1: \documentclass[aps,twocolumn,prb]{revtex4}%
2: \usepackage{amsfonts}
3: \usepackage{amsmath}
4: \usepackage{amssymb}
5: \usepackage{natbib}
6: \usepackage{graphicx}%
7: \setcounter{MaxMatrixCols}{30}
8: \begin{document}
9: \title[Pseudogap and high T$_{c}$]{Pseudogap and high-temperature superconductivity from weak to strong coupling.
10: Towards quantitative theory.}
11: \author{A.-M.S. Tremblay, B. Kyung, D. S\'{e}n\'{e}chal}
12: \affiliation{D\'{e}partement de physique and RQMP, Universit\'{e} de Sherbrooke,
13: Sherbrooke, QC J1K 2R1, Canada}
14: \keywords{Hubbard model, high temperature superconductivity, d-wave superconductivity,
15: pseudogap, Two-Particle Self-Consistent Approach, Quantum Cluster Approaches,
16: correlated electrons, quantum materials.}
17: \pacs{PACS number}
18:
19: \begin{abstract}
20: This is a short review of the theoretical work on the two-dimensional Hubbard
21: model performed in Sherbrooke in the last few years. It is written on the occasion of the twentieth
22: anniversary of the discovery of high-temperature superconductivity. We discuss
23: several approaches, how they were benchmarked and how they agree sufficiently
24: with each other that we can trust that the results are accurate solutions of
25: the Hubbard model. Then comparisons are made with experiment. We show that the
26: Hubbard model does exhibit d-wave superconductivity and antiferromagnetism
27: essentially where they are observed for both hole and electron-doped cuprates.
28: We also show that the pseudogap phenomenon comes out of these calculations. In
29: the case of electron-doped high temperature superconductors, comparisons with
30: angle-resolved photoemission experiments are nearly quantitative. The value of
31: the pseudogap temperature observed for these compounds in recent photoemission
32: experiments has been predicted by theory before it was observed
33: experimentally. Additional experimental confirmation would be useful. The
34: theoretical methods that are surveyed include mostly the Two-Particle
35: Self-Consistent Approach, Variational Cluster Perturbation Theory (or
36: variational cluster approximation), and Cellular Dynamical Mean-Field Theory.
37:
38: \end{abstract}
39: \date{November 2005}
40: \maketitle
41: \tableofcontents
42:
43:
44: \section{Introduction}
45:
46: In the first days of the discovery of high-temperature superconductivity,
47: Anderson\cite{Anderson:1987} suggested that the two-dimensional Hubbard model
48: held the key to the phenomenon. Despite its apparent simplicity, the
49: two-dimensional Hubbard model is a formidable challenge for theorists. The
50: dimension is not low enough that an exact solution is available, as in
51: one dimension. The dimension is not high enough that some mean-field theory,
52: like Dynamical Mean Field Theory\cite{Georges:1996, Jarrell:1992} (DMFT), valid in infinite
53: dimension, can come to the rescue. In two dimensions, both quantum and thermal
54: fluctuations are important. In addition, as we shall see, it turns out that
55: the real materials are in a situation where both potential and kinetic energy
56: are comparable. We cannot begin with the wave picture (kinetic energy
57: dominated, or so-called \textquotedblleft weak coupling\textquotedblright) and
58: do perturbation theory, and we cannot begin from the particle picture
59: (potential energy dominated, or so-called \textquotedblleft strong
60: coupling\textquotedblright) and do perturbation theory. In fact, even if one
61: starts from the wave picture, perturbation theory is not trivial in two
62: dimensions, as we shall see. Variational approaches on the ground state have
63: been proposed,\cite{Paramekanti:2004} but even if they capture key aspects of
64: the ground state, they say little about one-particle excitations.
65:
66: Even before the discovery of high-temperature superconductivity, it was
67: suggested that antiferromagnetic fluctuations present in the Hubbard model
68: could lead to d-wave superconductivity,\cite{Scalapino:1986, Beal-Monod:1986,
69: Miyake:1986} a sort of generalization of the Kohn-Luttinger
70: mechanism\cite{Kohn:1965} analogous to the superfluidity mediated by
71: ferromagnetic spin fluctuations in $^{3}$He.\cite{Leggett:1975} Nevertheless,
72: early Quantum Monte Carlo (QMC) simulations\cite{Hirsch:1988} gave rather
73: discouraging results, as illustrated in Fig.~\ref{fig_0}. In QMC, low
74: temperatures are inaccessible because of the sign problem. At accessible
75: temperatures, the d-wave pair susceptibility is smaller than the
76: non-interacting one, instead of diverging. Since the observed phenomenon
77: appears at temperatures that are about ten times smaller than what is
78: accessible with QMC, the problem was left open.
79: Detailed analysis of the irreducible vertex\cite{Bulut:1993} deduced from QMC did suggest the importance of d-wave pairing, but other numerical work\cite{Zhang:1997} concluded that long-range d-wave order is absent, despite the fact that slave-boson
80: approaches\cite{Kotliar:1988, Inui:1988} and many subsequent work suggested otherwise. The situation on the numerical side is changing since
81: more recent
82: variational,\cite{Paramekanti:2004} Dynamical Cluster
83: Approximation\cite{Maier:2000a} and exact diagonalization\cite{Poilblanc:2002}
84: results now point towards the existence of d-wave superconductivity in the
85: Hubbard model. Even more recently, new numerical approaches are making an even
86: more convincing case.\cite{Senechal:2005, Kancharla:2005, Maier:2005a}
87:
88: %-----------------------------------------------------------------------
89: %: FIG:fig_0
90: \begin{figure}[ptb]
91: \centerline{\includegraphics[width=7.5cm]{f0}} \caption{ The $d_{x^{2}-y^{2}}$
92: and extended $s$-wave susceptibilities obtained from QMC simulations for
93: $U=4t$ and a $4\times4$ lattice. The solid lines are the non-interacting
94: results. From Ref.~\onlinecite{Hirsch:1988}. The low temperature downturn of d-wave however seems to come from a mistreatment of the sign problem (D.J. Scalapino private communication).}%
95: \label{fig_0}%
96: \end{figure}
97: %-----------------------------------------------------------------------
98:
99: After twenty years, we should be as quantitative as possible.
100: How should we proceed to investigate a model without a small parameter?
101: We will try to follow this path: (1) Identify important physical principles and laws to constrain
102: non-perturbative approximation schemes, starting from both weak (kinetic
103: energy dominated) and strong (potential energy dominated) coupling. (2)
104: Benchmark the various approaches as much as possible against exact (or
105: numerically accurate) results. (3) Check that weak and strong coupling
106: approaches agree at intermediate coupling. (4) Compare with experiment.
107:
108: In brief, we are trying to answer the question, \textquotedblleft Is the
109: Hubbard model rich enough to contain the essential physics of the cuprates,
110: both hole and electron doped?"
111: The answer is made possible by new theoretical approaches, increased
112: computing power, and the reassurance that theoretical approaches, numerical
113: and analytical, give consistent results at intermediate coupling even if the
114: starting points are very different.
115:
116: This paper is a review of the work we have done in Sherbrooke on this subject.
117: In the short space provided, this review will not cover all of our work.
118: Needless to say, we will be unfair to the work of many other groups, even
119: though we will try to refer to the work of others that is directly relevant to
120: ours. We do not wish to make priority claims and we apologize to the authors
121: that may feel unfairly treated.
122:
123: Section \ref{Methodology} will introduce the methodology: First a method that
124: is valid at weak to intermediate coupling, the Two-Particle Self-Consistent
125: approach (TPSC), and then various quantum cluster methods that are better at
126: strong coupling, namely Cluster Perturbation Theory (CPT), the Variational
127: Cluster Approximation (VCA) also known as Variational Cluster Perturbation
128: Theory (VCPT), and Cellular Dynamical Mean Field Theory (CDMFT) with a brief
129: mention of the Dynamical Cluster Approximation (DCA). In all cases, we will
130: mention the main comparisons with exact or numerically accurate results that
131: have been used to benchmark the approaches. In Sect.~\ref{Results} we give
132: some of the results, mostly on the pseudogap and the phase diagram of
133: high-temperature superconductors. More importantly perhaps, we show the
134: consistency of the results obtained by both weak- and strong-coupling
135: approaches when they are used at intermediate coupling. Finally, we compare
136: with experiment in section \ref{Experiment}.
137:
138: \section{Methodology\label{Methodology}}
139:
140: We consider the Hubbard model%
141: \begin{equation}
142: H=-\sum_{i,j,\sigma}t_{ij}c_{i\sigma}^{\dagger}c_{j\sigma}+U\sum
143: _{i}n_{i\uparrow}n_{i\downarrow}%
144: \end{equation}
145: where $c_{i\sigma}^{\dagger}$ ($c_{i\sigma}$) are creation and annihilation
146: operators for electrons of spin $\sigma$, $n_{i\sigma}=c_{i\sigma}^{\dagger
147: }c_{i\sigma}$ is the density of spin $\sigma$ electrons, $t_{ij}=t_{ji}^{\ast
148: }$ is the hopping amplitude, and $U$ is the on-site Coulomb repulsion. In
149: general, we write $t,t^{\prime},t^{\prime\prime}$ respectively for the first-,
150: second- and third-nearest neighbor hopping amplitudes.
151:
152: In the following subsections, we first discuss how to approach the problem
153: from the weak coupling perspective and then from the strong coupling point of
154: view. The approaches that we will use in the end are non-perturbative, but in
155: general they are more accurate either at weak or strong coupling.
156:
157: \subsection{Weak coupling approach}
158:
159: Even at weak coupling, the Hubbard model presents difficulties specific to two dimensions.
160: The time-honored Random Phase Approximation (RPA) has the
161: advantage of satisfying conservation laws, but it violates the Pauli principle
162: and the Mermin-Wagner-Hohenberg-Coleman (or Mermin-Wagner, for short) theorem.
163: This theorem states that a continuous symmetry cannot be broken at finite
164: temperature in two dimensions. RPA gives a finite-temperature phase
165: transition. The Pauli principle means, in particular, that $\langle
166: n_{i\uparrow}n_{i\uparrow}\rangle=\langle n_{i\uparrow}\rangle$ in a model
167: with only one orbital per site. This is violated by RPA since it can be
168: satisfied only if all possible exchanges of electron lines are allowed (more
169: on this in the following section). Since the square of the density at a given
170: site is given by $\langle(n_{i\uparrow}+n_{i\downarrow})^{2}\rangle=2\langle
171: n_{\uparrow}n_{\uparrow}\rangle+2\langle n_{\uparrow}n_{\downarrow}\rangle$,
172: violating the Pauli condition $\langle n_{i\uparrow}n_{i\uparrow}%
173: \rangle=\langle n_{i\uparrow}\rangle$ will in general lead to large errors in
174: double occupancy, a key quantity in the Hubbard model since it is proportional
175: to the potential energy. Another popular approach is the
176: Moriya\cite{Moriya:1985} self-consistent spin-fluctuation
177: approach\cite{Markiewicz:2004} that uses a Hubbard-Stratonovich transformation
178: and a $\langle\phi^{4}\rangle\sim\phi^{2}\langle\phi^{2}\rangle$
179: factorization. This satisfies the Mermin-Wagner theorem but, unfortunately,
180: violates the Pauli principle and introduces an unknown mode-coupling constant
181: as well as an unknown renormalized $U$ in the second-order term. The
182: conserving approximation known as Fluctuation Exchange (FLEX)
183: approximation\cite{Bickers:1989} is an Eliashberg-type theory that is
184: conserving but violates the Pauli principle, assumes a Migdal theorem and does
185: not reproduce the pseudogap phenomenon observed in QMC. More detailed
186: criticism of this and other approaches may be found in
187: Refs.~\onlinecite{Vilk:1997,Vilk:1996}. Finally, the renormalization
188: group\cite{Honerkamp:2001, Halboth:2000, Furukawa:1998, Wegner:2003,
189: Hankevych:2003a} has the great advantage of being an unbiased method to look
190: for instabilities towards various ordered phases. However, it is quite
191: difficult to implement in two dimensions because of the proliferation of
192: coupling constants, and, to our knowledge, no one has yet implemented a
193: two-loop calculation without introducing additional
194: approximations.\cite{Zanchi:2001,Katanin:2004} Such a two-loop calculation is
195: necessary to observe the pseudogap phenomenon.
196:
197: \subsubsection{Two-Particle Self-consistent approach (TPSC)}
198:
199: The TPSC approach, originally proposed by Vilk, Tremblay and collaborators,
200: \cite{Vilk:1994, Vilk:1995} aims at capturing non-perturbative effects. It does not use
201: perturbation theory or, if you want, it drops diagrammatic expansions.
202: Instead, it is based on imposing constraints and sum rules: the theory should
203: satisfy (a) the spin and charge conservation laws (b) the Pauli principle in
204: the form $\langle n_{i\uparrow}n_{i\uparrow}\rangle=\langle n_{i\uparrow
205: }\rangle$ (c) the local-moment and the local-density sum rules. Without any
206: further explicit constraint, we find that the theory satisfies the
207: Mermin-Wagner theorem, that it satisfies consistency between one- and
208: two-particle quantities in the sense that $\frac{1}{2}\mathrm{Tr}(\Sigma
209: G)=U\langle n_{\uparrow}n_{\downarrow}\rangle$ and finally that the theory
210: contains the physics of Kanamori-Br\"{u}ckner screening (in other words,
211: scattering between electrons and holes includes T-matrix quantum fluctuation
212: effects beyond the Born approximation).
213:
214: Several derivations of our approach have been given,\cite{Vilk:1995,
215: Mahan:2000} including a quite formal one\cite{Allen:2003} based on the
216: functional derivative Baym-Kadanoff approach.\cite{Baym:1962} Here we only
217: give an outline\cite{Moukouri:2000} of the approach with a more
218: phenomenological outlook. We proceed in two steps. In the first step (in our
219: earlier work sometimes called zeroth step), the self-energy is
220: obtained by a Hartree-Fock-type factorization of the four-point function with
221: the \textit{additional constraint} that the factorization is exact when all
222: space-time coordinates coincide.\footnote{This additional constraint leads to
223: a degree of consistency between one- and two-particle quantities that is
224: absent from the standard Hartree-Fock factorization.} Functional
225: differentiation, as in the Baym-Kadanoff approach\cite{Baym:1962}, then leads
226: to a momentum- and frequency-independent irreducible particle-hole vertex for
227: the spin channel that satisfies\cite{Vilk:1994} $U_{sp}=U\langle n_{\uparrow
228: }n_{\downarrow}\rangle/(\langle n_{\uparrow}\rangle\langle n_{\downarrow
229: }\rangle)$. The local moment sum rule and the Pauli principle in the form
230: $\langle n_{\sigma}^{2}\rangle=\langle n_{\sigma}\rangle$ then determine
231: double occupancy and $U_{sp}$. The irreducible vertex for the charge channel
232: is too complicated to be computed exactly, so it is assumed to be constant and
233: its value is found from the Pauli principle and the local charge fluctuation
234: sum rule. To be more specific, let us use the notation, $q=(\mathbf{q,}%
235: iq_{n})$ and $k=(\mathbf{k,}ik_{n})$ with $iq_{n}$ and $ik_{n}$ respectively
236: bosonic and fermionic Matsubara frequencies. We work in units where $k_{B},$
237: $\hbar,$ and lattice spacing are all unity. The spin and charge
238: susceptibilities now take the form
239: \begin{equation}
240: \chi_{sp}^{-1}(q)=\chi_{0}(q)^{-1}-\frac12 U_{sp} \label{Chi_sp}%
241: \end{equation}
242: and
243: \begin{equation}
244: \chi_{ch}^{-1}(q)=\chi_{0}(q)^{-1}+\frac12 U_{ch}%
245: \end{equation}
246: with $\chi_{0}$ computed with the Green function $G_{\sigma}^{(1)}$ that
247: contains the self-energy whose functional differentiation gave the vertices.
248: This self-energy is constant, corresponding to the Hartree-Fock-type
249: factorization.\footnote{The constant self-energy is \textit{not} equal to
250: $Un_{-\sigma}$ as it would in the trivial Hartree-Fock factorization (see
251: previous note).} The susceptibilities thus satisfy conservation
252: laws\cite{Baym:1962}. One enforces the Pauli
253: principle $\langle n_{\sigma}^{2}\rangle=\langle n_{\sigma}\rangle$ implicit
254: in the following two sum rules,
255: \begin{align}
256: \frac{T}{N}\sum_{q}\chi_{sp}(q) & =\left\langle (n_{\uparrow}-n_{\downarrow
257: })^{2}\right\rangle =n-2\langle n_{\uparrow}n_{\downarrow}\rangle
258: \label{Suscep}\\
259: \frac{T}{N}\sum_{q}\chi_{ch}(q) & =\left\langle (n_{\uparrow}+n_{\downarrow
260: })^{2}\right\rangle -n^{2}=n+2\langle n_{\uparrow}n_{\downarrow}\rangle
261: -n^{2}\nonumber
262: \end{align}
263: where $n$ is the density. The above equations, in addition to\cite{Vilk:1994}
264: \begin{equation}
265: U_{sp}=\frac{U\langle n_{\uparrow}n_{\downarrow}\rangle}{\langle n_{\uparrow
266: }\rangle\langle n_{\downarrow}\rangle}, \label{ansatz}%
267: \end{equation}
268: suffice to determine the constant vertices $U_{sp}$ and $U_{ch}$.
269:
270: Once the two-particle quantities have been found as above, the next step of
271: the approach of Ref.~\onlinecite{Vilk:1997}, consists in improving the approximation
272: for the single-particle self-energy by starting from an exact expression where
273: the high-frequency Hartree-Fock behavior is explicitly factored out. One then
274: substitutes in the exact expression the irreducible low-frequency vertices
275: $U_{sp}$ and $U_{ch}$ as well as $G_{\sigma}^{(1)}(k+q)$ and $\chi
276: _{sp}(q),\chi_{ch}(q)$ computed above. The exact form for the self-energy
277: expression can however be obtained either in the longitudinal or in the
278: transverse channel. To satisfy crossing symmetry of the fully reducible vertex
279: appearing in the general expression and to preserve consistency between one-
280: and two-particle quantities, one averages the two possibilities to obtain
281: \cite{Moukouri:2000}%
282: \begin{align}
283: & \Sigma_{\sigma}^{(2)}(k)=Un_{-\sigma}\nonumber\label{Self-long}\\
284: & ~~+\frac{U}{8}\frac{T}{N}\sum_{q}\left[3U_{sp}\chi_{sp}(q)+U_{ch}%
285: \chi_{ch}(q)\right] G_{\sigma}^{(1)}(k+q).
286: \end{align}
287: The resulting self-energy $\Sigma_{\sigma}^{(2)}(k)$ on the left hand-side is
288: at the next level of approximation so it differs from the self-energy entering
289: the right-hand side. One can verify that the longitudinal spin fluctuations
290: contribute an amount $U\langle n_{\uparrow}n_{\downarrow}\rangle/4$ to the
291: consistency condition\cite{Vilk:1996} $\frac{1}{2}\mathrm{Tr}(\Sigma
292: ^{(2)}G^{(1)})=$ $U\langle n_{\uparrow}n_{\downarrow}\rangle$ and that each of
293: the two transverse spin components as well as the charge fluctuations also
294: each contribute $U\langle n_{\uparrow}n_{\downarrow}\rangle/4.$ In addition,
295: one verifies numerically that the exact sum rule\cite{Vilk:1997} $-\int
296: d\omega^{\prime}\operatorname{Im}[\Sigma_{\sigma}(\mathbf{k,}\omega^{\prime
297: })]/\pi=U^{2}n_{-\sigma}(1-n_{-\sigma})$ determining the high-frequency
298: behavior is satisfied to a high degree of accuracy.
299:
300: The theory also has a consistency check. Indeed, the exact expression for
301: consistency between one- and two-particle quantities should be written with
302: $G^{(2)}$ given by $(G^{-1})^{(2)}=(G^{-1})^{(0)}-\Sigma^{(2)}$ instead of
303: with $G^{(1)}$. In other words $\frac{1}{2}\mathrm{Tr}(\Sigma^{(2)}G^{(2)})=$
304: $U\langle n_{\uparrow}n_{\downarrow}\rangle$ should be satisfied instead of
305: $\frac{1}{2}\mathrm{Tr}(\Sigma^{(2)}G^{(1)})=$ $U\langle n_{\uparrow
306: }n_{\downarrow}\rangle,$ which is exactly satisfied here. We find through
307: QMC benchmarks that when the left- and right-hand side of the last equation differ
308: only by a few percent, then the theory is accurate.
309:
310: To obtain the thermodynamics, one finds the entropy by integrating $1/T$ times the
311: specific heat $(\partial E/\partial T)$ so that we know $F=E-TS$. There are
312: other ways to obtain the thermodynamics and one looks for consistency between
313: these.\cite{Roy:unpub} We will not discuss thermodynamic aspects in the
314: present review.
315:
316: At weak coupling in the repulsive model the particle-hole channel is the one
317: that is influenced directly. Correlations in crossed channels, such as pairing
318: susceptibilities, are induced indirectly and are harder to evaluate. This
319: simply reflects the fact the simplest Hartree-Fock factorization of the
320: Hubbard model does not lead to a d-wave order parameter (even though
321: Hartree-Fock factorization of its strong-coupling version does). The
322: $d_{x^{2}-y^{2}}$-wave susceptibility is defined by $\chi_{d}=\int_{0}^{\beta
323: }d\tau\langle T_{\tau}\Delta( \tau) \Delta^{\dagger}\rangle$ with the $d$-wave
324: order parameter equal to $\Delta^{\dagger}=\sum_{i}\sum_{\gamma}g( \gamma)
325: c_{i\uparrow}^{\dagger} c_{i+\gamma\downarrow}^{\dagger}$ the sum over
326: $\gamma$ being over nearest-neighbors, with $g( \gamma) =\pm1/2$ depending on
327: whether $\gamma$ is a neighbor along the $\hat{\mathbf{x}}$ or the
328: $\hat{\mathbf{y}}$ axis. Briefly speaking,\cite{Kyung:2003, Allen:unpub}
329: to extend TPSC to compute pairing susceptibility, we begin from the
330: Schwinger-Martin-Kadanoff-Baym formalism with both
331: diagonal\cite{Vilk:1997,Allen:2003} and off-diagonal\cite{Allen:2001} source
332: fields. The self-energy is expressed in terms of spin and charge fluctuations
333: and the irreducible vertex entering the Bethe-Salpeter equation for the
334: pairing susceptibility is obtained from functional differentiation. The final
335: expression for the $d$-wave susceptibility is,
336: %TCIMACRO{\TeXButton{%
337: %\begin{widetext}
338: %}{\begin{widetext}} }%
339: %BeginExpansion
340: \begin{widetext}
341: %EndExpansion
342: \begin{align}
343: \chi_{d}( \mathbf{q}=0,iq_{n}=0) & =\frac{T}{N}\sum_{k}\left( g_{d}%
344: ^{2}(\mathbf{k}) G_{\uparrow}^{(2)}( -k) G_{\downarrow}^{(2)}(k)\right)
345: -\frac{U}{4}\left( \frac{T}{N}\right) ^{2}\sum_{k,k^{\prime}}g_{d}(
346: \mathbf{k}) G_{\uparrow}^{( 2) }( -k) G_{\downarrow}^{( 2) }( k)\nonumber\\
347: & \times\left( \frac{3}{1-\frac12 U_{sp}\chi_{0}( k^{\prime}-k)} +\frac
348: {1}{1+\frac12 U_{ch}\chi_{0}(k^{\prime}-k)}\right) G_{\uparrow}^{( 1) }(
349: -k^{\prime}) G_{\downarrow}^{(1)}(k^{\prime}) g_{d}(\mathbf{k}^{\prime}).
350: \label{Suscep_d}%
351: \end{align}%
352: %TCIMACRO{\TeXButton{%
353: %\end{widetext}
354: %}{\end{widetext}}}%
355: %BeginExpansion
356: \end{widetext}%
357: %EndExpansion
358: In the above expression, $g_{d}( \mathbf{k}) $ is the
359: form factor for the gap symmetry, while $k$ and $k^{\prime}$ stand for both
360: wave-vector and fermionic Matsubara frequencies on a square-lattice with $N$
361: sites at temperature $T.$ The spin and charge susceptibilities take the form
362: $\chi_{sp}^{-1}( q) =\chi_{0}(q)^{-1}-\frac12 U_{sp}$ and $\chi_{ch}^{-1}(
363: q) =$ $\chi_{0}(q)^{-1}+\frac12 U_{ch}$ with $\chi_{0}$ computed with the
364: Green function $G_{\sigma}^{(1)}$ that contains the self-energy whose
365: functional differentiation gave the spin and charge vertices. The values of
366: $U_{sp},$ $U_{ch}$ and $\langle n_{\uparrow}n_{\downarrow}\rangle$ are
367: obtained\ as described above. In the pseudogap regime, one cannot use
368: $U_{sp}=U\langle n_{\uparrow}n_{\downarrow}\rangle/( \langle n_{\uparrow
369: }\rangle\langle n_{\downarrow}\rangle)$. Instead,\cite{Vilk:1997} one uses the
370: local-moment sum rule with the zero temperature value of $\langle n_{\uparrow
371: }n_{\downarrow}\rangle$ obtained by the method of Ref.~\onlinecite{Lilly:1990}
372: that agrees very well with QMC calculations at all values of $U.$ Also,
373: $G_{\sigma}^{(2)}$ contains self-energy effects coming from spin and charge
374: fluctuations, as described above.\cite{Moukouri:2000, Allen:2003}
375:
376: The same principles and methodology can be applied for the attractive Hubbard
377: model.\cite{Kyung:2001, Allen:2001, Allen:unpub} In that case, the dominant
378: channel is the s-wave pairing channel. Correlations in the crossed channel,
379: namely the spin and charge susceptibilities, can also be obtained
380: \textit{mutatis mutandi }along the lines of the previous paragraph.
381:
382: \subsubsection{Benchmarks for TPSC}
383:
384: To test any non-perturbative approach, we need reliable benchmarks. Quantum
385: Monte Carlo (QMC) simulations provide such benchmarks. The results of such numerical
386: calculations are unbiased and they can be obtained on much larger system sizes
387: than any other simulation method. The statistical uncertainty can be made as
388: small as required. The drawback of QMC is that the sign problem renders
389: calculations impossible at temperatures low enough to reach those that are
390: relevant for d-wave superconductivity. Nevertheless, QMC can be performed in
391: regimes that are non-trivial enough to allow us to eliminate some theories on
392: the grounds that they give qualitatively incorrect results. Comparisons with
393: QMC allow us to estimate the accuracy of the theory. An approach like TPSC can
394: then be extended to regimes where QMC is unavailable with the confidence
395: provided by agreement between both approaches in regimes where both can be performed.
396:
397: %-----------------------------------------------------------------------
398: %: FIG:fig_25
399: \begin{figure}[ptb]
400: \centerline{\includegraphics[width=5.0cm]{f25}} \caption{Comparisons between
401: the QMC simulations (symbols) and TPSC (solid lines) for the filling
402: dependence of the double occupancy. The results are for $T=t/6$ as a function
403: of filling and for various values of $U$ expect for $U=4t$ where the dashed
404: line shows the results of our theory at the crossover temperature $T=T_{X}$.
405: From Ref.~\onlinecite{Vilk:1997}.}%
406: \label{fig_25}%
407: \end{figure}
408: %-----------------------------------------------------------------------
409:
410:
411: %-----------------------------------------------------------------------
412: %: FIG:fig_26
413: \begin{figure}[ptb]
414: \centerline{\includegraphics[width=7cm]{f26}}
415: %and $T=0.2$. }
416: \caption{Wave vector ($\mathbf{q}$) dependence of the spin and charge structure
417: factors for different sets of parameters. Solid lines are from TPSC and
418: symbols are our QMC data. Monte Carlo data for $n=1$ and $U=8t$ are
419: for $6 \times6$ clusters and $T=0.5t$; all other data are for $8 \times8$
420: clusters and $T=0.2t$. Error bars are shown only when significant. From
421: Ref.~\onlinecite{Vilk:1994}.}%
422: \label{fig_26}%
423: \end{figure}
424: %-----------------------------------------------------------------------
425:
426:
427: %-----------------------------------------------------------------------
428: %: FIG:fig_27
429: \begin{figure}[ptb]
430: \centerline{\includegraphics[width=5.5cm]{f27}} \caption{Temperature
431: dependence of $S_{sp}(\pi,\pi)$ at half-filling $n=1$. The solid line is from
432: TPSC and symbols are Monte Carlo data from Ref.~\onlinecite{White:1989}. Taken
433: from Ref.~\onlinecite{Vilk:1994}.}%
434: \label{fig_27}%
435: \end{figure}
436: %-----------------------------------------------------------------------
437:
438:
439: %-----------------------------------------------------------------------
440: %: FIG:fig_28
441: \begin{figure}[ptb]
442: \centerline{\includegraphics[width=5.5cm]{f28}} \caption{Comparisons between
443: Monte Carlo simulations (BW), FLEX calculations and TPSC for the spin
444: susceptibility at $Q=(\pi,\pi)$ as a function of temperature at zero Matsubara
445: frequency. The filled circles (BWS) are from Ref.~\onlinecite{Bulut:1993}.
446: Taken from Ref.~\onlinecite{Vilk:1997}.}%
447: \label{fig_28}%
448: \end{figure}
449: %-----------------------------------------------------------------------
450:
451:
452: In order to be concise, details are left to figure captions. Let us first
453: focus on quantities related to spin and charge fluctuations. The symbols on
454: the figures refer to QMC results while the solid lines come from TPSC
455: calculations. Fig.~\ref{fig_25} shows double occupancy, a quantity that plays
456: a very important role in the Hubbard model in general and in TPSC in
457: particular. That quantity is shown as a function of filling for various values
458: of $U$ at inverse temperature $\beta=6$. Fig.~\ref{fig_26} displays the spin
459: and charge structure factors in a regime where size effects are not important.
460: Clearly the results are non-perturbative given the large difference between
461: the spin and charge structure factors, which are plotted here in units where
462: they are equal at $U=0$. In Fig.~\ref{fig_27} we exhibit the static structure
463: factor at half-filling as a function of temperature. Below the crossover
464: temperature $T_{X}$, there is an important size dependence in the QMC results.
465: The TPSC calculation, represented by a solid line, is done for an infinite
466: system. We see that the mean-field finite transition temperature $T_{MF}$ is
467: replaced by a crossover temperature $T_{X}$ at which the correlations enter an
468: exponential growth regime. One can show analytically\cite{Vilk:1994,
469: Vilk:1997} that the correlation length becomes infinite only at zero
470: temperature, thus satisfying the Mermin-Wagner theorem.
471: The QMC results approach the TPSC results as the system size
472: grows. Nevertheless, TPSC is in the $N=\infty$ universality
473: class\cite{Dare:1996} contrary to the Hubbard model for which $N=3,$ so one
474: expects quantitative differences to increase as the correlation length becomes
475: larger. It is important to note that $T_{X}$ does not coincide with the
476: mean-field transition temperature $T_{MF}$. This is because of
477: Kanamori-Brueckner screening\cite{Chen:1991, Vilk:1994} that manifests itself
478: in the difference between $U_{sp}$ and the bare $U$. Below $T_{X}$, the main
479: contribution to the static spin structure factor in Fig.~\ref{fig_27} comes
480: from the zero-Matsubara frequency component of the spin susceptibility. This
481: is the so-called renormalized classical regime where the characteristic spin
482: fluctuation frequency $\omega_{sp}$ is much less than temperature. Even at
483: temperatures higher than that, TPSC agrees with QMC calculation much better
484: than other methods, as shown in Fig.~\ref{fig_28}.
485:
486: %-----------------------------------------------------------------------
487: %: FIG:fig_29
488: \begin{figure}[ptb]
489: \centerline{\includegraphics[width=6.5cm]{f29}}
490: \caption{Single particle spectral weight $A(\mathbf{k},\omega)$ for $U=4t$,
491: $\beta=5/t$, $n=1$, and all independent wave vectors $\mathbf{k}$ of an $8
492: \times8$ lattice. Results obtained from Maximum Entropy inversion of QMC data
493: on the left panel and many-body TPSC calculations with Eq.(~\ref{Self-long}) on the
494: middle panel and with FLEX on the right panel. From
495: Ref.~\onlinecite{Moukouri:2000}.}%
496: \label{fig_29}%
497: %
498: %FLEX on the right panel. (Relative error in all cases is about 0.3\%).}
499: \end{figure}
500: %-----------------------------------------------------------------------
501:
502:
503: %-----------------------------------------------------------------------
504: %: FIG:fig_30
505: \begin{figure}[ptb]
506: \centerline{\includegraphics[width=7.5cm]{f30}}
507: \caption{Size dependent
508: results for various types of calculations for $U=4t$, $\beta=5/t$, $n=1$,
509: $\mathbf{k}=(0,\pi)$, $L=4,6,8,10$. Upper panels show $A(\mathbf{k},\omega)$
510: extracted from Maximum Entropy on $G(\tau)$ shown on the corresponding lower
511: panels. (a) QMC. (b) TPSC using Eq.~(\ref{Self-long}). (c) FLEX. From
512: Ref.~\onlinecite{Moukouri:2000}.}%
513: \label{fig_30}%
514: \end{figure}
515: %-----------------------------------------------------------------------
516:
517:
518: Below the crossover temperature to the renormalized classical regime, a
519: pseudogap develops in the single-particle spectral weight. This is illustrated
520: in Fig.~\ref{fig_29}.\cite{Moukouri:2000} Eliashberg-type approaches such as
521: FLEX do not show the pseudogap present in QMC. The size dependence of the
522: results is also quite close in TPSC and in QMC, as shown in Fig.~\ref{fig_30}.
523:
524: %-----------------------------------------------------------------------
525: %: FIG:fig_12
526: \begin{figure}[ptb]
527: \centerline{\includegraphics[width=6.5cm]{f12}}\caption{Comparisons between
528: the $d_{x^{2}-y^{2}}$ susceptibility obtained from QMC simulations (symbols)
529: and from the TPSC approach (lines) in the two-dimensional Hubbard model. Both
530: calculations are for $U=4t$, a $6\times6$ lattice. QMC error bars are smaller
531: than the symbols. Analytical results are joined by solid lines. The size
532: dependence of the results is small at these temperatures. The $U=0$ case is
533: also shown at $\beta=4/t$ as the upper line. The inset compares QMC and FLEX
534: at $U=4,\beta=4/t$. From Ref.~\onlinecite{Kyung:2003}.}%
535: \label{fig_12}%
536: \end{figure}
537: %-----------------------------------------------------------------------
538:
539:
540: The d-wave susceptibility\cite{Kyung:2003} shown in Fig.~\ref{fig_12} again
541: clearly demonstrates the agreement between TPSC and QMC. In particular, the
542: dome shape dependence of the QMC results is reproduced to within a few
543: percent. We will see in Sec. \ref{Results} how one understands the dome shape
544: and the fact that the d-wave susceptibility of the interacting system is
545: smaller than that of the non-interacting one in this temperature range.
546:
547: %-----------------------------------------------------------------------
548: %: FIG:fig_32
549: \begin{figure}[ptb]
550: \centerline{\includegraphics[width=8.5cm]{f32}} \caption{The crossover diagram
551: as a function of next-nearest-neighbor hopping $t^{\prime}$ from TPSC (left)
552: and from a temperature cutoff renormalization group technique from
553: Ref.~\onlinecite{Honerkamp:2001} (right). The corresponding Van Hove filling
554: is indicated on the upper horizontal axis. Crossover lines for magnetic
555: instabilities near the antiferromagnetic and ferromagnetic wave vectors are
556: represented by filled symbols while open symbols indicate instability towards
557: $d_{x^{2}-y^{2}}$-wave superconducting. The solid and dashed lines below the
558: empty symbols show, respectively for $U=3t$ and $U=6t$, where the
559: antiferromagnetic crossover temperature would have been in the absence of the
560: superconducting instability. The largest system size used for this calculation
561: is $2048\times2048$. From Ref.~\onlinecite{Hankevych:2003}.}%
562: \label{fig_32}%
563: \end{figure}
564: %-----------------------------------------------------------------------
565: %-----------------------------------------------------------------------
566: %: FIG:fig_33
567: \begin{figure}[ptb]
568: \centerline{\includegraphics[width=6.0cm]{f33}} \caption{TPSC s-wave paring
569: structure factor $S(\mathbf{q},\tau=0)$ (filled triangles)
570: and QMC $S(\mathbf{q},\tau=0)$ (open circles) for $U=-4t$ and various temperatures (a) at
571: $n=0.5$ and (b) at $n=0.8$ on a $8\times8$ lattice. The dashed lines are to
572: guide the eye. From Ref.~\onlinecite{Kyung:2001}.}%
573: \label{fig_33}%
574: \end{figure}
575: %-----------------------------------------------------------------------
576: %-----------------------------------------------------------------------
577: %: FIG:fig_34
578: \begin{figure}[ptb]
579: \centerline{\includegraphics[width=8.5cm]{f34}} \caption{Left: chemical
580: potential shifts $\mu^{(1)}-\mu_{0}$ (open diamonds) and $\mu^{(2)}-\mu_{0}$
581: (open squares) with the results of QMC calculations (open circles) for
582: $U=-4t$. Right: The momentum dependent occupation number $n(\mathbf{k})$.
583: Circles: QMC calculations from Ref.~\onlinecite{Trivedi:1995}. The solid
584: curve: TPSC. The dashed curve obtained by replacing $U_{pp}$ by $U$ in the
585: self-energy with all the rest unchanged. The long-dash line is the result of a
586: self-consistent T-matrix calculation, and the dot-dash line the result of
587: second-order perturbation theory. From Ref.~\onlinecite{Kyung:2001}.}%
588: \label{fig_34}%
589: \end{figure}
590: %-----------------------------------------------------------------------
591: %-----------------------------------------------------------------------
592: %: FIG:fig_35
593: \begin{figure}[ptb]
594: \centerline{\includegraphics[width=8.5cm]{f35}} \caption{Comparisons of local
595: density of states and single-particle spectral weight from TPSC (solid lines)
596: and QMC (dashed lines) on a $8\times8$ lattice. QMC data for the density of
597: states taken from Ref. \onlinecite{Moreo:1992}. Figures from
598: Ref.~\onlinecite{Kyung:2001}.}%
599: \label{fig_35}%
600: \end{figure}
601: %-----------------------------------------------------------------------
602:
603:
604: To conclude this section, we quickly mention a few other results obtained with
605: TPSC. Fig.~\ref{fig_32} contrasts the crossover phase diagram obtained for the
606: Hubbard model at the van Hove filling\cite{Hankevych:2003} with the results of
607: a renormalization group calculation.\cite{Honerkamp:2001} The difference
608: occurring in the ferromagnetic region is discussed in detail in
609: Ref.~\onlinecite{Hankevych:2003}. Finally, we point out various comparisons
610: for the attractive Hubbard model. Fig.~\ref{fig_33} shows the static s-wave
611: pairing susceptibility, Fig.~\ref{fig_34} the chemical potential and the
612: occupation number, and finally Fig.~\ref{fig_35} the local density of states
613: and the single-particle spectral weight at a given wave vector.
614:
615: \subsection{Strong-coupling approaches: Quantum clusters}
616:
617: DMFT\cite{Georges:1992, Jarrell:1992} has been extremely successful in helping
618: us understand the Mott transition, the key physical phenomenon that manifests
619: itself at strong coupling. However, in high dimension where this theory
620: becomes exact, spatial fluctuations associated with incipient order do not
621: manifest themselves in the self-energy. In low dimension, this is not the
622: case. The self-energy has strong momentum dependence, as clearly shown
623: experimentally in the high-temperature superconductors, and theoretically in
624: the TPSC approach, a subject we shall discuss again below. It is thus
625: necessary to go beyond DMFT by studying clusters instead of a single Anderson
626: impurity as done in DMFT. The simplest cluster approach that includes
627: strong-coupling effects and momentum dependence is Cluster Perturbation Theory
628: (CPT).\cite{Gros:1993, Senechal:2000} In this approach, an infinite set of
629: disconnected clusters are solved exactly and then connected to each other
630: using strong-coupling perturbation theory. Although the resulting theory turns
631: out to give the exact result in the $U=0$ case, its derivation clearly shows
632: that one expects reliable results mostly at strong coupling. This approach
633: does not include the self-consistent effects contained in DMFT.
634: Self-consistency or clusters was suggested in Ref. \onlinecite{Georges:1996,Schiller:1995} and a causal approach was first implemented within DCA,\cite{Hettler:1998} where a
635: momentum-space cluster is connected to a self-consistent momentum-space
636: medium. In our opinion, the best framework to understand all other cluster
637: methods is the Self-Energy Functional approach of
638: Potthoff.\cite{Potthoff:2003a, Potthoff:2003} The form of the lattice Green function obtained
639: in this approach is the same as that obtained in CPT, clearly exhibiting that
640: such an approach is better at strong-coupling, even though results often
641: extrapolate correctly to weak coupling. Amongst the special cases of this
642: approach, the Variational Cluster Approach (VCA), or Variational Cluster
643: Perturbation Theory (VCPT)\cite{Potthoff:2003} is the one closest to the
644: original approach. In a variant, Cellular Dynamical Mean Field
645: Theory\cite{Kotliar:2001} (CDMFT), a cluster is embedded in a self-consistent
646: medium instead of a single Anderson impurity as in DMFT (even though the latter approach is accurate in many realistic cases, especially in three dimensions). The strong-coupling
647: aspects of CDMFT come out clearly in
648: Refs.~\onlinecite{Stanescu:2004, Stanescu:2005}. A detailed review of quantum
649: cluster methods has appeared in Ref.~\onlinecite{Maier:2005}.
650:
651: \subsubsection{Cluster perturbation theory}
652:
653: Even though CPT does not have the self-consistency present in DMFT type
654: approaches, at fixed computing resources it allows for the best momentum
655: resolution. This is particularly important for the ARPES pseudogap in
656: electron-doped cuprates that has quite a detailed momentum space structure,
657: and for d-wave superconducting correlations where the zero temperature pair
658: correlation length may extend well beyond near-neighbor sites. CPT was
659: developed by Gros\cite{Gros:1993} and S\'{e}n\'{e}chal\cite{Senechal:2000}
660: independently. This approach can be viewed as the first term of a systematic
661: expansion around strong coupling.\cite{Senechal:2002} Let us write the hopping
662: matrix elements in the form
663: \begin{equation}
664: t_{\mu\nu}^{mn}=t_{\mu\nu}^{(c)}\delta_{mn}+V_{\mu\nu}^{mn}%
665: \end{equation}
666: where $m$ and $n$ label the different clusters, and $\mu,\nu$ label the sites
667: within a cluster. Then $t_{\mu\nu}^{(c)}$ labels all the hopping matrix
668: elements within a cluster and the above equation defines $V_{\mu\nu}^{mn}$.
669:
670: We pause to introduce the notation that will be used throughout for quantum
671: cluster methods. We follow the review article Ref.~\onlinecite{Maier:2005}. In
672: reciprocal space, any wave vector $\mathbf{k}$ in the Brillouin zone may be
673: written as $\mathbf{k=}\tilde{\mathbf{k}}+\mathbf{K}$ where both
674: $\mathbf{k}$ and $\tilde{\mathbf{k}}$ are continuous in the infinite size
675: limit, except that $\tilde{\mathbf{k}}$ is defined only in the reduced
676: Brillouin zone that corresponds to the superlattice. On the other hand,
677: $\mathbf{K}$ is discrete and denotes reciprocal lattice vectors of the
678: superlattice. By analogy, any position $\mathbf{r}$ in position space can be
679: written as $\tilde{\mathbf{r}}+\mathbf{R}$ where $\mathbf{R}$ is for
680: positions within clusters while $\tilde{\mathbf{r}}$ labels the origins of
681: the clusters, an infinite number of them. Hence, Fourier's theorem allows one
682: to define functions of $\mathbf{k,}$ $\tilde{\mathbf{k}}$ or $\mathbf{K}$
683: that contain the same information as functions of, respectively, $\mathbf{r,}$
684: $\tilde{\mathbf{r}}$ or $\mathbf{R.}$ Also, we have $\mathbf{K\cdot}$
685: $\tilde{\mathbf{r}}=2\pi n$ where $n$ is an integer. Sites within a
686: cluster are labelled by greek letters so that the position of site $\mu$
687: within a cluster is $\mathbf{R}_{\mu}$, while clusters are labelled by Latin
688: letters so that the origin of cluster $m$ is at $\tilde{\mathbf{r}}_{m}$.
689:
690: Returning to CPT, the Green function for the whole system is given by
691: \begin{equation}
692: \left[\hat{G}^{-1}(\tilde{\mathbf{k}},z)\right] _{\mu\nu}=\left[
693: \hat{G}^{(c)-1}(z)-\hat{V}(\tilde{\mathbf{k}})\right] _{\mu\nu}
694: \label{Gsuperlattice}%
695: \end{equation}
696: where hats denote matrices in cluster site indices and $z$ is the complex
697: frequency. At this level of approximation, the CPT Green function has the same
698: structure as in the Hubbard I approximation except that it pertains to a
699: cluster instead of a single site. Since $\hat{G}^{(c)-1}(z)=z+\mu
700: -\hat{t}^{(c)}-\hat{\Sigma}^{(c)}$ and $\hat{G}^{(0)-1}%
701: (\tilde{\mathbf{k}},z)=z+\mu-\hat{t}^{(c)}-$ $\hat{V}%
702: (\tilde{\mathbf{k}}),$ the Green function (\ref{Gsuperlattice}) may
703: also be written as
704: \begin{equation}
705: \hat{G}^{-1}(\tilde{\mathbf{k}}\mathbf{,}z)=\hat{G}^{(0)-1}%
706: (\tilde{\mathbf{k}}\mathbf{,}z)-\hat{\Sigma}^{(c)}(z).
707: \end{equation}
708: This form allows a different physical interpretation of the approach. In the
709: above expression, the self-energy of the lattice is approximated by the
710: self-energy of the cluster. The latter in real space spans only the size of
711: the cluster.
712:
713: We still need an expression to extend the above result to the lattice in a
714: translationally invariant way.
715: This is done by defining the following residual Fourier transform:
716: \begin{equation}
717: G_{\rm CPT}(\mathbf{k},z)=\frac{1}{N_{c}}\sum_{\mu,\nu}^{N_{c}}e^{i\mathbf{k}\cdot(\mathbf{R}_{\mu}-\mathbf{R}_{\nu})}G_{\mu\nu}(\tilde{\mathbf{k}},z).
718: \label{GCPT}%
719: \end{equation}
720: Notice that $G_{\mu\nu}(\tilde{\mathbf{k}},z)$ may be replaced by
721: $G_{\mu\nu}(\mathbf{k},z)$ in the above equation since $\hat{V}(\tilde{\mathbf{k}}+\mathbf{K}) = \hat{V}(\tilde{\mathbf{k}})$.
722:
723: \subsubsection{Self-energy functional approach}
724:
725: The self-energy functional approach, devised by Potthoff\cite{Potthoff:2003}
726: allows one to consider various cluster schemes from a unified point of view.
727: It begins with $\Omega_{\mathbf{t}}[G],$ a functional of the Green function%
728: \begin{equation}
729: \Omega_{\mathbf{t}}[G]=\Phi[G]-\mathrm{Tr}((G_{0\mathbf{t}}^{-1}%
730: -G^{-1})G)+\mathrm{Tr}\ln(-G). \label{GrandPotential}%
731: \end{equation}
732: The Luttinger Ward functional $\Phi[G]$ entering this equation is the
733: sum of connected vacuum skeleton diagrams. A diagram-free definition of this
734: functional is also given in Ref.~\onlinecite{Potthoff:2004}. For our purposes,
735: what is important is that (1) The functional derivative of $\Phi[G]$ is
736: the self-energy%
737: \begin{equation}
738: \frac{\delta\Phi[G]}{\delta G}=\Sigma\label{SelfLuttinger}%
739: \end{equation}
740: and (2) it is a universal functional of $G$ in the following sense: whatever
741: the form of the one-body Hamiltonian, it depends only on the interaction and,
742: functionnally, it has the same dependence
743: on $G$. The dependence of the functional $\Omega_{\mathbf{t}%
744: }[G]$ on the one-body part of the Hamiltonian is denoted by the subscript
745: $\mathbf{t}$ and it comes only through $G_{0\mathbf{t}}^{-1}$ appearing on the
746: right-hand side of Eq.~(\ref{GrandPotential}).
747:
748: The functional $\Omega_{\mathbf{t}}[G]$ has the important property that it is
749: stationary when $G$ takes the value prescribed by Dyson's equation. Indeed,
750: given the last two equations, the Euler equation takes the form%
751: \begin{equation}
752: \frac{\delta\Omega_{\mathbf{t}}[G]}{\delta G}=\Sigma-G_{0\mathbf{t}}%
753: ^{-1}+G^{-1}=0.
754: \end{equation}
755: This is a dynamic variational principle since it involves the frequency
756: appearing in the Green function, in other words excited states are involved in
757: the variation. At this stationary point, and only there, $\Omega_{\mathbf{t}%
758: }[G]$ is equal to the grand potential. Contrary to Ritz's variational
759: principle, this last equation does not tell us whether $\Omega_{\mathbf{t}%
760: }[G]$ is a minimum or a maximum or a saddle point there.
761:
762: There are various ways to use the stationarity property that we described
763: above. The most common one, is to approximate $\Phi[G]$ by a finite set
764: of diagrams. This is how one obtains the Hartree-Fock, the FLEX
765: approximation\cite{Bickers:1989} or other so-called thermodynamically
766: consistent theories. This is what Potthoff calls a type II
767: approximation strategy.\cite{Potthoff:2005} A type I approximation simplifies the Euler equation
768: itself. In a type III approximation, one uses the exact form of $\Phi[
769: G]$ but only on a limited domain of trial Green functions.
770:
771: Following Potthoff, we adopt the type III approximation on a functional
772: of the self-energy instead of on a functional of the Green function. Suppose
773: we can locally invert Eq.~(\ref{SelfLuttinger}) for the self-energy to write $G$ as a functional of $\Sigma.$ We can use this result to write,
774: \begin{equation}
775: \Omega_{\mathbf{t}}[\Sigma]=F[\Sigma]-\mathrm{Tr}\ln(-G_{0\mathbf{t}}%
776: ^{-1}+\Sigma).
777: \end{equation}
778: where we defined
779: \begin{equation}
780: F[\Sigma]=\Phi[G]-\mathrm{Tr}(\Sigma G).
781: \end{equation}
782: and where it is implicit that $G=G[\Sigma]$ is now a functional of $\Sigma$.
783: $F[\Sigma],$ along with the expression (\ref{SelfLuttinger}) for the derivative of the Luttinger-Ward functional, define the Legendre
784: transform of the Luttinger-Ward functional. It is easy to verify that%
785: \begin{equation}
786: \frac{\delta F[\Sigma]}{\delta\Sigma}=\frac{\delta\Phi[G]}{\delta
787: G}\frac{\delta G[\Sigma]}{\delta\Sigma}-\Sigma\frac{\delta G[\Sigma]}%
788: {\delta\Sigma}-G=-G
789: \end{equation}
790: hence, $\Omega_{\mathbf{t}}[\Sigma]$ is stationary with respect to $\Sigma$
791: when Dyson's equation is satisfied%
792: \begin{equation}
793: \frac{\delta\Omega_{\mathbf{t}}[\Sigma]}{\delta\Sigma}=-G+(G_{0\mathbf{t}}%
794: ^{-1}-\Sigma)^{-1}=0.
795: \end{equation}
796:
797:
798: %-----------------------------------------------------------------------
799: %: FIG:fig_42
800: \begin{figure}[ptb]
801: \centerline{\includegraphics[width=8.5cm]{f42b}} \caption{Various tilings used
802: in quantum cluster approaches. In these examples the grey and white sites are
803: inequivalent since an antiferromagnetic order is possible. }%
804: \label{fig_42}%
805: \end{figure}
806: %-----------------------------------------------------------------------
807:
808:
809: To perform a type III approximation on $F[\Sigma]$, we take advantage that it
810: is universal, i.e., that it depends only on the interaction part of the
811: Hamiltonian and not on the one-body part. This follows from the universal
812: character of its Legendre transform $\Phi[G]$. We thus evaluate
813: $F[\Sigma]$ exactly for a Hamiltonian $H^{\prime}$ that shares the same
814: interaction part as the Hubbard Hamiltonian, but that is exactly solvable.
815: This Hamiltonian $H^{\prime}$ is taken as a cluster decomposition of the
816: original problem, i.e., we tile the infinite lattice into identical,
817: disconnected clusters that can be solved exactly. Examples of such tilings are
818: given in Fig.~\ref{fig_42}. Denoting the corresponding quantities with a
819: prime, we obtain,
820: \begin{equation}
821: \Omega_{\mathbf{t}^{\prime}}[\Sigma^{\prime}]=F[\Sigma^{\prime}]-\mathrm{Tr}%
822: \ln(-G_{0\mathbf{t}^{\prime}}^{-1}+\Sigma^{\prime}).
823: \end{equation}
824: from which we can extract $F[\Sigma^{\prime}]$. It follows that
825: \begin{equation}
826: \Omega_{\mathbf{t}}[\Sigma^{\prime}]=\Omega_{\mathbf{t}^{\prime}}%
827: [\Sigma^{\prime}]+\mathrm{Tr}\ln(-G_{0\mathbf{t}^{\prime}}^{-1}+\Sigma
828: ^{\prime})-\mathrm{Tr}\ln(-G_{0\mathbf{t}}^{-1}+\Sigma^{\prime}).
829: \label{sef_eq}%
830: \end{equation}
831: The type III approximation comes from the fact that the self-energy
832: $\Sigma^{\prime}$ is restricted to the exact self-energy of the cluster
833: problem $H^{\prime}$, so that variational parameters appear in the definition
834: of the one-body part of $H^{\prime}$.
835:
836: In practice, we look for values of the cluster one-body parameters
837: $\mathbf{t}^{\prime}$ such that $\delta\Omega_{\mathbf{t}}[\Sigma^{\prime
838: }]/\delta\mathbf{t}^{\prime}=0$. It is useful for what follows to write the
839: latter equation formally, although we do not use it in actual calculations.
840: Given that $\Omega_{\mathbf{t}^{\prime}}[\Sigma^{\prime}]$ is the actual grand
841: potential evaluated for the cluster, $\partial\Omega_{\mathbf{t}^{\prime}%
842: }[\Sigma^{\prime}]/\partial\mathbf{t}^{\prime}$ is canceled by the explicit
843: $\mathbf{t}^{\prime}$ dependence of $\mathrm{Tr}\ln(-G_{0\mathbf{t}^{\prime}%
844: }^{-1}+\Sigma^{\prime})$ and we are left with
845: \begin{align}
846: 0 & =\frac{\delta\Omega_{\mathbf{t}}[\Sigma^{\prime}]}{\delta\Sigma^{\prime
847: }}\frac{\delta\Sigma^{\prime}}{\delta\mathbf{t}^{\prime}}\nonumber\\
848: & =-\mathrm{Tr}\left[\left( \frac{1}{G_{0\mathbf{t}^{\prime}}^{-1}%
849: -\Sigma^{\prime}}-\frac{1}{G_{0\mathbf{t}}^{-1}-\Sigma^{\prime}}\right)
850: \frac{\delta\Sigma^{\prime}}{\delta\mathbf{t}^{\prime}}\right] .
851: \end{align}
852: Given that the clusters corresponding to $\mathbf{t}^{\prime}$ are
853: disconnected and that translation symmetry holds on the superlattice of
854: clusters, each of which contains $N_c$ sites,
855: the last equation may be written
856: \begin{align}
857: & \sum_{\omega_{n}}\sum_{\mu\nu}\bigg[\frac{N}{N_{c}}\left( \frac
858: {1}{G_{0\mathbf{t}^{\prime}}^{-1}-\Sigma^{\prime}(i\omega_{n})}\right)
859: _{\mu\nu}\nonumber\\
860: & ~~ -\sum_{\tilde{\mathbf{k}}}\left( \frac{1}{G_{0\mathbf{t}}%
861: ^{-1}(\tilde{\mathbf{k}})-\Sigma^{\prime}(i\omega_{n})}\right) _{\mu\nu
862: }\bigg]\frac{\delta\Sigma_{\nu\mu}^{\prime}(i\omega_{n})}{\delta
863: \mathbf{t}^{\prime}}=0. \label{EulerVCA}%
864: \end{align}
865:
866:
867: \subsubsection{Variational cluster perturbation theory, or variational cluster
868: approximation}
869:
870: In Variational Cluster Perturbation Theory (VCPT), more aptly named the
871: Variational Cluster Approach (VCA), solutions to the Euler equations
872: (\ref{EulerVCA}) are found by looking for numerical minima (or more generally,
873: saddle-points) of the functional. Typically, the VCA
874: cluster Hamiltonian $H^{\prime}$ will have the same form as $H$ except that
875: there is no hopping between clusters and that long-range order is allowed by
876: adding some Weiss fields, for instance like in Eq.~(\ref{weiss_eq}) below. The
877: hopping terms and chemical potential within $H^{\prime}$ may also be treated
878: like additional variational parameters. In contrast with Mean-Field theory,
879: these Weiss fields are not mean fields, in the sense that they do not coincide
880: with the corresponding order parameters. The interaction part of $H$ (or $H$')
881: is not factorized in any way and short-range correlations are treated exactly.
882: In fact, the Hamiltonian $H$ is not altered in any way; the Weiss fields are
883: introduced to let the variational principle act on a space of self-energies
884: that includes the possibility of specific long-range orders, without imposing
885: those orders. Indeed, the more naturally an order arises in the system, the
886: smaller the Weiss field needs to be, and one observes that the strength of the
887: Weiss field at the stationary point of the self-energy functional generally
888: decreases with increasing cluster size, as it should since in the thermodynamic limit no
889: Weiss field should be necessary to establish order.
890:
891: \subsubsection{Cellular dynamical mean-field theory}
892:
893: The Cellular dynamical mean-field theory (CDMFT) is obtained by including in
894: the cluster Hamiltonian $H^{\prime}$ a bath of uncorrelated electrons that
895: somehow must mimic the effect on the cluster of the rest of the lattice.
896: Explicitly, $H^{\prime}$ takes the form
897: \begin{align}
898: H^{\prime} & =-\sum_{\mu,\nu,\sigma}t_{\mu\nu}^{\prime}c_{\mu\sigma}%
899: ^{\dagger}c_{\nu\sigma}+U\sum_{\mu}n_{\mu\uparrow}n_{\mu\downarrow}\nonumber\\
900: & ~~+\sum_{\mu,\alpha,\sigma}V_{\mu\alpha}(c_{\mu\sigma}^{\dagger}%
901: a_{\alpha\sigma}+\mathrm{H.c.})+\sum_{\alpha}\epsilon_{\alpha}a_{\alpha\sigma
902: }^{\dagger}a_{\alpha\sigma}%
903: \end{align}
904: where $a_{\alpha\sigma}$ annihilates an electron of spin $\sigma$ on a bath
905: orbital labelled $\alpha$. The bath is characterized by the energy of each
906: orbital ($\epsilon_{\alpha}$) and the bath-cluster hybridization matrix
907: $V_{\mu\alpha}$. This representation of the environment through an Anderson
908: impurity model was introduced in Ref.~\onlinecite{Caffarel:1994} in the
909: context of DMFT (i.e., a single site). The effect of the bath on the electron
910: Green function is encapsulated in the so-called hybridization function
911: \begin{equation}
912: \Gamma_{\mu\nu}(\omega)=\sum_{\alpha}{\frac{V_{\mu\alpha}V_{\nu\alpha}^{\ast}%
913: }{\omega-\epsilon_{\alpha}}}%
914: \end{equation}
915: which enters the Green function as
916: \begin{equation}
917: [G^{\prime-1}]_{\mu\nu}=\omega+\mu-t_{\mu\nu}^{\prime}-\Gamma_{\mu\nu
918: }(\omega)-\Sigma_{\mu\nu}(\omega).
919: \end{equation}
920:
921:
922: Moreover, the CDMFT does not look for a strict solution of the Euler equation
923: (\ref{EulerVCA}), but tries instead to set each of the terms between brackets to zero separately. Since the Euler
924: equation (\ref{EulerVCA}) can be seen as a scalar product, CDMFT requires
925: that the modulus of one of the vectors vanish to make the scalar product
926: vanish. From a heuristic point of view, it is as if each component of the
927: Green function in the cluster were equal to the corresponding component
928: deduced from the lattice Green function. This clearly reduces to single site
929: DMFT when there is only one lattice site.
930:
931: When the bath is discretized, i.e., is made of a finite number of bath
932: \textquotedblleft orbitals\textquotedblright, the left-hand side of
933: Eq.~(\ref{EulerVCA}) cannot vanish separately for each frequency, since the
934: number of degrees of freedom in the bath is insufficient. Instead, one adopts
935: the following self-consistent scheme: (1) one starts with a guess value of the
936: bath parameters $(V_{\mu\alpha},\epsilon_{\alpha})$ and solves the cluster
937: Hamiltonian $H^{\prime}$ numerically. (2) One then calculates the combination
938: \begin{equation}
939: \hat{\mathcal{G}}_{0}^{-1}=\left[\sum_{\tilde{\mathbf{k}}}\frac
940: {1}{\hat{G}_{0\mathbf{t}}^{-1}(\tilde{\mathbf{k}})-\hat{\Sigma}^{\prime}(i\omega_{n}%
941: )}\right] ^{-1}+\hat\Sigma^{\prime}(i\omega_{n})
942: \end{equation}
943: and (3) minimizes the following canonically invariant distance function:
944: \begin{equation}
945: d=\sum_{n,\mu,\nu}\left\vert\left( i\omega_{n}+\mu-\hat{t}^\prime-\hat\Gamma
946: (i\omega_{n})-\hat{\mathcal{G}}_{0}^{-1}\right)_{\mu\nu}\right\vert ^{2}\label{dist_func}%
947: \end{equation}
948: over the set of bath parameters (changing the bath parameters at this step
949: does not require a new solution of the Hamiltonian $H^{\prime}$, but merely a
950: recalculation of the hybridization function $\hat{\Gamma}$). The bath
951: parameters obtained from this minimization are then put back into step (1) and
952: the procedure is iterated until convergence.
953:
954: In practice, the distance function (\ref{dist_func}) can take various forms,
955: for instance by adding a frequency-dependent weight in order to emphasize
956: low-frequency properties\cite{Kancharla:2005, Bolech:2003, Stanescu:2005} or
957: by using a sharp frequency cutoff.\cite{Kyung:2005} These weighting factors
958: can be considered as rough approximations for the missing factor $\delta
959: \Sigma_{\nu\mu}^{\prime}(i\omega_{n})/\delta\mathbf{t}^{\prime}$ in the Euler
960: equation (\ref{EulerVCA}). The frequencies are summed over on a discrete,
961: regular grid along the imaginary axis, defined by some fictitious inverse
962: temperature $\beta$, typically of the order of 20 or 40 (in units of $t^{-1}%
963: $). Even when the total number of cluster plus bath sites in CDMFT equals the
964: number of sites in a VCA calculation, CDMFT is much faster than
965: the VCA since the minimization of a grand potential functional requires many exact diagonalizations of the cluster Hamiltonian $H'$.
966:
967: The final lattice Green function from which one computes observable quantities
968: may be obtained by periodizing the self-energy, as in
969: Ref.~\onlinecite{Kotliar:2001} or in the CPT manner described above in
970: Eq.~(\ref{GCPT}). We prefer the last approach because it corresponds to the
971: Green function needed to obtain the density from $\partial\Omega/\partial
972: \mu=-\mathrm{Tr}(G)$ and also because periodization of the self-energy gives
973: additional unphysical states in the Mott gap\cite{Senechal:2003} (see also Ref.~\onlinecite{Stanescu:2004}).
974:
975: \subsubsection{The Dynamical cluster approximation}
976:
977: The DCA\cite{Hettler:1998} cannot be formulated within the self-energy
978: functional approach.\footnote{Th. Maier, M. Potthoff and D. S\'{e}n\'{e}chal,
979: unpublished.}
980: It is based on the idea of discretizing irreducible quantities,
981: such as the energy, in reciprocal space. It is believed to converge faster for
982: $\mathbf{q=0}$ quantities whereas CDMFT converges exponentially fast for local
983: quantities.\cite{Biroli:2002, Aryanpour:2005, Biroli:2005}
984:
985: \subsubsection{Benchmarks for quantum cluster approaches}
986:
987: Since DMFT becomes exact in infinite dimension, the most difficult challenge
988: for cluster extensions of this approach is in one dimension. In addition,
989: exact results to compare with exist only in one dimension so it is mostly in
990: $d=1$ that cluster methods have been checked. In $d=2$ there have also been a
991: few comparisons with QMC as we shall discuss.
992:
993: %-----------------------------------------------------------------------
994: %: FIG:fig_53a
995: \begin{figure}[t]
996: \centerline {\includegraphics[width=7cm]{f53a}}\caption{The spectral function
997: of the $U\to\infty$ limit of the one- dimensional Hubbard model, as calculated
998: from (a) an exact diagonalization of the Hubbard model with $U/t=100$ on a
999: periodic 12- site cluster; (b) the same, but with CPT, on a 12-site cluster
1000: with open boundary conditions; (c) the exact solution, taken from
1001: Ref.~\onlinecite{Favand:1997}; beware: the axes are oriented differently. In
1002: (a) and (b) a finite width $\eta$ has been given to peaks that would otherwise
1003: be Dirac $\delta$-functions.}%
1004: \label{fig_53a}%
1005: \end{figure}
1006: %-----------------------------------------------------------------------
1007:
1008: %-----------------------------------------------------------------------
1009: %: FIG:fig_53b
1010: \begin{figure}[ptb]
1011: \centerline{\includegraphics[width=6.5cm]{f53b}} \caption{Chemical potential
1012: as a function of density in the one-dimensional Hubbard model, as calculated
1013: by CPT (from Ref.~\onlinecite{Senechal:2002}). The exact, Bethe-Ansatz result
1014: is shown as a solid line.}%
1015: \label{fig_53b}%
1016: \end{figure}
1017: %-----------------------------------------------------------------------
1018:
1019:
1020: %-----------------------------------------------------------------------
1021: %: FIG:fig_54
1022: \begin{figure}[ptb]
1023: \centerline{\includegraphics[width=6.5cm]{f54}} \caption{Top: Comparison
1024: (expressed in relative difference) between the ground-state energy density of
1025: the half-filled, one-dimensional Hubbard model calculated from the exact,
1026: Bethe-Ansatz result. The results are displayed as a function of the hopping
1027: $t$, for $U=2t$ and various cluster sizes $L$ (connected symbols). For
1028: comparison, the exact diagonalization values of finite clusters with periodic
1029: boundary conditions are also shown (dashed lines) for $L=8$ and $L=12$.
1030: Bottom: Same for the double occupancy. An extrapolation of the results to
1031: infinite cluster size ($L\to\infty$) using a quadratic fit in terms of $1/L$
1032: is also shown, and is accurate to within 0.5\%. Taken from
1033: Ref.~\onlinecite{Senechal:2002}.}%
1034: \label{fig_54}%
1035: \end{figure}
1036: %-----------------------------------------------------------------------
1037:
1038:
1039: CPT has been checked\cite{Senechal:2003} for example by comparing
1040: with exact results\cite{Favand:1997}
1041: for the spectral function at $U\rightarrow\infty$ in $d=1$
1042: as shown in Fig.~\ref{fig_53a}. Fig.~\ref{fig_53b} shows the chemical potential
1043: as a function of density for various values of $U$. Fig.~\ref{fig_54} shows
1044: the convergence rates for the total energy and for the double occupancy in the
1045: $d=1$ half-filled model. Clearly, there is a dramatic improvement compared
1046: with exact diagonalizations.
1047:
1048: %-----------------------------------------------------------------------
1049: %: FIG:fig_55b
1050: \begin{figure}[ptb]
1051: \centerline{\includegraphics[width=5.5cm]{f55b}} \caption{Ground state energy
1052: of the half-filled, two-dimensional Hubbard model ($t=1$) as a function of
1053: $U$, as obtained from various methods: Exact diagonalization (ED), CPT and
1054: VCPT on a 10-site cluster, quantum Monte Carlo (QMC) and variational Monte
1055: Carlo (VMC). Taken from Ref.~\onlinecite{Dahnken:2004}.}%
1056: \label{fig_55b}%
1057: \end{figure}
1058: %-----------------------------------------------------------------------
1059:
1060:
1061: The main weakness of CPT is that it cannot take into account tendency towards
1062: long-range order. This is remedied by VCPT, as shown in Fig.~\ref{fig_55b}
1063: where CPT, VCPT are both compared with QMC as a benchmark. Despite this
1064: agreement, we should stress that long wave length fluctuations are clearly
1065: absent from cluster approaches. Hence, the antiferromagnetic order parameter
1066: at half-filling, for example, does not contain the effect of zero-point long
1067: wave length transverse spin fluctuations. This is discussed for example in the
1068: context of Fig.~9 of Ref.~\onlinecite{Dahnken:2004}.
1069:
1070: %-----------------------------------------------------------------------
1071: %: FIG:fig_56
1072: \begin{figure}[ptb]
1073: \centerline{\includegraphics[width=6.5cm]{f56}} \caption{CDMFT calculation on
1074: a $2 \times2$ cluster with $8$ bath sites of the density as a function of the
1075: chemical potential in the one-dimensional Hubbard model for $U=4t$, as
1076: compared with the exact solution, DMFT and other approximation schemes. Taken
1077: from Ref.~\onlinecite{Capone:2004}.}%
1078: \label{fig_56}%
1079: \end{figure}
1080: %-----------------------------------------------------------------------
1081:
1082:
1083: CDMFT corrects the difficulties of CPT near half-filling by reproducing the
1084: infinite compressibility predicted by the Bethe ansatz in one dimension as
1085: shown in Fig.~\ref{fig_56}.\cite{Capone:2004} Detailed comparisons between the
1086: local and near-neighbor Green functions\cite{Capone:2004, Bolech:2003} have
1087: been performed. One should note that these results, obtained from exact
1088: diagonalization, also need the definition of a distance function (See
1089: Eq.~(\ref{dist_func}) above) that helps find the best bath parametrization to
1090: satisfy the self-consistency condition. This measure forces one to define
1091: calculational parameters such as a frequency cutoff and an fictitious
1092: temperature defining the Matsubara frequencies to sum over. The final results
1093: are not completely insensitive to the choice of fictitious temperature or
1094: frequency-weighing scheme but are usually considered reliable and consistent
1095: with each other when $\beta$ lies between $20$ and $40$. The cutoff procedures
1096: have been discussed in Ref.~\onlinecite{Kyung:2005}.
1097:
1098: The relative merits of DCA and CDMFT have been discussed for example in Refs~
1099: \onlinecite{Biroli:2002, Aryanpour:2005, Biroli:2005, Maier:2002a, Pozgajcic:2004}.
1100: Briefly speaking, convergence seems faster in DCA for long wave length
1101: quantities but CDMFT is faster (exponentially) for local quantities.
1102:
1103: \section{Results and concordance between different methods\label{Results}}
1104:
1105: In this section, we outline the main results we obtained concerning the
1106: pseudogap and d-wave superconductivity in the two-dimensional Hubbard model.
1107: Quantum cluster approaches are better at strong coupling while TPSC is best at
1108: weak coupling. Nevertheless, all these methods are non-perturbative, the
1109: intermediate coupling regime presenting the physically most interesting case.
1110: But it is also the regime where we have the least control over the
1111: approximations. As we will show, it is quite satisfying that, at
1112: intermediate coupling, weak-coupling and strong-coupling approaches give
1113: results that are nearly in quantitative agreement with each other. This gives
1114: us great confidence into the validity of the results. As an example of
1115: concordance, consider the fact that to obtain spectral weight near $(\pi
1116: /2,\pi/2)$ at optimal doping in the electron-doped systems, $U$ has to be
1117: roughly $6t$. For larger $U,$ ($U=8t$ in CPT) that weight, present in the experiments,
1118: disappears. Smaller values of $U$ ($U=4t$ in CPT)
1119: do not lead to a pseudogap. Other examples
1120: of concordance include the value of the superconducting transition
1121: temperature $T_{c}$ obtained with DCA and with TPSC as well as the
1122: temperature dependence of souble occupancy obtained with the same
1123: two methods.
1124:
1125: \subsection{Weak and strong-coupling pseudogap\label{WeakAndStrongCoupling}}
1126:
1127: To understand the pseudogap it is most interesting to consider both hole and
1128: electron-doped cuprates at once. This means that we have to include
1129: particle-hole symmetry breaking hoppings, $t^{\prime}$ and $t^{\prime\prime}$.
1130: We will see in the present section that it is possible to obtain a pseudogap
1131: at strong coupling without a large correlation length in the particle-hole or
1132: in the particle-particle channels. By contrast, at weak coupling one does need
1133: a long-correlation length and low dimension. So there appears to be
1134: theoretically two different mechanisms for the pseudogap.
1135:
1136: %-----------------------------------------------------------------------
1137: %: FIG:fig_62
1138: \begin{figure}[ptb]
1139: \centerline{\includegraphics[width=6.5cm]{f62}} \caption{ Single particle
1140: spectral weight, as a function of energy $\omega$ in units of $t$, for wave
1141: vectors along the high-symmetry directions shown in the inset. (a) CPT
1142: calculations on a $3 \times4$ cluster with ten electrons (17\% hole doped).
1143: (b) the same as (a), with 14 electrons (17\% electron doped). In all cases
1144: $t^{\prime}=-0.3t$ and $t^{\prime\prime}=0.2t$. A Lorentzian broadening
1145: $\eta=0.12t$ is used to reveal the otherwise delta peaks. From
1146: Ref.~\onlinecite{Senechal:2004}.}%
1147: \label{fig_62}%
1148: \end{figure}
1149: %-----------------------------------------------------------------------
1150:
1151:
1152: %-----------------------------------------------------------------------
1153: %: FIG:fig_63
1154: \begin{figure}[ptb]
1155: \centerline{\includegraphics[width=6.5cm]{f63}} \caption{Onset of the
1156: pseudogap as a function of $U$ corresponding to Fig.~\ref{fig_62}, taken from
1157: Ref.~\onlinecite{Senechal:2004}. Hole-doped case on the left, electron-doped
1158: case on the right panel}%
1159: \label{fig_63}%
1160: \end{figure}
1161: %-----------------------------------------------------------------------
1162:
1163:
1164: \subsubsection{CPT}
1165:
1166: The top panel in Fig.~\ref{fig_62} presents the single-particle spectral
1167: weight, $A(\mathbf{k},\omega)$ or imaginary part of the single-particle Green
1168: function, for the model with $t^{\prime}=-0.3t,$ $t^{\prime\prime}=0.2t$ in
1169: the hole-doped case, for about $17\%$ doping.\cite{Senechal:2004} Each curve
1170: is for a different wave vector (on a trajectory shown in the inset) and is
1171: plotted as a function of frequency in units of $t$. This kind of plot is known
1172: as Energy Dispersion Curves (EDC). It is important to point out that the
1173: theoretical results are obtained by broadening a set of delta function, so
1174: that the energy resolution is $\eta=0.12t$ corresponding roughly to the
1175: experimental resolution we will compare with in the next section. At small
1176: $U=2t$ on the top panel of Fig.~\ref{fig_62}, one recovers a Fermi liquid. At
1177: large $U$, say $U=8t$, the Mott gap at positive energy is a prominent feature.
1178: The pseudogap is a different feature located around the Fermi energy. To see
1179: it better, we present on the left-hand panel of Fig.~\ref{fig_63} a blow-up in
1180: the vicinity of the Fermi surface crossing occurring near $(\pi,0)$.
1181: Clearly, there is a minimum in $A(\mathbf{k},\omega)$ at the
1182: Fermi-surface crossing when $U$ is large enough instead of a maximum like in
1183: Fermi liquid theory.
1184:
1185: %-----------------------------------------------------------------------
1186: %: FIG:fig_64
1187: \begin{figure}[ptb]
1188: \centerline{\includegraphics[width=6.5cm]{f64}} \caption{MDC from CPT in the
1189: $t$-$t^{\prime}=-0.3t$, $t^{\prime\prime}=0.2t$ Hubbard model, taken from
1190: Ref.~\onlinecite{Senechal:2004}. }%
1191: \label{fig_64}%
1192: \end{figure}
1193: %-----------------------------------------------------------------------
1194:
1195:
1196: It is also possible to plot $A(\mathbf{k},\omega)$ at fixed frequency for
1197: various momenta. They are so-called Momentum Dispersion Curves (MDC). In
1198: Fig.~\ref{fig_64} we take the Fermi energy $\omega=0,$ and we plot the
1199: magnitude of the single-particle spectral weight in the first quadrant of the
1200: Brillouin zone using red for high-intensity and blue for low intensity. The
1201: figure shows that, in the hole-doped case (top panel), weight near $(\pi
1202: /2,\pi/2)$ survives while it tends to disappear near $(\pi,0)$ and $(0,\pi)$.
1203: That pseudogap phenomenon is due not only to large $U$ but also to the fact
1204: that the line that can be drawn between the points $(\pi,0)$ and $(0,\pi)$
1205: crosses the Fermi surface. When there is no such crossing, one recovers a
1206: Fermi surface (not shown here). The $(\pi,0)$ to $(0,\pi)$ line has a double
1207: role. It is the antiferromagnetic zone boundary, as well as the line that
1208: indicates where umklapp processes become possible, i.e., the line where we can
1209: scatter a pair of particles on one side of the Fermi surface to the other side
1210: with loss or gain of a reciprocal lattice vector. Large scattering rates
1211: explain the disappearance of the Fermi surface.\cite{Senechal:2004} We also
1212: note that the size of the pseudogap in CPT, defined as the distance between
1213: the two peaks, does not scale like $J=4t^{2}/U$ at large coupling. It seems to
1214: be very weakly $U$ dependent, its size being related to $t$ instead. This
1215: result is corroborated by CDMFT.\cite{Kyung:2005}
1216:
1217: %-----------------------------------------------------------------------
1218: %: FIG:fig_71
1219: \begin{figure}[ptb]
1220: \centerline{\includegraphics[width=8.5cm]{f71}} \caption{Right: The corresponding EDC in the $t$-$t^{\prime}$-$t^{\prime\prime}$ Hubbard model, calculated on a 16-site
1221: cluster in CPT, at $n=9/8$. Inset: the pseudogap. Left: The corresponding Fermi energy
1222: momentum distribution curve.}%
1223: \label{fig_71}%
1224: \end{figure}
1225: %-----------------------------------------------------------------------
1226:
1227:
1228: The EDC for the electron-doped case is shown on the bottom panels of
1229: Fig.~\ref{fig_62} near optimal doping again. This time, the Mott gap appears
1230: below the Fermi surface so that the lower Hubbard band becomes accessible to
1231: experiment. The EDC in Fig.~\ref{fig_71} shows very well both the Mott gap and
1232: the pseudogap. Details of that pseudogap can be seen both in the inset of Fig.
1233: \ref{fig_71} or on the right-hand panel of Fig. \ref{fig_63}. While in the
1234: hole-doped case the MDC appeared to evolve continuously as we increase $U$
1235: (top panel of Fig.~\ref{fig_64}), in the electron-doped case (bottom panel)
1236: the weight initially present near $(\pi/2,\pi/2)$ at $U=4t$ disappears by the
1237: time we reach $U=8t$.
1238:
1239: %-----------------------------------------------------------------------
1240: %: FIG:fig_110
1241: \begin{figure}[ptb]
1242: \centerline{\includegraphics[width=6.5cm]{f110}} \caption{EDC in the
1243: $t$-$t^{\prime}$-$t^{\prime\prime}$ Hubbard model, with $t'=-0.3t$ and $t''=0.2t$, calculated on a
1244: 8-site cluster for $U=8t$ in VCPT. d-wave superconductivity is present in the
1245: holde-doped case (left) and both antiferromagnetism and d-wave
1246: superconductivity in the electron-doped case. The resolution is not large
1247: enough in the latter case to see the superconducting gap. The Lorentzian
1248: broadening is $0.2t$. From Ref.~\onlinecite{Senechal:2005}. }%
1249: \label{fig_110}%
1250: \end{figure}
1251: %-----------------------------------------------------------------------
1252:
1253:
1254: In Fig.~\ref{fig_110} we show, with the same resolution as CPT, the MDC for
1255: VCPT.\cite{Senechal:2005} In this case the effect of long-range order is
1256: included and visible but, at this resolution, the results are not
1257: too different from those
1258: obtained from CPT in Fig.~\ref{fig_64}.
1259:
1260: %-----------------------------------------------------------------------
1261: %: FIG:fig_75
1262: \begin{figure}[ptb]
1263: \centerline{\includegraphics[width=7.5cm]{f75}} \caption{MDC in the
1264: $t$-$t^{\prime}$, $U=8t$ Hubbard model, calculated on a 4-site cluster in
1265: CDMFT. Energy resolution, $\eta=0.1t$ (left and middle). Left: Hole-doped dSC
1266: ($t'=-0.3t, n=0.96$), Middle: Electron-doped dSC ($t'=0.3t, n=0.93$), Right:
1267: Same as middle with $\eta=0.02t$. Note the particle-hole transformation in the
1268: electron-doped case. From Ref.~\onlinecite{Kancharla:2005}. }%
1269: \label{fig_75}%
1270: \end{figure}
1271: %-----------------------------------------------------------------------
1272:
1273:
1274: \subsubsection{CDMFT and DCA}
1275:
1276: CDMFT\cite{Kancharla:2005} gives MDC that, at comparable resolution,
1277: $\eta=0.1t$, are again compatible with CPT and with VCPT. The middle panel in
1278: Fig.~\ref{fig_75} is for the electron-doped case but with a particle-hole
1279: transformation so that $t^{\prime}=+0.3t$ and $\mathbf{k\rightarrow k}%
1280: +(\pi,\pi)$. Since there is a non-zero d-wave order parameter in this
1281: calculation, improving the resolution to $\eta=0.02t$ reveals the d-wave gap,
1282: as seen in the right most figure.
1283:
1284: %-----------------------------------------------------------------------
1285: %: FIG:fig_81
1286: \begin{figure}[ptb]
1287: \centerline{\includegraphics[width=8.5cm]{f81}} \caption{EDC in the
1288: $t$-$t^{\prime}=0$, $U=8t$ Hubbard model, calculated on a 4-site cluster in
1289: CDMFT. Top: normal (paramagnetic) state for various densities. Bottom: same
1290: for the antiferromagnetic state. From Ref.~\onlinecite{Kyung:2005}. }%
1291: \label{fig_81}%
1292: \end{figure}
1293: %-----------------------------------------------------------------------
1294:
1295:
1296: %-----------------------------------------------------------------------
1297: %: FIG:fig_55a
1298: \begin{figure}[ptb]
1299: \centerline{\includegraphics[width=6.5cm]{f55a}} \caption{EDC in the Hubbard
1300: model, $U=8t$, $t^{\prime}=0$ calculated in CPT, VCPT and QMC. From
1301: Ref.~\onlinecite{Dahnken:2004}. }%
1302: \label{fig_55a}%
1303: \end{figure}
1304: %-----------------------------------------------------------------------
1305:
1306:
1307: It has been argued for a while in DCA that there is a mechanism whereby
1308: short-range correlations at strong coupling can be the source of the pseudogap
1309: phenomenon.\cite{Huscroft:2001} To illustrate this mechanism in CDMFT, we take
1310: the case $t^{\prime}=t^{\prime\prime}=0$ and compare in Fig.~\ref{fig_81} the
1311: EDC for $U=8t$ without long-range order (top panels) and with long-range
1312: antiferromagnetic order (bottom panels).\cite{Kyung:2005} The four bands
1313: appearing in Figs~\ref{fig_81}a and \ref{fig_81}d are in agreement with what
1314: has been shown\cite{Dahnken:2004, Moreo:1995, Grober:2000} with CPT, VCPT and
1315: QMC in Fig.~\ref{fig_55a}. Evidently there are additional symmetries in the
1316: antiferromagnetic case. The bands that are most affected by the long-range
1317: order are those that are closest to the Fermi energy, hence they reflect spin
1318: correlations, while the bands far from the Fermi energy seem less sensitive to
1319: the presence of long-range order. These far away bands are what is left from
1320: the atomic limit where we have two dispersionless bands. As we dope, the
1321: chemical potential moves into the lower band closest to the Fermi energy. When
1322: there is no long-range order (Figs~\ref{fig_81}b and \ref{fig_81}c) the
1323: lower band closest to the Fermi energy moves very close to it, at the same
1324: time as the upper band closest to the Fermi energy looses weight, part of it
1325: moving closer to the Fermi energy. These two bands leave a pseudogap at the
1326: Fermi energy\cite{Harris:1967, Meinders:1993},
1327: although we cannot exclude that increasing the resolution would
1328: reveal a Fermi liquid at a very small energy scale. In the case when there is
1329: long-range antiferromagnetic order, (Figs~\ref{fig_81}e and \ref{fig_81}f) the
1330: upper band is less affected while the chemical potential moves in the lower
1331: band closest to the Fermi energy but without creating a pseudogap, as if we
1332: were doping an itinerant antiferromagnet. It seems that forcing the spin
1333: correlations to remain short range leads to the pseudogap phenomenon in this
1334: case. When a pseudogap appears, it is created again by very large scattering
1335: rates.\cite{Kyung:2005}
1336:
1337: %-----------------------------------------------------------------------
1338: %: FIG:fig_83
1339: \begin{figure}[ptb]
1340: \centerline{\includegraphics[width=6.5cm]{f83}} \caption{ Double occupancy
1341: $\left\langle n_{\uparrow}n_{\downarrow}\right\rangle $, in the
1342: two-dimensional Hubbard model for $n=1$, as calculated from TPSC (lines with
1343: x) and from DCA (symbols) from Ref.~\onlinecite{Moukouri:2001}. Taken from
1344: Ref.~\onlinecite{Kyung:2003a}.}%
1345: \label{fig_83}%
1346: \end{figure}
1347: %-----------------------------------------------------------------------
1348:
1349:
1350: \subsubsection{TPSC (including analytical results)}
1351:
1352: In Hartree-Fock theory, double occupancy is given by $n^{2}/4$ and is
1353: independent of temperature. The correct result does depend on temperature. One
1354: can observe in Fig.~\ref{fig_83} the concordance between the results for the
1355: temperature-dependent double occupancy obtained with DCA and with
1356: TPSC\cite{Kyung:2003a} for the $t^{\prime}=t^{\prime\prime}=0$ model. We have
1357: also done extensive comparisons between straight QMC calculations and
1358: TPSC.\cite{Roy:unpub} The downturn at low temperature has been confirmed by
1359: the QMC calculations. It comes from the opening of the pseudogap due to
1360: antiferromagnetic fluctuations, as we will describe below. The concomitant
1361: increase in the local moment corresponds to the decrease in double-occupancy.
1362: There seems to be a disagreement at low temperature between TPSC and DCA at
1363: $U=2t.$ In fact TPSC is closer to the direct QMC calculation. Since we expect
1364: quantum cluster methods in general and DCA in particular to be less accurate
1365: at weak coupling, this is not too worrisome. At $U=4t$ the density of states
1366: obtained with TPSC and with DCA at various temperatures are very close to each
1367: other.\cite{Kyung:2003a} We stress that as we go to temperatures well below
1368: the pseudogap, TPSC becomes less and less accurate, generally overemphasizing
1369: the downfall in double occupancy.
1370:
1371: %-----------------------------------------------------------------------
1372: %: FIG:fig_84
1373: \begin{figure}[ptb]
1374: \centerline{\includegraphics[width=8.5cm]{f84}} \caption{MDC at the Fermi
1375: energy for the two-dimensional Hubbard model for $U=6.25$, $t^{\prime
1376: }=-0.175t, t^{\prime\prime}=0.05t$ at various hole dopings, as obtained from
1377: TPSC. The far left from Ref.~\onlinecite{Ronning:2003} is the Fermi surface
1378: plot for $10 \%$ hole-doped Ca$_{2-x}$Na$_x$CuO$_2$Cl$_2$.}%
1379: \label{fig_84}%
1380: \end{figure}
1381: %-----------------------------------------------------------------------
1382: %-----------------------------------------------------------------------
1383: %: FIG:fig_85
1384: \begin{figure}[ptb]
1385: \centerline{\includegraphics[width=8.5cm]{f85}} \caption{MDC at the Fermi
1386: energy in the electron-doped case with $t^{\prime}=-0.175t, t^{\prime\prime
1387: }=0.05t$ and two different $U$'s, $U=6.25t$ and $U=5.75t$ obtained from TPSC.
1388: The first column is the corresponding experimental plots at $10\%$ and $15\%$
1389: doping in Ref.~\onlinecite{Armitage:2002}. From
1390: Refs.~\onlinecite{Kyung:2003,Kyung:2004}.}%
1391: \label{fig_85}%
1392: \end{figure}
1393: %-----------------------------------------------------------------------
1394:
1395:
1396: We will come back to more details on the predictions of TPSC for the
1397: pseudogap, but to illustrate the concordance with quantum cluster results
1398: shown in the previous subsection, we show in Fig.~\ref{fig_84} MDC obtained at
1399: the Fermi energy in the hole doped case for
1400: $t^{\prime}=-0.175t$, $t^{\prime\prime}=0.05t$. Again there is quasi-particle weight
1401: near $(\pi/2,\pi/2)$ and a pseudogap near $(\pi,0)$ and $(0,\pi)$. However, as
1402: we will discuss below, the antiferromagnetic correlation length necessary to
1403: obtain that pseudogap is too large compared with experiment. The
1404: electron-doped case is shown in Fig.~\ref{fig_85} near optimal doping
1405: and for different values of
1406: $U.$ As $U$ increases, the weight near $(\pi/2,\pi/2)$ disappears. That is in
1407: concordance with the results of CPT shown in Fig.~\ref{fig_64} where the
1408: weight at that location exists only for small $U.$ That also agrees with
1409: slave-boson calculations\cite{Yuan:2005} that found such weight for $U=6t$ and
1410: it agrees also with one-loop calculations\cite{Kusunose:2003} starting from a Hartree-Fock
1411: antiferromagnetic state that did not find weight at that location for $U=8t$.
1412: The simplest Hartree-Fock approach\cite{Kusko:2002, Markiewicz:2004} yields
1413: weight near $(\pi/2,\pi/2)$ only for unreasonably small values of $U$.
1414:
1415: %-----------------------------------------------------------------------
1416: %: FIG:fig_88
1417: \begin{figure}[ptb]
1418: \centerline{\includegraphics[width=6.5cm]{f88}} \caption{Cartoon explanation
1419: of the pseudogap in the weak-coupling limit. Below the dashed crossover line
1420: to the renormalized classical regime, when the antiferromagnetic correlation
1421: length becomes larger than the thermal de Broglie wave length, there appears
1422: precursors of the zero-temperature Boboliubov quasiparticles for the
1423: long-range ordered antiferromagnet.}%
1424: \label{fig_88}%
1425: \end{figure}
1426: %-----------------------------------------------------------------------
1427: A cartoon explanation of the pseudogap is given in Fig.~\ref{fig_88}. At high
1428: temperature we have a Fermi liquid, as illustrated in panel I. Now, suppose we
1429: start from a ground state with long-range order as in panel III, in other
1430: words at a filling between half-filling and $n_{c}$. In the Hartree-Fock
1431: approximation we have a gap and the fermion creation-annihilation operators
1432: now project on Bogoliubov-Valentin quasiparticles that have weight at both
1433: positive and negative energies. In two dimensions, the Mermin-Wagner theorem
1434: means that as soon as we raise the temperature above zero, long-range order
1435: disappears, but the antiferromagnetic correlation length $\xi$ remains large
1436: so we obtain the situation illustrated in panel II, as long as $\xi$ is much
1437: larger than the thermal de Broglie wave length $\xi_{th}\equiv v_{F}/(\pi T)$
1438: in our usual units. At the crossover temperature $T_{X}$ then the relative
1439: size of $\xi$ and $\xi_{th}$ changes and we recover the Fermi liquid. We now
1440: proceed to sketch analytically where these results come from starting from
1441: finite temperature. Details and more complete formulae may be found in
1442: Refs.~\onlinecite{Vilk:1994,Vilk:1996,Vilk:1997,Vilk:1995}. Note also that a
1443: study starting from zero temperature has also been performed in Ref. \onlinecite{borejsza:2004}.
1444:
1445: First we show how TPSC recovers the Mermin-Wagner theorem. Consider the
1446: self-consistency conditions given by the local moment sum rule Eq.~(\ref{Suscep})
1447: together with the expression for the spin-susceptibility
1448: Eq.~(\ref{Chi_sp}) and $U_{sp}$ in Eq.~(\ref{ansatz}). First, it is clear that
1449: if the left-hand side of the local moment sum rule Eq.~(\ref{Suscep}) wants to
1450: increase because of proximity to a phase transition, the right-hand side can
1451: do so only by decreasing $\langle n_{\uparrow}n_{\downarrow}\rangle$ which in
1452: turns decreases $U_{sp}$ through Eq.~(\ref{ansatz}) and moves the system away
1453: from the phase transition. This argument needs to be made more precise to
1454: include the effect of dimension. First, using the spectral representation one
1455: can show that every term of $\chi_{sp}({\bf q}, iq_n)$ is positive. Near a phase transition,
1456: the zero Matsubara frequency component of the susceptibility begins to
1457: diverge. On can check from the real-time formalism that the zero-Matsubara
1458: frequency contribution dominates when the characteristic spin fluctuation
1459: frequency $\omega_{sp}\sim\xi^{-2}$ becomes less than temperature, the
1460: so-called renormalized-classical regime. We isolate this contribution on the
1461: left-hand side of the local moment sum rule and we move the contributions from
1462: the non-zero Matsubara frequencies, that are non-divergent, on the right-hand
1463: side. Then, converting the wave vector sum to an integral and expanding the
1464: denominator of the susceptibility around the wave vector where the instability
1465: would occurs to obtain an Ornstein-Zernicke form, the local moment sum rule
1466: Eq.~(\ref{Suscep}) can be written in the form%
1467: \begin{equation}
1468: T\int q^{d-1}dq\frac{1}{q^{2}+\xi^{-2}}=C(T).
1469: \end{equation}
1470: The constant on the right-hand side contains only non-singular contributions
1471: and $\xi^{-2}$ contains $U_{sp}$ that we want to find. From the above
1472: equation, one finds immediately that in $d=2,$ $\xi\approx\exp(C(T)/T)$ so
1473: that the correlation length diverges only at $T=0.$ In three dimensions,
1474: isotropic or not, exponents correspond to those of the $N=\infty$ universality
1475: class.\cite{Dare:1996}
1476:
1477: %-----------------------------------------------------------------------
1478: %: FIG:fig_91
1479: \begin{figure}[ptb]
1480: \centerline{\includegraphics[width=6.5cm]{f91}}\caption{MDC plots at the Fermi
1481: energy (upper) and corresponding scattering rates (lower) obtained from TPSC.
1482: The red lines on the upper panel indicate the region where the scattering rate
1483: in the corresponding lower panels is large.}%
1484: \label{fig_91}%
1485: \end{figure}
1486: %-----------------------------------------------------------------------
1487: To see how the pseudogap opens up in the single-particle spectral weight,
1488: consider the expression (\ref{Self-long}) for the self-energy. Normally one
1489: has to do the sum over bosonic Matsubara frequencies first, but the zero
1490: Matsubara frequency contribution has the correct asymptotic behavior in
1491: fermionic frequencies $ik_{n}$ so that one can once more isolate on the
1492: right-hand side the zero Matsubara frequency contribution. This is confirmed
1493: by the real-time formalism\cite{Vilk:1997} (See also Eq.~(\ref{Sigma_Reel})
1494: below). In the renormalized classical regime then, we have
1495: \begin{equation}
1496: \Sigma(\mathbf{k}_{F},ik_{n})\propto T\int q^{d-1}dq\frac{1}{q^{2}+\xi^{-2}%
1497: }\frac{1}{ik_{n}-\varepsilon_{\mathbf{k}_{F}+\mathbf{Q+q}}}
1498: \label{RC-contribution-sigma}%
1499: \end{equation}
1500: where $\mathbf{Q}$ is the wave vector of the instability. Hence, when
1501: $\varepsilon_{\mathbf{k}_{F}+\mathbf{Q}}=0$, in other words at hot spots, we
1502: find after analytical continuation and dimensional analysis that
1503: \begin{align}
1504: \operatorname{Im}\Sigma^{R}(\mathbf{k}_{F},0) & \propto-\pi T\int
1505: d^{d-1}q_{\perp}dq_{||}\frac{1}{q_{\perp}^{2}+q_{||}^{2}+\xi^{-2}}\delta
1506: (v_{F}^{\prime}q_{||})\\
1507: & \propto\frac{\pi T}{v_{F}^{\prime}}\xi^{3-d}.
1508: \label{ImSigma2}
1509: \end{align}
1510: Clearly, in $d=4$, $\operatorname{Im}\Sigma^{R}(\mathbf{k}_{F},0)$ vanishes as
1511: temperature decreases, $d=3$ is the marginal dimension and in $d=2$ we have
1512: that $\operatorname{Im}\Sigma^{R}(\mathbf{k}_{F},0)\propto\xi/\xi_{th}$ that
1513: diverges at zero temperature. In a Fermi liquid that quantity vanishes at zero
1514: temperature. A diverging $\operatorname{Im}\Sigma^{R}(\mathbf{k}_{F},0)$
1515: corresponds to a vanishingly small $A(\mathbf{k}_{F},\omega=0)$ as we can see
1516: from%
1517: \begin{equation}
1518: A(\mathbf{k},\omega)=\frac{-2\operatorname{Im}\Sigma^{R}(\mathbf{k}_{F}%
1519: ,\omega)}{(\omega-\varepsilon_{\mathbf{k}}-\operatorname{Re}\Sigma
1520: ^{R}(\mathbf{k}_{F},\omega))^{2}+\operatorname{Im}\Sigma^{R}(\mathbf{k}%
1521: _{F},\omega)^{2}}. \label{Spectral-weight_1}%
1522: \end{equation}
1523: To see graphically this relationship between the location of the pseudogap and
1524: large scattering rates at the Fermi surface, we draw in Fig.~\ref{fig_91} both
1525: the Fermi surface MDC and, in the lower panels, the corresponding plots for
1526: $\operatorname{Im}\Sigma^{R}(\mathbf{k},0)$. Note that at stronger $U$ the
1527: scattering rate is large over a broader region, leading to a depletion of
1528: $A(\mathbf{k,}\omega)$ over a broader range of $\mathbf{k}$ values.
1529:
1530: An argument for the splitting in two peaks seen in Figs.~\ref{fig_29} and
1531: \ref{fig_88} is as follows. Consider the singular renormalized contribution
1532: coming from the spin fluctuations in Eq.~(\ref{RC-contribution-sigma}) at
1533: frequencies $\omega\gg v_{F}\xi^{-1}.$ Taking into account that contributions
1534: to the integral come mostly from a region $q\leq\xi^{-1}$, this expression
1535: leads to%
1536: \begin{align}
1537: \operatorname{Re}\Sigma^{R}( \mathbf{k}_{F},\omega) & =\left( T\int
1538: q^{d-1}dq\frac{1}{q^{2}+\xi^{-2}}\right) \frac{1}{ik_{n}-\varepsilon
1539: _{\mathbf{k}_{F}+\mathbf{Q}}}\nonumber\\
1540: & \equiv\frac{\Delta^{2}}{\omega-\varepsilon_{\mathbf{k}_{F}+\mathbf{Q}}}%
1541: \end{align}
1542: which, when substituted in the expression for the spectral weight
1543: (\ref{Spectral-weight_1}) leads to large contributions when
1544: \begin{equation}
1545: \omega-\varepsilon_{\mathbf{k}}-\frac{\Delta^{2}}{\omega-\varepsilon
1546: _{\mathbf{k}_{F}+\mathbf{Q}}}=0
1547: \end{equation}
1548: or, equivalently,%
1549: \begin{equation}
1550: \omega=\frac{( \varepsilon_{\mathbf{k}}+\varepsilon_{\mathbf{k}_{F}%
1551: +\mathbf{Q}}) \pm\sqrt{( \varepsilon_{\mathbf{k}}-\varepsilon_{\mathbf{k}%
1552: _{F}+\mathbf{Q}}) ^{2}+4\Delta^{2}}}{2},
1553: \end{equation}
1554: which corresponds to the position of the hot spots in Fig.~\ref{fig_85} for example.
1555:
1556: Note that analogous arguments hold for any fluctuation that becomes
1557: soft,\cite{Vilk:1997} including superconducting ones.\cite{Allen:2001,
1558: Kyung:2001}
1559: The wave vector $\mathbf{Q}$ would be different in each case.
1560:
1561: \subsubsection{Weak- and strong-coupling pseudogaps}
1562:
1563: The CPT results of Figs.~\ref{fig_62} and \ref{fig_71} clearly show that the
1564: pseudogap is different from the Mott gap. At finite doping, the Mott gap
1565: remains a local phenomenon, in the sense that there is a region in frequency
1566: space that is not tied to $\omega=0$ where \textit{for all wave vectors }there
1567: are no states. The peudogap by contrast is tied to $\omega=0$ and occurs in
1568: regions nearly connected by $(\pi,\pi),$ whether we are talking about hole- or
1569: about electron-doped cuprates. That the phenomenon is caused by short-range
1570: correlations can be seen in CPT from the fact that the pseudogap is
1571: independent of cluster shape and size (most of the results were presented for
1572: $3\times4$ clusters and we did not go below size $2\times2$). The
1573: antiferromagnetic correlations and any other two-particle correlations do not
1574: extend beyond the size of the lattice in CPT. Hence, the pseudogap phenomenon
1575: cannot be caused by antiferromagnetic long-range order since no such order
1576: exists in CPT. This is also vividly illustrated by the CDMFT results in
1577: Fig.~\ref{fig_81} that contrast the case with and without antiferromagnetic
1578: long-range order. The CDMFT results also suggest that the pseudogap appears in
1579: the bands that are most affected by antiferromagnetic correlations hence it
1580: seems natural to associate it with short-range spin correlations. The value of
1581: $t^{\prime}$ has an effect, but it mostly through the fact that it has a
1582: strong influence on the relative location of the antiferromagnetic zone
1583: boundary and the Fermi surface, a crucial factor determining where the
1584: pseudogap is. All of this as well as many results obtained earlier by
1585: DCA\cite{Huscroft:2001} suggest that there is a strong coupling mechanism that
1586: leads to a pseudogap in the presence of only short-range two-body
1587: correlations. However, the range cannot be zero. Only the Mott gap appears at
1588: zero range, thus the pseudogap is absent in single-site DMFT.
1589:
1590: In the presence of a pseudogap at strong coupling ($U>8t$), wave vector is not, so to speak, such a bad quantum number in certain directions. In other words the wave description is better in those directions.
1591: In other directions that are ``pseudogapped'', it is as if the localized, or particle description was better.
1592: This competition between wave and particle behavior is inherent to the Hubbard model.
1593: At the Fermi surface in low dimension, it seems that this competition is resolved by dividing (it is a crossover, not a real division) the Fermi surface in different sections.
1594:
1595: There is also a weak-coupling mechanism for the pseudogap. This has been
1596: discussed at length just in the previous section on TPSC. Another way to
1597: rephrase the calculations of the previous section is in the real frequency
1598: formalism. There one finds\cite{Vilk:1997} that
1599: \begin{align}
1600: &\Sigma^{\prime\prime R}(\mathbf{k}_{F},\omega)\nonumber\\
1601: %
1602: &~~ \propto\int\frac{d^{d-1}%
1603: q_{\perp}}{(2\pi)^{d-1}}\int\frac{d\omega^{\prime}}{\pi}[n(\omega^{\prime
1604: })+f(\omega+\omega^{\prime})]\chi_{sp}^{\prime\prime}(\mathbf{q}%
1605: ;\omega^{\prime}) \label{Sigma_Reel}%
1606: \end{align}
1607: so that if the characteristic spin fluctuation frequency in $\chi_{sp}%
1608: ^{\prime\prime}(\mathbf{q};\omega^{\prime})$ is much larger than temperature,
1609: then $[n(\omega^{\prime})+f(\omega+\omega^{\prime})]$ can be considered to act
1610: like a window of size $\omega$ or $T$ and $\chi_{sp}^{\prime\prime}%
1611: (\mathbf{q};\omega^{\prime})$ can be replaced by a function of ${\bf q}$ times
1612: $\omega^{\prime}$ which immediately leads to the Fermi liquid result
1613: $[\omega^{2}+(\pi T)^{2}].$ In the opposite limit where the characteristic
1614: spin fluctuation frequency in $\chi_{sp}^{\prime\prime}(\mathbf{q}%
1615: ;\omega^{\prime})$ is much less than temperature, then it acts as a window
1616: narrower than temperature and $[n(\omega^{\prime})+f(\omega+\omega^{\prime})]$
1617: can be approximated by the low frequency limit of the Bose factor, namely
1618: $T/\omega^{\prime}.$ Using the thermodynamic sum rule, that immediately leads
1619: to the result discussed before in Eq.(\ref{ImSigma2}),
1620: $\operatorname{Im}\Sigma(\mathbf{k}%
1621: _{F},0)\propto(\pi T/v_{F}^{\prime})\xi^{3-d}.$ This mechanism for the
1622: pseudogap needs long correlation lengths. In CPT, this manifests itself by the
1623: fact that the apparent pseudogap in Fig.~\ref{fig_64} at $U=4t$ is in fact
1624: mostly a depression in spectral weight that depends on system size and shape.
1625: In addition, in contrast to the short-range strong-coupling mechanism, at weak
1626: coupling the pseudogap is more closely associated with the intersection of the
1627: antiferromagnetic zone boundary with the Fermi surface.
1628:
1629: Which mechanism is important for the cuprates will be discussed below in the
1630: section on comparisons with experiments.
1631:
1632: \subsection{d-wave superconductivity}
1633:
1634: The existence of d-wave superconductivity at weak coupling in the Hubbard
1635: model mediated by the exchange of antiferromagnetic fluctuations
1636: \cite{Scalapino:1995, Carbotte:1999} had been proposed even before the
1637: discovery of high-temperature superconductivity.\cite{Scalapino:1986,
1638: Miyake:1986, Beal-Monod:1986} At strong-coupling, early
1639: papers\cite{Kotliar:1988, Inui:1988} also proposed that the superconductivity
1640: would be d-wave. The issue became extremely controversial, and even recently
1641: papers have been published\cite{Zhang:1997} that suggest that there is no
1642: d-wave superconductivity in the Hubbard model. That problem could have been
1643: solved very long ago through QMC calculations if it had been possible to do
1644: them at low enough temperature. Unfortunately, the sign problem hindered these
1645: simulations, and the high temperature
1646: results\cite{Hirsch:1988,White:1989,Moreo:1991,Scalettar:1991} were not
1647: encouraging: the d-wave susceptibility was smaller than for the
1648: non-interacting case. Since that time, numerical results from variational
1649: QMC,\cite{Paramekanti:2001, Paramekanti:2004} exact
1650: diagonalization\cite{Poilblanc:2002} and other numerical
1651: approaches\cite{Sorella:2002} for example, suggest that there is indeed d-wave
1652: superconductivity in the Hubbard model.
1653:
1654: In the first subsection, we show that VCPT leads to a zero-temperature phase
1655: diagram for both hole and electron-doped systems that does show the basic
1656: features of the cuprate phase diagram, namely an antiferromagnetic phase and a
1657: d-wave superconducting phase in doping ranges that are quite close to
1658: experiment\cite{Senechal:2005} (The following section will treat in more
1659: detail comparisons with experiment). The results are consistent with
1660: CDMFT.\cite{Kancharla:2005} The fall in the d-wave superconducting order
1661: parameter near half-filling is associated with the Mott phenomenon. The next
1662: subsection stresses the instability towards d-wave superconductivity as seen
1663: from the normal state and mostly at weak coupling. We show that TPSC can
1664: reproduce available QMC results and that its extrapolation to lower
1665: temperature shows d-wave superconductivity in the Hubbard model. The
1666: transition temperature found at optimal doping\cite{Kyung:2003} for $U=4t$
1667: agrees with that found by DCA,\cite{Maier:2005a} a result that could be
1668: fortuitous. But again the concordance between theoretical results obtained at
1669: intermediate coupling with methods that are best at opposite ends of the range
1670: of coupling strengths is encouraging.
1671:
1672: %-----------------------------------------------------------------------
1673: %: FIG:fig_96
1674: \begin{figure}[ptb]
1675: \centerline{\includegraphics[width=8cm]{f96}} \caption{Antiferromagnetic
1676: (bottom) and $d$-wave (top) order parameters for $U=8t$, $t^{\prime}=-0.3t$
1677: $t^{\prime\prime}=0.2t$ as a function of the electron density ($n$) for
1678: $2\times3$, $2\times4$ and 10-site clusters, calculated in VCPT. Vertical
1679: lines indicate the first doping where only $d$-wave order is non-vanishing.
1680: From Ref.~\onlinecite{Senechal:2005}. }%
1681: \label{fig_96}%
1682: \end{figure}
1683: %-----------------------------------------------------------------------
1684:
1685:
1686: %-----------------------------------------------------------------------
1687: %: FIG:fig_138
1688: \begin{figure}[ptb]
1689: \centerline{\includegraphics[width=7cm]{f138}} \caption{d-wave order parameter
1690: as a function of $n$ for various values of $t^{\prime}$, calculated in CDMFT
1691: on a $2 \times2$ cluster for $U=8t$. The positive $t^{\prime}$ case
1692: corresponds to the electron-doped case when a particle-hole transformation is
1693: performed. From Ref.~\onlinecite{Kancharla:2005}. }%
1694: \label{fig_138}%
1695: \end{figure}
1696: %-----------------------------------------------------------------------
1697:
1698:
1699: \subsubsection{Zero-temperature phase diagram}
1700:
1701: In VCPT, one adds to the cluster Hamiltonian the terms\cite{Senechal:2005}
1702: \begin{align}
1703: H'_{M} & =M\sum_{\mu}e^{i\mathbf{Q}\cdot\mathbf{R}_{\mu}}(n_{\mu\uparrow
1704: }-n_{\mu\downarrow})\label{weiss_eq}\\
1705: H'_{D} & =D\sum_{\mu\nu}g_{\mu\nu}(c_{\mu\uparrow}c_{\nu\downarrow
1706: }+\mathrm{H.c.})
1707: \end{align}
1708: with $M$ and $D$ are respectively antiferromagnetic and d-wave superconducting
1709: Weiss fields that are determined self-consistently and
1710: $g_{\mu\nu}$ equal to $\pm1$ on near-neighbor sites following the d-wave pattern.
1711: We recall that the cluster Hamiltonian should be understood in a variational
1712: sense. Fig.~\ref{fig_96} summarizes, for various cluster sizes, the results
1713: for the d-wave order parameter $D_{0}$ and for the antiferromagnetic order
1714: parameter $M_{0}$ for a fixed value of $U=8t$ and the usual hopping parameters
1715: $t^{\prime}=-0.3t$ and $t^{\prime\prime}=0.2t$. The results for
1716: antiferromagnetism are quite robust and extend over ranges of dopings that
1717: correspond quite closely to those observed experimentally. Despite the fact
1718: that the results for d-wave superconductivity still show some size dependence,
1719: it is clear that superconductivity alone without coexistence extends over a
1720: much broader range of dopings on the hole-doped than on the electron-doped
1721: side as observed experimentally. VCPT calculations on smaller system
1722: sizes\cite{Hanke:2005} but that include, for thermodynamic consistency,
1723: the cluster chemical potential as a
1724: variational parameter show superconductivity that extends over a much broader
1725: range of dopings. Also, for small $2\times2$ clusters, VCPT has stronger order
1726: parameter on the electron than on the hole-doped side, contrary to the results
1727: for the largest system sizes in Fig.~\ref{fig_96}. This is also what is found
1728: in CDMFT as can be seen in Fig.~\ref{fig_138}. It is quite likely that the
1729: zero-temperature Cooper pair size is larger than two sites, so we consider the
1730: results for $2\times2$ systems only for their qualitative value.
1731:
1732: Concerning the question of coexistence with antiferromagnetism, one can see
1733: that it is quite robust on the electron-doped side whereas on the hole-doped
1734: side, it is size dependent. That suggests that one should also look at
1735: inhomogeneous solutions on the hole-doped side since stripes are generally
1736: found experimentally near the regions where antiferromagnetism and
1737: superconductivity meet.
1738:
1739: %-----------------------------------------------------------------------
1740: %: FIG:fig_97
1741: \begin{figure}[ptb]
1742: \centerline{\includegraphics[width=8cm]{f97}} \caption{Antiferromagnetic
1743: (bottom) and $d$-wave (top) order parameters as a function of the electron
1744: density ($n$) for $t^{\prime}=-0.3t$ $t^{\prime\prime}=0.2t$ and various
1745: values of $U$ on a 8-site cluster, calculated in VCPT. From
1746: Ref.~\onlinecite{Senechal:2005}. }%
1747: \label{fig_97}%
1748: \end{figure}
1749: %-----------------------------------------------------------------------
1750:
1751:
1752: %-----------------------------------------------------------------------
1753: %: FIG:fig_139
1754: \begin{figure}[ptb]
1755: \centerline{\includegraphics[width=7cm]{f139}} \caption{d-wave order parameter
1756: as a function of $n$ for various values of $U$, and $t^{\prime}=t^{\prime
1757: \prime}=0$ calculated in CDMFT on a 4-site cluster. From
1758: Ref.~\onlinecite{Kancharla:2005}. }%
1759: \label{fig_139}%
1760: \end{figure}
1761: %-----------------------------------------------------------------------
1762:
1763:
1764: Fig.~\ref{fig_97} shows clearly that at strong coupling the size of the order
1765: parameter seems to scale with $J$, in other words it decreases with increasing
1766: $U.$ This is also found in CDMFT,\cite{Kancharla:2005} as shown in Fig.
1767: \ref{fig_139} for $t^{\prime}=t^{\prime\prime}=0$.
1768:
1769: %-----------------------------------------------------------------------
1770: %: FIG:fig_1
1771: \begin{figure}[ptb]
1772: \centerline{\includegraphics[width=6.5cm]{f1}} \caption{VCPT calculations for
1773: $U=4t, t^{\prime}=t^{\prime\prime}=0$ near half-filling on $2 \times4$
1774: lattice. Contrary to the strong coupling case, the d-wave order parameter $D_0$
1775: survives all the way to half-filling at weak coupling, unless we also allow
1776: for antiferromagnetism.}%
1777: \label{fig_1}%
1778: \end{figure}
1779: %-----------------------------------------------------------------------
1780:
1781:
1782: If we keep the antiferromagnetic order parameter to zero, one can check with
1783: both VCPT and CDMFT (Fig.~\ref{fig_138}) that the d-wave superconducting order
1784: parameter goes to zero at half-filling. This is clearly due to Mott
1785: localization. Indeed, at smaller $U=4t$ for example, the order parameter does
1786: not vanish at half-filling if we do not allow for long-range antiferromagnetic
1787: order, as illustrated in Fig.~\ref{fig_139} for CDMFT.\cite{Kancharla:2005}
1788: The same result was found in VCPT, as shown in Fig.\ref{fig_1}.\cite{Senechal:2005}
1789: Restoring long-range
1790: antiferromagnetic order does however make the d-wave order parameter vanish at half-filling.
1791:
1792: There are thus two ways to make d-wave superconductivity disappear at
1793: half-filling, either through long antiferromagnetic correlation lengths\cite{Micnas:2005}
1794: or
1795: through the Mott phenomenon. In the real systems, that are Mott insulators and
1796: also antiferromagnets at half-filling, both effects can contribute.
1797:
1798: \subsubsection{Instability of the normal phase}
1799:
1800: In the introduction to this section, we alluded to QMC calculations for the
1801: d-wave susceptibility.\cite{Hirsch:1988,White:1989,Moreo:1991,Scalettar:1991}
1802: Recent results\cite{Kyung:2003} for that quantity as a function of doping for
1803: various temperatures and for $U=4t,$ $t^{\prime}=t^{\prime\prime}=0$ are shown
1804: by symbols in Fig.~\ref{fig_12}. For lower temperatures, the sign problem
1805: prevents the calculation near half-filling. Yet, the lowest temperature is low
1806: enough that a dome shape begins to appear. Nevertheless, comparison with the
1807: non-interacting case, shown by the top continuous line, leads one to believe
1808: that interactions only suppress d-wave superconductivity. We can easily
1809: understand why this is so.
1810: As we already know,
1811: the TPSC results obtained from Eq.~(\ref{Suscep_d}) are very close to the QMC
1812: calculations, as shown by the solid lines in Fig.~\ref{fig_12}.
1813: In the temperature range of interest, the main
1814: contribution comes from the first term in Eq.~(\ref{Suscep_d}). That term
1815: represents the contribution to the susceptibility that comes from dressed
1816: quasiparticles that do not interact with each other. Since dressed
1817: quasiparticles have a lifetime, a pair breaking effect, it is normal that this
1818: contribution to the interacting susceptibility leads to a smaller contribution
1819: than in the non-interacting case. At the lowest temperature, $\beta=4/t$, the
1820: vertex contribution represented by the second term in Eq.~(\ref{Suscep_d})
1821: accounts for about $20\%$
1822: of the total. It goes in the direction of increasing the susceptibility. As we
1823: decrease the temperature further in TPSC, that term becomes comparable with
1824: the first one. Since the vertex in Eq.~(\ref{Suscep_d}) accounts for the
1825: exchange of a single spin wave, equality with the first term signals the
1826: divergence of the series, as in $1/(1-x)\sim1+x$. The divergence of that
1827: series represents physically the instability of the normal phase to a d-wave
1828: superconducting phase. This is analogous to the Thouless criterion and hence
1829: it gives an upper bound to $T_{c}$. In other words,
1830: Berezinskii-Kosterlitz-Thouless physics is not included.
1831:
1832: %-----------------------------------------------------------------------
1833: %: FIG:fig_104
1834: \begin{figure}[ptb]
1835: \centerline{\includegraphics[width=6.5cm]{f104}} \caption{$T_{c}$ as a
1836: function of doping, $\delta=1-n$, for $t^{\prime}=t^{\prime\prime}=0$
1837: calculated in TPSC using the Thouless criterion. From
1838: Ref.~\onlinecite{Kyung:2003}.}%
1839: \label{fig_104}%
1840: \end{figure}
1841: %-----------------------------------------------------------------------
1842:
1843:
1844: Fig.~\ref{fig_104} shows the TPSC transition temperature for $U=4t$ and for $U=6t$.
1845: As we move towards half-filling, located to the left of the diagram, starting
1846: from large dopings, $T_{c}$ goes up because of the increase in
1847: antiferromagnetic fluctuations. Eventually, $T_{c}$ turns around and decreases
1848: because of the opening of the pseudogap. Physically, when the pseudogap opens
1849: up, weight is removed from the Fermi level, the density of states becomes very
1850: small, and pairing cannot occur any more. In the FLEX
1851: approximation\cite{Bickers:1989, Pao:1994} that does not exhibit a
1852: pseudogap,\cite{Deisz:1996} that downturn, observed already in QMC at high
1853: temperature, does not occur. We have observed that in cases where $t^{\prime
1854: }\neq0$ so that the pseudogap opens only in a limited region around hot spots,
1855: the downturn can become less pronounced.
1856:
1857: The case $n=0.9=1-\delta$ that corresponds to optimal doping for $U=4t$ in
1858: Fig.~\ref{fig_104} has been studied by DCA. In an extensive and systematic
1859: study of the size dependence, Maier \textit{et al. }\cite{Maier:2005a}
1860: established the existence of a d-wave instability at a temperature that
1861: coincides to within a few percent with the result in Fig.~\ref{fig_104}. Since
1862: very few vortices can fit within even the largest cluster sizes studied in
1863: Ref.~\onlinecite{Maier:2005a}, it is quite likely that the $T_{c}$ that they find
1864: does not include Berezinskii-Kosterlitz-Thouless effects, just like ours.
1865: Despite the fact that, again, the concordance between weak and strong coupling
1866: methods at intermediate coupling comforts us, the uncertainties in the results
1867: found with TPSC and DCA force us to also allow for a fortuitous coincidence.
1868:
1869: \section{Comparisons with experiment\label{Experiment}}
1870:
1871: The reduction of the real problem of high-temperature superconducting
1872: materials to a one-band Hubbard is a non-trivial one. It has been discussed
1873: already in the early days of high $T_{c}$ superconductivity. The notion of a
1874: Zhang-Rice singlet\cite{Zhang:1988a} emerged for hole doped systems. The
1875: mapping to a one-band model has been discussed in many
1876: references,\cite{Andersen:1995, Macridin:2005} and we do not wish to discuss
1877: this point further here. In fact it is far from obvious that this mapping is
1878: possible. It is known that about $0.5$ eV below the Fermi surface, that
1879: mapping fails in hole-doped systems.\cite{Macridin:2005} Nevertheless, the
1880: one-band Hubbard model is in itself a hard enough problem for us. So it is
1881: satisfying to see that, in the end, it gives a picture that agrees with
1882: experiment in a quite detailed manner for the ARPES spectrum near the Fermi
1883: surface, for the phase diagram as well as for neutron scattering in cases
1884: where it can be calculated.
1885:
1886: Although we will not come back on this point at all, we briefly mention that
1887: fitting the spin wave spectrum\cite{Coldea:2001} for all energies and wave vectors at half
1888: filling in La$_{2}$CuO$_{4}$ gives values of $U,t,t^{\prime},t^{\prime\prime}$
1889: that are close to those used in the rest of this paper.\cite{Delannoy:2005,
1890: Delannoy:2005a, Toader:2005, Raymond:2005}
1891: It is in this context that ring exchange terms are usually discussed.
1892:
1893: \subsection{ARPES spectrum, an overview\label{ARPES_overview}}
1894:
1895: ARPES experiments have played a central role in the field of high-temperature
1896: superconductivity. We cannot expect to be able to present the vast
1897: experimental literature on the subject. We refer the reader to a very
1898: exhaustive review\cite{Damascelli:2003} and to some less complete but recent
1899: ones.\cite{Damascelli:2003a, Kordyuk:2005} The main facts about ARPES have
1900: been summarized in Ref.~\onlinecite{Damascelli:2003}. We comment on their main
1901: points one by one, using italics for our paraphrase of the reported
1902: experimental observations.
1903:
1904: \textit{(i) The importance of Mott Physics and the renormalization of the
1905: bandwidth from $t$ to $J$ for the undoped parent compounds}. This
1906: renormalization was clear already in early QMC
1907: calculations.\cite{Preuss:1995,Moreo:1995} We already discussed the presence
1908: of four peaks. The one nearest to the Fermi surface at negative energies is
1909: the one referred to by experimentalists when they refer to this
1910: renormalization. This band has a dispersion of order $J$ (not shown on
1911: Fig.~\ref{fig_55a},
1912: but see Ref. \onlinecite{Preuss:1995}). This result also agrees with
1913: VCPT as shown in Fig.~\ref{fig_55a} and CDMFT (Fig.~\ref{fig_81}a). As shown
1914: in Figs.~\ref{fig_81}a and \ref{fig_81}d, whether the state is ordered or not
1915: the band width is similar. Analytical strong-coupling
1916: expansions\cite{Pairault:1998, Pairault:2000} and exact diagonalizations also
1917: find the same result. To find detailed agreement with experiment, one needs to
1918: include $t^{\prime}$ and $t^{\prime\prime}$.\cite{Gooding:1994} The evolution
1919: of the position of chemical potential for extremely small dopings as discussed
1920: in Ref.~\onlinecite{Shen:2005} is not reproduced by the strong-coupling calculations,
1921: although the result on chemical potential is somewhat material
1922: dependent.\cite{Yoshida:2005}
1923:
1924: \textit{(ii) In the overdoped case, the Fermi surface is well defined.}
1925: Although we have not shown any figures concerning this point, VCPT and CDMFT
1926: show the same result.
1927:
1928: %-----------------------------------------------------------------------
1929: %: FIG:fig_107
1930: \begin{figure}[ptb]
1931: \centerline{\includegraphics[width=5cm]{f107}}\caption{MDC at the Fermi energy
1932: for $10 \%$ hole-doped Ca$_{2-x}$Na$_x$CuO$_2$Cl$_2$ from
1933: Ref.~\onlinecite{Ronning:2003}.}%
1934: \label{fig_107}%
1935: \end{figure}
1936: %-----------------------------------------------------------------------
1937:
1938:
1939: \textit{(iii) The evolution with doping of the electronic structure has been
1940: mapped. It has shown the importance of antiferromagnetic correlations in the
1941: p-type underdoped cuprates and especially in the n-type ones in which the
1942: hot-spot physics is still observed at optimal doping. } We will come back on
1943: the latter point for electron-doped cuprates in the following subsection. The
1944: strong-coupling results obtained with VCPT and CDMFT have a resolution of
1945: order $0.1t$, which translates into about $30$~meV. This is not enough to
1946: accurately measure the Fermi velocity, which was found to be doping
1947: independent in LSCO.\cite{Zhou:2003} However, this suffices to compare with
1948: MDC curves obtained experimentally by integrating over an energy range of
1949: about $60$~meV, as shown in Fig.~\ref{fig_107} obtained in
1950: Ref.~\onlinecite{Ronning:2003} on Calcium oxyclorate Ca$_{2-x}$Na$_{x}%
1951: $CuO$_{2}$Cl$_{2}$, a $10\%$ hole-doped high temperature superconductor. The
1952: similarities between that figure and the CPT (Fig.~\ref{fig_64}), VCPT
1953: (Fig.~\ref{fig_110}) and CDMFT (Fig.~\ref{fig_75}) results is striking. The
1954: agreement is better when no antiferromagnetic long-range order is assumed, as
1955: in the CPT case. The flattening of the band structure near $(\pi,0)$ observed
1956: experimentally, can also be seen in CPT by comparing the top and middle EDC's
1957: taken at small and large $U$ respectively on the left panel of
1958: Fig.~\ref{fig_63}. This flattening is associated with the pseudogap
1959: phenomenon. Recall that the theoretical results were obtained with $t^{\prime
1960: }=-0.3t$ and $t^{\prime\prime}=0.2t$. This in turn implies an electron-hole
1961: asymmetry that is observed experimentally. We come back to this in the
1962: following subsection.
1963:
1964: \textit{(iv) The overall d-wave symmetry of the superconducting gap has been
1965: observed for both hole and electron doping, supporting the universality of the
1966: pairing nature in the cuprates. } In the next to next subsection, we discuss
1967: the phase diagram for competing antiferromagnetism and d-wave
1968: superconductivity and show striking similarities with the observations.
1969:
1970: \textit{(v) A normal-state pseudogap has been observed to open up at a
1971: temperature }$T^{\ast}>T_{c}$ \textit{in the underdoped regime with a d-wave
1972: form similar to the one of the superconducting gap. }That statement is correct
1973: only in the hole-doped compounds. In electron-doped systems the pseudogap has
1974: a form that is not of d-wave shape. If $T_{c}$ comes from a universal pairing
1975: mechanism, a universal mechanism may also be behind the pseudogap. As we have
1976: already discussed however, there are quantitative differences between strong
1977: and weak coupling mechanisms for both $T_{c}$ and the pseudogap. For
1978: electron-doped systems, we made quantitative predictions for the value of
1979: $T^{\ast}$ that have later been confirmed experimentally. All this is
1980: discussed further below. To date, in cluster methods the pseudogap temperature
1981: has been studied only with DCA.\cite{Huscroft:2001}
1982:
1983: \textit{(vi) A coherent quasiparticle peak below }$T_{c}$\textit{ has been
1984: observed near }$(\pi,0)$\textit{ whose spectral weight scales with the doping
1985: level }$x$\textit{ in the underdoped regime. } We expect that it is a general
1986: result that long-range order will restore quasiparticle like excitations in
1987: strongly correlated systems because gaps remove scattering channels near the
1988: Fermi level. This is clearly illustrated by comparing the upper and lower
1989: panels in Fig.~\ref{fig_81} that contrast the same spectra with and without
1990: antiferromagnetic long-range order. We have not performed the analysis of our
1991: results yet that could tell us whether the spectral weight of the
1992: quasiparticle scales with $x$ in the hole-underdoped regime. Our resolution
1993: may not be good enough to see the quasiparticle peak. Sharpening of the
1994: quasiparticle excitations in the superconducting state has however been
1995: observed in DCA.\cite{Maier:2004}
1996:
1997: \textit{(vii) The presence of an energy scale of about }$40-80$~meV \textit{in
1998: the quasiparticle dynamics manifests itself through a sharp dispersion
1999: renormalization and drop in the scattering rate observed at those energies at
2000: different momenta. } In hole-doped systems there is a kink in the nodal
2001: direction that is already seen above $T_{c}$ while in the antinodal direction
2002: it appears only below $T_{c}$. The energy scales and doping dependences of
2003: these two kinks are also different.\cite{Gromko:2003} The energy resolution in
2004: VCPT and CDMFT is not sufficient to distinguish these subtleties. In
2005: electron-doped cuprates experiments\cite{Armitage:2003} suggest that there is
2006: no observable kink feature, in agreement with the results presented in the
2007: following subsection.
2008:
2009: \subsection{The pseudogap in electron-doped cuprates}
2010:
2011: %-----------------------------------------------------------------------
2012: %: FIG:fig_121a
2013: \begin{figure}[ptb]
2014: \centerline{\includegraphics[width=6.5cm]{f121a}} \caption{Doping dependence
2015: of the MDC from experiments on NCCO with the corresponding EDC. From
2016: Ref.~\onlinecite{Armitage:2002} }%
2017: \label{fig_121a}%
2018: \end{figure}
2019: %-----------------------------------------------------------------------
2020:
2021:
2022: The ARPES spectrum of electron-doped cuprates is strikingly different from
2023: that of their hole-doped counterpart. The Fermi energy MDC's for the first
2024: quadrant of the Brillouin zone\cite{Armitage:2002} are shown at the top of
2025: Fig.~\ref{fig_121a} for three different dopings. There is a very clear
2026: evolution with doping. At the lowest dopings, there is no weight near
2027: $(\pi/2,\pi/2),$ contrary to the hole-doped case shown in Fig.~\ref{fig_107}.
2028: For all dopings there is weight near $(\pi,0)$ instead of the pseudogap that
2029: appeared there in the hole-doped case. The EDC's, also shown on the bottom of
2030: Fig.~\ref{fig_121a}, are drawn for a trajectory in the Brillouin zone that
2031: follows what would be the Fermi surface in the non-interacting case. Regions
2032: that are more green than red on the corresponding MDC's along that trajectory
2033: are referred to as hot spots. On the EDC's we clearly see that hot spots do
2034: not correspond to simply a decrease in the quasiparticle weight $Z$. They
2035: truly originate from a pseudogap, in other words from the fact that the
2036: maximum is pushed away from zero energy. Even though the measurements are done
2037: at low temperature $(T=10-20K)$ the energy resolution of about $60$~meV makes
2038: the superconducting gap invisible. What is observed at this resolution is the pseudogap.
2039:
2040: %-----------------------------------------------------------------------
2041: %: FIG:fig_117
2042: \begin{figure}[ptb]
2043: \centerline{\includegraphics[width=8.5cm]{f117}} \caption{Intensity plot of
2044: the spectral function as a function of $\omega$ in units of $t$ and wave
2045: vector from VCPT for $U=8t$ $t^{\prime}=-0.3t, t^{\prime\prime}=0.2t$ and
2046: $n=0.93$ at the bottom and $n=1.10$ (electron-doped) at the top. The
2047: Lorentzian broadening is $0.12t$ in the main figure and $0.04t$ in the inset
2048: that displays the d-wave gap. Top panel is for the electron-doped case in the
2049: right-hand panel of Fig.~\ref{fig_110}, while bottom panel is for the
2050: hole-doped case on the left of Fig.~\ref{fig_110}. From
2051: Ref.~\onlinecite{Senechal:2005}.}%
2052: \label{fig_117}%
2053: \end{figure}
2054: %-----------------------------------------------------------------------
2055:
2056:
2057: %-----------------------------------------------------------------------
2058: %: FIG:fig_121
2059: \begin{figure}[ptb]
2060: \centerline{\includegraphics[width=6.5cm]{f121}} \caption{Experimental Fermi
2061: surface plot (MDC at the Fermi energy) for NCCO (left) and corresponding
2062: energy distribution curves (right) for $15 \%$ electron-doping. From
2063: Ref.~\onlinecite{Armitage:2002} }%
2064: \label{fig_121}%
2065: \end{figure}
2066: %-----------------------------------------------------------------------
2067:
2068:
2069: The contrast between the location of the pseudogap in the hole and
2070: electron-doped compounds is clearly seen in Fig.~\ref{fig_117} obtained from
2071: VCPT.\cite{Senechal:2005} In that figure, the magnitude of the spectral weight
2072: is represented by the different colors as a function of frequency (in units of
2073: $t$) along different cuts of the Brillouin zone. In the bottom panel, for the
2074: hole-doped case, one observes the pseudogap near $(\pi,0)$. In the top panel,
2075: for the electron-doped case, it is only by zooming (inset) on the region for the Fermi
2076: energy crossing near $(\pi,0)$ that one sees the d-wave superconducting gap.
2077: At $10\%$ electron-doping, the pseudogap near $(\pi/2,\pi/2)$ is apparent. In
2078: this case there is antiferromagnetic long-range order, but even if we use CPT
2079: that does not exhibit long-range order, there appears a pseudogap in that
2080: region.\footnote{The latter result is disputed by a recent DCA calculation by
2081: Alexandru Macridin, Mark Jarrell, Thomas Maier, P. R. C. Kent,
2082: cond-mat/0509166.} The main difference between CPT and VCPT results is the
2083: bending back of the bands (for example around the symmetry axis at $(\pi
2084: /2,\pi/2)$) caused by halving of the size of the Brillouin zone in the
2085: antiferromagnetic case. Form factors\cite{Kusko:2002} are such that the
2086: intensity is not symmetric even if the dispersion is. The faint band located
2087: at an energy about $t$ below the Fermi energy near $(\pi/2,\pi/2)$ was also
2088: found in Ref.~\onlinecite{Kusunose:2003} by a one-loop spin-wave calculation
2089: around the Hartree-Fock antiferromagnetic ground state at $U=8t$.
2090: Experimentalists\cite{Armitage:2002} have suggested the existence of these
2091: states. The VCPT results go well beyond the spin-wave calculation (dashed
2092: lines in Fig.~\ref{fig_117}) since one can also see numerous features in
2093: addition to remnants of the localized atomic levels around $+5t$ and $-10t$.
2094:
2095: The optimally doped case is the real challenge for strong-coupling
2096: calculations. The spin-wave approach in Ref.~\onlinecite{Kusunose:2003} never
2097: shows the weight near $(\pi/2,\pi/2)$ that is seen in experiment
2098: (Fig.~\ref{fig_121}). Early mean-field calculations by Kusko et
2099: al.\cite{Kusko:2002} suggest that this $(\pi/2,\pi/2)$ feature appears for
2100: $U=3t$. This is very small compared with $U$ of the order of the bandwidth
2101: $8t$ necessary to have a Mott insulator at half-filling. We already discussed
2102: in Sec. \ref{WeakAndStrongCoupling} that both CPT and TPSC show that weight
2103: near $(\pi/2,\pi/2)$ appears for $U$ not too large, say of order $6t$. This
2104: same result is also obtained in the Kotliar-Ruckenstein slave boson
2105: approach.\cite{Yuan:2005}
2106:
2107: %-----------------------------------------------------------------------
2108: %: FIG:fig_122
2109: \begin{figure}[ptb]
2110: \centerline{\includegraphics[width=6.5cm]{f122}}\caption{EDC $A_{<}(\mathbf{k},\omega)\equiv A(\mathbf{k},\omega)f(\omega)$ along the Fermi surface
2111: calculated in TPSC (left) at optimal doping for $t^{\prime}=-0.175t$,
2112: $t^{\prime\prime}=0.05t$, $t=350$~meV and corresponding ARPES data on NCCO
2113: (right). From Ref.~\onlinecite{Kyung:2004}.}%
2114: \label{fig_122}%
2115: \end{figure}
2116: %-----------------------------------------------------------------------
2117:
2118:
2119: %-----------------------------------------------------------------------
2120: %: FIG:fig_123
2121: \begin{figure}[ptb]
2122: \centerline{\includegraphics[width=6.5cm]{f123}}\caption{EDC $A_{<}(\mathbf{k},\omega)\equiv A(\mathbf{k},\omega)f(\omega)$ along two other directions
2123: calculated for $t^{\prime}=-0.175t$, $t^{\prime\prime}=0.05t$, $t=350$~meV in
2124: TPSC (left column) and corresponding ARPES data on NCCO (right column). From
2125: Ref.~\onlinecite{Kyung:2004}.}%
2126: \label{fig_123}%
2127: \end{figure}
2128: %-----------------------------------------------------------------------
2129:
2130:
2131: Since TPSC is valid for a system of infinite size, we present detailed
2132: comparisons\cite{Kyung:2004} with experiment\cite{Armitage:2002} on
2133: Nd$_{1.85}$Ce$_{0.15}$CuO$_{4}$, an electron-doped cuprate. We take
2134: $t^{\prime}=-0.175t$, $t^{\prime\prime}=0.05t.$ Results obtained with
2135: $t^{\prime}=-0.275$ are very close to those we present. With the values used
2136: in CPT, $t^{\prime}=-0.3t,$ $t^{\prime}=0.2t,$ $U=6t,$ TPSC does not lead to
2137: strong enough antiferromagnetic fluctuations to obtain non-trivial effects in
2138: the temperature range studied, $\beta=20t$. We take $t=350$~meV.
2139: Fig.~\ref{fig_121a} shows the correspondence between EDC and MDC. Comparisons
2140: with experimental EDC at wave vectors along the non-interacting Fermi surface
2141: appear in Fig.~\ref{fig_122} for $U=5.75t$ and $15\%$ doping $(n=1.15)$. The
2142: dashed lines indicate the quite detailed agreement between theory and
2143: experiment. At the hot spot, (middle dashed line), the weight is pushed back
2144: about $0.2$~eV and there is a very small peak left at the Fermi surface, as in
2145: the experiment. If $U$ is not large enough the antiferromagnetic fluctuations
2146: are not strong enough to lead to a pseudogap. As in CPT (Fig.~\ref{fig_64}), if
2147: $U$ is too large the $(\pi/2,\pi/2)$ weight disappears, as illustrated earlier
2148: in Fig.~\ref{fig_85}. In Fig.~\ref{fig_123}, cuts along the $(0,0)$ to
2149: $(\pi,\pi)$ and $(0.65\pi,0)$ to $(0.65\pi,\pi)$ directions are compared with
2150: experiment. Again the peak positions and widths are very close, except for
2151: some experimental tails extending in the large binding energy direction. The
2152: theoretical results have similar asymmetry, but not as pronounced.
2153: Experimentally, the large binding energy tails (\textquotedblleft the
2154: background\textquotedblright) are the least reproducible features from sample to sample, especially for wave vectors near the Fermi
2155: surface".
2156: \footnote{Kyle Shen, Private communication.} The experimental
2157: renormalized Fermi velocities are $3.31\times10^{5}$~m/s and $3.09\times
2158: 10^{5}$~m/s along the zone diagonal and along the $(\pi,0)$-$(\pi,\pi)$
2159: direction, respectively. The corresponding renormalized Fermi velocities
2160: obtained by TPSC are $3.27\times10^{5}$~m/s and $2.49\times10^{5}$~m/s,
2161: respectively. The agreement is very good, particularly along the diagonal
2162: direction. The bare Fermi velocities are renormalized in TPSC by roughly a
2163: factor of two.\cite{Hankevych:2005}
2164:
2165: %-----------------------------------------------------------------------
2166: %: FIG:fig_124
2167: \begin{figure}[ptb]
2168: \centerline{\includegraphics[width=6.5cm]{f124}} \caption{EDC $A_{<}(\mathbf{k},\omega)\equiv A(\mathbf{k},\omega)f(\omega)$ along the Fermi surface shown in
2169: the insets for (b) $n=1.10$, $U=6.25t$. Lines are shifted by a constant for
2170: clarity. From Ref.~\onlinecite{Kyung:2004}. }%
2171: \label{fig_124}%
2172: \end{figure}
2173: %-----------------------------------------------------------------------
2174:
2175:
2176: As we move towards half-filling, we have to increase $U$ slightly to find
2177: agreement with experiment, as discussed earlier in Fig.~\ref{fig_85}.
2178: Fig.~\ref{fig_124} shows how well the EDC's agree for a Fermi surface cut at
2179: $10\%$ doping $(n=1.10)$. The increase is expected physically from the fact
2180: that with fewer electrons the contribution to screening that comes from Thomas
2181: Fermi physics should not be as good. This is also consistent with the fact
2182: that a larger value of $U$ is necessary to explain the Mott insulator at
2183: half-filling. It would also be possible to mimic the ARPES spectrum by keeping
2184: $U$ fixed and changing the hopping parameters, but the changes would be of
2185: order $20\%,$ which does not appear realistic.\cite{Kyung:2004}
2186:
2187: %-----------------------------------------------------------------------
2188: %: FIG:fig_125
2189: \begin{figure}[ptb]
2190: \centerline{\includegraphics[width=5.5cm]{f125}}\caption{Hot spots from
2191: quasi-static scatterings off antiferromagnetic fluctuations (renormalized
2192: classical regime).}%
2193: \label{fig_125}%
2194: \end{figure}
2195: %-----------------------------------------------------------------------
2196: %-----------------------------------------------------------------------
2197: %: FIG:fig_126
2198: \begin{figure}[ptb]
2199: \centerline{\includegraphics[width=6.5cm]{f126}} \caption{Semi-log plot of the
2200: AFM correlation length (in units of the lattice constant) against inverse
2201: temperature (in units of $J=125$ meV). Filled symbols denote calculated
2202: results and empty ones experimental data of Ref.~\onlinecite{Mang:2004} and
2203: Ref.~\onlinecite{Matsuda:1992} ($x=0.15$). From Ref.~\onlinecite{Kyung:2004}.
2204: }%
2205: \label{fig_126}%
2206: \end{figure}
2207: %-----------------------------------------------------------------------
2208:
2209:
2210: We have already explained that the physics behind the pseudogap in TPSC is
2211: scattering by nearly critical antiferromagnetic fluctuations. This is
2212: illustrated in Fig.~\ref{fig_125}. If this explanation is correct, the
2213: antiferromagnetic correlation length measured by neutron scattering should be
2214: quite large. The results of the measurement\cite{Mang:2004, Matsuda:1992} and
2215: of the TPSC calculations are shown in Fig.~\ref{fig_126}. The agreement is
2216: again surprisingly good. As we move to smaller dopings $n=1.1$ (not shown) the
2217: agreement becomes less good, but we do expect TPSC to deteriorate as $U$
2218: increases with underdoping. The arrow points to the temperature where EDC's
2219: shown earlier were calculated. Note however that the neutron measurements were
2220: done on samples that were not reduced, by contrast with the ARPES measurements
2221: mentioned earlier. We are expecting experiments on this subject.\footnote{M.
2222: Greven, private communication.}. We should point out that the EDC's depend
2223: strongly on temperature and on the actual value of the antiferromagnetic
2224: correlation length only in the vicinity of the temperature where there is a
2225: crossover to the pseudogap regime. Decreasing the temperature makes the
2226: $\omega=0$ peaks near $(\pi,0)$ sharper\cite{Kyung:2004}
2227: as observed experimentally.\cite{Onose:2004}
2228:
2229: %-----------------------------------------------------------------------
2230: %: FIG:fig_127
2231: \begin{figure}[ptb]
2232: \centerline{\includegraphics[width=6.5cm]{f127}} \caption{Pseudogap
2233: temperature $T^{\ast}$ (filled circles denote $T^{\ast}$ calculated from TPSC,
2234: empty ones experimental data extracted from optical
2235: conductivity.\cite{Onose:2001}) Empty triangles are experimental N\'{e}el
2236: temperatures $T_{N}$. The samples are reduced.\cite{Mang:2004} From
2237: Ref.~\onlinecite{Kyung:2004}. }%
2238: \label{fig_127}%
2239: \end{figure}
2240: %-----------------------------------------------------------------------
2241:
2242:
2243: The ARPES pseudogap temperature $T^{\ast}$ has been predicted with
2244: TPSC.\cite{Kyung:2004} The predictions are shown by the solid line in
2245: Fig.~\ref{fig_127}. The pseudogap temperature observed in optical
2246: experiments\cite{Onose:2001} is shown by the open circles. It differs from the
2247: ARPES result, especially as we move towards optimal doping. The size of the
2248: pseudogap observed in the optical experiments\cite{Onose:2001} $(10T^{\ast})$
2249: is comparable to the ARPES pseudogap. The solid line in Fig.~\ref{fig_127}
2250: contains several predictions. If we look at $13\%$ doping $(n=1.13)$, the line
2251: predicts $T^{\ast}\sim250K$. Experiments that were done without being aware of
2252: this prediction\cite{Matsui:2005} have verified it. It would be most
2253: interesting to do neutron scattering experiments on the same samples to check
2254: whether the antiferromagnetic correlation length $\xi$ and the thermal de
2255: Broglie wave length $\xi_{th}$ are comparable at that temperature, as
2256: predicted by TPSC. Fig.~\ref{fig_127} also predicts that the pseudogap induced
2257: by antiferromagnetic fluctuations will disappear at the quantum critical point
2258: where long-range antiferromagnetic order disappears, in other words it will
2259: coincide with the crossing of the experimentally observed N\'{e}el temperature
2260: (dashed line with triangles in Fig.~\ref{fig_127}) with the zero temperature
2261: axis (if that crossing is not masked by the superconducting transition).
2262: Recent optical conductivity experiments\cite{Millis:2004, Lobo:2005} confirm
2263: this prediction as well.
2264:
2265: In TPSC, superconducting fluctuations can also lead to a pseudogap by an
2266: analogous mechanism.\cite{Vilk:1997}
2267:
2268: %-----------------------------------------------------------------------
2269: %: FIG:fig_133
2270: \begin{figure}[ptb]
2271: \centerline{\includegraphics[width=6.5cm]{f133}} \caption{The generic phase
2272: diagram of high-$T_{c}$ superconductors, from
2273: Ref.~\onlinecite{Damascelli:2003}. There should also be a pseudogap line on
2274: the electron-doped side. It was not well studied at the time of publication of
2275: that paper.}%
2276: \label{fig_133}%
2277: \end{figure}
2278: %-----------------------------------------------------------------------
2279:
2280:
2281: \subsection{The phase diagram for high-temperature superconductors}
2282:
2283: The main features appearing in the phase diagram of high-temperature
2284: superconductors are the pseudogap phase, the antiferromagnetic phase and the
2285: d-wave superconducting phase. Fig.~\ref{fig_133}\cite{Damascelli:2003} shows
2286: the typical diagram with hole doping to the right and electron doping to the
2287: left. Zero on the horizontal axis corresponds to half-filling. There are other
2288: features on the phase diagram, in particular checkerboard
2289: patterns\cite{Hanaguri:2004} or stripe phases\cite{Stock:2004} that appear in
2290: general close to the region where antiferromagnetism and superconductivity
2291: come close to each other. Before we try to understand these more detailed
2292: features, one should understand the most important phases. In the previous
2293: subsection we have discussed the pseudogap phase, in particular on the
2294: electron-doped side (not indicated on Fig. \ref{fig_133}). A recent review of
2295: the pseudogap appears in Ref. \onlinecite{Norman:2005}. In the following, we
2296: discuss in turn the phase diagram and then the nature of the superconducting phase
2297: itself and its relation to the Mott phenomenon.
2298:
2299: %-----------------------------------------------------------------------
2300: %: FIG:fig_136
2301: \begin{figure}[ptb]
2302: \centerline{\includegraphics[width=6.5cm]{f136}}\caption{Antiferromagnetic
2303: order parameter $m$ (dashed) and d-wave (solid) order parameter obtained from
2304: CDMFT on a $2 \times2$ cluster. The result obtained by forcing $m=0$ is also
2305: shown as a thin dashed line.}%
2306: \label{fig_136}%
2307: \end{figure}
2308: %-----------------------------------------------------------------------
2309:
2310:
2311: \subsubsection{Competition between antiferromagnetism and superconductivity}
2312:
2313: We have already shown in Fig.~\ref{fig_96} the prediction of VCPT for the
2314: zero-temperature phase diagram.\cite{Senechal:2005} Here, we just point out
2315: how closely the position of the antiferromagnetic phase boundary, appearing in
2316: the lower panel, coincides with the experimental phase diagram in
2317: Fig.~\ref{fig_133}. (Note that electron concentration increases from right to
2318: left on this experimental phase diagram). In particular, there is little size
2319: dependence to the position of this boundary, (6 to 10 sites) and in addition
2320: the dependence on the value of $U$ is also weak, as can be seen from
2321: Fig.~\ref{fig_97}. Hence, the positions of the antiferromagnetic phase
2322: boundaries is a robust prediction of VCPT. The CDMFT result for a four site
2323: cluster in a bath is shown in Fig.~\ref{fig_136} for $t'=-0.3t,t^{\prime
2324: \prime}=0$ and $U=8t$. The agreement with experiment is not as good. Despite
2325: the useful presence of a bath in CDMFT, the cluster itself is of size
2326: $2\times2,$ which is probably smaller than the Cooper pair size. We can obtain results closer to those of VCPT by increasing the variational space.
2327:
2328: The d-wave superconducting order parameter on the top panel of
2329: Fig.~\ref{fig_96} shows more size dependence than the antiferromagnetic order
2330: parameter. Nevertheless, there are some clear tendencies: (a) d-wave
2331: superconductivity can exist by itself, without antiferromagnetism. The
2332: vertical lines indicate the location of the end of the antiferromagnetic phase
2333: for the various system sizes to help this observation. (b) The range where
2334: d-wave-superconductivity exists without antiferromagnetism, is about three
2335: times larger on the hole than on the electron-doped side, as observed
2336: experimentally. (c) As system size increases, the maximum d-wave order
2337: parameter is larger on the hole than on the electron-doped side. (d) The
2338: tendency to have coexisting antiferromagnetism and d-wave superconductivity is
2339: rather strong on the electron-doped side of the phase diagram. This is
2340: observed experimentally\cite{Fournier:1998} but only over a rather narrow
2341: region near optimal doping. Recent experiments\cite{Motoyama:unpub} challenge
2342: this result, others\cite{Sonier:2003, Sonier:2004} indicate that
2343: antiferromagnetism can be induced from the d-wave superconducting phases with
2344: very small fields. (e) On the hole-doped side, d-wave superconductivity and
2345: antiferromagnetism coexist for a very narrow range of dopings for system size
2346: $N_{c}=6$, for a broad range extending to half-filling for $N_{c}=8$ and not
2347: at all for $N_{c}=10$. In other words, the tendency to coexistence is not even
2348: monotonic. We interpret this result as a reflection of the tendency to form
2349: stripes observed experimentally on the hole-doped side.\cite{Stock:2004,
2350: Wakimoto:2000, Wakimoto:2001} We cannot study systems large enough to allow
2351: for striped inhomogeneous states to check this statement.
2352:
2353: %-----------------------------------------------------------------------
2354: %: FIG:fig_136b
2355: \begin{figure}[ptb]
2356: \centerline{\includegraphics[width=6.5cm]{f136b}} \caption{Phase diagram
2357: obtained from DCA for $U=8t$ for the two-band model. From
2358: Ref.~\onlinecite{Macridin:2005}}%
2359: \label{fig_136b}%
2360: \end{figure}
2361: %-----------------------------------------------------------------------
2362:
2363:
2364: The more realistic two-band model has also been studied using
2365: DCA.\cite{Macridin:2005} The results are shown on Fig. \ref{fig_136b}.
2366: Electron concentration increases from right to left. This phase diagram is
2367: very close to that obtained with the same method from the one-band Hubbard
2368: model\cite{Macridin:2005} with $t^{\prime}=-0.3t,$ $t''=0$,
2369: $U=8t$. The qualitative results agree with the other calculations and with
2370: experiment: antiferromagnetism extends over a narrower doping range for hole
2371: than for electron doping and d-wave superconductivity by itself exists over a
2372: broader range for the hole-doped case than for the electron-doped case. The
2373: actual ranges where antiferromagnetism and d-wave superconductivity exist are
2374: not in as good an agreement with experiment as in the VCPT case. However, as
2375: in CDMFT, the system sizes, $2\times2$, are very small. Overall then, quantum
2376: cluster methods, VCPT in particular, allow us to obtain from the Hubbard model
2377: the two main phases, antiferromagnetic and d-wave superconducting, essentially
2378: in the observed doping range of the zero-temperature phase diagram. At finite
2379: temperature, DCA and TPSC agree on the value of $T_{c}$ for the particle-hole
2380: symmetric model at $10\%$ doping and $U=4t$. Recent studies of the irreducible
2381: vertex using DCA\cite{Maier:2005b} also show that in the weak-coupling
2382: limit the particle-particle d-wave
2383: channel leads to an instability driven by antiferromagnetic fluctuations as
2384: temperature decreases, as found in TPSC.
2385:
2386: To understand the effect of pressure on the phase diagram, note that $U/t$
2387: should decrease as pressure increases since the increase in the overlap
2388: between orbitals should lead mainly to an increase in $t$. Hence, as can be
2389: deduced from Fig.~\ref{fig_104}, applying pressure should lead to a decrease
2390: in the value of $T_{c}$ at weak coupling, concomitant with the decrease in
2391: antiferromagnetic fluctuations that lead to pairing in the weak coupling case.
2392: This is indeed what pressure does experimentally in the case of electron-doped
2393: high-temperature superconductors,\cite{Maple:1990} reinforcing our argument
2394: that near optimal doping they are more weakly coupled. It is
2395: widely known on the other hand that pressure \textit{increases} $T_{c}$ in
2396: hole-doped systems. That is consistent with the strong-coupling result that we
2397: found in VCPT and CDMFT, namely that the maximum d-wave order parameter in
2398: that case scales with $J=4t^{2}/U,$ a quantity that increases with $t$ and
2399: hence pressure. Whereas in the weak coupling case superconductivity is a
2400: secondary phenomenon that occurs after antiferromagnetic fluctuations have
2401: built up, in strong coupling they can be two distinct phenomena as can be seen
2402: from the phase diagram, even though they arise from the same microscopic
2403: exchange interaction represented by $J$.
2404:
2405: \subsubsection{Anomalous superconductivity near the\ Mott transition}
2406:
2407: Superconductivity in the underdoped regime is very much non-BCS. First of all,
2408: we notice in Fig.~\ref{fig_139} obtained in CDMFT\cite{Kancharla:2005} that at
2409: strong coupling the d-wave superconducting order parameter vanishes as we move
2410: towards half-filling even in the absence of long-range antiferromagnetic
2411: order. In other words, the Mott phenomenon by itself suffices to destroy
2412: d-wave superconductivity. This conclusion is reinforced by the fact that at
2413: weak coupling ($U=4t$) where there is no Mott localization, d-wave
2414: superconductivity survives at half-filling. In the presence of
2415: antiferromagnetic long-range order, that last statement would not be true, as
2416: confirmed by VCPT calculations in Fig.~\ref{fig_1}: at $U=4t$ d-wave
2417: superconductivity survives at half-filling if we do not allow for
2418: antiferromagnetic long-range order but it disappears if we do. In BCS theory,
2419: the presence of an interaction $J$ that leads to attraction in the d-wave
2420: channel would lead at $T=0$ to d-wave superconductivity at all dopings
2421: including half-filling, unless we allow for competing long-range order. At
2422: strong coupling, no long-range order is necessary to destroy d-wave superconductivity.
2423:
2424: Superconductivity at strong coupling\cite{Haslinger:2003, Marsiglio:2005} also
2425: differs from BCS in the origin of the condensation energy. Suppose we do BCS
2426: theory on the attractive Hubbard model. Then, as in the usual BCS model,
2427: kinetic energy is increased in the superconducting state because the Fermi
2428: surface is no-longer sharp. On the other hand, in the superconducting phase
2429: there is a gain in potential energy. The reverse is true at strong coupling. This result follows from DCA\cite{Maier:2004}
2430: and is in agreement with the kinetic energy drop in the superconducting state
2431: that has been estimated from the \textit{f-}sum rule in optical conductivity
2432: experiments.\cite{Santander-Syro:2003, Molegraaf:2002,Deutscher:2005}
2433: Photoemission data\cite{Norman:2000} had also suggested this kinetic energy
2434: drop in the superconducting state. A crossover from non-BCS-like to BCS
2435: behavior in the condensation mechanism as we go from underdoping to overdoping
2436: has also been seen recently experimentally.\cite{Deutscher:2005} We do not
2437: seem to have the resolution to find that crossover since the condensation
2438: energy becomes very small on the overdoped side. We expect that crossover from
2439: strong to weak coupling will also lead to a change from a kinetic-energy
2440: driven to a potential-energy driven pairing mechanism. This is confirmed by
2441: CDMFT calculations for the attractive Hubbard model.\cite{Kyung:2005a}
2442:
2443: %-----------------------------------------------------------------------
2444: %%: FIG:fig_141
2445: \begin{figure}[ptb]
2446: \centerline{\includegraphics[width=6.5cm]{f141}}\caption{The gap
2447: in the density of states of the dSC as a function of filling for $U=8t$, $t^{\prime}=-0.3t$ as calculated in CDMFT on a
2448: $2 \times2$ cluster. From Ref.~\onlinecite{Kancharla:2005}.}%
2449: \label{fig_141}%
2450: \end{figure}
2451: %-----------------------------------------------------------------------
2452: A third way in which superconductivity in the underdoped regime is non-BCS is
2453: that the drop in the order parameter as we go towards half-filling is
2454: accompanied by an increase in the gap as measured in the
2455: single-particle density of states. Fig.~6 of Ref.~\onlinecite{Sutherland:2003}
2456: summarizes the experimental evidence for the increase in the size of the
2457: gap. That increase, observed in the CDMFT calculation of the
2458: gap, is illustrated in Fig.~\ref{fig_141}.\cite{Kancharla:2005} That gap has essentially the same size as that observed in the normal pseudogap state\cite{Kyung:2005}.
2459:
2460: \section{Conclusion, open problems\label{Conclusion}}
2461:
2462: High-temperature superconductivity has forced both experimentalists and
2463: theorists to refine their tools and to develop new ones to solve the puzzles
2464: offered by this remarkable phenomenon. From a theoretical perspective, the
2465: original suggestion of Anderson\cite{Anderson:1987} that the physics was in
2466: the one-band Hubbard model is being confirmed. In the absence of \textit{ab
2467: initio} methods to tell us what is the correct starting point, such insight is
2468: essential. The non-perturbative nature of the phenomenon has however forced
2469: theorists to be extremely critical of each other's theories since none of them
2470: can pretend that a small parameter controls the accuracy of the approximations.
2471:
2472: If theorists are to convince each other and experimentalists that a solution
2473: of the high-temperature superconductivity problem has been found, then the
2474: theories have to give quantitative results and to make predictions. Unlike
2475: most traditional problems in condensed matter physics however, the
2476: non-perturbative nature of the problem means that no simple mean-field like
2477: theory can be trusted, even if it seems to agree qualitatively with
2478: experiment. In fact several such theories have been
2479: proposed\cite{Kotliar:1988, Inui:1988, Anderson:2005, Bickers:1987} not long after the
2480: experimental discovery of the phenomenon but they have not been accepted
2481: immediately. Theories have to be internally consistent, they have to agree
2482: with exact results whenever they are available, and then they can be compared
2483: with experiments. If there is a disagreement with experiment, the starting
2484: point (one-band Hubbard model) needs to be reconsidered. When approaches
2485: developed on the basis of weak-coupling ideas agree at intermediate coupling
2486: with approaches developed on the basis of strong-coupling ideas, then one
2487: gains confidence in the validity of the results. We have argued that such
2488: concordance is now found in a number of cases and that corresponding rather
2489: detailed quantitative agreement with experiment can be found. In a
2490: non-perturbative context it becomes essential to also cross check various approaches.
2491:
2492: The main theoretical methods that we have discussed are those that we have
2493: developed or perfected or simply used in our group: The Two-Particle
2494: Self-Consistent approach that is based on weak-coupling but non-perturbative
2495: ideas (no diagrams are involved), as well as heavily numerical approaches such
2496: as QMC and various quantum cluster methods, VCPT and CDMFT.
2497:
2498: Based on our own work and that of many others, we think the following
2499: experimental facts about high-temperature superconductivity can be reproduced
2500: very accurately by calculations for the one-band Hubbard model with $U$ in the
2501: intermediate coupling range ($U\sim8t$) with $t\sim350$~meV, and hopping
2502: parameters $t^{\prime}$ and $t^{\prime\prime}$ close to the values suggested
2503: by band structure calculations,\cite{Andersen:1995} namely $t^{\prime}=-0.3t,$
2504: $t^{\prime\prime}=0.2t$.
2505:
2506: (i) In the one-band Hubbard model the main phases of the zero-temperature
2507: phase diagram, namely antiferromagnetic and d-wave superconducting, appear
2508: very near the observed ranges for both the hole- and electron-doped cases.
2509:
2510: (ii) The normal state is unstable to a d-wave superconducting phase in a
2511: temperature range that has the correct order of magnitude. As usual the value
2512: of $T_{c}$ is the most difficult quantity to evaluate since one must take into
2513: account Kosterlitz-Thouless physics as well as the effect of higher dimensions
2514: etc, so this level of agreement must be considered satisfying.
2515:
2516: (iii) The ARPES MDC at the Fermi energy and the EDC near the Fermi energy are
2517: qualitatively well explained by cluster calculations for both hole- and
2518: electron-doped cases. These comparisons, made at a resolution of about $30$ to
2519: $60$ meV are not very sensitive to long-range order, although order does
2520: influence the results. One is mainly sensitive to the pseudogap, so this is
2521: the main phenomenon that comes out from the model. Energy resolution is not
2522: good enough to see a kink. More details about what aspects of ARPES are
2523: understood may be found in Sec. \ref{ARPES_overview}.
2524:
2525: (iv) In the case of electron-doped cuprates, the value of $U$ near optimal
2526: doping seems to be in the range $U\sim6t$, which means that it is accessible
2527: to studies with TPSC that have better resolution. In that case, the agreement
2528: with experiment is very accurate, even if there is room for improvement and a need for further experiments. In
2529: addition, the value of $T^{\ast}$ for 13\% doping has been predicted
2530: theoretically before it was observed experimentally, one of the very rare
2531: predictions in the field of high-temperature superconductors. All of this
2532: agreement with ARPES data is strong indication that $U\sim6t$ is appropriate
2533: to describe electron-doped superconductors near optimal doping. Additional
2534: arguments come from the pressure dependence of the superconducting transition
2535: temperature $T_{c},$ which increases with $t/U$ contrary to the
2536: strong-coupling result, and from simple ideas on Thomas-Fermi screening. The
2537: latter would predict that the screened interaction scales like $\partial
2538: \mu/\partial n$ and CPT results do lead to $\partial\mu/\partial n$ smaller on
2539: the electron- than on the hole-doped side.\cite{Senechal:2004} In addition, the optical gap at half-filling is smaller in electron- than in hole-doped systems.
2540:
2541: What is the physics? The physics of the antiferromagnetic phase at both weak
2542: and strong coupling is well understood and needs no further comment. For the
2543: pseudogap, we have argued that there seems to be two mechanisms, a weak
2544: coupling one that involves scattering off critical fluctuations and that is
2545: very well understood within TPSC, and a strong-coupling one where there is no
2546: need for large correlation lengths. There is no simple physical picture for
2547: the latter mechanism although the fact that it does not scale
2548: with $J$ but with $t$ seems to suggest forbidden hopping.
2549: The pseudogap is clearly different from the Mott gap.
2550: Whether there is a phase transition as a function of $U$ that separates the
2551: weak and strong coupling regimes or whether there is only a crossover is an
2552: open question. The shape of the MDC's at the Fermi energy clearly show in any
2553: case that in some directions wave vector is not such a bad quantum number
2554: whereas in the pseudogap direction, a \textquotedblleft
2555: localized\textquotedblright\ or \textquotedblleft almost
2556: localized\textquotedblright\ particle-like picture would be appropriate. In
2557: fact the pseudogap occurs near the intersection with the antiferromagnetic
2558: zone boundary that turns out to also be the place where umklapp processes are
2559: possible. In other words, the presence of a lattice is extremely important for
2560: the appearance of the pseudogap. We have seen that with spherical Fermi
2561: surfaces the Fermi liquid survives even for large $U$. The dichotomy between
2562: the wave description inherent to the Fermi liquid and the particle (localized)
2563: description inherent to the Mott phenomenon seems to be resolved in the
2564: pseudogap phase by having certain directions where electrons are more
2565: wave-like and other directions where particle-like (gapped) behavior appears.
2566: The latter behavior appears near regions where the presence of the lattice is
2567: felt through umklapp processes.
2568:
2569: It is clear that when weak-coupling-like ideas of quasiparticles scattering
2570: off each other and off collective excitations do not apply, a simple physical
2571: description becomes difficult. In fact, knowing the exact wave functions would
2572: give us the solution but we would not know how to understand \textquotedblleft
2573: physically\textquotedblright\ the results.
2574:
2575: This lack of simple physical images and the necessity to develop a new
2576: discourse is quite apparent for d-wave superconductivity. At weak coupling
2577: exchange of slow antiferromagnetic fluctuations is at the origin of the
2578: phenomenon, while at strong-coupling the fact that the maximum value of the
2579: d-wave order parameter scales with $J$ tells us that this microscopic coupling
2580: is important, even though there is no apparent boson exchange. This is where
2581: mean-field like theories\cite{Kotliar:1988, Inui:1988, Anderson:2005} or
2582: variational approaches\cite{Paramekanti:2004} can help when they turn out to
2583: give results that are confirmed by more accurate and less biased methods.
2584:
2585: There are many open problems, some of which are material dependent and hence
2586: may depend on interactions not included in the simplest Hubbard model. We have
2587: already mentioned the problem of the chemical potential shift in ARPES for
2588: very small dopings\cite{Shen:2005} that seems to be somewhat material
2589: dependent\cite{Yoshida:2005}. It would also be important to understand
2590: additional inhomogeneous phases that are observed in certain high-temperature
2591: superconductors. That is extremely challenging for quantum cluster methods and
2592: unlikely to be possible in the very near future, except for inhomogeneities of
2593: very short wave length. Also, we still need to improve concordance between the
2594: methods before we can make predictions that are quantitative at the few
2595: percent level for all physical quantities. Apart from DCA, there are no
2596: quantum cluster methods that have been developed yet to study two-particle
2597: response functions that are necessary to obtain results on the superfluid
2598: density and on transport in general. Transport studies are being completed in
2599: TPSC.\cite{Allen:unpub2} Such studies are crucial since they are needed to
2600: answer questions such as: (i) Why is it that for transport properties, such as
2601: optical conductivity, the number of carriers appears to scale with doping
2602: whereas in ARPES the surface of the Brillouin zone enclosed by the apparent
2603: Fermi surface appears to scale with the number of electrons? Is it because the
2604: weight of quasiparticles at the Fermi surface scales like the doping or
2605: because of vertex corrections or because of both? (ii) Can we explain a
2606: vanishing superfluid density as doping goes to zero\cite{Broun:2005} only
2607: through Mott physics or can competing order do the job.\footnote{I.~Herbut,
2608: private communication},\cite{Micnas:2005}
2609:
2610: After twenty years all the problems are not solved, but we think that we can
2611: say with confidence that the essential physics of the problem of
2612: high-temperature superconductivity is in the one-band Hubbard model. At least
2613: the pseudogap, the antiferromagnetic and the d-wave superconducting phases
2614: come out from the model. Refinements of that model may however be necessary as
2615: we understand more and more details of the material-specific experimental results.
2616:
2617: Has a revolution been necessary to understand the basic physics of
2618: high-temperature superconductors? Certainly, it has been necessary to change
2619: our attitude towards methods of solution. We have seen that to study
2620: intermediate coupling, even starting from weak coupling, it has been necessary
2621: to drop diagrams and to rely instead on sum rules and other exact results to
2622: devise a non-perturbative approach. At strong coupling we had to accept that
2623: numerical methods are essential for progress and that we need to abandon some
2624: of the traditional physical explanations of the phenomena in terms of
2625: elementary excitations. Even though progress has been relatively slow, the
2626: pace is accelerating in the last few years and there is hope that in a few
2627: years the problem will be considered for the most part solved. The theoretical
2628: methods (numerical and analytical) that have been developed and that still
2629: need to be developed will likely remain in the tool box of the theoretical
2630: physicist and will probably be useful to understand and perhaps even design
2631: other yet undiscovered materials with interesting properties. The success will
2632: have been the result of the patient and focused effort of a large community of
2633: scientists fascinated by the remarkable phenomenon of high-temperature superconductivity.
2634:
2635: \begin{acknowledgments}
2636: The present work was supported by NSERC (Canada), FQRNT (Qu\'{e}bec), CFI
2637: (Canada), CIAR, the Tier I Canada Research chair Program (A.-M.S.T.). We are
2638: grateful to our collaborators, G. Albinet, S.~Allen, M.~Boissonneault,
2639: C.~Brillon, M.~Capone, L.~Chen, M.~Civelli, A.-M. Dar\'{e}, B.~Davoudi, J.-Y.
2640: P. Delannoy, A.~Gagn\'e-Lebrun, M. J. P. Gingras, A.~Georges, M.~Guillot, V. Hankevych, P. C. W.
2641: Holdsworth, F.~Jackson, S.~Kancharla, G.~Kotliar, J.-S.~Landry, P.-L.~Lavertu, F.~Lemay, S.~Lessard,
2642: M.-A.~Marois, S.~Pairault, D.~Perez, M.~Pioro-Ladri\`{e}re, D.~Plouffe,
2643: D.~Poulin, L. Raymond, S.~Roy, P.~Sahebsara, H.~Touchette, and especially
2644: Y.M.~Vilk. We also acknowledge useful discussions with P.~Fournier, M. Greven,
2645: I.~Herbut, K. Shen and L.~Taillefer and we are grateful to V. Hankevych and S.
2646: Kancharla for permission to include some of their unpublished figures in this paper.
2647: \end{acknowledgments}
2648:
2649: \appendix
2650:
2651: \section{List of acronyms}
2652:
2653: \begin{description}
2654: \item ARPES: Angle Resolved Photoemission Spectroscopy: Experiment from which one can extract $A(\mathbf{k,}\omega)f(\omega)$.
2655:
2656: \item CPT: Cluster Perturbation Theory: Cluster method based on strong
2657: coupling perturbation theory.\cite{Gros:1993, Senechal:2000,Senechal:2002}
2658:
2659: \item CDMFT: Cellular Dynamical Mean Field Theory: A cluster generalization
2660: of DMFT that allows one to take into account both wave vector and frequency
2661: dependence of the self-energy.\cite{Kotliar:2001} It is best formulated in
2662: real space.
2663:
2664: \item DCA: Dynamical Cluster approximation: A cluster generalization of DMFT
2665: that allows one to take into account both wave vector and frequency dependence
2666: of the self-energy based on coarse graining of the self-energy in reciprocal
2667: space.\cite{Hettler:1998, Hettler:2000}
2668:
2669: \item DMFT: Dynamical Mean Field Theory: This approach is exact in infinite
2670: dimension. It takes the frequency dependence of the self-energy into account
2671: and includes both the Mott and the Fermi liquid
2672: limits.\cite{Georges:1996,Jarrell:1992}
2673:
2674: \item EDC: Energy Dispersion Curves: A representation of $A\left(
2675: \mathbf{k,}\omega\right) f\left( \omega\right) $ at fixed $\mathbf{k}$ as a
2676: function of $\omega.$
2677:
2678: \item FLEX: Fluctuation Exchange Approximation: A conserving many-body
2679: approach, similar in spirit to Eliashberg theory.\cite{Bickers:1989}
2680:
2681: \item MDC: Momentum Dispersion Curves: A representation of $A(
2682: \mathbf{k,}\omega)f(\omega)$ at fixed $\omega$ as a
2683: function of $\mathbf{k}$.
2684:
2685: \item QMC: Quantum Monte Carlo: Determinental
2686: approach\cite{Blankenbecler:1981}. This provides an essentially exact solution
2687: to the model for a given system size and within statistical errors that can be
2688: made smaller by performing more measurements.
2689:
2690: \item RPA: Random Phase Approximation.
2691:
2692: \item TPSC: Two-Particle Self-Consistent Approach: Based on sum rules and
2693: other constraints, allows to treat the Hubbard model non-perturbatively in the
2694: weak to intermediate coupling limit.\cite{Vilk:1997, Allen:2003}
2695:
2696: \item VCA: Variational Cluster Approach. Analogous to \textit{CDMFT} but
2697: with a convergence criterion based on an extremum principle. In the
2698: applications quoted here, there is no bath.\cite{Potthoff:2003a,
2699: Potthoff:2003b, Potthoff:2005, Dahnken:2004}
2700:
2701: \item VCPT: Variational Cluster Perturbation Theory. In the present paper
2702: synonymous with \textit{VCA.}
2703:
2704: \end{description}
2705:
2706: %==============================================================================
2707: \begin{thebibliography}{149}
2708:
2709: \bibitem{Anderson:1987} P.~ Anderson, Science \textbf{235}, 1196 (1987).
2710:
2711: \bibitem{Georges:1996} A.~ Georges, G.~ Kotliar, W.~ Krauth and M.~ Rozenberg, Rev. Mod. Phys. \textbf{68}, 13 (1996).
2712:
2713: \bibitem{Jarrell:1992} A.~ Georges and G.~ Kotliar, Phys. Rev. B \textbf{45}, 6479 (1992); M.~ Jarrell, Phys. Rev. Lett. \textbf{69}, 168 (1992).
2714:
2715: \bibitem{Paramekanti:2004} A.~ Paramekanti, M.~ Randeria and N.~ Trivedi, Phys. Rev. B \textbf{70}, 54504 (2004).
2716:
2717: \bibitem{Hirsch:1988} J.~ Hirsch and H.~ Lin, Phys. Rev. B \textbf{37}, 5070 (1988).
2718:
2719: \bibitem{Beal-Monod:1986} M.~Beal-Monod, C.~ Bourbonnais and V.~ Emery, Phys. Rev. B \textbf{34}, 7716 (1986).
2720:
2721: \bibitem{Miyake:1986} K.~ Miyake, S.~Schmitt-Rink and C.~ Varma, Phys. Rev. B \textbf{34}, 6554 (1986).
2722:
2723: \bibitem{Scalapino:1986} D.~ Scalapino, E.~Loh and J.~ Hirsch, Phys. Rev. B \textbf{34}, 8190 (1986).
2724:
2725: \bibitem{Kohn:1965} W.~ Kohn and J.~M. Luttinger, Phys. Rev. Lett. \textbf{15}, 524 (1965).
2726:
2727: \bibitem{Leggett:1975} A.~ Leggett, Rev. Mod. Phys. \textbf{47}, 331 (1975).
2728:
2729: \bibitem{Bulut:1993} N.~Bulut, D.J.~Scalapino, S.R.~White, Phys. Rev. B \textbf{47}, 14 599 (1993).
2730:
2731: \bibitem{Zhang:1997} S.~ Zhang, J.~ Carlson and J.~E. Gubernatis, Phys. Rev. Lett. \textbf{78}, 4486 (1997).
2732:
2733: \bibitem{Inui:1988} M.~ Inui, S.~ Doniach, P.~ Hirschfeld and A.~ Ruckenstein, Phys. Rev. B \textbf{37}, 2320 (1988).
2734:
2735: \bibitem{Kotliar:1988} G.~ Kotliar and J.~ Liu, Phys. Rev. Lett. \textbf{61}, 1784 (1988).
2736:
2737: \bibitem{Maier:2000a} T.~ Maier, M.~ Jarrell, T.~ Pruschke and J.~ Keller, Phys. Rev. Lett. \textbf{85}, 1524 (2000).
2738:
2739: \bibitem{Poilblanc:2002} D.~ Poilblanc and D.~ Scalapino, Phys. Rev. B \textbf{66}, 052513 (2002).
2740:
2741: \bibitem{Kancharla:2005} S.~S. Kancharla, M.~ Civelli, M.~ Capone, B.~ Kyung, D.~S\'en\'echal, G.~ Kotliar and A.-M. Tremblay, \emph{Anomalous superconductivity in doped Mott insulators} (2005), \eprint{cond-mat/0508205}.
2742:
2743: \bibitem{Maier:2005a} T.~ Maier, M.~ Jarrell, T.~ Schulthess, P.~ Kent and J.~ White, hys. Rev. Lett. 95, 237001 (2005).
2744:
2745: \bibitem{Senechal:2005} D.~S\'en\'echal, P.-L. Lavertu, M.-A. Marois and A.-M.~S. Tremblay, Phys. Rev. Lett. \textbf{94}, 156404 (2005).
2746:
2747: \bibitem{Moriya:1985} T.~ Moriya, \emph{Spin Fluctuations in Itinerant Electron Magnetism} (Springer-Verlag, 1985).
2748:
2749: \bibitem{Markiewicz:2004} R.~S. Markiewicz, Phys. Rev. B \textbf{70}, 174518 (2004).
2750:
2751: \bibitem{Bickers:1989} N.~ Bickers and D.~ Scalapino, Ann. Phys. (USA)\textbf{193}, 206 (1989).
2752:
2753: \bibitem{Vilk:1997} Y.~ Vilk and A.-M. Tremblay, J. Phys I (France) \textbf{7}, 1309 (1997).
2754:
2755: \bibitem{Vilk:1996} Y.~ Vilk and A.-M. Tremblay, Europhys. Lett. \textbf{33}, 159 (1996).
2756:
2757: \bibitem{Furukawa:1998} N.~ Furukawa, T.~M. Rice and M.~ Salmhofer, Phys. Rev. Lett. \textbf{81}, 3195 (1998).
2758:
2759: \bibitem{Halboth:2000} C.~ Halboth and W.~ Metzner, Phys. Rev. B \textbf{61}, 7364 (2000).
2760:
2761: \bibitem{Hankevych:2003a} V.~ Hankevych and F.~ Wegner, Eur. Phys. J. B \textbf{31}, 333 (2003).
2762:
2763: \bibitem{Honerkamp:2001} C.~ Honerkamp and M.~ Salmhofer, Phys. Rev. Lett. \textbf{87}, 187004 (2001).
2764:
2765: \bibitem{Wegner:2003} F.~ Wegner and V.~ Hankevych, Acta Phys. Pol. B \textbf{34}, 497 (2003).
2766:
2767: \bibitem{Katanin:2004} A.~A. Katanin and A.~P. Kampf, Phys. Rev. Lett. \textbf{93}, 106406 (2004).
2768:
2769: \bibitem{Zanchi:2001} D.~ Zanchi, Europhys. Lett. \textbf{55}, 376 (2001).
2770:
2771: \bibitem{Mahan:2000} G.~ Mahan, \emph{Many-Particle Physics, 3rd edition, Section 6.4.4} (Kluwer/Plenum, 2000).
2772:
2773: \bibitem{Vilk:1995} Y.~ Vilk and A.-M. Tremblay, J. Phys. Chem. Solids (UK) \textbf{56}, 1769 (1995).
2774:
2775: \bibitem{Allen:2003} S.~ Allen, A.-M. Tremblay and Y.~M. Vilk, in \emph{Theoretical Methods for Strongly Correlated Electrons}, edited by D.~S\'en\'echal, C.~Bourbonnais and A.-M. Tremblay, CRM Series in Mathematical Physics, Springer (2003).
2776:
2777: \bibitem{Baym:1962} G.~ Baym, Physical review \textbf{127}, 1391 (1962).
2778:
2779: \bibitem{Moukouri:2000} S.~ Moukouri, S.~ Allen, F.~ Lemay, B.~ Kyung, D.~ Poulin, Y.~ Vilk and A.-M. Tremblay, Phys. Rev. B \textbf{61}, 7887 (2000).
2780:
2781: \bibitem{Vilk:1994} Y.~ Vilk, L.~ Chen and A.-M. Tremblay, Phys. Rev. B \textbf{49}, 13267 (1994).
2782:
2783: \bibitem{Roy:unpub} S.~ Roy, C.~ Brillon and A.-M. Tremblay, unpublished.
2784:
2785: \bibitem{Allen:unpub} S.~ Allen, B.~ Kyung and A.-M. Tremblay, unpublished.
2786:
2787: \bibitem{Kyung:2003} B.~ Kyung, J.-S. Landry and A.~M.~S. Tremblay, Phys. Rev. B \textbf{68}, 174502 (2003).
2788:
2789: \bibitem{Allen:2001} S.~ Allen and A.-M. Tremblay, Phys. Rev. B \textbf{64}, 075115 (2001).
2790:
2791: \bibitem{Lilly:1990} L.~ Lilly, A.~ Muramatsu and W.~ Hanke, Phys. Rev. Lett. \textbf{65}, 1379 (1990).
2792:
2793: \bibitem{Kyung:2001} B.~ Kyung, S.~ Allen and A.-M. Tremblay, Phys. Rev. B \textbf{64}, 075116 (2001).
2794:
2795: \bibitem{White:1989} S.~ White, D.~J. Scalapino and R.~ Sugar, Phys. Rev. B \textbf{39}, 839 (1989).
2796:
2797: \bibitem{Bulut:1993} N.~ Bulut, D.~ Scalapino and S.~ White, Phys. Rev. B \textbf{47}, 2742 (1993).
2798:
2799: \bibitem{Dare:1996} A.~Dar\'e, Y.~ Vilk and A.-M. Tremblay, Phys. Rev. B \textbf{53}, 14236 (1996).
2800:
2801: \bibitem{Chen:1991} L.~ Chen, C.~ Bourbonnais, T.~ Li and A.-M.-S. Tremblay, Phys. Rev. Lett. \textbf{66}, 369 (1991).
2802:
2803: \bibitem{Hankevych:2003} V.~ Hankevych, B.~ Kyung and A.-M. Tremblay, Phys. Rev. B \textbf{68}, 214405 (2003).
2804:
2805: \bibitem{Trivedi:1995} N.~ Trivedi and M.~ Randeria, Phys. Rev. Lett. \textbf{75}, 312 (1995).
2806:
2807: \bibitem{Moreo:1992} A.~ Moreo, D.~J. Scalapino and S.~R. White, Phys. Rev. B \textbf{45}, 7544 (1992).
2808:
2809: \bibitem{Georges:1992} G . Kotliar, S. Y. Savrasov, K. Haule, V. S. Oudovenko, O. Parcollet, C.A. Marianetti, \emph{Electronic Structure Calculations with Dynamical Mean-Field Theory: A Spectral Density Functional Approach} (2005), \eprint{cond-mat/0511085}.
2810:
2811: \bibitem{Gros:1993} C.~ Gros and R.~ Valenti, Phys. Rev. B \textbf{48}, 418 (1993).
2812:
2813: \bibitem{Senechal:2000} D.~S\'en\'echal, D.~ Perez and M.~Pioro-Ladriere, Phys. Rev. Lett. \textbf{84}, 522 (2000).
2814:
2815: \bibitem{Schiller:1995} A.~Schiller and K.~Ingersent Phys. Rev. Lett. \textbf{75}, 113 (1975)
2816:
2817: \bibitem{Hettler:1998} M.~H. Hettler, A.~N. Tahvildar-Zadeh, M.~ Jarrell, T.~ Pruschke and H.~ Krishnamurthy, Phys. Rev. B \textbf{58}, R7475 (1998).
2818:
2819: \bibitem{Potthoff:2003a} M.~ Potthoff, Eur. Phys. J. B \textbf{36}, 335 (2003).
2820:
2821: \bibitem{Potthoff:2003} M.~ Potthoff, M.~ Aichhorn and C.~ Dahnken, Phys. Rev. Lett. \textbf{91}, 206402 (2003).
2822:
2823: \bibitem{Kotliar:2001} G.~ Kotliar, S.~ Savrasov, G.~P\'alsson and G.~ Biroli, Phys. Rev. Lett. \textbf{87}, 186401 (2001).
2824:
2825: \bibitem{Stanescu:2005} T.~ Stanescu and G.~ Kotliar, \emph{Cluster Dynamical Mean Field Theories: A strong coupling perspective} (2005), \eprint{cond-mat/0508302}.
2826:
2827: \bibitem{Stanescu:2004} T.~D. Stanescu and G.~ Kotliar, Phys. Rev. B \textbf{70}, 205112 (2004).
2828:
2829: \bibitem{Maier:2005} T.~ Maier, M.~ Jarrell, T.~ Pruschke and M.~H. Hettler, Rev. Mod. Phys. \textbf{77}, 1027 (2005).
2830:
2831: \bibitem{Senechal:2002} D.~S\'en\'echal, D.~ Perez and D.~ Plouffe, Phys. Rev. B \textbf{66}, 075129 (2002).
2832:
2833: \bibitem{Potthoff:2004} M.~ Potthoff, \emph{Non-perturbative construction of the Luttinger-Ward functional} (2004), \eprint{cond-mat/0406671}.
2834:
2835: \bibitem{Potthoff:2005} M.~ Potthoff, Adv. Solid State Phys. \textbf{45}, 135 (2005).
2836:
2837: \bibitem{Caffarel:1994} M.~ Caffarel and W.~ Krauth, Phys. Rev. Lett. \textbf{72}, 1545 (1994).
2838:
2839: \bibitem{Bolech:2003} C.~J. Bolech, S.~S. Kancharla and G.~ Kotliar, Phys. Rev. B \textbf{67}, 075110 (2003).
2840:
2841: \bibitem{Kyung:2005} B.~ Kyung, S.~ Kancharla, D.~S\'en\'echal, A.~M.~S. Tremblay, M.~ Civelli and G.~ Kotliar, \emph{Short-range correlation induced pseudogap in doped Mott insulators} (2005), \eprint{cond-mat/0502565}, Phys. Rev. B in press.
2842:
2843: \bibitem{Senechal:2003} D.~S\'en\'echal, in \emph{High Performance Computing Systems and Applications and OSCAR symposium}, edited by D.~S\'en\'echal (2003).
2844:
2845: \bibitem{Aryanpour:2005} K.~ Aryanpour, T.~A. Maier and M.~ Jarrell, Phys. Rev. B \textbf{71}, 037101 (2005).
2846:
2847: \bibitem{Biroli:2002} G.~ Biroli and G.~ Kotliar, Phys. Rev. B \textbf{65}, 155112 (2002).
2848:
2849: \bibitem{Biroli:2005} G.~ Biroli and G.~ Kotliar, Phys. Rev. B \textbf{71}, 037102 (2005); B.~Kyung, G.~Kotliar, A. -M. S.~Tremblay,
2850: \emph{Quantum Monte Carlo Study of Strongly Correlated Electrons: Cellular Dynamical Mean-Field Theory}, (2006) \eprint{cond-mat/0601271}.
2851:
2852: \bibitem{Favand:1997} J.~ Favand, S.~ Haas, K.~ Penc, F.~ Mila and E.~ Dagotto, Phys. Rev. B \textbf{55}, R4859 (1997).
2853:
2854: \bibitem{Dahnken:2004} C.~ Dahnken, M.~ Aichhorn, W.~ Hanke, E.~ Arrigoni and M.~ Potthoff, Phys. Rev. B \textbf{70}, 245110 (2004).
2855:
2856: \bibitem{Capone:2004} M.~ Capone, M.~ Civelli, S.~S. Kancharla, C.~ Castellani and G.~ Kotliar, Phys. Rev. B \textbf{69}, 195105 (2004).
2857:
2858: \bibitem{Maier:2002a} T.~A. Maier, O.~ Gonzalez, M.~ Jarrell and T.~ Schulthess, \emph{Two quantum cluster approximations} (2002), \eprint{cond-mat/0205460}.
2859:
2860: \bibitem{Pozgajcic:2004} K.~ Pozgajcic, \emph{Quantitative aspects of the dynamical impurity approach} (2004), \eprint{cond-mat/0407172}.
2861:
2862: \bibitem{Senechal:2004} D.~S\'en\'echal and A.-M.~S. Tremblay, Phys. Rev. Lett. \textbf{92}, 126401 (2004).
2863:
2864: \bibitem{Huscroft:2001} C.~ Huscroft, M.~ Jarrell, T.~ Maier, S.~ Moukouri and A.~ Tahvildarzadeh, Phys. Rev. Lett. \textbf{86}, 139 (2001).
2865:
2866: \bibitem{Grober:2000} C.~ Grober, R.~ Eder and W.~ Hanke, Phys. Rev. B \textbf{62}, 4336 (2000).
2867:
2868: \bibitem{Moreo:1995} A.~ Moreo, S.~ Haas, A.~ Sandvik and E.~ Dagotto, Phys. Rev. B \textbf{51}, 12045 (1995).
2869:
2870: \bibitem{Moukouri:2001} S.~ Moukouri and M.~ Jarrell, Phys. Rev. Lett. \textbf{87}, 167010 (2001).
2871:
2872: \bibitem{Harris:1967} A.Brooks~ Harris and Robert V.~ Lange, Phys. Rev. \textbf{157}, 295 (1967).
2873:
2874: \bibitem{Meinders:1993} M.B.J.~ Meinders, G.A.~ Sawatzky and G.A.~ Sawatzky, Phys. Rev. B \textbf{48}, 3916 (1993).
2875:
2876: \bibitem{Kyung:2003a} B.~ Kyung, J.~ Landry, D.~ Poulin and A.-M.S. Tremblay, Phys. Rev. Lett. \textbf{90}, 099702 (2003).
2877:
2878: \bibitem{Ronning:2003} F.~ Ronning et~al., Phys. Rev. B \textbf{67}, 165101 (2003).
2879:
2880: \bibitem{Armitage:2002} N.~ Armitage et~al., Phys. Rev. Lett. \textbf{88}, 257001 (2002).
2881:
2882: \bibitem{Kyung:2004} B.~ Kyung, V.~ Hankevych, A.-M. Dar\'e and A.-M.~S. Tremblay, Phys. Rev. Lett. \textbf{93}, 147004 (2004).
2883:
2884: \bibitem{Yuan:2005} Q.~ Yuan, F.~ Yuan and C.~S. Ting, Phys. Rev. B
2885: \textbf{72}, 054504 (2005).
2886:
2887: \bibitem{Kusunose:2003} H.~ Kusunose and T.~M. Rice, Phys. Rev. Lett. \textbf{91}, 186407 (2003).
2888:
2889: \bibitem{Kusko:2002} C.~ Kusko, R.~S. Markiewicz, M.~ Lindroos and A.~ Bansil, Phys. Rev. B \textbf{66}, 140513 (2002).
2890:
2891: \bibitem{borejsza:2004} K.~ Borejsza and N.~ Dupuis, Phys. Rev. B \textbf{69}, 085119 (2004).
2892:
2893: \bibitem{Carbotte:1999} J.~ Carbotte, E.~ Schachinger and D.~ Basov, Nature \textbf{401}, 354 (1999).
2894:
2895: \bibitem{Scalapino:1995} D.~ Scalapino, Physics Reports \textbf{250}, 329 (1995).
2896:
2897: \bibitem{Micnas:2005} B. Tobijaszewska and R.~ Micnas, Physica Status Solidi B \textbf{242}, 468 (2005).
2898:
2899: \bibitem{Moreo:1991} A.~ Moreo and D.~J. Scalapino, Phys. Rev. B \textbf{43}, 8211 (1991).
2900:
2901: \bibitem{Scalettar:1991} R.~ Scalettar and R.~ Singh, Phys. Rev. Lett. \textbf{67}, 370 (1991).
2902:
2903: \bibitem{Paramekanti:2001} A.~ Paramekanti, M.~ Randeria and N.~ Trivedi, Phys. Rev. Lett. \textbf{87}, 217002 (2001).
2904:
2905: \bibitem{Sorella:2002} S.~ Sorella, G.~B. Martins, F.~ Becca, C.~ Gazza, L.~ Capriotti, A.~ Parola and E.~ Dagotto, Phys. Rev. Lett. \textbf{88}, 117002 (2002).
2906:
2907: \bibitem{Hanke:2005} W.~ Hanke, M.~ Aichhorn, E.~ Arrigoni and M.~ Potthoff, \emph{\bibinfo{title}{Correlated band structure and the ground-state phase diagram in high-$T_c$ cuprates}} (2005), \eprint{cond-mat/0506364}.
2908:
2909: \bibitem{Pao:1994} C.~ Pao and N.~ Bickers, Phys. Rev. B \textbf{49}, 1586 (1994).
2910:
2911: \bibitem{Deisz:1996} J.~ Deisz, D.~ Hess and J.~ Serene, Phys. Rev. Lett. \textbf{76}, 1312 (1996).
2912:
2913: \bibitem{Zhang:1988a} F.~ Zhang and T.~M. Rice, Phys. Rev. B \textbf{37}, 3759 (1988).
2914:
2915: \bibitem{Andersen:1995} O.~ Andersen, A.~ Liechtenstein, O.~ Jepsen and F.~ Paulsen, Journal of the Physics and Chemistry of Solids \textbf{56}, 1573 (1995).
2916:
2917: \bibitem{Macridin:2005} A.~ Macridin, T.~A. Maier, M.~S. Jarrell and G.~ Sawatzky, Phys. Rev. B \textbf{71}, 134527 (2005).
2918:
2919: \bibitem{Coldea:2001} R.~ Coldea, S.~M. Hayden, G.~ Aeppli, T.~G. Perring, C.~D. Frost, T.~E. Mason, S.-W. Cheong and Z.~ Fisk, Phys. Rev. Lett. \textbf{86}, 5377 (2001).
2920:
2921: \bibitem{Delannoy:2005a} J.-Y. Delannoy, Ph.D. thesis, \'Ecole normale sup\'erieure de Lyon (2005).
2922:
2923: \bibitem{Delannoy:2005} J.-Y.~P. Delannoy, M.~J.~P. Gingras, P.~C.~W. Holdsworth and A.-M.~S. Tremblay, Phys. Rev. B \textbf{72}, 115114 (2005).
2924:
2925: \bibitem{Raymond:2005} L.~ Raymond, G.~ Albinet and A.-M. Tremblay, cond-mat/0510811.
2926:
2927: \bibitem{Toader:2005} A.~M. Toader, J.~P. Goff, M.~ Roger, N.~ Shannon, J.~R. Stewart and M.~ Enderle, Phys. Rev. Lett. \textbf{94}, 197202 (2005).
2928:
2929: \bibitem{Damascelli:2003} A.~ Damascelli, Z.~ Hussain and Z.-X. Shen, Rev. Mod. Phys. \textbf{75}, 473 (2003).
2930:
2931: \bibitem{Damascelli:2003a} A.~ Damascelli, \emph{Probing the low-energy electronic structure of complex systems by ARPES} (2003), \eprint{cond-mat/0307085}.
2932:
2933: \bibitem{Kordyuk:2005} A.~ Kordyuk and S.~V. Borisenko, \emph{ARPES on HTSC: simplicity vs. complexity} (2005), \eprint{cond-mat/0510218}.
2934:
2935: \bibitem{Preuss:1995} P.~ Preuss, W.~ Hanke and \bibinfo{author}{ W.~von~der Linden}, Phys. Rev. Lett. \textbf{75}, 1344 (1995).
2936:
2937: \bibitem{Pairault:2000} S.~ Pairault, D.~S\'en\'echal and A.-M. Tremblay, Eur. Phys. J. B (France) \textbf{16}, 85 (2000).
2938:
2939: \bibitem{Pairault:1998} S.~ Pairault, D.~S\'en\'echal and A.-M.~S. Tremblay, Phys. Rev. Lett. \textbf{80}, 5389 (1998).
2940:
2941: \bibitem{Gooding:1994} R.~ Gooding, K.~ Vos and P.~ Leung, Phys. Rev. B \textbf{50}, 12866 (1994).
2942:
2943: \bibitem{Shen:2005} K.~ Shen, F.~ Ronning, D.~ Lu, F.~ Baumberger, N.~ Ingle, W.~ Lee, W.~ Meevasana, Y.~ Kohsaka, M.~ Azuma, M.~ Takano et~al., Science \textbf{307}, 901 (2005).
2944:
2945: \bibitem{Yoshida:2005} T.~ Yoshida, X.~J. Zhou, K.~ Tanaka, W.~L. Yang, Z.~ Hussain, Z.-X. Shen, A.~ Fujimori, S.~ Komiya, Y.~ Ando, H.~ Eisaki et~al., cond-mat/0510608.
2946:
2947: \bibitem{Zhou:2003} X.~ Zhou, T.~ Yoshida, A.~ Lanzara, P.~V. Bogdanov, S.~A. Kellar, K.~M. Shen, W.~L. Yang, F.~ Ronning, T.~ Sasagawa, T.~ Kakeshita et~al., Nature \textbf{423}, 398 (2003).
2948:
2949: \bibitem{Maier:2004} T.~A. Maier, M.~ Jarrell, A.~ Macridin and C.~ Slezak, Phys. Rev. Lett. \textbf{92}, 027005 (2004); T.~Koretsune and M.~Ogata, J. Phys. Soc. Japan, \textbf{74}, 1390 (2005).
2950:
2951: \bibitem{Gromko:2003} A.~D. Gromko, A.~V. Fedorov, Y.-D. Chuang, J.~D. Koralek, Y.~ Aiura, Y.~ Yamaguchi, K.~ Oka, Y.~ Ando and D.~S. Dessau, Phys. Rev. B \textbf{68}, 174520 (2003).
2952:
2953: \bibitem{Armitage:2003} N.~P. Armitage, D.~H. Lu, C.~ Kim, A.~ Damascelli, K.~M. Shen, F.~ Ronning, D.~L. Feng, P.~ Bogdanov, X.~J. Zhou, W.~L. Yang et~al., Phys. Rev. B \textbf{68}, 064517 (2003).
2954:
2955: \bibitem{Hankevych:2005} V.~ Hankevych, B.~ Kyung, A.-M. Dar\'e, D.~S\'en\'echal and A.-M. Tremblay, in \emph{Proceedings of SNS2004} (2005), J. Chem. Phys. Sol., \textbf{67}, 189 (2006).
2956:
2957: \bibitem{Mang:2004} P.~K. Mang, O.~P. Vajk, A.~ Arvanitaki, J.~W. Lynn and M.~ Greven, Phys. Rev. Lett. \textbf{93}, 027002 (2004).
2958:
2959: \bibitem{Matsuda:1992} M.~ Matsuda, Y.~ Endoh, K.~ Yamada, H.~ Kojima, I.~ Tanaka, R.~J. Birgeneau, M.~A. Kastner and G.~ Shirane, Phys. Rev. B \textbf{45}, 12548 (1992).
2960:
2961: \bibitem{Onose:2004} Y.~ Onose, Y.~ Taguchi, K.~ Ishizaka and Y.~ Tokura, Phys. Rev. B. \textbf{69}, 024504 (2004).
2962:
2963: \bibitem{Onose:2001} Y.~ Onose, Y.~ Taguchi, K.~ Ishizaka and Y.~ Tokura, Phys. Rev. Lett. \textbf{87}, 217001 (2001).
2964:
2965: \bibitem{Matsui:2005} H.~ Matsui, K.~ Terashima, T.~ Sato, T.~ Takahashi, S.-C. Wang, H.-B. Yang, H.~ Ding, T.~ Uefuji and K.~ Yamada, Phys. Rev. Lett. \textbf{94}, 047005 (2005).
2966:
2967: \bibitem{Lobo:2005} R.~P. S.~M. Lobo, N.~ Bontemps, A.~ Zimmers, Y.~ Dagan, R.~ Greene, P.~ Fournier, C.~ Homes and A.~ Millis, unpublished.
2968:
2969: \bibitem{Millis:2004} A.J. Millis, A. Zimmers, R.P.S.M. Lobo, N. Bontemps, C.C. Homes Phys. Rev. B \textbf{72} 224517 (2005).
2970:
2971: \bibitem{Hanaguri:2004} T.~ Hanaguri, C.~ Lupien, Y.~ Kohsaka, D.~ Lee, M.~ Azuma, M.~ Takano, H.~ Takagi and J.~ Davis, Nature \textbf{430}, 1001 (2004).
2972:
2973: \bibitem{Stock:2004} C.~ Stock, W.~J.~L. Buyers, R.~ Liang, D.~ Peets, Z.~ Tun, D.~ Bonn, W.~N. Hardy and R.~J. Birgeneau, Phys. Rev. B \textbf{69}, 014502 (2004).
2974:
2975: \bibitem{Norman:2005} M.~ Norman, D.~ Pines and C.~ Kallin, Adv. Phys. \textbf{54}, 715 (2005).
2976:
2977: \bibitem{Fournier:1998} P.~ Fournier, P.~ Mohanty, E.~ Maiser, S.~ Darzens, T.~ Venkatesan, C.~ Lobb, G.~ Czjzek, R.~ Webb and R.~ Greene, Phys. Rev. Lett. \textbf{81}, 4720 (1998).
2978:
2979: \bibitem{Motoyama:unpub} E.~ Motoyama, P.~ Mang, D.~ Petitgrand, G.~ Yu, O.~ Vajk, I.~ Vishik and M.~ Greven, unpublished.
2980:
2981: \bibitem{Sonier:2003} J.~ Sonier, K.~ Poon, G.~ Luke, P.~ Kyriakou, R.~ Miller, R.~ Liang, C.~ Wiebe, P.~ Fournier and R.~ Greene, Phys. Rev. Lett. \textbf{91}, 147002 (2003).
2982:
2983: \bibitem{Sonier:2004} J.~ Sonier, K.~ Poon, G.~ Luke, P.~ Kyriakou, R.~ Miller, R.~ Liang, C.~ Wiebe, P.~ Fournier and R.~ Greene, Physica C: Superconductivity and its Applications \textbf{408-410}, 783 (2004).
2984:
2985: \bibitem{Wakimoto:2000} S.~ Wakimoto, R.~J. Birgeneau, M.~A. Kastner, Y.~S. Lee, R.~ Erwin, P.~M. Gehring, S.~H. Lee, M.~ Fujita, K.~ Yamada, Y.~ Endoh et~al., Phys. Rev. B \textbf{61}, 3699 (2000).
2986:
2987: \bibitem{Wakimoto:2001} S.~ Wakimoto, J.~M. Tranquada, T.~ Ono, K.~M. Kojima, S.~ Uchida, S.-H. Lee, P.~M. Gehring and R.~J. Birgeneau, Phys. Rev. B \textbf{64}, 174505 (2001).
2988:
2989: \bibitem{Maier:2005b} T.~A. Maier, M.~S. Jarrell and D.~J. Scalapino, Phys. Rev. Lett. \textbf{96}, 047005 (2006).
2990:
2991: \bibitem{Maple:1990} M.~ Maple, MRS Bulletin \textbf{15}, 60 (1990).
2992:
2993: \bibitem{Haslinger:2003} R.~ Haslinger and A.~ Chubukov, Phys. Rev. B \textbf{67}, 140504 (2003).
2994:
2995: \bibitem{Marsiglio:2005} F.~ Marsiglio, Phys. Rev. B \textbf{73}, 064507 (2006).
2996:
2997: \bibitem{Deutscher:2005} G.~ Deutscher, A.~F. Santander-Syro and N.~ Bontemps, Phys. Rev. B \textbf{72}, 092504 (2005).
2998:
2999: \bibitem{Molegraaf:2002} H.~ Molegraaf, C.~ Presura, \bibinfo{author}{ D.~van~der Marel}, P.~ Kes and M.~ Li, Science \textbf{295}, 2239 (2002).
3000:
3001: \bibitem{Santander-Syro:2003} A.~Santander-Syro, R.~ Lobo, N.~ Bontemps, Z.~ Konstatinovic, Z.~ Li and H.~ Raffy, Europhys. Lett. \textbf{62}, 568 (2003).
3002:
3003: \bibitem{Norman:2000} M.~ Norman, M.~ Randeria, B.~ Janko and J.~ Campuzano, Phys. Rev. B \textbf{61}, 14742 (2000).
3004:
3005: \bibitem{Kyung:2005a} B.~ Kyung, A.~ Georges and A.-M. Tremblay, \emph{Potential-energy (BCS) to kinetic-energy (BEC)-driven pairing in the attractive Hubbard model} (2005), \eprint{cond-mat/0508645}.
3006:
3007: \bibitem{Sutherland:2003} M.~ Sutherland, D.~G. Hawthorn, R.~W. Hill, F.~ Ronning, S.~ Wakimoto, H.~ Zhang, C.~ Proust, E.~ Boaknin, C.~ Lupien, L.~ Taillefer et~al., Phys. Rev. B \textbf{67}, 174520 (2003).
3008:
3009: \bibitem{Anderson:2005} P.~ Anderson, \emph{\bibinfo{title}{Present status of the theory of high-$T_c$ cuprates}} (2005), \eprint{cond-mat/0510053} published in this issue of Journal of Low Temperature Physics.
3010:
3011: \bibitem{Bickers:1987} N.E. Bickers, D.J. Scalapino and R.T. Scalettar, Int. J. Mod. Phys. B \textbf{1}, 687 (1987).
3012:
3013: \bibitem{Allen:unpub2} S.~ Allen, B.~ Kyung, D.~ Bergeron, V.~ Hankevych and A.-M. Tremblay, unpublished.
3014:
3015: \bibitem{Broun:2005} D.~M. Broun, P.~J. Turner, W.~A. Huttema, S.~ Ozcan, B.~ Morgan, L.~ R., W.~N. Hardy and D.~A. Bonn, \emph{In-plane superfluid density of highly underdoped YBa$_2$Cu$_3$O$_{6+x}$} (2005), \eprint{cond-mat/0509223}.
3016:
3017: \bibitem{Hettler:2000} M.~ Hettler, M.~ Mukherjee, M.~ Jarrell and H.~ Krishnamurthy, Phys. Rev. B \textbf{61}, 12739 (2000).
3018:
3019: \bibitem{Blankenbecler:1981} R.~ Blankenbecler, D.~J. Scalapino and R.~L. Sugar, Phys. Rev. D \textbf{24}, 2278 (1981).
3020:
3021: \bibitem{Potthoff:2003b} M.~ Potthoff, Eur. Phys. J. B (France) \textbf{32}, 429 (2003).
3022:
3023: \end{thebibliography}
3024: %============================================================================
3025: \end{document}
3026: