cond-mat0506354/vma.tex
1: \documentclass[10pt,aps,prb,showpacs,superscriptaddress,floatfix,
2: twocolumn]{revtex4}
3: \usepackage{amsmath}
4: \usepackage{citesort}
5: \usepackage{graphicx}
6: \usepackage{dcolumn}
7: %\renewcommand\baselinestretch{1.6}
8: \begin{document}
9: \title{Optical sum in Nearly Antiferromagnetic Fermi Liquid Model}
10: \author{E. Schachinger}
11: \email{schachinger@itp.tu-graz.ac.at}
12: \homepage{www.itp.tu-graz.ac.at/~ewald}
13: \affiliation{Institute of Theoretical and Computational Physics\\
14: Graz University of Technology, A-8010 Graz, Austria}
15: \author{J.P. Carbotte}
16: \affiliation{Department of Physics and Astronomy, McMaster University,\\
17: Hamilton, Ontario, L8S 4M1 Canada}
18: \date{\today}
19: \begin{abstract}
20: We calculate the optical sum (OS) and the kinetic energy (KE) for a
21: tight binding band in the Nearly Antiferromagnetic Fermi Liquid
22: (NAFFL) model which has had some success in describing the electronic
23: structure of the high $T_c$ cuprates. The interactions among electrons
24: due to the exchange
25: of spin fluctuations profoundly change the probability of occupation
26: $(n_{{\bf k},\sigma})$ of states of momentum {\bf k} and spin $\sigma$
27: which is the central quantity in the calculations of OS and KE. Normal and
28: superconducting states are considered as a function of temperature.
29: Both integrals are found to depend importantly on interactions
30: and an independent electron model is inadequate.
31: \end{abstract}
32: \pacs{74.20.Mn 74.25.Gz 74.72.-h}
33: \maketitle
34: \newpage
35: \section{Introduction}
36: 
37: The high $T_c$ oxides fall in the category of highly correlated systems.
38: A manifestation of this fact is that in the
39: underdoped regime there exists a pseudogap.\cite{r1,r2,r3,r4,r5}
40: Precisely how it is to be described
41: remains controversial.\cite{r5,r6,r7,r23a,r8,r9,r10,r11,r12,r13}
42: In certain theories it is closely related to
43: superconducting correlations,\cite{r5,r6,r7,r23a} and the
44: superconducting transition temperature $T_c$ is the temperature at which
45: phase coherence
46: is lost. In other theories the
47: pseudogap has its origin in completely different correlations
48: and is a manifestation of a
49: competing order such as a $d$-density wave (DDW).\cite{r8,r9,r10,r11,%
50: r12,r13} In any case interactions among charge carriers play an
51: important role in such systems and cannot be ignored in any realistic
52: approach to their properties even around optimum doping which is the
53: case of interest here.
54: 
55: While many aspects of the superconducting state can be understood
56: qualitatively on the basis of extensions of BCS theory,
57: particularly around optimum doping,
58: the search
59: for essential differences\cite{r14} has remained an important
60: avenue of investigation. In particular the idea of kinetic energy
61: as opposed to potential energy driven superconductivity
62: (i.e.: the kinetic energy is reduced at the transition temperature)
63: and its relation to the OS has
64: recently been given serious consideration in
65: theory\cite{r14,r15,r16,r17,r18,r19,r20,r21,r22,r23}
66: and experimentally.\cite{r24,r25,r26,r27,r28,r29}
67: An issue of importance is the relation of the
68: kinetic energy (KE) to the optical sum (OS). A recent paper\cite{r28}
69: has provided some insight into this relationship and has given a
70: comparison with experiment for the temperature variation of the OS
71: in the normal state, its change at $T_c$, and its further
72: evolution in the superconducting state. Correlations beyond
73: BCS pairing were not considered, however.
74: There have also been several recent studies of the
75: temperature dependence of the OS, experimentally for the system
76: LA$_{2-x}$Sr$_x$CuO$_4$, Ref.~\onlinecite{r42a}, and, theoretically,
77: using other models.\cite{r42b,r42c,r42d}
78: 
79: Within a BCS model, the condensation is potential energy driven and,
80: in fact, the KE increases because the probability of occupation
81: of the state {\bf k} $(n_{\bf k})$ becomes smeared at the Fermi
82: energy by the opening of the gap. This translates into a decrease in the
83: OS as compared with its normal state value. But this is opposite to
84: the behavior obtained experimentally in Ref.~\onlinecite{r26}.
85: However, as mentioned, the theoretical discussion of Ref.~\onlinecite{r28}
86: is based on a non interacting model in the normal state and includes
87: pairing correlations in the superconducting state only at the level
88: of BCS. More sophisticated formulations of the theory of
89: superconductivity could give different results.
90: Also, interactions can significantly modify the results even
91: in the normal state, as has been demonstrated recently by
92: Knigavko {\it et al.}\cite{r30} in a simplified model in which
93: the charge carriers are coupled to a single Einstein mode. While
94: only the normal state was considered, it was found that
95: boson hardening (softening) results in an increase (decrease)
96: in the OS and that interactions
97: play an important role in determining temperature dependences.
98: 
99: In this paper we study a tight binding model with an emphasis on the
100: effect interactions can have on the OS, particularly on its
101: temperature dependence and its relationship to the KE. Here,
102: the interactions among the charge carriers are treated in the
103: Nearly Antiferromagnetic Fermi Liquid (NAFFL) model.\cite{r31}
104: A review of its main properties and successes is given in
105: Ref.~\onlinecite{r31}. The model is phenomenological and falls into
106: the general class of boson exchange models where the interaction
107: between electrons proceeds through the exchange of spin
108: fluctuations.\cite{r32} The imaginary part of the spin susceptibility
109: replaces the phonon propagator of the classic electron-phonon
110: Eliashberg theory. The spin susceptibility could be calculated
111: from microscopic theory. More usually, however, it is fit to
112: experimental data, specifically to NMR in Ref.~\onlinecite{r32}.
113: The basic idea is that the doped metallic cuprates are near an
114: antiferromagnetic phase boundary and that coupling to spin
115: fluctuations is therefore strong.
116: While there is no consensus as to the validity of such a model when
117: applied to the oxides the NAFFL model has been widely discussed and
118: has had considerable
119: successes particularly in correlating
120: superconducting properties,\cite{r33,r34,r35,r35a,r35b,r35c} and has
121: been used to
122: give a detailed description of the optical properties\cite{r35,r36,%
123: r37,r38,r39,r40,r43a} in the high $T_c$ oxides at optimum
124: doping. In any case, the NAFFL model\cite{r41}
125: provides a convenient and specific
126: framework within which  we can study the effect of correlations on the
127: OS and on related properties.
128: 
129: In Sec.~\ref{sec:2} we summarize the basic equations that are needed to
130: compute the optical sum integral as well as the KE. They are a set
131: of three coupled generalized Eliashberg equations written for any
132: momentum {\bf k} in the two dimensional CuO$_2$ Brillouin zone.
133: They involve renormalized Matsubara frequencies, the renormalized
134: quasiparticle energies as well as the superconducting energy gap.
135: The interaction
136: between the charge carriers is mediated by the exchange of spin
137: fluctuations and involve the spin susceptibility. Fast Fourier
138: transforms (FFT) provide solutions which give a $d$-wave gap as
139: observed in experiments. In Sec.~\ref{sec:3} we apply our solutions to
140: evaluate the probability of occupation of the state of momentum
141: {\bf k} and spin $\sigma$ $(n_{{\bf k},\sigma})$ from which the OS
142: and the KE follow. Their temperature dependence is studied. Comparison
143: with the non interacting case is made and it is concluded that
144: interactions can profoundly modify results. Different behaviors
145: can result depending on the choice of microscopic parameters. In
146: Sec.~\ref{sec:4} we consider explicitly the superconducting state. Here,
147: again, generalized Eliashberg equations give variations with $T$
148: which are considerably different from earlier BCS  results.
149: When the electron-exchange boson interaction is taken as
150: temperature independent and independent of state, the OS is lower
151: in the superconducting than in the normal state. Nevertheless,
152: it can keep increasing with decreasing temperature in contrast
153: to BCS where it was found to decrease. This increase can be traced to the
154: underlying temperature dependence of the OS which depends on
155: details of the band structure and interactions involved, for
156: example on the model spin susceptibility. Further, if we take account
157: of the low energy gaping of the spin susceptibility
158: which is brought
159: about by the superconducting transition (a process which is not
160: operative in the normal state) the KE can be further decreased
161: and there is an additional increase in the OS which can effectively
162: increase faster than it does in the normal state. In this case
163: the KE in the superconducting state with low energy gaping of the
164: spin susceptibility can be less
165: than in the normal state without low energy gaping.
166: An important conclusion
167: of our work is that the observation of a faster increase in the OS with
168: decreasing temperature in the superconducting state than in the
169: normal state cannot unambiguously be taken
170: to be an indication of kinetic energy driven superconductivity
171: in contrast to a recent claim by van der Marel {\it et al.}\cite{r28}
172: In Sec.~\ref{sec:5} we provide a brief conclusion.
173: 
174: We use units in which $\hbar = c = 1$ throughout this paper.
175: 
176: \section{Formalism}
177: \label{sec:2}
178: 
179: In the NAFFL model the interaction between holes proceeds through the
180: exchange of spin fluctuations and the spin susceptibility
181: $\chi({\bf q},\omega)$ plays a central role. The three Eliashberg
182: equations for the renormalized frequencies $\tilde{\omega}({\bf k},i\omega_n)$,
183: the energy renormalization $\xi({\bf k},i\omega_n)$ and the pairing energy
184: $\phi({\bf k},i\omega_n)$ as a function of momentum {\bf k} in the
185: two dimensional CuO$_2$ Brillouin zone, and of fermionic Matsubara
186: frequencies $i\omega_n = i\pi T(2n+1),\,n=0, \pm 1, \pm 2,\ldots$
187: and the temperature $T$ are\cite{r35a,r35b,r35c}
188: \begin{subequations}
189: \label{eq:1}
190: \begin{eqnarray}
191:   \tilde{\omega}({\bf k},i\omega_n) &=& \omega_n+
192:   T\sum\limits_m\sum\limits_{{\bf k}'}
193:   \lambda_{SF}({\bf k}-{\bf k}',i\omega_n-i\omega_m)ΒΈ\nonumber\\
194:   &&\times
195:   \frac{\tilde{\omega}({\bf k}',i\omega_m)}{D({\bf k}',i\omega_m)},
196:   \label{eq:1a}\\
197:   \xi({\bf k},i\omega_n) &=&
198:   -T\sum\limits_m\sum\limits_{{\bf k}'}
199:   \lambda_{SF}({\bf k}-{\bf k}',i\omega_n-i\omega_m)\nonumber\\
200:   &&\times
201:   \frac{\epsilon_{{\bf k}'}+\xi({\bf k}',i\omega_m)}{D({\bf k}',i\omega_m)},
202:   \label{eq:1b}\\
203:   \phi({\bf k},i\omega_n) &=&
204:   -T\sum\limits_m\sum\limits_{{\bf k}'}
205:   \lambda_{SF}({\bf k}-{\bf k}',i\omega_n-i\omega_m)\nonumber\\
206:   &&\times
207:   \frac{\phi({\bf k}',i\omega_m)}{D({\bf k}',i\omega_m)}.
208:   \label{eq:1c}
209: \end{eqnarray}
210: \end{subequations}
211: In Eqs.~\eqref{eq:1} $D({\bf k},i\omega_n)$ is given by
212: \begin{equation}
213:   \label{eq:2}
214:   D({\bf k},i\omega_n) = \tilde{\omega}^2({\bf k},i\omega_n)+
215:   \left[\epsilon_{\bf k}+\xi({\bf k},i\omega_n)\right]^2+
216:   \phi^2({\bf k},i\omega_n),
217: \end{equation}
218: and $\epsilon_{\bf k}$ is the charge carrier dispersion relation. In a
219: tight binding model without the inclusion of the coupling to the
220: spin fluctuations it is given by:
221: \begin{eqnarray}
222:   \epsilon_{\bf k} &=& -2\left\{t\left[\cos(ak_x)+\cos(ak_y)\right]
223:   \right.\nonumber\\ &&-\left.
224:   2t'\cos(ak_x)\cos(ak_y)\right\}-\mu^\ast.
225:   \label{eq:3}
226: \end{eqnarray}
227: Here, $a$ is the lattice parameter in the copper oxide plane,
228: $t$ the nearest neighbor hopping,
229: $t'$ the next nearest neighbor hopping, and $\mu^\ast$ the chemical
230: potential. We will discuss within this
231: context, two tight binding models with the parameters given in
232: Table~\ref{tab:1}. The corresponding Fermi surface is presented in
233: Fig.~\ref{fig:2}a for model A and in Fig.~\ref{fig:2}b for model B.
234: These figures define also the various points in the Brillouin zone.
235: The dotted lines indicate the antiferromagnetic Brillouin zone
236: boundary. For model A, the Fermi surface crosses the anti-ferromagnetic
237: \begin{table}[tp]
238: \caption{\label{tab:1}The two tight binding models used within this
239: paper. Model A corresponds to the tight binding model discussed by
240: van der Marel {\it et al.}\protect{\cite{r28}} $t$ and $t'$ are
241: given in meV, the critical temperature $T_c$ in K, and the filling
242: $\langle n\rangle$ is defined in Eq.~\protect{\eqref{eq:9}}.
243: }
244: \begin{ruledtabular}
245: \begin{tabular}{ldddd}
246: Model & t & t' & \langle n\rangle & T_c\\
247: \hline
248: A & 148.8 & 40.9 & 0.425 & 90\\
249: B & 100.0 & 16.0 & 0.4 & 100\\
250: \end{tabular}
251: \end{ruledtabular}
252: \end{table}
253: Brillouin zone around $X$ and symmetry related points. These points
254: are `hot spots' for which Fermi surface to Fermi surface transitions
255: are possible with momentum transfer $(\pi/a,\pi/a)$ (nesting
256: property). Model A has
257: been used previously in Ref.~\onlinecite{r28} to discuss the OS and
258: KE in Bi$_2$Sr$_2$Cu$_2$O$_{8+\delta}$ (BSCCO) assuming no interactions
259: between the electrons in the
260: normal state. Therefore, it is natural to employ this same model
261: in the present study which extends the previous calculations to
262: include exchange of spin fluctuations between charge carriers in
263: the NAFFL model. The second model, Model B, was chosen to contrast
264: with the first. It has no hot spots and its Fermi surface is closed
265: around the $\Gamma$-point in contrast to Model A which has a Fermi surface
266: which is closed around the $M$-point. It also has a smaller value of
267: $t$ which, on its own, would imply a smaller absolute value of KE.
268: These differences in band parameters lead, as we shall see, to
269: some differences in KE and optical sum at $T=0$ in the
270: normal state.
271: 
272: In the phenomenological model of Millis {\it et al.}\cite{r31,r32} and
273: Monthoux {\it et al.}\cite{r41} (MMP-model) the magnetic susceptibility
274: was fit to NMR data and its imaginary part takes on the form
275: \begin{equation}
276:   \label{eq:4}
277:   \Im{\rm m}\chi_{MMP}({\bf q},\omega) =
278:   \frac{\chi_{\bf Q}\,(\omega/\omega_{SF})}{\left[1+\zeta^2({\bf q}-{\bf Q})^2
279:   \right]^2+(\omega/\omega_{SF})^2},
280: \end{equation}
281: where $\chi_{\bf Q}$ is the static susceptibility, {\bf Q} is the
282: commensurate antiferromagnetic wave vector $(\pi/a,\pi/a)$
283: in the upper right hand quadrant of the CuO$_2$-plane Brillouin zone
284: and symmetry related points. $\zeta$
285: \begin{figure}[tp]
286: \vspace*{-1cm}
287:   \includegraphics[width=9cm]{vmFig1a.eps}\\
288: \vspace*{-1cm}
289:   \includegraphics[width=9cm]{vmFig1b.eps}
290: \vspace*{-10mm}
291:   \caption{The Fermi surface for the non interacting system (dashed line)
292: and the system with interaction (solid line). a) model A of Table~\ref{tab:1},
293: b) for model B.}
294:   \label{fig:2}
295: \end{figure}
296: is the magnetic coherence length, and $\omega_{SF}$ a characteristic
297: spin fluctuation frequency. We set $\zeta = 2.5\,a$ throughout this
298: paper and various values for $\omega_{SF}$ are investigated.
299: 
300: The kernel $\lambda_{SF}({\bf q},i\nu_{n-m})$ in Eqs.~\eqref{eq:1}
301: with momentum transfer ${\bf q} = {\bf k}-{\bf k}'$ and the
302: bosonic Matsubara frequency $i\nu_{n-m} = i\omega_n-i\omega_m$ is
303: given as
304: \begin{equation}
305:   \label{eq:5}
306:   \lambda_{SF}({\bf q},i\nu_n) = \frac{g^2\chi_{\bf Q}}
307:   {1+\zeta^2({\bf q}-{\bf Q})^2+(\vert\nu_n\vert/\omega_{SF})},
308: \end{equation}
309: with $g^2\chi_{\bf Q}$ adjusted to get the desired value of the critical
310: temperature $T_c$ for a certain value of $\omega_{SF}$ from the
311: solution of the linearized Eqs.~\eqref{eq:1}. This defines the model.
312: 
313: The aim of this paper is to investigate the effect interactions have on
314: the OS defined as
315: \begin{equation}
316:   \label{eq:6}
317:   %\frac{\pi\, e^2}{\hbar^2}
318:    \pi\,e^2\,I_\sigma = 
319:   \int\limits_{-\Omega}^\Omega\!d\omega\,\Re{\rm e}\sigma_{xx}(\omega) =
320:   \frac{\pi\,e^2}{V}\sum\limits_{{\bf k},\sigma}
321:   n_{{\bf k},\sigma}\frac{\partial^2\epsilon_{\bf k}}{\partial k_x^2},
322: \end{equation}
323: where $e$ is the charge on the electron, $V$ the volume, and
324: $n_{{\bf k},\sigma}$ is the probability of occupation of a state of
325: momentum {\bf k} and spin $\sigma$. Finally, $\sigma_{xx}(\omega)$ is
326: the optical conductivity. The integral in Eq.~\eqref{eq:6} is to be
327: taken over the single band with $\Omega$, the upper limit in the
328: integral of Eq.~\eqref{eq:6}, large enough to include all
329: possible transitions in that band. We are also interested
330: in the relationship between the OS and the kinetic energy. By
331: definition
332: \begin{equation}
333:   \label{eq:7}
334:   I_{\rm KE} = \left\langle H_{\rm KE}\right\rangle = \frac{a^2}{V}
335:   \sum\limits_{{\bf k},\sigma} n_{{\bf k},\sigma}\epsilon_{\bf k}.
336: \end{equation}
337: We will see that, to a good approximation $I_\sigma$ and
338: $-I_{\rm KE}$ are nearly proportional to each other. Thus,
339: \begin{equation}
340:   \label{eq:8}
341:   \rho_L \propto \frac{1}{\pi e^2}\int\limits_{-\Omega}^\Omega\!
342:   d\omega\,\Re{\rm e}\,\sigma_{xx}(\omega) \approx -\frac{1}{2}\left\langle
343:   H_{\rm KE}\right\rangle.
344: \end{equation}
345: holds approximately. (An equal sign would be appropriate if the
346: dispersion relation \eqref{eq:3} contained only nearest neighbor
347: interaction, i.e.: $t'=0$.) Here, $\rho_L$ is the experimentally
348: determined value of the OS $(I_\sigma)$.
349: 
350: Interactions have a profound effect on the probability of occupation of
351: the state $\vert{\bf k},\sigma\rangle$ which would be one or zero for
352: occupied and unoccupied states respectively in the non interacting case.
353: In Fig.~\ref{fig:1} we show results for $n_{\bf k}$, along
354: certain selected
355: \begin{figure}[tp]
356: \vspace*{-3mm}
357:   \includegraphics[width=9cm]{vmFig2.eps}
358:   \caption{The occupation number $n_{\bf k}$ for both spin states for
359: selected directions in the CuO$_2$ Brillouin zone. Model A of
360: Table~\protect{\ref{tab:1}} was used.
361: Top frame: the non interacting case. Center frame: The interacting
362: case at a temperature $T=20\,$K. We show normal state (solid line)
363: and superconducting state (dashed line) results. Bottom frame: The
364: temperature influence on the normal state $n_{\bf k}$ for
365: $T=20\,$K (solid line) and $T=150\,$K (dashed line). }
366:   \label{fig:1}
367: \end{figure}
368: directions in the CuO$_2$ Brillouin zone. In all cases Model A of
369: Table~\ref{tab:1} is used for the
370: electronic dispersion \eqref{eq:3} with $\mu^\ast$ adjusted to the
371: required filling which is is defined as
372: \begin{equation}
373:   \label{eq:9}
374:   \langle n\rangle = \frac{1}{2}-\sum\limits_{\bf k}\sum\limits_{n\ge 0}
375:   \frac{\epsilon_{\bf k}+\xi({\bf k},i\omega_n)}
376:   {\tilde{\omega}^2({\bf k},i\omega_n)+\left[\epsilon_{\bf k}+
377:   \xi({\bf k},i\omega_n)\right]^2+\phi^2({\bf k},i\omega_n)},
378: \end{equation}
379: and the charge carrier spin fluctuation strength
380: $g^2\chi_{\bf Q}$ is adjusted to get a $T_c = 90\,$K for the
381: superconducting state. In the top frame of Fig.~\ref{fig:1} we show
382: $n_{\bf k}$ in the non interacting case as we go from $\Gamma$ to $X$
383: and from $X$ to $M$ in the Brillouin zone with the Fermi surface defined
384: as the value of {\bf k} at which $n_{\bf k}$ jumps from one to zero.
385: It is obvious from Fig.~\ref{fig:2}a (dashed line) that
386: the path $\overline{XM}$ crosses the Fermi surface. A second
387: crossing of the Fermi surface can be observed along the path from $M$
388: to $\Gamma$.
389: 
390: The center frame of Fig.~\ref{fig:1} shows $n_{\bf k}$
391: when interactions are taken into account. (The corresponding Fermi surface
392: is shown as the solid line in Fig.~\ref{fig:2}a.) We see a drastic
393: difference in the value of $n_{\bf k}$ which is now of the order
394: 0.9 at the $\Gamma$ point indicating that the effect of the
395: interaction is very significant even in the center of the Brillouin zone.
396: Also, $n_{\bf k}$ is of the order 0.1 outside the non interacting
397: Fermi surface where $n_{\bf k}\approx 0$ for the non interacting case.
398: The solid line applies to the normal state at $T=20\,$K and the dashed
399: line to the superconducting state at the same temperature. On comparing
400: these two cases we see that the transition to the superconducting state
401: depletes even further the non interacting sea in the sense that it further
402: reduces $n_{\bf k}$ at ${\bf k}$s below the Fermi surface and
403: correspondingly increases the probability of occupation of states right
404: above the Fermi surface beyond the effect of interactions in the
405: normal state. It is to
406: be noted, however, that the onset of superconductivity results in rather
407: modest changes in $n_{\bf k}$ as compared to the difference between
408: interactions and no interactions in the normal state.
409: 
410: The bottom frame of Fig.~\ref{fig:1} gives results in the normal state
411: but compares two temperatures, namely $T=20\,$K (solid line) and
412: $T=150\,$K (dashed line). Comparison with the middle frame shows that
413: increasing the temperature has roughly the same qualitative effect
414: on $n_{\bf k}$ as does the transition to the superconducting state. In
415: both cases, the KE given by Eq.~\eqref{eq:7} increases because the states
416: of lower $\epsilon_{\bf k}$ get depleted while states with higher
417: $\epsilon_{\bf k}$ are occupied with increasing probability. This will
418: also hold for the OS according to Eq.~\eqref{eq:6} which will now
419: depend on interactions and on temperature.
420: 
421: \section{Results for the optical sum in the normal state}
422: \label{sec:3}
423: 
424: In Fig.~\ref{fig:4} we show results for the
425: \begin{figure}[tp]
426: \vspace*{-5mm}
427:   \includegraphics[width=9cm]{vmFig3.eps}
428:   \caption{Optical sum $I_\sigma$ and kinetic energy, $-I_{\rm KE}/2$,
429: as a function of $T^2$. Solid circles and squares are for the non
430: interacting case while solid up-triangles and solid down-triangles
431: include interactions.
432: The top frame applies to
433: Model A of Table~\protect{\ref{tab:1}} and $\omega_{SF}=82\,$meV was
434: used. Here, the interacting and non interacting cases show similar
435: temperature dependencies. The bottom frame is for
436: Model B of Table~\protect{\ref{tab:1}} with an MMP-model
437: $\omega_{SF}=10\,$meV. Note the difference in temperature dependence
438: between interacting and non interacting case.
439: }
440:   \label{fig:4}
441: \end{figure}
442: optical sum $(I_\sigma)$ Eq.~\eqref{eq:6} and compare with the kinetic
443: energy $(I_{\rm KE})$, Eq.~\eqref{eq:7}. The top frame is based on
444: Model A and the bottom frame on Model B of Table~\ref{tab:1}.
445: The solid squares and circles are $-I_{\rm KE}/2$
446: and $I_\sigma$ respectively in the free tight binding case, i.e.:
447: no interactions, plotted as a function of the square of the temperature
448: for the normal state. (A $128\times 128$ sampling of the
449: ${\bf k}$-space, $ak_x, ak_y\in [0,\pi]$, was used but going to
450: a $256\times 256$ sampling did not influence the results.)
451: In both models variation with $T$ over the range $0$ to
452: $200\,$K is small (of order 1\% for Model A and 2\% for Model B)
453: as was also found in the work of
454: Molegraaf {\it et al.}\cite{r26} Also, the two integrals
455: $(I_\sigma$, $-I_{\rm KE}/2)$ track each
456: other closely even though they are not equal in magnitude.
457: (They would be equal for $t'=0$. We tried other Fermi surfaces, even
458: one with perfect nesting, i.e.: $t'=0$ and $\langle n\rangle = 0.5$,
459: and found no qualitative changes.) These results are for
460: comparison with results indicated by up/down-triangles
461: which include interactions in the NAFFL model.
462: The magnitude of both, the optical
463: sum integral and the kinetic energy has been changed considerably by
464: the interactions although the order remains the same, i.e.: $-I_{\rm KE}/2$
465: is greater than $I_\sigma$ and, again, they track each other.
466: More importantly, for the discussion here,
467: the temperature variation has been changed. Both integrals now
468: show variation of the order 8 to $9\%$. Also the dependence on
469: $T^2$ is not linear at small values of $T^2$. Model A shows similar
470: behavior for small values of $\omega_{SF}$.
471: It is clear that any estimate
472: based on the independent particle tight binding model is unreliable.
473: It is, however,
474: possible to chose specific parameters in the MMP-model which show
475: variations in the interacting case that are much closer to the non
476: interacting case. This is illustrated in the top frame of Fig.~\ref{fig:4}.
477: Here we used Model A of Table~\ref{tab:1}.
478: Again, results with and without interaction
479: are compared and both show little temperature variation. To get
480: this we used $\omega_{SF}=82\,$meV in our MMP form of Eq.~\eqref{eq:4}
481: without a change in the magnetic
482: coherence length. Both results, with and without interaction, show
483: little variation with temperature.
484: What this shows is that the magnitude as well as
485: the temperature variation of KE and of OS depends significantly on the
486: parameters used to characterize their electronic structure,
487: particularly the spin susceptibility.
488: 
489: We have done additional calculations for Model A with
490: $\omega_{SF} = 40,\,20,$ and $10\,$meV.
491: In all cases the change in KE due to interactions at $T=0$
492: increases with decreasing values of $\omega_{SF}$. In particular,
493: it changes at $T=0$ by 5.3\% when compared with the non interacting
494: case, for $\omega_{SF}=82\,$meV and by 15.8\% for $\omega_{SF}=10\,$meV.
495: The corresponding temperature changes from $T=0$ to $T=200\,$K are
496: roughly a factor of 5 smaller, more precisely, they are
497: 0.7\% and 3.4\% respectively. 
498: Thus, a change in KE at $T=0$ due
499: to interactions also implies a corresponding change in temperature
500: dependence with both changes tracking each other.
501: For Model B the change in KE due to interactions is 34\% for
502: $\omega_{SF} = 10\,$meV with a 8.7\% increase in KE from $T=0$
503: to $T=200\,$K. These variations are about a factor of two larger
504: than for the equivalent case of Model A with a comparable value
505: of $\omega_{SF}$. Despite the fact that the two models
506: represent very different band structures $I_\sigma$ and
507: $-I_{KE}/2$ show essentially the same qualitative features in their
508: temperature and $\omega_{SF}$ dependence. However, the quantitative
509: differences are important.
510: 
511: \section{Results for the optical sum in the superconducting state}
512: \label{sec:4}
513: 
514: Results for the superconducting state are illustrated in Fig.~\ref{fig:5}
515: \begin{figure}[tp]
516:   \includegraphics[width=9cm]{vmFig4.eps}
517:   \caption{Comparison of normal and superconducting state for the
518: optical sum and the kinetic energy. The top frame applies to
519: Model A of Table~\ref{tab:1} and $\omega_{SF}=82\,$meV and the
520: bottom frame is for the band structure Model B of Table~\ref{tab:1} and
521: $\omega_{SF} = 10\,$meV.
522: }
523:   \label{fig:5}
524: \end{figure}
525: which has two frames. The top frame applies to the band structure
526: Model A of Table~\ref{tab:1} and is for $\omega_{SF}=82\,$meV
527: as in the top frame of Fig.~\ref{fig:4}. We have also included
528: 2\% impurities in the
529: unitary limit but this serves mainly to illustrate that impurities
530: introduce no qualitative differences into our results. We see that,
531: as we expect, superconductivity reduces the optical integral
532: (open triangles) as compared with its normal state (solid
533: triangles) value at the same temperature.
534: This reduction is small. For the top frame which shows the least
535: temperature dependence, the KE integral shows a reduction of about
536: 0.25\% below its normal state value which can be compared with the
537: results shown in the bottom frame of Fig.~6 of Ref.~\onlinecite{r29}
538: where the difference is 0.2\% in their BCS calculations. On the
539: other hand, in the bottom frame of our Fig.~\ref{fig:5} for Model
540: B the reduction is about 0.8\% (four times larger). This shows
541: that the Eliashberg results depend on band structure as well as on
542: the details of the interactions involved, in particular on the
543: value of $\omega_{SF}$. In the BCS limit the increase in KE
544: normalized to the absolute value of the condensation energy is
545: given by the formula $\left[\ln\left(\frac{\omega_D}{T_c}\right)%
546: -0.38\right]$ for both $s$- and $d$-wave superconductors. Here
547: $\omega_D$ is the Debye energy. This shows a strong dependence on
548: $\omega_D/T_c$. The formula itself, however, is valid only for
549: $\omega_D/T_c\gg 1$ and cannot be used to understand our Eliashberg
550: results. The NAFFL model includes interactions which,
551: as we have seen, change importantly the probability of occupation
552: $n_{\bf k}$ and consequently the optical integral as well as the
553: kinetic energy. For the parameters
554: of Model B and $\omega_{SF} = 10\,$meV we find
555: that $I_\sigma$ and $-I_{\rm KE}/2$ can keep increasing with
556: decreasing temperature in the superconducting state
557: (bottom frame of Fig.~\ref{fig:5}). The open squares and
558: triangles (superconducting state) are below their solid
559: counterparts (normal state) but still keep growing as the temperatures
560: is reduced. This does not indicate an exotic mechanism
561: but comes directly from a
562: generalization of Eliashberg theory that includes anisotropy in the band
563: structure and, more importantly, the interaction due to coupling to
564: the spin fluctuations.
565: 
566: In Fig.~\ref{fig:6} we show additional results where the OS is seen to
567: \begin{figure}[tp]
568:   \includegraphics[width=9cm]{vmFig5.eps}
569:   \caption{The optical sum as a function of the square of the temperature
570: for the band structure Model A of Table~\protect{\ref{tab:1}}
571: with different values of $\omega_{SF}$. Also in
572: one case a low frequency cutoff is applied to the spin susceptibility.
573: Note the significance of the $T^2$-variation on the value of
574: $\omega_{SF}$. The solid line indicates experimental normal
575: state results reported by Molegraaf {\it et al.}\protect{\cite{r26}}
576: }
577:   \label{fig:6}
578: \end{figure}
579: increase even more rapidly, with reduced temperature below the onset
580: of superconductivity than it does in the normal state above $T_c$.
581: What is shown is $\rho_L$ which is $I_\sigma$ or $-I_{\rm KE}/2$
582: scaled to agree with experiment as discussed below; either,
583: $I_\sigma$ or $-I_{\rm KE}/2$, will do since these differ
584: mainly by a different scaling factor. Only
585: the OS is considered but the KE integral follows the same trend and
586: therefore the optical measurement can again be used to get information on
587: KE and its variation with temperature.
588: 
589: While in obtaining Fig.~\ref{fig:6} we applied Model A which
590: was used by van der Marel {\it et al.}\cite{r28} to describe
591: their optimally doped and underdoped BSCCO samples,
592: we now vary, in contrast to the top frame of Fig.~\ref{fig:5},
593: the value of $\omega_{SF}$ used in the MMP-model for
594: the spin susceptibility Eq.~\eqref{eq:4}. Results are presented
595: for $\omega_{SF}=20\,$meV (dashed line), $10\,$meV (dotted line),
596: $13\,$meV (solid squares) for the normal state, and
597: solid triangles for the
598: superconducting state. Also shown as the thick solid line are the
599: experimental results of Ref.~\onlinecite{r26} for their optimally
600: doped sample. We have scaled our
601: theoretical results to agree with experiment at $T=120\,$K. We first
602: note that varying $\omega_{SF}$ in the normal state can strongly
603: influence the temperature dependence obtained for $\rho_L$ (in meV).
604:  The value
605: of $\omega_{SF}=13\,$meV was chosen from a best fit in the region
606: $120\,{\rm K}\le T\le 200\,$K. The scaling factor required to get
607: agreement with this data is approximately 2. (When interactions are
608: neglected, as in Ref.~\onlinecite{r28}, the scaling factor is 
609: approximately 1.5.)
610: To reduce this
611: discrepancy, the value of $t$ would need to be increased but this
612: would also decrease the sensitivity of $I_\sigma(T)$ to temperature
613: variations and, thus, $\omega_{SF}$ would need to be adjusted as well.
614: To prove this we employ results of local density
615: approximation calculations by Markievicz {\it et al.}\cite{r42}
616: which suggest significantly bigger values for $t$ in BSCCO with
617: a Fermi surface which is little different from the one presented
618: in Fig.~\ref{fig:1}a. Using the dispersion relation, Eq.~(3),
619: and the parameter values of Table I of Ref.~\onlinecite{r42}
620: we find $I_\sigma(T=0) = 267.34\,$meV for the non-interacting system,
621: well above the value of $\rho_L(T=0) \approx 171.53\,$meV as has been
622: extrapolated from the normal state experimental data of
623: Ref.~\onlinecite{r26}. In order to reproduce the experimentally
624: observed temperature dependence $\rho_L(T)/\rho_L(T=0)$ we have
625: to introduce interactions and a value $\omega_{SF} = 8\,$meV is
626: found to give excellent agreement of $I_\sigma(T)/I_\sigma(T=0)$
627: with experiment. In this case $I_\sigma(T=0) = 230.34\,$meV, still
628: well above the experimental value and a down-scaling of
629: $I_\sigma(T)$ by a factor of 0.745 is required to achieve agreement
630: with experiment. Ultimately, a dispersion relation somewhere
631: between Model A and the one reported by Markievicz {\it et al.}\cite{r42}
632: and an $\omega_{SF}$ between $8$ and $13\,$meV will result in
633: a $I_\sigma(T)$ from theory which agrees with the experimental
634: $\rho_L(T)$ without scaling.
635: However, our main aim is not to treat a specific case but to
636: understand better the role interactions between the charge carriers
637: can play in the OS. Interactions introduce a new energy scale into
638: the problem, namely $\omega_{SF}$ for the NAFFL model. This
639: energy scale is additional to the chemical potential or the hopping
640: parameter $t$. With the values of the microscopic
641: parameters associated with the NAFFL model just described,
642: we proceed to compute
643: the OS for temperatures at and below $T_c$. The solid squares give
644: the continuation of the normal state curve and are presented for
645: comparison with the solid triangles which are the equivalent results
646: in the superconducting state. Again, superconducting state results
647: fall below the normal state ones but they are seen to, nevertheless,
648: increase with decreasing $T$. This occurs even if an Eliashberg
649: formulation is
650: used which represents a generalization of BCS theory and, in that
651: sense, is not exotic. The mechanism is the exchange of
652: antiferromagnetic spin fluctuations.
653: On comparing the top frame of Fig.~\ref{fig:5} with Fig.~\ref{fig:6},
654: we note that whether or not the superconducting state results keep increasing
655: with decreasing temperature is, for a given band structure, governed
656: by the value of $\omega_{SF}$.
657: 
658: In Fig.~\ref{fig:6} we show additional results, open squares for a
659: model normal state and open triangles for the superconducting
660: state. Now, there is a further dramatic increase in the OS both in the
661: normal and the superconducting state
662: as compared to its value in the normal state at $T=T_c=90\,$K. A
663: detailed explanation of how these results were arrived at is required.
664: In their analysis of optical data Carbotte {\it et al.}\cite{r34}
665: found that the spin fluctuations themselves are modified when the
666: superconducting state sets in. To carry out their analysis these
667: authors used a simplified version of our Eliashberg Eqs.~\eqref{eq:1}
668: which follows when the sum over {\bf k} is changed to an energy
669: integral as well as an angular average and the energy integral is done
670: analytically in a constant density of states approximation for an
671: infinite band model with interactions pinned to the Fermi surface.
672: Here we have been more realistic but what is important for us in
673: the work of Ref.~\onlinecite{r34} is that they find that the spin
674: fluctuation spectrum is gaped at low energies, or at
675: the very least loses intensity
676: and a spin resonance or peak forms at higher energy. This readjustment
677: in the spin susceptibility is not unexpected
678: and is a characteristic that should
679: be seen in any electronic mechanism for superconductivity.\cite{r37,%
680: r38,r39,r40,r43a,r41,r44,r45,r46,r47,r48}
681: Details are not important for the present discussion beyond the fact
682: that some adjustment of the spin susceptibility
683: $\Im{\rm m}\chi({\bf q},\omega)$ at small $\omega$ is expected,
684: which weakens the inelastic scattering.
685: Here we simply use the same low $\omega$ $[\omega_c(T)]$ cutoff 
686: applied to Eq.~\eqref{eq:4} which was determined by
687: E.~Schachinger {\it et al.},\cite{r43a} through consideration of
688: microwave data. Another approach would be to calculate the low
689: energy gaping of the spin susceptibility
690: from first principles but this would go beyond the scope of this
691: work and would introduce additional uncertainties.
692: The temperature dependence of $\omega_c(T)$
693: follows the temperature dependence of the superconducting gap
694: with a maximum value of $24\,$meV.
695: Application of this cutoff in
696: otherwise standard Eliashberg calculations based on our
697: Eqs.~\eqref{eq:1} yield the open triangles (superconducting state)
698: and open squares (normal state) of Fig.~\ref{fig:6}. The physics
699: underlying these curves has been made clear in a simple model
700: recently studied by Knigavko {\it et al.}\cite{r30} These authors
701: studied a model in which the charge carriers are coupled to a single
702: Einstein mode of unspecified origin. What they found was that stiffening
703: of this mode decreases the kinetic energy and hence increases the OS.
704: This is precisely the same mechanism that is operative in Fig.~\ref{fig:6}.
705: By applying a low frequency cutoff to the spin fluctuations in our
706: MMP-model we are decreasing the KE. This decrease in KE, present in the
707: underlying normal state below $T_c$, compensates for the increase in
708: KE intrinsic to the superconducting transition which results from the opening
709: of the superconducting gap. We note, however, that in our formulation
710: the OS, at any given temperature, is always below (although not very
711: much) its normal state value at this same temperature calculated with the
712: spectrum with a low frequency cutoff (open triangles). But this cutoff
713: is only operative in the superconducting state and is responsible for
714: making the open triangles fall above the solid squares.
715: The kinetic energy in the superconducting state with low
716: frequency cutoff is now less than the normal state kinetic energy
717: without cutoff.
718: Including the feedback mechanism of the formation of the
719: superconducting state on the spin susceptibility itself has the
720: net effect, at zero temperature (where it is largest),
721: of changing the sign of the KE
722: contribution to the condensation energy from that in BCS.
723: 
724: \section{Conclusion}
725: \label{sec:5}
726: 
727: We have used the Nearly Antiferromagnetic Fermi Liquid model to study
728: the effect of interactions on the optical sum and on the kinetic energy
729: in tight binding bands. Comparison of normal state results with
730: equivalent results when interactions are neglected showed that
731: temperature variations can be strongly affected by details of the
732: microscopic parameters involved in the spin fluctuation exchange
733: mechanism. Behaviors are possible which
734: can be quite different from the non interacting independent
735: particle model. Comparison with normal state
736: experimental data proves that the tight binding model of non
737: interacting particles is certainly not adequate to describe
738: properly the temperature dependence of the optical sum. (This
739: has also been observed by Benfatto {\it et al.}\cite{r42b}) Taking
740: into account interactions between the charge carriers makes the
741: tight binding model a viable model for the analysis of the
742: temperature dependence of the normal state optical sum. This
743: was demonstrated for the particular case of BSCCO and a particle
744: interaction modeled on the NAFFL. Other models,
745: like the one presented by Toschi {\it et al.}\cite{r42c} are also
746: capable to reproduce the temperature dependence
747: $I_\sigma(T)/I_\sigma(T=0)$ but they lack agreement with the
748: value of the optical sum at zero temperature.
749: 
750: When superconductivity is considered within an
751: Eliashberg formalism, the superconducting gap has $d$-wave
752: symmetry as a function of momentum in the two dimensional
753: Brillouin zone. The optical sum is found to decrease
754: with decreasing temperature for some range of parameters characterizing
755: the spin susceptibility but can also increase. This increase
756: cannot necessarily be interpreted as kinetic energy driven superconductivity.
757: In fact, in all cases considered, the optical sum is always lower,
758: at a given temperature in the superconducting state, than it is
759: in the corresponding normal state but, in some cases not by much.
760: Correspondingly, the kinetic energy is increased in the
761: superconducting state. What makes the optical sum and KE integral
762: continue to go up (in some cases) with decreasing temperature
763: is the fact that the interactions themselves introduce a
764: temperature dependence in the underlying normal state.
765: 
766: The results just described were obtained for a fixed
767: (i.e.: temperature independent) value
768: of the spin susceptibility. If
769: we consider the possibility that the spin fluctuation spectrum may itself
770: be modified\cite{r33,r34,r35,r44,r45,r46,r47,r48} by the onset of
771: superconductivity
772: and by temperature, even larger increases in the optical sum with
773: decreasing temperature can be obtained.
774: It is widely recognized that a generic feature of an electronic
775: mechanism of superconductivity is the possible gaping of the
776: excitation spectrum itself at small energies due to the opening
777: of the superconducting gap. This leads to a weakening of
778: interactions at small $\omega$ and to the so called collapse of
779: the inelastic scattering rate\cite{r44,r45,r46,r47,r48} which manifests
780: itself as a large peak in the temperature dependence of the
781: microwave conductivity. The weakening of the interaction
782: in the superconducting state through the opening of a low energy
783: gap in the spin susceptibility
784: corresponds to a decrease in KE in the superconducting state which can,
785: in the case considered, more than compensate for the intrinsic
786: increase that accompanies the formation of Cooper pairs
787: and, consequently, the OS rises with a larger slope in the
788: superconducting state than in the normal state just above $T_c$.
789: At $T_c$ there is no low energy gaping of the spin susceptibility and,
790: therefore, the mechanism for
791: KE reduction just described is not operative. In this sense
792: our model does not describe KE driven superconductivity.
793: 
794: \section*{Acknowledgment}
795:  
796: Research supported by the Natural Sciences and Engineering
797: Research Council of Canada (NSERC) and by the Canadian
798: Institute for Advanced Research (CIAR).
799: 
800: \begin{thebibliography}{99}
801: \bibitem{r1} T.~Timusk and B.~Statt, Rep.~Prog.~Phys. {\bf 62}, 61
802: (1999) and references therein.
803: \bibitem{r2} H.~Ding, T.~Yokoya, J.C.~Campuzano, T.~Takahashi,
804: M.~Randeria, M.R.~Norman, T.~Mochiku, K.~Kadowaki, and
805: J.~Giapintzakis, Nature (London) {\bf 382}, 51 (1996).
806: \bibitem{r3} A.G.~Loeser, Z.-X.~Shen, D.S.~Dessau, D.S.~Marshall,
807: C.H.~Park, P.~Fournier, and A.~Kapitulnik, Science {\bf 273}, 325 (1996).
808: \bibitem{r4} J.C.~Campuzano, M.R.~Norman, and M.~Randeria, in
809: {\it The Physics of Superconductivity: Conventional and Unconventional},
810: edited by K.H.~Bennemann and J.B.~Ketterson (Springer, Berlin, 2003)
811: Vol. 2, p. 167.
812: \bibitem{r5} Ch.~Renner, B.~Revaz, J.-Y.~Genoud, K.~Kadowaki, and
813: O.~Fischer, \prl {\bf 80}, 149 (1998).
814: \bibitem{r6} V.J.~Emery and S.A.~Kivelson, Nature (London) {\bf 374},
815: 134 (1995).
816: \bibitem{r7} E.~Carlson, V.J.~Emery, S.A.~Kivelson, and D.~Orgard, in
817: {\it The Physics of Superconductivity: Conventional and Unconventional},
818: edited by K.H.~Bennemann and J.B.~Ketterson (Springer, Berlin, 2003)
819: Vol. 2, p. 275.
820: \bibitem{r23a} T.~Eckl, W.~Hanke, and E.~Arrigoni, \prb {\bf 68},
821: 014505 (2003).
822: \bibitem{r8} Q.~Chen, I.~Kosztin, B.~Jank\'{o}, and K.~Levin, \prl
823: {\bf 81}, 4708 (1998).
824: \bibitem{r9} S.~Chakravarty, R.B.~Laughlin, D.K.~Morr, and
825: C.~Nayak, \prb {\bf 63}, 094503 (2001).
826: \bibitem{r10} B.~D\`{o}ra, A.~Virosztek, and K.~Maki, \prb
827: {\bf 65}, 155119 (2002), K.~Maki, B.~D\`{o}ra, M.~Kartsovnik,
828: A.~Virosztek, B.~Korin-Hamzic, and M.~Basletic, \prl {\bf 90},
829: 256402 (2003).
830: \bibitem{r11} S.~Chakravarty, H.-Y.~Kee, and C.~Nayak, Int.\ J.\ Mod.\
831: Phys.\ B {\bf 15}, 2901 (2001).
832: \bibitem{r12} J.-X.~Zhu, W.~Kim, C.S.~Ting, and J.P.~Carbotte,
833: \prl {\bf 87}, 197001 (2001).
834: \bibitem{r13} X.~Yang and C.~Nayak, \prb {\bf 65}, 064523 (2002).
835: \bibitem{r14} P.W.~Anderson, in {\it The Theory of Superconductivity
836: in the High-$T_c$ Cuprates\/} (Princeton University Press, Princeton,
837: 1998).
838: \bibitem{r15} J.E.~Hirsch, Physica C {\bf 199}, 305 (1992); Physica C
839: {\bf 201}, 347 (1992).
840: \bibitem{r16} J.E.~Hirsch and F.~Marsiglio, \prb {\bf 62}, 15131 (2000).
841: \bibitem{r17} J.E.~Hirsch, Science {\bf 295}, 2226 (2002).
842: \bibitem{r18} Wonkee Kim and J.P.~Carbotte, \prb {\bf 64}, 104501 (2001);
843: \prb {\bf 61}, R11886 (2000); \prb {\bf 63}, 140505(R) (2001).
844: \bibitem{r19} S.~Chakravarty, Euro. Phys. J. B {\bf 5}, 337 (1998).
845: \bibitem{r20} M.R.~Norman and C.~P\'epin, \prb {\bf 66}, 100506(R) (2002).
846: \bibitem{r21} M.R.~Norman and C.~P\'epin, Rep.\ Prog.\ Phys. {\bf 66},
847: 1547 (2003).
848: \bibitem{r22} L.~Benfatto, S.G.~Sharapov, and H.~Beck, Eur. Phys. J.
849: {\bf 39}, 469 (2004).
850: \bibitem{r23} A.E.~Karakozov, E.G.~Maksimov, and O.V.~Dologov,
851: Solid State Comm. {\bf 124}, 119 (2002).
852: \bibitem{r24} D.N.~Basov, S.I.~Woods, A.S.~Katz, E.J.~Singley, R.C.~Dynes,
853: M.~Xu, D.G.~Hinks, C.C.~Homes, and M.~Strongin, Science {\bf 283}, 49 (1999).
854: \bibitem{r25} A.S.~Katz, S.I.~Woods, E.J.~Singley, T.W.~Li,
855: M.~Xu, D.G.~Hinks, R.C.~Dynes, and D.N.~Basov, \prb {\bf 61}, 5930 (2000).
856: \bibitem{r26} H.J.A.~Molegraaf, C.~Presura, D.~van der Marel, P.H.~Kes,
857: and M.~Li, Science {\bf 295}, 22 (2002).
858: \bibitem{r27} A.F.~Santander-Syro, R.P.M.S.~Lobo, W.~Bontemps,
859: Z.~Konstantinovic, Z.Z.~Li, and H.~ Raffy, Europhys. Lett. {\bf 62},
860: 568 (2003).
861: \bibitem{r28} D.~van der Marel, H.J.A.~Molegraaf, C.~Presura, and
862: L.~Santoso, in {\it Concepts in Electron Correlations}, edited by
863: A.~Hewson and V.~Zlatic (Kluwer, 2003) p. 7.
864: \bibitem{r42a}M.~Ortolani, P.~Calvani, and S.~Lupi, \prl {\bf 94},
865: 067002 (2005).
866: \bibitem{r42b}B.L.~Benefatto, S.~Sharapov, N.~Andrenacci, and
867: H.~Beck, \prb {\bf 71}, 104511 (2005).
868: \bibitem{r42c}A.~Toschi, M.~Capone, M.~Ortoloani, S.~Lupi, P.~Calvani,
869: and C.~Castellani, cond-mat/0502528 (unpublished).
870: \bibitem{r42d}A.J.~Millis, A.~Zimmers, R.P.S.M.~Lobo, and N.~Bontemps,
871: cond-mat/0411172 (unpublished).
872: \bibitem{r29} A.V.~Boris, N.N.~Kovalera, O.V.~Dolgov, T.~Holden,
873: C.T.~Jin, B.~Keimer, and C.~Bernhard, Science {\bf 304}, 708 (2004).
874: \bibitem{r30} A.~Knigavko, J.P.~Carbotte, and F.~Marsiglio,
875: \prb {\bf 70}, 224501 (2004).
876: \bibitem{r31} A.V.~Chubukov, D.~Pines, and J.~Schmalian, in
877: {\it The Physics of Superconductivity: Conventional and Unconventional},
878: edited by K.H.~Bennemann and J.B.~Ketterson (Springer-Verlag, Berlin, 2003),
879: Vol. 2, p. 495.
880: \bibitem{r32}A.J.~Millis, H.~Monien, and D.~Pines,
881: \prb {\bf 42}, 167 (1990).
882: \bibitem{r33} E.~Schachinger, J.P.~Carbotte, and D.N.~Basov,
883: Europhys.~Lett. {\bf 54}, 380 (2001).
884: \bibitem{r34}J.P.~Carbotte, E.~Schachinger, and D.N.~Basov,
885: Nature (London) {\bf 401}, 354 (1999).
886: \bibitem{r35} E.~Schachinger and J.P.~Carbotte, \prb {\bf 62}, 9054 (2000).
887: \bibitem{r35a}D.~Branch and J.P.~Carbotte, Can.\ J.\ Phys. {\bf 77},
888: 531 (1999); J.\ Superconductivity {\bf 12}, 667 (1999);
889: {\bf 13}, 535 (2000).
890: \bibitem{r35b}P.~Monthoux and D.~Pines, \prb {\bf 47}, 6069 (1993);
891: {\bf 49}, 4261 (1994).
892: \bibitem{r35c}A.~Abanov, A.V.Chubukov, and J.~Schmalian,
893: Adv. in Physics {\bf 52}, 119 (2003).
894: \bibitem{r36} E.~Schachinger, J.J.~Tu, and J.P.~Carbotte. \prb {\bf 67},
895: 214508 (2003).
896: \bibitem{r37} E.~Schachinger and J.P.~Carbotte, \prb {\bf 64}, 094501
897: (2001); \prb {\bf 65}, 064514 (2002).
898: \bibitem{r38}A.~Abanov, A.V.~Chubukov, and J.~Schmalian, \prb {\bf 63},
899: 180510 (2001); J.\ Electron Spectrosc. {\bf 117}, 129 (2001).
900: \bibitem{r39} J.P.~Carbotte and E.~Schachinger, \prb {\bf 69}, 224501
901: (2004).
902: \bibitem{r40} E.~Schachinger and J.P.~Carbotte, 
903: in: {\it Models and Methods of High-TC Superconductivity: some Frontal
904: Aspects}, edited by J.K.~Srivastava and S.M.~Rao,
905: (Nova Science, Hauppauge, NY, 2003),  Vol. II, p. 73.
906: \bibitem{r43a}E.~Schachinger, J.P.~Carbotte, and F.~Marsiglio,
907: \prb {\bf 56}, 2738 (1997).
908: \bibitem{r41} P.~Monthoux, A.V.~Balatsky, and D.~Pines, \prl {\bf 67},
909: 3448 (1991); \prb {\bf 46}, 14 803 (1992).
910: \bibitem{r42}R.S.~Markiewicz, S.~Sahrakorpi, M.~Lindroos, Hsin Lin,
911: and A.~Bansil, cond-mat/0503064 (unpublished) and references therein.
912: \bibitem{r44}M.C.~Nuss, P.M.~Mankiewich, N.L. O'Malley, E.H.~Westwick,
913: and P.B.~Littlewood, \prl {\bf 66}, 3305 (1991).
914: \bibitem{r45}D.B.~Romero, C.D.~Porter, D.B.~Tanner, L.~Forro, D.~Mandrus,
915: L.~Mihaly, C.C.~Carr, and G.P.~Williams, \prl {\bf 68}, 1590 (1992).
916: \bibitem{r46}D.A.~Bonn, P.~Dosanjh, R.~Liang, and W.N.~Hardy, \prl
917: {\bf 68}, 2390 (1992).
918: \bibitem{r47}A.~Hosseini, R.~Harris, S.~Kamal, P.~Dosanjh, J.~Preston,
919: R.~Liang, W.N.~Hardy, and D.A.~Bonn, \prb {\bf 60}, 1349 (1999).
920: \bibitem{r48}E.J.~Nicol and J.P.~Carbotte, \prb {\bf 44}, R7741 (1991).
921: \end{thebibliography}
922: \end{document}
923: 
924: %%% Local Variables: 
925: %%% mode: latex
926: %%% TeX-master: t
927: %%% End: 
928: