T_{c},c_{n}) and μ(cs)\mu(c_{s}) as solid, dashed-dotted, dotted and dashed lines, respectively; μcs\frac{\partial\mu}{\partial c_{s}} has been taken to be constant for simplicity; the origin EF=μ=0E_{F}=\mu=0 is set at the bottom of the conduction band; the crossing points A,BA,B of EF(0,cn),EF(Tc,cn)E_{F}(0,c_{n}),E_{F}(T_{c},c_{n}), respectively, with μ(cs)\mu(c_{s}), exemplify stable solutions of Eq.(1); the tiny differences EF(T,cn)μ(c0cn)E_{F}(T,c_{n})-\mu(c_{0}-c_{n}) have been hugely magnified for the reader’s convenience.

The bound and independent electrons contribute, respectively,

s(cs)=0csμ(u)𝑑un(T,cn)=ϵbϵuϵρ(ϵ)f(ϵEF,T)𝑑ϵ,\begin{array}[]{l}\mathcal{E}_{s}(c_{s})=\int_{0}^{c_{s}}\mu(u)du\\ \mathcal{E}_{n}(T,c_{n})=\int_{\epsilon_{b}}^{\epsilon_{u}}\epsilon\rho(\epsilon)f(\epsilon-E_{F},T)d\epsilon\end{array}\quad,

to the total electronic energy =n+s\mathcal{E}=\mathcal{E}_{n}+\mathcal{E}_{s}. The symbols ϵ,f(ϵEF,T)\epsilon,f(\epsilon-E_{F},T) refer to the one-electron energy and Fermi-Dirac distribution, while ϵb,ϵu\epsilon_{b},\epsilon_{u} designate the lower and upper limits of the conduction band. Then, thanks to Eq.(1) (dcn+dcs=dEFdμ=0\Rightarrow dc_{n}+dc_{s}=dE_{F}-d\mu=0), Cs(T)=ddTC_{s}(T)=\frac{d\mathcal{E}}{dT} is inferred to read

Cs=nTEFcnT+dEFdT(nEFEFcnEF),C_{s}=\frac{\partial\mathcal{E}_{n}}{\partial T}-E_{F}\frac{\partial c_{n}}{\partial T}+\frac{dE_{F}}{dT}\left(\frac{\partial\mathcal{E}_{n}}{\partial E_{F}}-E_{F}\frac{\partial c_{n}}{\partial E_{F}}\right)\quad, (4)

with cn=cn(T),cs=cs(T),EF=EF(T,cn(T))c_{n}=c_{n}(T),c_{s}=c_{s}(T),E_{F}=E_{F}\left(T,c_{n}(T)\right). Because the independent electrons make up a degenerate Fermi gas (T<<TF=EF/kB\Rightarrow T<<T_{F}=E_{F}/k_{B}), the following expressions can be obtained owing to the Sommerfeld expansionash up to T2T^{2}

nEF=EFρ+(2ρ+EFρ′′)(πkBT)26nT=(ρ+EFρ)(πkB)23TcnEF=ρ+ρ′′(πkBT)26,cnT=ρ(πkB)23T,\begin{array}[]{c}\frac{\partial\mathcal{E}_{n}}{\partial E_{F}}=E_{F}\rho+\left(2\rho^{\prime}+E_{F}\rho^{\prime\prime}\right)\frac{\left(\pi k_{B}T\right)^{2}}{6}\\ \frac{\partial\mathcal{E}_{n}}{\partial T}=\left(\rho+E_{F}\rho^{\prime}\right)\frac{\left(\pi k_{B}\right)^{2}}{3}T\\ \frac{\partial c_{n}}{\partial E_{F}}=\rho+\rho^{\prime\prime}\frac{\left(\pi k_{B}T\right)^{2}}{6}\quad,\quad\frac{\partial c_{n}}{\partial T}=\rho^{\prime}\frac{\left(\pi k_{B}\right)^{2}}{3}T\end{array}\quad, (5)

with ρ=ρ(EF),ρ=dρdEF(EF),ρ′′=d2ρdEF2(EF)\rho=\rho(E_{F}),\rho^{\prime}=\frac{d\rho}{dE_{F}}(E_{F}),\rho^{\prime\prime}=\frac{d^{2}\rho}{dE_{F}^{2}}(E_{F}). Then Eq.4 is finally recast into

Cs(T)=(πkB)23ρT(1+dEFdTρρT).C_{s}(T)=\frac{\left(\pi k_{B}\right)^{2}}{3}\rho T\left(1+\frac{dE_{F}}{dT}\frac{\rho^{\prime}}{\rho}T\right)\quad. (6)

Applying Eq.6 at T=TcT=T_{c} yields

Cs(Tc)=Cn(Tc)(1+dEFdT(Tc)ρρTc).C_{s}(T_{c})=C_{n}(T_{c})\left(1+\frac{dE_{F}}{dT}(T_{c}^{-})\frac{\rho^{\prime}}{\rho}T_{c}\right)\quad. (7)

Hence it is in order to work out the expressions of dEFdT(T>Tc)\frac{dE_{F}}{dT}(T>T_{c}) and dEFdT(TTc)\frac{dE_{F}}{dT}(T\leq T_{c}).

Due to cn(T>Tc)=c0c_{n}(T>T_{c})=c_{0}, dEFdT\frac{dE_{F}}{dT} is deducedash to read

dEFdT(T>Tc)=cnTcnEF=(πkB)23ρρT,\frac{dE_{F}}{dT}(T>T_{c})=-\frac{\frac{\partial c_{n}}{\partial T}}{\frac{\partial c_{n}}{\partial E_{F}}}=-\frac{\left(\pi k_{B}\right)^{2}}{3}\frac{\rho^{\prime}}{\rho}T\quad, (8)

which is integrated with respect to TT to yield

EF(T=0,c0)EF(T,c0)=(πkB)26ρρT2.E_{F}(T=0,c_{0})-E_{F}(T,c_{0})=\frac{\left(\pi k_{B}\right)^{2}}{6}\frac{\rho^{\prime}}{\rho}T^{2}\quad. (9)

Then consistency with Fig.1 requires ρ(EF)>0\rho^{\prime}(E_{F})>0 so that CC goes toward BB for TTcT\searrow T_{c}. Assuming ρ(ϵ)=ρf(ϵ)ϵρf(ϵ)>0,ϵ\rho(\epsilon)=\rho_{f}(\epsilon)\propto\sqrt{\epsilon}\Rightarrow\rho_{f}^{\prime}(\epsilon)>0,\forall\epsilon, with ρf(ϵ)\rho_{f}(\epsilon) being the density of states of three-dimensional free electrons, leads to

1EF(T>Tc)EF(0,c0)=π212(TTF)2.1-\frac{E_{F}(T>T_{c})}{E_{F}(0,c_{0})}=\frac{\pi^{2}}{12}\left(\frac{T}{T_{F}}\right)^{2}\quad.

A numerical application with a typical value TF=3×104KT_{F}=3\times 10^{4}K yields TF(300K)TF(0)3K<<TF|dTFdT(T>Tc)|<<1T_{F}(300K)-T_{F}(0)\approx 3K<<T_{F}\Rightarrow\left|\frac{dT_{F}}{dT}\left(T>T_{c}\right)\right|<<1.

Taking advantage of Eq.1, the expression of dEFdT(TTc)\frac{dE_{F}}{dT}(T\leq T_{c}) is obtained to read

dcn=cnEFdEF+cnTdTdcs=csμdμ=dcn}dEFdT=cnTβ(T),\left.\begin{array}[]{l}dc_{n}=\frac{\partial c_{n}}{\partial E_{F}}dE_{F}+\frac{\partial c_{n}}{\partial T}dT\\ dc_{s}=\frac{\partial c_{s}}{\partial\mu}d\mu=-dc_{n}\end{array}\right\}\Rightarrow\frac{dE_{F}}{dT}=-\frac{\frac{\partial c_{n}}{\partial T}}{\beta(T)}\quad, (10)

with β(T)=cnEF+csμ\beta(T)=\frac{\partial c_{n}}{\partial E_{F}}+\frac{\partial c_{s}}{\partial\mu}. The Sommerfeld expansion (see Eq.(5)) leads to

α=dEFdT(Tc)ρρTc=(πkBρTc)23ρβ(Tc).\alpha=\frac{dE_{F}}{dT}(T_{c}^{-})\frac{\rho^{\prime}}{\rho}T_{c}=-\frac{\left(\pi k_{B}\rho^{\prime}T_{c}\right)^{2}}{3\rho\beta(T_{c})}\quad. (11)

Thus, looking back at Eq.7, it is realized that the observedpar ; ash relation Cs(Tc)3Cn(Tc)C_{s}(T_{c})\approx 3C_{n}(T_{c}) requires α>0β(Tc)<0\alpha>0\Rightarrow\beta(T_{c})<0, which had been already identifiedsz4 as a necessary condition for the superconducting state to be at thermal equilibrium. At last, α\alpha reads in case of ρ=ρf\rho=\rho_{f}

α=π212(TTF)2ρEFcn(1+EFcncsμ)1.\alpha=\frac{\pi^{2}}{12}\left(\frac{T}{T_{F}}\right)^{2}\rho\frac{\partial E_{F}}{\partial c_{n}}\left(1+\frac{\partial E_{F}}{\partial c_{n}}\frac{\partial c_{s}}{\partial\mu}\right)^{-1}\quad.

Due to TTF<<1\frac{T}{T_{F}}<<1 and ρEFcn1\rho\frac{\partial E_{F}}{\partial c_{n}}\approx 1, getting α2\alpha\approx 2 requires β(Tc)0EFcn+μcs0\beta(T_{c})\approx 0\Rightarrow\frac{\partial E_{F}}{\partial c_{n}}+\frac{\partial\mu}{\partial c_{s}}\approx 0, so that the stability criterion of the superconducting state reads finally

EFcn(Tc,c0)=μcs(0),ρ(EF(Tc,c0))>0.\frac{\partial E_{F}}{\partial c_{n}}(T_{c},c_{0})=-\frac{\partial\mu}{\partial c_{s}}(0),\quad\rho^{\prime}(E_{F}(T_{c},c_{0}))>0\quad. (12)

Because of EFcn(Tc,c0)1ρ>0\frac{\partial E_{F}}{\partial c_{n}}(T_{c},c_{0})\approx\frac{1}{\rho}>0, Eq.(12) is seen to be consistent with μcs(cs)<0\frac{\partial\mu}{\partial c_{s}}(c_{s})<0, established previously as a prerequisite for persistent currentssz4 and the Josephson effectsz6 . At last, note that there is dTFdT(TTc)>>1\frac{dT_{F}}{dT}(T\leq T_{c})>>1 but inversely 0<dTFdT(T>Tc)<<10<-\frac{dT_{F}}{dT}(T>T_{c})<<1.

In order to grasp the significance of the constraint expressed by Eq.(12), let us elaborate the case for which Eq.(12) is not fulfilled (Cs(T<Tc)<Cn(T)\Rightarrow C_{s}(T<T_{c})<C_{n}(T)). Accordingly the hatched area in Fig.1 is equal to the difference in free energy at T=0T=0 between the superconducting phase and the normal one, and thence also equal to Eb(0)>0E_{b}(0)>0 because the entropy of the normal state vanisheslan at T=0T=0. However applying Eq.(3) with Cs(T<Tc)<Cn(T)C_{s}(T<T_{c})<C_{n}(T) yields Eb(0)<0E_{b}(0)<0, which contradicts the above opposite conclusion Eb(0)>0E_{b}(0)>0, and thereby entails that the MBE state, associated with AA in Fig.1, is not observable at thermal equilibrium in case of unfulfilled Eq.(12), even though it is definitely a MBE eigenstateja1 ; ja2 ; ja3 of the Hubbard Hamiltonian, accounting for the motion of correlated electrons, and its energy is indeed lower than that of the Fermi gas n(T=0,c0)\mathcal{E}_{n}(T=0,c_{0}).

Since energy and free energy are equallan at T=0T=0, Eb(0)E_{b}(0) reads

Eb(0)=0cs(0)(EF(0,c0cs)μ(cs))𝑑cs.E_{b}(0)=\int_{0}^{c_{s}(0)}\left(E_{F}(0,c_{0}-c_{s})-\mu(c_{s})\right)dc_{s}\quad.

In order to work out an upper bound for Eb(0)E_{b}(0), EF(T,c0cs)μ(cs)E_{F}(T,c_{0}-c_{s})-\mu(c_{s}) will be approximated by its Taylor expansion at first order with respect to cscs(T)c_{s}-c_{s}(T), which yields

EF(T,c0cs)μ(cs)me2γ(cs(T)cs),E_{F}(T,c_{0}-c_{s})-\mu(c_{s})\approx\frac{m}{e^{2}}\gamma\left(c_{s}(T)-c_{s}\right)\quad, (13)

with γ=EFcn(cn(T))+μcs(cs(T))\gamma=\frac{\partial E_{F}}{\partial c_{n}}(c_{n}(T))+\frac{\partial\mu}{\partial c_{s}}(c_{s}(T)). Since it has been shownsz5 that EF(T,c0cs)μ(cs)<105eVE_{F}(T,c_{0}-c_{s})-\mu(c_{s})<10^{-5}eV, Eq.(13) turns out to be very accurate. Likewise, due to cs(T)cs0c_{s}(T)\geq c_{s}\geq 0 (see Fig.1) and γ>0\gamma>0 being a necessary conditionsz4 for AA in Fig.1 to correspond to a stable equilibrium, Eq.(13) entails

EF(T,c0cs)μ(cs)EF(T,c0)μ(0).E_{F}(T,c_{0}-c_{s})-\mu(c_{s})\leq E_{F}(T,c_{0})-\mu(0)\quad.

Then by taking advantage of Eqs.(2,9), we get

EF(T,c0)μ(0)EF(0,c0)EF(Tc,c0)=(πkBTc)26ρρ,E_{F}(T,c_{0})-\mu(0)\leq E_{F}(0,c_{0})-E_{F}(T_{c},c_{0})=\frac{\left(\pi k_{B}T_{c}\right)^{2}}{6}\frac{\rho^{\prime}}{\rho},

with ρ>0\rho^{\prime}>0 as required by Eq.(12). At last, assuming ρ(ϵ)=ρf(ϵ)\rho(\epsilon)=\rho_{f}(\epsilon), the searched upper bound per electron is obtained to read

Eb(0)c0EF(Tc,c0)π212(TcTF)2.\frac{E_{b}(0)}{c_{0}E_{F}(T_{c},c_{0})}\leq\frac{\pi^{2}}{12}\left(\frac{T_{c}}{T_{F}}\right)^{2}\quad.

Applying this formula to AlAl (Tc=1.2K,TF3×104KT_{c}=1.2K,T_{F}\approx 3\times 10^{4}K) gives Eb(0)c0EF(Tc,c0)<108\frac{E_{b}(0)}{c_{0}E_{F}(T_{c},c_{0})}<10^{-8}. Moreover, that latter result had enabled us to realizesz2 that the formula Eb(0)=μ0Hc(0)2/2E_{b}(0)=\mu_{0}H_{c}(0)^{2}/2, albeit ubiquitous in textbookspar ; sch ; tin (Hc(TTc),μ0H_{c}(T\leq T_{c}),\mu_{0} refer to the critical magnetic field and the magnetic permeability of vacuum, respectively), underestimates Eb(0)E_{b}(0) by ten orders of magnitude.

Since fulfilling Eq.(12) is tantamount to β(Tc)=0\beta(T_{c})=0, which entails dEFdT(TTc)\frac{dE_{F}}{dT}(T\rightarrow T_{c}^{-})\rightarrow\infty and thence Cs(TTc)C_{s}(T\rightarrow T_{c}^{-})\rightarrow\infty, it must be checked that (T)=0TCs(u)𝑑u\mathcal{E}(T)=\int_{0}^{T}C_{s}(u)du remains still finite for TTcT\rightarrow T_{c}^{-}. To that end, let us work out the Taylor expansion of μ(cs),EF(T,cn)\mu(c_{s}),E_{F}(T,c_{n}) up to the second order around T=0,cs=0T=0,c_{s}=0

μ(cs)=μ(0)+μcs(0)cs+2μcs2(0)cs22EF(T,cn)=EF(Tc,c0)csρρρ3cs22+(πkB)26ρρ(Tc2T2),\begin{array}[]{l}\mu(c_{s})=\mu(0)+\frac{\partial\mu}{\partial c_{s}}(0)c_{s}+\frac{\partial^{2}\mu}{\partial c_{s}^{2}}(0)\frac{c_{s}^{2}}{2}\\ E_{F}(T,c_{n})=E_{F}(T_{c},c_{0})-\frac{c_{s}}{\rho}-\frac{\rho^{\prime}}{\rho^{3}}\frac{c_{s}^{2}}{2}\\ \quad\quad\quad\quad\quad+\frac{\left(\pi k_{B}\right)^{2}}{6}\frac{\rho^{\prime}}{\rho}\left(T_{c}^{2}-T^{2}\right)\end{array}\quad,

for which we have used cn=c0cs,cs=cs(T),EFcn=1ρ2EFcn2=ρρ3c_{n}=c_{0}-c_{s},c_{s}=c_{s}(T),\frac{\partial E_{F}}{\partial c_{n}}=\frac{1}{\rho}\Rightarrow\frac{\partial^{2}E_{F}}{\partial c_{n}^{2}}=-\frac{\rho^{\prime}}{\rho^{3}}. Then taking advantage of Eqs.(1,2) (EF(T,cn)=μ(cs),EF(Tc,c0)=μ(0)\Rightarrow E_{F}(T,c_{n})=\mu(c_{s}),E_{F}(T_{c},c_{0})=\mu(0)) and Eq.(12) (β(Tc)=μcs(0)+1ρ=0\Rightarrow\beta(T_{c})=\frac{\partial\mu}{\partial c_{s}}(0)+\frac{1}{\rho}=0) results into

cs(TTc)=πkBρ(Tc2T2)3(ρ2μcs2(0)+ρρ2)TcT.c_{s}(T\rightarrow T_{c}^{-})=\pi k_{B}\sqrt{\frac{\rho^{\prime}\left(T_{c}^{2}-T^{2}\right)}{3\left(\rho\frac{\partial^{2}\mu}{\partial c_{s}^{2}}(0)+\frac{\rho^{\prime}}{\rho^{2}}\right)}}\propto\sqrt{T_{c}-T}\quad.

It should be noticed that the GL equation predictstin rather cs(TTc)TcTc_{s}(T\rightarrow T_{c}^{-})\propto T_{c}-T.

Likewise, let us calculate similarly the Taylor expansion of β(T)EFcn+μcs\beta(T)\propto\frac{\partial E_{F}}{\partial c_{n}}+\frac{\partial\mu}{\partial c_{s}} up to the first order around T=Tc,cs=0T=T_{c},c_{s}=0

β(TTc)(2μcs2(0)2EFcn2(Tc,c0))csβ(TTc)TcT(TTc)TcT,\begin{array}[]{c}\beta(T\rightarrow T_{c}^{-})\propto\left(\frac{\partial^{2}\mu}{\partial c_{s}^{2}}(0)-\frac{\partial^{2}E_{F}}{\partial c_{n}^{2}}(T_{c},c_{0})\right)c_{s}\Rightarrow\\ \beta(T\rightarrow T_{c}^{-})\propto\sqrt{T_{c}-T}\Rightarrow\mathcal{E}(T\rightarrow T_{c}^{-})\propto\sqrt{T_{c}-T}\end{array}\quad,

whence (TTc)\mathcal{E}(T\rightarrow T_{c}^{-}) is concluded to remain indeed finite.

At last, we shall work out the expression of jM(TTc)j_{M}(T\rightarrow T_{c}^{-}), the maximum current density jsj_{s}, conveyed by bound electrons which was shownsz5 to read

jM=ecm(T)2m(EF(T,c0cm(T))μ(cm(T))),j_{M}=ec_{m}(T)\sqrt{\frac{2}{m}\left(E_{F}\left(T,c_{0}-c_{m}(T)\right)-\mu(c_{m}(T))\right)}\quad,

with e,me,m standing for the charge and effective mass of the electron, while cm(T)=23cs(T)c_{m}(T)=\frac{2}{3}c_{s}(T) designates the corresponding value of csc_{s}, i.e. js(cm)=jMj_{s}(c_{m})=j_{M}. Hence jMj_{M} readssz5 finally

jM(T)=erm(23cs(T))1.5r=EFcn(cn(T))+μcs(cs(T)).\begin{array}[]{c}j_{M}(T)=\frac{er}{\sqrt{m}}\left(\frac{2}{3}c_{s}(T)\right)^{1.5}\\ r=\sqrt{\frac{\partial E_{F}}{\partial c_{n}}(c_{n}(T))+\frac{\partial\mu}{\partial c_{s}}(c_{s}(T))}\end{array}\quad.

It ensues from β(Tc)=0\beta(T_{c})=0 that the leading term of the Taylor expansion of rr around T=Tc,cs=0T=T_{c},c_{s}=0 reads

r(TTc)cs(T)r(TcT)14jM(TTc)TcT,\begin{array}[]{c}r\left(T\rightarrow T_{c}^{-}\right)\propto\sqrt{c_{s}(T)}\Rightarrow r\propto\left(T_{c}-T\right)^{\frac{1}{4}}\Rightarrow\\ j_{M}\left(T\rightarrow T_{c}^{-}\right)\propto T_{c}-T\end{array}\quad,

which is to be compared with the maximum persistent current densitysz5 jc(TT)TTj_{c}\left(T\rightarrow T_{*}^{-}\right)\propto\sqrt{T_{*}-T} with T<TcT_{*}<T_{c}.

III Isotope Effect

Substituting, in a superconducting material, an atomic species of mass MM by an isotope, is well-knownpar ; sch ; tin to alter TcT_{c}. This isotope effect was ascribed to the electron-phonon coupling, on the basis of the observed relation TcM=constantT_{c}\sqrt{M}=constant. The ensueing theoretical treatmentpar ; sch ; tin capitalisedfro on Froehlich’s perturbationlan2 calculation of the self-energy of an independent electron induced by the electron-phonon coupling. However since the BCS picturebcs has subsequently ascertained the paramount role of inter-electron coupling, we shall rather focus hereafter on the effective phonon-mediated interaction between two electrons.

Thus let us consider independent electrons of spin σ=±1/2\sigma=\pm 1/2, moving in a three-dimensional crystal, containing NN sites. The dispersion of the one-electron band reads ϵ(k)\epsilon(k) with ϵ(k),k\epsilon(k),k being the electron, spin-independent (ϵ(k)=ϵ(k)\Rightarrow\epsilon(-k)=\epsilon(k)) energy and a vector of the Brillouin zone, respectively. Their motion is governed, in momentum space, by the Hamiltonian HdH_{d}

Hd=k,σϵ(k)ck,σ+ck,σ,H_{d}=\sum_{k,\sigma}\epsilon(k)c^{+}_{k,\sigma}c_{k,\sigma}\quad,

with the sum over kk to be carried out over the whole Brillouin zone. Then the ck,σ+,ck,σc^{+}_{k,\sigma},c_{k,\sigma}’s are Fermi-like creation and annihilation operatorssch on the Bloch state |k,σ\left|k,\sigma\right\rangle

|k,σ=ck,σ+|0,|0=ck,σ|k,σ,\left|k,\sigma\right\rangle=c^{+}_{k,\sigma}\left|0\right\rangle\quad,\quad\left|0\right\rangle=c_{k,\sigma}\left|k,\sigma\right\rangle\quad,

with |0\left|0\right\rangle being the no electron state. Let us introduce now the electron-phononpar ; sch ; tin ; fro coupling heϕh_{e-\phi}

heϕ=gqNk,k,σck,σ+ck,σ(aq++aq),h_{e-\phi}=\frac{g_{q}}{\sqrt{N}}\sum_{k,k^{\prime},\sigma}c^{+}_{k,\sigma}c_{k^{\prime},\sigma}\left(a^{+}_{q}+a_{-q}\right)\quad,

with q=kkq=k^{\prime}-k and gq(ωqM)1/2g_{q}\propto\left(\omega_{q}M\right)^{-1/2} being the coupling constant characterising the electron-phonon interaction. Likewise, ωq\omega_{q} is the phonon frequency, while the aq+,aqa^{+}_{q},a_{q}’s are Bose-like creation and annihilation operatorssch on the nq𝒩n_{q}\in\mathcal{N} phonon state |nq\left|n_{q}\right\rangle

aq+|nq=nq+1|nq+1,aq|nq=nq|nq1.a^{+}_{q}\left|n_{q}\right\rangle=\sqrt{n_{q}+1}\left|n_{q}+1\right\rangle\quad,\quad a_{q}\left|n_{q}\right\rangle=\sqrt{n_{q}}\left|n_{q}-1\right\rangle.

Because of k|heϕ|k=0,k,k\left\langle k\left|h_{e-\phi}\right|k^{\prime}\right\rangle=0,\forall k,k^{\prime} with |k=ck,++ck,+|0,|k=ck,++ck,+|0\left.|k\right\rangle=c^{+}_{k,+}c^{+}_{-k,-}\left.|0\right\rangle,\left.|k^{\prime}\right\rangle=c^{+}_{k^{\prime},+}c^{+}_{-k^{\prime},-}\left.|0\right\rangle, we shall deal with heϕh_{e-\phi} as a perturbation with respect to HdH_{d}, in order to reckon k|k2\left\langle k\left|k^{\prime}_{2}\right.\right\rangle with |k2\left|k^{\prime}_{2}\right\rangle denoting |k\left|k^{\prime}\right\rangle perturbed at second orderlan2 . Accordingly, we first introduce the unperturbed electron-phonon eigenstates

|k~=|k|nq+|nq2,|k~=|k|nq+|nq2,\left|\widetilde{k}\right\rangle=\left|k\right\rangle\otimes\frac{\left|n_{q}\right\rangle+\left|n_{-q}\right\rangle}{\sqrt{2}},\quad\left|\widetilde{k^{\prime}}\right\rangle=\left|k^{\prime}\right\rangle\otimes\frac{\left|n_{q}\right\rangle+\left|n_{-q}\right\rangle}{\sqrt{2}}\quad,

with nq=nq=nn_{q}=n_{-q}=n. Their respective energies read E(k)=2ϵ(k)+nωqE(k)=2\epsilon(k)+n\hbar\omega_{q}, E(k)=2ϵ(k)+nωqE(k^{\prime})=2\epsilon(k^{\prime})+n\hbar\omega_{q}. Then we reckon |k2~\left|\widetilde{k^{\prime}_{2}}\right\rangle and further project it onto |k~\left|\widetilde{k}\right\rangle, which yields

k~|k2~=gq22N(k~|heϕ|φ+φ+|heϕ|k~+k~|heϕ|φφ|heϕ|k~)φ+=ck,++ck,+|0(n+1D+|nq+1+nD|nq1)φ=ck,++ck,+|0(n+1D+|nq+1+nD|nq1),\begin{array}[]{l}\left\langle\widetilde{k}\left|\widetilde{k^{\prime}_{2}}\right.\right\rangle=\frac{g^{2}_{q}}{2N}\left(\left\langle\widetilde{k}\left|h_{e-\phi}\right|\varphi_{+}\right\rangle\left\langle\varphi_{+}\left|h_{e-\phi}\right|\widetilde{k^{\prime}}\right\rangle\right.\\ \quad\quad\quad\quad\quad\quad\left.+\left\langle\widetilde{k}\left|h_{e-\phi}\right|\varphi_{-}\right\rangle\left\langle\varphi_{-}\left|h_{e-\phi}\right|\widetilde{k^{\prime}}\right\rangle\right)\\ \varphi_{+}=c^{+}_{k,+}c^{+}_{-k^{\prime},-}\left|0\right\rangle\otimes\left(\frac{\sqrt{n+1}}{D_{+}}\left|n_{q}+1\right\rangle+\frac{\sqrt{n}}{D_{-}}\left|n_{-q}-1\right\rangle\right)\\ \varphi_{-}=c^{+}_{k^{\prime},+}c^{+}_{-k,-}\left|0\right\rangle\otimes\left(\frac{\sqrt{n+1}}{D_{+}}\left|n_{-q}+1\right\rangle+\frac{\sqrt{n}}{D_{-}}\left|n_{q}-1\right\rangle\right)\end{array},

with D±=ϵkϵk±ωqD_{\pm}=\epsilon_{k}-\epsilon_{k^{\prime}}\pm\hbar\omega_{q}. The searched xk,k=Nk|k2x_{k,k^{\prime}}=N\left\langle k\left|k^{\prime}_{2}\right.\right\rangle is then inferred to read

xk,k=(2n(T)+1)gq2((ϵkϵk)2(ωq)2),x_{k,k^{\prime}}=\frac{\left(2n(T)+1\right)g^{2}_{q}}{\left(\left(\epsilon_{k}-\epsilon_{k^{\prime}}\right)^{2}-\left(\hbar\omega_{q}\right)^{2}\right)}\quad,

with n(T)=(eωqkBT1)1n(T)=\left(e^{\frac{\hbar\omega_{q}}{k_{B}T}}-1\right)^{-1} being the thermal average of n±qn_{\pm q}. Moreover it can be checked that xk,k=xk,kx_{k,k^{\prime}}=x_{k^{\prime},k}. Thus, for qq not close to the Brillouin zone center (the most likely occurence), there is xk,k>0x_{k,k^{\prime}}>0, whereas xk,k<0x_{k,k^{\prime}}<0 can be found only for q0q\approx 0. Likewise, though the hereabove expression is redolent of one derived by Froehlichfro , their respective significances are unrelated, since Froehlich interpreted the self-energy of one electron and one phonon bound together in terms of virtual transitions between various electron-phonon states, whereas xk,kx_{k,k^{\prime}} refers to the dot product of two-electron-states.

Projecting the hermitian BCS Hamiltonianbcs ; ja1 ; ja2 ; ja3 HH onto the basis {|k2,|k2}\left\{\left|k_{2}\right\rangle,\left|k^{\prime}_{2}\right\rangle\right\} yields

Hk2,k2=2(ϵk+xk,kUN2+xk,k2N2ϵk)Hk2,k2=UN(1+xk,k2N2)+2xk,kN(ϵk+ϵk)Hk2,k2=2(ϵk+xk,kUN2+xk,k2N2ϵk),\begin{array}[]{l}H_{k_{2},k_{2}}=2\left(\epsilon_{k}+\frac{x_{k,k^{\prime}}U}{N^{2}}+\frac{x^{2}_{k,k^{\prime}}}{N^{2}}\epsilon_{k^{\prime}}\right)\\ H_{k_{2},k^{\prime}_{2}}=\frac{U}{N}\left(1+\frac{x^{2}_{k,k^{\prime}}}{N^{2}}\right)+2\frac{x_{k,k^{\prime}}}{N}\left(\epsilon_{k}+\epsilon_{k^{\prime}}\right)\\ H_{k^{\prime}_{2},k^{\prime}_{2}}=2\left(\epsilon_{k^{\prime}}+\frac{x_{k,k^{\prime}}U}{N^{2}}+\frac{x^{2}_{k,k^{\prime}}}{N^{2}}\epsilon_{k}\right)\end{array}\quad,

whence it can be concluded within the thermodynamic limit (NN\rightarrow\infty) that the diagonal matrix elements Hk,kH_{k,k} remains unaltered by the electron-phonon coupling, whereas UU is slightly renormalised to U+2xk,k(ϵk+ϵk)U+2x_{k,k^{\prime}}\left(\epsilon_{k}+\epsilon_{k^{\prime}}\right). Anyhow, since, as noted above, xk,k>0x_{k,k^{\prime}}>0 is the most likely case, it is hard to figure out how the phonon-mediated isotope effect could lessen UU, as concluded by Froehlichfro .

Because, in some materials, the observed isotope effect does not comply with TcM=constantT_{c}\sqrt{M}=constant, it has been ascribed tentativelysab to the hyperfineabr interaction, coupling the nuclear and electron spin, provided the electron wave-function includes some ss-like character. We shall derive the corresponding xk,kx_{k,k^{\prime}}, by proceeding similarly as above for the electron-phonon one and keeping the same notations.

The Hamiltonian reads for nuclear spins =1/2=1/2 in momentum space

Hh=ANk,kck,++ck,Iq+ck,+ck,+Iq+,H_{h}=\frac{A}{\sqrt{N}}\sum_{k,k^{\prime}}c^{+}_{k,+}c_{-k^{\prime},-}I_{q}^{-}+c^{+}_{-k,-}c_{k^{\prime},+}I_{q}^{+}\quad,

with AA being the hyperfine constant, ±\pm referring to the two spin directions and q=k+kq=k+k^{\prime}. Likewise, the I±=σx±iσy2I^{\pm}=\frac{\sigma_{x}\pm i\sigma_{y}}{2}’s, with σx,σy\sigma_{x},\sigma_{y} being Pauli’s matricesabr characterising the nuclear spin, operate on nuclear spin states |±\left|\pm\right\rangle. Note that the term σz\propto\sigma_{z} has been dropped because it turned out to contribute nothing to xk,kx_{k,k^{\prime}}. The unperturbed eigenstates read

|k~=|k|+q+|q2,|k~=|k|+q+|q2.\left|\widetilde{k}\right\rangle=\left|k\right\rangle\otimes\frac{\left|+\right\rangle_{q}+\left|-\right\rangle_{q}}{\sqrt{2}},\quad\left|\widetilde{k^{\prime}}\right\rangle=\left|k^{\prime}\right\rangle\otimes\frac{\left|+\right\rangle_{q}+\left|-\right\rangle_{q}}{\sqrt{2}}\quad.

Their respective energies are E(k)=2ϵ(k)E(k)=2\epsilon(k), E(k)=2ϵ(k)E(k^{\prime})=2\epsilon(k^{\prime}). Then xk,k,k|k2x_{k,k^{\prime}},\left\langle k\left|k^{\prime}_{2}\right.\right\rangle read in this case

xk,k=A24(ϵkϵk)2k|k2=xk,kNk~|hh|φφ|hh|k~φ=ck,++ck,++|0|q+ck,+ck,+|0|+q.\begin{array}[]{l}x_{k,k^{\prime}}=-\frac{A^{2}}{4\left(\epsilon_{k^{\prime}}-\epsilon_{k}\right)^{2}}\\ \left\langle k\left|k^{\prime}_{2}\right.\right\rangle=\frac{x_{k,k^{\prime}}}{N}\left\langle\widetilde{k}\left|h_{h}\right|\varphi\right\rangle\left\langle\varphi\left|h_{h}\right|\widetilde{k^{\prime}}\right\rangle\\ \varphi=c^{+}_{k,+}c^{+}_{k^{\prime},+}\left|0\right\rangle\otimes\left|-\right\rangle_{q}+c^{+}_{-k,-}c^{+}_{-k^{\prime},-}\left|0\right\rangle\otimes\left|+\right\rangle_{q}\end{array}.

Except for having the opposite sign, xk,kx_{k,k^{\prime}} has the same properties as in case of the electron-phonon coupling, which causes UU to be renormalised to a slightly lesser value.

IV Experimental Outlook

Three experiments, enabling one to assess the validity of this analysis, will be discussed below. The first one addresses the determination of μcs\frac{\partial\mu}{\partial c_{s}}, which plays a key role for the existence of persistent currentssz4 and the stability of the superconducting phase (see Eq.(12)). As shown elsewheresz5 , the partial pressure p(TTc)p(T\leq T_{c}), exerted by the conduction electrons, and their associated compressibility coefficientlan3 χ(T)\chi(T) read

p=cnEF(cn)Fn(cn)+csμ(cs)s(cs)χ=dVVdp=(cn2EFcn+cs2μcs)1,\begin{array}[]{l}p=c_{n}E_{F}(c_{n})-F_{n}(c_{n})+c_{s}\mu(c_{s})-\mathcal{E}_{s}(c_{s})\Rightarrow\\ \chi=-\frac{dV}{Vdp}=\left(c_{n}^{2}\frac{\partial E_{F}}{\partial c_{n}}+c_{s}^{2}\frac{\partial\mu}{\partial c_{s}}\right)^{-1}\end{array}\quad, (14)

with cn=cn(T),cs=cs(T)c_{n}=c_{n}(T),c_{s}=c_{s}(T) and VV being the sample volume. For TTcT\rightarrow T_{c}, there is cs0c_{s}\rightarrow 0, so that it might be impossible to measure the contribution of bound electrons cs2μcs(0)\propto c_{s}^{2}\frac{\partial\mu}{\partial c_{s}}(0) to χ\chi in Eq.(14). Such a hurdle might be dodged by making the kind of differential measurement to be described now. A square-wave current I(t+tp)=I(t),tI(t+t_{p})=I(t),\forall t, such that I(t[tp2,0])=0,I(t[0,tp2])=IcI\left(t\in\left[-\frac{t_{p}}{2},0\right]\right)=0,I\left(t\in\left[0,\frac{t_{p}}{2}\right]\right)=I_{c} (IcI_{c} stands for the critical current), is flown through the sample, so that the sample switches periodically from superconducting to normal. Then using a lock-in detection procedure for the χ\chi measurement might enable one to discriminate cs2μcsc_{s}^{2}\frac{\partial\mu}{\partial c_{s}} against cn2EFcnc_{n}^{2}\frac{\partial E_{F}}{\partial c_{n}}, despite cs(TTc)<<cnc0c_{s}\left(T\rightarrow T_{c}\right)<<c_{n}\approx c_{0} and thence to check the validity of Eq.(12).

The validity of Eq.(1) can be assessed by shining UVUV light of variable frequency ω\omega onto the sample and measuring the electron work functionash w(TTc)w(T\leq T_{c}) by observing two distinct photoemission thresholds w1=ω1=EF(T),w2=ω2=2μ(T)w_{1}=\hbar\omega_{1}=E_{F}(T),w_{2}=\hbar\omega_{2}=2\mu(T), associated respectively with single electron and electron pair excitation. Observing ω2=2ω1\omega_{2}=2\omega_{1} would validate Eq.(1). Besides, if cs(T)c_{s}(T) is known from skin-depth measurementssz1 , μ(cs)\mu(c_{s}) could be charted. Note also that, if such an experiment were to be carried out in a material, exhibiting a superconducting gap EgE_{g}, a large decrease of EFE_{F} from EF(Tc)E_{F}(T_{c}) down to EF(0)=μ(c0)=ϵbEgE_{F}(0)=\mu(c_{0})=\epsilon_{b}-E_{g} should be expected (ϵb\epsilon_{b} designates the bottom of the conduction band).

For T>10KT>10K, the electron specific heat is overwhelmedash by the lattice contribution CϕC_{\phi}, so that there are no accurate experimental datalor for Cs(T)C_{s}(T). Such a difficulty might be overcome by using again the differential technique, described above. A constant heat power WW is fed into a thermally insulated sample, while its time-dependent temperature T(t)T(t) is monitored. Thus T(t)T(t) can be obtained owing to

W=(Cϕ(T)+Cs(T))T˙(t[tp2,0])W=(Cϕ(T)+Cn(T))T˙(t[0,tp2]),\begin{array}[]{l}W=\left(C_{\phi}(T)+C_{s}(T)\right)\dot{T}\left(t\in\left[-\frac{t_{p}}{2},0\right]\right)\\ W=\left(C_{\phi}(T)+C_{n}(T)\right)\dot{T}\left(t\in\left[0,\frac{t_{p}}{2}\right]\right)\end{array}\quad,

with T˙=dTdt\dot{T}=\frac{dT}{dt}. Feeding again the square-wave current, mentioned above, into the sample, while using the same lock-in detection technique, could enable one to extract Cs(T)Cn(T)C_{s}(T)-C_{n}(T) from the measured signal T˙(t)\dot{T}(t), despite Cϕ>>Cs,CnC_{\phi}>>C_{s},C_{n}. Notesz5 that Cϕ,CnC_{\phi},C_{n}, unlike CsC_{s}, do not depend on the current II and CnC_{n} can always be measured at low TT and then extrapolatedash up to TcT_{c} thanks to Cn(T)=(πkB)23ρ(EF)TC_{n}(T)=\frac{\left(\pi k_{B}\right)^{2}}{3}\rho(E_{F})T.

V Conclusion

A criterion, warranting the stability of the superconducting phase, has been worked out and found to agree with a prerequisite μcs<0\frac{\partial\mu}{\partial c_{s}}<0, established previously for persistent currentssz4 , thermal equilibriumsz5 and the Josephson effectsz6 . The temperature dependence at TTcT\rightarrow T_{c} has been given for the specific heat, concentration and maximum current density, conveyed by superconducting electrons. At last, an original derivation of the isotope effect has been given.

Due to the inequality Uμcs<0U\frac{\partial\mu}{\partial c_{s}}<0, shown elsewheresz5 , the necessary condition μcs<0\frac{\partial\mu}{\partial c_{s}}<0 entails U>0U>0, i.e. a repulsive inter-electron force, such as the Coulomb one, is needed for superconductivity to occur, if the Hubbard model is taken to describe the correlated electron motion. Note that, in the mainstream interpretationdag ; arm of the properties of high-TcT_{c} materials, such a repulsive force is also believed to be instrumental above TcT_{c} but not below TcT_{c} due to the BCS assumption U<0U<0, although the nature of the inter-electron coupling remains unaltered at TcT_{c}. Thence, the BCS model is found not to be consistent with persistent currentssz4 , thermal equilibriumsz5 and a stable superconducting phase, as shown hereabove, due to U<0μcs>0U<0\Rightarrow\frac{\partial\mu}{\partial c_{s}}>0.

References

  • (1) R.D. Parks, Superconductivity, ed. CRC Press (1969)
  • (2) J.R. Schrieffer, Theory of Superconductivity, ed. Addison-Wesley (1993)
  • (3) M. Tinkham, Introduction to Superconductivity, ed. Dover Books (2004)
  • (4) V.L. Ginzburg and L.D. Landau, Zh.Eksperim.i.Teor.Fiz., 20 (1950) 1064
  • (5) J. Bardeen, L.N. Cooper and J.R. Schrieffer, Phys.Rev., 108 (1957) 1175
  • (6) L.D. Landau and E.M. Lifshitz, Statistical Physics, ed. Pergamon Press (1959)
  • (7) N.W. Ashcroft and N. D. Mermin, Solid State Physics, ed. Saunders College (1976)
  • (8) J. Szeftel, N. Sandeau and M. Abou Ghantous, Eur.Phys.J.B, 92 (2019) 67
  • (9) J. Szeftel, N. Sandeau and M. Abou Ghantous, J.Supercond.Nov.Mag., 33 (2020) 1307
  • (10) J. Szeftel, N. Sandeau and A. Khater, Phys.Lett.A, 381 (2017) 1525
  • (11) J. Szeftel, N. Sandeau and A. Khater, Prog.In.Electro.Res.M, 69 (2018) 69
  • (12) J. Szeftel, M. Abou Ghantous and N. Sandeau, Prog.In.Electro.Res.L, 81 (2019) 1
  • (13) J. Szeftel, N. Sandeau, M. Abou Ghantous and A. Khater, EPL, 131 (2020) 17003
  • (14) C.J. Gorter and H. Casimir, Physica, 1 (1934) 306
  • (15) J. Szeftel, N. Sandeau and M. Abou Ghantous, arXiv : 2003.07196
  • (16) J. Szeftel and A. Khater, Phys.Rev.B, 54 (1996) 13581
  • (17) J. Szeftel, Electron Correlations and Material Properties 2, eds. A. Gonis, N. Kioussis, M. Ciftan, (Kluwer Academic, New York, 2003)
  • (18) J. Szeftel and M. Caffarel, J.Phys. A, 37 (2004) 623
  • (19) H. Froehlich, Phys.Rev., 79 (1950) 845
  • (20) L.D. Landau and E.M. Lifshitz, Quantum Mechanics, ed. Pergamon Press (1965) 845
  • (21) M.H. El Saba, private communication
  • (22) A. Abragam, Nuclear Magnetism, ed. Oxford Press (1961)
  • (23) L.D. Landau and E.M. Lifshitz, Theory of Elasticity, ed. Pergamon Press (1970)
  • (24) J.W. Loram, K.A. Mirza and P.F. Freeman, Physica C, 171 (1990) 243
  • (25) E. Dagotto, Rev.Mod.Phys., 66 (1994) 763
  • (26) N.P. Armitage, P. Fournier and R.L. Greene, Rev.Mod.Phys., 82 (2010) 2421