1: % INJ
2:
3: %\documentclass[epj,referee]{svjour}
4: \documentclass[epj]{svjour}
5:
6: \usepackage{epsfig}
7: \usepackage{xspace}
8: \usepackage{amsmath}
9: \usepackage{amssymb}
10:
11: \let\onlinecite\cite
12:
13: %\newcommand{\bk}{{\bf k}}
14: \newcommand{\bk}{\vec{k}}
15: \newcommand{\kB}{{k_{\rm B}}}
16: \newcommand{\sgn}{\mathop{\rm sgn}\nolimits}
17: \newcommand{\BSCCO}{Bi$_2$\-Sr$_2$\-Ca\-Cu$_2$\-O$_{8+\delta}$\xspace}%{} }
18: \newcommand{\XCOC}{X$_2$\-Cu\-O$_2$\-Cl$_2$\xspace}%{} }
19:
20: \begin{document}
21:
22: \title{Extended $d_{x^2 - y^2}$-wave superconductivity}
23: \subtitle{Flat nodes in the gap and the low-temperature asymptotic
24: properties of high-$T_c$ superconductors}
25: \author{Giuseppe G. N. Angilella\inst{1} \and Asle Sudb\o\inst{2} \and
26: Renato Pucci\inst{1}}
27: \institute{Dipartimento di Fisica dell'Universit\`a di Catania
28: and Istituto Nazionale di Fisica della Materia, U. d. R.
29: di Catania,\\
30: 57, Corso Italia, I-95129 Catania, Italy, EU \and
31: Institutt for Fysikk, Norges Teknisk-Naturvitenskapelige
32: Universitet NTNU,\\
33: Sem S\ae landsvei 9, Gl\o shaugen, N-7491 Trondheim, Norway}
34: \date{Received: November 19, 1999 / Revised version: December 20, 1999}
35:
36: \abstract{%
37: Remarkable anisotropic structures have been recently observed
38: in the order parameter $\Delta_\bk$ of the underdoped superconductor
39: \BSCCO.
40: Such findings are strongly suggestive of deviations from a
41: simple $d_{x^2 - y^2}$-wave picture of high-$T_c$
42: superconductivity, \emph{i.e.} $\Delta_\bk \sim \cos k_x - \cos k_y$.
43: In particular, flatter nodes in $\Delta_\bk$ are observed along the
44: $k_x = \pm k_y$ directions in $\bk$-space, than within this simple
45: model for a $d$-wave gap.
46: We argue that nonlinear corrections in the $\bk$-dependence of
47: $\Delta_\bk$ near the nodes introduce new energy scales, which
48: would lead to deviations in the predicted power-law asymptotic
49: behaviour of several measurable quantities, at low or intermediate
50: temperatures.
51: We evaluate such deviations, either analytically or numerically,
52: within the interlayer pair-tunneling model, and within yet another
53: phenomenological model for a $d$-wave order parameter.
54: We find that such deviations are expected to be of different sign in
55: the two cases.
56: Moreover, the doping dependence of the flatness of the gap near the
57: nodes is also attributable to Fermi surface effects, in addition to
58: possible screening effects modifying the in-plane pairing kernel,
59: as recently proposed.
60: %
61: \PACS{%
62: {74.25.-q}{General properties; correlations between physical
63: properties in normal and superconducting states} \and
64: {74.25.Jb}{Electronic structure} \and
65: {74.20.Mn}{Nonconventional mechanisms} \and
66: {74.72.Hs}{Bi-based cuprates}
67: }}
68:
69: \maketitle
70:
71: \section{Introduction}
72:
73: Power laws in the low-temperature asymptotic behaviour of several
74: linear response electronic properties provide
75: complementary evidence for $d$-wave symmetry of the order parameter
76: (OP) $\Delta_\bk$ of high-$T_c$
77: superconductors~\cite{Annett:90,Annett:96} as well as preliminary
78: evidence of `exotic' shapes in the OP of heavy fermion
79: superconductors, such as UPt$_3$~\cite{Sigrist:91}.
80: This has to be contrasted to an ``activated'' behaviour $\propto\exp
81: (-\beta\Delta_{\rm min} )$, appropriate of $s$-wave superconductors,
82: or, in the case of mixed symmetry,
83: of superconductors with a non-vanishing $s$-wave contribution to
84: their OP, where $\Delta_{\rm min} = \min_\bk |\Delta_{\bk} | > 0$.
85: In the case of a non-empty nodal manifold for the superconducting
86: excitation spectrum $E_\bk$, defined as the locus of points in
87: $\bk$-space such that $E_\bk =0$, a large number of quasiparticles
88: can be created near such nodes, thus dominating all the low-temperature
89: electronic properties~\cite{Lee:97}.
90: An exact analysis allows one to relate the exponent of the leading
91: power of the low-$T$ expansion of a given linear response function
92: to the dimension of the Fermi manifold (defined as the locus of
93: states in $\bk$-space with vanishing dispersion relative to the
94: Fermi level in the normal state, $\xi_\bk =0$) and the topological
95: nature of the nodal manifold, \emph{viz.} a collection of points, of line
96: segments, or of surface patches~\cite{Volovik:96,Annett:90}.
97:
98: On the basis of group theoretical arguments, the simplest choice for a
99: $d$-wave gap function on a square lattice is $\Delta_\bk = \Delta
100: g(\bk)$, where $\Delta$ is a $T$-dependent parameter, and
101: \begin{equation}
102: g(\bk) = \frac{1}{2} (\cos k_x - \cos k_y )
103: \label{eq:g}
104: \end{equation}
105: is the first basis function associated with the $d$-wave irreducible
106: representation of the appropriate crystal point group,
107: $C_{4v}$~\cite{Annett:90}.
108: We remark that $g(\bk)$ is generated, together with an extended
109: $s$-wave term proportional to
110: \begin{equation}
111: h(\bk) = \frac{1}{2} (\cos k_x + \cos k_y ),
112: \label{eq:h}
113: \end{equation}
114: by a nearest-neighbour interaction term in real space.
115: Here and in the following we shall measure the wavevectors in units of
116: the appropriate inverse lattice spacings.
117: Proportionality to Eq.~(\ref{eq:g}) allows $\Delta_\bk$ to vanish
118: linearly at a given point along the Fermi line, which for most cuprate
119: superconductors can be modelled by the tight-binding expansion:
120: \begin{equation}
121: \xi_\bk = -2t(\cos k_x + \cos k_y ) + 4t^\prime \cos k_x \cos k_y -
122: \mu = 0,
123: \label{eq:disp}
124: \end{equation}
125: where $t=0.25$~eV, $t^\prime = 0.45 t$ measure nearest and
126: next-nearest neighbour hopping, respectively, and $\mu$ is the
127: chemical potential.
128:
129: On the other hand, increasing experimental evidence above all from
130: angle-resolved photoemission spectroscopy (ARPES) suggests a richer
131: structure in $\bk$-space for the OP of the underdoped high-$T_c$
132: superconductor \BSCCO~\cite{Mesot:99}.
133: In particular, the superconducting gap near the nodal points turns
134: out to be flatter than predicted by the simple assumption
135: $\Delta_\bk \propto g(\bk)$~\cite{Mesot:99}.
136: Such a feature is consistent with the observation of whole ungapped
137: segments of the Fermi line above $T_c$ in the pseudogap regime of
138: underdoped \BSCCO~\cite{Norman:98}, and will of course serve as a
139: constraint for a microscopic understanding of the pairing
140: mechanism.
141:
142: Quite remarkably, qualitatively similar deviations from a
143: $g(\bk)$-like dispersion have been evidenced in the
144: $\bk$-depend\-ence of the antiferromagnetic gap in the related
145: insulating compounds \XCOC (X = Ca, Sr) \cite{Ronning:98}.
146: Such a finding has been interpreted in terms of an interrelation
147: between the antiferromagnetic phase of the parent insulator and the
148: underdoped regime of the intervening superconductor \cite{Zacher:99}.
149:
150: In this paper, we argue that such extended structures in the
151: superconducting OP, interpolating between point and line nodes, can
152: be included in the definition of $\Delta_\bk$ as higher order terms
153: in $g(\bk)$.
154: We shall then look for their signatures in the low-temperature
155: asymptotic electronic properties of the superconducting cuprates,
156: as corrections to the predicted power-law behaviour.
157: In deriving our results analytically, we will specifically consider
158: the interlayer pair-tunneling (ILT) mechanism of high-$T_c$
159: superconductivity~\cite{Chakravarty:93}, which has been shown to
160: accurately reproduce most of the observed gap
161: features~\cite{Angilella:99}.
162:
163: \section{Extended $d$-wave gap within the ILT model}
164:
165: A distinguishing feature of the ILT mechanism, compared to other
166: proposed models of HTSC, is that superconductivity is driven by a
167: gain in kinetic, rather than potential, energy as temperature is
168: lowered below the critical temperature $T_c$.
169: It is assumed that coherent single particle hopping between adjacent
170: CuO$_2$ layers in the cuprates is suppressed by the non-Fermi
171: liquid character of the normal state (\emph{e.g.} due to spin-charge
172: separation), while interlayer coherent tunneling of Cooper pairs is
173: allowed as soon as a superconducting condensate is established.
174: Confined coherence~\cite{Clarke:97} within
175: CuO$_2$ layers in the
176: normal state is indeed largely motivated by the absence of coherent
177: transport along the $c$-axis, whereas a comprehensive theoretical
178: understanding of it is still lacking.
179: However, there is now abundant \emph{experimental} evidence that
180: $c$-axis transport in the normal state indeed is incoherent, while
181: that in the superconducting state may not be~\cite{vanderMarel}.
182: This seems to warrant attention being paid to unconventional models of
183: high-$T_c$ superconductivity based on relieving $c$-axis frustrated
184: kinetic energy.
185: Recent findings~\cite{Moler:98,Tsvetkov:98} suggest,
186: however, that the ILT mechanism alone is not sufficient to account
187: for the large condensation energy $E_c$, as extracted
188: experimentally from measurements of the penetration length $\lambda_c$
189: of several single layered compounds, such as
190: Tl$_2$Ba$_2$CuO$_{6+\delta}$~\cite{Moler:98,Tsvetkov:98},
191: whereas the predictions of the ILT model agrees with the measured
192: value of $E_c$ for
193: La$_{2-x}$Sr$_x$CuO$_4$~\cite{Panagopoulos:97,Panagopoulos:99,Leggett:98,Anderson:98}.
194: It has been pointed out, however, that while considerable experimental
195: effort has
196: been devoted to the determination of $\lambda_c$, extracting $E_c$
197: from existing data on electronic specific heat is by no means
198: straightforward~\cite{Chakravarty:99}.
199: A direct evaluation of $E_c$ from its mean-field expression at $T=0$
200: \cite{Schrieffer:64} would relieve the complications arising from
201: thermal fluctuation effects, inherent in the method of integrating
202: specific heat data, from $T=0$ through $T_c$, recently pointed out
203: in Ref.~\onlinecite{Chakravarty:99}.
204: By utilizing the gap equation, Eq.~(\ref{eq:delta}), within the ILT
205: model, and of the expression relating $\lambda_c$ at $T=0$ to $E_c$
206: \cite{Chakravarty:98}, we find results for $\lambda_c (T=0)$ in
207: Bi2212 which are within factors of order unity from the
208: experimental values, rather than factors of 10 to 20
209: \cite{Angilella:00}.
210: The observed doping dependence of $\lambda_c$ \cite{Panagopoulos:99}
211: is also qualitatively reproduced \cite{Angilella:00}.
212:
213: The emerging scenario suggests therefore that some in-plane effective
214: interaction might co-operate with the ILT mechanism in establishing
215: the superconducting state~\cite{note:int-kin}.
216: One could think of such a mechanism as a seed for the Cooper
217: instability, and the origin of the gap's dominant $d$-wave symmetry.
218: Once Cooper pairs are formed in the appropriate symmetry channel(s)
219: via such in-plane effective interaction, the ILT mechanism would
220: allow the condensate for an additional energy gain, by releasing
221: the constraint of in-plane segregation.
222:
223: Without explicitly specifying the microscopic origin of the in-plane
224: mechanism, we therefore assume the in-plane pairing potential to
225: be given by $V_{\bk\bk^\prime} = V g(\bk) g(\bk^\prime )$ ($V <0$),
226: thus allowing for $d$-wave symmetry of the order parameter.
227: The issue of the competition with other subdominant ($s$-wave)
228: symmetry channels in the presence of ILT has been addressed in
229: Ref.~\onlinecite{Angilella:99}, showing that the $d$-wave
230: contribution wins out at optimal doping and in the underdoped
231: regime.
232: Despite its kinetic nature, ILT can be absorbed in the interacting
233: part of the Hamiltonian as an effective term
234: $T_J (\bk)\delta_{\bk\bk^\prime}$, whose $\bk$-space locality
235: enforces in-plane momentum conservation during a tunneling
236: process~\cite{Chakravarty:93}.
237: Following Ref.~\onlinecite{Chakravarty:93}, we assume $T_J (\bk) =
238: t_\perp^2 (\bk)/t \equiv T_J g^4 (\bk)$, being $t_\perp (\bk)$ the
239: single-particle interlayer hopping amplitude, with $t_\perp (\bk)
240: \propto g^2 (\bk)$, as suggested by ARPES as well as by band
241: structure calculations \cite{Chakravarty:93,Andersen:96}.
242: A standard mean-field diagonalization technique then yields the
243: following expression for the energy
244: gap~\cite{Sudbo:95,Angilella:99}:
245: \begin{equation}
246: \Delta_\bk = \frac{\Delta g(\bk)}{1-T_J (\bk)\chi_\bk} ,
247: \label{eq:delta}
248: \end{equation}
249: where $\chi_\bk = (2E_\bk )^{-1} \tanh (\beta E_\bk /2)$ is the
250: superconducting pair susceptibility, and $E_\bk = (\xi_\bk^2 +
251: |\Delta_\bk |^2 )^{1/2}$ is the upper branch of the superconducting
252: elementary excitation spectrum.
253:
254: Along the Fermi line ($\xi_\bk =0$) at $T=0$, one immediately sees
255: that:
256: \begin{equation}
257: \Delta_\bk = \Delta g(\bk) + \frac{1}{2} T_J g^4 (\bk) \sgn
258: [g(\bk)].
259: \label{eq:deltaFL}
260: \end{equation}
261: Such an expression, together with manifestly fulfilling the
262: requirement of $d$-wave symmetry, also endows the superconducting
263: gap with a richer structure near the nodal points along the $k_x =
264: \pm k_y$ directions.
265: This is probably best seen by considering the Fourier expansions:
266: \begin{subequations}
267: \label{eqs:expansion}
268: \begin{eqnarray}
269: \label{eq:expansion-g}
270: g(\bk) &&= -2\sum_{m=1}^\infty J_{4m-2} (k) \cos[(4m-2)\phi],\\
271: \label{eq:expansion-h}
272: h(\bk) &&= J_0 (k) + 2 \sum_{m=1}^\infty J_{4m} (k) \cos(4m\phi),\\
273: \label{eq:expansion-TJ}
274: T_J (\bk)/T_J &&= \frac{9}{64} + \frac{a_0}{2} +
275: \sum_{m=1}^\infty a_{4m} \cos (4m\phi),
276: \end{eqnarray}
277: \end{subequations}
278: with
279: \begin{eqnarray}
280: a_{4m} &&= \frac{1}{32} J_{4m} (4k) + \frac{1}{2} J_{4m} (2k) +
281: \frac{3}{16} (-1)^m J_{4m} (2k\sqrt{2}) \nonumber \\
282: && - \frac{3}{4} (-1)^m J_{4m} (k\sqrt{2}) -\frac{1}{4} J_{4m}
283: \left( \frac{k}{\sin\phi_0} \right) \cos(4m\phi_0 ). \nonumber \\
284: \end{eqnarray}
285: Here, the generic wavevector $\bk$ is expressed in terms of its
286: modulus $k$ and of the angle $\phi$ formed with the $\Gamma X$
287: direction in the first Brillouin zone (1BZ), $\bk =
288: (k\cos\phi,k\sin\phi)$, $J_\alpha (x)$ are Bessel functions of the
289: first kind and order $\alpha$, and $\tan\phi_0 = \frac{1}{3}$.
290:
291: Eq.~(\ref{eq:deltaFL}) is to be contrasted to the phenomenological fit
292: \begin{equation}
293: \Delta_\bk = \Delta [B\cos(2\phi) +(1-B)\cos(6\phi)]
294: \label{eq:Mesot}
295: \end{equation}
296: proposed in Ref.~\onlinecite{Mesot:99} for $\Delta_\bk$ along the
297: Fermi line:
298: Instead of requiring an in-plane interaction extended to further
299: neighbours, Eq.~(\ref{eq:delta}) endows the superconducting gap with
300: the observed flat structure around the nodes, through the ILT term
301: $T_J (\bk)$.
302: In Fig.~\ref{fig:fit}, we fit Eq.~(\ref{eq:deltaFL}) against Mesot {\em
303: et al.}'s experimental data for one of the underdoped Bi2212
304: samples in Ref.~\cite{Mesot:99}, having $T_c = 75$~K. % (sample UD75K).
305: A remarkable agreement follows already by fixing $\Delta$ so that
306: $|\Delta_{\bf k} |$ reproduces the maximum datum at $\bk = (\pi,\pi)$,
307: whereas $T_J$ is taken to be 0.04~eV \cite{Chakravarty:93}.
308: In particular, besides obtaining an enhanced maximum value of $|\Delta_{\bf k}
309: |$ at $\bk=(\pi,\pi)$, we are thus able to recover the anomalously
310: flat region around the node at $\phi=45^\circ$ in a rather natural way.
311: We note, however, that our fit requires $\Delta\approx T_J /2$ around
312: optimal doping, which will not be without consequences in
313: evaluating other fundamental quantities \cite{Angilella:00}.
314:
315: \begin{figure}
316: \centering
317: \epsfig{width=0.95\columnwidth,angle=-90,file=cc_UD75K_2.ps}
318: \caption{%
319: Fit for $|\Delta_{\bf k} |$ within the ILT model,
320: Eq.~(\protect\ref{eq:deltaFL}) (solid line), and in the case of a
321: simple $d$-wave gap $|\Delta_{\bf k}| = \Delta g({\bf k})$
322: (dashed line), against Mesot {\em et al.}'s ARPES data for
323: underdoped Bi2212 ($T_c = 75$~K, Ref.~\protect\cite{Mesot:99}).}
324: \label{fig:fit}
325: \end{figure}
326:
327: Eq.~(\ref{eq:deltaFL}) already contains the doping dependence of the
328: observed gap anisotropy, although in a hidden way.
329: As pointed out in Ref.~\onlinecite{Angilella:99}, the auxiliary parameter
330: $\Delta$ is to be self-consistently determined by solving the
331: appropriate gap equation.
332: Besides being intrinsically doping dependent, this equation is
333: unconventionally modified by the presence of a $\bk$-local
334: effective interaction, as induced by the ILT mechanism.
335: Moreover, the role of the contribution $\propto T_J g^4 (\bk)$ in
336: Eq.~(\ref{eq:deltaFL}) is strongly influenced by the actual
337: location of the Fermi line, as $g^4 (\bk)$ is sharply peaked at
338: $\bk = (0,\pi)$ (and symmetry related points).
339:
340: Eq.~(\ref{eq:deltaFL}) also facilitates the evaluation of the slope of the
341: superconducting gap $v_\Delta = (1/2) d|\Delta_\bk |/d\phi$ at the
342: nodal point along the Fermi line.
343: Such a quantity is related to the temperature derivative of the
344: superfluid stiffness at $T=0$.
345: In particular, it is seen that the ratio $v_\Delta /\Delta_{\rm max}$
346: decreases with underdoping~\cite{Mesot:99}.
347: From Eq.~(\ref{eq:deltaFL}), one derives that
348: $v_\Delta$ is independent of $T_J$, and that therefore a doping induced
349: change of $v_\Delta$ through $\Delta$ essentially can be traced
350: back to the actual position of the Fermi line, as discussed above,
351: within the ILT model.
352: The ratio $v_\Delta / \Delta_{\rm max}$ will anyway deviate from its
353: value within simple $d$-wave (BCS-like) models, as a function of doping,
354: due to the enhancement of $\Delta_{\rm max}$ induced by ILT.
355:
356: \section{Low-temperature asymptotic behaviour of electronic properties}
357:
358: We now address the issue, whether such extended features of the OP near
359: the nodes, as those described in the previous section, induce
360: deviations in the low or intermediate temperature asymptotic
361: behaviour of linear response electronic properties in the
362: superconducting state.
363: In what follows, we shall limit our discussion to clean
364: superconductors, and neglect impurity effects altogether.
365: Mean-field (BCS or BCS-like) expressions for most linear response
366: electronic properties are available also in the case of
367: anisotropic, \emph{i.e.} non $s$-wave, superconductors.
368: In particular, we have in mind observable quantities such as the
369: superconducting density~\cite{Leggett:75}, the electronic
370: specific heat~\cite{Leggett:75}, the spin
371: susceptibility~\cite{Leggett:75}, the penetration
372: depth~\cite{Scalapino:92}, the thermal
373: conductivity~\cite{Bardeen:59}, and so on.
374: Their expressions basically involve the evaluation of some integral of
375: the kind:
376: \begin{equation}
377: {\cal F}[\beta; \varphi_\bk (\beta)] = \frac{1}{(2\pi)^2} \int d^2 \bk
378: \varphi_\bk (\beta) e^{-\beta E_\bk} ,
379: \label{eq:integral}
380: \end{equation}
381: where $\beta=(k_{\rm B} T)^{-1}$, $\varphi_\bk (\beta)$ is a
382: (dimensional) function of wavevector $\bk$ and temperature, related
383: to the electronic quantity of interest, and the integration is
384: extended to the 1BZ, $\bk\in[-\pi,\pi]\times[-\pi,\pi]$ (see
385: App.~\ref{app:electronic}).
386: In the case of $d$-wave superconductors, $E_\bk$ is allowed to vanish
387: at the intersection between the Fermi line and the nodal lines of
388: the gap function.
389: Around such points, quasiparticles can be created in large numbers.
390: In the limit of low temperatures ($\beta\to\infty$), therefore, the
391: value of the integral in Eq.~(\ref{eq:integral}) is dominated by
392: the contributions from wavevectors $\bk$ close to such point nodes.
393: Around such nodes, it is useful to introduce the new sets of
394: coordinates $(k_1 ,k_2 )$ or $(\epsilon,\theta)$, defined
395: as~\cite{Lee:97}:
396: \begin{subequations}
397: \label{eq:conecoords}
398: \begin{eqnarray}
399: \xi_\bk &&\simeq {\bf v}_{\rm F} \cdot \bk \equiv v_{\rm F} k_1 =
400: \epsilon\cos\theta , \\
401: \Delta g(\bk) &&\simeq {\bf v}_2 \cdot \bk \equiv v_2 k_2 =
402: \epsilon\sin\theta,
403: \end{eqnarray}
404: \end{subequations}
405: in units where $\hbar=1$.
406: Here, $v_{\rm F}$ and $v_2$ are the Fermi velocity and a suitable
407: `gap' velocity, respectively, evaluated at $E_\bk =0$, and
408: $\epsilon$ measures the distance in energy from a given dispersionless
409: point implicitly defined by $E_\bk =0$.
410: In terms of the new coordinates, the superconducting spectrum for a
411: simple $d$-wave superconductor near a node therefore looks
412: like an anisotropic Dirac cone~\cite{Lee:97},
413: \begin{equation}
414: E_\bk \sim \left(v_{\rm F}^2 k_1^2 + v_2^2 k_2^2 \right)^{1/2} =\epsilon.
415: \label{eq:Dirac}
416: \end{equation}
417:
418: The observation of flatter structures near the nodes~\cite{Mesot:99}
419: not only implies a more significant anisotropy ratio $v_{\rm F} /
420: v_2$, but also the possibility that higher order terms in
421: $\epsilon$ may contribute to $E_\bk$, Eq.~(\ref{eq:Dirac}).
422: Indeed, within the ILT model, from Eq.~(\ref{eq:delta}) at $T=0$ one obtains
423: \begin{equation}
424: E_\bk \sim \epsilon \left[1 + \left(\frac{\epsilon}{\epsilon_\star}
425: \right)^3 \sin^6 \theta \right],
426: \label{eq:EkILT}
427: \end{equation}
428: to lowest order in $\epsilon/\epsilon_\star$, with $1/\epsilon_\star^3
429: = (1/2)(T_J /\Delta^4 )$ related to the pair-tunneling amplitude
430: $T_J$ and to the auxiliary gap parameter $\Delta$
431: (Figs.~\ref{fig:dispersion} and \ref{fig:cones}).
432:
433: \begin{figure}
434: \centering
435: \epsfig{file=ek.ps,height=0.8\columnwidth,width=0.8\columnwidth,angle=-90}
436: \caption{Typical contour lines of the superconducting spectrum $E_\bk$ in the
437: simple $d$-wave case (dashed lines) and in presence of ILT
438: (continuous line).}
439: \label{fig:dispersion}
440: \end{figure}
441:
442: Other models, based on extended in-plane
443: pairing mechanisms, would in general yield different polynomial
444: corrections in $\epsilon$ to $E_\bk$.
445: For instance, within the spin fluctuation theory~\cite{Monthoux:91},
446: the following phenomenological expansion holds for the momentum
447: distribution of the superconducting energy gap~\cite{Ghosh:99}
448: \begin{equation}
449: \Delta_\bk = \Delta g(\bk) \sum_{n=0}^N d_n h^n (\bk),
450: \label{eq:SF}
451: \end{equation}
452: with all coefficients $d_n =1$.
453: We explicitly observe that for $N=0$, $d_0 =1$, one recovers the
454: simple $d$-wave gap $\Delta_\bk \propto g(\bk)$, while the
455: case $N=1$, with the identifications $\Delta \mapsto B\Delta$,
456: $d_0 = 1$, $d_1 = 4(1-B)/B$, maps to~\cite{Mesot:99,Ghosh:99}:
457: \begin{equation}
458: \Delta_\bk = \Delta [ Bg(\bk) + (1-B)g(2\bk)],
459: \label{eq:Mesot2}
460: \end{equation}
461: which is compatible with the phenomenological fit Eq.~(\ref{eq:Mesot})
462: proposed by Mesot \emph{et al.} in Ref.~\onlinecite{Mesot:99} for
463: their experimental data of $|\Delta_\bk |$ along the Fermi
464: line~\cite{Mesot:99,Ghosh:99}.
465: In particular, Eq.~(\ref{eq:Mesot2}) would follow from a correction
466: $\delta V_{\bk\bk^\prime} \propto g(2\bk)g(2\bk^\prime )$ to the
467: in-plane coupling, corresponding to next-nearest neighbours
468: interaction.
469:
470: In such a particular case, and
471: assuming for simplicity $t^\prime = \mu =0$ in Eq.~(\ref{eq:disp}),
472: one straightforwadly obtains
473: \begin{equation}
474: E_\bk \sim \epsilon \left[ \cos^2 \theta + \sin^2 \theta \left( 1 -
475: \frac{\epsilon}{\tilde\epsilon_\star} \cos\theta \right)^2
476: \right]^{1/2} ,
477: \label{eq:Ekext}
478: \end{equation}
479: where $\tilde\epsilon_\star = tB/(1-B)$ is now related to the ratio of
480: nearest \emph{vs} next-nearest neighbours coupling.
481: Therefore, both within the ILT model and within other models, based on
482: extended in-plane pairing, the additional mechanism responsible for
483: the nonlinear correction to $E_\bk$ away from its nodes introduces
484: new energy scales (here, $\epsilon_\star$ or
485: $\tilde\epsilon_\star$, respectively).
486: Fig.~\ref{fig:cones} depicts the two different ways in which $E_\bk$
487: deviates from the cone-like shape, Eq.~(\ref{eq:Dirac}), near a
488: node, in the two cases given by Eqs.~(\ref{eq:EkILT}) and
489: (\ref{eq:Ekext}).
490:
491: \begin{figure}
492: \centering
493: \epsfig{file=surfaces_mod.ps,width=0.95\columnwidth,angle=-90}
494: \caption{Deviations from the simple $d$-wave case,
495: Eq.~(\protect\ref{eq:Dirac}) (continuous line), of the
496: superconducting spectrum $E_\bk$ around a node, within the ILT
497: model, Eq.~(\protect\ref{eq:EkILT}) (dashed line), and in the case
498: of an extended $d$-wave gap, Eq.~(\protect\ref{eq:Ekext})
499: (dashed-dotted line), as a function of the reduced coordinates
500: $\epsilon/\epsilon_\star$ ($\epsilon/\tilde\epsilon_\star$,
501: respectively).
502: Such deviations are most easily seen along the direction of the nodal
503: line (left panel, $v_2 k_2 =0$ or $\theta=\pi$), and along the Fermi
504: line (right panel, $v_{\rm F} k_1 = 0$ or $\theta=\pi/2$).
505: Note that along $v_2 k_2 =0$, one has $E_\bk =\epsilon$ also within
506: the ILT model.
507: }
508: \label{fig:cones}
509: \end{figure}
510:
511: In the absence of any such additional mechanism ($\epsilon_\star,
512: \tilde\epsilon_\star = 0$), the leading contribution to
513: Eq.~(\ref{eq:integral}) for the simplest, reference case $\varphi_\bk
514: (\beta)\equiv 1$ is:
515: \begin{equation}
516: {\cal F}_1 (\beta) \equiv {\cal F}[\beta;1] \doteq \frac{A}{\beta^2} ,
517: \label{eq:conventional}
518: \end{equation}
519: where $A = (2\pi v_{\rm F} v_2 \Delta )^{-1}$ is a doping-dependent
520: factor, and $\doteq$ denotes equality up to terms vanishing
521: exponentially with $\beta$ at \emph{all} energy scales, as
522: $\beta\to\infty$ ($T\to0$).
523: Eq.~(\ref{eq:conventional}) should be regarded as typical of the
524: power-law asymptotic low-temperature behaviour of the superconducting
525: electronic properties within a simple $d$-wave BCS-like model.
526:
527: In order to obtain an asymptotic expansion for ${\cal F}_1 (\beta)$ as
528: $\beta\to\infty$ ($T\to0$), including the corrections due to ILT,
529: Eq.~(\ref{eq:EkILT}), we observe that the integration over
530: $\epsilon$ in Eq.~(\ref{eq:integral}) is actually made of two
531: contributions:
532: \begin{equation}
533: \int_0^\infty d\epsilon = \int_0^{\epsilon_\star} d\epsilon +
534: \int_{\epsilon_\star}^\infty d\epsilon.
535: \end{equation}
536: In the first integral, we may safely retain only the linear term $E_\bk
537: \sim \epsilon$ in the exponent, since $\epsilon\leq\epsilon_\star$.
538: In the second contribution, this is no longer possible, and
539: Eq.~(\ref{eq:EkILT}) has to be retained in full.
540: However, since $\epsilon\geq\epsilon_\star > 0$, one can make use of
541: Laplace's (saddle point) method for the integral over angles around
542: $\theta=0$.
543: The final result is:
544: \begin{eqnarray}
545: {\cal F}_1 (\beta) &&\sim \frac{A}{\beta^2} \left[ 1 -
546: (1+\beta\epsilon_\star ) e^{-\beta\epsilon_\star} + \frac{1}{3\pi}
547: \Gamma \left( \frac{1}{6} \right) \beta\epsilon_\star \Gamma \left(
548: \frac{4}{3} , \beta\epsilon_\star \right) \right] \nonumber \\
549: &&\sim \frac{A}{\beta^2} \left[ 1 - \left( 1 + \beta\epsilon_\star -
550: \frac{1}{3\pi} \Gamma \left( \frac{1}{6} \right)
551: (\beta\epsilon_\star )^{5/6} \right) e^{-\beta\epsilon_\star}
552: \right],
553: \label{eq:asymptotic}
554: \end{eqnarray}
555: where $\Gamma(x)$, $\Gamma(\alpha,x)$ are Euler gamma and incomplete
556: gamma functions, respectively~\cite{Gradshteyn:94}.
557: A comparison of Eq.~(\ref{eq:conventional}) and
558: Eq.~(\ref{eq:asymptotic}) is provided by Fig.~\ref{fig:comparison},
559: and shows that ${\cal F}_1 (\beta)$ gets effectively \emph{suppressed}
560: in the presence of flat nodes in the order parameter, as provided
561: by the ILT mechanism, with respect
562: to the simple $d$-wave case, at an energy scale
563: $\sim\epsilon_\star$.
564:
565: \begin{figure}
566: \centering
567: \epsfig{file=deviation.ps,width=0.95\columnwidth,angle=-90}
568: \caption{Asymptotic power-law ($\propto T^2$, solid line) and modified
569: power-law behaviours of ${\cal F}_1 (\beta)$, in the presence of ILT
570: (dashed line), and in the extended $d$-wave case (dashed-dotted
571: line), as $(\beta\epsilon_\star )^{-1} \to0$.
572: Given the values used for the fits of $|\Delta_\bk |$ along the Fermi
573: line in Refs.~\protect\onlinecite{Mesot:99} and
574: \protect\onlinecite{Angilella:99}, it turns out that
575: $\epsilon_\star \sim 250$~K.}
576: \label{fig:comparison}
577: \end{figure}
578:
579: No such simple asymptotic expansion for ${\cal F}_1 (\beta)$ is available
580: in the extended $d$-wave case described by Eq.~(\ref{eq:Ekext}),
581: and the integrations have to be performed numerically.
582: Fig.~\ref{fig:comparison} shows the result, with the identifications
583: $A\mapsto \tilde A = (2\pi v_{\rm F} v_2 B\Delta )^{-1}$ and
584: $\epsilon_\star \mapsto \tilde\epsilon_\star$.
585: In this case, Eq.~(\ref{eq:Ekext}) provides $E_\bk$ with a different
586: kind of anisotropy with respect to the simple case,
587: Eq.~(\ref{eq:Dirac}), than Eq.~(\ref{eq:EkILT}) does.
588: While in the latter case one always has $E_\bk \geq \epsilon$, here one
589: has $E_\bk \gtreqless\epsilon$, depending on the angle $\theta$
590: (cfr. Fig.~\ref{fig:cones}).
591: As a consequence, ${\cal F}_1 (\beta)$ is \emph{enhanced}
592: with respect to the simple $d$-wave case, at an energy scale
593: $\sim\tilde\epsilon_\star$.
594:
595: \section{Conclusions}
596:
597: Motivated by recent experimental findings of extended flat structures
598: in the order parameters of the underdoped $d$-wave superconductor
599: \BSCCO~\cite{Mesot:99}, we have addressed the issue of whether
600: nonlinear, high-energy corrections to the superconducting energy
601: spectrum $E_\bk$ around the gap nodes induce deviations in
602: the predicted power-law behaviour of several electronic properties
603: at low or intermediate temperatures.
604: We have shown that nonlinear corrections to $E_\bk$
605: in general introduce additional energy scales in the problem.
606: Deviations from the usual power-law behaviour of the
607: superconducting electronic properties are indeed to be expected at such
608: energy scales, but the actual value and \emph{sign} of such
609: deviations are specific to the model under consideration.
610: In particular, within the ILT model, we have explicitly derived the
611: expected corrections to a typical power-law asymptotic behaviour as
612: $T\to0$, showing these to be \emph{negative,} whereas within a
613: phenomenological model of extended $d$-wave
614: superconductivity~\cite{Mesot:99} such corrections are predicted to
615: be \emph{positive.}
616: Whether such deviations will actually be observable in real
617: measurements of superconducting electronic properties, will of
618: course depend on the effective values of the additional energy
619: scales $\epsilon_\star$ or $\tilde\epsilon_\star$ in real compounds.
620:
621: \begin{acknowledgement}
622: G. G. N. A. thanks P. Falsaperla, J. O. Fj\ae restad, and Ch.~W\"alti for
623: valuable discussions, and acknowledges the NTNU (Trondheim, Norway)
624: for warm hospitality and financial support during the period in
625: which the present work was brought to completion.
626: A. S. acknowledges support from Norges Forskningsr{\aa}d through
627: Grants No.~110566/410 and No.~110569/410.
628: \end{acknowledgement}
629:
630: \appendix
631:
632: \section{Low-temperature superconducting electronic properties}
633: \label{app:electronic}
634:
635: We now give a sketch of how the low-temperature asymptotic behaviour of
636: several electronic properties of interest can be reduced to that of
637: ${\cal F}_1 (\beta)$ or its derivatives.
638: Most electronic quantities in the superconducting state are in fact
639: given by Eq.~(\ref{eq:integral}), with $\varphi_\bk (\beta)$
640: actually depending on $\bk$ only through $E_\bk$.
641: In what follows, $f(\epsilon) = [1+\exp(\beta\epsilon)]^{-1}$ denotes
642: the Fermi function.
643:
644: Within BCS theory, the electronic specific heat is given
645: by~\cite{Leggett:75}:
646: \begin{eqnarray}
647: C_V &&= \sum_\bk 2 k_{\rm B} \beta E_\bk \left[ E_\bk + \frac{\partial
648: E_\bk}{\partial\beta} \right] \left( -\frac{\partial f}{\partial
649: E_\bk} \right) \nonumber \\
650: &&\sim \sum_\bk 2 k_{\rm B} \beta E_\bk^2 \left( -\frac{\partial f}{\partial
651: E_\bk} \right) \nonumber\\
652: &&\sim 2k_{\rm B} \beta^2 {\cal F}_1^{\prime\prime} (\beta),
653: \end{eqnarray}
654: where apices denote derivatives with respect to $\beta$.
655: Here, $\varphi_\bk (\beta) = 2k_{\rm B} \beta^2 E_\bk^2$, and we have
656: made use of the fact that $(-\partial f/\partial E_\bk ) \doteq
657: \beta\exp(\beta E_\bk)$.
658:
659: Analogously, the unrenormalized, static, isotropic spin susceptibility
660: $\chi_0 = \chi_0 ({\bf q}\to 0,\omega\to 0)$, which is directly
661: related to the Knight shift, is simply given
662: by~\cite{Schrieffer:64,Sudbo:94}:
663: \begin{equation}
664: \chi_0 = \sum_\bk \left( -\frac{\partial f}{\partial E_\bk} \right)
665: \doteq \beta {\cal F}_1 (\beta).
666: \end{equation}
667:
668: The expression of the electronic thermal conductivity for an
669: anisotropic $d$-wave
670: superconductor also involves an average of $(-\partial f/\partial
671: E_\bk )$ over the 1BZ~\cite{Bardeen:59,Krishana:97}:
672: \begin{equation}
673: \kappa_e = \frac{1}{T} \sum_\bk \left( -\frac{\partial f}{\partial
674: E_\bk} \right) E_\bk^2 \left( \frac{\partial E_\bk}{\partial k_x}
675: \right)^2 \tau (\bk),
676: \end{equation}
677: where $\tau(\bk)$ is the superconducting quasiparticles lifetime.
678: Due to the presence of the $x$ component of the group velocity
679: $\nabla_\bk E_\bk$, however, its expression in our notation reduces
680: to:
681: \begin{equation}
682: \kappa_e = \frac{1}{8\pi^2} \frac{\ell_0}{v_2} \frac{1}{T}
683: \int_0^\infty d\epsilon\,\epsilon^3 \int_0^{2\pi} d\theta \left(
684: -\frac{\partial f}{\partial\epsilon} \right) \left( \cos\theta +
685: \frac{v_2}{v_{\rm F}} \sin\theta \right)^2 ,
686: \end{equation}
687: where $\ell_0 = v_{\rm F} \tau(\bk_{\rm F} )$ is the quasiparticle
688: mean free path at the nodes.
689: The final result crucially depends on the anisotropy ratio $v_2
690: /v_{\rm F}$, and would be different in the two cases given by
691: Eqs.~(\ref{eq:EkILT}) and (\ref{eq:Ekext}), due to their different
692: $\theta$ dependence.
693: This has to be contrasted with the result obtained in the simple
694: $d$-wave case, where~\cite{Krishana:97}:
695: \begin{equation}
696: \kappa_e = \eta k_{\rm B}^3 T^2 \frac{\ell_0}{v_2} \left( 1 +
697: \frac{v_2^2}{v_{\rm F}^2} \right),
698: \end{equation}
699: with $\eta = (8\pi)^{-1} \int_0^\infty d x\, x^3 (-\partial
700: f/\partial x)$.
701:
702: \section{A limiting case}
703:
704: In the absence of in-plane coupling, a spurious solution of the
705: mean-field gap equation at $T=0$ can be
706: implicitly expressed via~\cite{Angilella:99}:
707: \begin{equation}
708: E_\bk = \frac{1}{2} T_J (\bk) = \frac{1}{2} T_J g^4 (\bk).
709: \end{equation}
710: In such a limiting case, the superconducting energy spectrum would have
711: purely kinetic origin, and would be identified with the interlayer
712: pair-tunneling amplitude, divided by two.
713: A closed expression can then be obtained for ${\cal F}_1 (\beta)$, by
714: utilizing the useful result:
715: \begin{equation}
716: \frac{1}{(2\pi)^2} \int d^2 \bk {\cal G} [\eta(\bk)] = \frac{2}{\pi^2}
717: \int_{-1}^1 dx {\cal G} (x) K(\sqrt{1-x^2}),
718: \end{equation}
719: where ${\cal G}[\eta(x)]$ is any continuous functional of $\eta(\bk) =
720: h(\bk)$ or $g(\bk)$ alone, and $K(x^\prime )$ ($x^\prime =
721: \sqrt{1-x^2}$) is the complete elliptic integral of first
722: kind~\cite{Gradshteyn:94}.
723: From Eq.~(\ref{eq:integral}), expanding $K(x^\prime )$ around $x=0$,
724: one eventually arrives at the closed
725: expression:
726: \begin{equation}
727: {\cal F}_1 (\beta) \sim \frac{1}{\pi^2} \Gamma \left( \frac{1}{4} \right)
728: \frac{1}{\zeta^{1/4}} \left( 2\log 2 -\frac{1}{4} \psi \left(
729: \frac{1}{4} \right) + \log \zeta^{1/4} \right),
730: \end{equation}
731: where $\psi(x)$ is the digamma function~\cite{Gradshteyn:94}, and the
732: ILT amplitude $T_J$ itself here fixes the appropriate energy scale,
733: through $\zeta = \frac{1}{2} \beta T_J$.
734:
735: %\vspace{-2\baselineskip}
736:
737: \bibliography{r}
738: \bibliographystyle{prsty}
739:
740:
741:
742: \end{document}
743:
744: