1: \documentclass[draft,prb,twocolumn,showpacs,nobibnotes]{revtex4}
2:
3: \usepackage{bm}
4: \usepackage[dvips,final]{graphics}
5:
6: \begin{document}
7:
8: \title{Violation of the Wiedemann-Franz Law in a Large-N Solution of the t-J Model}
9:
10: \author{A. Houghton}
11: \author{S. Lee}
12: \author{J. B. Marston}
13: \affiliation{Department of Physics, Brown University, Providence, RI
14: 02912-1843}
15:
16: \date{\today}
17:
18: \begin{abstract}
19: We show that the Wiedemann-Franz law, which holds for Landau Fermi liquids,
20: breaks down in a large-n treatment of the t-J model.
21: The calculated ratio of the in-plane
22: thermal and electrical conductivities agrees quantitatively with
23: experiments on the normal state of the electron-doped Pr$_{2-x}$Ce$_x$CuO$_4$ ($x = 0.15$)
24: cuprate superconductor.
25: The violation of the Wiedemann-Franz law in the uniform phase contrasts with
26: other properties of the phase that are Fermi liquid like.
27: \end{abstract}
28:
29: \pacs{71.10.-w, 71.10.Hf, 71.27.+a, 72.10.-d}
30:
31: \maketitle
32:
33: A recent experiment that measured the electrical and thermal conductivities of
34: the copper-oxide superconductor Pr$_{2-x}$Ce$_x$CuO$_4$ in its normal state
35: found striking deviations from the Wiedemann-Franz law\cite{hill}. Simply
36: stated, the Wiedemann-Franz law says that fermion quasiparticles transport
37: both electrical and heat currents, with the ratio of the
38: heat conductivity $\kappa$ to the electrical conductivity $\sigma$ given by:
39: \begin{equation}
40: {{\kappa}\over{\sigma T}} = {{\pi^2}\over{3}}~ \bigg{(} {{k_B}\over{e}} \bigg{)}^2
41: \end{equation}
42: Ordinary Landau Fermi liquids respect the Wiedemann-Franz law, so deviations from
43: it indicate the presence of non-Fermi liquid physics.
44: The results of Ref.~\onlinecite{hill}
45: are therefore broadly consistent with other experimental
46: evidence that points to non-Fermi
47: liquid behavior in the cuprate phase diagram.
48:
49: One possible interpretation of the breakdown of the Wiedemann-Franz law
50: is that the quasiparticle fractionalizes into separate carriers of charge
51: and spin. To see what effects such fractionalization might have,
52: consider first heat transport in a
53: system of weakly-interacting electrons, which could be viewed as a crude approximation
54: to electrons in the highly overdoped region of the cuprate phase diagram.
55: The electrons transport both charge and heat. The electrical conductivity is given by
56: the Drude formula $\sigma = n e^2 \tau / m$ and the thermal conductivity $\kappa \propto T$ because
57: while each quasiparticle carries fixed charge $e$, it only carries an
58: energy of order the temperature. Next consider a model for the undoped cuprates: the N\'eel
59: ordered antiferromagnetic insulator with zero electrical conductivity.
60: Phonons and spinwaves transport heat, and as both excitations are bosonic in character
61: with linear dispersions at low energy, each contributes
62: similarly, yielding a thermal conductivity $\kappa \propto T^3$.
63:
64: Leaving aside these ordinary states of matter, consider the case in which the spins in
65: the insulator, instead of ordering, fractionalize into spinons with an extended
66: Fermi surface and a non-zero density of states\cite{PWA,BZA}.
67: Now spins transport heat in much the same way as charges do
68: in the non-interacting electron system, with $\kappa \propto T$. The Lorenz ratio
69: is infinite, and remains significantly larger than one upon doping with holes or electrons.
70:
71: Just this scenario is predicted in a large-n treatment of the t-J model on the
72: square lattice. We follow the approach of Refs.~\onlinecite{AM,MA,Ted}.
73: Implementing the single-occupancy constraint by introducing slave-boson operators
74: $b_i$, the t-J model may be written:
75: \begin{eqnarray}
76: H &=& -t~ \sum_{<i,j>}~ (c^{\dagger \alpha}_i b_i c_{j \alpha} b^\dagger_j + H.c.)
77: \nonumber \\
78: &+& J~ \sum_{<i,j>}~ (\vec{S}_i \cdot \vec{S}_j - \frac{1}{4} n_i n_j)
79: \nonumber \\
80: &+& {{1}\over{2}}~ \sum_{i \neq j}~ V(|\vec{r}_i - \vec{r}_j|)~ n_i n_j \ .
81: \label{tJ}
82: \end{eqnarray}
83: Number and spin operators are related to the electron creation and annihilation
84: operators by $n_i \equiv c^{\dagger \alpha}_i c_{i \alpha}$ and
85: $\vec{S}_i \equiv (1/2) c^{\dagger \alpha}_i \vec{\sigma}_\alpha^\beta c_{i \beta}$
86: where there is an implicit sum over pairs of raised and lowered Greek indices.
87: The single-occupancy constraint is now holonomic,
88: $b^\dagger_i b_i + c^{\dagger \alpha}_i c_{i \alpha} = 1$,
89: with the physical meaning that only a hole, or a single electron, may occupy each site
90: of the lattice. (In the following
91: we generally refer to hole doping with the understanding
92: that our calculations apply equally well to electron doped systems.)
93: Included in Eq.~\ref{tJ} is the off-site Coulomb repulsion $V(r)$; however, in the uniform
94: and staggered flux phases discussed below it plays no role other than to contribute an
95: additive constant to the total energy.
96:
97: Because spin-exchange involves no net flow of charge, electrical current
98: only arises from the hopping term. The continuity equation relates the time
99: rate of change of the occupancy on a given site to the lattice divergence of the
100: current flowing into the site:
101: \begin{equation}
102: e * {{d n_j(t)}\over{dt}} = -i * e * [n_j,~ H]
103: = - \sum_{\hat{e} = \hat{x}, \hat{y}} {{j^e_{j, j + \hat{e}} - j^e_{j - \hat{e}, j}}\over{a}}
104: \end{equation}
105: where $a$ is the lattice spacing between copper atoms. Thus
106: \begin{equation}
107: j^e_{j, j + \hat{e}} = -i e * t * a * (c^{\dagger \alpha}_{j} b_j
108: c_{j + \hat{e} \alpha} b^\dagger_{j + \hat{e}} - H.c.)
109: \end{equation}
110: is the electric current flowing from site $j$ into site $j+\hat{e}$ along the link connecting
111: the two sites. We emphasize that neither the spin-spin exchange interaction $J$
112: nor the Coulomb interaction $V(r)$ appear in the expression
113: for the electrical current. This result, which is a direct consequence
114: of the gauge invariance of the spin-exchange and Coulomb interactions
115: in our microscopic calculation, contrasts with that obtained recently\cite{yang}
116: within the more phenomenological ``d-density wave'' picture\cite{nayak}.
117:
118: The heat current $j^q_{j, j+\hat{e}}$ can be found by taking the time derivative of the
119: Hamiltonian density:
120: \begin{equation}
121: {{d h_j(t)}\over{dt}} =
122: - \sum_{\hat{e} = \hat{x}, \hat{y}} {{j^q_{j, j+\hat{e}} - j^q_{j - \hat{e}, j}}\over{a}}
123: \end{equation}
124: where
125: \begin{equation}
126: h_j = \sum_{i = j \pm \hat{x}, \hat{y}}~ (-t~ c^{\dagger \alpha}_i b_i c_{j \alpha} b^\dagger_j + H.c.)
127: + J~ \vec{S}_i \cdot \vec{S}_j
128: \label{ham-density}
129: \end{equation}
130: and the sum is only over the four sites $i$ that are nearest-neighbors of site $j$.
131: We have dropped the $n_i n_j$ interaction terms. These do not contribute to
132: the DC thermal conductivity because $\langle n_i \rangle$ remains unchanged in
133: the presence of currents. However, in contrast to the electrical current,
134: the heat current has contributions both from hopping, and from spin-exchange:
135: \begin{eqnarray}
136: j^q_{j, j+\hat{e}} &=& t * a * [ c^{\dagger \alpha}_{j} b_j
137: \partial_t (c_{j + \hat{e} \alpha} b^\dagger_{j + \hat{e}}) + H.c.]
138: \nonumber \\
139: &+& {{J * a}\over{2}} * [ c^{\dagger \alpha}_j (\partial_t c_{j + \hat{e} \alpha})
140: c^{\dagger \beta}_{j + \hat{e}} c_{j \beta}
141: \nonumber \\
142: &+& c^{\dagger \alpha}_j c_{j + \hat{e} \alpha}
143: (\partial_t c^{\dagger \beta}_{j + \hat{e}}) c_{j \beta} ]\ .
144: \label{jq}
145: \end{eqnarray}
146:
147: The model generalizes from the physical case of $n = 2$ (up and down spins)
148: to arbitrary (even integer) values of $n$ by letting the Greek indices run over $1, \cdots, n$.
149: In the large-n limit the spin-spin interaction factorizes in the particle-hole channel.
150: Formally this factorization is implemented via a Hubbard-Stratonovich transformation
151: within the functional integral approach. Complex-valued mean-fields along the bonds
152: are introduced:
153: \begin{equation}
154: \chi_{i j} = \frac{J}{n} \langle c^{\dagger \alpha}_i c_{j \alpha} \rangle\ .
155: \end{equation}
156: The $\chi_{ij}$ fields function as the order parameter, and as they are spin-rotation
157: invariant, there is no possibility
158: of N\'eel or other spin order, and the mean-field Hamiltonian respects spin-rotational symmetry.
159: Furthermore, at sufficiently low temperatures the (holon) boson
160: fields condense\cite{condense},
161: and we may make the replacement $b_i = b^\dagger_i = \sqrt{x}$ where
162: $x$ is the hole density. The mean-field Hamiltonian, which is exact in the
163: $n \rightarrow \infty$ limit, then takes the form:
164: \begin{equation}
165: H_{MF} = \sum_{<i,j>}~ [ {{n}\over{J}}~ |\chi_{i j}|^2 -
166: (t~ x + \chi_{i j}) (c^{\dagger \alpha}_j c_{i \alpha} + H.c.) ]\ .
167: \label{H_mf}
168: \end{equation}
169: For parameters appropriate to the cuprates, $t = 0.44$ eV (following
170: Hybertsen {et al.}\cite{hybertsen}) and $J = 0.13$ eV (obtained by
171: Singh {\it et al.}\cite{singh}),
172: the minimum energy configuration has $\chi_{ij}$ both real and constant
173: when the doping exceeds $x > 0.12$. There are no broken symmetries in this
174: ``uniform'' phase. Upon suppressing dimerization with the addition of a
175: biquadratic spin exchange interaction\cite{MA,Vojta},
176: a staggered flux (SF) phase with counter-circulating
177: orbital currents\cite{AM,MA,Ted,john} occurs for $x < 0.12$, that is, in the
178: underdoped region of the phase diagram. (The biquadratic interaction which simultaneously
179: exchanges four fermions disappears in the
180: physical $n = 2$ limit of up and down spins,
181: and thus does not alter the physics of the t-J model\cite{MA}.)
182: More realistic models include non-zero next- and third-nearest neighbor
183: hopping amplitudes\cite{OKA} but these terms do not qualitatively affect
184: the phase diagram or transport behavior.
185:
186: The expressions for the two currents simplify in the large-n limit. As the boson operators
187: may be replaced by the c-number $\sqrt{x}$, the electrical current becomes:
188: \begin{equation}
189: j^e_{j, j+\hat{e}} = -i e * t * x * a * (c^{\dagger \alpha}_{j} c_{j + \hat{e} \alpha} - H.c.) \ .
190: \label{je-mf}
191: \end{equation}
192: Upon further replacing the fermion bilinear operator $c^{\dagger \alpha}_i c_{j \alpha}$
193: with ${{n}\over{J}}~ \chi$, the heat current also simplifies\cite{why}:
194: \begin{equation}
195: j^q_{j, j+\hat{e}} = (t * x + \chi) * a * (c^{\dagger \alpha}_{j} \partial_t c_{j + \hat{e} \alpha} + H.c.)\ .
196: \label{jq-mf}
197: \end{equation}
198: As the heat current differs only by the $(t * x + \chi)$ prefactor from that of a
199: non-interacting tight-binding system, in the low-temperature limit
200: the Lorenz ratio for in-plane transport is simply:
201: \begin{equation}
202: {{\kappa}\over{\sigma T}} = {{\pi^2}\over{3}}~ \bigg{(} {{k_B}\over{e}} \bigg{)}^2
203: * \bigg{(} {{t x + \chi}\over{t x}} \bigg{)}^2\ .
204: \label{wf}
205: \end{equation}
206: Thus for any $\chi \neq 0$ the ratio differs from unity, indicating a breakdown of
207: Fermi liquid theory. Note that all details of the scattering mechanisms cancel out in
208: the ratio. Direct calculation of the two conductivities in linear response shows that
209: the integrals over momentum have identical form. Only the frequency integrals differ;
210: hence for static impurities in the weak-scattering limit
211: the Lorenz ratio is given by Eq.~\ref{wf}. We note that while the order
212: parameter $\chi$ is perturbed by the application of external electric fields
213: or thermal gradients, the perturbation in $\chi$ does not alter the DC response\cite{caroli}.
214:
215: In the SF phase the $\chi$-fields are complex numbers\cite{AM,MA,Ted},
216: and the prefactor $(t x + \chi)^2$ should be replaced by $|t x + \chi |^2$.
217: Specifically, for $\chi = |\chi| \exp(i \theta)$, with phase $\theta$,
218: the Lorenz ratio generalizes to\cite{in_prep}:
219: \begin{equation}
220: {{L}\over{L_0}} =
221: {{(t x)^2 + |\chi|^2 + 2 t x |\chi| \cos(\theta)}\over{(t x)^2}}
222: \label{wf-sf}
223: \end{equation}
224: where $L \equiv \kappa / (\sigma T)$
225: and $L_0 \equiv (\pi^2/3) (k_B /e)^2 \approx 2.45 \times 10^{-8} {\rm W \Omega K^{-2}}$
226: is the Lorenz number.
227:
228: We turn now to a comparison of our predictions, Eqs.~\ref{wf} and \ref{wf-sf},
229: with experiments. In Fig.~\ref{fig:ratio-x}
230: we plot the Lorenz ratio as a function of the doping. As expected, the
231: ratio diverges as the insulating limit $x \rightarrow 0$ is approached because
232: the spinons transport only heat, not charge. In the opposite limit of large doping
233: $\chi \rightarrow 0$, and the ratio approaches unity. Landau Fermi liquid
234: theory is recovered in the dilute limit of widely spaced electrons.
235: We emphasize that the uniform phase with
236: $\chi_{ij}$ constant and real does not break any symmetries.
237: It exhibits weak pseudogap behavior
238: because, according to the mean-field equations, $|\chi|$ increases
239: slightly in size at low temperatures, which in turn increases
240: the quasiparticle bandwidth (see Eq.~\ref{H_mf})
241: and decreases the density of states (DOS)\cite{chung}.
242: For example, at a hole concentration of
243: $x = 0.15$, $|\chi| = 0.024$ eV at $T = 500$ K rising
244: to $|\chi| = 0.026$ eV at zero temperature; consequently the DOS drops by 2\%.
245: This contrasts with the strong pseudogap behavior of the SF phase which has a gap along
246: most of the Fermi surface and which breaks time reversal invariance.
247: In either phase, however, the fermionic quasiparticles are non-interacting in
248: the $n \rightarrow \infty$ limit and hence behave as long-lived Landau quasiparticles
249: such as those found in ordinary Fermi liquids.
250:
251: We note that the Wiedemann-Franz law is strongly violated in s-wave superconducting
252: states because while Cooper pairs carry charge, the condensate has no entropy.
253: In a d-wave superconductor, quasiparticle excitations at the nodes
254: result in a modified Wiedemann-Franz law\cite{durst}.
255: The violation that we find occurs in the normal state, and is a consequence of
256: the spin-charge separation inherent in the large-n solution of the t-J model,
257: and not of any incipient superconducting tendencies.
258:
259: \begin{figure}[h,t]
260: \resizebox{8cm}{!}{\includegraphics{ratio-x.eps}}
261: \caption{The Lorenz ratio (Eqs.~\ref{wf} and \ref{wf-sf}) as
262: a function of the doping for $t = 0.44$ eV, $J = 0.13$ eV.
263: The ratio approaches unity in the dilute limit, $x \rightarrow 1$.}
264: \label{fig:ratio-x}
265: \end{figure}
266:
267: \begin{figure}[h,t]
268: \resizebox{8cm}{!}{\includegraphics{ratio-T.eps}}
269: \caption{The Lorenz ratio (Eqs.~\ref{wf} and \ref{wf-sf}) as a function of temperature
270: at dopings $x = 0.06$, $0.15$, and $0.26$
271: for $t = 0.44$ eV, $J = 0.13$ eV. In the underdoped $x = 0.06$ case
272: there is a transition to the staggered flux phase at a temperature of
273: approximately $300$ K.}
274: \label{fig:ratio-T}
275: \end{figure}
276:
277: In Fig.~\ref{fig:ratio-T} we plot the temperature dependence of the
278: Lorenz ratio for three dopings at which transport experiments have been
279: conducted: $x = 0.06$, $0.15$, and $0.26$. For a single crystal of the
280: La$_{2-x}$Sr$_x$CuO$_4$ material with hole doping $x = 0.06$, the resistivity
281: was measured upon suppressing the superconductivity by application of a
282: 18 T magnetic field along the c-axis\cite{takeya}. The thermal conductivity
283: was, however, measured in the superconducting state, so it is not possible to
284: extract a real Lorenz ratio. Nevertheless it is intriguing that
285: $L / L_0 \approx 5$ at low temperatures, based on the numbers appearing in the
286: inset to Fig. 3 of Ref.~\onlinecite{takeya}. This compares
287: reasonably well with the theoretical value of $4.1$ seen in Fig.~\ref{fig:ratio-T}.
288:
289: At optimal doping, experimentally available magnetic fields can only eliminate
290: superconductivity in electron-doped compounds. In Ref.~\onlinecite{hill} a 13 T field
291: was applied to Pr$_{2-x}$Ce$_x$CuO$_4$ at $x = 0.15$ to access the normal state.
292: The measured ratio of $L / L_0 \approx 2$ found at temperatures
293: above $0.3$ K is again in reasonable quantitative agreement with
294: the theoretical value of $1.95$. At temperatures below $0.18$ K, however,
295: the experimentally determined ratio drops rapidly below one. We have no explanation for
296: the observed behavior at the lowest temperatures\cite{si}.
297:
298: Finally, in the highly overdoped regime Proust {\it et al.}\cite{proust}
299: have studied the Tl$_2$Ba$_2$CuO$_{6+\delta}$ material at a hole concentration
300: of $x \approx 0.26$. Superconductivity was suppressed in a 13 T field, and
301: $L / L_0 = 0.99 \pm 0.01$ in good agreement with the
302: Wiedemann-Franz law for Fermi liquids. The theoretical value of the ratio is $1.5$.
303: We speculate that the persistence of non-Fermi liquid behavior at large doping
304: in the mean-field theory is an artifact of the large-n approximation. Finite-n
305: corrections could possibly restore Fermi liquid behavior in the overdoped region.
306: At large doping $|\chi|$ is small compared to the effective hole hopping amplitude $t x$,
307: so fluctuations in $\chi$
308: may be expected to be relatively more important than at low doping.
309:
310: In summary we have shown that the Wiedemann-Franz law is
311: violated in a mean-field treatment of the t-J model.
312: Our analysis, which holds for weak scattering, is exact in the
313: $n \rightarrow \infty$ limit. The Lorenz ratio is significantly larger than one
314: both in the uniform phase ($x > 0.12$) and in the SF phase ($x < 0.12$).
315: The theoretical prediction is in reasonably good quantitative agreement
316: with existing experimental measurements on the cuprate materials.
317:
318: {\it Note added: } After this work was completed a paper by Kim and Carbotte (KC)
319: appeared\cite{kim} that examined the Wiedemann-Franz law within the context of
320: the phenomenological d-density wave picture. There are several differences between
321: their work and ours. The main difference is that we study both the
322: uniform phase which has no broken symmetries, and the SF or d-density phase with
323: time-reversal breaking counter-circulating currents.
324: We find that the Lorenz ratio is significantly larger than one in both phases.
325: Furthermore, at low temperatures KC find only small deviations
326: from the Wiedemann-Franz law. This is due in part to the fact that their d-density
327: order parameter (the analog of our $\chi_{ij}$) was chosen to be purely imaginary
328: (equivalent to setting $\theta = \pi/2$ in our Eq.~\ref{wf-sf}) and also because
329: their kinetic energy is not rescaled by the slave-boson doping factor, $x$, as it
330: is in our microscopic analysis of the t-J model. KC also find a
331: large temperature variation in the Lorenz ratio for the case of
332: strong scattering because the quasiparticle lifetime has a strong frequency dependence.
333:
334: %%%%%%%%%%%%%%%%%%%%%%
335: We thank John Fj{\ae}restad and Louis Taillefer for helpful comments.
336: This work was supported in part by the NSF under grant No. DMR-9712391.
337:
338: \begin{thebibliography}{}
339:
340: \bibitem{hill} R. W. Hill, C. Proust, L. Taillefer, P. Fournier, and R. L. Greene,
341: Nature {\bf 414}, 711 (2001).
342:
343: \bibitem{PWA} P. W. Anderson, Science {\bf 235}, 1196 (1987).
344:
345: \bibitem{BZA} G. Baskaran, Z. Zou, and P. W. Anderson, Solid State Comm. {\bf 63},
346: 973 (1987).
347:
348: \bibitem{AM} I. Affleck and J. B. Marston, Phys. Rev. B{\bf 37}, 3774 (1988).
349:
350: \bibitem{MA} J. B. Marston and I. Affleck, Phys. Rev. B{\bf 39}, 11538 (1989).
351:
352: \bibitem{Ted} T. C. Hsu, J. B. Marston, and I. Affleck,
353: Phys. Rev. B {\bf 43}, 2866 (1991).
354:
355: \bibitem{yang} X. Yang and C. Nayak, cond-mat/0108407.
356:
357: \bibitem{nayak} S. Chakravarty, R. B. Laughlin, D. K. Morr, and C. Nayak,
358: Phys. Rev. B {\bf 63}, 094503 (2001).
359:
360: \bibitem{condense} Strictly speaking bosons cannot condense in a purely
361: two-dimensional system at any non-zero temperature. In practice, however, weak
362: interlayer tunneling stabilizes the condensate. As the temperature scale for
363: condensation is determined primarily by the in-plane hopping amplitude $t$,
364: and since $t \gg k_B T$, we
365: assume that the bosons remain condensed at the temperatures relevant for the
366: experiments we consider here.
367:
368: \bibitem{hybertsen} M. S. Hybertsen, E. B. Stechel, M. Schluter, and
369: D. R. Jennison, Phys. Rev. B {\bf 41}, 11068 (1990).
370:
371: \bibitem{singh} R. R. P. Singh, P. A. Fleury, K. B. Lyons, and P. E. Sulewski,
372: Phys. Rev. Lett. {\bf 62}, 2736 (1989).
373:
374: \bibitem{Vojta} M. Vojta, Y. Zhang, and S. Sachdev, Phys. Rev. B{\bf 62} 6721 (2000).
375:
376: \bibitem{john} The SF phase was recently shown to occur in a half-filled
377: two-leg ladder. See: J. O. Fj{\ae}restad and J. B. Marston, cond-mat/0107094 (to
378: appear in Phys. Rev. B) and J. B. Marston, J. O. Fj{\ae}restad, and A. Sudb{\o},
379: cond-mat/0202188.
380:
381: \bibitem{OKA} O. K. Andersen, A. I. Liechtenstein, O. Jepsen, and F. Paulsen,
382: J. Phys. Chem. Solids. {\bf 56}, 1573 (1995).
383:
384: \bibitem{why} It is straightforward to verify that the heat
385: current-current correlation function is the same in the large-n limit
386: for either form of the current operator, Eq.~\ref{jq} or Eq.~\ref{jq-mf}.
387:
388: \bibitem{caroli} The BCS pairing order parameter $\Delta$ is likewise perturbed in a
389: superconductor, but in the absence of an external magnetic field, the perturbation
390: has no effect on the low-frequency conductivities.
391: See: C. Caroli and K. Maki, Phys. Rev. {\bf 159}, 306 (1967); {\bf 159}, 316 (1967).
392:
393: \bibitem{in_prep} A. Houghton, S. Lee, and J. B. Marston, in preparation.
394:
395: \bibitem{chung} C. H. Chung, J. B. Marston, and R. H. McKenzie,
396: J. Phys. C. {\bf 13}, 5159 (2001).
397:
398: \bibitem{durst} A. C. Durst and P. A. Lee, Phys. Rev. B {\bf 62}, 1270 (2000).
399:
400: \bibitem{takeya} J. Takeya, Y. Ando, S. Komiya, and X. F. Sun, Phys. Rev. Lett.
401: {\bf 88}, 077001 (2002).
402:
403: \bibitem{si} For one possible explanation see: Q. Si, Physica C {\bf 364-365},
404: 9 (2001).
405:
406: \bibitem{proust} C. Proust, E. Boaknin, R. W. Hill, L. Taillefer, and A. P. Mackenzie,
407: cond-mat/0202101.
408:
409: \bibitem{kim} W. Kim and J. P. Carbotte, cond-mat/0202514.
410:
411: \end{thebibliography}
412: \end{document}
413: