1: \documentclass[10pt,aps,prb,showpacs,superscriptaddress,twocolumn,floatfix]{revtex4}
2: \usepackage{amsmath}
3: \usepackage{citesort}
4: %\usepackage{fullpage}
5: \usepackage{graphicx}
6: \usepackage{dcolumn}
7: %\renewcommand\baselinestretch{1.6}
8: \begin{document}
9: \title{Optical Sum Rule in Finite Bands}
10: \author{J.P. Carbotte}
11: \affiliation{Department of Physics and Astronomy, McMaster University,\\
12: Hamilton, Ontario, L8S 4M1 Canada}
13: \author{E. Schachinger}
14: \email{schachinger@itp.tu-graz.ac.at}
15: \homepage{www.itp.tu-graz.ac.at/~ewald}
16: \affiliation{Institute of Theoretical and Computational Physics\\
17: Graz University of Technology, A-8010 Graz, Austria}
18: \date{\today}
19: \begin{abstract}
20: In a single finite electronic band the total optical spectral weight
21: or optical sum carries information on the interactions involved
22: between the charge carriers as well as on their band structure.
23: It varies with temperature as well as with impurity scattering.
24: The single band optical sum also bears some relationship to the charge
25: carrier kinetic energy and, thus, can potentially provide
26: useful information, particularly on its change as the charge
27: carriers go from normal to superconducting state. Here we
28: review the considerable advances that have recently been made in
29: the context of high $T_c$ oxides, both theoretical and experimental.
30: \end{abstract}
31: \pacs{74.20.Mn, 74.25.Gz, 74.72.-h.}
32:
33: \maketitle
34: %\newpage
35:
36: \section{Introduction}
37:
38: Much is known about the properties of the superconducting state
39: in the cuprates, yet after 20 years of intense research there is,
40: as yet, no consensus as to the driving interactions responsible
41: for the pairing (mechanism). It is well established that the
42: superconducting order parameter has $d$-wave rather than
43: conventional $s$-wave symmetry. Angular resolved photoemission
44: spectroscopy\cite{ref1,ref2} (ARPES) finds that the leading
45: edge of the electron spectral density at the Fermi energy shifts
46: below the chemical potential as a result of the onset of superconductivity.
47: The observed shift is zero in the nodal direction $(\pi,\pi)$
48: and maximum in the antinodal direction $(0,\pi)$.
49: More recently a non monotonic increase of
50: the gap amplitude as a function of direction has been observed
51: in an electron doped cuprate but the symmetry is again $d$-wave%
52: .\cite{ref3}
53:
54: Many other independent experimental techniques have provided
55: additional compelling evidence for $d$-wave symmetry. One such
56: technique is microwave spectroscopy which gives a linear in
57: temperature $(T)$ dependence of the superfluid density at low
58: $T$ [\onlinecite{ref4}]. Complimentary to the above techniques are
59: phase sensitive experiments\cite{ref5} based on Josephson
60: tunneling and flux quantization which provide direct evidence for a
61: change in sign of the order parameter not just that it has a zero.
62:
63: Another important observation which points directly to a non
64: conventional mechanism is the so called collapse of the
65: inelastic scattering rate\cite{ref6,ref7,ref8,ref9} in the
66: superconducting state which results in a large peak at intermediate
67: $T$ below the critical temperature $(T_c)$ in the real part of the
68: microwave conductivity.\cite{ref6,ref7,ref10,ref11} The origin of this
69: peak is understood in terms of a readjustment in the excitation spectrum
70: involved in the interaction as superconductivity develops. This
71: is expected in electronic mechanisms\cite{ref6,ref9} involving
72: the spin or charge susceptibility which looses weight at small
73: energies corresponding to a hardening of the spectrum and little
74: inelastic scattering remains at low $T$.
75:
76: All the evidence reviewed above points to deviations from a simple
77: BCS $s$-wave phonon driven superconductivity, but, nevertheless,
78: the search goes on for other essential differences which could
79: help identify the underlying mechanism. An avenue to explore is
80: the idea of kinetic as opposed to potential energy driven
81: superconductivity. The idea is that it is a reduction in kinetic
82: energy (KE) that drives the condensation into the superconducting state
83: in contrast to BCS theory where the potential is decreased
84: sufficiently to overcome a KE increase and provides as well a
85: reduction in total energy. The possibility of KE driven
86: superconductivity was considered by Hirsch\cite{ref12} and
87: explicitely demonstrated for the hole mechanism of
88: superconductivity\cite{ref13,ref14,ref15} for which the effective
89: mass of the holes decreases in the superconducting state by
90: pairing. It is also expected in other theoretical frameworks
91: for highly correlated systems.\cite{ref16,ref17}
92:
93: In a single finite tight binding band with nearest neighbors
94: hopping only, the KE
95: is related to the optical sum defined as the total optical spectral
96: weight under the real part of the optical conductivity
97: $\sigma_1(T,\omega)$ integrated over energy $\omega$.\cite{ref19} This fact
98: applies whatever the interactions involved, the temperature,
99: and also when impurities are present. Of course, for this to hold
100: the conductivity needs to be computed exactly including vertex
101: corrections. At first sight the
102: restriction to a tight binding band with nearest neighbors only
103: appears restrictive but in reality it has been found that KE
104: and optical sum (OS) track each other reasonably well
105: even when higher neighbors are considered (second, third, etc.).
106: In view of this fact several recent experimental and theoretical
107: papers have appeared concerning the temperature dependence of the
108: OS in the normal state and its change in the superconducting state.
109: Here, we review this body of work with the aim of understanding
110: what it tells us about the underlying interactions involved.
111: We will restrict our discussion to the cuprates and to their
112: in-plane response. For the $c$-axis motion a sum rule
113: violation was noticed by Basov {\it et al.}\cite{ref18}
114: but the KE involved in the $c$-axis motion is small.\cite{ref20,ref21}
115: To investigate its relation to the total condensation energy
116: of the transition to the superconducting state, it is necessary to
117: consider the $ab$-plane of the cuprates. Exactly how to treat the
118: $c$-axis transport also complicates the interpretation. Some
119: possible mechanisms include strong intra layer scattering,\cite{ref22}
120: non Fermi liquid ground states,\cite{ref23} confinement,\cite{ref16}
121: inter-plane and in-plane charge fluctuations,\cite{ref24} indirect
122: $c$-axis coupling through the particle-particle channel,\cite{ref25}
123: resonant tunneling on localized states in the blocking layer,\cite{ref26}
124: two band models,\cite{ref27,ref28} and coherent and incoherent
125: tunneling.\cite{ref29,ref30,ref31,ref32,ref33} Clearly this problem
126: is worth further study but is not part of this review which will deal
127: only with the cuprates and their $ab$-plane response.
128:
129: The in-plane optical sum (OS) in the cuprates is observed to increase
130: as the temperature is decreased below room temperature.
131: In the normal phase it is often, but not always, found to
132: follow a $T^2$ law. The changes are
133: small but larger than is expected on a rigid band model neglecting
134: interactions. % and often, but not always, found to follow a $T^2$ law.
135: In the superconducting phase, on the other hand,
136: %Below the critical temperatureRe: Schachinger - Weber
137: a change in slope of this $T^2$
138: behavior is seen. While there remain some differences in details
139: between the various experimental groups, for underdoped samples it is
140: agreed that the OS falls above the extrapolated normal state value.
141: This is clearly seen in Fig.~\ref{fig:14} which was reproduced from
142: van der Marel {\it et al.}\cite{ref34} and deals with
143: Bi$_2$Sr$_2$CaCu$_2$O$_{8+x}$
144: (BSCCO, Bi2212). This behavior is opposite to that expected on the
145: basis of BCS theory for which the OS is predicted to decrease. This
146: fact can be traced to an increase in KE due to the
147: formation of Cooper pairs. On the other hand, such a conventional
148: behavior of the OS or change in KE on entering the superconducting
149: state was observed in overdoped samples of BSCCO by Deutscher
150: {\it et al.}\cite{ref87a} Their data is reproduced in our
151: Fig.~\ref{fig:19}. These authors also find that the crossover
152: from negative to positive KE change occurs around, but slightly
153: above optimum doping (see Fig.~\ref{fig:20}). These findings
154: have been largely confirmed by Carbone {\it et al.}\cite{ref33a} It is
155: these facts that this review is directly concerned with and seeks
156: to understand. Theoretically we will find that the temperature
157: dependence of the OS can in some cases be dominated by a term which is
158: proportional to a specific average over energy of the quasiparticle
159: inelastic scattering rates and this need not give a $T^2$ law. This
160: term is missing in all theories that do not explicitly treat damping
161: effects. Further, this non $T^2$ dependence has implications for the
162: accuracy of experimental results on the OS difference between
163: superconducting and normal state which require an extrapolation to
164: low temperatures of normal state data known only above $T_c$.
165:
166: The paper is structured as follows. We begin with theoretical
167: considerations and then review experimental information as it relates
168: to the calculations.
169: Section \ref{ssec:2a} introduces the optical sum and its relation to the
170: kinetic energy. In
171: Sec.~\ref{ssec:2d} we provide simple analytic formulas for the KE and
172: OS in a free electron model which allows one to understand some
173: qualitative aspects of this relationship. It also contains a general
174: formulation of the OS for a
175: simplified tight binding model which averages over anisotropies and
176: greatly simplifies the mathematics. We argue that the temperature
177: dependence of the OS is governed mainly by the inelastic scattering.
178: Both, real and imaginary part of the charge carrier self energy are
179: important. In Sec.~\ref{ssec:2b} we describe the Nearly
180: Antiferromagnetic Fermi Liquid model (NAFFL), give the set of generalized
181: Eliashberg equations needed for numerical work, and also specify the
182: electron-spin fluctuations interaction (MMP model). We give results for the
183: temperature dependence of the OS and of the KE in the normal and
184: superconducting state. In Sec.~\ref{ssec:2c} we introduce the Hubbard
185: model and Dynamical Mean Field Theory (DMFT) and give results.
186:
187: In Sec.~\ref{ssec:3b} we summarize some of the important
188: effects a finite band cutoff has on the self energy. Section~\ref{ssec:3c}
189: is devoted to results for the OS when a delta function, and
190: Sec.~\ref{ssec:3d} when an MMP form
191: (spin fluctuations) is used for the electron-boson interaction.
192: Section~\ref{ssec:3e} provides an analysis of the temperature
193: dependence of the OS for coupling to a low energy boson which
194: results in a linear
195: rather than quadratic in $T$ law. It is argued that other temperature
196: dependences could arise when different model interactions are used.
197:
198: In Sec.~\ref{ssec:4a} we investigate within the NAFFL model the effect on
199: the OS of the collapse of the inelastic scattering on entering the
200: superconducting state which can lead to an increase in the OS rather than
201: the usual BCS behavior. We also present experimental results in the
202: BSCCO cuprates. Section~\ref{ssec:4c} deals
203: with a simplified qualitative model based on a temperature dependent
204: scattering rate
205: which decreases with $T$ as $T^4$. While this model is not
206: accurate, it shows clearly how the KE in the superconducting state can
207: decrease below its normal state value due to the collapse of the
208: inelastic scattering.
209: In Sec.~\ref{ssec:4b} we describe a related,
210: more phenomenological model due to Norman and P\'epin which has several
211: common elements with the model discussed in
212: Sec.~\ref{ssec:4a}. We also provide comparison with
213: experiment and additional theoretical results for the KE change on
214: condensation into the superconducting state in Sec.~\ref{ssec:4e}.
215: Section~\ref{ssec:4d}
216: gives DMFT results for the normal as well as superconducting state
217: in the $t$-$J$ model. KE and potential energy are discussed and
218: compared with experiment. Further results based on the Hubbard
219: model are commented on as are those based on the negative $U$
220: Hubbard model
221: used to describe the BCS - BE (Bose - Einstein) crossover.
222:
223: Section~\ref{sec:5} deals with models of the pseudogap state above
224: $T_c$ (the superconducting critical temperature) that exists up
225: to a temperature $T^\ast>T_c$, with $T^\ast$ the pseudogap temperature.
226: In Sec.~\ref{ssec:5a} we discuss KE changes in the preformed pair model
227: in which pairs form at the pseudogap temperature $T^\ast$ and
228: superconductivity sets in only at a lower temperature when phase
229: coherence is established. In Sec.~\ref{ssec:5b} we consider an alternative
230: model to phase fluctuations of the pseudogap state, namely the
231: $D$-density wave model, a competing ordered state with bond currents
232: and associated magnetic moments. In Sec.~\ref{sec:6} we deal briefly with
233: the problem of spectral weight distribution as a function of energy.
234: A short summary is provided in Sec.~\ref{sec:7}.
235:
236: \section{THEORY}
237: \label{sec:2}
238: \subsection{General considerations}
239: \label{ssec:2a}
240:
241: The single band OS is defined as\cite{ref34}
242: \begin{equation}
243: \label{eq:1}
244: \int\limits_{-\Omega}^\Omega\!d\omega\,\Re{\rm e}[
245: \sigma_{1xx}(\omega)] = \frac{\pi e^2}{\hbar^2V}
246: \sum\limits_{{\bf k},\sigma}n_{{\bf k},\sigma}
247: \frac{\partial^2\varepsilon_{\bf k}}{\partial k_x^2} =
248: \frac{\pi e^2}{\hbar^2} W.
249: \end{equation}
250: Here $\sigma_{1xx}(\omega)$ is the real part of the
251: $(x,x)$-component of the optical conductivity,
252: $e$ is the electron charge, $\hbar$ Planck's constant, {\bf k}
253: momentum, $\sigma$ spin, and $V$ the crystal volume. The limits
254: $-\Omega$ to $\Omega$ on the conductivity integral are to be taken
255: to include all the contributions to $\sigma_{1xx}(\omega)$ from the
256: band of interest and from no other.
257: The integration from $-\Omega$ to $\Omega$
258: can be restricted to $(0,\Omega)$ by making use of the fact that
259: the real part of the conductivity is an even function of $\omega$.
260: Often the integration on $\Omega$ is also extended to infinity under
261: the condition that only transitions from the one band of interest
262: are included.
263: The sum over {\bf k} extends over the
264: entire first Brillouin zone, $n_{{\bf k},\sigma}$ is the probability
265: of occupation of a state $\vert{\bf k},\sigma\rangle$, and
266: $\varepsilon_{\bf k}$ is the electron dispersion relation.
267: $W$ is the OS divided by $\pi e^2/\hbar^2$ and will be quoted in
268: meV. In deriving this OS rule the one body part of the Hamiltonian
269: gives the single particle orbitals of energy
270: $\varepsilon_{\bf k}$ and the two body piece
271: provides correlation effects beyond those included in
272: \begin{table}[tp]
273: \caption{\label{tab:1}The two tight binding models used within this
274: paper. Model A corresponds to the tight binding model discussed by
275: van der Marel {\it et al.}\protect{\cite{ref34}} $t$ and $t'$ are
276: given in meV, the critical temperature $T_c$ in K, and the filling
277: $\langle n\rangle$ is defined in Eq.~\protect{\eqref{eq:8}}.
278: }
279: \begin{ruledtabular}
280: \begin{tabular}{ldddd}
281: Model & t & t' & \langle n\rangle & T_c\\
282: \hline
283: A & 148.8 & 40.9 & 0.425 & 90\\
284: B & 100.0 & 16.0 & 0.4 & 100\\
285: \end{tabular}
286: \end{ruledtabular}
287: %\vspace*{10cm}
288: \end{table}
289: $\varepsilon_{\bf k}$ (referred to as kinetic energy even though it
290: can contain effects of interactions).
291: If all bands are included rather than just the one in Eq.~\eqref{eq:1}
292: one would obtain the exact sum rule proportional to $n/m$ with $m$
293: the bare electron mass and $n$ the total electron density. This
294: is a constant independent of temperature. It does not change when
295: the system undergoes a phase transition to the superconducting state
296: by charge conservation. This also applies to the free electron
297: infinite band and is the basis for the Ferrell-Glover-Tinkham
298: sum rule\cite{ref123,ref124} which will be elaborated upon in Sec.~\ref{sec:6}.
299:
300: \begin{figure}[tp]
301: %\centering
302: \vspace*{-4mm}
303: \includegraphics[width=9cm]{srFig1.eps}
304: % \includegraphics[width=13cm]{smFig1.eps}
305: \vspace*{-8mm}
306: \caption{Optical sum $W$ and kinetic energy, $-W_{\rm KE}/2$,
307: as a function of $T^2$. Solid circles and squares are for the non
308: interacting case while solid up-triangles and solid down-triangles
309: include interactions.
310: The left hand frame applies to
311: Model A of Table~\protect{\ref{tab:1}} and an MMP model
312: parameter $\omega_{SF}=82\,$meV [see Eq.~\protect{\eqref{eq:5}}] was
313: used. Here, the interacting and non interacting cases show similar
314: temperature dependencies. The right hand frame is for
315: Model B of Table~\protect{\ref{tab:1}} with an MMP model parameter
316: $\omega_{SF}=10\,$meV. Note the difference in temperature dependence
317: between interacting and non interacting case.
318: }
319: \label{fig:1}
320: \end{figure}
321: The
322: important observation for this review is that for tight binding bands
323: with nearest neighbors hopping, $t$, only, i.e.:
324: \begin{equation}
325: \label{eq:2}
326: \varepsilon_{\bf k} = -2t\left[\cos(ak_x)+\cos(ak_y)\right],
327: \end{equation}
328: where $a$ is the CuO$_2$ plane lattice parameter,
329: \begin{equation}
330: \label{eq:3}
331: W\equiv-\frac{1}{2}W_{KE}\equiv -\frac{1}{2}
332: \frac{a^2}{V}\sum\limits_{{\bf k},\sigma} n_{{\bf k},\sigma}
333: \varepsilon_{\bf k}.
334: \end{equation}
335: $W_{KE}$ is the kinetic energy per copper atom.
336: Thus, for this simple case an experimental measurement of $W$
337: gives a measure of the KE. The band structure in the oxides is not,
338: in general, limited to first neighbor hopping and one may well
339: wonder if the above observation is of any practical use.
340:
341: For second neighbor hopping ($t,\, t'$ model)
342: \begin{eqnarray}
343: \varepsilon_{\bf k} &=& -2\left\{
344: t\left[\cos(ak_x)+\cos(ak_y)\right]\right.\nonumber\\
345: &&\left.- 2t'\cos(ak_x)\cos(ak_y)
346: \right\} - \mu,
347: \label{eq:4}
348: \end{eqnarray}
349: with $\mu$ the chemical potential. In Fig.~\ref{fig:1} we
350: compare results for $W$ (solid circles) with results for
351: $-W_{KE}/2$ (solid squares) for two models, A and B, left and right hand
352: frame, respectively. Parameters of the models are defined in
353: Table~\ref{tab:1} where $\langle n\rangle$ is the filling factor
354: which determines the chemical potential; with $\langle n\rangle = 0.5$,
355: half filling, $\mu = 0$ and $t'=0$ in Eq.~\eqref{eq:4}. No interactions are
356: included beyond those giving rise to the tight binding band structure.
357: It is clear that while $W$ in both cases is a few percent
358: smaller than $-W_{KE}/2$ their dependence on temperature track
359: each other fairly well. Thus, it makes sense to pursue this avenue as
360: a means to get information on the KE and its variation with temperature
361: and also when it undergoes a phase transition to the superconducting
362: state. One should be aware, however, that for some sets of tight
363: binding parameters the difference between OS and KE can be much larger
364: and they may no longer track each other (see the very recent
365: paper by Marsiglio {\it et al.}\cite{ref124a} for more discussion).
366: A final point should be stressed. To compute the OS in Eq.~\eqref{eq:1}
367: it is not necessary, although one can if one wishes, to calculate
368: the conductivity which depends on a two-particle Green's function
369: and is to be calculated with vertex corrections. The right hand side
370: of Eq.~\eqref{eq:1}, instead, depends only on the one particle
371: spectral density $n_{{\bf k},\sigma}$ which determines the probability
372: of occupation of the state $\vert\textbf{k},\sigma\rangle$.
373: On the other hand, if one is interested in the optical spectral
374: weight integrated only to $\omega_M$ with $\omega_M<\Omega$ one
375: can no longer avoid calculating the conductivity. For this reason
376: the problem of optical weight distribution is left to a last
377: section \ref{sec:6}.
378:
379: \subsection{The Continuum Limit and Quadratic Dispersion; Simple Results}
380: \label{ssec:2d}
381:
382: To get some insight into the significance of the OS, we
383: consider next several simplifications. The probability of occupation
384: of the state $\vert{\bf k}\rangle$ (both spins)
385: is related to the charge carrier
386: spectral density $A({\bf k},\omega)$ by
387: \begin{equation}
388: \label{eq:10}
389: n_{\bf k}(T) = \int\limits_{-\infty}^\infty\!d\omega\,f(T,\omega)
390: A({\bf k},\omega),
391: \end{equation}
392: where $f(T,\omega)$ is the Fermi-Dirac distribution function at
393: temperature $T$.
394: For tight binding electrons with no correlation effects the spectral
395: function $A({\bf k},\omega)=\delta(\varepsilon_{\bf k}-\omega)$ and
396: $n_{\bf k}(T) = f(\varepsilon_{\bf k},T)$. Before proceeding to
397: include interactions it is convenient in what follows to rewrite
398: Eq.~\eqref{eq:1} in an equivalent form through integration by parts
399: with the surface integral equal to zero by symmetry in a periodic
400: band:
401: \begin{equation}
402: \label{eq:11}
403: \bar{W} = -\frac{2}{V}
404: \sum\limits_{\bf k}\left(\frac{\partial\varepsilon_{\bf k}}{\partial
405: {\bf k}}\right)^2\frac{dn_{\bf k}}{d\varepsilon_{\bf k}}.
406: \end{equation}
407:
408: Knigavko {\it et al.}\cite{ref49} have found this second form to be
409: very useful when making approximations. In the Kubo formula which
410: defines the conductivity $\sigma_{1xx}(\omega)$ which appears on the
411: left hand side of Eq.~\eqref{eq:1} it is the velocity squared,
412: $(\partial\varepsilon_{\bf k}/\partial k_x)^2$ that enters naturally.
413: In making approximations to the band structure in order to get
414: simplified expressions it is important to have the same quantity
415: $(\partial\varepsilon_{\bf k}/\partial k_x)^2$ on both sides of
416: Eq.~\eqref{eq:1} and, indeed, in Ref.~\onlinecite{ref49} it is verified
417: that the right hand side (RHS)
418: of Eq.~\eqref{eq:1} and $\bar{W}$ of Eq.~\eqref{eq:11} agree
419: to high accuracy for the two simplified band structure models they
420: considered, namely free electron bands cut off to $-D/2$ and $D/2$ at
421: half filling and another model designed to treat the tight binding case
422: which we will define later. In both cases the density of states was
423: approximated by a constant. Such an approximation, however, is not
424: completely compatible with the assumed Fermi velocity model and this
425: explains why we use $\bar{W}$ rather than $W$.
426: In the continuum limit of Eq.~\eqref{eq:4}
427: with $t' = \mu = 0$ as the lattice parameter $a\to 0$ one gets
428: \[
429: \varepsilon_{\bf k} = -4t+\frac{\hbar^2}{2m}\left(k_x^2+k_y^2\right),
430: \]
431: with $\hbar^2/2m = t a^2$ (where this product is assumed to be finite
432: as $a\to 0$) or $a^2 = \frac{\hbar^2}{m}\frac{4}{D}$
433: and $D/2 = 4t$. For these approximations
434: \[
435: W = \bar{W} = \frac{(\hbar\Omega_p)^2}{4\pi e^2},
436: \]
437: the well known result for free electron bands, with $\Omega^2_p =
438: 4\pi ne^2/m$ the plasma frequency squared and $n$ the electron density.
439: We can also show that
440: \[
441: W_{KE} = -W,
442: \]
443: so that the relationship between KE and OS found for
444: tight binding bands with first neighbors is profoundly changed
445: when these are approximated by their continuum
446: limit (free electron case). This is, in a sense, the extreme opposite
447: case to the nearest neighbor only tight binding model.
448: These results were obtained at zero
449: temperature. At finite $T$ things are not quite as simple. $W$
450: and $\bar{W}$ undergo no change with temperature but
451: $W_{KE}$ does and is
452: \[
453: W_{KE} = \frac{N(0)D\hbar^2}{m}\left[\frac{\pi^2}{3}\left(
454: \frac{k_BT}{D/2}\right)^2-1\right],
455: \]
456: which is the well known result that the KE increases as the square of
457: the temperature normalized to the half band width $D/2$. These results
458: show that the correspondence between $W$ and $W_{KE}$ can be
459: rather subtle and it can be lost when approximations to the tight
460: binding band are made. Here $N(0)$ is the charge carrier density
461: of states at the Fermi energy.
462:
463: Knigavko {\it et al.}\cite{ref49} used a
464: somewhat more sophisticated model to approximate a tight binding
465: band. This model was used previously by Marsiglio and Hirsch.\cite{ref50}
466: In this approach the square of the electron velocity is replaced by
467: \[
468: % \label{eq:14}
469: \left(\frac{1}{\hbar}\frac{\partial\varepsilon_{\bf k}}
470: {\partial k_x}\right)^2 =
471: \frac{D}{4m_b}\left[1-\left(\frac{\varepsilon_{\bf k}}{D/2}
472: \right)^2\right]
473: \]
474: and
475: \begin{equation}
476: \label{eq:15}
477: \bar{W}(T) = \frac{\hbar^2}{m_b}\left(-\int\limits_{-\infty}^\infty
478: \frac{d\omega}{D/2}\,f(T,\omega)
479: \int\limits_{-D/2}^{D/2}\frac{d\epsilon}{D/2}\,
480: \epsilon A(\epsilon,\omega)\right).
481: \end{equation}
482: Here, $m_b$ is a band electronic mass. In this model the rule that
483: $\bar{W} = -W_{KE}/2$ still holds.
484: For $A(\epsilon,\omega) = \delta(\epsilon-\omega)$ (no correlations)
485: we recover
486: \begin{equation}
487: \label{eq:16}
488: \bar{W}(T) = \frac{\hbar^2}{m_b}\left[1-\frac{1}{3}\left(
489: \frac{k_BT}{D/2}\right)^2\right],
490: \end{equation}
491: which is a result obtained by van der Marel {\it et al.}\cite{ref34}
492: and often used to interpret experiments. It is important to contrast this
493: result with the free electron case for which
494: \[
495: \left(\frac{1}{\hbar}\frac{\partial\varepsilon_{\bf k}}{\partial
496: k_x}\right)^2 = \frac{D/2+\varepsilon_{\bf k}}{m}
497: \]
498: and
499: \begin{equation}
500: \label{eq:16a}
501: \bar{W}(T) = \frac{\hbar^2 n}{m}\left[1-\frac{2}{n}
502: \int\limits_{-\infty}^\infty\!d\omega\,f(T,\omega)A(D/2,\omega)
503: \right],
504: \end{equation}
505: where $n$ is the charge carrier density.
506:
507: To treat interactions it is convenient to rewrite Eq.~\eqref{eq:15}
508: in the form
509: \begin{equation}
510: \label{eq:17}
511: \bar{W}(T) = \frac{\hbar^2}{2m_b}\left[1-\int\limits_{-\infty}^\infty
512: \!d\omega\,f(T,\omega)h(T,\omega)\right],
513: \end{equation}
514: where
515: \begin{equation}
516: \label{eq:18}
517: h(T,\omega) = \frac{4}{(D/2)^2}\int\limits_0^{D/2}\!d\epsilon\,
518: \epsilon A(\epsilon,\omega).
519: \end{equation}
520: Note also that the probability of occupation of the state $\epsilon$
521: given in Eq.~\eqref{eq:10}
522: %\begin{equation}
523: % \label{eq:18a}
524: % n(\epsilon,T) = \int\limits_{-\infty}^\infty\!d\epsilon\,
525: % f(T,\epsilon)A(\epsilon,\omega)
526: %\end{equation}
527: is closely related to Eqs.~\eqref{eq:17} and \eqref{eq:18}. We can
528: apply the Sommerfeld expansion to Eq.~\eqref{eq:18} to get
529: \begin{eqnarray}
530: \bar{W}(T) &=& \frac{\hbar^2}{2m_b}\left[1-\int\limits_{-\infty}^0\!
531: d\omega\,h(T,\omega)-\frac{\pi^2}{6}\left(k_BT\right)^2\right.
532: \nonumber \\
533: && \times\left.
534: \left.\frac{d\,h(T,\omega)}{d\,\omega}\right\vert_{\omega=0}
535: \right].
536: \label{eq:19}
537: \end{eqnarray}
538: For no interactions $h(T,\omega) = 4\omega/(D/2)^2$ for $\omega\ge 0$
539: and zero for $\omega<0$
540: [note that the derivative in Eq.~\eqref{eq:19} (non interacting case)
541: is to be weighted by $1/2$ at $\omega=0$]
542: so that the second term in Eq.~\eqref{eq:19}
543: vanishes and the third term
544: gives the first correction in \eqref{eq:16}. In terms of the real and
545: imaginary part of the self energy we can work out $h(T,\omega)$ to
546: get
547: \begin{eqnarray}
548: h(T,\omega) &=& \frac{4}{\left(\frac{D}{2}\right)^2}\left\{\frac{\chi}{\pi}
549: \left[\tan^{-1}\left(\frac{\frac{D}{2}-\chi}{y}\right)+
550: \tan^{-1}\left(\frac{\chi}{y}\right)\right]\right.\nonumber\\
551: &&\left.+
552: \frac{y}{2\pi}\ln\left\vert\frac{y^2+\left(\frac{D}{2}-\chi\right)^2}
553: {y^2+\chi^2}\right\vert\right\},
554: \label{eq:20}
555: \end{eqnarray}
556: where $\chi=\omega-\Sigma_1(\omega)$ and $y = -\Sigma_2(\omega)$ with
557: $\Sigma_1(\omega)$ [$\Sigma_2(\omega)$] the real [imaginary] part of
558: the self energy $\Sigma(\omega)$. If the real part of $\Sigma$ is
559: neglected and its imaginary part is assumed to be a constant $\Gamma$
560: as it would be for impurities in an infinite band we can work out the
561: integral in Eq.~\eqref{eq:19} and get:
562: \begin{equation}
563: \label{eq:21}
564: \bar{W}(T) = \frac{\hbar^2}{2m_b}\left[1-\frac{8\Gamma}{D\pi}-
565: \frac{\pi^2}{3}\left(\frac{k_BT}{D/2}\right)^2\right].
566: \end{equation}
567: Note that the second term which deals directly
568: with interactions between
569: the electrons is of order some scattering rate over $D/2$ while the
570: third which has often been emphasized in literature, is of order
571: $[k_BT/(D/2)]^2$ and, therefore, can be expected to be much smaller
572: than the first. In the oxides, as an example, $\Gamma$ is known to be
573: of order $T$.\cite{ref126}
574: Eq.~\eqref{eq:21} was arrived at independently by
575: Benfatto {\it et al.}\cite{ref52} and by Karakozov and Maksimov.\cite{ref53}
576: It is central to any discussion of single band sum rule. While the
577: approach taken in Refs.~\onlinecite{ref52} and \onlinecite{ref53} are
578: somewhat different much of the basic content is similar.
579:
580: Finally, we consider the superconducting state at $T=0$ for which in
581: BCS theory
582: \begin{equation}
583: \label{eq:11a}
584: n_{\bf k} = \frac{1}{2}\left(1-\frac{\varepsilon_{\bf k}}
585: {\sqrt{\varepsilon_{\bf k}^2+\Delta_{\bf k}^2}}\right),
586: \end{equation}
587: where the gap $\Delta_{\bf k}$ can have $d$-wave symmetry of the
588: form $\Delta_{\bf k} = \Delta\cos(2\phi)$. Here $\Delta$ is the
589: gap amplitude and $\phi$ an angle along the cylindrical Fermi
590: surface. To keep things simple, we start with the $s$-wave case
591: and obtain
592: \begin{subequations}
593: \begin{equation}
594: \label{eq:12}
595: W^S_{KE}-W^N_{KE} = \frac{4\hbar^2}{m}\frac{N(0)}{D}
596: \Delta^2\left[\ln\left(\frac{D}{\Delta}
597: \right)-\frac{1}{2}\right].
598: \end{equation}
599: Eq.~\eqref{eq:12}
600: is the difference in KE between superconducting (S) and normal (N)
601: state which has increased as expected since $n_{\bf k}$ given by
602: Eq.~\eqref{eq:11a}
603: populates states above the chemical potential while the step function
604: of the normal state
605: does not. We have also worked out the difference in OS to give
606: \begin{equation}
607: \label{eq:13}
608: \bar{W}^S -\bar{W}^N = -\frac{N(0)\hbar^2}{m}
609: \frac{2\Delta^2}{D},
610: \end{equation}
611: \end{subequations}
612: which has dropped in the superconducting state.
613: Knigavko {\it et al.}\cite{ref49} argued that $\bar{W}$ is the
614: quantity to use when discussing the OS. Here we note that $W$
615: itself shows no change with superconducting transition. Taking ratios
616: with the normal state in Eqs.~\eqref{eq:12} and \eqref{eq:13} we obtain
617: \begin{subequations}
618: \begin{equation}
619: \label{eq:13a}
620: \frac{W^S_{KE}-W^N_{KE}}{\left\vert W^N_{KE}\right\vert} =
621: \left(\frac{2\Delta}{D}\right)^2\left[\ln\left(\frac{D}
622: {\Delta}\right)-\frac{1}{2}\right],
623: \end{equation}
624: and
625: \begin{equation}
626: \label{eq:13b}
627: \frac{\bar{W}^S-\bar{W}^N}{\bar{W}^N} =
628: -\frac{1}{2}\left(\frac{2\Delta}{D}\right)^2.
629: \end{equation}
630: \end{subequations}
631: Note that because $\ln\left(D/\Delta\right)$ is expected to be
632: greater than one, the normalized KE change is greater than the value of the
633: OS drop in this very simplified model. So far we have considered only
634: $s$-wave. The formulas given above hold for $d$-wave with
635: $\Delta\to\Delta\cos(2\phi)$ and an average over angles is required.
636: When this is done
637: \begin{subequations}
638: \begin{equation}
639: \label{eq:13c}
640: \frac{W^S_{KE}-W^N_{KE}}{\left\vert W^N_{KE}\right\vert} =
641: \left(\frac{2\Delta}{D}\right)^2\frac{1}{2}\left[\ln\left(\frac{2D}
642: {\Delta}\right)-\frac{1}{2}\right],
643: \end{equation}
644: and
645: \begin{equation}
646: \label{eq:13d}
647: \frac{\bar{W}^S-\bar{W}^N}{\bar{W}^N} =
648: -\frac{1}{4}\left(\frac{2\Delta}{D}\right)^2.
649: \end{equation}
650: \end{subequations}
651: These are useful expressions to understand qualitatively the physics
652: but are not believed to be accurate. They do represent an extreme
653: case where the OS does not follow the temperature dependence of the
654: KE and they differ in the superconducting case as well.
655:
656: \subsection{The Nearly Antiferromagnetic Fermi Liquid Model}
657: \label{ssec:2b}
658:
659: So far we have not considered interactions yet the oxides are highly
660: correlated systems. A phenomenological approach to correlations
661: is embodied in the Nearly Antiferromagnetic Fermi Liquid model (NAFFL)
662: of Pines and coworkers\cite{ref35,ref36,ref37} and this approach is
663: convenient to obtain some information on the effect of interactions.
664: In this approach the important interactions proceed through the
665: exchange of spin fluctuations and the imaginary part of the spin
666: susceptibility enters a generalized set of Eliashberg equations.
667: A model susceptibility often used is\cite{ref37}
668: \begin{equation}
669: \label{eq:5}
670: \Im{\rm m}\left\{\chi({\bf q},\omega)\right\}
671: = \frac{(\omega/\omega_{SF})\chi_{\bf Q}}
672: {\left[1+\zeta^2({\bf q}-{\bf Q})^2\right]^2+(\omega/\omega_{SF})^2}.
673: \end{equation}
674: The parameters are $\chi_{\bf Q}$ the static susceptibility,
675: {\bf Q} the commensurate antiferromagnetic wave vector $(\pi/a,\pi/a)$
676: in the upper right hand quadrant of the first Brillouin zone and
677: symmetry related points. $\zeta$ is the magnetic coherence length
678: set at $\zeta = 2.5\,a$ in this paper and $\omega_{SF}$ is a
679: characteristic spin fluctuation frequency. Finally, $g$ is the
680: coupling between charge carriers and the spin fluctuations.
681: The Eliashberg equations for renormalized Matsubara frequencies
682: $\tilde{\omega}({\bf k},i\omega_n)$, renormalized energies
683: $\xi({\bf k},i\omega_n)$, and pairing energy (gap) in the
684: \begin{figure}[tp]
685: \vspace*{-5mm}
686: \includegraphics[width=9cm]{srFig2.eps}
687: \caption{The occupation number $n_{\bf k}$ for
688: selected directions in the CuO$_2$ Brillouin zone. Model A of
689: Table~\protect{\ref{tab:1}} was used.
690: Top frame: the non interacting case at $T=0$. It is included for
691: comparison. Center frame: The interacting
692: case at a temperature $T=20\,$K. We show normal state (solid line)
693: and superconducting state (dashed line) results. Bottom frame: The
694: temperature influence on the normal state $n_{\bf k}$ for
695: $T=20\,$K (solid line) and $T=150\,$K (dashed line). }
696: \label{fig:2}
697: \end{figure}
698: superconducting state
699: $\phi({\bf k},i\omega_n)$
700: are\cite{ref35,ref38,ref39,ref40,ref40a}
701: \begin{widetext}
702: \begin{subequations}
703: \label{eq:6}
704: \begin{eqnarray}
705: \tilde{\omega}({\bf k},i\omega_n) &=& \omega_n+
706: T\sum\limits_m\sum\limits_{{\bf k}'}
707: \lambda_{SF}({\bf k}-{\bf k}',i\omega_n-i\omega_m)
708: \frac{\tilde{\omega}({\bf k}',i\omega_m)}{D({\bf k}',i\omega_m)},
709: \label{eq:6a}\\
710: \xi({\bf k},i\omega_n) &=&
711: -T\sum\limits_m\sum\limits_{{\bf k}'}
712: \lambda_{SF}({\bf k}-{\bf k}',i\omega_n-i\omega_m)
713: \frac{\epsilon_{{\bf k}'}+\xi({\bf k}',i\omega_m)}{D({\bf k}',i\omega_m)},
714: \label{eq:6b}\\
715: \phi({\bf k},i\omega_n) &=&
716: -T\sum\limits_m\sum\limits_{{\bf k}'}
717: \lambda_{SF}({\bf k}-{\bf k}',i\omega_n-i\omega_m)
718: \frac{\phi({\bf k}',i\omega_m)}{D({\bf k}',i\omega_m)},
719: \label{eq:6c}
720: \end{eqnarray}
721: with
722: \begin{equation}
723: \label{eq:6d}
724: D({\bf k},i\omega_n) = \tilde{\omega}^2({\bf k},i\omega_n)+
725: \left[\epsilon_{\bf k}+\xi({\bf k},i\omega_n)\right]^2+
726: \phi^2({\bf k},i\omega_n),
727: \end{equation}
728: \end{subequations}
729: and
730: \begin{equation}
731: \label{eq:7}
732: \lambda_{SF}({\bf q},i\nu_n) = \frac{g^2\chi_{\bf Q}}
733: {1+\zeta^2({\bf q}-{\bf Q})^2+(\vert\nu_n\vert/\omega_{SF})}.
734: \end{equation}
735: Here {\bf q} is the momentum transfer and
736: $i\omega_n$, $i\nu_n$ are the electron and boson Matsubara frequencies,
737: respectively. Finally, the filling $\langle n \rangle$ is determined
738: from
739: \begin{equation}
740: \label{eq:8}
741: \langle n\rangle = \frac{1}{2}-\sum\limits_{\bf k}\sum\limits_{n\ge 0}
742: \frac{\epsilon_{\bf k}+\xi({\bf k},i\omega_n)}
743: {\tilde{\omega}^2({\bf k},i\omega_n)+\left[\epsilon_{\bf k}+
744: \xi({\bf k},i\omega_n)\right]^2+\phi^2({\bf k},i\omega_n)}.
745: \end{equation}
746: \end{widetext}
747: For fixed filling $\langle n\rangle$ this last equation determines the
748: chemical potential which changes with temperature for bands which do
749: not have particle-hole-symmetry.
750:
751: We solved the Eliashberg
752: equations\cite{ref40} for the two models of Table~\ref{tab:1} with
753: the single parameter not yet specified, $g$, adjusted to get a
754: critical temperature of $90\,$K and $100\,$K, respectively, for model A
755: and model B. The results for $W$ (solid down-triangles) and
756: $-W_{KE}/2$ (solid up-triangles) are shown in Fig.~\ref{fig:1} in which
757: they are compared with the non interacting results. [A $128\times 128$
758: sampling of the \textbf{k}-space, $ak_x$, $ak_y\in[0,\pi]$ was used
759: which accounts for the wiggles. In one case (not shown in the figure)
760: the \textbf{k}-space sampling was increased to $512\times 512$. This
761: made the curves smoother but there were no other changes.]
762: In both models
763: the interactions have lowered the value of each integral but the
764: temperature dependence of $W$ and $-W_{KE}/2$ still largely
765: track each other. The role played by the interactions can be
766: traced simply in Eqs.~\eqref{eq:1} and \eqref{eq:3} as changing the
767: probability of occupation of the state $\vert{\bf k},\sigma\rangle$
768: factor $n_{{\bf k},\sigma}$.
769: This factor depends on correlations as seen in Fig.~\ref{fig:2}
770: where we show results for model A. The top frame gives $n_{\bf k}$
771: (for both spins)
772: along $\Gamma\to X\to M\to \Gamma$ in the first Brillouin zone for the
773: non interacting case while the bottom frame shows results at $T=20\,$K
774: (solid curve) and $T=150\,$K (dotted curve) in the normal state. We
775: see that the effect of interactions is to make the probability of
776: occupation of the state $\vert{\bf k}\rangle$
777: non zero even in regions where there
778: is no occupation in the non interacting case. Also, there is
779: considerable ``smearing'' of the interacting case distribution which
780: changes with temperature and with the onset of superconductivity (center
781: frame). On comparing bottom and center frame we note an increase
782: in temperature of about $100\,$K in the normal state
783: corresponds roughly to the same amount
784: of extra smearing as is due to the onset of superconductivity. This
785: smearing in the occupation factor means an increase in kinetic
786: energy. This is expected in a BCS mechanism. In the Cooper pair
787: model two electrons are introduced at the Fermi surface of a
788: quiescent Fermi sphere. Because of an effective attractive potential
789: between them they prefer to go into a superposition of plane wave
790: states with $\vert{\bf k}\vert$ equal or slightly greater than
791: $k_F$, the Fermi momentum,
792: and so reduce their potential energy. Although the kinetic
793: energy in the process is increased over the free electron value
794: of $2\varepsilon(k_F)$, the potential is sufficiently reduced to
795: compensate for this increase.
796:
797: Returning to Fig.~\ref{fig:1} several features are to be noted.
798: For the case of no interactions the numerical data for both
799: $W$ and $W_{KE}$ is well fit by a $T^2$ law as is also
800: the interacting case based on Model A. But this is clearly not
801: so for Model B for which the OS and also $-W_{KE}/2$ turns up
802: from a $T^2$ law as $T$ is lowered towards zero with turn up onset
803: occurring around $T=100\,$K. In this case the change in $W$
804: between $T=0$ and $200\,$K is 6.4\% which is much larger than for
805: the non interacting case (1.5\%). While the two models A and B have a
806: different Fermi surface, dispersion relation, and filling, the
807: most important parameter
808: \begin{figure}[tp]
809: \vspace*{-3mm}
810: %\centering
811: \includegraphics[width=9cm]{srFig3.eps}
812: \vspace*{-5mm}
813: \caption{Comparison of normal and superconducting state for the
814: optical sum and the kinetic energy. The left hand frame applies to
815: Model A of Table~\ref{tab:1} and $\omega_{SF}=82\,$meV and the
816: right hand frame is for the band structure Model B of Table~\ref{tab:1} and
817: $\omega_{SF} = 10\,$meV. The dotted line in this frame shows
818: a $T^2$ law extrapolation of the normal state data for $T>T_c$ to
819: zero temperature. Adapted from Ref.~\protect{\onlinecite{ref40}}.
820: }
821: \label{fig:3}
822: %\vspace*{5cm}
823: \end{figure}
824: which gives the deviation from the $T^2$ law noted above is the small
825: value of the spin fluctuation frequency $\omega_{SF}=10\,$meV
826: used in Model B. This implies that the thermally activated scattering is
827: larger in this model than it is for Model A. Later, when we consider
828: simpler versions of our interaction model, we will return to this
829: issue where it will become much easier to trace these dependencies.
830:
831: In Fig.~\ref{fig:3} we compare normal and superconducting state for the
832: two models of Table~\ref{tab:1}. We see that both KE and OS (open squares
833: and up triangles, respectively) fall below their normal state
834: values (solid squares and up triangles, respectively) at the same
835: temperature. This corresponds to an increase in KE as the
836: superconducting state is entered. The changes are small in all
837: cases. The KE in meV has increased by 0.2\% for Model A and
838: by 0.77\% in Model B between superconducting and normal state at
839: $T=0$. For the OS the differences are 0.22\% and 0.6\%, respectively,
840: very comparable to what is found for the KE but not identical. We
841: note that in the NAFFL model once the susceptibility \eqref{eq:7}
842: is specified, superconducting solutions with $d$-wave symmetry result.
843: These are not put in by hand as to symmetry or functional form.
844: The gap involves a harmonic $[\cos(ak_x)-\cos(ak_y)]$ as well as
845: many of the higher harmonics consistent with the
846: $d_{x^2-y^2}$ irreducible representation of the symmetry group
847: for the square CuO$_2$ lattice. The solutions are far from simple
848: BCS. In $d$-wave BCS an ansatz on the pairing potential
849: $V_{{\bf k},{\bf k}'}$ is made that it have the form
850: $\eta_{\bf k}V \eta_{{\bf k}'}$ with $\eta_{\bf k}\sim[\cos(ak_x)-\cos(ak_y)]$.
851: This leads to a $d$-wave gap consisting of the lowest harmonic only.
852: Early solutions\cite{ref38}
853: based on Model A with $\omega_{SF} = 7.76\,$meV showed that the gap
854: had maximum amplitude at the $X$-point $(\pi,0)$
855: of the two dimensional CuO$_2$
856: Brillouin zone with $\Delta_{max}^{BZ} = 33\,$meV. On the other
857: hand the maximum gap on the Fermi surface was $\Delta_{max}^{FS}=27\,$meV at
858: $T=20\,$K far below its BZ maximum. The value of
859: $2\Delta_{max}^{FS}/(k_BT_c) = 6.27$ is quite different from the BCS
860: value of 4.28. Also shown on the right hand panel is a dotted
861: straight line which indicates the least squares fit extrapolation of
862: the normal state data to zero temperature from $T>T_c$. It represents
863: a $T^2$ law. It is important to notice, for later reference, that both,
864: the superconducting as well as normal state $W$ data are above this
865: dotted line for all temperatures $T<T_c$.
866:
867: \subsection{The Hubbard Model and Dynamical Mean Field}
868: \label{ssec:2c}
869:
870: %Throughout this section we follow the notation of Toschi
871: %{\it et al.}\cite{ref42} and use the symbol $D$ for half the bandwidth
872: %in contrast to the main body of this review in which $D$ symbolizes the full
873: %band width.
874: %
875: Another approach used to treat strongly correlated systems such as
876: the cuprates is Dynamical Mean Field Theory (DMFT).\cite{ref41} The
877: approach is numerical and based on the Hubbard model with hopping
878: $t$, onsite Coulomb repulsion $U$, and chemical potential $\mu$.
879: The half bandwidth $D = 8\,t$ is taken to be $1.2\,$eV in what
880: follows with $U = 3D/2 = 12\,t$, so that the antiferromagnetic
881: super exchange $J = 4\,t^2/U\simeq 100\,$meV. For large values of
882: $4U/D$ at half filling the system is a Mott insulator. Away from
883: half filling, the weight of the quasiparticle peak at the Fermi
884: level denoted by $Z$ and related to the self energy $\Sigma(\omega)$
885: by
886: \[
887: % \label{eq:9}
888: Z = \left[1-\left.\frac{\partial\Sigma(\omega)}{\partial \omega}
889: \right\vert_{\omega=0}\right]^{-1}
890: \]
891: is non zero but small and is a measure of the metalicity.
892:
893: The Hamiltonian used has the form
894: \begin{equation}
895: \label{eq:9}
896: H = -t\sum\limits_{\langle ij\rangle\sigma}c^\dagger_{i\sigma}
897: c_{j\sigma}+U\sum\limits_i n_{i\uparrow} n_{j\downarrow}-
898: \mu\sum\limits_i\left(n_{i\uparrow}+n_{i\downarrow}\right),
899: \end{equation}
900: where $\sigma$ is spin $\uparrow$, $\downarrow$, $c_{i\sigma}$
901: ($c^\dagger_{i\sigma}$) are annihilation (creation) operators
902: for fermions of spin $\sigma$ on site $i$, $n_{i\sigma} =%
903: c^\dagger_{i\sigma}c_{i\sigma}$, and the sum $\langle ij\rangle$
904: \begin{figure}[tp]
905: % \vspace*{0.5cm}
906: % \centering
907: \includegraphics[width=8.3cm]{srFig4a.eps}
908: \includegraphics[width=7.9cm]{srFig4b.eps}
909: % \includegraphics[width=14cm]{Toschi1.eps}
910: % \includegraphics[width=8cm]{Toschi2.eps}
911: \caption{(Color online) $W(\Omega,T)$ normalized to its $T=0$ value
912: as a function of $T^2$ for the Hubbard model,
913: for $\Omega = 0.05\,D$ (top frame) and $\Omega = 0.75\,D$
914: (bottom frame).
915: Symbols are the results of the DMFT calculations; lines are best
916: fits to them. The various symbols give four dopings $(x)$.
917: Adapted from Ref.~\protect{\onlinecite{ref42}}.
918: }
919: \label{fig:4}
920: \end{figure}
921: is restricted to nearest neighbors only. In DMFT the full many body
922: system with strong interactions is modeled as a single site
923: problem with relevant interactions between, say, two electrons at
924: that site plus coupling to a bath which represents on
925: average the remaining degrees of freedom. The bath and local
926: problem is to be solved for in a self consistent way. The method is
927: now well developed and has proved its ability to simultaneously
928: describe low and high energy features in Mott systems.
929: Toschi {\it et al.}\cite{ref42} have calculated the real part
930: of the optical conductivity in this model and obtained the results
931: \begin{figure}[tp]
932: % \vspace*{5mm}
933: \includegraphics[width=8.5cm]{srFig5.eps}
934: \caption{The relative variation of spectral weight between the
935: lowest $T$ and $300\,$K as a function of doping $x$ for various
936: cuprates (full symbols) is compared with DMFT calculations
937: (open squares) and with the predictions of non interacting models
938: [tight binding model (solid line), constant density of states
939: approximation (dashed line)]. The
940: dotted line is a guide to the eye. The simple inclusion of
941: correlation effects allows one to reproduce the observed absolute values
942: with no need of fitting parameters. Data for LSCO are obtained from
943: Refs.~\protect{\onlinecite{ref43}} and \protect{\onlinecite{ref44}},
944: and for BSCCO from
945: Refs.~\protect{\onlinecite{ref45,ref46,ref47}}
946: (BSCCO$_1$) and from Ref.~\protect{\onlinecite{ref48}} (BSCCO$_2$).
947: Adapted from Ref.~\protect{\onlinecite{ref42}}.
948: }
949: \label{fig:5}
950: \end{figure}
951: \begin{figure}[tp]
952: \centering
953: \includegraphics[width=9cm]{srFig6.eps}
954: \caption{The $T=0$ spectral weight $W_0$ for $\Omega = 0.75D$ as a
955: function of doping $x$ for the Hubbard model (open squares) and the
956: experimental observation of Lucarelli {\it et al.}
957: Ref.~\protect{\onlinecite{ref44}} (solid circles) obtained by integrating
958: data on $\sigma_1(\omega)$ for LSCO. The dotted curve is for reference
959: and gives results for $U=0$ (no correlations included). Adapted by
960: A. Toschi from Ref.~\protect{\onlinecite{ref42}}.
961: }
962: \label{fig:4a}
963: \end{figure}
964: shown in Fig.~\ref{fig:4} for various values of the doping $x$ as
965: indicated in the figure which gives $W(\Omega,T)/W(\Omega,T=0)$
966: as a function of $(2T/D)^2$. The left hand frame employs a cutoff
967: $\Omega$ in Eq.~\eqref{eq:1} of $0.05D$ while the right hand
968: frame is for $\Omega = 0.75D$. All curves appear to follow a $T^2$
969: behavior and the slope of
970: these lines becomes smaller with increasing $\Omega$.
971: A comparison of their DMFT results with experiment is also
972: offered by the same authors and we reproduce it in Fig.~\ref{fig:5}
973: where we show
974: the difference $\Delta W \equiv W(T=0)-W(T=300\,{\rm K})$ for
975: $\Omega=1.5\,D$ renormalized to $W_0=W(T=0)$ denoted by $\Delta W/W_0$
976: on the figure as a function of doping $x$ (open squares).
977:
978: Experimental
979: results are for La$_{2-x}$Sr$_x$CuO$_4$ (LSCO, solid circles), Refs.~%
980: \onlinecite{ref43} and \onlinecite{ref44},
981: and Bi$_2$Sr$_2$CaCu$_2$O$_{8+x}$. Solid squares
982: are from Refs.~\onlinecite{ref45,ref46,ref47} (BSCCO$_1$) and solid
983: triangles from Ref.~\onlinecite{ref48} (BSCCO$_2$). Also shown for
984: comparison are results for tight binding with no interactions (solid line)
985: and in a constant density of states approximation (dashed line).
986: It is clear that
987: correlations dominate the observed change in $W$ between zero and
988: $300\,$K to its $T=0$ value.
989: DMFT provides a reasonable fit to the data while tight binding fails badly,
990: giving values that are too small.
991:
992: In Fig.~\ref{fig:4a} DMFT results (open squares) as a function of
993: doping $x$ are compared with the experimental results (sold circles)
994: of Lucarelli {\it et al.}\cite{ref44} for the optical sum
995: of LSCO at $T=0$.
996: The theory reproduces well the doping dependence of
997: $W_0\equiv W(\Omega=0.75D,0)$. While the absolute value of $W_0$ is
998: somewhat underestimated its large decrease as the Mott transition
999: is approached is well captured by the model calculations. Also shown
1000: on the same figure (dotted line) are results obtained in the limit
1001: $U=0$ (no Hubbard $U$). We see that in this case the observed trend
1002: with doping is not reproduced and $W_0$ is much too large, particularly
1003: at small dopings indicating, as expected, that near the Mott transition
1004: $W_0$ goes like $x$ rather than $1-x$. Finally, we note that the
1005: inclusion of $U$ (correlation effects) reduces $W_0$ just as we have
1006: seen in the NAFFL model. We note, however, that this model is not well
1007: suited to describe doping differences because the phenomenological
1008: spin susceptibility \eqref{eq:5} really needs to be fixed to some
1009: experimental observation at each doping level, particularly as the
1010: Mott transition is approached.
1011:
1012: One thing to note about the DMFT results shown in Fig.~\ref{fig:4} is
1013: that even for the smaller value of the cutoff $\Omega$ on the OS, a
1014: $T^2$ behavior is observed at least for the limited data available
1015: in the figure. By contrast, we have seen in the NAFFL calculations
1016: of Figs.~\ref{fig:2} and \ref{fig:3} that deviations from $T^2$ can
1017: occur when the characteristic spin fluctuation energy $\omega_{SF}$
1018: is small even when the sum is taken over the entire band, i.e.:
1019: $\Omega$ is large enough to include all contributions. Thus, the
1020: two calculations differ in this important point. We will see later
1021: that there is no guarantee that a given model for the interactions
1022: should always give a $T^2$ law.
1023:
1024: To end we mention related Hubbard
1025: model work by Maier {\it et al.}\cite{ref89c} They work directly
1026: with the KE rather than with the conductivity as did Toschi
1027: {\it et al.}\cite{ref42} and use the dynamical cluster approximation
1028: (DCA). They present results in both normal and superconducting
1029: state for two doping levels, namely $x = 0.05$ and
1030: $x=0.2$. In both cases the KE is found to decrease with
1031: decreasing temperature even in the superconducting state and this
1032: decrease, in fact, drives superconductivity. The changes are of order
1033: a few percent at most.
1034:
1035: As we have been so far mainly interested in
1036: the normal state we postpone further discussion of this work to
1037: a later section which deals explicitly with the superconducting state.
1038: We will also describe the recent work of Haule and Kotliar\cite{ref89b}
1039: which deals with the $t$-$J$ model (see Sec.~\ref{sec:4}).
1040:
1041: \section{The effect of interactions in isotropic boson
1042: exchange models}
1043: \label{sec:3}
1044: \subsection{Self Energy in Finite Bands}
1045: \label{ssec:3b}
1046:
1047: Recent studies of finite band effects have revealed that the self
1048: energy due to impurities or to interaction with bosons is profoundly
1049: changed by the application of a cutoff $\pm D/2$ in a half filled
1050: band.\cite{ref54,ref55,ref56,ref57,ref58,ref59,ref62}
1051: In a finite band, the normal state self energy due to the
1052: electron-phonon (or some other boson) interaction is given by
1053: \begin{subequations}
1054: \label{eq:23s}
1055: \begin{eqnarray}
1056: \Sigma(z) &=& T\sum\limits_m\lambda(z-i\omega_m)\eta(i\omega_m)\nonumber\\
1057: &&+
1058: \int\limits_0^\infty\!d\nu\,\alpha^2F(z)\left\{
1059: [f(T,\nu-z)+n(T,\nu)]\right.\nonumber\\
1060: &&\left.\times\eta(z-\nu)+[f(T,\nu+z)+n(T,\nu)]\eta(z+\nu)\right\},
1061: \nonumber\\
1062: \label{eq:23}
1063: \end{eqnarray}
1064: in the mixed representation of Marsiglio {\it et al.}\cite{ref60}
1065: Here,
1066: \begin{equation}
1067: \label{eq:24}
1068: \lambda(z) = \int\limits_0^\infty\!d\nu\alpha^2F(\nu)\frac{2\nu}
1069: {\nu^2-z^2}
1070: \end{equation}
1071: and
1072: \begin{equation}
1073: \label{eq:25}
1074: \eta(z) = \int\limits_{-\infty}^\infty\!d\epsilon\,
1075: \frac{N_0(\epsilon)}{N_0(0)}\frac{1}{z-\epsilon-\Sigma(z)}.
1076: \end{equation}
1077: \end{subequations}
1078: These equations need to be solved self consistently
1079: as $\Sigma(z)$ itself depends on itself through Eq.~\eqref{eq:25}.
1080: Here $z$ is a complex variable, $n(T,\nu)$ is the Bose thermal factor,
1081: $\alpha^2F(z)$ the electron-phonon spectral density and
1082: $N_0(\epsilon)/N_0(0)$ the normalized bare electronic density of
1083: states taken here to be constant and confined to $[-D/2,D/2]$.
1084: This provides a band cutoff in $\eta(z)$ of Eq.~\eqref{eq:25}
1085: and makes Eq.~\eqref{eq:23} self consistent. These equations are also
1086: given by Karakozov and Maksimov\cite{ref53} where they are written
1087: either in pure Matsubara notation or fully on the real frequency
1088: axis rather than in the mixed notation of Eqs.~\eqref{eq:23s} %to
1089: %\eqref{eq:25}
1090: which includes both versions, Matsubara and real axis.
1091: We have found the mixed notation more convenient for numerical
1092: work. For coupling to a
1093: single Einstein boson mode $\alpha^2F(\omega) = a\delta(\omega-\Omega_E)$
1094: where $\Omega_E$ and $a$ are taken in units of $D/2$. The electron mass
1095: renormalization is given by $\lambda = 2a/\Omega_E$ which is
1096: dimensionless. For an
1097: infinite band at $T=0$ we recover from Eqs.~\eqref{eq:23s} %to
1098: %\eqref{eq:25}
1099: the very familiar result\cite{ref62}
1100: \begin{equation}
1101: \label{eq:26}
1102: \Sigma_1(\omega) = a\ln\left\vert\frac{\omega-\Omega_E}{\omega+
1103: \Omega_E}\right\vert,
1104: \end{equation}
1105: and
1106: \begin{equation}
1107: \label{eq:27}
1108: \Sigma_2(\omega) = -\pi a \theta\left(\vert\omega\vert-\Omega_E\right),
1109: \end{equation}
1110: where $\theta(x)$ is the theta function equal to one for $x>0$ and
1111: zero otherwise. For positive values of $\omega$ the real part
1112: $\Sigma_1(\omega)\le 0$ and approaches zero as $\omega\gg\Omega_E$.
1113: The imaginary part $\Sigma_2(\omega)$ is zero till
1114: $\vert\omega\vert = \Omega_E$ at which point it jumps to a value
1115: of $\pi a$ and then stays constant. Finite band effects change this
1116: behavior radically.
1117:
1118: In Fig.~\ref{fig:7} we show results of Knigavko and Carbotte\cite{ref58}
1119: for minus the imaginary part $[-\Sigma_2(\omega)]$ of the self energy
1120: in frame (a) and its real part $[\Sigma_1(\omega)]$ in frame (b) for
1121: \begin{figure}[tp]
1122: % \centering
1123: \includegraphics[width=9cm]{srFig7.eps}
1124: \caption{Evolution of the self energy vs frequency dependence with
1125: temperature for the Einstein spectrum with $a=0.1$ and $\Omega_E =
1126: 0.1$ $(\lambda = 2)$. Frame (a) is for minus the imaginary part,
1127: $-\Sigma_2(\omega)$, while frame (b) is for the real part,
1128: $\Sigma_1(\omega)$. In each frame the different curves correspond
1129: to temperatures $t\equiv 2T/D =0.1$ (dash-dotted line),
1130: 0.05 (dotted line),
1131: 0.01 (dashed line), 0.001. All energies are in units of $D/2$.
1132: Adapted from Ref.~\protect{\onlinecite{ref58}}.
1133: }
1134: \label{fig:7}
1135: \end{figure}
1136: %\clearpage
1137: coupling to a single Einstein phonon at $\Omega_E = 0.1$ with
1138: mass enhancement parameter $\lambda = 2$ for different temperatures,
1139: all in units of $D/2$.
1140: The temperatures are as given in the figure caption.
1141: The sharper curves correspond to the lower temperatures.
1142: For $T=0$, $\Sigma_1(\omega)$ has a
1143: logarithmic like singularity at $\omega = \Omega_E$ as in the
1144: infinite band case of Eq.~\eqref{eq:26} but now, rather than remain
1145: negative as it goes towards zero for $\omega\gg \Omega_E$,
1146: $\Sigma_1(\omega)$ crosses the $\omega$-axis and takes on large
1147: positive values before dropping towards zero from above. Equally
1148: different from the infinite band case, $-\Sigma_2(\omega)$ is not
1149: a constant equal to $\pi a$ [see Eq.~\eqref{eq:27}] above
1150: $\omega>\Omega_E$ but rather drops as we approach the bare band
1151: edge at $\omega=1.0$ after which it becomes small for
1152: $\omega\stackrel{>}{\sim}1.5$ which is where the renormalized
1153: density of states $\tilde{N}(\omega)$ also becomes small. For further
1154: discussion of finite band features we refer the reader to Refs.
1155: \onlinecite{ref54} to \onlinecite{ref62} as well as to Ref.~%
1156: \onlinecite{ref53} where a somewhat different density of states
1157: model is used but this does not change the qualitative behavior of the
1158: self energy seen in Fig.~\ref{fig:7}. An advantage of Karakozov and
1159: Maksimov's choice of
1160: DOS, however, is that they can get a simple analytic expression
1161: for the self energy which they also show to be accurate by comparing
1162: with numerical results based on the full Eliashberg equations.
1163:
1164: We return next to the OS of Eq.~\eqref{eq:17}. % which can be rewritten
1165: %as:
1166: %\begin{equation}
1167: % \label{eq:28}
1168: % \bar{W}(T) = \frac{\hbar^2}{2m_b}\left[1-\frac{4}{(D/2)^2}
1169: % \int\limits_0^{D/2}\!d\epsilon\,\epsilon\int\limits_{-\infty}^\infty\!
1170: % d\omega\,f(T,\omega)A(\epsilon,\omega)\right].
1171: %\end{equation}
1172: One can get insight into the relative importance of the temperature
1173: dependence carried by the thermal factor $f(T,\omega)$ in
1174: $\bar{W}(T)$ and the temperature
1175: \begin{figure}[tp]
1176: % \centering
1177: \includegraphics[width=9cm]{srFig8.eps}
1178: \caption{The probability of occupation $n(\epsilon)$ of a state
1179: of energy $\epsilon$ vs
1180: normalized energy $2\epsilon/D$. (a) The results for the case when only
1181: the temperature dependence of the self energy is included. The normalized
1182: temperatures are $t \equiv 2T/D = 0.0$ (solid line),
1183: 0.01 (dashed line), 0.03 (dotted line), and 0.05
1184: (dash-dotted line). (b) Comparison of the
1185: results when the complete temperature dependence of $n(\epsilon)$
1186: [see Eq.~\protect{\eqref{eq:17}}] is accounted for (solid lines)
1187: with the case presented in frame (a). The normalized
1188: temperatures are $t=0.01$ (dashed line) and 0.05 (dash-dotted
1189: line). (All energies are in units of $D/2$.) A constant density of states
1190: model with cutoff $\pm D/2$ is used. Adapted from
1191: Ref.~\protect{\onlinecite{ref49}}.
1192: }
1193: \label{fig:8}
1194: \end{figure}
1195: dependence solely due
1196: to the interactions in $A(\epsilon,\omega)$. In Fig.~\ref{fig:8}
1197: we show results obtained by Knigavko {\it et al.}\cite{ref49}
1198: for $n(\epsilon)$, Eq.~\eqref{eq:10}.
1199: Frame (a) shows $n(\epsilon)$
1200: vs the renormalized energy $2\epsilon/D$ for four values of reduced
1201: temperatures $t=2T/D$,
1202: namely 0.0 (solid line), 0.01 (dashed line),
1203: 0.03 (dotted line), and 0.05 (dash-dotted line)
1204: when only the temperature dependence
1205: in $A(\epsilon,\omega)$ is included, i.e.: the $f(T,\omega)$ factor
1206: in $n(\epsilon)$ was excluded. Here we have considered coupling to a
1207: single phonon of energy $\Omega_E=0.04$ with mass enhancement
1208: $\lambda = 1$.
1209: It is clear that considerable temperature dependence comes from this
1210: source alone
1211: and will influence the temperature dependence of the
1212: OS. In frame (b) of Fig.~\ref{fig:8} we show results
1213: (solid curves)
1214: when both sources of $T$ dependence in $n(\epsilon)$ are included
1215: and compare with the two equivalent curves of frame (a) for
1216: $t=0.01$ (dashed line) and 0.05 (dash-dotted line).
1217: It is clear from these figures that the temperature
1218: dependence of the self energy is always important in determining
1219: the probability of occupation, $n(\epsilon)$, of a state of energy
1220: $\epsilon$ and hence the
1221: OS's $T$ dependence.
1222:
1223: \subsection{Results for a delta-function boson model}
1224: \label{ssec:3c}
1225:
1226: \begin{figure}[tp]
1227: % \centering
1228: \vspace*{-7mm}
1229: \includegraphics[width=9cm]{srFig9.eps}
1230: \caption{The variation of the optical sum $S$ vs reduced temperature
1231: $t = 2T/D$ for the
1232: interaction strengths (a) $a=0.02$ and (b) $a=0.1$. The results are for
1233: both the quadratic (solid symbols) and tight binding
1234: (open symbols) bands, as indicated. In the top frame
1235: the mass enhancement factor $\lambda=0.8$ (squares) and
1236: $\lambda=1$ (triangles), while in the bottom frame $\lambda=2$ (squares)
1237: and $\lambda=4$ (triangles).
1238: The inset in the top frame illustrates
1239: the behavior of the optical sum during a sudden ``undressing
1240: transition'' at $t_{undress}=0.02$ with 20\% hardening of the
1241: normalized boson frequency $\Omega$ and a corresponding reduction
1242: of the mass enhancement factor from $\lambda=1.0$ to 0.8. Adapted
1243: from Ref.~\protect{\onlinecite{ref49}}.
1244: }
1245: \label{fig:9}
1246: \vspace*{-3mm}
1247: \end{figure}
1248: Results for the OS normalized by a factor
1249: of $\hbar^2/(2m_b)$ for the tight binding case \eqref{eq:17}
1250: and by $\hbar^2\, n/m$ for the quadratic dispersion, both
1251: denoted by $S(t)$, are given in Fig.~\ref{fig:9} reproduced
1252: from Knigavko {\it et al.}\cite{ref49} The parameters are as shown on the
1253: figure and given in the caption. The top frame is for $a=0.02$ and
1254: the bottom frame for $a=0.1$. The former corresponds
1255: to a conventional metal while the latter may be more characteristic of
1256: the cuprates. Two representative values of the boson energy
1257: $\Omega_E$ have been chosen to get $\lambda = 0.8$
1258: (squares) and 1.0 (triangles)
1259: in the top frame and $\lambda = 2$ (squares) and
1260: $\lambda = 4$ (triangles) in the bottom frame.
1261: Both tight binding (open symbols)
1262: and quadratic (solid symbols) dispersion
1263: relations are considered. It is quite clear from
1264: these results that the temperature dependence of the OS needs not be
1265: quadratic and depends on the size of the coupling to the phonons
1266: ($\lambda$ value) and also on the dispersion relation used to describe
1267: the electron dynamics. An important observation for what will follow
1268: is contained in the inset of the top frame of Fig.~\ref{fig:9}. The
1269: heavy solid line follows the temperature evolution of the OS when
1270: a sharp ``undressing'' transition is assumed to take place at $t=0.02$
1271: where a 20\% hardening of the phonon spectrum occurs.
1272: This corresponds to a shift in the mass enhancement parameter from
1273: $\lambda=1$ to 0.8. Such a hardening of the phonon energy leads to
1274: an increase in the OS corresponding to a decrease in the KE of the
1275: charge carriers.\cite{ref49}
1276:
1277: \subsection{Results for an MMP spin fluctuation model}
1278: \label{ssec:3d}
1279:
1280: The delta function spectra used in the work of Knigavko {\it et al.}%
1281: \cite{ref49} while giving us insight
1282: into the mechanism that leads to temperature dependences in the OS is not
1283: realistic for the cuprates. If one wishes to remain within the framework of
1284: a boson exchange mechanism one should at the very least choose a different
1285: form for $\alpha^2F(\omega)$. For spin fluctuation exchange a much
1286: used choice is the MMP model of Millis {\it et al.}\cite{ref37,ref61,ref63}
1287: In its simplest form $\alpha^2F(\omega)$ which will now be denoted
1288: $I^2\chi(\omega)$ is taken as
1289: \begin{equation}
1290: \label{eq:29}
1291: I^2\chi(\omega) = I^2\frac{\omega/\omega_{SF}}{1+(\omega/\omega_{SF})^2},
1292: \end{equation}
1293: with $I$ a coupling between charge carriers and spin susceptibility
1294: and $\omega_{SF}$ a spin fluctuation frequency. Here, the coupling $I$ is
1295: adjusted to get a superconducting transition temperature $T_c$ of
1296: \begin{figure}[tp]
1297: % \centering
1298: \includegraphics[width=9cm]{srFig10.eps}
1299: \caption{Top frame: minus the imaginary part of the carrier self
1300: energy $-\Sigma_2(\omega,T)$ vs $\omega$ at the various temperatures
1301: shown. Middle frame: minus the real part of the carrier self energy
1302: $\Sigma_1(\omega,T)$ vs $\omega$. Bottom frame: the function
1303: $h(T,-\omega)$ of Eq.~\protect{\eqref{eq:18}} vs $\omega$. The
1304: calculation is for an MMP model, Eq.~\protect{\eqref{eq:29}}, with
1305: $\omega_{SF} = 20\,$meV, $I^2=0.82$, and $D=800\,$meV.
1306: }
1307: \label{fig:10}
1308: \end{figure}
1309: $100\,$K when used in the corresponding Eliashberg
1310: equations generalized for a $d$-wave gap symmetry. In this section we
1311: concentrate on the
1312: normal state only. Thus, Eqs.~\eqref{eq:23s} %and \eqref{eq:24}
1313: apply with
1314: $\alpha^2F(\omega)$ replaced by Eq.~\eqref{eq:29}. Results for a case
1315: with $\omega_{SF}=20\,$meV, $I^2=0.82$, and $D/2 = 400\,$meV are shown in
1316: Fig.~\ref{fig:10} which has three frames. In the top frame we show
1317: $-\Sigma_2(T,\omega)$ for four temperatures as labeled, in the center
1318: frame $\Sigma_1(T,\omega)$,
1319: and in the bottom frame $h(T,-\omega)$
1320: vs $\omega$. It is clear that both real and imaginary part of
1321: $\Sigma(T,\omega)$ have important $T$ dependencies which get
1322: reflected in $h(T,\omega)$ and consequently in the OS given by
1323: Eq.~\eqref{eq:17} and more approximately by Eq.~\eqref{eq:19} with
1324: \begin{figure}[tp]
1325: % \centering
1326: \includegraphics[width=9cm]{srFig11.eps}
1327: \caption{The reduced optical sum $S(T)$ according to
1328: Eq.~\protect{\eqref{eq:29a}}
1329: vs temperature $(T)$ squared. The left hand frame is for
1330: an MMP model \protect{\eqref{eq:29}} with $\omega_{SF} = 20\,$meV and
1331: $I^2=0.82$ while the right hand frame is for $\omega_{SF} = 82\,$meV
1332: and $I^2 = 0.655$.
1333: }
1334: \label{fig:11}
1335: \end{figure}
1336: the last term dropped. In this instance it is simply the area under
1337: $h(T,\omega)$ for negative $\omega$ that determines the OS while in
1338: Eq.~\eqref{eq:17} it is its overlap with the Fermi distribution and this
1339: has an additional temperature dependence due to $f(T,\omega)$ but it
1340: is small. Results for
1341: \begin{equation}
1342: \label{eq:29a}
1343: S(T) = 2\bar{W}(T)\frac{m_b}{\hbar^2}
1344: \end{equation}
1345: are shown in Fig.~\ref{fig:11} which has two frames. In the left hand
1346: frame we show results for $S(T)$ vs $T^2$ up to $40 000\,$K$^2$ for our
1347: MMP model with $\omega_{SF} = 20\,$meV and $I^2 = 0.82$ while in the
1348: right hand frame $\omega_{SF}$ is increased to $82\,$meV with
1349: $I^2=0.655$. In both cases the full band width $D$ ranges over 2500
1350: (dashed line), 800 (solid line), and $500\,$meV (dash-dotted line).
1351: It is clear that decreasing $D$ reduces the value of $S(T=0)$ and
1352: gives a stronger $T$ dependence which, nevertheless, remains near
1353: a $T^2$ behavior for the right hand frame but not in the left hand
1354: frame. These results are in qualitative agreement with those shown
1355: in the previous section, Figs.~\ref{fig:1} and \ref{fig:3} obtained
1356: in a tight binding model with a {\bf k}-point sampling method of
1357: the Brillouin zone and the full momentum dependent Eliashberg
1358: equations \eqref{eq:6}. Also shown in both frames as a dotted curve for the
1359: case $D=800\,$meV are straight lines which allow us to judge better
1360: the deviations from a $T^2$ law at small $T$ in the curves of the left
1361: hand frame. For the right hand frame there is little deviation.
1362: Note that both these behaviors are completely consistent with
1363: numerical results of Fig.~\ref{fig:3} and show that the upward
1364: deviation at small $T$ of the solid curve as compared with the
1365: dotted $(T^2)$ is a robust result not dependent on the model
1366: used, provided that the spin fluctuation frequency is small.
1367: Note also that Eq.~\eqref{eq:17} includes $T$ variations in $h(T,\omega)$
1368: as well as the thermal factor $f(T,\omega)$, but this latter factor
1369: can be neglected. The differences are very small
1370: confirming our previous claim that the temperature dependence
1371: in $S(T)$ is mainly due to variations in the interaction term
1372: at least for these cases. Finally, we stress again that the $T^2$
1373: law found in the DMFT calculations for the Hubbard model and
1374: often confirmed in experiments also arises in the NAFFL model with
1375: an MMP interaction provided the spin fluctuation energy is fairly large
1376: and $D$ not too small. In principle, however, there is no fundamental
1377: reason for an exact $T^2$ law. The actual $T$ dependence of $S(T)$
1378: comes from the $T$ dependence of the area under the curves for
1379: $h(T,-\omega)$ (bottom frame of Fig.~\ref{fig:10}). We see that
1380: these curves change most at small $\omega$ as $T$ is increased but
1381: there are also important changes at large $\omega$ and we are sampling
1382: an average of these changes. Nevertheless, the $\omega$
1383: dependence of $\alpha^2F(\omega)$ or $I^2\chi(\omega)$ at low energies
1384: $\omega$ will determine most strongly the temperature dependence of
1385: $h(T,-\omega)$ at those energies and, thus, plays an important role
1386: in the temperature dependence of $S(T)$. However, even at higher
1387: energies there is still significant $T$ dependence in $h(T,-\omega)$.
1388:
1389: \subsection{The non $T^2$ temperature dependence of the optical sum}
1390: \label{ssec:3e}
1391:
1392: Benfatto {\it et al.}\cite{ref52} have reconsidered the effect of
1393: \begin{figure}[tp]
1394: % \centering
1395: % \includegraphics[width=9cm]{fig3bw.eps}
1396: \includegraphics[width=8cm]{srFig12.eps}
1397: \caption{(Color online)
1398: The negative of the KE, $-\langle K\rangle$, vs temperature
1399: showing the breakdown for the various contributions. The solid
1400: curves are for a small boson frequency $\omega_E=2\,$meV with
1401: $\lambda=1$. The two thin lines show how the zero temperature value
1402: is modified by the Sommerfeld term by a very small amount. A similar
1403: size contribution (actually a bit larger) is illustrated for the
1404: noninteracting case by dotted curves, where only the
1405: Sommerfeld term is responsible for the temperature variation. Finally,
1406: for higher boson frequencies, the dashed curves illustrate
1407: the amount of temperature variation due to the Sommerfeld term
1408: compared with the rest. The band width $W \equiv D = 1\,$eV.
1409: Adapted from Ref.~\protect{\onlinecite{ref52}}.
1410: }
1411: \label{fig:12}
1412: \end{figure}
1413: interaction on the temperature dependence of $S(T)$ with an aim at
1414: providing simple analytic expressions that would help trace more
1415: directly the dependence on the interactions.
1416: In Fig.~\ref{fig:12} we reproduce some of their results for the
1417: temperature dependence of the negative of the KE, $-\langle K\rangle \equiv
1418: K(T)$, in meV. The two dotted lines at the top of the graph represent
1419: the noninteracting case $(\lambda=0)$ and are for reference. As expected,
1420: there is little difference between $K(T)$ and its zero temperature
1421: value $K(T=0)$. The three solid curves are for a system with coupling
1422: to a boson of frequency $\omega_E=2\,$meV and $\lambda=1$. We see that
1423: $K(T=0)$ and $K(T=0)+K_{\rm Somm}(T)$, the Sommerfeld contribution, are
1424: not significantly different. Thus, $K_{\rm Somm}(T)$ is not the
1425: important contribution to the temperature dependence of the complete KE
1426: [heavy solid line, denoted $K(T)$] which develops a linear temperature
1427: dependence for $T\sim20\,$K. The same holds for the case
1428: $\omega_E=25\,$meV (three dashed curves) but the variation in $T$ of
1429: $K(T)$ in this case is less pronounced and closer to a $T^2$ law as
1430: was found for the MMP spectrum. The band width in all three cases was
1431: $W \equiv D = 1\,$eV. Benfatto {\it et al.}\cite{ref52} trace these
1432: $T$ dependences in detail using analytic as well as numerical techniques.
1433: They concluded that both the real and imaginary parts of $\Sigma$
1434: contribute. Karakozov and Maksimov\cite{ref53} also find deviations
1435: from a $T^2$ law and provide an approximate analytic formula for
1436: these deviations in a particular case.
1437: For different models of $\alpha^2F(\omega)$ other
1438: $T$ dependencies are possible. This means that in principle one
1439: can learn about some features of this underlying $\alpha^2F(\omega)$
1440: from a study of the temperature dependence of the OS, but the
1441: correspondence is not necessarily simple or unique.
1442: On the other hand, a complete study of the
1443: temperature and frequency dependence of the optical self energy\cite{ref64}
1444: as is now done routinely, provides much more detailed information
1445: than does the OS.
1446: Recall that $h(T,\omega)$ is dependent on an average over the
1447: real and imaginary part of $\Sigma$, the quasiparticle self energy.
1448: This information is accessible directly from angular resolved
1449: photoemission spectroscopy (ARPES) at each point in the Brillouin
1450: zone separately.
1451:
1452: \begin{figure}[tp]
1453: \vspace*{-16mm}
1454: % \centering
1455: % \includegraphics[width=10cm]{srFig13.eps}
1456: \includegraphics[width=9cm]{srFig13.eps}
1457: \vspace*{-2cm}
1458: \caption{(Color online)
1459: The partial spectral weight integrated up to various
1460: frequencies as a function of temperature. Below $16\,000\,$cm$^{-1}$
1461: there is an increase in spectral weight as the temperature is lowered
1462: signaling a line narrowing on this frequency scale. Below $T_c$ there
1463: is strong loss of spectral weight to the superconducting condensate.
1464: There is no evidence of any precursors to superconductivity at
1465: $67\,$K. In the inset we show $N_{eff}(\omega)$ at $295\,$K. Adapted
1466: from Ref.~\protect{\onlinecite{ref64}}.
1467: }
1468: \label{fig:13}
1469: %\vspace*{-1cm}
1470: \end{figure}
1471: While many experiments give a $T^2$ dependence in the normal state
1472: within the precision of measurement there is some evidence for
1473: other dependences in the cuprates. In Fig.~\ref{fig:13} we reproduce
1474: data from the paper of Hwang {\it et al.}\cite{ref64} for the partial sum
1475: to $\omega_c$ in units of carriers per Cu atom denoted by
1476: $N_{eff}(T)$. What is shown is the difference $N_{eff}(T)-N_{eff}(300\,
1477: {\rm K})$ as a function of temperature $T$ to $300\,$K. The data is for
1478: underdoped
1479: YBa$_2$Cu$_3$O$_{6.50}$ (YBCO$_{6.50}$) Ortho II material of high
1480: quality and purity. There are 5 values of $\omega_c$, namely 1000, 5000,
1481: $10\,000$, $16\,000$, and $20\,000\,$cm$^{-1}$. It is clear that the
1482: temperature dependence of $N_{eff}(T)$ is not necessarily quadratic
1483: in $T$ for the partial sums and also for the case most relevant for
1484: our discussions with $\omega_c = 16\,000\,$cm$^{-1}$. These samples
1485: are, however, underdoped and a pseudogap is involved which could
1486: have some effect on the temperature dependence of the OS as we will
1487: discuss later within a phase fluctuation model for the pseudogap and
1488: alternatively a $D$-density wave model.
1489:
1490: \section{The Optical Sum in the Superconducting State}
1491: \label{sec:4}
1492: \subsection{Superconducting optical sum including of inelastic
1493: scattering collapse}
1494: \label{ssec:4a}
1495:
1496: \begin{figure}[tp]
1497: \vspace*{-4mm}
1498: \includegraphics[width=8.5cm]{srFig14.eps}
1499: \caption{Experimental values of the $ab$-plane spectral function
1500: $\rho_L$ defined in Eq.~\protect{\eqref{eq:30}}.
1501: Adapted from Ref.~\protect{\onlinecite{ref34}}.
1502: }
1503: \label{fig:14}
1504: %\vspace*{5cm}
1505: \end{figure}
1506: Next we return to a more detailed look at the superconducting state.
1507: In Fig.~\ref{fig:14} we reproduce the original results of
1508: Molegraaf {\it et al.}\cite{ref48} for two samples of BSCCO
1509: as they were presented by van der Marel {\it et al.}\cite{ref34}
1510: The top frame is for optimally doped and the bottom frame for
1511: an underdoped sample. What is shown is
1512: \begin{equation}
1513: \label{eq:30}
1514: \rho_L \equiv C\frac{\hbar^2}{\pi e^2}\int\limits_{-\Omega}^\Omega\!
1515: d\omega\,\Re{\rm e}[\sigma(\omega)]
1516: \end{equation}
1517: in meV with $C$ a constant to be specified shortly.
1518: It is striking that in these two samples the OS appears to
1519: increase faster in the superconducting state than in the underlying
1520: normal state extrapolated to low temperatures (dotted line). This
1521: was interpreted by Molegraaf {\it et al.}\cite{ref48} as indicating
1522: a decrease in KE in the superconducting state as compared with the
1523: underlying normal state which also shows a decrease in KE as $T\to 0$
1524: but by less than in the superconducting state. This behavior
1525: clearly goes beyond ordinary BCS theory and also Eliashberg theory as
1526: formulated for phonons.
1527:
1528: There remains some controversy about the interpretation of such data
1529: with some authors finding the opposite effect.\cite{ref65} We will not go into
1530: these details here but refer the reader to some relevant
1531: literature\cite{ref65,ref66,ref67} and note that Santander-Syro%
1532: \cite{ref47} confirm the basic results of Molegraaf {\it et al.}\cite{ref48}
1533: See also the recent work of Carbone {\it et al.}\cite{ref33a}
1534:
1535: We have already seen in Fig.~\ref{fig:9} [inset in frame (a)] for
1536: the normal state, that undressing effects associated with a hardening of
1537: the phonon spectra at a particular onset temperature can give an OS curve
1538: that behaves very much like those seen in Fig.~\ref{fig:14}.
1539: While in principle
1540: phonon frequencies can shift as a result of the onset of superconductivity
1541: these effects are small and usually negligible. For an electronic mechanism,
1542: however, we have already discussed in the introduction the idea of the
1543: collapse of the inelastic scattering as superconductivity sets in which
1544: leads directly to a peak in the real part of the microwave conductivity
1545: at some intermediate temperature below $T_c$. Within a spin fluctuation
1546: mechanism this translates into a hardening of the spin fluctuation spectrum
1547: and the modification of $I^2\chi(\omega)$ of Eq.~\eqref{eq:29} as the
1548: $d$-wave superconducting gap opens. An analysis of the optical scattering
1549: rates in YBCO$_{6.95}$ by Carbotte {\it et al.}\cite{ref68} showed that
1550: \begin{figure}[tp]
1551: \vspace*{7mm}
1552: \vspace*{-2mm}
1553: \includegraphics[width=8.5cm]{srFig15.eps}
1554: \caption{The charge carrier-spin spectral density $I^2\chi(\omega)$
1555: determined from optical scattering data of YBCO$_{6.95}$
1556: at various temperatures.
1557: Solid gray curve $T=90\,$K, dash-dotted $T=80\,$K, dashed $T=60\,$K,
1558: dotted $T=40\,$K, and black solid $T=10\,$K. Note the growth in strength
1559: of the $41\,$meV optical resonance as the temperature is lowered.
1560: Adapted from Ref.~\protect{\onlinecite{ref70}}.
1561: }
1562: \label{fig:15}
1563: \end{figure}
1564: there is a reduction in $I^2\chi(\omega)$ at small $\omega$ as the
1565: temperature is reduced. Results obtained by Schachinger
1566: {\it et al.}\cite{ref69,ref70} for the temperature evolution of
1567: $I^2\chi(\omega)$
1568: in YBCO$_{6.95}$ are reproduced in Fig.~\ref{fig:15}. We see a reduction in
1569: spectral weight at low $\omega$ as well as the growth of an optical
1570: resonance at $41\,$meV, the energy of the spin one resonance seen in
1571: neutron scattering.\cite{ref71} When these spectra are used in a
1572: \begin{figure}[tp]
1573: % \centering
1574: \vspace*{8mm}
1575: \includegraphics[width=8cm]{srFig16.eps}
1576: \caption{
1577: Temperature dependence of the conductivity $\sigma_1(\omega)$ at
1578: microwave frequency $\omega=0.144\,$meV in YBCO$_{6.95}$.
1579: The open circles are results based on
1580: our model spectral density (see Fig.~\protect{\ref{fig:15}}) obtained
1581: from inversion of optical conductivity data and the dashed line is
1582: based on an
1583: MMP model with low frequency cutoff. The solid triangles include impurities
1584: with the solid line from Ref.~\protect{\onlinecite{ref74}}. The solid
1585: squares represent experimental data by Bonn {\it et al.}\protect{%
1586: \cite{ref75}} Adapted from Ref.~\protect{\onlinecite{ref69}}.
1587: }
1588: \label{fig:16}
1589: \end{figure}
1590: generalization of the ordinary Eliashberg equations that
1591: includes the possibility of $d$-wave symmetry with gap $\Delta(\phi) =
1592: \Delta_0\cos(2\phi)$ where $\phi$ is an angle on the Fermi surface
1593: taken to be cylindrical, Schachinger and Carbotte,\cite{ref70,ref72}
1594: find excellent agreement for the microwave data of Hosseini {\it et al.}%
1595: \cite{ref73} taken at five different microwave frequencies on high quality
1596: samples of YBCO$_{6.99}$. In Fig.~\ref{fig:16}, reproduced from
1597: Ref.~\onlinecite{ref69}, we show older results obtained by Bonn {\it et
1598: al.}\cite{ref75} fit by Schachinger {\it et al.}\cite{ref74} by simply
1599: applying a low frequency cutoff to an MMP form.\cite{ref74,ref76} The data
1600: are given as the solid squares while the theoretical calculations without
1601: impurities (open circles and dashed line) and with impurities
1602: (solid triangles and solid line) are shown for comparison. Here
1603: $t^+$ gives the impurity scattering rate in Born approximation.
1604: Open circles and solid triangles are based on the spectra of Fig.~\ref{fig:15}
1605: which include the $41\,$meV spin resonance. The dashed and solid lines
1606: are from earlier calculations based on an MMP spectrum with application
1607: of a low frequency cutoff without resonance.\cite{ref74,ref76} We see
1608: that such a procedure gives results that are very close to those based
1609: on the more exact results of Fig.~\ref{fig:15}.
1610:
1611: Calculations in the model of Eqs.~\eqref{eq:6} of the OS and KE
1612: of optimally doped BSCCO have
1613: been carried out by Schachinger and Carbotte\cite{ref40} who simulated
1614: \begin{figure}[tp]
1615: % \centering
1616: \includegraphics[width=9cm]{srFig17.eps}
1617: \caption{The optical sum of optimally doped BSCCO
1618: as a function of the square of the temperature
1619: for the band structure Model A of Table~\protect{\ref{tab:1}} with
1620: different values of $\omega_{SF}$. Also in one case a low frequency
1621: cutoff is applied to the spin susceptibility. Note the significance
1622: of the $T^2$-variation on the value of $\omega_{SF}$. The thick solid
1623: line represents experimental normal state data of Molegraaf {\it et al.}%
1624: \protect{\cite{ref48}} also shown in the top frame of
1625: Fig.~\protect{\ref{fig:14}}.
1626: Adapted from Ref.~\protect{\onlinecite{ref40}}.
1627: }
1628: \label{fig:17}
1629: \end{figure}
1630: the expected hardening of the spin fluctuation spectra in the
1631: superconducting state by applying a low frequency cutoff to the
1632: interaction of Eq.~\eqref{eq:7}. This cutoff varies with temperature.
1633: It is zero at $T_c$ and has a maximum amplitude of $23\,$meV at $T=0$ with
1634: intensity changed according to a mean field BCS $d$-wave order parameter
1635: temperature dependence. Their results are shown in Fig.~\ref{fig:17}.
1636: The quantity $\rho_L$ according to Eq.~\eqref{eq:30} on the vertical
1637: axis is in meV and the horizontal scale is $T^2$ in K$^2$. The heavy
1638: solid line are normal state experimental data of Molegraaf {\it et al.},%
1639: \cite{ref48} also shown in the top frame of Fig.~\ref{fig:14} of this
1640: review. The theoretical results for the OS are all based on Model A
1641: of Table~\ref{tab:1} but were scaled upward by a factor
1642: [$C$ of Eq.~\eqref{eq:30}] of approximately
1643: two to fit the data. The spin fluctuation frequency $\omega_{SF}$
1644: was also adjusted to improve the fit; $\omega_{SF}=13\,$meV is best.
1645: To reduce the discrepancy in absolute value of the OS the value of
1646: the nearest neighbor hopping parameter $t$ in our band structure model
1647: would need to be increased. This would, however, also decrease the
1648: sensitivity of the OS to temperature variations and so $\omega_{SF}$
1649: would have to be adjusted downwards as well. Markiewicz {\it et al.}%
1650: \cite{ref77} have suggested significantly larger values of nearest
1651: neighbor hopping $t$ for
1652: the bare band structure of BSCCO than used to fit ARPES data. As is
1653: well known, there can be a factor of two or more. We have
1654: used the dispersion relation of Table I of Ref.~\onlinecite{ref77}
1655: (specifically we used $t=360\,$meV, $t'=-100\,$meV, and
1656: $\langle n\rangle = 0.4$. We also included further neighbors, namely
1657: $t''=35\,$meV and $t'''=20\,$meV)
1658: and found a value of $W(T=0)$ of $230\,$meV well above the
1659: experimental value with $\omega_{SF}=8\,$meV giving the proper
1660: temperature dependence. It is clear that a value of $t$ intermediate between
1661: that of Markiewicz {\it et al.}\cite{ref77} and the one in Table~\ref{tab:1}
1662: for Model A would be needed to get agreement with the OS data without any
1663: adjustment and a $\omega_{SF}$ value between 8 and $13\,$meV. Schachinger
1664: and Carbotte\cite{ref40} did not attempt such a fit
1665: as their aim was not
1666: to fit any particular case but rather see how a low frequency cutoff
1667: applied to a spin fluctuation model changes the OS. The solid
1668: squares in Fig.~\ref{fig:17} show the normal state results while the
1669: solid triangles are in the superconducting state. These are obtained
1670: without cutoff and so the KE has increased with respect to its normal
1671: state. The open squares and triangles show results for the same two
1672: cases but now the low frequency cutoff is applied to simulate the
1673: hardening of the spectrum as $T$ is reduced below $T_c$. We see a
1674: large decrease in KE as compared to the case without cutoff and the
1675: superconducting state would then show a decrease in KE as compared to
1676: the normal state without cutoff in agreement with the data of
1677: Fig.~\ref{fig:14}. It is the modification of the underlying interactions
1678: that have lead to this effect. It is clear that a modest hardening of
1679: the spectra consistent with the changes seen in the electron-boson
1680: spectral functions of Fig.~\ref{fig:15} would be sufficient to explain
1681: the data in Fig.~\ref{fig:14}. Of course, in a complete theory as yet
1682: not attempted, it would be necessary to find an interaction which
1683: can give not only the correct OS but also the microwave peak
1684: and all the
1685: other properties of the superconducting state. Finally, we note that
1686: in Fig.~\ref{fig:17} the OS shows a non $T^2$ behavior at low
1687: temperatures even without the low frequency cutoff which has the effect
1688: of making it more pronounced.
1689: If we did not have the normal state data below $T_c$ and simply
1690: extrapolated the normal state data above $T_c$ with a $T^2$ law
1691: to $T=0$ we would have to conclude that the OS is higher in
1692: the superconducting state
1693: than in the extrapolated ``normal state'' yet in this case the mechanism
1694: for pairing is not exotic in any way, i.e.: there is no low frequency cut
1695: off. (See also Fig.~\ref{fig:3}.) With the introduction of a low frequency
1696: cutoff this effect becomes even more pronounced and
1697: for the case shown the rise in the superconducting state is larger
1698: than seen in the experiments of Fig.~\ref{fig:14}. We caution the reader,
1699: however, that while this rise is seen by several experimental groups and is
1700: not controversial, its actual size is.
1701:
1702: \subsection{The Temperature Dependent Scattering Time Model}
1703: \label{ssec:4c}
1704:
1705: Recently Marsiglio\cite{ref90} has
1706: given calculations that provide support for the results of Fig.~\ref{fig:17}
1707: using a related but much simpler model. He notes that in an infinite band for
1708: an elastic scattering rate $\Gamma$ the probability of occupation of the
1709: state $\vert{\bf k}\rangle$ at finite temperature is given by
1710: \begin{equation}
1711: \label{eq:32}
1712: n_N(\epsilon_{\bf k}) = \frac{1}{2}-\frac{1}{\pi}\,\Im{\rm m}\psi\left(
1713: \frac{1}{2}+\frac{\Gamma}{4\pi T}+i\frac{\epsilon_{\bf k}}{2\pi T}
1714: \right)
1715: \end{equation}
1716: for the normal state and, to a good approximation, by
1717: \begin{equation}
1718: \label{eq:33}
1719: n_S(\epsilon_{\bf k}) = \frac{1}{2}-\frac{\epsilon_{\bf k}}{E_{\bf k}}
1720: \frac{1}{\pi}\Im{\rm m}\,\psi\left(
1721: \frac{1}{2}+\frac{\Gamma}{4\pi T}+i\frac{E_{\bf k}}{2\pi T}
1722: \right)
1723: \end{equation}
1724: for the superconducting state with $E_{\bf k} = \sqrt{\epsilon_{\bf k}^2+
1725: \Delta_{\bf k}^2}$,
1726: with $\Delta_{\bf k}$ the superconducting gap and
1727: $\psi(z)$ is the digamma function. Note that at zero temperature
1728: \begin{equation}
1729: \label{eq:34}
1730: n_S(\epsilon_{\bf k}) = \frac{1}{2}\left[1-\frac{\epsilon_{\bf k}}
1731: {E_{\bf k}}\frac{2}{\pi}\tan^{-1}\left(\frac{2E_{\bf k}}{\Gamma}\right)
1732: \right],
1733: \end{equation}
1734: which, in the limit $\Gamma\to 0$, reduces to the well known expression
1735: $(1-\epsilon_{\bf k}/E_{\bf k})/2$, Eq.~\eqref{eq:11a}
1736: introduced in Sec.~\ref{ssec:2d}.
1737: It is clear from Eq.~\eqref{eq:34} that both, the appearance of a
1738: gap in $E_{\bf k}$ and the elastic scattering $\Gamma$ smear
1739: $n_S(\epsilon_{\bf k})$.
1740: \begin{figure}[tp]
1741: % \centering
1742: % \includegraphics[width=11cm]{srFig21.eps}
1743: \includegraphics[width=9cm]{srFig18.eps}
1744: \caption{(Color online)
1745: Minus the kinetic energy, $-\langle T_{xx}\rangle$,
1746: vs temperature squared for various
1747: degrees of elastic scattering. Below $T_c$ we change the elastic
1748: scattering rate smoothly to zero [as $\Gamma_0 (T/T_c)^4$] as the
1749: temperature is lowered.
1750: Adapted from Ref.~\protect{\onlinecite{ref90}}
1751: }
1752: \label{fig:21}
1753: \end{figure}
1754: From these expressions for $n(\epsilon_{\bf k})$ it is easy
1755: to work out the average KE which Marsiglio denotes by $-\langle T_{xx}
1756: \rangle$. He uses a tight binding band and a constant
1757: $\Gamma_0 = 1/\tau = 0$, 2, 5, 10, and $15\,$meV from top to bottom
1758: in Fig.~\ref{fig:21}. For the superconducting state $\Gamma_0$ is
1759: changed to $\Gamma(T)$ with $\Gamma(T) = \Gamma_0(T/T_c)^4$ which is
1760: observed in the work of Hosseini {\it et al.}\cite{ref73} This
1761: introduces phenomenologically the idea of the collapse of the
1762: scattering rate. We see in Fig.~\ref{fig:21} that for $1/\tau = 5\,$meV
1763: and larger the KE decreases below its normal state extrapolation as the
1764: system becomes superconducting. The change is of the order of a few
1765: percent for the largest $1/\tau$ considered and is large enough to
1766: explain the measured KE changes in the underdoped cuprates.
1767:
1768: \subsection{The model of Norman and P\'epin}
1769: \label{ssec:4b}
1770:
1771: The mechanism described above which leads to an OS increase as
1772: superconductivity sets in has some common elements with the ideas
1773: of Norman and P\'epin\cite{ref78,ref79} although there are also
1774: important differences. These authors use a less fundamental, more
1775: phenomenological approach based on ARPES\cite{ref80} and optical
1776: conductivity data from which they construct directly a model for
1777: the self energy $\Sigma(\omega)$. For the normal state they begin
1778: with a constant, frequency independent scattering rate
1779: $\Gamma_0 = \Im{\rm m}\Sigma(\omega)$ (taken to be of order
1780: $100\,$meV) to simulate the very broad (incoherent) line shapes seen
1781: in ARPES at the anti nodal point $(0,\pi)$. In the superconducting
1782: state a sharp coherence peak appears which signals increased coherence.
1783: To model this fact a low frequency cutoff $\omega_0$ is applied
1784: to the imaginary part of $\Sigma(\omega)$ making it zero for
1785: $\omega\le\omega_0$ and $\Gamma_0$ above. The value of $\omega_0$
1786: is chosen as the energy of the spectral dip feature seen in ARPES
1787: in the superconducting state. Here, as in the model described
1788: in Sec.~\ref{ssec:4c},
1789: the underlying normal state on top of which superconductivity
1790: develops effectively has reduced scattering, i.e.: is more coherent
1791: which is a critical additional feature not contained in ordinary
1792: BCS and this leads to a reduction in KE and, consequently, in an increase
1793: of the OS. To arrive at their final estimates for the KE increase
1794: Norman and P\'epin include further complications in their model
1795: such as the anisotropy observed in scattering rates\cite{ref81,ref82,ref83,%
1796: ref84} ($\Gamma_{\bf k}$ instead of $\Gamma_0$)
1797: \begin{figure}[tp]
1798: % \centering
1799: %\vspace*{5mm}
1800: \includegraphics[width=7cm]{srFig19a.eps}
1801: \includegraphics[width=7cm]{srFig19b.eps}
1802: \caption{Top frame:
1803: The optical scattering rate $1/\tau(\omega)$ vs $\omega$
1804: for various BSCCO samples from
1805: Ref.~\protect{\onlinecite{ref87}} (OD overdoped, OPT optimal doped,
1806: UD underdoped). Bottom frame:
1807: Calculated sum rule violation $(-\Delta E_K)$ vs
1808: doping $x$ (solid circles).
1809: The curve is $T_c$. Also shown in this frame are the
1810: experimental results (open squares from Ref.~\protect{\onlinecite{ref46}},
1811: open diamonds from Ref.~\protect{\onlinecite{ref48}}). The theoretical
1812: doping trend in (b) is due to the increasing offset in $1/\tau$ seen
1813: in (a). Adapted from Ref.~\protect{\onlinecite{ref78}}.
1814: }
1815: \label{fig:18}
1816: \end{figure}
1817: going from anti nodal to nodal direction as well as an
1818: $\omega$ dependence (proportional to a momentum independent parameter
1819: $\alpha$) in the self energy modeled on the Marginal
1820: Fermi Liquid model\cite{ref85,ref86} with parameters determined
1821: from optical
1822: data on scattering rates. We reproduce in Fig.~\ref{fig:18} their
1823: final estimates for the KE change $-\Delta E_K$ between superconducting
1824: and normal state. The left hand frame shows the optical scattering rate
1825: denoted $1/\tau(\omega)$ for four
1826: BSCCO samples from Puchkov {\it et al.}\cite{ref87} which they use
1827: to fit parameters. The right hand frame presents the results for the change in
1828: KE $(-\Delta E_K)$ associated with the formation of the superconducting
1829: state in
1830: meV as a function of doping $x$. The solid circles give the theoretical
1831: results and the open squares and open diamonds experimental data from
1832: Santander-Syro {\it et al.}\cite{ref46} and Molegraaf {\it et al.},\cite{ref48}
1833: respectively. The theoretical estimates are deemed reasonable
1834: but not accurate. On the overdoped side the OS behaves in a conventional
1835: fashion while for the underdoped side it is anomalous representing a
1836: lowering of KE as compared with what is expected in conventional BCS.
1837: Note that the lowest open square in this graph shown as being zero for
1838: the overdoped sample represents an early not very accurate estimate. In
1839: Ref.~\onlinecite{ref46} it is actually negative rather than zero
1840: bringing it closer in agreement with the lowest solid circle (theory).
1841: This point was further discussed by Deutscher {\it et al.}\cite{ref87a}
1842: %taken from
1843: %Ref.~\onlinecite{ref48} for an overdoped sample falls at negative values
1844: %of $-\Delta E_K$ rather than at zero\cite{ref87a,ref88} as shown in
1845: %Fig.~\ref{fig:18} bringing it in closer agreement with the lowest solid
1846: %circle (theory).
1847:
1848: \subsection{Additional data on kinetic energy changes}
1849: \label{ssec:4e}
1850:
1851: \begin{figure}[tp]
1852: % \centering
1853: \includegraphics[width=8cm]{srFig20.eps}
1854: \caption{Spectral weight of the overdoped Bi2212 sample, integrated
1855: up to $1\,$eV, plotted vs $T^2$, from Ref.~\protect{\onlinecite{ref46}}.
1856: Closed symbols: spectral weight in the normal state, open symbols:
1857: spectral weight in the superconducting state, including the weight
1858: of the superfluid. The errors in the {\it relative} variations of the
1859: spectral weight are of the size of the symbols. Adapted from
1860: Ref.~\protect{\onlinecite{ref87a}}.
1861: }
1862: \label{fig:19}
1863: \end{figure}
1864: We reproduce in Fig.~\ref{fig:19} the data of Santander-Syro {\it et al.}%
1865: \cite{ref46} for their overdoped sample as presented by Deutscher
1866: {\it et al.}\cite{ref87a} What is shown is the optical spectral weight
1867: in units $10^6\,\Omega^{-1}\,{\rm cm}^{-2}$ as a function of $T^2$.
1868: The solid circles are in the normal state and the open circles in the
1869: superconducting state below $T_c = 63\,$K. A clear $T^2$ law is noted
1870: above $T_c$ and a reduction in spectral weight below, as expected in
1871: ordinary BCS theory. For the optimally doped sample
1872: \begin{figure}[tp]
1873: % \centering
1874: \includegraphics[width=8cm]{srFig21.eps}
1875: \caption{Change $\Delta E_{kin}$ of the KE, in meV per copper site
1876: vs the charge $p$ per copper site with
1877: respect to $p_{opt}$ [Eq.~\protect{\eqref{eq:31}}] in BSCCO.
1878: Full diamonds: data from
1879: Ref.~\protect{\onlinecite{ref47}}, high-frequency cutoff $1\,$eV. Open
1880: circles: data from Ref.~\protect{\onlinecite{ref48}}, high frequency
1881: cutoff $1.25\,$eV. Error bars: vertical, uncertainties due to the
1882: extrapolation of the temperature dependence of the normal state spectral
1883: weight down to zero temperature; horizontal, uncertainties resulting
1884: from $T_c/T_{c,max}$ through Eq.~\protect{\eqref{eq:31}}.
1885: Deutscher {\it et al.}\protect{\cite{ref87a}} took
1886: $T_{c,max} = (83\pm2)\,$K for films and $(91\pm2)\,$K for crystals.
1887: Adapted from Ref.~\protect{\onlinecite{ref87a}}.
1888: }
1889: \label{fig:20}
1890: \end{figure}
1891: (not shown) the normal state temperature dependence is closer to
1892: linear than quadratic and the underdoped sample shows a very
1893: flat region before entering the superconducting state. This demonstrates
1894: once more that there is as yet no strong consensus in the literature
1895: as to the $T$ dependence of the normal state. (Please see also
1896: Ref.~\onlinecite{ref89a}.)
1897: The same paper also analyses the data in terms
1898: of KE change for three samples UND70K, OPT80K, and OVR63K, as shown
1899: in Fig.~\ref{fig:20} reproduced from Deutscher {\it et al.}\cite{ref87a}
1900: The horizontal axis is $p-p_{opt}$ where $p$ is the charge per Cu atom related
1901: to $T_c$ by
1902: \begin{equation}
1903: \label{eq:31}
1904: \frac{T_c}{T_{c,opt}} = 1-86.2\left(p-p_{opt}\right)^2,
1905: \end{equation}
1906: with $T_{c,opt}$ the maximum critical temperature for the Bi2212 series.%
1907: \cite{ref89} What is clear from this figure is that there is a smooth
1908: crossover from standard behavior on the overdoped side to anomalous
1909: behavior on the underdoped side. The very recent data of
1910: Carbone {\it et al.}\cite{ref33a} lends further support to this
1911: conclusion.
1912:
1913: \subsection{Cluster dynamical mean field of the $t$--$J$ model}
1914: \label{ssec:4d}
1915:
1916: Recently Haule and Kotliar\cite{ref89b} calculated the optical
1917: conductivity of the $t$--$J$ model within a cluster DMFT (CDMFT).
1918: (Please see also earlier work by Maier {\it et al.}\cite{ref89c}
1919: \begin{figure}[tp]
1920: % \centering
1921: % \includegraphics[width=12cm]{srFig32.eps}
1922: \includegraphics[width=8cm]{srFig22.eps}
1923: \caption{(Color online)
1924: The difference between the superconducting and normal state
1925: energies as a function of temperature. The following curves are shown:
1926: up-triangles - $E_{kin-S}-E_{kin-N}$; squares - $E_{x-S}-E_{x-N}$;
1927: diamonds - $E_{tot-S}-E_{tot-N}$; circles - $\mu_S-\mu_N$.
1928: Adapted from Ref.~\protect{\onlinecite{ref89b}}.
1929: }
1930: \label{fig:32}
1931: \end{figure}
1932: based on the Hubbard model.) Haule and Kotliar\cite{ref89b} treat
1933: the temperature and doping
1934: dependence and address the issue of the change in KE of the holes
1935: when superconductivity sets in. They find, in agreement with
1936: experiment, that on the overdoped side the KE makes a negative
1937: contribution to the condensation energy as in conventional BCS theory
1938: but that there is a crossover to the opposite case on the underdoped
1939: side. This is shown in Fig.~\ref{fig:32} reproduced from
1940: Ref.~\onlinecite{ref89b}. What is shown is the difference between the
1941: superconducting and normal state energies $(E)$ as a function of
1942: temperature, both in units of $t$ (nearest neighbor hopping). The
1943: up-triangles give the KE contribution $(E_{kin})$,
1944: the squares are the super exchange $(E_x)$,
1945: and the diamonds the total energy $(E_{tot})$.
1946: We see a change in sign in the KE
1947: contribution upon condensation as we go from the overdoped to
1948: the underdoped regime. ($\delta$ denotes the doping.) These results
1949: contrast with earlier work of Maier {\it et al.}\cite{ref89c} also based
1950: on the Hubbard model. There, the KE is found to be lower in the
1951: superconducting state than in the normal state for both values
1952: of doping presented, namely $\delta=0.05$ and $\delta=0.2$ (overdoped).
1953: In the underdoped case the potential energy increases slightly
1954: above its normal state value below $T=T_c$. For the overdoped case
1955: it does decrease very slightly but plays a much smaller role in the
1956: condensation than the corresponding drop in KE.
1957:
1958: We end this section by mentioning two related works\cite{ref127,ref128}
1959: based on the negative $U$ Hubbard model which has been used
1960: successfully to describe the BCS - Bose Einstein (BE) crossover.
1961: Toschi {\it et al.}\cite{ref127} employ DMFT and Kyung {\it et al.}%
1962: \cite{ref128} a cellular DMFT and obtain very similar results. Both
1963: normal and superconducting states are considered. Both groups find
1964: a change of sign in the KE difference between superconducting and
1965: \begin{figure}[tp]
1966: \centering
1967: \includegraphics[width=9cm]{srFig23.eps}
1968: \caption{(Color online) Low-temperature behavior of the normalized kinetic
1969: energy $E_{\rm kin}/D$ (upper panels) and the normalized potential energy
1970: $E_{\rm pot}/\vert U\vert$ (lower panels) as a function of the normalized
1971: temperature $T/D$. Here $U$ is the Hubbard attraction.
1972: The critical temperature is marked by arrows. Adapted from
1973: Ref.~\protect{\onlinecite{ref127}}.}
1974: \label{fig:32a}
1975: \end{figure}
1976: normal state as one goes from underdoping (anomalous) to overdoping
1977: (conventional BCS behavior) as seen in Fig.~\ref{fig:20}. In
1978: Fig.~\ref{fig:32a} we reproduce from Ref.~\onlinecite{ref127} results
1979: for KE $(E_{\rm kin})$ in units of the band width $D$ vs the
1980: normalized temperature $T/D$ (upper frames) and for the potential energy
1981: $E_{\rm pot}$ in units of $\vert U\vert$ (lower frames). Three values
1982: of $U$ are considered, $U=0.4\,D$ in the left frames and $U=1.2\,D$
1983: and $3.2\,D$ in the frames on the right. These values correspond to
1984: BCS, intermediate, and BE (Bose - Einstein) regimes, respectively. The frames
1985: on the left show conventional behavior with a small increase in KE
1986: in the superconducting state and a decrease in potential energy.
1987: On the other hand, in the frames on the right the KE decreases
1988: while the potential
1989: energy increases which is referred to as KE driven superconductivity.
1990:
1991: \section{Models of the Pseudogap State}
1992: \label{sec:5}
1993: \subsection{The Preformed Pair Model}
1994: \label{ssec:5a}
1995:
1996: Another model for superconductivity in the oxides is the preformed
1997: pair model.\cite{ref91,ref92,ref93,ref94,ref95,ref96} The idea is
1998: that the pairs form at a temperature $T^\ast$, the pseudogap formation
1999: temperature. The superconducting state emerges from the preformed
2000: pair state at a lower temperature $T_c$ when phase coherence sets in. The
2001: pseudogap regime is then due to superconducting phase fluctuations.
2002: In a recent paper Eckl {\it et al.}\cite{ref97}
2003: considered the effect of phase fluctuations on the OS, i.e.: the KE
2004: in the pseudogap regime. [See also the related work of
2005: Kope\'c, Ref.~\onlinecite{ref97a}.]
2006: They start with a Hamiltonian which contains
2007: two terms, the KE
2008: \begin{equation}
2009: \label{eq:35}
2010: K = -t\sum\limits_{\langle ij\rangle\sigma}\left(
2011: c^\dagger_{i\sigma}c_{j\sigma}+c^\dagger_{j\sigma}c_{i\sigma}\right)
2012: \end{equation}
2013: and a pairing term
2014: \begin{equation}
2015: \label{eq:36}
2016: -\frac{1}{4}\sum\limits_{i\delta}\left(\Delta_{i\delta}\left\langle
2017: \Delta^\dagger_{i\delta}\right\rangle+\Delta^\dagger_{i\delta}
2018: \left\langle\Delta_{i\delta}\right\rangle\right),
2019: \end{equation}
2020: where $\delta$ connects the site $i$ to its nearest neighbors only and
2021: $\langle ij\rangle$, as before, is limited to nearest neighbors as well.
2022: Furthermore,
2023: \begin{equation}
2024: \label{eq:37}
2025: \Delta^\dagger_{i\delta} = \frac{1}{\sqrt{2}}\left(c^\dagger_{i\uparrow}
2026: c^\dagger_{i+\delta\downarrow}-c^\dagger_{i\downarrow}
2027: c^\dagger_{i+\delta\uparrow}\right).
2028: \end{equation}
2029: Its average $\left\langle\Delta^\dagger_{i\delta}\right\rangle\equiv\Delta\,
2030: \exp\left(i\Phi_{i\delta}\right)$ with $\Delta$ the gap amplitude is
2031: assumed to have $d$-wave symmetry and the phase
2032: \begin{equation}
2033: \label{eq:38}
2034: \Phi_{i\delta} = \begin{cases}
2035: \left(\phi_i+\phi_{i+\delta}\right)/2 & \mbox{for $\delta$ along $x$},\\
2036: \left(\phi_i+\phi_{i+\delta}\right)/2+\pi &
2037: \mbox{for $\delta$ along $y$}.
2038: \end{cases}
2039: \end{equation}
2040: The phases are assumed to fluctuate according to classical XY free energy.
2041: In this model the Kosterlitz-Thouless transition $T_{KT}$ is identified with
2042: \begin{figure}[tp]
2043: % \centering
2044: \includegraphics[width=8cm]{srFig24.eps}
2045: \caption{Kinetic energy per bond, $\langle k_x\rangle$, as a function of
2046: temperature for non interacting tight-binding electrons (TB), the BCS
2047: solution (BCS), and the phase fluctuation (PP) model for $\mu=0$
2048: $(\langle n\rangle = 1)$. The large vertical arrows indicate the increase
2049: in KE upon pairing, relative to the free tight-binding model, and
2050: the small arrows indicate the additional increase due to phase
2051: fluctuations. This additional phase fluctuation energy rapidly
2052: vanishes near $T_c=T_{KT}$, which causes the significant change in the
2053: OS upon entering the superconducting state at $T_{KT}=0.1\,t$. Note
2054: that the thick line follows the actual KE encountered in this model, when
2055: going from the pseudogap to the superconducting regime. Adapted from
2056: Ref.~\protect{\onlinecite{ref97}}.
2057: }
2058: \label{fig:22}
2059: \end{figure}
2060: the superconducting transition temperature $T_c$ and the mean field
2061: temperature $T_{MF}$ at which the pairs form, is identified with the
2062: pseudogap temperature $T^\ast$. They take $T_{KT}\simeq T_{MF}/4$ with
2063: $T_{KT}=0.1\,t$. In Fig.~\ref{fig:22}, reproduced from Ref.~\onlinecite{%
2064: ref97}, we show the KE per bond, $\langle k_x\rangle$ as a function
2065: of the reduced temperature $T/T_{KT}$. Here
2066: \begin{equation}
2067: \label{eq:39}
2068: \langle k_x\rangle = -t\sum\limits_\sigma\left\langle
2069: c^\dagger_{i\sigma}c_{i+x,\sigma}+c^\dagger_{i+x.\sigma}c_{i\sigma}
2070: \right\rangle.
2071: \end{equation}
2072: The dashed curve is the result of the simple tight-binding
2073: Hamiltonian [$K$ of Eq.~\eqref{eq:35}
2074: only] for zero chemical potential and $\langle n\rangle=1$. The light
2075: solid line is the result of BCS mean field and the heavy solid
2076: line the result obtained by taking into account phase fluctuations
2077: within the preformed pair model. The mean field BCS condensation
2078: increases the KE above its
2079: tight-binding value and the phase fluctuations provide an additional
2080: KE increase which vanishes at $T_c=T_{KT}$ and this causes significant
2081: \begin{figure}[tp]
2082: % \centering
2083: \includegraphics[width=8cm]{srFig25.eps}
2084: \caption{Kinetic energy contribution from phase fluctuations
2085: $\delta\langle k_x\rangle \equiv \langle k_x\rangle_{PP}-
2086: \langle k_x\rangle_{BCS}$. One can clearly see the sharp decrease of KE
2087: near the Kosterlitz-Thouless transition $T_{KT}=0.1\,t\equiv T_c$.
2088: $\Delta E_k$ gives an estimate of the kinetic condensation energy.
2089: Adapted from Ref.~\protect{\onlinecite{ref97}}.
2090: }
2091: \label{fig:23}
2092: \end{figure}
2093: change in the OS as the superconducting state is entered at
2094: $T_{KT}=0.1\,t$ in this model calculations. The KE gain from the
2095: phase fluctuations $\delta\langle k_x\rangle \equiv \langle k_x\rangle_{PP}-
2096: \langle k_x\rangle_{BCS}$ is shown in Fig.~\ref{fig:23} which shows
2097: a sharp decrease in KE near the Kosterlitz-Thouless transition and
2098: $\Delta E_k$ gives an estimate of the KE condensation $\propto 0.003\,t$
2099: of the same order as was found in other models. However, this is not the
2100: KE change between normal and superconducting state at $T=0$
2101: as discussed previously.
2102: Rather it is a change in KE due to the suppression of phase
2103: fluctuations present in the pseudogap state. At $T_c=T_{KT}$ the
2104: phases become locked in and $\delta\langle k_x\rangle$ consequently
2105: becomes zero. The change in KE on formation of the Cooper pairs is
2106: never explicitly considered in this model as the pairs are assumed to
2107: form at a much higher temperature $T=T^\ast = 0.4\,t$.
2108:
2109: \subsection{The $D$-density Waves, Competing Order Model}
2110: \label{ssec:5b}
2111:
2112: There are other very different models that have been proposed for the
2113: pseudogap state which exists at temperatures above the superconducting
2114: state in the underdoped regime. One proposal is $D$-density waves\cite{%
2115: ref98,ref99,ref100,ref101,ref102,ref103,ref103a,ref104,ref105,ref106}
2116: (DDW) which
2117: falls into the general category of competing interactions. The view is
2118: that a new phase, not superconducting, but having a gap with $d$-wave
2119: symmetry,
2120: forms at $T^\ast$ and the superconductivity which arises only at some
2121: lower temperature $T_c$ is based on this new ground state.\cite{%
2122: ref98,ref99,ref100,ref101,ref102,ref103,ref103a,ref104,ref105,ref106,%
2123: ref107,ref108,ref109,ref110,ref111,ref112,ref113,ref114} The DDW state breaks
2124: time reversal symmetry because it introduces bond currents\cite{ref115,%
2125: ref116} with associated small magnetic moments and a gap forms at the
2126: antiferromagnetic Brillouin zone boundary. The mean field DDW Hamiltonian is
2127: \begin{equation}
2128: \label{eq:40}
2129: H = \sum\limits_{{\bf k},\sigma}\left[\left(\epsilon_{\bf k}-
2130: \mu\right) c^\dagger_{{\bf k}\sigma}c_{{\bf k}\sigma}+i D_{\bf k}
2131: c^\dagger_{{\bf k}\sigma}c_{{\bf k}+{\bf Q}\sigma}\right],
2132: \end{equation}
2133: where the sum on {\bf k} ranges over the entire Brillouin zone and
2134: {\bf Q} is the commensurate wave vector $(\pi/a,\pi/a)$. Here
2135: $D_{\bf k} = D_0\left[\cos(k_xa)-\cos(k_ya)\right]/2$ with $D_0$ the
2136: gap amplitude. Within a mean field approximation
2137: the probability of occupation of
2138: the state $\vert{\bf k}\sigma\rangle$ is given by\cite{ref112}
2139: \begin{eqnarray}
2140: n_{{\bf k}\sigma}(T) &=& \frac{1}{2E_{\bf k}}\left\{E_{\bf k}\left[
2141: f(T,\xi_{+{\bf k}})+f(T,\xi_{-{\bf k}})\right]\right.\nonumber\\
2142: &&\left.+\epsilon_{\bf k}\left[
2143: f(T,\xi_{+{\bf k}})-f(T,\xi_{-{\bf k}})\right]\right\},
2144: \label{eq:41}
2145: \end{eqnarray}
2146: where $\xi_{\pm{\bf k}} = -\mu\pm E_{\bf k}$ and $E_{\bf k} = \sqrt{%
2147: \epsilon_{\bf k}^2+D_{\bf k}^2}$. The number density is
2148: \begin{equation}
2149: \label{eq:42}
2150: n(T) = \frac{2}{N}\sum\limits_{{\bf k}\in{\rm MBZ}}\left[f(T,\xi_{-{\bf k}})+
2151: f(T,\xi_{+{\bf k}})\right],
2152: \end{equation}
2153: where the sum over {\bf k} is restricted to half the original
2154: Brillouin zone, i.e.: to the magnetic Brillouin zone (MBZ).
2155: The KE from Refs. \onlinecite{ref108}, \onlinecite{ref113}, and
2156: \onlinecite{ref114} is denoted by $W(D,T)$ and is
2157: \begin{equation}
2158: \label{eq:43}
2159: W(D,T) = \frac{2}{N}\sum\limits_{{\bf k}\in{\rm MBZ}}
2160: \frac{\epsilon_{\bf k}^2}
2161: {E_{\bf k}}\left[f(T,\xi_{+{\bf k}})-f(T,\xi_{-{\bf k}})\right].
2162: \end{equation}
2163: Here and later in this section we give formulas only for the simpler
2164: case of $t'=0$, i.e.: only nearest neighbors. The results to be presented,
2165: however, are based on straightforward generalizations that include
2166: second nearest neighbors hopping.
2167:
2168: Aristov and Zeyher\cite{ref119} calculated the optical conductivity in
2169: the DDW model with and without vertex corrections to the usual current
2170: operator. The previous work of Valenzuela {\it et al.}\cite{ref110} was
2171: without vertex corrections as is the more recent work of Gerami and
2172: Nayak\cite{ref120} who, however, consider the effect of anisotropic
2173: scattering. These authors use a current operator which includes a term
2174: proportional to the gap velocity $\nabla_{\bf k}D_{\bf k}$ which is
2175: \begin{figure}[tp]
2176: \vspace*{-4mm}
2177: \includegraphics[width=9cm]{srFig26.eps}
2178: \caption{Restricted optical sums with (solid line) and without
2179: (dot-dashed line) vertex corrections, and Drude weights with (dashed
2180: line) and without (dotted line) vertex corrections.
2181: Adapted from Ref.~\protect{\onlinecite{ref119}}.
2182: }
2183: \label{fig:24}
2184: \end{figure}
2185: introduced so as to ensure charge conservation as is discussed in more
2186: detail by Benfatto {\it et al.}\cite{ref113} Here we follow
2187: Aristov and Zeyher\cite{ref119} and show in Fig.~\ref{fig:24} their
2188: results for the optical spectral weight as a function of temperature
2189: $T$ in units of the hopping parameter $t$. The solid curve is with
2190: and the dash-dotted curve is without vertex corrections.
2191: We see that while vertex corrections
2192: have changed the magnitude of the OS they have changed much less its
2193: temperature dependence which is somewhat more pronounced in the solid
2194: curve. In both cases the OS decreases with decreasing temperature as
2195: was the case in BCS theory. Also shown in Fig.~\ref{fig:24} are results
2196: for the Drude weight separately, dashed and dotted lines with and
2197: without vertex corrections. We see that it is also strongly enhanced by
2198: vertex corrections. In Fig.~\ref{fig:25} we reproduce
2199: results of Aristov and Zeyher\cite{ref119} for the conductivity
2200: $\sigma(\omega)$ vs frequency $\omega$ for $t=0.25\,$eV, $t'=0.076$,
2201: $T=0.001$, $n=0.4$ and a scattering rate of $1/\tau = 0.01$. It is
2202: seen that both the Drude region (intra-band) and the inter-band
2203: transitions which set in at higher $\omega\ge 0.2$ are larger when
2204: vertex corrections are included but that beyond $\omega\ge0.3$ the two
2205: curves merge and the vertex corrections are no longer important.
2206:
2207: In the continuum limit of the DDW model Valenzuela {\it et al.}%
2208: \cite{ref110} have obtained useful analytic results for intra
2209: (Drude) and inter (verticle transitions) separately. Vertex
2210: \begin{figure}[tp]
2211: \vspace*{-4mm}
2212: \includegraphics[width=9cm]{srFig27.eps}
2213: \caption{Optical conductivity $\sigma(\omega)$ with (solid line)
2214: and without (dashed line) vertex corrections.
2215: Adapted from Ref.~\protect{\onlinecite{ref119}}.
2216: }
2217: \label{fig:25}
2218: \end{figure}
2219: corrections are neglected and the limit of zero impurity scattering
2220: was taken. They find
2221: \begin{equation}
2222: \label{eq:46}
2223: \sigma_{intra}(T,\omega) = -\delta(\omega)\pi e^2
2224: \int\limits_{-\infty}^\infty\!d\nu\,\left[\frac{\partial f(t,\nu)}
2225: {\partial \nu}\right]g_1(\nu),
2226: \end{equation}
2227: with
2228: \begin{equation}
2229: \label{eq:47}
2230: g_1(\nu) = (\hbar v_F)^2 N(0)\frac{2}{\pi}\int\limits_0^{\pi/2}\!
2231: d\theta\,\Re{\rm e}\sqrt{1-\frac{D_0^2}{(\mu+\nu)^2}\cos^2\theta},
2232: \end{equation}
2233: which can be written in terms of elliptic integrals. Here, $v_F$
2234: is the Fermi velocity. Note
2235: also that the Fermi factor $f(T,\nu)$ has $\mu=0$ and the chemical
2236: potential has been transfered to $g_1(\nu)$. For the inter-band
2237: \begin{widetext}
2238: \begin{equation}
2239: \label{eq:48}
2240: \sigma_{inter}(T,\omega) = (\hbar v_F)^2N(0)\frac{1}{\pi}\left(\frac{2D_0}
2241: {\omega}\right)^2\int_0^{\pi/2}\!d\theta\,\Re{\rm e}
2242: \left[\frac{\cos^2\theta}{\sqrt{1-\left(\frac{2\Delta}{\omega}\right)^2
2243: \cos^2\theta}}\right]B(T,\omega),
2244: \end{equation}
2245: \end{widetext}
2246: where $B(T,\omega)$ is a universal thermal factor
2247: \begin{equation}
2248: \label{eq:49}
2249: B(T,\omega) = \frac{\pi e^2}{\omega}\frac{\sinh(\beta\omega/2)}
2250: {\cosh(\beta\mu)+\cosh(\beta\omega/2)},\qquad \beta=\frac{1}{k_B T},
2251: \end{equation}
2252: which at $T=0$ becomes proportional to a theta function
2253: $\theta(\omega-2\vert\mu\vert)$. This provides a low energy cutoff to the
2254: inter-band transitions. When impurity scattering is included the
2255: delta function $\delta(\omega)$ in Eq.~\eqref{eq:46} broadens into a
2256: Lorentzian and temperature leads to the overlap of the two contributions
2257: as seen in Fig.~\ref{fig:25}.
2258:
2259: \begin{figure}[tp]
2260: % \centering
2261: % \includegraphics[width=10cm,angle=270]{srFig26.eps}
2262: \includegraphics[width=6cm,angle=270]{srFig28.eps}
2263: \caption{(Color online)
2264: Optical spectral weight in the presence of a $t'$ term in the band
2265: dispersion.
2266: Here $W(D,T)$ (Eq.~\protect{\eqref{eq:43}}, dash-dotted line),
2267: $W^{DDW}(D,T)$ (Eq.~\protect{\eqref{eq:49a}} (solid line) are shown
2268: together with $W(T)$ (dashed line) which is for the non interacting
2269: case. The parameters are $t'=0.3\,t$, $T_{DDW}=0.12\,t$,
2270: $D(0) = 4T_{DDW}$, and the doping $\delta = 0.1$. The chemical potential is
2271: evaluated self consistently at each temperature.
2272: Observe that near $T_{DDW}$ a small decrease of $W^{DDW}$ with
2273: respect to $W(T)$ is seen, due to the change of chemical potential
2274: near $T_{DDW}$. Adapted from Ref.~\protect{\onlinecite{ref113}}.
2275: }
2276: \label{fig:26}
2277: \end{figure}
2278: We note that the kinetic energy shows the same trend in its temperature
2279: dependence as does the OS of Fig.~\ref{fig:24}. This is shown in
2280: Fig.~\ref{fig:26} which we took from the work of Benfatto
2281: {\it et al.}\cite{ref113} and where it is denoted by $W(D,T)$ as
2282: the dash-dotted curve for the parameters shown in the figure and described
2283: in the caption. We see the same decreasing trend with decreasing
2284: temperatures as for the OS. There are two other curves, the dashed
2285: one is for
2286: a pure tight-binding band with no DDW and is included for comparison.
2287: The solid curve is for $W^{DDW}(D,T)$ given by
2288: \begin{equation}
2289: \label{eq:49a}
2290: W^{DDW}(D,T) = -\frac{1}{N}\sum\limits_{\bf k}^{MBZ}E_{\bf k}
2291: \left[f(T,\xi_{+{\bf k}})-f(T,\xi_{-{\bf k}})\right]
2292: \end{equation}
2293: (again in units of $\pi e^2/\hbar^2$)
2294: in the notation of Benfatto
2295: {\it et al.}\cite{ref113,ref114} and requires explanation. It was obtained
2296: from a current operator which was modified to ensure charge
2297: conservation without the need for vertex corrections. This current
2298: operator has been used in other works\cite{ref108,ref120} as well.
2299: As noted by Aristov and Zeyher\cite{ref119}, however, such a
2300: procedure tends to overestimate the conductivity at higher
2301: frequencies and the OS now shows an increase as $T$ is decreased.
2302:
2303: We can also add a mean field BCS term
2304: \begin{equation}
2305: \label{eq:44}
2306: H' = \sum\limits_{\bf k}\left[\Delta^\ast_{\bf k}c_{-{\bf k}\downarrow}
2307: c_{{\bf k}\uparrow}+ h.c.\right],
2308: \end{equation}
2309: to the Hamiltonian \eqref{eq:40},
2310: where $\Delta_{\bf k}$ is the superconducting order parameter.
2311: $\Delta_{\bf k}=\Delta_0\left[\cos(k_xa)-\cos(k_ya)\right]$ with
2312: $\Delta_0$ the superconducting gap amplitude. In this case the
2313: OS in units of $\pi e^2/\hbar^2$ is given by
2314: \begin{eqnarray}
2315: W^{DDW}(D,\Delta,T) &=& \frac{2}{N}
2316: \sum\limits_{\bf k}^{\rm MBZ} E_{\bf k}\left[
2317: \frac{\xi_{+{\bf k}}}{E_{+{\bf k}}}\tanh\left(\frac{E_{+{\bf k}}}{2T}\right)
2318: \right.\nonumber\\
2319: &&\left.
2320: -\frac{\xi_{-{\bf k}}}{E_{-{\bf k}}}\tanh\left(\frac{E_{-{\bf k}}}{2T}\right)
2321: \right],
2322: \label{eq:45}
2323: \end{eqnarray}
2324: with $E_{\pm{\bf k}} = \sqrt{\xi^2_{\pm{\bf k}}+\Delta^2_{\bf k}}$.
2325:
2326: \begin{figure}[tp]
2327: % \centering
2328: % \includegraphics[width=14cm]{srFig27.eps}
2329: \includegraphics[width=8cm]{srFig29.eps}
2330: \caption{(Color online)
2331: Optical spectral weight in units of $e^2\pi/\hbar^2$ in the
2332: normal state
2333: (dash double-dotted line), in the DDW state
2334: Eq.~\protect{\eqref{eq:49a}} (solid line)
2335: and in DDW+SC state Eq.~\protect{\eqref{eq:45}} (dashed line).
2336: The values of parameters are
2337: for the doping $\delta = 0.13$ [$D_0(T=0)=0.92\,t$,
2338: $\Delta(T=0)=0.064\,t$, see Refs.~\protect{\onlinecite{ref107}} and
2339: \protect{\onlinecite{ref108}}
2340: for further details. The critical temperature is marked by the arrow.
2341: Observe that the decrease of $W^{DDW}(D,\Delta,T)$ below $T_c$ is
2342: small. Inset: spectral weight plotted as a function of
2343: $(T/t)^2$. Adapted from Ref.~\protect{\onlinecite{ref114}} (see also
2344: Ref.~\onlinecite{ref108}).
2345: }
2346: \label{fig:27}
2347: \end{figure}
2348: In Fig.~\ref{fig:27} we show results reproduced from Benfatto and
2349: Sharapov\cite{ref114} for the effect of combined DDW and superconducting
2350: transition described by Eq.~\eqref{eq:45}.
2351: The optical weight is given in units of $t$ (nearest neighbor hopping)
2352: as is the temperature. The dash-double-dotted curve $[W(T)]$ is for
2353: reference
2354: and gives results for the tight binding band without interactions while
2355: the solid lines gives $W^{DDW}(D,T)$ as before and the dashed
2356: curve is for $W^{DDW}(D,\Delta,T)$ which includes the effect of a $d$-wave
2357: superconducting gap of amplitude $\Delta$. We note that, as expected,
2358: this leads to an increase in KE with respect to the pure DDW case and,
2359: therefore, a drop in the OS.
2360:
2361: It is clear that, as yet, a complete theory of the OS in the DDW model
2362: does not
2363: exist. The calculations of Aristov and Zeyher\cite{ref119} with vertex
2364: corrections properly accounted for give the opposite temperature
2365: dependence than found experimentally. This theory, however, does not
2366: include correlations beyond those directly responsible for the DDW
2367: transition. It is clear from what we have described here that correlations
2368: leading to lifetime effects need to be included as these are very
2369: closely related to the observed temperature dependence of the OS
2370: and cannot be ignored.
2371:
2372: \section{Optical Spectral Weight Distribution}
2373: \label{sec:6}
2374:
2375: In this review we focused mainly on the OS for a single band
2376: \begin{figure}[tp]
2377: % \centering
2378: \includegraphics[width=8cm]{srFig30.eps}
2379: \caption{Selection of conductivity spectra for
2380: three BSCCO samples, underdoped, $T_c=70\,$K (B70KUND, top frame),
2381: optimally doped, $T_c=80\,$K (B80KOPT, middle frame), and overdoped,
2382: $T_c=63\,$K (B63KOVR, bottom frame). For the
2383: B70KUND sample, within the showed spectral range, the spectra at
2384: $50\,$K and $10\,$K are indistinguishable.
2385: Adapted from Ref.~\protect{\onlinecite{ref47}}.
2386: }
2387: \label{fig:28}
2388: \end{figure}
2389: integrated over all energies of relevance. As is seen from Eq.~\eqref{eq:1}
2390: this quantity can be computed from a knowledge of the single electron
2391: spectral density $A({\bf k},\omega)$ which determines the probability
2392: of occupation $n_{{\bf k},\sigma}$ of the state $\vert{\bf k},\sigma\rangle$.
2393: This is a much simpler problem than computing the frequency dependent
2394: conductivity from a Kubo formula which involves the two-particle Green's
2395: functions. However, calculating the full frequency dependent conductivity
2396: cannot be avoided if one wishes to discuss the optical spectral weight
2397: distribution. The partial integration of
2398: $\sigma_1(T,\omega)$ to a maximum $\omega$ equal to $\omega_M$ has
2399: proved be very useful and has provided valuable information
2400: about normal and superconducting state beyond what is obtained from the
2401: OS itself. We give here only two
2402: examples. In Fig.~\ref{fig:28} we reproduce the conductivity
2403: data of Santander-Syro
2404: {\it et al.}\cite{ref47} in three BSCCO samples, underdoped $T_c=70\,$K
2405: (B70KUND), optimally doped at $T_c=80\,$K (B80KOPT), and overdoped
2406: at $T_c=63\,$K (B63KOVR) at the various temperatures noted in the figure.
2407: From these data it is possible to calculate
2408: $W(T,\omega_M)=\int_{0^+}^{\omega_M}%
2409: d\omega\,\sigma_1(T,\omega)$ at various temperatures.
2410: Here the $0^+$ indicates that
2411: \begin{figure}[tp]
2412: % \centering
2413: \includegraphics[width=8cm]{srFig31.eps}
2414: \caption{Rato $\Delta W/W_S$ vs frequency showing the exhaustion of the
2415: Ferrell-Glover-Tinkham sum rule at conventional energies for the OVR (diamonds)
2416: and OPT (triangles) samples. An
2417: unconventional ($\sim 16 000\,$cm$^{-1}$ or $2\,$eV) energy scale is
2418: required for the UND sample (circles). Note that the frequency scale changes at
2419: 800 and $8 000\,$cm$^{-1}$. The changes in spectral weight are taken
2420: between $80\,$K - $10\,$K, $91\,$K - $10\,$K, and $100\,$K - $10\,$K
2421: for the OVR, OPT, and UND samples, respectively. Adapted from
2422: Ref.~\protect{\onlinecite{ref47}}.
2423: }
2424: \label{fig:29}
2425: \end{figure}
2426: the delta function representing the condensate has been left out.
2427: In Fig.~\ref{fig:29}
2428: we reproduce the experimental results of Santander-Syro {\it et
2429: al.}\cite{ref47} for the normalized change in spectral weight
2430: $\Delta W/W_S$ vs $\omega_M$ in $10^3\,$cm$^{-1}$ based on the data of
2431: Fig.~\ref{fig:28}. Open diamonds,
2432: triangles, and circles are for B63KOVR, B80KOPT, and B70KUND, respectively.
2433: Note the two breaks in the horizontal scale for $\omega_M$ at 800 and
2434: $8 000\,$cm$^{-1}$. The change in optical spectral weight is slightly
2435: different
2436: \begin{figure}[tp]
2437: % \centering
2438: % \includegraphics[width=12cm]{srFig30.eps}
2439: \includegraphics[width=9cm]{srFig32.eps}
2440: \caption{(Color online)
2441: The normalized weight of the condensate [$N_n(\omega)-
2442: N_s(\omega)]/\rho_{s,a}$ vs $\omega$ for optimally doped YBCO$_{6.95}$
2443: (solid line)
2444: and underdoped YBCO$_{6.60}$ (dotted line) along the $a$-axis direction.
2445: The condensate for the optimally
2446: doped material has saturated by $\simeq 800\,$cm$^{-1}$, while in the
2447: underdoped material the condensate is roughly 80\% formed by this
2448: frequency, but the other 20\% is not recovered until much higher
2449: frequencies. The error bars on the curve for the underdoped material
2450: indicate the uncertainty associated with the Ferrell-Glover-Tinkham
2451: sum rule. Inset:
2452: The low-frequency region. Adapted from Ref.~\protect{\onlinecite{ref121}}.
2453: }
2454: \label{fig:30}
2455: \end{figure}
2456: for the OVR, OPT, and UND samples as indicated in the caption. Note
2457: the approach to its asymptotic value of one. For the overdoped case
2458: the scale is $\sim 600\,$cm$^{-1}$ while for the UND sample it is
2459: much larger and of order $2\,$eV. These experiments clearly reveal a
2460: fundamental difference in behavior between overdoped and underdoped
2461: samples. This was also observed in the YBCO series. Data from Homes
2462: {\it et al.}\cite{ref121} are reproduced in Fig.~\ref{fig:30} for electric
2463: field {\bf E} parallel to the $a$-axis in YBCO$_{6.95}$ (solid line)
2464: and YBCO$_{6.60}$ (dotted line). For the optimally doped sample the OS
2465: is rapidly saturated ($\omega_M\simeq 800\,$cm$^{-1}$) while for the
2466: underdoped case a frequency of about $9 000\,$cm$^{-1}$ is needed.
2467:
2468: For a BCS $s$-wave superconductor the expectation is that the saturation
2469: of the OS should occur at a frequency $\omega$ a few times the gap $\Delta$
2470: even if
2471: the system is dirty with scattering rates $\stackrel{>}{\sim}2\Delta$,
2472: Refs.~\onlinecite{ref121} and \onlinecite{ref122}.
2473: However, superconductors are better
2474: described by Eliashberg theory which properly accounts for coupling
2475: of the electrons to phonons. In this case the weight in the coherent
2476: quasiparticle part of the spectral function is
2477: $Z = 1/(1+\lambda)$ where $\lambda$ is the mass enhancement factor. The rest
2478: of the spectral weight lies in an incoherent phonon induced band at
2479: higher energy, usually in the infrared. This part of the spectral
2480: function $A({\bf k},\omega)$ contributes the so called Holstein band to
2481: the optical conductivity. Only the quasiparticle part is included in
2482: BCS theory, yet for $\lambda > 1$, say, more of the electron spectral
2483: weight is in the incoherent part than one finds in the coherent part.
2484: This part introduces a new energy scale into the problem, namely, an
2485: average phonon energy and it is no longer true to say that readjustment
2486: of optical spectral weight on entering the superconducting state
2487: can only occur on the scale of twice the gap. In fact, one should expect
2488: that when the electrons pair, an absorption process involving a phonon
2489: $\omega_E$ would be changed and shifted to an energy of $2\Delta+\omega_E$,
2490: thus shifting the Holstein band to higher energies and, hence, the incoherent
2491: part of the optical spectral weight in the superconducting state
2492: extends to higher energies. This implies that
2493: \begin{figure}[tp]
2494: \vspace*{-4mm}
2495: \includegraphics[width=9cm]{srFig33.eps}
2496: \caption{Top frame: Optical spectral weight $W(\omega,T) = \int_{0^+}^\omega\!
2497: d\nu\,\sigma_1(\nu)$ for various cases as a function of $\omega$. The
2498: dotted (dash-double-dotted) curve is for the normal state at $T=10\,$K
2499: ($T=95\,$K), the solid curve for the superconducting state at $T=10\,$K.
2500: The dashed (dash-dotted) curve is the difference curve between superconducting
2501: at $T=10\,$K and normal states
2502: at $T=10\,$K ($T=95\,$K). The approach of the difference curve to its
2503: saturated large $\omega$ value depends significantly on the temperature
2504: used for the subtracted normal state. The thin dash-double-dotted
2505: horizontal line is the value of the condensate contribution
2506: (penetration depth). The bottom frame
2507: shows the real part of the conductivity for the normal state at
2508: $T=293\,$K (dashed curve), $T=95\,$K (dash-dotted curve), $T=10\,$K
2509: (dotted curve), and for the superconducting state at a $T=10\,$K
2510: (solid curve). All curves are for YBCO$_{6.95}$ with the impurity parameter
2511: given in Ref.~\protect{\onlinecite{ref122a}}.
2512: }
2513: \label{fig:31}
2514: \end{figure}
2515: $\int_{0^+}^\omega\!d\nu\,[\sigma_N(T,\nu)-\sigma_S(T,\nu)]$
2516: for large $\omega$ should saturate from above rather than from below as
2517: is the case in BCS. This is known from the $s$-wave phonon mediated
2518: case\cite{ref129}
2519: and is illustrated in Fig.~\ref{fig:31} taken from
2520: Carbotte and Schachinger\cite{ref122a} for a $d$-wave superconductor
2521: using for $I^2\chi(\omega)$ an MMP form. What is shown in the top frame are
2522: numerical results for $W(T,\omega)$ defined as $W(T,\omega) =
2523: \int_{0^+}^{\omega}\!d\nu\,\sigma_1(T,\nu)$. % where the $0^+$ indicates that
2524: %the delta function representing the condensate has been left out.
2525: The real part of the conductivity $\sigma_1(T,\omega)$ in arbitrary
2526: units on which these
2527: various curves are based are shown in the bottom frame. These calculations
2528: are all done for infinite bands and corresponding Eliashberg equations
2529: for a $d$-wave superconductor so that the Ferrell-Glover-Tinkham
2530: sum rule holds, i.e.: the total optical spectral weight is conserved
2531: between normal and superconducting state.\cite{ref123,ref124} Parameters
2532: were varied to get a good fit to data on YBCO$_{6.95}$. The reader is
2533: referred to the paper of Schachinger and Carbotte\cite{ref33} for
2534: details. Some impurity scattering in the unitary limit is included to
2535: get Fig.~\ref{fig:31}. $W(T,\omega)$ is shown for
2536: $\omega$ up to $250\,$meV. The solid curve is in the superconducting state at
2537: $T=10\,$K and the dotted the normal state at the same temperature.
2538: $W_N(T)$ (normal state) rises much more rapidly at small $\omega$
2539: than does $W_S(T)$ (supercond. state) and goes
2540: to much larger values. The difference $W_N(T=10\,{\rm K},\omega)-
2541: W_S(T=10\,{\rm K},\omega)$ (dashed curve) is the amount of optical
2542: spectral weight that has been transferred to the condensate between
2543: $(0^+,\omega)$. This curve rapidly grows to a value slightly below
2544: the horizontal line representing the condensate contribution to the
2545: total sum rule. After this the remaining variation is small with a shallow
2546: minimum around $30\,$meV followed by a broad peak around $60\,$meV which
2547: falls above the thin dash-double-dotted horizontal line before gradually
2548: falling again towards its asymptotic value which must be equal to the
2549: condensate contribution. All these features can be understood from a
2550: consideration of the curves for $\sigma_1(T,\omega)$ given in the
2551: bottom frame. Comparing dotted and solid curves, we see that they cross
2552: at three places on the frequency axis at $\omega_1\approx 8\,$meV,
2553: $\omega_2\approx 32\,$meV, and $\omega_3\approx 130\,$meV. These
2554: features are a result of the shift in incoherent background towards
2555: higher energies due to the opening of the superconducting gap.
2556:
2557: In an actual experiment it is usually not possible to access the
2558: normal state at $T=10\,$K (say) so that $W_N(T=95\,{\rm K},\omega)$
2559: just above $T_c = 92\,$K needs to be used (dash-double dotted curve).
2560: The difference
2561: $W_N(T=95\,{\rm K},\omega)-W_S(T=10\,{\rm K},\omega)$ is shown as the
2562: dash-dotted curve in the top frame and is seen to merge with the
2563: dashed curve only at higher values of $\omega$. Reference to the
2564: bottom frame shows that the real part of the normal state conductivity
2565: at $T=95\,$K (dash-dotted curve) is much broader than at
2566: $T=10\,$K (dotted curve) and this accounts for the slower rise towards
2567: saturation of the dash-dotted as compared to the dashed curve in the
2568: top frame. We note, however, that
2569: $W_N(T=95\,{\rm K},\omega)-W_S(T=10\,{\rm K},\omega)$ still approaches
2570: the penetration depth curve from above but much of the structure seen
2571: in the dashed curve is lost by using the data for the normal state at
2572: $T=95\,$K rather than at $T=10\,$K. Nevertheless, the energy scale
2573: over which the condensate is formed is set by the energy of the
2574: spin fluctuation spectra which extends up to $400\,$meV in our
2575: calculations and not by the gap value $2\Delta_0$.
2576:
2577: %We give one more example of the usefulness of considering the spectral
2578: %weight up to some finite, small $\omega_N$ rather than considering the
2579: %total spectral weight for the particular band of interest. As we have
2580: %said before, many experiments show that a pseudogap develops in
2581: %underdoped systems. One might expect that its opening should lead to
2582: %loose spectral weight at small frequencies as it opens.
2583:
2584: \section{Summary}
2585: \label{sec:7}
2586:
2587: The temperature dependence of the OS has recently been measured up to
2588: $\sim 300\,$K in the normal state of several cuprates as well as in
2589: the superconducting state. Some controversy remains about details such
2590: as the exact temperature dependence followed in the normal state.
2591: However, it seems established that underdoped samples show anomalous
2592: behavior when they enter the superconducting state. For a normal
2593: BCS superconductor the KE should increase while it is seen to decrease
2594: over its extrapolated normal state value. This, on its own, does not
2595: mean that the condensation is purely driven by the kinetic energy but it
2596: does mean that we are faced with non BCS behavior. For optimally
2597: doped samples the change in KE is close to zero with a definite
2598: crossover to conventional behavior in overdoped samples.
2599: For the normal state many experiments, but not all, give a $T^2$ dependence
2600: of the OS. Recent
2601: DMFT calculations for the Hubbard model provide evidence for a
2602: universal $T^2$ behavior in the normal state.
2603: We argue, however, that such a law is not robust when
2604: electron-boson theories for the interaction are considered. In this
2605: case, if the boson energy is low, a linear in $T$ dependence can result and
2606: by implication other dependences on $T$ can arise for different
2607: electron-boson spectral densities. The same holds for coupling to
2608: spin fluctuations in the NAFFL model. The issue of the temperature
2609: dependence of the underlying normal state OS or KE impacts the
2610: analysis of the change in KE that results from the superconducting
2611: condensation. If the normal state KE decreases faster
2612: at low temperatures than it does above $T_c$ (as we have seen in some
2613: models) this could be interpreted as an anomalous superconducting
2614: state, yet it would not be. On the other hand, in the
2615: NAFFL model as well as in electronic models more generally, anomalous
2616: behavior of the optical sum can be understood as due to the so called
2617: collapse of the inelastic scattering rates in the superconducting
2618: state if this reduces the effect of spin fluctuations at small
2619: $\omega$. % In this model
2620: %anomalous behavior of the OS can be understood on the basis of the
2621: %so called collapse of the inelastic scattering rates as
2622: %superconductivity sets in and reduces the effect of spin fluctuations
2623: %at small $\omega$. Coupling to an optical resonance is also observed
2624: %in some cases.
2625: Such a hardening of the spin fluctuation spectrum is
2626: thought to be a viable explanation for the peak in the
2627: microwave conductivity which is observed at a temperature
2628: considerably below $T_c$. This effect, which can be modeled by a
2629: low energy cutoff in the electron-spin fluctuation spectral density,
2630: results directly in a decrease in KE of the underlying normal state
2631: as additional coherence sets in with the establishment of
2632: superconductivity. This idea is related to the work of
2633: Norman and P\'epin\cite{ref78,ref79} who model directly the
2634: charge carrier self
2635: energy rather than go through the electron-boson spectral function.
2636: They take the imaginary part of the self
2637: energy $(\Sigma_2)$ to be large in the normal state (incoherence)
2638: and increase coherence in the superconducting state through
2639: a low frequency cutoff
2640: below which $\Sigma_2$ is zero. Another related model, which is even simpler,
2641: is to use a temperature dependent elastic scattering time which collapses
2642: in the superconducting state as $\alpha T^4$ with $\alpha$ taken from
2643: microwave experiments. All these mechanisms can easily provide savings in KE
2644: sufficiently large to explain the OS experiments.
2645: None, however, are fully quantitative at present. They also need to be
2646: extended to the overdoped side of the phase diagram. In this case one
2647: would expect the corresponding microwave peak to be greatly reduced
2648: and be more conventional. More sophisticated,
2649: but perhaps less transparent numerical calculations based on the
2650: Hubbard model and on the $t$-$J$ model have also found that, as a
2651: function of doping, the KE can favor pairing in underdoped systems
2652: while for the overdoped case it behaves as for a conventional
2653: superconductor. Earlier calculations based on the Hubbard model, however,
2654: give kinetic energy pairing even on the overdoped side of the phase
2655: diagram. By contrast, calculations based on the negative $U$ Hubbard model
2656: used to describe the BCS-BE crossover do capture the observed change
2657: in sign of the kinetic energy difference.
2658:
2659: In Ref.~\onlinecite{ref122} and, more recently, in Ref.~\onlinecite{ref53}
2660: the possibility is raised
2661: that some of the effects studied in this review may arise from infinite
2662: bands when a finite cutoff is applied to the optical sum integral as
2663: it must in the analysis of experimental data. Certainly, for the high
2664: $T_c$ oxides there is no clear `ending' of one band at a given energy
2665: followed by a gap before the next band sets in at higher energies.
2666: Instead, only a minimum is observed in the real part of the conductivity
2667: as a function of $\omega$ at about $1.2\,$eV. This energy has been taken
2668: as the cutoff on the transitions associated with the single band of
2669: interest. This cutoff is also roughly consistent with what is known
2670: about the width in energy of the bands in the oxides. Nevertheless,
2671: ambiguity remains and a theoretical study of overlapping bands and what
2672: this might imply for the single band sum rule would be valuable in
2673: clarifying this situation further.
2674:
2675: Much remains to be done to connect weak coupling
2676: approaches with the more numerical strong coupling results. There
2677: is also a need to get more accurate data which could resolve the
2678: experimental debate that still goes on and to achieve quantitative
2679: agreement with experiments. What is clear, however, is that such
2680: experiments can give useful information on the correlation effects
2681: that are involved and they certainly have shown that
2682: superconductivity on the underdoped side of the phase diagram of the
2683: cuprates is likely to be unconventional. It cannot be understood on the basis
2684: of a simple BCS model or an Eliashberg model with conventional phonons
2685: as these are not expected to change much as superconductivity sets in
2686: and they are distributed over a large energy range rather than peaked at low
2687: $\omega$. However, we stress again that as we have seen in one case,
2688: a small upturn in the
2689: OS at low temperature in the superconducting state can arise in spin
2690: fluctuation models without a readjustment of the electron-spin fluctuation
2691: spectrum, when the spin fluctuation energy in the MMP spin
2692: susceptibility is sufficiently small. While this upturn is less in the
2693: superconducting state than in its normal state at this same temperature
2694: (below $T_c$) it is above the value extrapolated to $T=0$ from the
2695: normal state above $T_c$. A second important conclusion of this review
2696: is that lifetime effects are
2697: important in determining the temperature dependence of the OS in
2698: both normal and superconducting state. These have to be accounted for in
2699: any definitive theory of this effect.
2700:
2701: There is a need to continue to explore other models such as
2702: phase fluctuations. In particular, the $D$-density wave model as a
2703: possible competing order for the pseudogap phase has not yet included
2704: effects of correlations beyond those that lead to the DDW order,
2705: i.e: lifetime effects which we have argued to be central to this
2706: problem. More work involving other models of interaction effects
2707: so as to understand what other temperature laws are possible for the
2708: OS vs $T$ would also be useful.\\[3ex]
2709:
2710: \begin{acknowledgments}
2711: Research supported in part by the Natural Sciences and Engineering
2712: Research Council (NSERC) of Canada and the Canadian Institute of Advanced
2713: Research (CIAR). We thank, E. Arrigoni, L. Benfatto, N. Bontemps,
2714: R.P.S.M. Lobo, D. van der Marel, F. Marsiglio,
2715: E. Nicol, S. Sharapov, A. Toschi, and R. Zeyher for discussion
2716: and/or correspondence.
2717: \end{acknowledgments}
2718:
2719: \begin{thebibliography}{999}
2720:
2721: \bibitem{ref1} C. Campuzano, M.R. Norman, and M. Randeria, in: The Physics
2722: of Superconductivity; Conventional and High-T$_c$ Superconductors,
2723: edited by K.-H. Bennemann and J. B. Ketterson (Springer, Berlin, 2003),
2724: Vol. II, p. 167.
2725: \bibitem{ref2}A. Damascelli, Z. Hussain, and Z.X. Shen, Rev. Mod. Phys.
2726: {\bf 75}, 473 (2003).
2727: \bibitem{ref3}H. Matsui, K. Terashima, T. Sato, T. Takahashi,
2728: M. Fujita, and K. Yamada, Phys. Rev. Lett. {\bf 95}, 017003 (2005).
2729: \bibitem{ref4}D.A. Bonn and W.N. Hardy, in: Physical Properties of
2730: High Temperature Superconductors, edited by D.M. Ginsberg (World
2731: Scientific, Singapore, 1996), Vol. 5, p. 7.
2732: \bibitem{ref5} C.C. Tsuei and J.R. Kirtley, in: The Physics
2733: of Superconductivity; Conventional and High-T$_c$ Superconductors,
2734: edited by K.-H. Bennemann and J. B. Ketterson (Springer, Berlin, 2003),
2735: Vol. I, p. 648.
2736: \bibitem{ref6}M.C. Nuss, P.M. Mankiewich, M.L. O'Malley, E.H.
2737: Westerwick, and P.B. Littlewood, Phys. Rev. Lett. {\bf 66}, 3305
2738: (1991).
2739: \bibitem{ref7}D.A. Bonn, P. Dosanjh, R. Liang, and W.N. Hardy,
2740: Phys. Rev. Lett. {\bf 68}, 2390 (1992).
2741: \bibitem{ref8}D.B. Romero, C.D. Porter, D.B. Tanner, L. Forro,
2742: D. Mandrus, L. Mihaly, G.L. Carr, and G.P. Williams, Phys. Rev. Lett.
2743: {\bf 68}, 1590 (1992).
2744: \bibitem{ref9}E.J. Nicol and J.P. Carbotte, Phys. Rev. B {\bf 44},
2745: 7741 (1991).
2746: \bibitem{ref10}Shih-Fu Lee, D.C. Morgan, R.J. Ormeno, D.M. Broun,
2747: R.A. Doyle, J.R. Waldram, and K. Kadowaki, Phys. Rev. Lett. {\bf 77},
2748: 735 (1996).
2749: \bibitem{ref11}S.M. Broun, D.C. Morgan, R.J. Ormeno, S.F. Lee,
2750: A.W. Tyler, A.P. Mackenzie, and J.R. Waldram, Phys. Rev. B {\bf 56},
2751: R11443 (1997).
2752: \bibitem{ref12}J.E. Hirsch, Phys. Lett. A {\bf 134}, 451 (1989).
2753: \bibitem{ref13}J.E. Hirsch, Physica C {\bf 199}, 305 (1992);
2754: Physica C {\bf 201}, 347 (1992).
2755: \bibitem{ref14}J.E. Hirsch and F. Marsiglio, Phys. Rev. B {\bf 39},
2756: 11515 (1989).
2757: \bibitem{ref15}J.E. Hirsch and F. Marsiglio Physics C {\bf 331}, 150
2758: (2000); Phys. Rev. B {\bf 62}, 15131 (2000).
2759: \bibitem{ref16}P.W. Anderson, Science {\bf 279}, 1196 (1998);
2760: in: The Theory of Superconductivity in the
2761: High $T_c$ Cuprates (Princeton University Press, Princeton, 1997).
2762: \bibitem{ref17}S. Chakravarty, Eur. Phys. J. B {\bf 5}, 337 (1998).
2763: \bibitem{ref19}R. Kubo, J. Phys. Soc. Japan {\bf 12}, 570 (1957).
2764: \bibitem{ref18}D.N. Basov, S.I. Woods, A.S. Katz, E.J. Singley,
2765: R.C. Dynes, M. Xu, D.G. Hinks, C.C. Homes, and M. Strongin, Science
2766: {\bf 283} 49 (1999).
2767: \bibitem{ref20}L.B. Ioffe and A.J. Millis, Science {\bf 285}, 1241 (1999);
2768: Phys. Rev. B {\bf 61}, 9077 (2000).
2769: \bibitem{ref21}Wonkee Kim and J.P. Carbotte, Phys. Rev B {\bf 61}, 11886R
2770: (2000); {\bf 62}, 8661 (2000); {\bf 63}, 140505R (2001);
2771: {\bf 63}, 054526 (2001); {\bf 64}, 104501 (2001).
2772: \bibitem{ref22}N. Kumar and A.M. Jayannavar, Phys. Rev. B {\bf 45},
2773: 5001 (1992).
2774: \bibitem{ref23}P.W. Anderson and Z. Zou, Phys. Rev. Lett. {\bf 60},
2775: 132 (1988); {\bf 60}, 2557 (1988).
2776: \bibitem{ref24}A.J. Leggett, Braz. J. Phys. {\bf 22}, 129 (1992);
2777: M. Turlabov and A.J. Leggett, Phys. Rev. B {\bf 62}, 064518 (2001).
2778: \bibitem{ref25}Z. Tesanovic, Phys. Rev. B {\bf 36}, 2364 (1987);
2779: {\bf 38}, 2489 (1988).
2780: \bibitem{ref26}J. Halbritter, J. Supercond. {\bf 11}, 231 (1998).
2781: \bibitem{ref27}W.A. Atkinson and J.P. Carbotte, Phys. Rev. B {\bf 52},
2782: 10 601 (1995); {\bf 55}, 3230 (1997); {\bf 55}, 12 748 (1997).
2783: \bibitem{ref28}W.C. Wu, W.A. Atkinson, and J.P. Carbotte, J.
2784: Supercond. {\bf 11}, 305 (1997).
2785: \bibitem{ref29}T. Xiang and J.M. Wheatley, Phys. Rev. Lett. {\bf 77},
2786: 4632 (1996).
2787: \bibitem{ref30}P.J. Hirschfeld, S.M. Quinlan, and D.J. Scalapino,
2788: Phys. Rev. B {\bf 55}, 12 742 (1997).
2789: \bibitem{ref31}E.H. Kim, Phys. Rev. B {\bf 58}, 2452 (1997).
2790: \bibitem{ref32}M.J. Graf, M. Palumbo, D. Rainer, and J.A. Sauls,
2791: Phys. Rev. B {\bf 52}, 10 588 (1995).
2792: \bibitem{ref33}E. Schachinger and J.P. Carbotte, Phys. Rev. B {\bf 64},
2793: 094501 (2001).
2794: \bibitem{ref34} D.~van der Marel, H.J.A.~Molegraaf, C.~Presura, and
2795: L.~Santoso, in {\it Concepts in Electron Correlations}, edited by
2796: A.~Hewson and V.~Zlatic (Kluwer, 2003) p. 7.
2797: \bibitem{ref87a}G. Deutscher, A.F. Santander-Syro, and N. Bontemps,
2798: Phys. Rev. B {\bf 72}, 092504 (2005).
2799: \bibitem{ref33a} F. Carbone, A.B. Kuzmenko, H.J.A. Molegraaf,
2800: E. van Heumen, V. Lukovac, F. Marsiglio, D. van der Marel,
2801: K. Haule, G. Kotliar, H. Berger, S. Courjault, P.H. Kes,
2802: M. Li, cond-mat/0605209 (unpublished).
2803: \bibitem{ref123}R.A. Ferrell and R.E. Glover, Phys. Rev. {\bf 109}, 1398
2804: (1958).
2805: \bibitem{ref124}M. Tinkham and R.B. Ferrell, Phys. Rev. Lett. {\bf 2},
2806: 331 (1959).
2807: \bibitem{ref124a}F. Marsiglio, F. Carbone, A. Kuzmenk, and D.
2808: van der Marel, cond-mat/0606688 (unpublished).
2809: \bibitem{ref49}A. Knigavko, J.P. Carbotte, and F. Marsiglio, Phys. Rev.
2810: B. {\bf 70}, 224501 (2004).
2811: \bibitem{ref50}F. Marsiglio and J.E. Hirsch, Phys. Rev. B {\bf 41}, 6435
2812: (1990); Physica C {\bf 165}, 71 (1990).
2813: \bibitem{ref126} M.A. Quijada, D.B. Tanner, R.J. Kelley, M. Onellion,
2814: H. Berger, and G. Margaritondo, Phys. Rev. B \textbf{60}, 14917 (1999).
2815: \bibitem{ref52}L. Benfatto, F. Marsiglio, and J.P. Carbotte,
2816: cond-mat/0603661 (unpublished).
2817: \bibitem{ref53}A.E. Karakozov and E.G. Maksimov, cond-mat/0511185
2818: (unpublished).
2819: \bibitem{ref35}P.~Monthoux and D.~Pines, Phys. Rev. B {\bf 47}, 6069 (1993);
2820: {\bf 49}, 4261 (1994).
2821: \bibitem{ref36} P.~Monthoux, A.V.~Balatsky, and D.~Pines, Phys. Rev.
2822: Lett. {\bf 67}, 3448 (1991); Phys. Rev. B {\bf 46}, 14 803 (1992).
2823: \bibitem{ref37}A.J.~Millis, H.~Monien, and D.~Pines,
2824: Phys. Rev. B {\bf 42}, 167 (1990).
2825: \bibitem{ref38}D.~Branch and J.P.~Carbotte, Can.\ J.\ Phys. {\bf 77},
2826: 531 (1999); J.\ Superconductivity {\bf 12}, 667 (1999);
2827: {\bf 13}, 535 (2000).
2828: \bibitem{ref39}A.~Abanov, A.V.Chubukov, and J.~Schmalian,
2829: Adv. in Physics {\bf 52}, 119 (2003).
2830: \bibitem{ref40}E. Schachinger and J.P. Carbotte, Phys. Rev. B {\bf 72},
2831: 014535 (2005).
2832: \bibitem{ref40a} V. Chubukov, D. Pines, and J. Schmalian, in: The Physics
2833: of Superconductivity; Conventional and High-T$_c$ Superconductors,
2834: edited by K.-H. Bennemann and J. B. Ketterson (Springer, Berlin, 2003),
2835: Vol. I, p. 495.
2836: \bibitem{ref42}A. Toschi, M. Capone, M. Ortolani, P. Calvani, S. Lupi,
2837: C. Castellani, Phys. Rev. Lett. {\bf 95}, 097002 (2005).
2838: \bibitem{ref41}A. Georges, G. Kotliar, M.J. Rozenberg, and W. Krauth,
2839: Rev. Mod. Phys. {\bf 68}, 13 (1996).
2840: \bibitem{ref43}M. Ortolani, P. Calvani, and S. Lupi, Phys. Rev. Lett.
2841: {\bf 94}, 067002 (2005).
2842: \bibitem{ref44}A. Lucarelli, S. Lupi, M. Ortolani, P. Calvani, P. Maselli,
2843: M. Capizzi, P. Giura, H. Eisaki, N. Kikugawa, T. Fujita, M. Fujita, and
2844: K. Yamada, Phys. Rev. Lett. {\bf 90}, 037002 (2003).
2845: \bibitem{ref45}A.F. Santander-Syro, R.P.S.M. Lobo, N. Bontemps, Z.
2846: Konstantinovic, and H. Raffy, Phys. Rev. Lett. {\bf 88}, 097005 (2002).
2847: \bibitem{ref46}A.F. Santander-Syro, R.P.S.M. Lobo, N. Bontemps,
2848: Z. Konstantinovic, Z.Z. Li, and H. Raffy,
2849: Europhys. Lett. {\bf 62}, 568 (2003).
2850: \bibitem{ref47}A.F. Santander-Syro, R.P.S.M. Lobo, N. Bontemps, W. Lopera,
2851: D. Girat\'{a}, Z. Konstantinovic, Z.Z. Li, and H. Raffy, Phys. Rev.
2852: {\bf 70}, 134504 (2004).
2853: \bibitem{ref48}H.J.A. Molegraaf, C. Presura, D. van der Marel, P.H. Kes,
2854: and M. Li, Science {\bf 295}, 22 (2002).
2855: \bibitem{ref89c}Th.A. Maier, M. Jarrell, A. Macridin, and
2856: C. Slezak, Phys. Rev. Lett. {\bf 92}, 027005 (2004)
2857: \bibitem{ref89b}K. Haule and G. Kotliar, cond-mat/0601478 (unpublished).
2858: \bibitem{ref54}G. Grimaldi, E. Cappelutti, and L. Pietrobero,
2859: Europhys. Lett. {\bf 42}, 667 (1998).
2860: \bibitem{ref55}E. Cappelluti, C. Grimaldi, and L. Pietronero,
2861: Phys. Rev. B {\bf 64}, 125104 (2001).
2862: \bibitem{ref56}E. Capeppelluti and L. Pietronero, Phys. Rev. B
2863: {\bf 68}, 224511 (2003).
2864: \bibitem{ref57}F. Do\v{g}an and F. Marsiglio, Phys. Rev. B {\bf 68},
2865: 165102 (2003).
2866: \bibitem{ref58}A. Knigavko and J.P. Carbotte, Phys. Rev. B {\bf 72},
2867: 035125 (2005).
2868: \bibitem{ref59}A. Knigavko and J.P. Carbotte, Phys. Rev. B {\bf 73},
2869: 125114 (2006).
2870: \bibitem{ref62}A. Knigavko, J.P. Carbotte, and F. Marsiglio,
2871: Europhys. Lett. {\bf 71}, 776 (2005).
2872: \bibitem{ref60}F. Marsiglio, M. Schossmann, and J.P. Carbotte, Phys.
2873: Rev. B {\bf 32}, 4965 (1988).
2874: \bibitem{ref63}J.P. Carbotte, E. Schachinger, and J. Hwang, Phys.
2875: Rev. B {\bf 71}, 054506 (2005).
2876: \bibitem{ref61}E. Schachinger and J.P. Carbotte, Phys. Rev. B {\bf 62},
2877: 9054 (2000).
2878: \bibitem{ref64}J. Hwang, J. Yang, T. Timusk, S.G. Sharapov, J.P. Carbotte,
2879: D.A. Bonn, Ruixing Liang, and W.N. Hardy, Phys. Rev. B {\bf 73},
2880: 014508 (2006).
2881: %\bibitem{ref51} D. van der Marel, {\bf \ldots ??????????????}
2882: %\bibitem{ref60a}E. Schachinger, J.J. Tu, and J.P. Carbotte, Phys.
2883: %Rev. B {\bf 67}, 214508 (2003).
2884: \bibitem{ref65}A.V. Boris, N.N. Kovaleva, D.V. Dolgov, T. Holden,
2885: C.T. Lin, B. Keimer, and C. Bernhard, Science {\bf 304}, 708 (2004).
2886: \bibitem{ref66}A.B. Kuzmenko, H.J.A. Molegraaf, F. Carbone, and
2887: D. van der Marel, Phys. Rev. B {\bf 72}, 144503 (2005).
2888: \bibitem{ref67}A.F. Santander-Syro and N. Bontemps, Science {\bf 304},
2889: 708 (2004).
2890: \bibitem{ref68}J.P. Carbotte, E. Schachinger, and D.N. Basov,
2891: Nature (London) {\bf 401}, 354 (1999).
2892: \bibitem{ref69}E. Schachinger, J.P. Carbotte, and D.N. Basov,
2893: Europhys. Lett. {\bf 54}, 380 (2001).
2894: \bibitem{ref70} E.~Schachinger and J.P.~Carbotte,
2895: in: {\it Models and Methods of High-TC Superconductivity: some Frontal
2896: Aspects}, edited by J.K.~Srivastava and S.M.~Rao,
2897: (Nova Science, Hauppauge, NY, 2003), Vol. II, p. 73.
2898: \bibitem{ref71}P. Bourges, in: The Gap Symmetry and Fluctuations in
2899: High Temperature Superconductors, edited by J. Bok, G. Deutscher,
2900: D. Pavuna, and S.A. Wolf (Plenum Press, London, 1998), p. 349.
2901: \bibitem{ref72}E. Schachinger and J.P. Carbotte, Phys. Rev. B {\bf 65},
2902: 064514 (2002).
2903: \bibitem{ref73}A. Hosseini, R. Harris, Saeid Kamal, P. Dosanjh,
2904: J. Preston, Ruixing Liang, W.N. Hardy, and D.A. Bonn, Phys. Rev. B {\bf 60},
2905: 1349 (1999).
2906: \bibitem{ref75}D.A. Bonn, Ruixing Liang, T.M. Risemann, D.J. Baar,
2907: D.C. Morgan, K. Zhang, P. Dosanjh, T.L. Duty, A. MacFarlane,
2908: G.D. Morris, J.H. Brewer, W.H. Hardy, C. Kallin, and A.J.
2909: Berlinsky, Phys. Rev. B {\bf 47}, 11 314 (1993).
2910: \bibitem{ref74}E. Schachinger and J.P. Carbotte, Phys. Rev. B {\bf 57},
2911: 13 773 (1998); {\bf 57}, 7970 (1998).
2912: \bibitem{ref76}E. Schachinger, J.P. Carbotte, and F. Marsiglio,
2913: Phys. Rev. B {\bf 56}, 2738 (1997).
2914: \bibitem{ref77}R.S.~Markiewicz, S.~Sahrakorpi, M.~Lindroos, Hsin Lin,
2915: and A.~Bansil, cond-mat/0503064 (unpublished) and references therein.
2916: \bibitem{ref90}F. Marsiglio, Phys. Rev. B {\bf 73} 064507 (2006);
2917: Erratum to be published.
2918: \bibitem{ref78}M.R. Norman and C. P\'epin, Phys. Rev. B {\bf 66},
2919: 100506(R) (2002).
2920: \bibitem{ref79}M.R. Norman and C. P\'epin, Rep. Progr. Phys. {\bf 66},
2921: 1547 (2003).
2922: \bibitem{ref80}M.R. Norman and H. Ding, Phys. Rev. B {\bf 57},
2923: R11089 (1998).
2924: \bibitem{ref81}R. Hlubina and T.M. Rice, Phys. Rev. B {\bf 51},
2925: 9253 (1995).
2926: \bibitem{ref82}T. Valla, A.V. Fedorov, P.D. Johnson, Q. Li, G.D. Gu,
2927: and N. Koshizuka, Phys. Rev. Lett. {\bf 85}, 828 (2000).
2928: \bibitem{ref83}K.G. Sandeman and A.J. Schofield, Phys. Rev. B {\bf 63},
2929: 094510 (2001).
2930: \bibitem{ref84}B.P. Stojkovic and D. Pines, Phys. Rev. Lett. {\bf 76},
2931: 811 (1996).
2932: \bibitem{ref85}C.M. Varma, P.B. Littlewood, S. Schmitt-Rink,
2933: A. Abrahams, and A.E. Ruckenstein, Phys. Rev. Lett. {\bf 63},
2934: 1996 (1989).
2935: \bibitem{ref86}C.M. Varma, Phys. Rev. B {\bf 55}, 14 554 (1997).
2936: \bibitem{ref87}A.V. Puchkov, D.N. Basov, and T. Timusk, J. Phys.:
2937: Condens. Matter {\bf 8}, 10 049 (1996).
2938: \bibitem{ref88}This was pointed out in footnote 15 of
2939: Ref.~\onlinecite{ref87a}.
2940: \bibitem{ref89a}N. Bontemps, R.P.S.M. Lobo, A.F. Santander-Syro,
2941: A. Zimmers, cond-mat/0603024 (unpublished).
2942: \bibitem{ref89}J.L. Tallon and J.W. Loram, Physica C {\bf 349}, 53
2943: (2001).
2944: \bibitem{ref127}A. Toschi, M. Capone, and C. Castellani, Phys. Rev.
2945: B \textbf{72}, 235118 (2005).
2946: \bibitem{ref128}B. Kyung, A. Georges, and A.-M. S. Tremblay,
2947: cond-mat/0508645 (unpublished).
2948: \bibitem{ref91}V.J. Emery and S.A. Kivelson, Nature (London) {\bf 374},
2949: 434 (1995).
2950: \bibitem{ref92}M. Randeria, N. Trivedi, A. Moreo, and R.T. Scalettar,
2951: Phys. Rev. Lett. {\bf 69}, 2001 (1992).
2952: \bibitem{ref93}M. Franz and A.J. Millis, Phys. Rev. B {\bf 58}, 14 572
2953: (1998).
2954: \bibitem{ref94}N.-J. Kwon and A.T. Dorsey, Phys. Rev. B {\bf 59},
2955: 6438 (1999).
2956: \bibitem{ref95}I.F. Herbut, Phys. Rev. Lett. {\bf 88}, 047006 (2002).
2957: \bibitem{ref96}T. Eckl, D.J. Scalapino, E. Arrigoni, and W. Hanke,
2958: Phys. Rev. B {\bf 66}, 140510(R) (2002).
2959: \bibitem{ref97}T. Eckl, W. Hanke, and E. Arrigoni, Phys. Rev. B {\bf 68},
2960: 014505 (2003).
2961: \bibitem{ref97a}T.K. Kope\'c, Phys. Rev. B {\bf 67}, 014520 (2003).
2962: \bibitem{ref98}S. Chakravarty, R.B. Laughlin, D.K. Morr, and C. Nayak,
2963: Phys. Rev. B {\bf 63}, 094503 (2001).
2964: \bibitem{ref99}B. D\`ora, A. Virosztek, and K. Maki, Phys. Rev. B {\bf 65},
2965: 155119 (2002); K. Maki, B. D\`ora, M. Kartsovnik, A. Virosztek, B.
2966: Korin-Hamzic, and M. Basletic, Phys. Rev. Lett. {\bf 90}, 256402 (2003).
2967: \bibitem{ref100}S. Chakravarty, H.-Y. Kee, and C. Nayak, Int. J. Mod. Phys.
2968: B {\bf 15}, 2901 (2001).
2969: \bibitem{ref101}J.-X. Zhu, W. Kim, C.S. Ting, and J.P. Carbotte,
2970: Phys. Rev. Lett. {\bf 87}, 197001 (2001).
2971: \bibitem{ref102}X. Yang and C. Nayak, Phys. Rev. B {\bf 65}, 064523 (2002).
2972: \bibitem{ref103}Q.H. Wang, J.H. Han, and D.H. Lee, Phys. Rev. Lett.
2973: {\bf 87}, 077004 (2001).
2974: \bibitem{ref103a}S. Chakravarty, C. Nayak, and S. Tewari, Phys. Rev. B
2975: {\bf 68}, 100504 (2003).
2976: \bibitem{ref104}S. Chakravarty, C. Nayak, S. Tewari, and X. Yang,
2977: Phys. Rev. Lett. {\bf 89}, 277003 (2002).
2978: \bibitem{ref105}S. Tewari, H.-Y. Kee, C. Nayak, S. Chakravarty, Phys.
2979: Rev. B {\bf 64}, 224516 (2001).
2980: \bibitem{ref106}E. Capelutti and R. Zehyer, Phys. Rev. B {\bf 59},
2981: 6475 (1999).
2982: \bibitem{ref107}L. Benfatto, S. Caprara, and C. DiCastro, Eur. Phys. J. B
2983: {\bf 17}, 95 (2000).
2984: \bibitem{ref108}L. Benfatto, S. Sharapov, and H. Beck, Eur. Phys. J. B
2985: {\bf 34}, 469 (2004).
2986: \bibitem{ref109}D.N. Aristov and R. Zeyher, Phys. Rev. B {\bf 70},
2987: 212511 (2004).
2988: \bibitem{ref110}B. Valenzuela, E.J. Nicol, and J.P. Carbotte, Phys.
2989: Rev. B {\bf 71}, 134503 (2005).
2990: \bibitem{ref111}Wonkee Kim and J.P. Carbotte, Phys. Rev. B {\bf 66},
2991: 033104 (2002).
2992: \bibitem{ref112}Wonkee Kim, J.X. Zhu, J.P. Carbotte, and C.S. Ting,
2993: Phys. Rev. B {\bf 65}, 064502 (2002).
2994: \bibitem{ref113}L. Benfatto, S. Sharapov, N. Andrenacci, and H. Beck,
2995: Phys. Rev. B {\bf 71}, 104511 (2005).
2996: \bibitem{ref114}L. Benfatto and S. Sharapov, Low Temp. Phys.
2997: \textbf{32}, 533 (2006).
2998: \bibitem{ref115}B.I. Halperin and T.M. Rice, Solid State Phys. {\bf 21},
2999: 115 (1968).
3000: \bibitem{ref116}I. Affleck and J.B. Marston, Phys. Rev. B {\bf 37},
3001: 3774 (1988).
3002: \bibitem{ref119}D.N. Aristov and R. Zeyher, Phys. Rev. B {\bf 72},
3003: 115118 (2005).
3004: \bibitem{ref120}R. Gerami and C. Nayak, Phys. Rev. B {\bf 73}, 024505
3005: (2006).
3006: \bibitem{ref121}C.C. Homes, S.V. Dordevic, D.A. Bonn, Ruixing Liang,
3007: W.N. Hardy, Phys. Rev. B {\bf 69}, 024514 (2004).
3008: \bibitem{ref122}A.E. Karakozov, F.G. Maksimov, and O.V. Dolgov,
3009: Solid State Commun. {\bf 124}, 119 (2002).
3010: \bibitem{ref129}W. Shaw and J.C. Swihart, \prl {\bf 20}, 1000 (1968).
3011: \bibitem{ref122a}J.P. Carbotte and E. Schachinger, Phys. Rev. B {\bf 69},
3012: 224501 (2004).
3013: %\bibitem{ref125}E. Schachinger and J.P. Carbotte, Phys. Rev. B {\bf 64},
3014: %094501 (2001).
3015: \end{thebibliography}
3016: %\newpage
3017: %\section*{Tables}
3018: %\begin{table}[h]
3019: %\caption{\label{tab:1}The two tight binding models used within this
3020: %paper. Model A corresponds to the tight binding model discussed by
3021: %van der Marel {\it et al.}\protect{\cite{ref34}} $t$ and $t'$ are
3022: %given in meV, the critical temperature $T_c$ in K, and the filling
3023: %$\langle n\rangle$ is defined in Eq.~\protect{\eqref{eq:8}}.
3024: %}
3025: %\begin{ruledtabular}
3026: %\begin{tabular}{ldddd}
3027: %Model & t & t' & \langle n\rangle & T_c\\
3028: %\hline
3029: %A & 148.8 & 40.9 & 0.425 & 90\\
3030: %B & 100.0 & 16.0 & 0.4 & 100\\
3031: %\end{tabular}
3032: %\end{ruledtabular}
3033: %\vspace*{10cm}
3034: %\end{table}
3035: %\mbox{ }
3036: %\clearpage
3037: %\begin{table}
3038: %\caption{\label{tab:2}Results of the fits in Fig.~\protect{\ref{fig:4}}
3039: %for selected doping values. First and second columns: values of the
3040: %coefficient $B$ (in units of $1/D$) for low and high frequency cutoff,
3041: %respectively (note that, for $U=0$, $B\simeq 2/D$ in the whole doping
3042: %range considered here). Third and forth column: corresponding spectral
3043: %weights. Last two columns: kinetic energy and quasiparticle weight.
3044: %Adapted from Ref.~\protect{\onlinecite{ref42}}.
3045: %}
3046: %\begin{ruledtabular}
3047: %\begin{tabular}{ldddddd}
3048: %\begin{center}
3049: %\begin{tabular}{lcccccc}
3050: % & B(0.1D) & B(1.5D) & W_0(0.1D) & W_0(1.5D) & -E_{kin} & Z\\
3051: %\hline
3052: %x=0.07 & 58\pm5 & 27\pm3 & 0.05D & 0.11D & 0.17D & 0.12\\
3053: %x=0.12 & 48\pm5 & 16\pm4 & 0.08D & 0.16D & 0.21D & 0.20\\
3054: %x=0.15 & 43\pm5 & 15\pm4 & 0.09D & 0.18D & 0.22D & 0.22\\
3055: %x=0.19 & 51\pm5 & 13\pm5 & 0.11D & 0.20D & 0.24D & 0.27\\
3056: %x=0.26 & 33\pm7 & 11\pm3 & 0.14D & 0.24D & 0.27D & 0.35\\
3057: %\\
3058: %\end{tabular}
3059: %\end{center}
3060: %\end{ruledtabular}
3061: %\end{table}
3062: %\mbox{ }
3063: %\clearpage
3064: %\section*{Figures}
3065: %\begin{figure}[h]
3066: %\centering
3067: % \vspace*{-3cm}
3068: % \includegraphics[width=13cm]{vmFig3.eps}
3069: % \includegraphics[width=13cm]{smFig1.eps}
3070: %\vspace*{-5mm}
3071: % \caption{Optical sum $W$ and kinetic energy, $-W_{\rm KE}/2$,
3072: %as a function of $T^2$. Solid circles and squares are for the non
3073: %interacting case while solid up-triangles and solid down-triangles
3074: %include interactions.
3075: %The left hand frame applies to
3076: %Model A of Table~\protect{\ref{tab:1}} and $\omega_{SF}=82\,$meV was
3077: %used. Here, the interacting and non interacting cases show similar
3078: %temperature dependencies. The right hand frame is for
3079: %Model B of Table~\protect{\ref{tab:1}} with an MMP model
3080: %$\omega_{SF}=10\,$meV. Note the difference in temperature dependence
3081: %between interacting and non interacting case.
3082: %}
3083: % \label{fig:1}
3084: %\end{figure}
3085: %\mbox{ }
3086: %\clearpage
3087: %\begin{figure}[h]
3088: % \centering
3089: % \includegraphics[width=10cm]{vmFig2.eps}
3090: % \caption{The occupation number $n_{\bf k}$ for
3091: %selected directions in the CuO$_2$ Brillouin zone. Model A of
3092: %Table~\protect{\ref{tab:1}} was used.
3093: %Top frame: the non interacting case. Center frame: The interacting
3094: %case at a temperature $T=20\,$K. We show normal state (solid line)
3095: %and superconducting state (dashed line) results. Bottom frame: The
3096: %temperature influence on the normal state $n_{\bf k}$ for
3097: %$T=20\,$K (solid line) and $T=150\,$K (dashed line). }
3098: % \label{fig:2}
3099: %\end{figure}
3100: %\clearpage
3101: %\begin{figure}[tp]
3102: %\vspace*{-1cm}
3103: %\centering
3104: % \includegraphics[width=13cm]{vmFig5a.eps}
3105: % \includegraphics[width=13cm]{smFig3.eps}
3106: % \caption{Comparison of normal and superconducting state for the
3107: %optical sum and the kinetic energy. The left hand frame applies to
3108: %Model A of Table~\ref{tab:1} and $\omega_{SF}=82\,$meV and the
3109: %right hand frame is for the band structure Model B of Table~\ref{tab:1} and
3110: %$\omega_{SF} = 10\,$meV. The dotted line in this frame shows
3111: %a $T^2$ law extrapolation of the normal state data for $T>T_c$ to
3112: %zero temperature. Adapted from Ref.~\protect{\onlinecite{ref40}}.
3113: %}
3114: % \label{fig:3}
3115: %\vspace*{5cm}
3116: %\end{figure}
3117: %\clearpage
3118: %\begin{figure}[tp]
3119: % \vspace*{0.5cm}
3120: % \centering
3121: % \includegraphics[width=8cm]{srFig4a.eps}
3122: % \includegraphics[width=8cm]{srFig4b.eps}
3123: % \caption{$W(\Omega,T)$ normalized to its $T=0$ value
3124: %as a function of $T^2$ for the Hubbard model,
3125: %for $\Omega = 0.1\,D$ (left hand frame) and $\Omega = 1.5\,D$
3126: %(right hand frame).
3127: %Symbols are the results of the DMFT calculations; lines are best
3128: %fits to them. The various symbols give four dopings $(x)$.
3129: %Adapted from Ref.~\protect{\onlinecite{ref42}}.
3130: %}
3131: % \label{fig:4}
3132: %\end{figure}
3133: %\clearpage
3134: %\begin{figure}[tp]
3135: % \vspace*{-2cm}
3136: % \centering
3137: % \includegraphics[width=13cm]{srFig5.eps}
3138: % \caption{The relative variation of spectral weight between the
3139: %lowest $T$ and $300\,$K as a function of doping $x$ for various
3140: %cuprates (full symbols) is compared with DMFT calculations
3141: %(open squares) and with the predictions of non interacting models
3142: %[tight binding model (solid line), flat band (dashed line)]. The
3143: %dotted line is a guide to the eye. The simple inclusion of
3144: %correlation effects allows one to reproduce the observed absolute values
3145: %with no need of fitting parameters. Data for LSCO are obtained from
3146: %Refs.~\protect{\onlinecite{ref43}} and \protect{\onlinecite{ref44}},
3147: %and for BSCCO from
3148: %Refs.~\protect{\onlinecite{ref45,ref46,ref47}}
3149: %(BSCCO$_1$) and from Ref.~\protect{\onlinecite{ref48}} (BSCCO$_2$).
3150: %Adapted from Ref.~\protect{\onlinecite{ref42}}.
3151: %}
3152: % \label{fig:5}
3153: %\end{figure}
3154: %\clearpage
3155: %\begin{figure}[tp]
3156: % \centering
3157: % \includegraphics[width=10cm]{srFig6.eps}
3158: % \caption{$T=0$ spectral weight $W_0$ for $\Omega=1.5D$ as a function of
3159: %doping $x$, for both the Hubbard model (solid circles) and for the
3160: %observations in LSCO (open circles). The latter data are obtained by
3161: %integrating the $\sigma_1(\omega)$ of LSCO in Ref.~\protect{
3162: %\onlinecite{ref44}}. Adapted from Ref.~\protect{\onlinecite{ref42}}.
3163: %}
3164: % \label{fig:6}
3165: %\end{figure}
3166: %\clearpage
3167: %\begin{figure}[tp]
3168: % \centering
3169: % \includegraphics[width=12cm]{srFig7.eps}
3170: % \caption{Evolution of the self energy vs frequency dependence with
3171: %temperature for the Einstein spectrum with $a=0.1$ and $\Omega_E =
3172: %0.1$ $(\lambda = 2)$. Frame (a) is for the minus imaginary part,
3173: %$-\Sigma_2(\omega)$, while frame (b) is for the real part,
3174: %$\Sigma_1(\omega)$. In each frame different curves correspond
3175: %to temperatures $t\equiv 2T/D =0.1$ (dash-dotted line),
3176: %0.05 (dotted line),
3177: %0.01 (dashed line), 0.001. All energies are in units of $D/2$.
3178: %Adapted from Ref.~\protect{\onlinecite{ref58}}.
3179: %}
3180: % \label{fig:7}
3181: %\end{figure}
3182: %\clearpage
3183: %\begin{figure}[tp]
3184: % \centering
3185: % \includegraphics[width=11cm]{srFig8.eps}
3186: % \caption{The probability of occupation of the state $n(\omega)$ vs
3187: %normalized energy $2\omega/D$. (a) The results for the case when only
3188: %the temperature dependence of the self energy is included. The normalized
3189: %temperatures are $t \equiv 2T/D = 0.0$ (solid line),
3190: %0.01 (dashed line), 0.03 (dotted line), and 0.05
3191: %(dash-dotted line). (b) Comparison of the
3192: %results when the complete temperature dependence of $n(\omega)$
3193: %[see Eq.~\protect{\eqref{eq:17}}] is accounted for (solid lines)
3194: %with the case presented in frame (a). The normalized
3195: %temperatures are $t=0.01$ (dashed line) and 0.05 (dash-dotted
3196: %line). (All energies are in units of $D/2$.) Adapted from
3197: %Ref.~\protect{\onlinecite{ref49}}.
3198: %}
3199: % \label{fig:8}
3200: %\end{figure}
3201: %\clearpage
3202: %\begin{figure}[tp]
3203: % \centering
3204: % \includegraphics[width=11cm]{srFig9.eps}
3205: % \caption{The variation of the optical sum $S$ vs temperature for the
3206: %interaction strengths (a) $a=0.02$ and (b) $a=0.1$. The results are for
3207: %both the quadratic (solid symbols) and tight binding
3208: %(open symbols) bands, as indicated. In the top frame
3209: %the mass enhancement factor $\lambda=0.8$ (squares) and
3210: %$\lambda=1$ (triangles), while in the bottom frame $\lambda=2$ (squares)
3211: %and $\lambda=4$ (triangles).
3212: %The insert in the top frame illustrates
3213: %the behavior of the optical sum during a sudden ``undressing
3214: %transition'' at $t_{undress}=0.02$ with 20\% hardening of the
3215: %normalized boson frequency $\Omega$ and a corresponding reduction
3216: %of the mass enhancement factor from $\lambda=1.0$ to 0.8. Adapted
3217: %from Ref.~\protect{\onlinecite{ref49}}.
3218: %}
3219: % \label{fig:9}
3220: %\end{figure}
3221: %\clearpage
3222: %\begin{figure}[tp]
3223: % \centering
3224: % \includegraphics[width=11cm]{srFig10.eps}
3225: % \caption{Top frame: minus the imaginary part of the carrier self
3226: %energy $-\Sigma_2(\omega,T)$ vs $\omega$ at the various temperatures
3227: %shown. Middle frame: minus the real part of the carrier self energy
3228: %$-\Sigma_1(\omega,T)$ vs $\omega$. Bottom frame: the function
3229: %$h(T,-\omega)$ of Eq.~\protect{\eqref{eq:18}} vs $\omega$. The
3230: %calculation is for an MMP model, Eq.~\protect{\eqref{eq:29}}, with
3231: %$\omega_{SF} = 20\,$meV, $I^2=0.82$, and $D=800\,$meV.
3232: %}
3233: % \label{fig:10}
3234: %\end{figure}
3235: %\clearpage
3236: %\begin{figure}[tp]
3237: % \centering
3238: % \includegraphics[width=14cm]{srFig11.eps}
3239: % \caption{The reduced optical sum $S(T)$ according to
3240: %Eq.~\protect{\eqref{eq:29a}}
3241: %vs temperature $(T)$ squared. The left hand frame is for
3242: %an MMP model \protect{\eqref{eq:29}} with $\omega_{SF} = 20\,$meV and
3243: %$I^2=0.82$ while the right hand frame is for $\omega_{SF} = 82\,$meV
3244: %and $I^2 = 0.655$.
3245: %}
3246: % \label{fig:11}
3247: %\end{figure}
3248: %\clearpage
3249: %\begin{figure}[tp]
3250: % \centering
3251: % \includegraphics[width=12cm]{fig3bw.eps}
3252: % \caption{The negative of the KE, $-\langle K\rangle$, vs temperature
3253: %showing the breakdown for the various contributions. The solid
3254: %curves are for a small boson frequency $\omega_E=2\,$meV with
3255: %$\lambda=1$. The two thin lines show how the zero temperature value
3256: %is modified by the Sommerfeld term by a very small amount. A similar
3257: %size contribution (actually a bit larger) is illustrated for the
3258: %noninteracting case by dotted curves, where only the
3259: %Sommerfeld term is responsible for the temperature variation. Finally,
3260: %for higher boson frequencies, the dashed curves illustrate
3261: %the amount of temperature variation due to the Sommerfeld term
3262: %compared with the rest. The band width $W = 1\,$eV.
3263: % Adapted from Ref.~\protect{\onlinecite{ref52}}.
3264: %}
3265: % \label{fig:12}
3266: %\end{figure}
3267: %\clearpage
3268: %\begin{figure}[tp]
3269: % \centering
3270: % \includegraphics[width=13cm]{srFig13.eps}
3271: % \caption{The partial spectral weight integrated up to various
3272: %frequencies as a function of temperature. Below $16\,000\,$cm$^{-1}$
3273: %there is an increase in spectral weight as the temperature is lowered
3274: %signaling a line narrowing on this frequency scale. Below $T_c$ there
3275: %is strong loss of spectral weight to the superconducting condensate.
3276: %There is no evidence of any precursors to superconductivity at
3277: %$67\,$K. In the inset we show $N_{eff}(\omega)$ at $295\,$K. Adapted
3278: %from Ref.~\protect{\onlinecite{ref64}}.
3279: %}
3280: % \label{fig:13}
3281: %\vspace*{-1cm}
3282: %\end{figure}
3283: %\clearpage
3284: %\begin{figure}[tp]
3285: % \centering
3286: % \includegraphics[width=9cm]{srFig14.eps}
3287: % \caption{Experimental values of the $ab$-plane spectral function
3288: %defined in Eq.~\protect{\eqref{eq:30}}.
3289: %Adapted from Ref.~\protect{\onlinecite{ref34}}.
3290: %}
3291: % \label{fig:14}
3292: %\vspace*{5cm}
3293: %\end{figure}
3294: %\endfloats
3295: %\begin{figure}[tp]
3296: % \centering
3297: % \includegraphics[width=11cm]{srFig15.eps}
3298: % \caption{The charge carrier-spin spectral density $I^2\chi(\omega)$
3299: % determined from optical scattering data of YBCO$_{6.95}$
3300: %at various temperatures.
3301: % Solid gray curve $T=90\,$K, dash-dotted $T=80\,$K, dashed $T=60\,$K,
3302: %dotted $T=40\,$K, and black solid $T=10\,$K. Note the growth in strength
3303: %of the $41\,$meV optical resonance as the temperature is lowered.
3304: %Adapted from Ref.~\protect{\onlinecite{ref70}}.
3305: %}
3306: % \label{fig:15}
3307: %\end{figure}
3308: %\clearpage
3309: %\begin{figure}[tp]
3310: % \centering
3311: % \includegraphics[width=11cm]{srFig16.eps}
3312: % \caption{
3313: %Temperature dependence of the conductivity $\sigma_1(\omega)$ at
3314: %microwave frequency $\omega=0.144\,$meV in YBCO$_{6.95}$.
3315: %The open circles are results based on
3316: %our model spectral density (see Fig.~\protect{\ref{fig:15}}) obtained
3317: %from inversion of optical conductivity data and the dashed line is
3318: %based on an
3319: %MMP model with low frequency cutoff. The solid triangles include impurities
3320: %with the solid line from Ref.~\protect{\onlinecite{ref74}}. The solid
3321: %squares represent experimental data by Bonn {\it et al.}\protect{%
3322: %\cite{ref75}} Adapted from Ref.~\protect{\onlinecite{ref69}}.
3323: %}
3324: % \label{fig:16}
3325: %\end{figure}
3326: %\clearpage
3327: %\begin{figure}[tp]
3328: % \centering
3329: % \includegraphics[width=12cm]{srFig17.eps}
3330: % \caption{The optical sum of optimally doped BSCCO
3331: %as a function of the square of the temperature
3332: %for the band structure Model A of Table~\protect{\ref{tab:1}} with
3333: %different values of $\omega_{SF}$. Also in one case a low frequency
3334: %cutoff is applied to the spin susceptibility. Note the significance
3335: %of the $T^2$-variation on the value of $\omega_{SF}$. The thick solid
3336: %line represents experimental normal state data of Molegraaf {\it et al.}%
3337: %\protect{\cite{ref48}} also shown in the top frame of
3338: %Fig.~\protect{\ref{fig:14}}.
3339: %Adapted from Ref.~\protect{\onlinecite{ref40}}.
3340: %}
3341: % \label{fig:17}
3342: %\end{figure}
3343: %\clearpage
3344: %\begin{figure}[tp]
3345: % \centering
3346: % \includegraphics[width=7cm]{srFig18a.eps}
3347: % \includegraphics[width=7cm]{srFig18b.eps}
3348: % \caption{(a) $1/\tau(\omega)$ vs $\omega$ for various BSCCO samples from
3349: %Ref.~\protect{\onlinecite{ref87}} (OD overdoped, OPT optimal doped,
3350: %UD underdoped). (b) Calculated sum rule violation $(-\Delta E_K)$ vs
3351: %doping $x$ (solid circles).
3352: %The curve is $T_c$. The parameters (meV) extracted from
3353: %(a) are $\Gamma_N$ (1, 22, 27, 37), $\alpha$ (.65, .75, .88, .98),
3354: %$\omega_0$ (54, 71, 76, 83), and $\Delta_{max}$ (24, 32, 41, 54) for
3355: %OD70, OPT90, UD82, and UD67, respectively. Also shown in (b) are the
3356: %experimental results (open squares from Ref.~\protect{\onlinecite{ref46}},
3357: %open diamonds from Ref.~\protect{\onlinecite{ref48}}). The theoretical
3358: %doping trend in (b) is due to the increasing offset in $1/\tau$ seen
3359: %in (a). Adapted from Ref.~\protect{\onlinecite{ref78}}.
3360: %}
3361: % \label{fig:18}
3362: %\end{figure}
3363: %\clearpage
3364: %\begin{figure}[tp]
3365: % \centering
3366: % \includegraphics[width=11cm]{srFig19.eps}
3367: % \caption{Spectral weight of the overdoped Bi2212 sample, integrated
3368: %up to $1\,$eV, plotted vs $T^2$, from Ref.~\protect{\onlinecite{ref46}}.
3369: %Closed symbols: spectral weight in the normal state, open symbols:
3370: %spectral weight in the superconducting state, including the weight
3371: %of the superfluid. The errors in the {\it relative} variations of the
3372: %spectral weight are of the size of the symbols. Adapted from
3373: %Ref.~\protect{\onlinecite{ref87a}}.
3374: %}
3375: % \label{fig:19}
3376: %\end{figure}
3377: %\clearpage
3378: %\begin{figure}[tp]
3379: % \centering
3380: % \includegraphics[width=11cm]{srFig20.eps}
3381: % \caption{Change $\Delta E_{kin}$ of the KE, in meV per copper site
3382: %vs the charge $p$ per copper with
3383: %respect to $p_{opt}$ [Eq.~\protect{\eqref{eq:31}}] in BSCCO.
3384: %Full diamonds: data from
3385: %Ref.~\protect{\onlinecite{ref47}}, high-frequency cutoff $1\,$eV. Open
3386: %circles: data from Ref.~\protect{\onlinecite{ref48}}, high frequency
3387: %cutoff $1.25\,$eV. Error bars: vertical, uncertainties due to the
3388: %extrapolation of the temperature dependence of the normal state spectral
3389: %weight down to zero temperature; horizontal, uncertainties resulting
3390: %from $T_c/T_{c,max}$ through Eq.~\protect{\eqref{eq:31}}.
3391: %Deutscher {\it et al.}\protect{\cite{ref87a}} took
3392: %$T_{c,max} = (83\pm2)\,$K for films and $(91\pm2)\,$K for crystals.
3393: %Adapted from Ref.~\protect{\onlinecite{ref87a}}.
3394: %}
3395: % \label{fig:20}
3396: %\end{figure}
3397: %\clearpage
3398: %\begin{figure}[tp]
3399: % \centering
3400: % \includegraphics[width=12cm]{srFig32.eps}
3401: % \caption{The difference between the superconducting and normal state
3402: %energies as a function of temperature. The following curves are shown:
3403: %up-triangles - $E_{kin-S}-E_{kin-N}$; squares - $E_{x-S}-E_{x-N}$;
3404: %diamonds - $E_{tot-S}-E_{tot-N}$; circles - $\mu_S-\mu_N$.
3405: %Adapted from Ref.~\protect{\onlinecite{ref89b}}.
3406: %}
3407: % \label{fig:32}
3408: %\end{figure}
3409: %\clearpage
3410: %\begin{figure}[tp]
3411: % \centering
3412: % \includegraphics[width=11cm]{srFig21.eps}
3413: % \caption{Minus the kinetic energy, $-\langle T_{xx}\rangle$,
3414: %vs temperature for various
3415: %degrees of elastic scattering. Below $T_c$ we change the elastic
3416: %scattering rate smoothly to zero [as $\Gamma_0 (T/T_c)^4$] as the
3417: %temperature is lowered.
3418: %Adapted from Ref.~\protect{\onlinecite{ref90}}
3419: %}
3420: % \label{fig:21}
3421: %\end{figure}
3422: %\clearpage
3423: %\begin{figure}[tp]
3424: % \centering
3425: % \includegraphics[width=10cm]{srFig22.eps}
3426: % \caption{Kinetic energy per bond, $\langle k_x\rangle$, as a function of
3427: %temperature for non interacting tight-binding electrons (TB), the BCS
3428: %solution (BCS), and the phase fluctuation (PP) model for $\mu=0$
3429: %$(\langle n\rangle = 1)$. The large vertical arrows indicate the increase
3430: %in KE upon pairing, relative to the free tight-binding model, and
3431: %the small arrows indicate the additional increase due to phase
3432: %fluctuations. This additional {\it phase fluctuation energy} rapidly
3433: %vanishes near $T_c=T_{KT}$, which causes the significant change in the
3434: %OS upon entering the superconducting state at $T_{KT}=0.1\,t$. Note
3435: %that the thick line follows the actual KE encountered in our model, when
3436: %going from the pseudogap to the superconducting regime. Adapted from
3437: %Ref.~\protect{\onlinecite{ref97}}.
3438: %}
3439: % \label{fig:22}
3440: %\end{figure}
3441: %\clearpage
3442: %\begin{figure}[tp]
3443: % \centering
3444: % \includegraphics[width=11cm]{srFig23.eps}
3445: % \caption{Kinetic energy contribution from phase fluctuations
3446: %$\delta\langle k_x\rangle \equiv \langle k_x\rangle_{PP}-
3447: %\langle k_x\rangle_{BCS}$. One can clearly see the sharp decrease of KE
3448: %near the Kosterlitz-Thouless transition $T_{KT}=0.1\,t\equiv T_c$.
3449: %$\Delta E_k$ gives an estimate of the kinetic condensation energy.
3450: %Adapted from Ref.~\protect{\onlinecite{ref97}}.
3451: %}
3452: % \label{fig:23}
3453: %\end{figure}
3454: %\clearpage
3455: %\begin{figure}[tp]
3456: % \centering
3457: % \includegraphics[width=12cm]{srFig24.eps}
3458: % \caption{Restricted optical sums with (solid line) and without
3459: %(dot-dashed line) vertex corrections, and Drude weights with (dashed
3460: %line) and without (dotted line) vertex corrections.
3461: %Adapted from Ref.~\protect{\onlinecite{ref119}}.
3462: %}
3463: % \label{fig:24}
3464: %\end{figure}
3465: %\clearpage
3466: %\begin{figure}[tp]
3467: % \centering
3468: % \includegraphics[width=12cm]{srFig25.eps}
3469: % \caption{Optical conductivity $\sigma(\omega)$ with (solid line)
3470: %and without (dashed line) vertex corrections.
3471: %Adapted from Ref.~\protect{\onlinecite{ref119}}.
3472: %}
3473: % \label{fig:25}
3474: %\end{figure}
3475: %\clearpage
3476: %\begin{figure}[tp]
3477: % \centering
3478: % \includegraphics[width=13cm]{srFig26.eps}
3479: % \caption{Spectral weight in the presence of a $t'$ term in the band
3480: %dispersion.
3481: %Here $W(D,T)$ (Eq.~\protect{\eqref{eq:43}}, dash-dotted line),
3482: %$W^{DDW}(D,T)$ (Eq.~\protect{\eqref{eq:49a}} (solid line) are shown
3483: %together with $W(T)$ (dashed line) which is for the non interacting
3484: %case. The parameters are $t'=0.3\,t$, $T_{DDW}=0.12\,t$,
3485: %$D(0) = 4T_{DDW}$, and the doping $\delta = 0.1$. The chemical potential is
3486: %evaluated self consistently at each temperature.
3487: %Observe that near $T_{DDW}$ a small decrease of $W^{DDW}$ with
3488: %respect to $W(T)$ is seen, due to the change of chemical potential
3489: %near $T_{DDW}$. Adapted from Ref.~\protect{\onlinecite{ref113}}.
3490: %}
3491: % \label{fig:26}
3492: %\end{figure}
3493: %\clearpage
3494: %\begin{figure}[tp]
3495: % \centering
3496: % \includegraphics[width=14cm]{srFig27.eps}
3497: % \caption{Spectral weight in units of $e^2\pi/\hbar^2$ in the normal state
3498: %(dash double-dotted line), in the DDW state
3499: %Eq.~\protect{\eqref{eq:49a}} (solid line)
3500: %and in DDW+SC state Eq.~\protect{\eqref{eq:45}} (dashed line).
3501: %The values of parameters are
3502: %for the doping $\delta = 0.13$ [$D_0(T=0)=0.92\,t$,
3503: %$\Delta(T=0)=0.064\,t$, see Refs.~\protect{\onlinecite{ref107}} and
3504: %\protect{\onlinecite{ref108}}
3505: %for further details. The critical temperature is marked by the arrow.
3506: %Observe that the decrease of $W^{DDW}(D,\Delta,T)$ below $T_c$ is
3507: %small. Inset: spectral weight plotted as a function of
3508: %$(T/t)^2$. Adapted from Ref.~\protect{\onlinecite{ref114}} (see also
3509: %Ref.~\onlinecite{ref108}).
3510: %}
3511: % \label{fig:27}
3512: %\end{figure}
3513: %\clearpage
3514: %\begin{figure}[tp]
3515: % \centering
3516: % \includegraphics[width=10cm]{srFig28.eps}
3517: % \caption{Selection of conductivity spectra for
3518: %three BSCCO samples, underdoped, $T_c=70\,$K (B70KUND, top frame),
3519: %optimally doped, $T_c=80\,$K (B80KOPT, middle frame), and overdoped,
3520: %$T_c=63\,$K (B63KOVR, bottom frame). For the
3521: %B70KUND sample, within the showed spectral range, the spectra at
3522: %$50\,$K and $10\,$K are indistinguishable.
3523: % Adapted from Ref.~\protect{\onlinecite{ref47}}.
3524: %}
3525: % \label{fig:28}
3526: %\end{figure}
3527: %\clearpage
3528: %\begin{figure}[tp]
3529: % \centering
3530: % \includegraphics[width=11cm]{srFig29.eps}
3531: % \caption{Rato $\Delta W/W_S$ vs frequency showing the exhaustion of the
3532: %FGT sum rule at conventional energies for the OVR (diamonds)
3533: %and OPT (triangles) samples. An
3534: %unconventional ($\sim 16 000\,$cm$^{-1}$ or $2\,$eV) energy scale is
3535: %required for the UND sample (circles). Note that the frequency scale changes at
3536: %800 and $8 000\,$cm$^{-1}$. The changes in spectral weight are taken
3537: %between $80\,$K - $10\,$K, $91\,$K - $10\,$K, and $100\,$K - $10\,$K
3538: %for the OVR, OPT, and UND samples, respectively. Adapted from
3539: %Ref.~\protect{\onlinecite{ref47}}.
3540: %}
3541: % \label{fig:29}
3542: %\end{figure}
3543: %\clearpage
3544: %\begin{figure}[tp]
3545: % \centering
3546: % \includegraphics[width=12cm]{srFig30.eps}
3547: % \caption{The normalized weight of the condensate [$N_n(\omega)-
3548: %N_s(\omega)]/\rho_{s,a}$ vs $\omega$ for optimally doped YBCO$_{6.95}$
3549: %(solid line)
3550: %and underdoped YBCO$_{6.60}$ (dotted line) along the $a$-axis direction.
3551: %The condensate for the optimally
3552: %doped material has saturated by $\simeq 800\,$cm$^{-1}$, while in the
3553: %underdoped material the condensate is roughly 80\% formed by this
3554: %frequency, but the other 20\% is not recovered until much higher
3555: %frequencies. The error bars on the curve for the underdoped material
3556: %indicate the uncertainty associated with the FGT sum rule. Inset:
3557: %The low-frequency region. Adapted from Ref.~\protect{\onlinecite{ref121}}.
3558: %}
3559: % \label{fig:30}
3560: %\end{figure}
3561: \clearpage
3562: %\begin{figure}[tp]
3563: % \centering
3564: % \includegraphics[width=10cm]{srFig31.eps}
3565: % \caption{Top frame: Optical spectral weight $W(\omega,T) = \int_{0^+}^\omega\!
3566: %d\nu\,\sigma_1(\nu)$ for various cases as a function of $\omega$. The
3567: %dotted (dash-double-dotted) curve is for the normal state at $T=10\,$K
3568: %($T=95\,$K), the solid curve for the superconducting state at $T=10\,$K.
3569: %The dashed (dash-dotted) curve is the difference curve between superconducting
3570: %at $T=10\,$K and normal states
3571: %at $T=10\,$K ($T=95\,$K). The approach of the difference curve to its
3572: %saturated large $\omega$ value depends significantly on the temperature
3573: %used for the subtracted normal state. The thin dash-double-dotted
3574: %horizontal line is the value of the condensate contribution
3575: %(penetration depth). The bottom frame
3576: %shows the real part of the conductivity for the normal state at
3577: %$T=293\,$K (dashed curve), $T=95\,$K (dash-dotted curve), $T=10\,$K
3578: %(dotted curve), and for the superconducting state at a $T=10\,$K
3579: %(solid curve). All curves are for YBCO$_{6.95}$ with the impurity parameter
3580: %given in Ref.~\protect{\onlinecite{ref122a}}.
3581: %}
3582: % \label{fig:31}
3583: %\end{figure}
3584: \end{document}
3585: