1: % INJ
2:
3: %\documentclass[aps,superscriptaddress,showpacs,preprint]{revtex4}
4: \documentclass[aps,superscriptaddress,showpacs,twocolumn]{revtex4}
5: %\documentclass[aps,prb,superscriptaddress,showpacs,twocolumn]{revtex4}
6: %\documentclass[aps,superscriptaddress,showpacs,preprint]{revtex4}
7: \usepackage{graphicx}
8: \usepackage{color}
9: \usepackage{amssymb}
10: \usepackage{amsmath}
11:
12: \newcommand{\bk}{{\bf k}}
13: \newcommand{\bq}{{\bf q}}
14: \newcommand{\bQ}{{\bf Q}}
15: \newcommand{\bp}{{\bf p}}
16: \newcommand{\bx}{{\bf x}}
17: \newcommand{\br}{{\bf r}}
18: \newcommand{\Li}{{\mathop{\rm{Li}}\nolimits}}
19: \renewcommand{\Im}{{\mathop{\rm{Im}}\nolimits\,}}
20: \renewcommand{\Re}{{\mathop{\rm{Re}}\nolimits\,}}
21: \newcommand{\sgn}{{\mathop{\rm{sgn}}\nolimits\,}}
22: \newcommand{\Tr}{{\mathop{\rm{Tr}}\nolimits\,}}
23: \newcommand{\EF}{E_{\mathrm{F}}}
24: \newcommand{\kB}{k_{\mathrm{B}}}
25: \newcommand{\Green}{{\mathcal G}}
26: \newcommand{\dSC}{{\mathrm{dSC}}}
27: \newcommand{\dPG}{{\mathrm{dPG}}}
28: \newcommand{\dDW}{{\mathrm{dDW}}}
29: \newcommand{\Ret}{{\mathrm{R}}}
30: \newcommand{\Tau}{T_\tau}
31:
32: %\newcommand{\modified}[1]{\textcolor{red}{#1}}
33: %\newcommand{\added}[1]{\textcolor{red}{#1}}
34: %\newcommand{\deleted}[1]{\textcolor{yellow}{#1}}
35: %\newenvironment{emodified}{\color{red}}{\normalcolor}
36:
37: \newcommand{\modified}[1]{\relax #1}
38: \newcommand{\added}[1]{\relax #1}
39: \newcommand{\deleted}[1]{\relax #1}
40: \newenvironment{emodified}{\relax}{\relax}
41:
42:
43: \begin{document}
44:
45: \title{Comment on ``Variation of the superconducting transition
46: temperature of hole-doped copper oxides''.}
47:
48:
49:
50: \author{G. G. N. Angilella}
51: \email{Giuseppe.Angilella@ct.infn.it}
52: \affiliation{Dipartimento di Fisica e Astronomia, Universit\`a di
53: Catania,\\ and Istituto Nazionale per la Fisica della Materia,
54: UdR di Catania,\\ Via S. Sofia, 64, I-95123 Catania, Italy}
55: \author{R. Pucci}
56: \affiliation{Dipartimento di Fisica e Astronomia, Universit\`a di
57: Catania,\\ and Istituto Nazionale per la Fisica della Materia,
58: UdR di Catania,\\ Via S. Sofia, 64, I-95123 Catania, Italy}
59: \author{A. Sudb\o}
60: \affiliation{ Department of Physics, Norwegian University of Science
61: and Technology, N-7491 Trondheim, Norway}
62:
63: \date{\today}
64:
65:
66: \begin{abstract}
67: \medskip
68: We point out the incorrect derivation of the gap equation in X.-J. Chen
69: and H. Q. Lin [Phys. Rev. B {\bf 69}, 104518 (2994)] within the
70: interlayer tunneling (ILT) model for multilayered cuprates.
71: There, the \emph{local} structure in $\bk$-space of the ILT effective
72: interaction has not been taken into due account when the ILT model
73: is generalized to the case of $n$ layers per unit cell.
74: This is a specific characteristic of the ILT model that, apart from giving
75: rise to a highly nontrivial $\bk$-dependence of the gap function, is
76: known to enhance the critical temperature $T_c$ in a natural way.
77: As a consequence, we argue that Chen and Lin's results cannot be
78: employed, in their present form, for a quantitative interpretation
79: of the high-pressure dependence of $T_c$ in Bi-2212, as is done by
80: X.-J. Chen \emph{et al.} [{\tt cond-mat/0408587}, to appear in
81: Phys. Rev. B].
82: Moreover, when the generalization of X.-J. Chen \emph{et al.} [{\tt
83: cond-mat/0408587}] is applied to the case $n=2$, it fails to
84: reproduce the original ILT gap equation.
85: However, a more careful analysis of the ILT model for multilayered
86: cuprates, taking into account the nonuniform hole distribution
87: among inequivalent layers, has been earlier suggested to describe
88: the observed pressure dependence of $T_c$ in homologous series of
89: high-$T_c$ cuprates.
90: \\
91: \pacs{%
92: 74.62.-c, % Tc variations
93: 74.72.-h, % Cuprate superconductors (high-Tc and insulating parent compounds)
94: %74.72.Hs, % Bi-based cuprates
95: 74.62.Fj, % pressure effects
96: 74.20.-z % Theories and models of superconducting state
97: %Theories and models of superconducting state
98: }
99: \end{abstract}
100:
101: \maketitle
102:
103: %\section{Introduction}
104:
105: In Ref.~\onlinecite{Chen:04}, Chen and Lin reconsider the dependence
106: of $T_c$ on doping and on the number of layers in a homologous series of
107: multilayered high-$T_c$ cuprates within the interlayer tunneling
108: (ILT) model \cite{Chakravarty:93}.
109: However, in deriving their gap equation, Chen and Lin erroneusly
110: neglect the intrinsic \emph{local} structure in momentum ($\bk$)
111: space of the effective ILT coupling.
112: This is a specific characteristic of the ILT model, which is known to give
113: rise to highly nontrivial features in the $\bk$-dependence of the
114: gap function already for a bilayer complex
115: \cite{Angilella:99,Angilella:00}.
116: Moreover, a local term in the gap equation has been shown to provide a
117: lower bound for $T_c$ at all dopings, which is the precise way in
118: which the ILT mechanism enhances $T_c$ \cite{Angilella:99}.
119: The consequences of such an incorrect analysis of the ILT model are
120: both qualitative and quantitative.
121: Therefore, the recent use of Chen and Lin's results to interpret the
122: high-pressure dependence of $T_c$ in Bi-2212 \cite{Chen:04a} can be
123: questioned.
124: In this context, we point out that a more careful analysis of the ILT
125: model for layered cuprates has been presented elsewhere
126: \cite{Sudboe:94c}, and successfully applied to study the pressure
127: dependence of $T_c$ in homologous series of layered cuprates, by
128: explicitly taking
129: into account the inhomogeneous hole-doping in inequivalent
130: layers \cite{Angilella:99b,Wijngaarden:99}.
131:
132: Superconductivity in the high-$T_c$ layered cuprates is characterized
133: by \emph{(i)} a non-monotonic dependence of $T_c$ on the overall hole-doping
134: $\delta$; \emph{(ii)} a monotonic increase of $T_c$ with the number of
135: layers $n$, for moderately low $n$ ($n\lesssim 3$).
136: While \emph{(i)} is a generic consequence of the quasi-bidimensional nature
137: of these compounds (see \emph{e.g.} Ref.~\onlinecite{Angilella:01}), the
138: latter fact has suggested that coherent tunneling of
139: superconducting pairs between adjacent CuO$_2$ layers may
140: considerably enhance $T_c$ \cite{Chakravarty:93}.
141: Within the ILT model, it is postulated that strong in-plane
142: correlations forbid coherent hopping of single
143: particles between adjacent CuO$_2$ planes.
144: Such a restriction is removed when accessing the superconducting
145: state, where interlayer Josephson tunneling of Cooper pairs is
146: allowed.
147: This results in a net gain in kinetic energy, as compared to the
148: normal state.
149: Thus, within the ILT model, superconductivity is stabilized \emph{via} a
150: kinetic mechanism, as opposed to conventional BCS
151: superconductivity, where the enhancement in kinetic energy is
152: overcompensated by a reduction in the potential energy
153: \cite{Chakravarty:98a}.
154:
155: However, after its original formulation more than a decade ago
156: \cite{Chakravarty:93}, the relevance of the ILT mechanism at least
157: for single-layer cuprates has been called into question by
158: experiments \cite{Anderson:98,Moler:98}.
159: Recently, Chakravarty \emph{et al.} \cite{Chakravarty:04} have revived
160: the ILT model in connection with multilayered cuprates.
161: There, ILT needs not be the sole source of superconducting
162: condensation energy.
163: Charge carriers require a `seed' in-plane interaction to form Cooper
164: pairs in a given symmetry channel, before they can actually tunnel
165: between adjacent layers \cite{Angilella:99}.
166: Such in-plane interaction would then provide the missing condensation
167: energy \cite{Chakravarty:04}.
168:
169: Moreover, it has been suggested that the competition with a `hidden'
170: order parameter, such as a $d$-density-wave (dDW)
171: \cite{Chakravarty:01}, could be responsible for the downturn of $T_c$
172: with $n$, for $n\gtrsim 3$.
173: Indeed, in multilayered cuprates, due to the different proximity to
174: the `charge reservoir' blocks, experiments \cite{Trokiner:91} as
175: well as density functional theory calculations
176: \cite{Ambrosch-Draxl:04} revealed a nonuniform hole-content
177: distribution between inner and outer layers.
178: Since this usually places inner (outer) layers in the underdoped
179: (overdoped) region of the cuprate phase diagram, competition with
180: the dDW order would be stronger in inner layers than in outer
181: layers, thus depressing $T_c$ with increasing $n$.
182: Hydrostatic pressure could then be used to tune both the overall
183: hole-content content and its distribution among inequivalent
184: layers, thus inducing an `exchange of roles' between inner and outer
185: layers with respect to the onset of superconductivity
186: \cite{Angilella:99b}, which is observed as `kinks' in the
187: pressure dependence of $T_c$ in layered cuprates
188: \cite{Wijngaarden:99}.
189:
190: The effective Hamiltonian considered by Chen and Lin \cite{Chen:04}
191: (see also \cite{Angilella:99,Angilella:99b}) is
192: \begin{eqnarray}
193: H &=& \sum_{\ell \bk\sigma} \xi_\bk c_{\bk\sigma}^{\ell\dag}
194: c_{\bk\sigma}^\ell - \sum_{\ell\bk\bk^\prime} V_{\bk\bk^\prime}
195: c_{\bk\uparrow}^{\ell\dag} c_{-\bk\downarrow}^{\ell\dag}
196: c_{-\bk^\prime \downarrow}^\ell c_{\bk^\prime \uparrow}^\ell
197: \nonumber\\
198: &&+ \sum_{\langle\ell\ell^\prime \rangle} \sum_\bk T_J (\bk)
199: c_{\bk\uparrow}^{\ell\dag} c_{-\bk\downarrow}^{\ell\dag}
200: c_{-\bk\downarrow}^{\ell^\prime} c_{\bk\uparrow}^{\ell^\prime} ,
201: \label{eq:H}
202: \end{eqnarray}
203: where $\xi_\bk$ is the in-plane quasiparticle dispersion measured with
204: respect to the chemical potential $\mu^\ell$ ($\mu^\ell \equiv \mu$
205: for all layers, in Ref.~\onlinecite{Chen:04}), and
206: $c_{\bk\sigma}^{\ell\dag}$
207: is a quasiparticle creation operator with wave-vector $\bk$ and spin
208: $\sigma$ on layer $\ell$.
209: It should be emphasized that in Eq.~(\ref{eq:H}) the first interaction
210: term ($V_{\bk\bk^\prime}$) pertains to a single layer and governs the
211: overall symmetry of the order parameter (\emph{i.e.}, $d$-wave, if
212: $V_{\bk\bk^\prime} = V g_\bk g_{\bk^\prime}$, with $g_\bk =
213: \frac{1}{2} (\cos k_x - \cos k_y )$) \cite{Sudbo:95b}, while the
214: second term applies to adjacent layers
215: $\langle\ell\ell^\prime \rangle$, and is \emph{local} in
216: $\bk$-space, with $T_J (\bk) = \frac{1}{16} T_J (\cos k_x - \cos
217: k_y )^4$ \cite{Chakravarty:93}.
218: This enforces momentum conservation for the interlayer pair tunneling
219: process.
220: (The effect of $\bk$-space broadening of the ILT kernel, \emph{e.g.}
221: due to impurities, has been considered in
222: Ref.~\onlinecite{Fjaerestad:98}.)
223:
224: A straightforward mean-field analysis of Eq.~(\ref{eq:H}) for a
225: bilayer complex and an in-plane superconducting instability in the
226: $d$-wave channel yields the gap equation \cite{Angilella:99}:
227: \begin{equation}
228: \Delta_\bk = \frac{\Delta_0 g_\bk}{1-T_J (\bk)\chi_\bk} ,
229: \label{eq:gap1}
230: \end{equation}
231: where $\Delta_0$ is determined self-consistently from
232: \begin{equation}
233: 1 = \frac{V}{N} \sum_{\bk^\prime} g_{\bk^\prime}^2
234: \frac{\chi_{\bk^\prime}}{1-T_J (\bk^\prime )\chi_{\bk^\prime} }.
235: \label{eq:gap2}
236: \end{equation}
237: Here, $\chi_\bk = (2E_\bk )^{-1} \tanh (\beta E_\bk /2)$ is the pair
238: susceptibility at the inverse temperature $\beta = (\kB T)^{-1}$,
239: $E_\bk = \sqrt{\xi_\bk^2 + |\Delta_\bk |^2}$ is the upper branch of
240: the superconducting spectrum, and $N$ is the number of lattice sites.
241:
242: Eqs.~(\ref{eq:gap1}) and (\ref{eq:gap2}) should be immediately compared and
243: contrasted with Eq.~(9) in Ref.~\onlinecite{Chen:04} (for a
244: multilayered complex) and Eq.~(1) in Ref.~\onlinecite{Chen:04a}
245: (for a bilayer complex).
246: Even without going into the subtleties of the more general derivation
247: for an $n$-layered complex (for which, see
248: Refs.~\onlinecite{Sudboe:94c,Angilella:99b}), or with the
249: competition among several in-plane pairing channels
250: (Ref.~\onlinecite{Angilella:99}), it is apparent that the gap
251: function within
252: the ILT model, Eq.~(\ref{eq:gap1}), is characterized by a
253: \emph{local} prefactor $[1-T_J (\bk)\chi_\bk ]^{-1}$ which, albeit
254: linked self-consistently to $\Delta_0$ \emph{via}
255: Eq.~(\ref{eq:gap2}), is responsible of most of the quantitative and
256: qualitative features of the model.
257: Such a structure is missing in Refs.~\onlinecite{Chen:04,Chen:04a}.
258: Even though the actual symmetry of the gap function is independent of
259: the ILT kernel, and is rather determined by the $d$-wave nature of
260: the in-plane coupling, the ILT mechanism endows the gap function
261: with a nontrivial structure in $\bk$-space \cite{Angilella:99},
262: which has been shown to be consistent with ARPES results
263: \cite{Angilella:00}.
264: Moreover, the `renormalized' pair susceptibility in the summand of
265: Eq.~(\ref{eq:gap2}), \emph{viz.} $\chi_\bk \mapsto \chi_\bk /
266: [1-T_J (\bk)\chi_\bk ]$, which is due to the local ILT tunneling
267: amplitude, gives rise to additional, algebraic divergences
268: in the energy dependence of the integrated pair susceptibility, as
269: opposed to the logarithmic one, typical of BCS theory \cite{AGD}.
270: This is directly responsible of the enhancement of $T_c$ within the
271: ILT model.
272: In particular, in the case of a bilayer complex, one analytically
273: finds a lower bound for $T_c$ as
274: \begin{equation}
275: \kB T^\ast (\mu) =
276: \begin{cases}
277: \displaystyle
278: \frac{T_J}{64} \left( \frac{\mu_\perp - \mu}{\mu_\perp + 2t}
279: \right)^4 , & \mu_\perp \leq \mu < \mu_{\mathrm{VH}} ,\\
280: \displaystyle
281: \frac{T_J}{64} \left( \frac{\mu_\top - \mu}{\mu_\top - 2t}
282: \right)^4 , & \mu_{\mathrm{VH}} \leq \mu \leq \mu_\top ,
283: \end{cases}
284: \end{equation}
285: where nearest ($t$) and next-nearest ($t^\prime$) hopping have been
286: assumed, and $\mu_\perp = -4t+4t^\prime$, $\mu_\top =
287: 4t+4t^\prime$, and $\mu_{\mathrm{VH}} = -4t^\prime$
288: denote the bottom, the top of the band, and the location of the
289: Van~Hove singularity, respectively \cite{Angilella:99}.
290:
291: On the contrary, the ILT kernel enters Chen \emph{et al.}'s gap
292: equation in Eq.~(9) of Ref.~\onlinecite{Chen:04} and Eq.~(1) of
293: Ref.~\onlinecite{Chen:04} as an additional contribution to the
294: non-local in-plane coupling term, \emph{i.e.} it amounts to
295: defining another in-plane interaction, with no reference to
296: interlayer tunneling.
297: \emph{This same observation applies to the general case of an
298: $n$-layered complex.}
299: In that case, the gap equation for each layer should also contain a
300: local contribution due to the ILT mechanism between adjacent
301: layers (again, absent in Ref.~\onlinecite{Chen:04}), with an ILT
302: renormalized pair susceptibility $\chi_\bk^\ell / [1-T_J
303: (\bk)\hat{\chi}_\bk^\ell ]$ for each layer
304: \cite{Sudboe:94c,Angilella:99b}, with
305: \begin{eqnarray}
306: \hat{\chi}_\bk^\ell &=& \left[\sin\left( \frac{\ell\pi}{n+1}
307: \right)\right]^{-1} \left[ \chi_\bk^{\ell+1} \sin \left(
308: \frac{(\ell+1)\pi}{n+1} \right) \right. \nonumber\\
309: &&+ \left. \chi_\bk^{\ell-1} \sin \left(
310: \frac{(\ell-1)\pi}{n+1} \right) \right],
311: \end{eqnarray}
312: which can be further simplified in the limit of uniform hole-content
313: in all layers (as is tacitly assumed in
314: Ref.~\onlinecite{Chen:04}).
315: In analogy to the bilayer case, the condition
316: \begin{equation}
317: \min_\bk [1-T_J (\bk)\chi_\bk^\ell ] = 0
318: \end{equation}
319: then implicitly defines a lower bound $T_c^{\ast\ell}$ for the
320: critical temperature
321: corresponding to the onset of superconductivity \emph{in the given
322: layer $\ell$.}
323: Therefore, for nonuniform hole-content among inequivalent layers, as
324: is the case for the multilayered cuprates
325: \cite{Trokiner:91,Ambrosch-Draxl:04}, one can estimate a lower
326: bound to $T_c$ as $\max_\ell T_c^{\ast\ell}$.
327: A nonuniform distribution of the overall hole-content among
328: inequivalent layers can be conveniently described by means of
329: appropriate models \cite{Angilella:99b}.
330: This then enables us to identify whether the superconducting instability
331: first sets in in inner or outer layers.
332: One finds a crossover as function of the overall hole-content
333: \cite{Angilella:99b}, which has been related to the observed
334: `kinks' in the pressure dependence of $T_c$ in several layered
335: cuprates \cite{Wijngaarden:99}.
336:
337: In conclusion, we have pointed out an incorrect derivation of the gap
338: equation(s) for layered cuprates within the ILT model
339: \cite{Chen:04,Chen:04a} for the general case of $n$ superconducting
340: layers per unit cell.
341: This in turn leads to a failure to capture most of the
342: qualitative and quantitative features of the theory, both for
343: bilayered and multilayered compounds.
344: As a consequence, the theoretical analysis of the high pressure data
345: in Ref.~\onlinecite{Chen:04a} is not consistent with the ILT
346: mechanism.
347: On the other hand, a more careful analysis of the ILT model
348: \cite{Sudboe:94c}, when taking into account a nonuniform
349: hole-content distribution among inequivalent layers
350: \cite{Angilella:99b}, is indeed able to reproduce the observed
351: pressure dependence of $T_c$ in multilayered cuprates
352: \cite{Wijngaarden:99}.
353:
354:
355:
356: \begin{acknowledgments}
357: We thank S. Chakravarty and J. O. Fj\ae{}restad for useful discussions
358: and correspondence.
359: \end{acknowledgments}
360:
361: \begin{small}
362: \bibliographystyle{apsrev}
363: \bibliography{a,b,c,d,e,f,g,h,i,j,k,l,m,n,o,p,q,r,s,t,u,v,w,x,y,z,zzproceedings,Angilella}
364: \end{small}
365:
366: \end{document}
367: