cond-mat0207186/rev.tex
1: \documentstyle[amssymb,fullpage,11pt,subfigure,epsfig]{article}
2: %\documentstyle[amssymb,a4,11pt]{article}
3: 
4: \newif\iffigure
5: 
6: % vklop in izklop slik:
7: \figuretrue
8: %\figurefalse
9: 
10: \setcounter{topnumber}{4}
11: \setcounter{totalnumber}{5}
12: \renewcommand{\topfraction}{1.0}
13: \renewcommand{\textfraction}{0.0}
14: \renewcommand{\floatpagefraction}{0.8}
15: 
16: \def\f#1#2{\frac{#1}{#2}}
17: \def\r#1{{\rule[-2.5mm]{0mm}{7mm}#1}}
18: \def\vc#1{{\bf{#1}}}
19: \def\agt{\gtrsim}
20: \def\alt{\lesssim}
21: \def\tsig{\hbox{\boldmath$\sigma$}}
22: \def\tchi{\hbox{\boldmath$\chi$}}
23: \def\ttau{\hbox{\boldmath$\tau$}}
24: 
25: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
26: \def\incl#1{\setcounter{equation}{0}\setcounter{figure}{0}\input{#1}}
27: \renewcommand{\thefigure}{\arabic{section}.\arabic{figure}}
28: 
29: \title{\bf {\boldmath$d_{x^2-y^2}$}-wave superconductivity 
30: and the Hubbard model}
31: \author{N. BULUT\\
32: Department of Physics, Ko\c{c} University \\
33: Sariyer, 80910 Istanbul, Turkey\\ }
34: 
35: \date{} 
36: 
37: \begin{document}
38: 
39: \maketitle
40: 
41: %\centerline{DRAFT}
42: \centerline{to appear in {\it Advances in Physics}, {\bf 51}, no. 6 (2002)}
43: 
44: \bigskip
45: 
46: \begin{abstract}
47: 
48: The numerical studies of $d_{x^2-y^2}$-wave pairing in the 
49: two-dimensional (2D) and the 2-leg Hubbard models
50: are reviewed.
51: For this purpose, the results obtained from the 
52: determinantal Quantum Monte Carlo 
53: and the density-matrix renormalization-group 
54: calculations are presented.
55: These are calculations which 
56: were motivated by the discovery of the 
57: high-$T_c$ cuprates.
58: In this review, the emphasis is placed on 
59: the microscopic many-body processes which are responsible for 
60: the $d_{x^2-y^2}$-wave pairing correlations observed in the 2D and 
61: the 2-leg Hubbard models.
62: In order to gain insight into these processes, 
63: the results on the effective pairing interaction
64: as well as the magnetic, 
65: density and the single-particle excitations will be reviewed.
66: In addition,
67: comparisons will be made with the other numerical approaches
68: to the Hubbard model and the numerical results on the
69: $t$-$J$ model.
70: The results reviewed here 
71: indicate that an effective pairing
72: interaction which is repulsive at $(\pi,\pi)$ momentum transfer
73: and enhanced single-particle spectral weight near the 
74: $(\pi,0)$ and $(0,\pi)$ points of the Brillouin zone 
75: create optimum conditions for $d_{x^2-y^2}$-wave pairing.
76: These are two effects which act to enhance the $d_{x^2-y^2}$-wave 
77: pairing correlations in the Hubbard model.
78: Finding additional ways is an active research problem.
79: 
80: \end{abstract}
81: 
82: %\newpage
83: 
84: \tableofcontents
85: 
86: \newpage
87: %\incl{rev1}
88: %\incl{rev2}
89: %\incl{rev3}
90: %\incl{rev4}
91: %\incl{rev5}
92: %\incl{rev6}
93: %\incl{rev7}
94: %\newpage
95: %\incl{revref}
96: %\incl{revfig}
97: %\end{document}
98: \setcounter{equation}{0}\setcounter{figure}{0}
99: \section{Introduction}
100: 
101: Since the discovery of high-temperature superconductivity
102: in the layered cuprates in 1986
103: [Bednorz and M\"uller 1986],
104: it has been established that 
105: the superconducting order parameter has the $d_{x^2-y^2}$-wave 
106: symmetry in a number of these materials
107: [Schrieffer 1994, Scalapino 1995, van Harlingen 1995,
108: Tsuei and Kirtley 2000]. 
109: This is important, because 
110: the $d_{x^2-y^2}$-wave symmetry of the order parameter
111: suggests the possibility of 
112: an electronically mediated pairing mechanism.
113: Perhaps, the simplest model used for modelling the low-energy
114: electronic correlations of the layered cuprates is 
115: the two-dimensional (2D) Hubbard model.
116: Within this context, the nature of the pairing correlations 
117: in the Hubbard model as well as the nature of its low-lying electronic
118: excitations has received considerable attention. 
119: 
120: In 1987 Anderson suggested that the 2D Hubbard model 
121: is relevant to the cuprates 
122: [Anderson 1987].
123: However, 
124: even today questions remain about this model.
125: In this article, 
126: what has been learned about the physical properties 
127: of the 2D and the 2-leg Hubbard models from the numerical studies 
128: will be reviewed.
129: The emphasis will be placed on the $d_{x^2-y^2}$ pairing 
130: correlations seen in these models and their microscopic origin. 
131: The implications of these calculations 
132: for the $d_{x^2-y^2}$-wave pairing 
133: in the high-$T_c$ cuprates is of current interest. 
134: In particular, 
135: one is interested in knowing whether the 
136: $d_{x^2-y^2}$-wave superconducting order 
137: could exist in the ground
138: state of the 2D Hubbard model, and, if it does, whether
139: it would have sufficient strength to explain the 
140: superconducting transition temperatures as high as
141: those seen in the cuprates.
142: There is also much interest in finding ways of enhancing the 
143: strength of the $d_{x^2-y^2}$-wave pairing correlations observed
144: in the Hubbard model. 
145: Within the past ten to fifteen years, 
146: the determinantal Quantum Monte Carlo (QMC) and 
147: the density-matrix renormalization-group (DMRG) 
148: techniques have been used to address these questions. 
149: These numerical studies are the subject of this review article.
150: 
151: For the 2D Hubbard model, the QMC data 
152: on the magnetic, charge and the single-particle excitations 
153: will be presented.
154: In order to investigate the pairing correlations, 
155: the QMC data on the irreducible particle-particle 
156: interaction and the solutions 
157: of the particle-particle Bethe-Salpeter equation will be shown. 
158: Using the DMRG method, the equal-time pair-field correlation function
159: has been calculated in the ground state of the 2-leg 
160: Hubbard ladder. 
161: These DMRG data along with the QMC results 
162: on the 2-leg ladder will be shown 
163: and compared with each other. 
164: In addition, 
165: these results will be compared with the numerical studies of 
166: the $t$-$J$ model.
167: 
168: Figure 1.1 shows a schematic drawing of the CuO$_2$ plane
169: of the cuprates consisting of the one-electron Cu($3d_{x^2-y^2}$)
170: and O($p_x,p_y$) orbitals that give rise to the 
171: band in which the $d_{x^2-y^2}$-wave superconducting pairs form.
172: In an electronic model of the CuO$_2$ plane, 
173: there would be an onsite Coulomb repulsion at the 
174: Cu($3d_{x^2-y^2}$) and the O($p_x,p_y$) orbitals and one-electron 
175: hopping matrix elements between the neighbouring 
176: Cu-O and the O-O orbitals
177: as well as longer-range hoppings and Coulomb interactions.
178: A simplified approach is to use the single-band Hubbard model
179: on the 2D square lattice which has an onsite Coulomb repulsion
180: and one-electron hopping matrix elements between the 
181: nearest-neighbour sites, as illustrated in Fig.~1.2(a).
182: To some extent, 
183: this is motivated by the notion that the 
184: nonperturbative effects due to the onsite Coulomb repulsion
185: will dominate at low temperatures and low energies.
186: Extensive numerical calculations have been 
187: carried out for studying the properties of the 
188: one-band 2D Hubbard model.
189: Another model which has proven useful for studying 
190: $d_{x^2-y^2}$-wave pairing is the 2-leg Hubbard ladder
191: illustrated in Fig.~1.2(b).
192: This model has an onsite Coulomb repulsion $U$ and 
193: the intrachain and interchain hopping matrix elements 
194: $t$ and $t_{\perp}$, respectively.
195: The 2-leg Hubbard ladder is an important model 
196: where the pairing correlations can be studied in detail 
197: in the ground state for systems with up to 32 rungs.
198: Here, the numerical results on the 2D and the 2-leg 
199: Hubbard models will be used to discuss the nature of the 
200: $d_{x^2-y^2}$-wave pairing found in the CuO$_2$ layers of the 
201: high-$T_c$ cuprates.
202: 
203: \begin{figure}
204: \centering
205: \iffigure
206: \epsfig{file=ch1-fig1.ps,height=7cm,angle=0}
207: \fi
208: \caption{
209: Schematic drawing of the CuO$_2$ lattice consisting of the 
210: one-electron Cu($3d_{x^2-y^2}$) and the O($p_x,p_y$)
211: orbitals.
212: }
213: \label{1.1}
214: \end{figure}
215: 
216: The Hubbard model was introduced 
217: for describing the Coulomb correlation 
218: effects in the $d$-bands of the transition metals, 
219: which show both band-like and atomic-like behaviour
220: [Hubbard 1963].
221: The Hubbard Hamiltonian has a one-body kinetic term
222: and an interaction term which represents the onsite Coulomb
223: repulsion when two electrons occupy the same orbital.
224: In spite of studies covering four decades, 
225: questions remain about the Hubbard model.
226: In two dimensions, the Hubbard Hamiltonian 
227: with only nearest-neighbor hopping is given by
228: \begin{equation}
229: \label{Hubbard}
230: H = -t \sum_{\langle i,j\rangle, \sigma}
231: ( c^{\dagger}_{i\sigma} c_{j\sigma} + 
232: c^{\dagger}_{j\sigma} c_{i\sigma} )
233: + U \sum_i n_{i\sigma} n_{i-\sigma}
234: - \mu \sum_{i\sigma} n_{i\sigma},
235: \end{equation}
236: where the sum over $i$ and $j$ is done over the nearest-neighbour 
237: sites on a square lattice.
238: The one-electron hopping matrix element is $t$, 
239: the onsite Coulomb repulsion is $U$ 
240: and the chemical potential $\mu$ is used for 
241: controlling the electron occupation in the grand canonical 
242: ensemble. 
243: Here, $c^{\dagger}_{i\sigma}$ ($c_{i\sigma}$) 
244: creates (annihilates) an electron of spin $\sigma$ at site $i$
245: and $n_{i\sigma}=c^{\dagger}_{i\sigma} c_{i\sigma}$ 
246: is the occupation number of electrons with spin 
247: $\sigma$ at site $i$.
248: In addition to the 2D Hubbard model, the numerical results 
249: on the 2-leg Hubbard ladder with anisotropic hopping 
250: will be reviewed.
251: In this case, the Hamiltonian is 
252: \begin{eqnarray}
253: \label{Ladder}
254: H = -t \sum_{i,\lambda,\sigma}
255: ( c^{\dagger}_{i,\lambda,\sigma} c_{i+1,\lambda,\sigma} + 
256: c^{\dagger}_{i+1,\lambda,\sigma} c_{i,\lambda,\sigma} )
257: -t_{\perp} \sum_{i,\sigma}
258: ( c^{\dagger}_{i,1,\sigma} c_{i,2,\sigma} + 
259: c^{\dagger}_{i,2,\sigma} c_{i,1,\sigma} ) \nonumber \\ 
260: + U \sum_{i,\lambda} 
261: n_{i,\lambda,\sigma} n_{i,\lambda,-\sigma}
262: - \mu \sum_{i,\lambda,\sigma} n_{i,\lambda,\sigma},
263: \end{eqnarray}
264: where $t$ is the hopping matrix element parallel to the 
265: chains and $t_{\perp}$ is the inter-chain hopping,
266: as illustrated in Fig.~1.2(b).
267: The operators $c^{\dagger}_{i,\lambda,\sigma}$ 
268: create an electron of spin $\sigma$ at site $i$ of the 
269: $\lambda$'th leg, and 
270: $n_{i,\lambda,\sigma}= c^{\dagger}_{i,\lambda,\sigma}
271: c_{i,\lambda,\sigma}$
272: is the electron number operator with spin $\sigma$
273: at site $(i,\lambda)$.
274: 
275: \begin{figure}
276: \centering
277: \iffigure
278: \mbox{
279: \subfigure[]{
280: \epsfysize=6cm
281: \epsffile[100 200 480 560]{ch1-fig2a.ps}}
282: \quad
283: \subfigure[]{
284: \epsfysize=6cm
285: \epsffile[50 200 600 560]{ch1-fig2b.ps}}}
286: \fi
287: \caption{
288: (a) Sketch of the 2D Hubbard lattice with 
289: isotropic nearest-neighbour hopping $t$ and the 
290: onsite Coluomb repulsion $U$.
291: (b) Sketch of the 2-leg Hubbard ladder with onsite Coulomb
292: repulsions $U$, and the intrachain and interchain
293: hopping matrix elements $t$ and $t_{\perp}$, respectively.
294: }
295: \label{1.2}
296: \end{figure}
297: 
298: The QMC and the DMRG are powerful numerical methods,
299: which have been developed for studying 
300: strongly-correlated systems such as the Hubbard model.
301: These techniques allow for the possibility to determine whether
302: superconductivity exists in the ground state of an interacting 
303: system without first having to construct an approximate 
304: theoretical framework for describing its elementary 
305: excitations.
306: The determinantal QMC method was introduced in 
307: Ref.~[Blankenbecler {\it et al.} 1981].
308: The QMC results reviewed here were obtained by using the 
309: algorithm described in Ref.~[White {\it et al.} 1989b].
310: Reviews of this method can be found in 
311: Refs.~[Scalapino 1993, Muramatsu 1999].
312: The DMRG method was developed by White 
313: [White 1992 and 1993],
314: and since then has been used to 
315: calculate the equal-time correlation functions in the 
316: ground state of interacting systems.
317: Here, 
318: in Sections 3 through 7,
319: numerical results on the 2D and 
320: the 2-leg Hubbard models, 
321: which were obtained by using these two algorithms, 
322: will be presented. 
323: In Appendix, 
324: the determinantal QMC and the DMRG techniques 
325: will be described briefly.
326: 
327: In Section 2, an introduction to 
328: the spin-fluctuation mediated $d_{x^2-y^2}$-wave 
329: superconductivity will be given.
330: In addition, here, the experimental evidence 
331: for $d_{x^2-y^2}$-wave superconductivity in the cuprates 
332: and the theoretical studies 
333: of $d_{x^2-y^2}$-wave pairing in the cuprates will be reviewed 
334: briefly.
335: 
336: In order to understand the microscopic 
337: many-body processes causing 
338: $d_{x^2-y^2}$-wave pairing correlations in the 
339: 2D Hubbard model, it is useful to first discuss the magnetic, 
340: density and the single-particle excitations.
341: The review of the numerical studies on the 2D Hubbard model 
342: will begin in Section 3 by presenting QMC 
343: data on the magnetic properties.
344: Here, QMC results on the magnetic susceptibility will be shown
345: at and near half-filling.
346: In addition, an RPA-like model for the magnetic 
347: susceptibility will be introduced, and 
348: the effective particle-hole vertex will be discussed.
349: The NMR experiments which support the existence of short-range 
350: AF fluctuations in the high-$T_c$ cuprates will also be
351: reviewed briefly.
352: 
353: In Section 4, 
354: the QMC results on the charge susceptibility 
355: will be shown and its dependence on the strength of the 
356: Coulomb repulsion will be discussed.
357: Here, 
358: the possibility of any "$4{\bf k}_F$" 
359: charge-density-wave (CDW) fluctuations in the 
360: 2D Hubbard model will be investigated.
361: In this section, 
362: numerical results on the optical conductivity will also be discussed.
363: 
364: The nature of the single-particle excitations in the 2D Hubbard 
365: model has received much attention within the context 
366: of the cuprates.
367: In Section 5, the single-particle 
368: density of states $N(\omega)$ and spectral weight 
369: $A({\bf p},\omega)$ obtained from the maximum-entropy 
370: analytic continuation of the QMC data on the single-particle
371: Green's function will be reviewed.
372: At half-filling, these calculations find 
373: a Mott-Hubbard gap, 
374: lower and upper Hubbard bands and quasiparticle bands.
375: These quasiparticle bands are similar to those found in a
376: spin-density-wave (SDW) insulator.
377: It will be seen that,
378: upon doping, a redistribution of the spectral weight takes
379: place forming a narrow metallic band at the Fermi level.
380: This band has unusually flat dispersion near the 
381: $(\pi,0)$ and $(0,\pi)$ points of the 
382: Brillouin zone leading to large amount of 
383: single-particle spectral weight available 
384: for scatterings in the particle-particle channel.
385: These results on $A({\bf p},\omega)$ will also be
386: compared with the results of the angular-resolved
387: photoemission spectroscopy (ARPES) 
388: measurements on the cuprates.
389: 
390: After reviewing the numerical results on the 
391: magnetic, density and the single-particle properties, 
392: in Section 6 the $d_{x^2-y^2}$-wave pairing correlations in the 
393: 2D Hubbard model will be discussed.
394: The QMC simulations have found that 
395: when doped away from half-filling
396: there are short-range $d_{x^2-y^2}$-wave pairing correlations 
397: in the 2D Hubbard model, but no long-range superconducting order
398: has been observed in the parameter regime where the simulations 
399: are carried out.
400: The irreducible particle-particle
401: interaction $\Gamma_I$ has been also calculated 
402: with the QMC simulations and it gives
403: information on the microscopic many-body
404: processes causing the attraction in the 
405: $d_{x^2-y^2}$-wave pairing channel.
406: In Section 6, these QMC results will be reviewed.
407: It will be seen that $\Gamma_I$ is repulsive 
408: at ${\bf q}=(\pi,\pi)$ momentum transfers 
409: and that it is the short-range 
410: AF correlations which are responsible for this behavior.
411: Comparisons with the various 
412: diagrammatic approaches will also be carried out, 
413: which indicate that in the intermediate-coupling regime 
414: the momentum and the Matsubara-frequency structure in 
415: $\Gamma_I$ can be described by a properly-renormalized 
416: single spin-fluctuation exchange interaction. 
417: Using the QMC data on $\Gamma_I$ and the 
418: single-particle Green's functions, the Bethe-Salpeter
419: equation in the particle-particle channel has been solved. 
420: These calculations allow for a comparison of the strength of the 
421: pairing correlations in the various channels.
422: Here, it will be seen that as the temperature is lowered, 
423: the fastest growing pairing correlations occur in the 
424: singlet $d_{x^2-y^2}$-wave channel.
425: 
426: While these calculations are not carried out at sufficiently 
427: low temperatures to determine whether the
428: $d_{x^2-y^2}$-wave long-range superconducting order 
429: exists in the ground state of the doped 2D Hubbard model,
430: the results on $\Gamma_I$ are useful in the following sense.
431: Consider a phonon-mediated superconductor
432: such as Pb, where the effective interaction mediating 
433: the pairing forms already at temperatures of order the 
434: characteristic phonon frequency $\omega_D$.
435: Hence, in this case, 
436: at $T\sim \omega_D$ it would be possible to study 
437: the effective particle-particle interaction responsible 
438: for superconductivity, 
439: even though the superconducting long-range order 
440: takes place at much lower temperatures.
441: Similarly, 
442: the QMC simulations are carried out at temperatures 
443: of order or less than the characteristic 
444: energy scale $J\sim 4t^2/U$ of the AF correlations.
445: Hence, it is possible to probe $\Gamma_I$ in the 
446: temperature regime where short-range AF correlations have formed,
447: and see what type of many-body processes 
448: are important in mediating the pairing
449: and which pairing channels are favored.
450: 
451: The 2-leg Hubbard ladder is a model where the 
452: $d_{x^2-y^2}$-wave superconducting correlations can be 
453: studied in detail. 
454: In Section 7, the DMRG and the QMC results on the 
455: 2-leg Hubbard ladder will be reviewed.
456: The DMRG calculations have found enhanced power-law decaying 
457: $d_{x^2-y^2}$-wave superconducting correlations in the 
458: ground state of the 2-leg Hubbard ladder. 
459: At the same time, with the QMC simulations,
460: it is possible to calculate the magnetic susceptibility, 
461: the single-particle spectral weight and the 
462: particle-particle vertex $\Gamma_I$ at sufficiently low
463: temperatures.
464: The comparisons of the DMRG and 
465: the QMC results show that it is
466: the strong short-range AF fluctuations which mediate 
467: the $d_{x^2-y^2}$-wave pairing correlations 
468: in this system.
469: Furthermore, in this case, it is possible to 
470: study the dependence of the strength of the 
471: superconducting correlations on the model parameters
472: such as $t_{\perp}/t$ and $U/t$.
473: It will be seen that the superconducting correlations are 
474: strongest in the intermediate-coupling regime and 
475: when $t_{\perp}/t$ is such that 
476: there is large amount of single-particle 
477: spectral weight pinned near the $(\pi,0)$ and 
478: $(0,\pi)$ points of the Brillouin zone.
479: In addition, the QMC calculations find that $\Gamma_I$
480: peaks at $(\pi,\pi)$ momentum transfer.
481: These features of $A({\bf p},\omega)$ and 
482: $\Gamma_I$ create optimum conditions for 
483: $d_{x^2-y^2}$-wave superconducting correlations.
484: In Section 7.3, these numerical results on the 2D 
485: and the 2-leg Hubbard models will be compared
486: with each other.
487: This comparison suggests that the $d_{x^2-y^2}$-wave 
488: pairing correlations seen in these models do not require
489: a particularly sharp peak in 
490: $\Gamma_I$ but rather simply weight at momentum transfers
491: near $(\pi,\pi)$.
492: 
493: In Section 8, 
494: the QMC and the DMRG results reviewed above
495: will be compared with the diagrammatic 
496: and the other numerical approaches to the Hubbard model
497: as well as with the numerical results on the $t$-$J$ model.
498: First, in Section 8.1, 
499: the QMC results on the 2D Hubbard model 
500: will be compared with the results 
501: of the fluctuation exchange (FLEX) approach.
502: Here, the purpose will be to make simple 
503: estimates for the maximum $T_c$ possible 
504: in the 2D Hubbard model, 
505: if a superconducting phase were to occur.
506: In particular, 
507: the effects of the system being near an AF Mott-Hubbard
508: insulator on the strength of the $d_{x^2-y^2}$-wave 
509: pairing will be explored.
510: In Section~8.2, the results from the variational and 
511: the projector Monte Carlo studies of the Hubbard model 
512: will be reviewed briefly.
513: Recently,
514: the dynamical cluster approximation (DCA) 
515: and the one-loop renormalization-group (RG)
516: technique employing a 2D Fermi surface were applied
517: to study $d_{x^2-y^2}$ pairing 
518: in the 2D Hubbard model.
519: The results of these calculations will be discussed 
520: in Section~8.3.
521: The origin of the normal state properties 
522: of the cuprates in the pseudogap regime 
523: remains an important problem in this field.
524: In Section~8.4, 
525: the findings of various calculations
526: in the low-doping regime of the Hubbard model 
527: will be compared with the pseudogap regime of the cuprates. 
528: 
529: The Hubbard model is closely related to the $t$-$J$ model,
530: for which various Monte Carlo, exact diagonalization and
531: DMRG calculations have been carried out.
532: There is much interest in the ground-state phase diagram
533: and the nature of the density and the pairing correlations 
534: in this model.
535: In Section~8.5, the numerical studies on phase separation, 
536: stripe correlations and the $d_{x^2-y^2}$-wave pairing 
537: correlations in the $t$-$J$ model will be briefly reviewed
538: and compared with the results on the Hubbard model.
539: Comparisons will also be made with the 2-leg $t$-$J$ model.
540: 
541: In Section~8.6, the implications of 
542: the numerical results on the Hubbard model for the 
543: nature of the 
544: $d_{x^2-y^2}$-wave pairing in the high-$T_c$ 
545: cuprates will be discussed.
546: In Section 9, 
547: the summary and the conclusions will be given.
548: 
549: \setcounter{equation}{0}\setcounter{figure}{0}
550: \section{{\boldmath$d_{x^2-y^2}$}-wave superconductivity 
551: in the cuprates}
552: 
553: \subsection{Spin-fluctuation mediated $d_{x^2-y^2}$-wave 
554: superconductivity} 
555: 
556: Since the development of the BCS theory,
557: it has been of interest to see whether the effective interaction 
558: which is responsible for pairing could be mediated 
559: by excitations other than phonons.
560: The superfluidity of $^3$He is an example where 
561: it is believed that the pairing is due to the exchange 
562: of ferromagnetic spin fluctuations resulting in triplet 
563: $p$-wave superconductivity [Leggett 1975].
564: Within the context of organic superconductors, 
565: the possibility of the pairing being mediated by spin fluctuations 
566: had been noted in Ref.~[Emery 1986].
567: For the heavy fermion materials, 
568: the possibility of $d$-wave superconductivity due to the exchange 
569: of AF spin fluctuations was proposed in 
570: Refs.~[Scalapino {\it et al.} 1986, 
571: Miyake {\it et al.} 1986, Cyrot 1986].
572: For the high-$T_c$ superconductors,
573: the possibility of $d_{x^2-y^2}$-wave superconductivity 
574: was first studied in
575: Ref.~[Bickers {\it et al.} 1987], 
576: where the framework of the two-dimensional Hubbard model 
577: was used near an SDW phase.
578: 
579: Here, a discussion of these ideas on how the exchange of the 
580: AF spin-fluctuations leads to a $d_{x^2-y^2}$-wave gap 
581: function will be given, 
582: since this picture will be useful for the 
583: remainder of the article.
584: In the paramagnetic phase
585: of the 2D Hubbard model and within RPA,
586: the single spin-fluctuation exchange interaction
587: between opposite-spin particles at zero frequency transfer is  
588: given by 
589: [Berk and Schrieffer 1966, Doniach and Engelsberg 1966]
590: \begin{equation} 
591: V({\bf p'}|{\bf p}) = U + 
592: { U^3\chi_0^2({\bf p'}-{\bf p}) 
593: \over 1-U^2\chi_0^2({\bf p'}-{\bf p}) } +
594: { U^2\chi_0({\bf p'}+{\bf p}) 
595: \over 1-U \chi_0({\bf p'}+{\bf p}) },
596: \end{equation}
597: where $\chi_0({\bf q})$ is the usual Lindhard function.
598: The longitudinal and the transverse spin fluctuations 
599: contributing to $V$ are illustrated in Fig.~2.1.
600: Here, $V({\bf p'}|{\bf p})$ is the interaction for the scattering of 
601: a pair of opposite-spin electrons at states $({\bf p},-{\bf p})$ 
602: to $({\bf p'},-{\bf p'})$.
603: Near half-filling and for a tight-binding band structure
604: $\varepsilon_{\bf p} = -2t (\cos{p_x} + \cos{p_y} ) - \mu $, 
605: where $\mu$ is the chemical potential, 
606: the Lindhard function peaks 
607: at ${\bf q}\sim (\pi,\pi)$.
608: Consequently,
609: for a system with Stoner-enhanced AF correlations,
610: the spin-fluctuation exchange interaction 
611: $ V({\bf p'}|{\bf p})$ 
612: is large and repulsive at 
613: ${\bf p'}-{\bf p} \sim (\pi,\pi)$ momentum transfers.
614: Now, consider the BCS gap equation
615: \begin{equation}
616: \Delta_{\bf p} = - \sum_{\bf p'} 
617: { V({\bf p}|{\bf p'})  \Delta_{\bf p'} \over 
618: 2E_{\bf p'} },
619: \end{equation}
620: where ${\bf p}$ and ${\bf p'}$ are restricted to being 
621: on the Fermi surface of the doped system 
622: as illustrated in Fig.~2.2.
623: When this equation is solved,
624: it is found that 
625: the leading superconducting instability occurs for
626: a gap function with the $d_{x^2-y^2}$-wave symmetry,
627: \begin{equation}
628: \Delta_{\bf p} = {\Delta_0 \over 2}
629: ( \cos{p_x} - \cos{p_y} ).
630: \end{equation}
631: In order to see why 
632: the $d_{x^2-y^2}$-wave form is a solution, consider 
633: a quasiparticle pair occupying $({\bf p},-{\bf p})$
634: with ${\bf p}$ near $(\pi,0)$,
635: as shown in Fig.~2.2.
636: This pair can scatter to 
637: $({\bf p'},-{\bf p'})$ where ${\bf p'}\sim(0,\pi)$
638: through the interaction $V({\bf p'}|{\bf p})$, 
639: which is enhanced and positive 
640: for momentum transfers near $(\pi,\pi)$,
641: because $\Delta_{\bf p}$ changes sign between 
642: ${\bf p}$ and ${\bf p'}$.
643: This is basically the reason for why the $d_{x^2-y^2}$-wave gap 
644: symmetry is favored by the AF spin fluctuations. 
645: Since $V({\bf p'}|{\bf p})$ is always positive, 
646: there is no singlet solution with the 
647: usual $s$-wave symmetry.
648: For simplicity, 
649: these arguments were given at zero frequency,
650: where ${\bf p}$ and ${\bf p'}$ are restricted to being on the 
651: Fermi surface,
652: but the same arguments 
653: hold when the frequency dependencies of the 
654: gap equation and of the effective interaction are 
655: taken into account.
656: In Section~6, the QMC data on the irreducible particle-particle 
657: interaction $\Gamma_I$ 
658: and the solutions of the Bethe-Salpeter equation
659: will be reviewed for the 2D Hubbard model.
660: There, 
661: it will be seen that $\Gamma_I$ is similar to the 
662: single spin-fluctuation exchange interaction 
663: and 
664: the leading singlet pairing instability occurs
665: in the $d_{x^2-y^2}$-wave channel in the parameter regime where
666: the QMC simulations are carried out.
667: 
668: \begin{figure}
669: \centering
670: \iffigure
671: \epsfig{file=ch2-fig1.ps,height=5.5cm,angle=0}
672: \fi
673: \caption{
674: Feynman diagrams illustrating the effective 
675: particle-particle interaction $V({\bf p'}|{\bf p})$ 
676: within the single spin-fluctuation exchange approximation. 
677: }
678: \label{2.1}
679: \end{figure}
680: 
681: \begin{figure}
682: \centering
683: \iffigure
684: \epsfysize=8cm
685: \epsffile[100 150 550 610]{ch2-fig2.ps}
686: \fi
687: \caption{
688: Sketch of the Fermi surface of the 2D tight-binding 
689: model doped near half-filling.
690: The two-particle momentum states $({\bf p},-{\bf p})$
691: and $({\bf p'},-{\bf p'})$ are indicated by the filled circles.
692: }
693: \label{2.2}
694: \end{figure}
695: 
696: \subsection{Experimental evidence for $d_{x^2-y^2}$-wave superconductivity
697: in the cuprates} 
698: 
699: Early experimental evidence for 
700: $d_{x^2-y^2}$-wave pairing in the cuprates arose from the 
701: measurement of the NMR longitudinal nuclear relaxation 
702: rate $T_1^{-1}$ for Cu(2) and O(2,3) nuclei in 
703: YBa$_2$Cu$_3$O$_7$.
704: In particular,
705: the anisotropy of $T_1^{-1}$ for Cu(2) 
706: below $T_c$ provided a signature for $d_{x^2-y^2}$-wave pairing
707: in this compound 
708: [Takigawa {\it et al.} 1991, 
709: Martindale {\it et al.} 1992, Bulut and Scalapino 1992].
710: These NMR measurements provided evidence for both the phase 
711: and the nodes of the $d_{x^2-y^2}$-wave gap function.
712: The temperature dependence of the Cu(2) transverse 
713: nuclear relaxation rate $T_2^{-1}$ was also used for the 
714: identification of the gap symmetry
715: [Bulut and Scalapino 1991, Itoh {\it et al.} 1992].
716: These NMR measurements were also supported by the 
717: ARPES experiments which extracted the 
718: magnitude of the gap function
719: $|\Delta_{\bf p}|$ on the Fermi surface
720: [Shen {\it et al.} 1993].
721: The microwave cavity measurements of the superconducting 
722: penetration depth $\lambda(T)$ in single crystals of 
723: YBa$_2$Cu$_3$O$_7$ found a linear temperature variation which also gave 
724: support to a $d_{x^2-y^2}$-wave gap [Hardy {\it et al.} 1993].
725: The phase-coherence measurements on 
726: YBCO-Pb dc SQUIDS [Wollman {\it et al.} 1993]
727: and the tricrystal experiments of Tsuei {\it et al.}
728: [Tsuei {\it et al.} 1994] established that in 
729: YBa$_2$Cu$_3$O$_7$ the order parameter has the 
730: $d_{x^2-y^2}$-wave symmetry.
731: Reviews of these results on the symmetry of the gap function 
732: are given in Refs.~[Schrieffer 1994, Scalapino 1995, 
733: van Harlingen 1995, Tsuei and Kirtley 2000].
734: Today, it is established that in a number of the hole 
735: and the electron doped cuprates the gap function has the 
736: $d_{x^2-y^2}$-wave symmetry [Tsuei and Kirtley 2000].
737: 
738: \subsection{Theoretical studies of $d_{x^2-y^2}$-wave pairing
739: in the cuprates} 
740: 
741: For the high-$T_c$ cuprates, 
742: the possibility of $d_{x^2-y^2}$-wave 
743: superconductivity was first studied by Bickers {\it et al.} 
744: [Bickers {\it et al.} 1987]
745: using the framework of the 2D Hubbard model.
746: In particular, the single spin-fluctuation exchange 
747: interaction was used for studying the $d_{x^2-y^2}$-wave pairing 
748: near an SDW instability within the random-phase approximation.
749: In these calculations, the effect of the single-particle self-energy 
750: corrections due to the spin fluctuations was taken into account.
751: 
752: The next level of calculations for 
753: the $d_{x^2-y^2}$-wave superconductivity 
754: were carried out using the fluctuation-exchange approximation (FLEX)
755: within the 2D Hubbard model 
756: [Bickers {\it et al.} 1989].
757: This method self-consistently treats the fluctuations in the 
758: magnetic, density and the pairing channels.
759: Within this approach it was found that the AF spin fluctuations
760: lead to a $d_{x^2-y^2}$-wave superconducting phase which
761: neighbors the SDW phase.
762: In these calculations, superconducting transition temperatures 
763: as high as $0.025t$ were found 
764: [Bickers {\it et al.} 1989, Bickers and White 1991].
765: Since the hopping matrix element within a one-band description 
766: of the cuprates is estimated to be about 0.45~eV
767: [Hybertson {\it et al.} 1990], this value of $T_c$ corresponds to 
768: about 130~K.
769: However, the need to carry out exact calculations 
770: was also noted.
771: 
772: Another approach to $d_{x^2-y^2}$-wave pairing in the cuprates 
773: was from a phenomenological point of view
774: [Moriya {\it et al.} 1990].
775: In this approach, a spin-fluctuation exchange 
776: interaction was used to calculate the 
777: superconducting transition temperatures. 
778: The spin susceptibility and the parameters 
779: used in the model were obtained from
780: fitting the NMR and the electrical resistivity 
781: data on the normal state of the cuprates
782: within the self-consistent renormalization theory.
783: In these calculations, 
784: the single-particle self-energy corrections due to the 
785: spin fluctuations were also included, and $T_c$'s as high as 
786: those in the cuprates were obtained.
787: A review of these calculations can be found in 
788: Ref.~[Moriya and Ueda 2000].
789: 
790: A similar phenomenological approach was taken 
791: in Ref.~[Monthoux {\it et al.} 1991].
792: In this approach, the spin susceptibility used in the 
793: effective interaction was also taken from fitting the NMR data, 
794: and $T_c$'s comparable to those in the cuprates were obtained. 
795: In these calculations, the $T_c$ was initially calculated without 
796: including the self-energy effects. 
797: Later, the self-energy corrections were included self-consistently
798: [Monthoux and Pines 1992].
799: 
800: These spin-fluctuation theories for superconductivity differ from 
801: phonon-mediated superconductivity in a fundamental way.
802: Here, the effective attractive interaction which 
803: is responsible for pairing the electrons is generated 
804: by the electronic correlations rather than by an external system
805: such as the lattice vibrations. 
806: Hence, it is an important question whether in a microscopic model
807: the electronic correlations can indeed produce 
808: an effective attractive interaction which leads to 
809: superconductivity, and, if so, what would be 
810: the nature and the strength 
811: of such pairing correlations.
812: The QMC and the DMRG methods were developed to address these
813: types of questions.
814: The numerical studies of the Hubbard model carried out 
815: with this purpose are the subject of this review article.
816: 
817: \setcounter{equation}{0}\setcounter{figure}{0}
818: \section{Magnetic correlations in the 2D Hubbard model}
819: 
820: The parent compounds of the cuprates are 
821: AF insulators, 
822: and the AF phase extends up to $\sim $5\% doping. 
823: For dopings beyond this, there is clear evidence from the 
824: NMR and the inelastic neutron scattering experiments that the 
825: system exhibits short-range AF spin fluctuations
826: [Pennington and Slichter 1990, Birgeneau 1990].
827: These are properties which support 
828: using the Hubbard framework for studying the 
829: electronic correlations of the cuprates.
830: 
831: The QMC calculations have found that 
832: the ground state of the Hubbard model at half-filling has 
833: AF long-range order 
834: [Hirsch and Tang 1989, White {\it et al.} 1989b].
835: Away from half-filling, the QMC calculations 
836: find short range AF correlations [Hirsch 1985].
837: In this section, 
838: the QMC results on the magnetic susceptibility 
839: $\chi$ of the 2D Hubbard model will be reviewed, and 
840: compared with an RPA-like approach for modelling it.
841: In addition, the QMC results on the irreducible 
842: particle-hole vertex will be discussed.
843: These will be useful for gaining a microscopic
844: understanding of the magnetic correlations in the Hubbard model.
845: 
846: \begin{figure}[ht]
847: \centering
848: \iffigure
849: \mbox{
850: \subfigure{
851: \epsfysize=8cm
852: \epsffile[100 150 480 610]{ch3-fig1a.ps}}
853: \quad
854: \subfigure{
855: \epsfysize=8cm
856: \epsffile[50 150 600 610]{ch3-fig1b.ps}}}
857: \fi
858: \caption{
859: (a) Static magnetic susceptibility $\chi({\bf q},0)$ versus
860: ${\bf q}$.
861: (b) Equal-time correlation function 
862: $|\langle m^z({\bf r})m^z(0)\rangle |$
863: versus ${\bf r}$ along $(1,0)$.
864: These results are for $U=8t$ 
865: and $\langle n\rangle=1.0$ on an $8\times 8$ lattice.
866: }
867: \label{3.1}
868: \end{figure}
869: 
870: \subsection{Antiferromagnetic long-range order at half-filling} 
871: 
872: The transverse magnetic susceptibility is defined by
873: \begin{equation}
874: \label{chiqw}
875: \chi({\bf q},\omega) = 
876: {1\over N} \sum_{\ell} \int_0^{\beta} \, d\tau \,
877: e^{i\omega_m\tau} \, e^{-i{\bf q}\cdot \ell} \,
878: \langle m^-(i+\ell,0) m^+(i,0) \rangle .
879: \end{equation}
880: Here, 
881: $m^+(i,0) = c^{\dagger}_{i\uparrow} c_{i\downarrow}$,
882: and $m^-(i,\tau) = e^{H\tau} m^-(i,0) e^{-H\tau}$,
883: where $m^-(i,0)$ is the hermitian conjugate of 
884: $m^+(i,0)$.
885: Ref.~[Mahan 1981] can be consulted for the finite-temperature 
886: Green's function formalism.
887: In the following, $\chi$ will be plotted in units of $t^{-1}$.
888: Figure 3.1(a) shows 
889: $\chi({\bf q},i\omega_m=0)$ versus ${\bf q}$ 
890: along various cuts in the Brillouin zone
891: as the temperature is lowered.
892: Here,
893: it is seen that as $T$ is lowered, 
894: $\chi({\bf q},0)$ develops a sharp peak at the antiferromagnetic 
895: wave vector ${\bf q}=(\pi,\pi)$.
896: 
897: Figure 3.1(b) shows the equal-time correlation function 
898: \begin{equation}
899: \label{mz}
900: |\langle m^z({\bf r}) m^z(0) \rangle |,
901: \end{equation}
902: where 
903: $m^z({\bf r}) = 
904: c^{\dagger}_{{\bf r}\uparrow} c_{{\bf r}\uparrow}
905: - c^{\dagger}_{{\bf r}\downarrow} c_{{\bf r}\downarrow}$,
906: versus ${\bf r}$ as the temperature is lowered on an $8\times 8$
907: lattice at half-filling. 
908: In this figure, it is seen that as $T$ is lowered below $0.5t$,
909: the AF correlation length $\xi_{AF}$ reaches 
910: the size of the system and long-range AF order is established 
911: on the $8\times 8$ lattice. 
912: For an infinite lattice, the long-range AF order 
913: would take place at $T=0$.
914: 
915: \begin{figure}[ht]
916: \centering
917: \iffigure
918: \mbox{
919: \subfigure{
920: \epsfysize=8cm
921: \epsffile[100 150 480 610]{ch3-fig2a.ps}}
922: \quad
923: \subfigure{
924: \epsfysize=8cm
925: \epsffile[50 150 600 610]{ch3-fig2b.ps}}}
926: \fi
927: \caption{
928: (a) Static magnetic susceptibility $\chi({\bf q},0)$ versus
929: ${\bf q}$.
930: (b) Equal-time correlation function 
931: $|\langle m^z({\bf r})m^z(0)\rangle |$
932: versus ${\bf r}$ along $(1,0)$.
933: These results are for $U=8t$ and  
934: $\langle n\rangle=0.87$ on an $8\times 8$ lattice.
935: }
936: \label{3.2}
937: \end{figure}
938: 
939: \subsection{Short-range antiferromagnetic fluctuations 
940: away from half-filling}
941: 
942: Figure 3.2(a) shows $\chi({\bf q},0)$ 
943: versus ${\bf q}$ for $\langle n\rangle =0.87$ 
944: and $U=8t$ on the $8\times 8$ lattice.
945: Here, as $T$ decreases, 
946: the development of a broad peak centred at 
947: ${\bf q}=(\pi,\pi)$ is seen. 
948: Figure 3.2(b) shows the corresponding 
949: $|\langle m^z({\bf r}) m^z(0) \rangle |$ 
950: versus ${\bf r}$. 
951: At the lowest temperature of 0.33t, 
952: the AF correlation length $\xi$ is about 0.6 
953: lattice spacing for this filling. 
954: Here, $\xi$ was obtained by fitting the decay 
955: of $|\langle m^z({\bf r})m^z(0)\rangle |$ 
956: to $e^{-|{\bf r}|/\xi}$.
957: 
958: The Monte Carlo results on $\chi({\bf q},i\omega_m)$ 
959: have been analytically continued to the real frequency axis 
960: using the Pade approximation.
961: The results on the spin-fluctuation spectral weight 
962: ${\rm Im}\,\chi({\bf q},\omega)$ versus $\omega$
963: at ${\bf q}=(\pi,\pi)$ obtained in 
964: this way are shown in Fig.~3.3(a). 
965: While the Pade approximation can only resolve the crude features of the 
966: spectrum, the softening of the AF spin fluctuations at low temperatures
967: is clearly seen. 
968: 
969: \begin{figure}[ht]
970: \centering
971: \iffigure
972: \mbox{
973: \subfigure{
974: \epsfysize=8cm
975: \epsffile[100 150 480 610]{ch3-fig3a.ps}}
976: \quad
977: \subfigure{
978: \epsfysize=8cm
979: \epsffile[50 150 600 610]{ch3-fig3b.ps}}}
980: \fi
981: \caption{
982: (a) Spin-fluctuation spectral weight 
983: ${\rm Im}\,\chi({\bf q},\omega)$
984: versus $\omega$ at ${\bf q}=(\pi,\pi)$.
985: (b) $F({\bf q})$ versus ${\bf q}$
986: for the same temperatures as in (a).
987: These results are for $U=8t$ and  
988: $\langle n\rangle=0.87$ on an $8\times 8$ lattice.
989: }
990: \label{3.3}
991: \end{figure}
992: 
993: Another quantity which can be used for characterising 
994: the low-frequency magnetic correlations is 
995: \begin{equation}
996: \label{Fq}
997: F({\bf q}) = 
998: { {\rm Im}\,\chi({\bf q},\omega) \over \omega}
999: \bigg|_{\omega\rightarrow 0}.
1000: \end{equation}
1001: This quantity is of interest because it is used to calculate 
1002: the longitudinal nuclear relaxation rates of the 
1003: planar Cu and O nuclear spins 
1004: in the layered cuprates.
1005: Figure 3.3(b) shows $F({\bf q})$ versus ${\bf q}$ for 
1006: $\langle n\rangle =0.87$ and $U=8t$.
1007: As $T$ is lowered, 
1008: $F({\bf q})$ gets enhanced at $(\pi,\pi)$, and more weakly at 
1009: ${\bf q}\sim 0$.
1010: 
1011: \subsection{RPA-like model for the magnetic susceptibility}
1012: 
1013: An RPA-like approach has been used for modelling
1014: $\chi({\bf q},\omega)$ 
1015: in the intermediate coupling regime with the form
1016: [Bulut {\it et al.} 1990, Bulut 1990, Chen {\it et al.} 1991, 
1017: Bulut {\it et al.} 1993]
1018: \begin{equation}
1019: \label{chirpa}
1020: \chi({\bf q},\omega) = 
1021: { \chi_0({\bf q},\omega) \over 
1022: 1 - \overline{U}\chi_0({\bf q},\omega) }
1023: \end{equation}
1024: where $\chi_0({\bf q},\omega)$ is the Lindhard function for the 
1025: band electrons 
1026: \begin{equation}
1027: \label{lindhard}
1028: \chi_0({\bf q},\omega) = 
1029: {1 \over N} \sum_{\bf p}
1030: { 
1031: f(\varepsilon_{{\bf p}+{\bf q}}) - f(\varepsilon_{\bf p}) 
1032: \over 
1033: \omega - ( \varepsilon_{{\bf p}+{\bf q}} - \varepsilon_{\bf p} ) + i\delta 
1034: },
1035: \end{equation}
1036: $\varepsilon_{\bf p}=-2t(\cos{p_x} + \cos{p_y})-\mu$
1037: and $f(\varepsilon_{\bf p})$ is the usual Fermi factor.
1038: The parameter $\overline{U}$ entering Eq.~(\ref{chirpa})
1039: represents the reduced Coulomb repulsion, which is assumed to 
1040: be renormalized due to the electronic correlations.
1041: In this simple approach, $\overline{U}$ is taken to be 
1042: independent of momentum and temperature.
1043: This turns out to be a reasonable approximation in the 
1044: temperature range where the Monte Carlo simulations 
1045: have been carried out.
1046: 
1047: Figure 3.4(a) compares this RPA-like approximation for 
1048: $\chi({\bf q},0)$ with the Monte Carlo data for 
1049: $U=4t$, $T=0.25t$ and $\langle n\rangle =0.87$. 
1050: A similar comparison is given in Fig.~3.4(b) for the 
1051: $T$ dependence of $\chi({\bf q}=(\pi,\pi),0)$.
1052: In these figures $\overline{U}=2t$ has been used.
1053: This choice will be discussed further in the following subsection,
1054: where results on the irreducible particle-hole vertex 
1055: will be reviewed.
1056: 
1057: \begin{figure}[ht]
1058: \centering
1059: \iffigure
1060: \mbox{
1061: \subfigure{
1062: \epsfysize=8cm
1063: \epsffile[100 150 480 610]{ch3-fig4a.ps}}
1064: \quad
1065: \subfigure{
1066: \epsfysize=8cm
1067: \epsffile[50 150 600 610]{ch3-fig4b.ps}}}
1068: \fi
1069: \caption{
1070: (a) Comparison of the QMC and the RPA 
1071: results on $\chi({\bf q},0)$ versus ${\bf q}$ at $T=0.25t$.
1072: (b) Comparison of the QMC and the RPA 
1073: results on $\chi({\bf Q}=(\pi,\pi),0)$
1074: versus $T/t$.
1075: Here, the QMC data (filled circles) are for $U=4t$ and 
1076: the RPA results (solid curves) were obtained 
1077: using $\overline{U}=2t$.
1078: In both figures, 
1079: the results are shown for $\langle n\rangle=0.87$.
1080: }
1081: \label{3.4}
1082: \end{figure}
1083: 
1084: \subsection{Irreducible particle-hole interaction}
1085: 
1086: In the RPA form discussed above, 
1087: a reduced Coulomb repulsion $\overline{U}$ which is 
1088: independent of ${\bf q}$ is used. 
1089: In order to test this approximation, 
1090: the effective particle-hole vertex defined by 
1091: \begin{equation}
1092: \overline{U}({\bf q},i\omega_m) = 
1093: {1 \over \overline{\chi}({\bf q},i\omega_m)} - 
1094: {1 \over \chi({\bf q},i\omega_m)},
1095: \end{equation}
1096: was calculated in Ref.~[Bulut {\it et al.} 1995].
1097: Here, 
1098: $\overline{\chi}({\bf q},i\omega_m)$ 
1099: is the piece of the magnetic susceptibility 
1100: which does not include the reducible particle-hole vertex,
1101: and it is defined by 
1102: \begin{equation}
1103: \overline{\chi} ({\bf q},i\omega_m) = 
1104: - {T\over N} \sum_{{\bf p},i\omega_m} 
1105: G({\bf p}+{\bf q},i\omega_n + i\omega_m)
1106: G({\bf p},i\omega_n),
1107: \end{equation}
1108: where $G({\bf p},i\omega_n)$ is the exact single-particle 
1109: Green's function calculated with QMC.
1110: In Fig.~3.5(a), 
1111: the QMC data on $\overline{\chi}({\bf q},0)$ versus 
1112: ${\bf q}$ is compared with the Lindhard function
1113: $\chi_0({\bf q},0)$ at $T=0.25t$.
1114: Here, 
1115: one sees that $\overline{\chi}({\bf q},0)$ 
1116: is suppressed with respect to $\chi_0({\bf q},0)$,
1117: and this is 
1118: because of the single-particle self-energy effects.
1119: The effective irreducible vertex 
1120: $\overline{U}({\bf q},0)$ obtained 
1121: from the QMC data on $\chi({\bf q},0)$ and 
1122: $\overline{\chi}({\bf q},0)$ is plotted in Fig.~3.5(b) 
1123: as a function of ${\bf q}$ at various temperatures.
1124: In this figure it is seen that 
1125: $\overline{U}({\bf q},0)$ shows little dependence on 
1126: ${\bf q}$ and $T$.
1127: This is the reason for the good agreement between the QMC data 
1128: and the RPA results obtained using a reduced $\overline{U}$
1129: in Eq.~(3.4).
1130: However, 
1131: the values of $\overline{U}({\bf q},0)$ are larger than 
1132: $\overline{U}=2t$ used in the RPA.
1133: For instance, at $T=0.25t$ one has 
1134: $\overline{U}({\bf q}=(\pi,\pi),0)=3.2t$.
1135: This is because 
1136: in the RPA form the Lindhard susceptibility $\chi_0$ is used rather 
1137: than $\overline{\chi}$, which has the effects 
1138: of the single-particle self-energy corrections. 
1139: At this point, it should be noted that 
1140: the momentum dependence of the effective particle-hole
1141: vertex within the context of the cuprates was also
1142: discussed in Ref.~[Anderson 1997].
1143: 
1144: \begin{figure}
1145: \centering
1146: \iffigure
1147: \mbox{
1148: \subfigure{
1149: \epsfysize=8cm
1150: \epsffile[100 150 480 610]{ch3-fig5a.ps}}
1151: \quad
1152: \subfigure{
1153: \epsfysize=8cm
1154: \epsffile[50 150 600 610]{ch3-fig5b.ps}}}
1155: \fi
1156: \caption{
1157: (a) Comparison of the QMC data on $\overline{\chi}({\bf q},0)$
1158: versus ${\bf q}$ with the Lindhard function $\chi_0({\bf q},0)$ 
1159: at $T=0.25t$.
1160: (b) Effective particle-hole vertex
1161: $\overline{U}({\bf q},0)$ versus ${\bf q}$ along $(1,1)$.
1162: These results were obtained for 
1163: $U=4t$ and $\langle n\rangle =0.87$ on an 
1164: $8\times 8$ lattice.
1165: }
1166: \label{3.5}
1167: \end{figure}
1168: 
1169: With the QMC simulations the irreducible 
1170: particle-hole vertex $\overline{\Gamma}_I$ was also calculated
1171: [Bulut {\it et al.} 1995].
1172: The calculation of $\overline{\Gamma}_I$ is similar to 
1173: that of the irreducible particle-particle vertex $\Gamma_I$,
1174: which will be discussed in Section~6.
1175: In this approach, $\overline{\Gamma}_I$ is 
1176: obtained from the solution of the particle-hole $t$-matrix
1177: equation illustrated in Fig.~3.6.
1178: These calculations showed 
1179: that the QMC data on $\overline{\Gamma}_I$
1180: are in agreement with the results on the 
1181: effective particle-hole vertex 
1182: $\overline{U}({\bf q},0)$ discussed in this section.
1183: In particular, $\overline{\Gamma}_I$ with the center-of-mass
1184: momentum ${\bf q}=(\pi,\pi)$ and frequency $\omega_m=0$
1185: was averaged over the incoming and outgoing 
1186: momenta near the Fermi surface, 
1187: and it was found that this average 
1188: $\langle \overline{\Gamma}_I\rangle$ is $3.6t \pm 0.7t$
1189: for $U=4t$, $T=0.25t$ and $\langle n\rangle=0.87$.
1190: This agrees with 
1191: $\overline{U}({\bf q}=(\pi,\pi),0)=3.2t$
1192: found above.
1193: Hence, the effective particle-hole vertex for 
1194: center-of-mass momentum ${\bf q}=(\pi,\pi)$ and
1195: frequency $\omega_m=0$ is about $3.2t$ for a bare Coulomb
1196: repulsion of $U=4t$.
1197: The difference reflects the renormalization of the bare
1198: Coulomb repulsion in the particle-hole channel due to the 
1199: higher-order many-body processes such as the Kanamori type 
1200: of particle-particle scatterings
1201: [Kanamori 1963].
1202: However, in the RPA form of Eq.~(3.4), 
1203: $\overline{U}=2t$ needs to be used in order to take 
1204: into account the single-particle self-energy corrections 
1205: which are neglected within RPA.
1206: 
1207: These results on the effective particle-hole vertex will 
1208: also be useful when the irreducible particle-particle interaction
1209: $\Gamma_I$ is compared with the single spin-fluctuation
1210: exchange interaction $\Gamma^{SF}_{Is}$
1211: in Section~6.4.
1212: There,
1213: it will be seen that when the effective coupling $gU$ 
1214: between the electrons and the spin fluctuations is taken
1215: to be about $3.2t$, the resulting $\Gamma^{SF}_{Is}$
1216: agrees with the QMC data on $\Gamma_I$. 
1217: One expects $gU$ to be closely related to 
1218: $\overline{\Gamma}_I$ and, hence, 
1219: the QMC data on the effective particle-particle 
1220: and particle-hole vertices are consistent.
1221: 
1222: \begin{figure}
1223: \centering
1224: \iffigure
1225: \epsfig{file=ch3-fig6.ps,height=8cm,angle=90}
1226: \fi
1227: \caption{
1228: Feynman diagrams for the particle-hole $t$-matrix equation.
1229: }
1230: \label{3.6}
1231: \end{figure}
1232: 
1233: \subsection{Comparison with the NMR experiments on the cuprates}
1234: 
1235: The measurements of the longitudinal NMR rate 
1236: $T_1^{-1}$ have provided evidence for the existence of 
1237: short-range AF fluctuations in the normal state of the 
1238: high-$T_c$ cuprates.
1239: These experiments are reviewed in 
1240: Refs.~[Pennington and Slichter 1990, Takigawa 1990].
1241: The relaxation rate $T_1^{-1}$ is given by 
1242: \begin{equation}
1243: \label{T1}
1244: T_1^{-1} = {T\over N} \sum_{\bf q} 
1245: |A({\bf q})|^2 F({\bf q}),
1246: \end{equation}
1247: where $|A({\bf q})|^2$ is the hyperfine form factor of the 
1248: particular nuclear spin, and $F({\bf q})$ has been defined in 
1249: Section~3.2.
1250: The hyperfine form factors 
1251: used in calculating the various NMR rates for the cuprates 
1252: were derived in Ref.~[Mila and Rice 1989].
1253: 
1254: An additional experimental quantity which supports the existence 
1255: of the short-range AF correlations in the cuprates is the 
1256: transverse relaxation rate $T_2^{-1}$, which is given by 
1257: [Pennington and Slichter 1991]
1258: \begin{equation}
1259: \label{T2}
1260: \bigg( 
1261: {1\over T_2} \bigg)^2 =
1262: {1\over N} \sum_{\bf q} 
1263: |A({\bf q})|^4 \chi^2({\bf q},0)
1264: - \bigg( {1\over N}\sum_{\bf q} 
1265: |A({\bf q})|^2 \chi({\bf q},0) \bigg)^2.
1266: \end{equation}
1267: The measurements of $T^{-1}_2$ for $^{63}$Cu(2)
1268: in the normal state of the YBa$_2$Cu$_3$O$_7$,
1269: YBa$_2$Cu$_4$O$_8$ and the 
1270: La$_{2-x}$Sr$_x$CuO$_4$ systems have 
1271: provided evidence for the presence of the short-range 
1272: AF correlations in these compounds.
1273: [Pennington and Slichter 1991, Itoh {\it et al.} 1992, 
1274: Imai {\it et al.} 1993].
1275: 
1276: Since the Monte Carlo calculations cannot be carried out at 
1277: low temperatures, it is difficult to make direct comparisons with 
1278: the $T_1^{-1}$ and the $T_2^{-1}$ measurements on the cuprates using 
1279: the Monte Carlo data.
1280: However, the RPA-like model described above in Section~3.3 
1281: has been used to calculate the NMR rates in the normal state.
1282: Even though it is a simple approximation,
1283: this approach has been useful for analysing the 
1284: $T_1^{-1}$ and the $T_2^{-1}$ data on 
1285: YBa$_2$Cu$_3$O$_7$
1286: [Bulut {\it et al.} 1990, 
1287: Bulut and Scalapino 1991].
1288: These calculations imply that the low-energy AF spin fluctuations 
1289: estimated for the 2D Hubbard model have the right order of magnitude
1290: for describing the general features of the AF correlations in 
1291: optimally doped 
1292: YBa$_2$Cu$_3$O$_7$.
1293: The NMR data on YBa$_2$Cu$_3$O$_7$ were also fit 
1294: phenomenologically
1295: [Millis {\it et al.} 1990]
1296: and by using self-consistent renormalization theory
1297: [Moriya {\it et al.} 1990].
1298: 
1299: \setcounter{equation}{0}\setcounter{figure}{0}
1300: \section{Charge fluctuations in the 2D Hubbard model}
1301: 
1302: In this section, the QMC data on the charge susceptibility 
1303: and the optical conductivity of the 2D Hubbard model will be reviewed. 
1304: 
1305: \subsection{Charge susceptibility}
1306: 
1307: The momentum and frequency dependent charge susceptibility is defined by 
1308: \begin{equation}
1309: \label{Pi}
1310: \Pi({\bf q},i\omega_m) = \int_0^{\beta} \, d\tau \,
1311: e^{i\omega_m\tau} \, 
1312: S({\bf q},\tau)
1313: \end{equation}
1314: where
1315: \begin{equation}
1316: S({\bf q},\tau) = -
1317: \langle T \rho({\bf q},\tau) \rho^{\dagger} ({\bf q},0) \rangle ,
1318: \end{equation}
1319: with 
1320: $\rho^{\dagger}({\bf q})= 
1321: {1\over \sqrt{N}}  \sum_{{\bf p}\sigma} 
1322: c^{\dagger}_{{\bf p}+{\bf q}\sigma} c_{{\bf p}\sigma}$
1323: and $\rho({\bf q},\tau)= e^{H\tau} \rho({\bf q}) e^{-H\tau}$.
1324: The equal-time density-density correlation function 
1325: $S({\bf q})\equiv S({\bf q},\tau=0)$
1326: and the charge compressibility 
1327: $\kappa = \partial \langle n\rangle / \partial \mu$
1328: given by $\Pi({\bf q}\rightarrow 0,0)$
1329: are also of interest.
1330: In the following $\Pi$ and $S$ 
1331: will be plotted in units of $t^{-1}$.
1332: The nature of the density correlations in the 2D Hubbard model was
1333: studied with various numerical techniques.
1334: The QMC 
1335: [Moreo and Scalapino 1991, Moreo {\it et al.} 1991] 
1336: and the exact diagonalization [Dagotto {\it et al.} 1992b] 
1337: calculations did not find 
1338: any indication of phase separation in the 2D Hubbard model.
1339: The charge compressibility of the Hubbard model 
1340: near half-filling was calculated 
1341: by zero-temperature QMC simulations 
1342: [Furukawa and Imada 1991 and 1992].
1343: At fillings $\langle n\rangle=0.5$ and 0.2, 
1344: $S({\bf q})$ was calculated with the determinantal QMC for 
1345: $U=4t$ and $8t$ on up to $16\times 16$ lattices
1346: [Chen {\it et al.} 1994].
1347: The Green's function Monte Carlo (GFMC) method was also 
1348: used for calculating $S({\bf q})$ for the 
1349: 2D Hubbard model on lattices with up to 162 sites in the ground state
1350: [Becca {\it et al.} 2000].
1351: These calculations were carried out for $U=4t$ and $10t$
1352: at and near half-filling.
1353: In this section, 
1354: the QMC results on $\Pi({\bf q},\omega)$ and $S({\bf q})$
1355: from Ref.~[Bulut 1996] will be shown
1356: for various values of $U/t$ near half-filling.
1357: 
1358: Extensive numerical calculations have been 
1359: also carried out for studying 
1360: the density correlations in the ground state
1361: of the $t$-$J$ model.
1362: These calculations have found phase-separated, striped, and uniform
1363: phases, and they will be briefly described in Section~8.
1364: The results on $\Pi({\bf q},\omega)$ 
1365: and $S({\bf q})$ presented here were obtained at relatively high
1366: temperatures and it is difficult to reach conclusions about the 
1367: ground state of the Hubbard model using them.
1368: Nevertheless, here, the QMC data on 
1369: the density correlations will be 
1370: compared with the results on the $t$-$J$ model
1371: obtained with the high-temperature series expansions
1372: [Putikka {\it et al.} 1994].
1373: This gives valuable information on the
1374: dynamics of the density fluctuations in the 2D Hubbard model
1375: 
1376: The high-temperature series expansions have
1377: found that the structure in $S({\bf q})$ 
1378: can be described by the noninteracting spinless-fermion model.
1379: This is due to the fact that the double-occupancy of the sites 
1380: is not allowed in the $t$-$J$ model, and once this 
1381: constraint is taken into account, the density 
1382: correlations are similar to those of the noninteracting particles.
1383: Consequently, for a filling $\langle n\rangle$, 
1384: $S({\bf q})$ has structure at wave vectors which are 
1385: associated with the Fermi surface corresponding to a filling 
1386: of $2\langle n\rangle$.
1387: In other words, 
1388: the structure in $S({\bf q})$ follows the 
1389: "$4{\bf k}_F$" wave vectors
1390: rather than the "$2{\bf k}_F$" wave vectors
1391: of the system.
1392: 
1393: First, results at half-filling will be presented.
1394: Figure~4.1 shows QMC results on 
1395: $\Pi({\bf q},0)$ versus ${\bf q}$ at various 
1396: temperatures for $U=8t$.
1397: Here, 
1398: $\Pi({\bf q},0)$ is rather featureless in ${\bf q}$, and it gets 
1399: further suppressed as $T$ is lowered.
1400: This is due to the Mott-Hubbard gap which exists at 
1401: half-filling.
1402: Next, results for $\langle n\rangle=0.87$ are shown.
1403: Figure~4.2(a) shows the temperature evolution of 
1404: $\Pi({\bf q},0)$ versus ${\bf q}$ for $U=8t$.
1405: In this case, the ${\bf q}\sim 0$ component of 
1406: $\Pi({\bf q},0)$ gets enhanced as $T$ decreases.
1407: The $T$ dependence of $\Pi({\bf q}\rightarrow 0,0)$ 
1408: seen in Fig.~4.2(a) is consistent with the results of  
1409: the zero-temperature QMC calculations
1410: of the charge compressibility
1411: [Furukawa and Imada 1991 and 1992].
1412: Figure~4.2(b) shows ${\rm Im}\,\Pi({\bf q},\omega)$ versus
1413: $\omega$ for ${\bf q}=(\pi/4,0)$,
1414: where it is seen that the density fluctuations soften 
1415: as $T$ is lowered.
1416: The spectral weight 
1417: ${\rm Im}\,\Pi$ was obtained by the Pade analytic 
1418: continuation of the QMC data.
1419: 
1420: The evolution of the ${\bf q}$ structure in $\Pi({\bf q},0)$ 
1421: with $U/t$ is shown in Fig.~4.3(a) for $T=0.5t$
1422: and $\langle n\rangle=0.87$.
1423: For comparison, in Fig.~4.3(b) results  are shown for 
1424: $\Pi_0({\bf q},i\omega_m=0)$ of the $U=0$ system
1425: given by 
1426: \begin{equation}
1427: \label{P0}
1428: \Pi_0({\bf q},i\omega_m) = 2
1429: {1 \over N} \sum_{\bf p}
1430: { f(\varepsilon_{{\bf p}+{\bf q}}) - f(\varepsilon_{\bf p}) 
1431: \over 
1432: i\omega_m - ( \varepsilon_{{\bf p}+{\bf q}} 
1433: - \varepsilon_{\bf p} ) + i\delta 
1434: }.
1435: \end{equation}
1436: In this figure,
1437: the solid curve represents $\Pi_0({\bf q},0)$ for 
1438: $\langle n\rangle=0.87$ and the dotted curve is for 
1439: $\langle n\rangle=1.74$, 
1440: which corresponds to $\Pi({\bf q},0)$ 
1441: of the noninteracting spinless-fermion system for
1442: $\langle n\rangle=0.87$.
1443: As $U/t$ increases, it is seen that 
1444: $\Pi({\bf q},0)$ develops features which are more similar to 
1445: those seen for the noninteracting spinless-fermion system.
1446: In Fig.~4.3(a), 
1447: it is seen that the peak in $\Pi({\bf q},0)$ shifts 
1448: from ${\bf q}=(\pi,\pi)$ to ${\bf q}\sim 0$ as $U/t$ 
1449: increases.
1450: It is not possible to obtain this type of 
1451: momentum dependence for $\Pi({\bf q},0)$ 
1452: from the simple RPA form
1453: \begin{equation}
1454: \Pi_{\rm RPA}({\bf q},i\omega_m) =
1455: { \Pi_0({\bf q},i\omega_m) \over 
1456: 1 + {1\over 2} \overline{U} 
1457: \Pi_0({\bf q},i\omega_m) },
1458: \end{equation}
1459: where $\overline{U}$ is a constant 
1460: representing the effective coupling 
1461: in the density channel.
1462: 
1463: \begin{figure}
1464: \centering
1465: \iffigure
1466: \epsfysize=8cm
1467: \epsffile[100 150 550 610]{ch4-fig1.ps}
1468: \fi
1469: \caption{
1470: Momentum dependence of $\Pi({\bf q},0)$
1471: at half-filling for $U=8t$ and various temperatures.
1472: }
1473: \label{4.1}
1474: \end{figure}
1475: 
1476: \begin{figure}
1477: \centering
1478: \iffigure
1479: \mbox{
1480: \subfigure{
1481: \epsfysize=8cm
1482: \epsffile[100 150 480 610]{ch4-fig2a.ps}}
1483: \quad
1484: \subfigure{
1485: \epsfysize=8cm
1486: \epsffile[50 150 600 610]{ch4-fig2b.ps}}}
1487: \fi
1488: \caption{
1489: (a) Momentum dependence of $\Pi({\bf q},0)$.
1490: (b) ${\rm Im}\,\Pi({\bf q},\omega)$ versus $\omega$
1491: at ${\bf q}=(\pi/4,0)$.
1492: These results are
1493: for $\langle n\rangle=0.87$, $U=8t$ and various temperatures.
1494: }
1495: \label{4.2}
1496: \end{figure}
1497: 
1498: \begin{figure}
1499: \centering
1500: \iffigure
1501: \mbox{
1502: \subfigure[]{
1503: \epsfysize=8cm
1504: \epsffile[100 150 480 610]{ch4-fig3a.ps}}
1505: \quad
1506: \subfigure[]{
1507: \epsfysize=8cm
1508: \epsffile[50 150 600 610]{ch4-fig3b.ps}}}
1509: \fi
1510: \caption{
1511: (a) $\Pi({\bf q},0)$ versus ${\bf q}$ 
1512: at $\langle n\rangle=0.87$ 
1513: for various values of $U$.
1514: (b) $\Pi_0({\bf q},0)$ versus ${\bf q}$ 
1515: for the $U=0$ system at
1516: $\langle n\rangle=0.87$ (solid curve)
1517: and $\langle n\rangle=1.74$ (dotted curve).
1518: These results are for $T=0.5t$.
1519: }
1520: \label{4.3}
1521: \end{figure}
1522: 
1523: Next, QMC results on $S({\bf q})$ will be discussed.
1524: Figure~4.4 shows $S({\bf q})$ versus ${\bf q}$ at various 
1525: temperatures for $U=8t$. 
1526: Here, it is seen that 
1527: $S({\bf q})$ exhibits small variation with $T/t$.
1528: Figure~4.5(a) shows the evolution of $S({\bf q})$ with 
1529: $U/t$ for $T=0.5t$ and $\langle n\rangle=0.87$.
1530: Here, it is seen that with $U/t$ increasing, $S({\bf q})$ gets
1531: suppressed.
1532: However,
1533: in contrast to the case of $\Pi({\bf q},0)$, 
1534: there is no qualitative change in the ${\bf q}$ structure 
1535: of $S({\bf q})$ as $U/t$ increases,
1536: and the peak remains at $(\pi,\pi)$.
1537: It is useful to compare these QMC results 
1538: with $S_0({\bf q})$ of the noninteracting system.
1539: Figure~4.5(b) shows $S_0({\bf q})$ versus ${\bf q}$ 
1540: at $T=0.5t$ for $\langle n\rangle=0.87$ (solid curve)
1541: and $\langle n\rangle=1.74$ (dotted curve).
1542: Comparing Figs.~4.5(a) and (b), one observes that the QMC data on 
1543: $S({\bf q})$ is more similar to $S_0({\bf q})$ for 
1544: $\langle n\rangle=0.87$,
1545: and the ${\bf q}$ structure in $S({\bf q})$ 
1546: does not follow that of the 
1547: noninteracting spinless-fermion system.
1548: These results on $S({\bf q})$ are consistent with 
1549: those found by the GFMC technique for the ground state
1550: [Becca {\it et al.} 2000],
1551: and by the QMC method at finite temperatures
1552: [Chen {\it et al.} 1994].
1553: 
1554: \begin{figure}
1555: \centering
1556: \iffigure
1557: \epsfysize=8cm
1558: \epsffile[100 150 550 610]{ch4-fig4.ps}
1559: \fi
1560: \caption{
1561: Equal-time density correlation function $S({\bf q})$
1562: versus ${\bf q}$
1563: at $\langle n\rangle=0.87$ for $U=8t$ and various temperatures.
1564: }
1565: \label{4.4}
1566: \end{figure}
1567: 
1568: \begin{figure}
1569: \centering
1570: \iffigure
1571: \mbox{
1572: \subfigure[]{
1573: \epsfysize=8cm
1574: \epsffile[100 150 480 610]{ch4-fig5a.ps}}
1575: \quad
1576: \subfigure[]{
1577: \epsfysize=8cm
1578: \epsffile[50 150 600 610]{ch4-fig5b.ps}}}
1579: \fi
1580: \caption{
1581: (a) $S({\bf q})$ versus ${\bf q}$ 
1582: at $\langle n\rangle=0.87$ 
1583: for various values values of $U$.
1584: (b) $S_0({\bf q})$ versus ${\bf q}$ 
1585: of the $U=0$ system for 
1586: $\langle n\rangle=0.87$ (solid curve)
1587: and $\langle n\rangle=1.74$ (dotted curve).
1588: These results are for $T=0.5t$.
1589: }
1590: \label{4.5}
1591: \end{figure}
1592: 
1593: Here, it has been seen that
1594: the results on $S({\bf q})$ for the 2D Hubbard model  
1595: are quite different from those calculated
1596: for the 2D $t$-$J$ model for $J\sim 0.5t$ with the 
1597: high-temperature series expansion
1598: [Putikka {\it et al.} 1994]
1599: or by the GFMC technique [Calandra {\it et al.} 1998].
1600: In the 2D Hubbard model near half-filling,
1601: $S({\bf q})$ does not exhibit 
1602: obvious features which might be associated
1603: with the "$4{\bf k}_F$" wave vectors of the Fermi surface.
1604: However, $\Pi({\bf q},0)$ has structure, 
1605: which might be associated with the "$4{\bf k}_F$" wave vectors, 
1606: when $U/t$ is large.
1607: Hence, 
1608: the dynamical density fluctuations in 2D Hubbard model
1609: with large $U/t$ have features 
1610: which might be related to the "$4{\bf k}_F$" wave vectors,
1611: while the static density correlations 
1612: appear to follow the 
1613: "$2{\bf k}_F$" wave vectors.
1614: However, results on $\Pi({\bf q},\omega)$ 
1615: at lower temperatures and on bigger lattices
1616: are necessary before reaching conclusions.
1617: 
1618: The nature of the density correlations 
1619: in the Hubbard model will be discussed further
1620: in the remainder of the article.
1621: In Section~7.3, comparisons will be made with the 2-leg
1622: Hubbard ladder, in which case power-law decaying 
1623: "$4{\bf k}_F$" CDW correlations have been found
1624: [Noack {\it et al.} 1996].
1625: In addition, in Section~8
1626: there will be further comparisons with the 
1627: density correlations in the 2D $t$-$J$ model.
1628: 
1629: \subsection{Optical conductivity}
1630: 
1631: In this section, 
1632: the results of the numerical calculations on
1633: the optical conductivity 
1634: $\sigma_1(\omega)$ of the 2D Hubbard model will be discussed.
1635: The optical conductivity for the 2D Hubbard model 
1636: was calculated with the exact diagonalization technique 
1637: for a $4\times 4$ lattice with $U=10t$
1638: at half-filling and in the doped case
1639: [Dagotto {\it et al.} 1992c].
1640: In these calculations,
1641: the insulating gap in $\sigma_1(\omega)$
1642: at half-filling is clearly seen. 
1643: Upon doping to $\langle n\rangle = 0.875$,
1644: the amount of the spectral weight 
1645: above the insulating gap is reduced. 
1646: In this case, the additional features found in $\sigma_1(\omega)$
1647: are a Drude peak at $\omega=0$ and spectral weight induced 
1648: at intermediate energies.
1649: The Drude weight of the Hubbard model was also calculated 
1650: with the QMC simulations
1651: [Scalapino {\it et al.} 1992 and 1993].
1652: A review of the exact diagonalization calculations 
1653: for $\sigma_1(\omega)$ and of comparisons with the 
1654: experimental data on the cuprates can be found 
1655: in Ref.~[Dagotto 1994].
1656: 
1657: The real part of the frequency-dependent 
1658: ${\bf q}=0$ conductivity is given by 
1659: \begin{equation} 
1660: \sigma_1(\omega) = 
1661: {\rm Re}\,{\Lambda_{xx}(i\omega_m) \over i\omega_m}
1662: \bigg|_{i\omega_m \rightarrow \omega+i\delta}
1663: \end{equation}
1664: where the current-current correlation function 
1665: $\Lambda_{xx}$ is defined as
1666: \begin{equation}
1667: \Lambda_{xx}(i\omega_m) = 
1668: \int_0^{\beta} d\tau\, 
1669: e^{-i\omega_m\tau} 
1670: \langle j_x(\tau) j_x(0) \rangle
1671: \end{equation}
1672: with
1673: \begin{equation}
1674: j_x = -itea\sum_{is} 
1675: ( c^{\dagger}_{is} c_{i+xs} - 
1676: c^{\dagger}_{i+xs}c_{is} ).
1677: \end{equation}
1678: The analytic continuation in Eq.~(4.5) is carried out using 
1679: the maximum entropy procedure. 
1680: In the following, the hopping $t$, the electron charge $e$ and 
1681: the lattice constant $a$ are set to unity.
1682: 
1683: The solid curve in Fig.~4.6 shows 
1684: $\sigma_1(\omega)$ versus $\omega$ 
1685: at half-filling for $U=8t$, $T=0.125t$ 
1686: and an $8\times 8$ lattice. 
1687: These are data from Ref.~[Bulut {\it et al.} 1994a].
1688: Here, 
1689: in spite of the limited resolution 
1690: of the analytic continuation procedure, 
1691: the insulating gap in $\sigma_1(\omega)$ is clearly seen. 
1692: The dotted curve in this graph 
1693: is the mean-field SDW result given by 
1694: \begin{eqnarray} 
1695: \sigma_1(\omega) = {2\pi \over N}
1696: \sum_{\bf p} \, 
1697: {1\over 2}
1698: \left( 1 - {\varepsilon^2_{\bf p} - \Delta^2 \over E^2_{\bf p}} \right) 
1699: \left( 1 - 2f(E_{\bf p}) \over E_{\bf p} \right)
1700: \sin^2{p_x}
1701: \end{eqnarray}
1702: which was calculated using $\Delta=2.4t$.  
1703: Near the threshold $\omega \simeq 2\Delta$, 
1704: the SDW coherence factor 
1705: ${1\over 2}[ 1 - (\varepsilon^2_{\bf p}-\Delta^2)/E^2_{\bf p} ]$ 
1706: in Eq.~4.8 goes to 1 for $\varepsilon_{\bf p}$ near zero, and 
1707: hence the peak at the threshold does not vanish within SDW. 
1708: A check of the maximum entropy analytic continuation 
1709: is the $f$-sum rule which has the form
1710: \begin{equation}
1711: \int_0^{\infty} d\omega\, \sigma_1(\omega) = 
1712: {\pi \over 2} \langle -k_x \rangle, 
1713: \end{equation}
1714: where $\langle k_x \rangle$ is the average kinetic 
1715: energy per site associated 
1716: with hopping in the $x$ direction. 
1717: The area under the solid curve in Fig.~4.6 is 0.74, 
1718: where the separate QMC measurement 
1719: of ${\pi \over 2} \langle -k_x \rangle$ gives 0.77.
1720: The difference reflects the difficulty in analytically continuing 
1721: the numerical data. 
1722: When the system is doped, 
1723: it is difficult to estimate the accuracy 
1724: of the analytic continuation procedure for $\sigma_1(\omega)$,
1725: especially in the limit $\omega\rightarrow 0$.
1726: Hence, 
1727: away from half-filling, 
1728: the maximum entropy or the Pade techniques were not used to 
1729: extract $\sigma_1(\omega)$ or the dc resistivity $\rho(T)$
1730: from the QMC data on 
1731: the current-current correlation function $\Lambda_{xx}$.
1732: 
1733: \begin{figure}
1734: \centering
1735: \iffigure
1736: \epsfysize=8cm
1737: \epsffile[100 150 550 610]{ch4-fig6.ps}
1738: \fi
1739: \caption{
1740: Real part of the frequency dependent conductivity of the half-filled
1741: Hubbard model with $U=8t$ (solid curve).
1742: These results were obtained for an $8\times 8$ lattice
1743: at $T=0.125t$.
1744: The dotted curve is the mean-field SDW result with $\Delta=2.4t$.
1745: A finite broadening of $\Gamma=0.5t$ has been used 
1746: in plotting the SDW result.
1747: }
1748: \label{4.6}
1749: \end{figure}
1750: 
1751: Even though it has not been possible to reliably extract 
1752: $\rho(T)$ for the 2D Hubbard model,
1753: for 2D $t$-$J$ clusters
1754: $\sigma_1(\omega)$ and $\rho(T)$ 
1755: were calculated at various fillings
1756: by using the finite-temperature Lanczos technique
1757: [Jaklic and Prelovsek 2000].
1758: For $J/t=0.3$, the results for $\sigma_1(\omega)$ and $\rho(T)$ 
1759: were compared with the experimental data 
1760: on cuprates in various doping regimes. 
1761: In some cases, 
1762: the agreement obtained with the experimental data
1763: was at a quantitative level. 
1764: In the intermediate doping regime, 
1765: where $0.75 < \langle n\rangle < 0.85$, 
1766: it was shown that $\sigma_1(\omega)$ 
1767: is not consistent with the usual Drude
1768: form but rather with the marginal Fermi 
1769: liquid concept where the only $\omega$ 
1770: scale is given by $T$ 
1771: [Varma {\it et al.} 1989].
1772: The finite-temperature Lanczos calculations 
1773: find also that $\rho(T)$ is proportional 
1774: to $T$ in this regime but the slope of 
1775: $\rho(T)$ changes at $T \sim J$.
1776: The underdoped and the overdoped regimes 
1777: were also studied with this technique.
1778: Jaklic and Prelovsek concluded that 
1779: the main features of the unusual normal state properties 
1780: of the cuprates could be attributed 
1781: to the large degeneracy of states 
1782: and the frustration induced by doping the antiferromagnet. 
1783: 
1784: \setcounter{equation}{0}\setcounter{figure}{0}
1785: \section{Single-particle properties of the 2D Hubbard model}
1786: 
1787: The angular-resolved photoemission spectroscopy (ARPES) measurements
1788: on the high-$T_c$ cuprates probe the single-particle properties 
1789: of these compounds and have found a number of 
1790: interesting features.
1791: In the normal state of the optimally doped cuprates,
1792: the ARPES experiments 
1793: have found that there are quasiparticle-like bands which cross
1794: the Fermi level leading to a large Fermi surface.
1795: An unusual feature is that these bands have
1796: extended flat dispersion near the $(\pi,0)$ and 
1797: $(0,\pi)$ points in the Brillouin zone
1798: [Dessau {\it et al.} 1993, Gofron {\it et al.}
1799: 1993].
1800: The ARPES experiments have also provided valuable 
1801: information about the evolution of the single-particle
1802: spectral weight with doping in the underdoped regime as the 
1803: insulating state is approached.
1804: Reviews of these experiments can be found in 
1805: Refs.~[Shen and Dessau 1995, Damascelli {\it et al.} 2001].
1806: Below, the numerical studies of the single-particle 
1807: spectral weight in the 2D Hubbard model will be reviewed.
1808: In particular, 
1809: the origin of the correlated metallic band 
1810: which develops at the Fermi level 
1811: upon doping the AF Mott-Hubbard insulator 
1812: will be investigated.
1813: It will be seen that the 2D Hubbard model provides an explanation 
1814: for some of the features found in the ARPES data.
1815: 
1816: The single-particle properties of the 2D Hubbard model 
1817: were studied with various many-body techniques.
1818: The single-particle spectral weight 
1819: was calculated phenomenologically by taking into account the 
1820: scattering of the quasiparticles by the AF spin fluctuations
1821: [Kampf and Schrieffer 1990a and 1990b].
1822: The evolution of the single-particle spectral weight 
1823: with doping was 
1824: also studied with the exact diagonalization technique 
1825: [Dagotto {\it et al.} 1991].
1826: There have been extensive numerical calculations 
1827: of $A({\bf p},\omega)$ by analytically continuing the QMC data 
1828: on the imaginary-time single-particle Green's function
1829: [White {\it et al.} 1989c, White 1991, Vekic and White 1993,
1830: Bulut {\it et al.} 1994a, Bulut and Scalapino 1995, 
1831: Moreo {\it et al.} 1995, Haas {\it et al.} 1995, 
1832: Duffy and Moreo 1995, Preuss {\it et al.} 1995 and 1997, 
1833: Gr\"ober {\it et al.} 2000].
1834: In these studies, 
1835: the maximum-entropy technique has been the main
1836: algorithm for the analytic continuation procedure.
1837: This technique
1838: was also used for calculating $A({\bf p},\omega)$ 
1839: for the three-band CuO$_2$ model [Dopf {\it et al.} 1992b] and 
1840: the one-dimensional Hubbard model 
1841: [Preuss {\it et al.} 1994, Zacher {\it et al.} 1998].
1842: These studies have provided valuable 
1843: information about the single-particle spectral weight in the 
1844: Hubbard model.
1845: In the following, the maximum-entropy results on the 
1846: 2D Hubbard model will be reviewed.
1847: Here, the emphasis will be on the 
1848: narrow correlated band which forms at the Fermi level 
1849: upon doping the insulator, 
1850: and on the flat bands which are
1851: observed near the $(\pi,0)$ and $(0,\pi)$ points 
1852: in the Brillouin zone.
1853: 
1854: The single-particle spectral weight 
1855: $A({\bf p},\omega)$ is given by
1856: \begin{equation}
1857: \label{Apw}
1858: A({\bf p},\omega) = - {1\over \pi}
1859: {\rm Im}\, G({\bf p},i\omega_n \rightarrow \omega+i\delta)
1860: \end{equation}
1861: where
1862: \begin{equation}
1863: G({\bf p},i\omega_n) = \int_0^{\beta} \, d\tau \, 
1864: e^{i\omega_n \tau} \,
1865: G({\bf p},\tau)
1866: \end{equation}
1867: and
1868: \begin{equation}
1869: \label{Gpt}
1870: G({\bf p},\tau) = - \langle T \,
1871: c_{{\bf p}\sigma}(\tau) 
1872: c^{\dagger}_{{\bf p}\sigma}(0) \rangle.
1873: \end{equation}
1874: With the maximum-entropy technique, one uses the QMC data on 
1875: $G({\bf p},\tau)$ to invert the integral equation
1876: \begin{equation}
1877: G({\bf p},\tau) = - \int_{-\infty}^{\infty} \, 
1878: d\omega  \,
1879: { e^{-\omega\tau} \over 1 + e^{-\beta\omega} } \,
1880: A({\bf p},\omega),
1881: \end{equation}
1882: in order to obtain $A({\bf p},\omega)$. 
1883: By applying the same procedure to 
1884: \begin{equation}
1885: G_{ii}(\tau) = - \langle T\, 
1886: c_{i\sigma}(\tau) c^{\dagger}_{i\sigma}(0) \rangle,
1887: \end{equation}
1888: the single-particle density of states 
1889: $N(\omega)$ is obtained. 
1890: The maximum-entropy technique for analytically continuing
1891: QMC data is described in 
1892: Refs.~[Silver {\it et al.} 1990, Jarrell and Gubernatis 1996].
1893: The statistical errors for $G_{ii}(\tau)$ are smaller than for 
1894: $G({\bf p},\tau)$,
1895: hence the results for $N(\omega)$ shown below have higher resolution. 
1896: Here, 
1897: first the results on the evolution of $N(\omega)$ with doping 
1898: will be reviewed, and later $A({\bf p},\omega)$ will be discussed.
1899: In the following, 
1900: $N(\omega)$ and $A({\bf p},\omega)$
1901: will be plotted in units of $t^{-1}$.
1902: The data which will be shown in this section are from 
1903: Refs.~[Bulut {\it et al.} 1994a, Bulut and Scalapino 1995].
1904: 
1905: \begin{figure}[ht]
1906: \centering
1907: \iffigure
1908: \epsfysize=8cm
1909: \epsffile[100 150 550 610]{ch5-fig1.ps}
1910: \fi
1911: \caption{
1912: Single-particle density of states $N(\omega)$ versus
1913: $\omega$.
1914: These results were obtained for 
1915: $U=8t$ and $\langle n\rangle =1.0$ on a
1916: $12\times 12$ lattice.
1917: }
1918: \label{5.1}
1919: \end{figure}
1920: 
1921: \subsection{Single-particle density of states}
1922: 
1923: Figure 5.1 shows the single-particle 
1924: density of states $N(\omega)$ versus $\omega$
1925: at half-filling for $U=8t$ on a $12\times 12$ lattice
1926: at $T=0.5$ and $0.125t$.
1927: Here, the chemical potential $\mu$ 
1928: is located at $\omega=0$.
1929: In this figure, 
1930: it is seen that at $T=0.5t$, 
1931: $N(\omega)$
1932: consists of two peaks corresponding to the lower 
1933: and the upper Hubbard bands, 
1934: which are separated by the Mott-Hubbard pseudogap.
1935: At this temperature, the AF correlation length 
1936: is still less than the system size, 
1937: as it was seen in Section 3.1.
1938: As $T$ is lowered down to $0.125t$, 
1939: the AF correlation length reaches the system size, and 
1940: in this case additional sharp peaks appear at the upper edge of the 
1941: lower Hubbard band and at the lower edge of 
1942: the upper Hubbard band.
1943: At $T=0.125t$,
1944: the pseudogap has become a full gap 
1945: in the single-particle spectrum 
1946: with a magnitude of $2\Delta \approx 4.5t$.
1947: The size of the gap $2\Delta$ was also calculated within 
1948: an SDW approach where the single-particle self-energy 
1949: corrections were included [Schrieffer {\it et al.} 1989].
1950: Within this approach, for $U=8t$, 
1951: $2\Delta$ is found to be about $4.8t$, while for
1952: $U=4t$, $2\Delta$ is about $2t$.
1953: 
1954: The sharp peaks which are located below and above the 
1955: Mott-Hubbard gap appear when 
1956: the system has long-range AF order.
1957: These narrow bands have a bandwidth of about $1t$, 
1958: corresponding to $\approx 2J$ where 
1959: $J=4t^2/U$, and they exhibit SDW-like dispersion, 
1960: which will be discussed in the next section.
1961: It is known that in the half-filled 2D $t$-$J$ model 
1962: the quasiparticle bandwidth is about $2J$
1963: [Liu and Manousakis 1992].
1964: The exact diagonalization calculations for the 2D $t$-$J$ 
1965: model also find SDW-like quasiparticle
1966: excitations with a bandwidth of about $2.2J$ at half-filling
1967: [Dagotto 1994].
1968: In the $t$-$J$ model at half-filling, 
1969: one doped hole propagates by flipping the spins in the AF background 
1970: around it.
1971: For this reason, 
1972: the bandwidth for the hole motion is determined by 
1973: the magnetic exchange $J$ rather 
1974: than the hopping matrix element $t$.
1975: The QMC data implies that 
1976: the hole propagation is accompanied by 
1977: similar many-body processes in the 
1978: half-filled 2D Hubbard model.
1979: 
1980: \begin{figure}[ht]
1981: \centerline{
1982: \epsfysize=6cm \epsffile[-207 164 367 598]{ch5-fig2a.ps}
1983: \epsfysize=6cm \epsffile[18 164 592 598]{ch5-fig2b.ps}
1984: \epsfysize=6cm \epsffile[243 164 817 598]{ch5-fig2c.ps}}
1985: \caption{
1986: Single-particle density of states $N(\omega)$ versus
1987: $\omega$ at fillings (a) $\langle n\rangle=0.94$, 
1988: (b) 0.87 and (c) 0.70.
1989: These results are for $T=0.5t$ and $U=8t$.
1990: }
1991: \label{5.2}
1992: \end{figure}
1993: 
1994: Next, results on the evolution of $N(\omega)$ 
1995: with doping are shown.
1996: Figure~5.2 shows $N(\omega)$ versus $\omega$ at fillings 
1997: $\langle n\rangle =0.94$, 0.87 and 0.70
1998: for $U=8t$ and $T=0.5t$.
1999: Upon hole doping, the chemical potential moves
2000: rapidly from $\omega=0$ to the top of the lower Hubbard band.
2001: This is accompanied with a gradual transfer of spectral
2002: weight from above the Mott-Hubbard gap to the Fermi level.
2003: The transferred spectral weight goes into forming 
2004: a narrow metallic band at the Fermi level.
2005: These data are in agreement 
2006: with the exact diagonalization calculations
2007: [Dagotto {\it et al.} 1991].
2008: In the overdoped regime at $\langle n\rangle=0.70$, 
2009: the Fermi level lies below the metallic band.
2010: The temperature variation of $N(\omega)$ is shown in 
2011: Fig.~5.3 for $\langle n\rangle=0.87$ and $U=8t$.
2012: Here, the build up of a narrow band with a width 
2013: between $4J$ and $5J$ is clearly seen as $T$ decreases.
2014: The comparison of Figs.~5.3 and 5.1 indicates that 
2015: this correlated metallic band forms out
2016: of the SDW-like bands of the half-filled case.
2017: 
2018: At this point,
2019: it should be noted that the general features of 
2020: $N(\omega)$ seen in Fig.~5.3 are similar to what is found for the 
2021: infinite-dimensional Hubbard model away from half-filling
2022: [Jarrell 1992].
2023: $N(\omega)$ for the infinite-dimensional Hubbard model 
2024: is calculated by first mapping the model to an 
2025: Anderson impurity problem, which is then solved exactly.
2026: In this case,
2027: a narrow peak at the Fermi level is also found in addition to 
2028: the lower and upper Hubbard bands.
2029: 
2030: \begin{figure}[ht]
2031: \centering
2032: \iffigure
2033: \epsfysize=8cm
2034: \epsffile[100 150 550 610]{ch5-fig3.ps}
2035: \fi
2036: \caption{
2037: Development of the 
2038: single-particle density of states $N(\omega)$ as the 
2039: temperature is lowered for 
2040: $\langle n\rangle=0.87$ and $U=8t$.
2041: }
2042: \label{5.3}
2043: \end{figure}
2044: 
2045: \begin{figure}
2046: \centering
2047: \iffigure
2048: \mbox{
2049: \subfigure{
2050: \epsfysize=8cm
2051: \epsffile[100 150 480 610]{ch5-fig4a.ps}}
2052: \quad
2053: \subfigure{
2054: \epsfysize=8cm
2055: \epsffile[50 150 600 610]{ch5-fig4b.ps}}}
2056: \fi
2057: \caption{
2058: $N(\omega)$ versus $\omega$ for (a) $U=4t$ and 
2059: (b) $U=12t$ at $\langle n\rangle=0.87$ and $T=0.5t$.
2060: }
2061: \label{5.4}
2062: \end{figure}
2063: 
2064: It is useful to compare these results on $N(\omega)$ for 
2065: $U=8t$ with the results for $U=4t$ and $12t$, 
2066: which are shown in Fig.~5.4.
2067: For $U=12t$, the narrow metallic band is further separated 
2068: from the lower and the upper Hubbard bands,
2069: and the pseudogap is bigger.
2070: On the other hand,
2071: for $U=4t$, the pseudogap is not observed at these 
2072: temperatures, even though there are hump-like structures which 
2073: might be attributed to the lower and the upper Hubbard bands.
2074: In addition, the peak at the Fermi level is broader.
2075: Hence, as $U$ increases from $4t$ to $12t$, 
2076: the distribution of the weight in $N(\omega)$ changes
2077: considerably.
2078: However, the density of states at the Fermi level 
2079: $N(\mu)$ varies by a small amount.
2080: 
2081: It is also useful to discuss how $N(\mu)$ varies with doping 
2082: at fixed $U$ and $T$.
2083: Figure~5.5 shows $N(\mu)$ versus $\langle n\rangle$ for 
2084: $U=8t$ at $T=0.5t$ (filled circles) and $1.0t$ (open circles).
2085: It is clearly seen that at these temperatures $N(\mu)$ 
2086: has a peak at finite doping, and 
2087: a depression in $N(\mu)$ exists near half-filling.
2088: However, it is not possible to extract 
2089: the $T\rightarrow 0$ behavior.
2090: 
2091: \begin{figure}[ht]
2092: \centering
2093: \iffigure
2094: \epsfysize=8cm
2095: \epsffile[100 150 550 610]{ch5-fig5.ps}
2096: \fi
2097: \caption{
2098: Filling dependence of the density of states at the Fermi level
2099: $N(\mu)$ for $U=8t$ at temperatures $T=0.5t$ (filled circles)
2100: and $1.0t$ (open circles).
2101: }
2102: \label{5.5}
2103: \end{figure}
2104: 
2105: \subsection{Single-particle spectral weight}
2106: 
2107: In this section, results on the single-particle spectral weight 
2108: $A({\bf p},\omega)$ will be discussed.
2109: It is useful to first study the effects of varying $U$ on 
2110: $A({\bf p},\omega)$.
2111: Figure~5.6 shows $A({\bf p},\omega)$ versus $\omega$ at various
2112: temperatures for $U=8t$ and $4t$.
2113: These results were obtained for the ${\bf p}=(\pi/2,\pi/2)$ 
2114: point on an $8\times 8$ lattice.
2115: For both values of $U$, a heavily-damped 
2116: quasiparticle-like peak develops near the
2117: Fermi level as the temperature decreases.
2118: However, for $U=8t$ this peak is broader and it has reduced weight.
2119: In addition, in this case, the peak location varies with the
2120: temperature.
2121: Even though the maximum-entropy technique has an intrinsic broadening 
2122: which is difficult to estimate, 
2123: these results suggest that the damping of the quasiparticles
2124: is stronger for larger $U/t$.
2125: 
2126: \begin{figure}
2127: \centering
2128: \iffigure
2129: \mbox{
2130: \subfigure{
2131: \epsfysize=8cm
2132: \epsffile[100 150 480 610]{ch5-fig6a.ps}}
2133: \quad
2134: \subfigure{
2135: \epsfysize=8cm
2136: \epsffile[50 150 600 610]{ch5-fig6b.ps}}}
2137: \fi
2138: \caption{
2139: Comparison of
2140: $A({\bf p}=(\pi/2,\pi/2),\omega)$ versus $\omega$ 
2141: for (a) $U=8t$ and (b) $U=4t$.
2142: Here results are shown at various temperatures 
2143: for $\langle n\rangle=0.87$, 
2144: and the frequency axis is shifted such that the Fermi level 
2145: occurs at $\omega=0$.
2146: }
2147: \label{5.6}
2148: \end{figure}
2149: 
2150: In the previous section, it was seen that for 
2151: $U=8t$ a narrow metallic band develops at the Fermi level 
2152: upon doping.
2153: In order to gain insight into the origin of this band, 
2154: in Fig.~5.7, $A({\bf p},\omega)$ versus $\omega$ is plotted for 
2155: ${\bf p}$ taken along various cuts in the Brillouin zone 
2156: as indicated in the insets.
2157: These data were obtained on a $12 \times 12$ lattice for
2158: $U=8t$, $\langle n\rangle=0.87$ and T$=0.5t$.
2159: In this figure, spectral weight representing the upper 
2160: and the lower Hubbard bands are seen in addition to 
2161: a quasiparticle band which crosses the Fermi level.
2162: The quasiparticle peak is especially of interest.
2163: When ${\bf p}$ is on the Fermi surface, this peak is 
2164: centred at $\omega=0$.
2165: As ${\bf p}$ moves outside of the Fermi surface, 
2166: the quasiparticle peak is clearly resolved.
2167: However, when ${\bf p}$ is below the Fermi surface, 
2168: for instance at ${\bf p}=(0,0)$,
2169: the quasiparticle peak is obscured by the lower Hubbard band, 
2170: even though the exact diagonalization calculations
2171: [Dagotto {\it et al.} 1992a]
2172: find a quasiparticle peak at this momentum in addition to the 
2173: lower Hubbard band.
2174: This is because the maximum-entropy algorithm used here has 
2175: poor resolution when $\omega$ is away from the Fermi level.
2176: Indeed, 
2177: similar calculations with higher resolution were able to observe
2178: the quasiparticle peak below the Fermi surface
2179: [Preuss {\it et al.} 1995].
2180: An unresolved issue is whether the dispersing band near the Fermi level 
2181: corresponds to a true quasiparticle band.
2182: However, the exact determination of the energy and the temperature
2183: scaling of the spectral weight cannot be carried out 
2184: because the maximum-entropy technique has finite resolution and, 
2185: in addition, bigger lattices and lower temperatures
2186: are required.
2187: 
2188: \begin{figure}
2189: \centerline{
2190: \epsfysize=9cm \epsffile[-207 124 367 778]{ch5-fig7a.ps} 
2191: \epsfysize=9cm \epsffile[18 124 592 778]{ch5-fig7b.ps} 
2192: \epsfysize=9cm \epsffile[243 124 817 778]{ch5-fig7c.ps}}
2193: \caption{
2194: Single-particle spectral weight $A({\bf p},\omega)$ versus
2195: $\omega$ for various momentum cuts through the 
2196: Brillouin zone for $U=8t$ and $\langle n\rangle = 0.87$.
2197: Here, the frequency axis is shifted such that the Fermi level 
2198: occurs at $\omega=0$.
2199: }
2200: \label{5.7}
2201: \end{figure}
2202: 
2203: An important feature of the quasiparticle band 
2204: is that it is unusually flat near the $(\pi,0)$ 
2205: and $(0,\pi)$ points of the Brillouin zone. 
2206: This causes large amount of spectral weight 
2207: to be near the Fermi level.  
2208: Even though the weight in the quasiparticle peak near 
2209: the Fermi level decreases
2210: as $U$ increases from $4t$ to $8t$, 
2211: the density of states at
2212: the Fermi level stays nearly the same. 
2213: This is because of the flat bands which are pinned near 
2214: the Fermi level. 
2215: The presence of the flat bands is especially important 
2216: because they generate phase space 
2217: for the scattering of the quasiparticles 
2218: in the $d_{x^2-y^2}$-wave pairing channel.
2219: These features found in the 2D Hubbard model 
2220: are similar to the ARPES data on the cuprates 
2221: [Dessau {\it et al.} 1993, Gofron {\it et al.} 1993].
2222: 
2223: In addition to the QMC data,
2224: in Ref.~[Dagotto {\it et al.} 1994], 
2225: the relation with the flat bands 
2226: seen in the ARPES data were noted by using 
2227: the dispersion of one hole in the half-filled $t$-$J$
2228: model calculated on a $16\times 16$ lattice.
2229: It should also be noted that Beenen and Edwards
2230: have found good agreement with the main features of the 
2231: QMC data using a two-pole approximation in 
2232: calculating the single-particle Green's function
2233: [Beenen and Edwards 1995].
2234: Dorneich {\it et al.} have extended this type of calculations 
2235: for the Hubbard model for the case of $\langle n\rangle=1$
2236: [Dorneich {\it et al.} 2000].
2237: Finally, it should be noted that 
2238: a number of transport measurements
2239: on the cuprates have been interpreted in terms of the Fermi
2240: level being close to a van Hove singularity
2241: [Tsuei {\it et al.} 1992, Newns {\it et al.} 1994].
2242: Here, 
2243: it is seen that in the 2D Hubbard model 
2244: there can be extended flat bands near the Fermi level and 
2245: they are produced by the many-body effects 
2246: rather than being due to the one-electron band structure.
2247: 
2248: These calculations for $A({\bf p},\omega)$ were also 
2249: carried out at half-filling, and the results were fitted 
2250: with the SDW form [Bulut {\it et al.} 1994a],
2251: \begin{equation}
2252: A({\bf p},\omega) = u^2_{\bf p} \delta(\omega-E_{\bf p}) +
2253: v^2_{\bf p}\delta(\omega+E_{\bf p})
2254: \end{equation}
2255: where $u^2_{\bf p}={1\over 2}( 1 + \gamma_{\bf p}/E_{\bf p} )$,
2256: $v^2_{\bf p} = {1\over 2}(1 - \gamma_{\bf p}/E_{\bf p})$,
2257: $\gamma_{\bf p} = -2t(\cos{p_x} + \cos(p_y))$, 
2258: and $E_{\bf p}=\sqrt{\gamma^2_{\bf p} + \Delta^2}$,
2259: For large $\Delta/t$,
2260: \begin{equation}
2261: E_{\bf p} = \sqrt{ \gamma^2_{\bf p} + \Delta^2 }
2262: \simeq \Delta + J(\cos{p_x} + \cos{p_y})^2. 
2263: \end{equation}
2264: Similar calculations with higher resolution 
2265: [Preuss {\it et al.} 1995] were carried out, which
2266: showed that, in fact, the spectrum is better described by 
2267: $E_{\bf p}\simeq \Delta + {J \over 2}( \cos{p_x} + \cos{p_y} )^2$,
2268: hence, 
2269: the quasiparticle band at half-filling has a width of $2J$ 
2270: in agreement with the results on 
2271: $N(\omega)$ at half-filling and 
2272: with the calculations on the $t$-$J$ model with one hole.
2273: 
2274: Based on the results discussed above, 
2275: schematic dispersion 
2276: relations were constructed for the 2D Hubbard model, 
2277: which are shown in Fig.~5.8
2278: for $\langle n\rangle=0.87$ and $\langle n\rangle=1.0$.
2279: In Fig.~5.8(a),
2280: the dashed horizontal line represents the chemical potential,
2281: and the solid curve denotes the narrow metallic band located
2282: at the Fermi level.
2283: The shaded regions represent the lower and the upper 
2284: Hubbard bands.
2285: The width of the quasiparticle band is of order $4J$.
2286: Figure~5.8(b) illustrates the distribution of the spectral 
2287: weight at half-filling.
2288: Here, the solid and the dotted lines denote 
2289: the SDW-like quasiparticle bands.
2290: The picture which emerges from these calculations is that the 
2291: narrow metallic band which lies at the Fermi level
2292: for $\langle n\rangle=0.87$ forms out of the SDW-like bands
2293: which exist in the half-filled case.
2294: 
2295: The QMC results seen in Fig.~5.7 were obtained at $T=0.5t$ for 
2296: $\langle n\rangle =0.87$, where the AF correlation length 
2297: $\xi$ is less than one lattice spacing. 
2298: The QMC simulations were also carried out at $T=0.25t$
2299: for $\langle n \rangle =0.93$, where $\xi$ is about 1.2
2300: lattice spacing 
2301: [Preuss {\it et al.} 1997].
2302: In this case, 
2303: spectral weight has been observed at wave vectors 
2304: which are displaced by ${\bf Q}=(\pi,\pi)$ with respect to the 
2305: narrow metallic band at the Fermi level.
2306: These results are consistent with the calculations 
2307: where "shadow bands" due to the scattering 
2308: of the quasiparticles by the AF spin fluctuations were found
2309: [Kampf and Schrieffer 1990b].
2310: This type of shadow structures were also observed in the 
2311: photoemission experiments 
2312: [Aebi {\it et al.} 1995].
2313: Detailed single-particle spectra 
2314: in the low-doping regime of the 2D Hubbard 
2315: model were obtained recently
2316: [Gr\"ober {\it et al.} 2000].
2317: 
2318: \begin{figure}[ht]
2319: \centering
2320: \iffigure
2321: \mbox{
2322: \subfigure[]{
2323: \epsfysize=8cm
2324: \epsffile[100 150 480 610]{ch5-fig8a.ps}}
2325: \quad
2326: \subfigure[]{
2327: \epsfysize=8cm
2328: \epsffile[50 150 600 610]{ch5-fig8b.ps}}}
2329: \fi
2330: \caption{
2331: Schematic drawing of 
2332: $E_{\bf p}$ versus ${\bf p}$ at $U=8t$ 
2333: for (a) $\langle n\rangle =0.87$ 
2334: and (b) $\langle n\rangle=1.0$.
2335: Here, the shaded areas represent the lower and the upper Hubbard
2336: bands, and the solid and the dotted curves indicate the 
2337: quasiparticle bands.
2338: In (a), the solid points indicate the position 
2339: of the quasiparticle peak obtained from Fig.~5.7
2340: when the quasiparticle peak is not obscured by the lower Hubbard band,
2341: and the horizontal dashed line denotes the chemical potential
2342: $\mu$.
2343: }
2344: \label{5.8}
2345: \end{figure}
2346: 
2347: In this section, 
2348: it was seen that the single-particle 
2349: properties in the 2D Hubbard model 
2350: are determined by the AF spin fluctuations and the 
2351: Coulomb correlations.
2352: An especially important feature of the single-particle 
2353: spectral weight is the flat bands near 
2354: ${\bf p}=(\pi,0)$ and $(0,\pi)$.
2355: They create large amount of phase space for the scattering of the 
2356: quasiparticles in the $d_{x^2-y^2}$-wave pairing channel.
2357: In Section 7, it will be seen that similar flat bands exist in the 
2358: 2-leg Hubbard ladder, and they play a key role 
2359: in determining the strength of the $d_{x^2-y^2}$-like 
2360: superconducting correlations in this system.
2361: In Section~8.1, these QMC results on the single-particle 
2362: spectral weight will be compared with those obtained from the 
2363: FLEX approximation, and the implications for the 
2364: $d_{x^2-y^2}$-wave pairing in the 2D Hubbard model will be 
2365: discussed.
2366: 
2367: \setcounter{equation}{0}\setcounter{figure}{0}
2368: \section{Pairing correlations in the 2D Hubbard model}
2369: 
2370: An important question for the high-$T_c$ cuprates is 
2371: whether the 2D Hubbard model exhibits superconductivity, 
2372: and if does, what is the nature of the pairing. 
2373: Various views were expressed on this subject, and 
2374: some of them are as follows.
2375: In one view, 
2376: the effective interaction between the particles is dominated 
2377: by the exchange of an AF spin fluctuation, and 
2378: this leads to $d_{x^2-y^2}$-wave pairing
2379: [Bickers {\it et al.} 1987].
2380: In an alternative view, 
2381: the magnetic correlations are responsible for the pairing,
2382: but it is because of the "bag effect" which leads to pairing
2383: with the extended $s$ or the $d_{x^2-y^2}$-wave symmetry
2384: [Schrieffer {\it et al.} 1988 and 1989].
2385: In this approach, the effective pairing interaction is
2386: qualitatively different than the single spin-fluctuation
2387: exchange interaction.
2388: Another view is that in the ground state of the 
2389: doped 2D Hubbard model there are no single-particle 
2390: excitations carrying both charge and spin and 
2391: the one-layer Hubbard model does not exhibit superconductivity
2392: [Anderson 1987, Anderson {\it et al.} 1987, Anderson and Zou 1988].
2393: In this view, 
2394: the pairing is mediated by interlayer hopping.
2395: In order to differentiate among these theories, 
2396: QMC simulations were carried out for the 2D Hubbard model.
2397: In this section, these QMC calculations will be reviewed.
2398: 
2399: In Section~6.1, the QMC results on the pair-field susceptibilities
2400: will be reviewed. 
2401: These calculations have found that there is an effective 
2402: attractive interaction in the singlet $d_{x^2-y^2}$-wave channel
2403: [White {\it et al.} 1989a].
2404: In addition, it is found that there is an 
2405: attractive interaction in the extended $s$-wave channel.
2406: However, at the temperatures where the simulations are carried out,
2407: the pairing correlations are short range and do not show scaling with 
2408: the lattice size
2409: [Moreo and Scalapino 1991, Moreo 1992].
2410: QMC simulations were also carried out for the pair-field 
2411: susceptibilities in the three-band CuO$_2$ model
2412: and similar results were found
2413: [Scalettar {\it et al.} 1991, Dopf {\it et al.} 1992a].
2414: 
2415: Using QMC simulations, the irreducible particle-particle 
2416: interaction $\Gamma_I$ was calculated
2417: [Bulut {\it et al.} 1993, 1994b, 1995] 
2418: and these results will be reviewed in Section~6.2.
2419: In these calculations, it is found that the momentum and the 
2420: Matsubara-frequency structure in $\Gamma_I$ follows closely 
2421: that of the magnetic susceptibility, 
2422: which means that the AF spin fluctuations dominate the 
2423: effective interaction in the parameter regime 
2424: where the simulations are carried out.
2425: In Section~6.3, the Bethe-Salpeter equation in the 
2426: particle-particle channel will be solved using the 
2427: QMC data on $\Gamma_I$ and the single-particle Green's function $G$.
2428: The solution of the Bethe-Salpeter equation makes 
2429: it possible to determine the strength of the various pairing channels
2430: quantitatively.
2431: Finite-size scaling of these results shows that as $T$ is lowered, 
2432: the fastest growing pairing instability 
2433: is in the singlet $d_{x^2-y^2}$-wave channel.
2434: Here, it is also found that as $U/t$ increases from 4 to 8, 
2435: the strength of the $d_{x^2-y^2}$ pairing correlations grows.
2436: Later in Section~8.1, these calculations will be compared with the 
2437: results of the FLEX approximation.
2438: In Section~6.4, the results on $\Gamma_I$ will be compared with the
2439: single spin-fluctuation exchange approximation.
2440: In Section~6.5, comparisons will be made with the perturbation theory
2441: results for $\Gamma_I$, which are third order in $U$.
2442: These comparisons are useful for gaining insight into the 
2443: effects of the various subgroups of many-body scattering processes 
2444: contributing to $\Gamma_I$.
2445: 
2446: \subsection{Pair-field susceptibilities}
2447: 
2448: In order to study the pairing correlations in the 2D Hubbard model,
2449: the pair-field susceptibilities defined by 
2450: \begin{equation}
2451: \label{P}
2452: P_{\alpha} = \int_0^{\beta} \, d\tau \,
2453: {1\over N} \sum_{\ell} \,
2454: \langle \Delta_{\alpha}(\ell,\tau) 
2455: \Delta^{\dagger}_{\alpha}(0,0) \rangle,
2456: \end{equation}
2457: where
2458: $\Delta^{\dagger}_{\alpha}$ is a pair creation 
2459: operator with $\alpha$ symmetry,
2460: were calculated.
2461: The QMC calculations of the pair-field susceptibilities 
2462: for the Hubbard model were first carried out 
2463: in Ref.~[Hirsch 1985].
2464: In these calculations, it was found that there is an 
2465: attractive effective interaction between two electrons 
2466: with antiparallel spins when they are separated by one lattice 
2467: spacing.
2468: In this reference, 
2469: the pair-field susceptibilities in the singlet $s$ and 
2470: extended $s$-wave ($s^*$) channels and the triplet channels were 
2471: calculated, however the singlet $d_{x^2-y^2}$-wave 
2472: channel was not considered.
2473: Later in Ref.~[White {\it et al.} 1989a], 
2474: various pair-field susceptibilities including the 
2475: one with the singlet $d_{x^2-y^2}$-wave symmetry were 
2476: calculated with QMC.
2477: In this section, results from Ref.~[White {\it et al.} 1989a]
2478: will be shown for $P_{\alpha}$ which were calculated using the 
2479: following pair-field operators in the 
2480: $d_{x^2-y^2}$, $s$, extended $s$ ($s^*$) and 
2481: $p$-wave channels,
2482: \begin{eqnarray}
2483: \Delta^{\dagger}_d && = {1\over 4} \sum_{\ell=1}^4 \,
2484: (-1)^{\ell} 
2485: c^{\dagger}_{i+\ell \uparrow}
2486: c^{\dagger}_{i \downarrow}, \\
2487: \Delta^{\dagger}_s && = 
2488: c^{\dagger}_{i \uparrow}
2489: c^{\dagger}_{i \downarrow}, \\
2490: \Delta^{\dagger}_{s^*} && = {1\over 4} \sum_{\ell=1}^4 \,
2491: c^{\dagger}_{i+\ell \uparrow}
2492: c^{\dagger}_{i \downarrow}, \\
2493: \Delta^{\dagger}_{p_x} && = {1\over 2} 
2494: (c^{\dagger}_{i+{\bf x} \uparrow}
2495: c^{\dagger}_{i \downarrow} 
2496: - c^{\dagger}_{i-{\bf x} \uparrow}
2497: c^{\dagger}_{i \downarrow}).
2498: \end{eqnarray}
2499: Here $\ell$ sums over the four neighbours
2500: of site $i$. 
2501: 
2502: \begin{figure}
2503: \centering
2504: \iffigure
2505: \epsfig{file=ch6-fig1.ps,height=3.5cm}
2506: \fi
2507: \caption{
2508: Feynman diagrams illustrating the pair-field
2509: susceptibilities 
2510: (a) $P_{\alpha}$ and
2511: (b) $\overline{P}_{\alpha}$.
2512: Here, 
2513: $\Gamma$ represents the reducible particle-particle interaction.
2514: }
2515: \label{6.1}
2516: \end{figure}
2517: 
2518: In order to see whether a given pairing channel 
2519: with the symmetry $\alpha$ is attractive, 
2520: $P_{\alpha}$ was compared with 
2521: $\overline{P}_{\alpha}$, 
2522: where $\overline{P}_{\alpha}$ is the 
2523: component of $P_{\alpha}$ which does not include
2524: the reducible particle-particle interaction $\Gamma$
2525: as illustrated in Fig.~6.1.
2526: Figure 6.2 compares $P_{\alpha}$ and $\overline{P}_{\alpha}$ 
2527: as a function of $T/t$ for $U=4t$ and $\langle n\rangle=0.87$
2528: on an $8\times 8$ lattice.
2529: Here, it is seen that $P_{\alpha}$ is enhanced 
2530: with respect to $\overline{P}_{\alpha}$ by the particle-particle
2531: reducible vertex for the $d_{x^2-y^2}$-wave and weakly for the extended
2532: $s$-wave symmetries.
2533: On the other hand, in the $s$- and $p$-wave 
2534: channels the pair-field susceptibility is suppressed
2535: with respect to $\overline{P}_{\alpha}$.
2536: This means that the effective particle-particle interaction 
2537: is attractive in the $d_{x^2-y^2}$ and $s^*$  channels,
2538: and it is repulsive in the $s$ and $p$-wave channels.
2539: However, at the lowest temperatures where 
2540: the QMC simulations can be carried out, 
2541: the $d_{x^2-y^2}$-wave pairing correlations are only short range:
2542: $P_d$ does not grow as the system size increases.  
2543: Furthermore, 
2544: while $P_d$ is enhanced with respect to $\overline{P}_d$,
2545: it is suppressed with respect to $P_d^0$, 
2546: the $d_{x^2-y^2}$-wave pair-field susceptibility of the 
2547: $U=0$ system, as seen in Fig.~6.2(a).
2548: So, at these temperatures, the $d_{x^2-y^2}$-wave  
2549: pair-field susceptibility gets suppressed when the Coulomb 
2550: repulsion is turned on,
2551: which is not encouraging for $d_{x^2-y^2}$-wave 
2552: superconductivity in the 2D Hubbard model.
2553: 
2554: \begin{figure}
2555: \centering
2556: \iffigure
2557: \mbox{
2558: \subfigure{
2559: \epsfysize=8cm
2560: \epsffile[100 150 480 610]{ch6-fig2a.ps}}
2561: \quad
2562: \subfigure{
2563: \epsfysize=8cm
2564: \epsffile[50 150 600 610]{ch6-fig2b.ps}}}
2565: \fi
2566: \centering
2567: \iffigure
2568: \mbox{
2569: \subfigure{
2570: \epsfysize=8cm
2571: \epsffile[100 150 480 610]{ch6-fig2c.ps}}
2572: \quad
2573: \subfigure{
2574: \epsfysize=8cm
2575: \epsffile[50 150 600 610]{ch6-fig2d.ps}}}
2576: \fi
2577: \caption{
2578: Pair-field susceptibilities $P_{\alpha}$ and
2579: $\overline{P}_{\alpha}$ versus $T/t$ for 
2580: the $d_{x^2-y^2}$, $s^*$, $s$ and $p$-wave 
2581: pairing channels.
2582: These results are for 
2583: $U=4t$ and $\langle n\rangle=0.87$ on an $8\times 8$ lattice.
2584: In (a), $P_0$ denotes the $d_{x^2-y^2}$-wave 
2585: pair-field susceptibility of the $U=0$ system.
2586: From [White {\it et al.} 1989a].
2587: }
2588: \label{6.2}
2589: \end{figure}
2590: 
2591: \subsection{Irreducible particle-particle interaction}
2592: 
2593: In this section, the QMC data on the 
2594: irreducible particle-particle vertex $\Gamma_I$
2595: from Ref.~[Bulut {\it et al.} 1993, 1994b, 1995] 
2596: will be reviewed.
2597: Using QMC simulations it is possible 
2598: to calculate the two-particle Green's function 
2599: \begin{equation}
2600: \Lambda(x_4,x_3|x_2,x_1) = - \langle T \,
2601: c_{\uparrow}(x_4) c_{\downarrow}(x_3)
2602: c^{\dagger}_{\downarrow}(x_2) c^{\dagger}_{\uparrow}(x_1)
2603: \rangle,
2604: \end{equation}
2605: where $c^{\dagger}_{\sigma}(x_i)$ with $x_i=({\bf x}_i,\tau_i)$ 
2606: creates an electron with spin $\sigma$ at site ${\bf x}_i$
2607: and imaginary time $\tau_i$.
2608: By Fourier transforming on both the space and the 
2609: imaginary-time variables one obtains
2610: \begin{equation}
2611: \Lambda(p',k'|p,k) = -\delta_{pp'} \delta_{kk'} 
2612: G_{\uparrow}(p) G_{\downarrow}(k) + 
2613: {T\over N} \delta_{k',p+k-p'}
2614: G_{\uparrow}(p') G_{\downarrow}(k') 
2615: \Gamma(p',k'|p,k) 
2616: G_{\uparrow}(p) G_{\downarrow}(k),
2617: \end{equation}
2618: where $p=({\bf p},i\omega_n)$,
2619: $G_{\sigma}(p)$ is the single-particle Green's function, and 
2620: $\Gamma(p',k'|p,k)$ is the reducible particle-particle
2621: vertex.
2622: Hence, using the QMC data on $\Lambda$ and $G$,
2623: $\Gamma$ can be calculated.
2624: This equation is illustrated in Fig.~6.3 in terms of 
2625: the Feynman diagrams.
2626: Here, the particle-particle interaction will be studied in the 
2627: zero center-of-mass momentum and energy channel, 
2628: and hence $k$ and $k'$ will be set to $-p$ and $-p'$, respectively.
2629: At the next stage, 
2630: the irreducible particle-particle vertex 
2631: $\Gamma_I$ is obtained from 
2632: the Monte Carlo results on $\Gamma$ and $G$ 
2633: by solving the particle-particle $t$-matrix equation,
2634: \begin{equation}
2635: \Gamma_I(p'|p) = \Gamma(p'|p) 
2636: + {T\over N} \sum_k \,
2637: \Gamma_I(p'|k) G_{\uparrow}(k) G_{\downarrow}(-k)
2638: \Gamma(k|p), 
2639: \end{equation}
2640: which is illustrated in Fig.~6.4.
2641: In Eq.~(6.8), $\Gamma(p'|p)$ is used as a short notation for 
2642: $\Gamma(p',-p'|p,-p)$.
2643: This procedure for calculating $\Gamma_I$ is essentially the 
2644: opposite of the usual diagrammatic approach in which $\Gamma_I$ 
2645: is used to solve for $\Gamma$.
2646: In solving the $t$-matrix equation, 
2647: an upper frequency cut-off of order the bandwidth 
2648: is used.
2649: Finally, 
2650: the singlet component of the irreducible vertex 
2651: is obtained from 
2652: \begin{equation} 
2653: \Gamma_{Is}(p'|p) = {1\over 2} 
2654: \bigg[ \Gamma_I(p'|p) + \Gamma_I(-p'|p) \bigg].
2655: \end{equation}
2656: In the following, the QMC data on $\Gamma_{Is}$ and 
2657: $\Gamma_s$ will be plotted 
2658: in units of $t$ as a function of the momentum 
2659: transfer ${\bf q}={\bf p'}-{\bf p}$
2660: and of the Matsubara-frequency transfer
2661: $\omega_m=\omega_{n'}-\omega_n$.
2662: Here, ${\bf p}$ will be kept fixed at $(\pi,0)$ 
2663: and $\omega_n$ at $\pi T$.
2664: 
2665: \begin{figure}
2666: \centering
2667: \iffigure
2668: \epsfig{file=ch6-fig3.ps,height=1.5cm}
2669: \fi
2670: \caption{
2671: Feynman diagrams illustrating the 
2672: correlation function 
2673: $\Lambda(p',k'|p,k)$
2674: in terms of the single-particle Green's functions $G$ and 
2675: the reducible particle-particle interaction $\Gamma$.
2676: }
2677: \label{6.3}
2678: \end{figure}
2679: 
2680: \begin{figure}
2681: \centering
2682: \iffigure
2683: \epsfig{file=ch6-fig4.ps,height=8cm,angle=+90}
2684: \fi
2685: \caption{
2686: Feynman diagrams illustrating the $t$-matrix equation 
2687: which relates the irreducible particle-particle interaction
2688: $\Gamma_I$ to the reducible particle-particle interaction
2689: $\Gamma$. 
2690: }
2691: \label{6.4}
2692: \end{figure}
2693: 
2694: It is useful to start the discussion by first showing 
2695: results on the singlet reducible vertex $\Gamma_s$.
2696: In Fig.~6.5, 
2697: $\Gamma_s({\bf q},i\omega_m=0)$ 
2698: versus ${\bf q}$ is shown for $U=4t$ and $8t$.
2699: As the temperature is lowered, 
2700: the ${\bf q}\sim (\pi,\pi)$ component of 
2701: $\Gamma_s$ increases,
2702: and the ${\bf q}\sim 0$ component is suppressed.
2703: Here, it is seen that $\Gamma_s$ becomes quite large especially 
2704: for $U=8t$.
2705: In Fig.~6.5(b), the ${\bf q}=(0,0)$ point is not shown 
2706: because of large error bars
2707: for this point due to the way $\Gamma$ is calculated 
2708: from $\Lambda$.
2709: It is also not possible to show results at lower temperatures
2710: since the error bars grow rapidly.
2711: For instance, at $T=0.33t$ and $U=8t$, 
2712: $\Gamma_s({\bf q}=(\pi,\pi),0)$ was calculated to be 
2713: $60t\pm 20t$ after long simulation times,
2714: while $\Gamma_s({\bf q}=(\pi/4,\pi/4),0)$
2715: was $10t\pm 10t$. 
2716: 
2717: \begin{figure}
2718: \centering
2719: \iffigure
2720: \mbox{
2721: \subfigure[]{
2722: \epsfysize=8cm
2723: \epsffile[100 150 480 610]{ch6-fig5a.ps}}
2724: \quad
2725: \subfigure[]{
2726: \epsfysize=8cm
2727: \epsffile[50 150 600 610]{ch6-fig5b.ps}}}
2728: \fi
2729: \caption{
2730: Reducible particle-particle scattering vertex in the singlet channel
2731: $\Gamma_s({\bf q},i\omega_m=0)$ versus ${\bf q}$
2732: for (a) $U=4t$ and (b) $U=8t$.
2733: Here, ${\bf q}={\bf p'}-{\bf p}$ and ${\bf p}$ is kept fixed at 
2734: $(\pi,0)$.
2735: These results were obtained 
2736: at $\langle n\rangle=0.87$ on an $8\times 8$ lattice.
2737: For $U=8t$, 
2738: $\Gamma_s({\bf q},0)$ at ${\bf q}=(0,0)$ is not shown 
2739: because of large error bars in the data in this case.
2740: }
2741: \label{6.5}
2742: \end{figure}
2743: 
2744: Next, in Fig.~6.6(a) and (b), the momentum 
2745: and the Matsubara-frequency 
2746: dependence of the irreducible particle-particle vertex
2747: $\Gamma_{Is}({\bf q},i\omega_m)$ is shown for $U=4t$.
2748: At ${\bf q}=(\pi,\pi)$ momentum transfer, 
2749: $\Gamma_{Is}({\bf q},0)$ reaches values larger than twice the 
2750: bandwidth. 
2751: It is useful to compare Figs.~6.6(a) and 6.5(a)
2752: in order to understand the effect of the repeated 
2753: particle-particle scatterings on $\Gamma_{Is}$.
2754: At these temperatures, the effect of these scatterings is 
2755: basically to suppress the momentum and frequency-independent 
2756: background in $\Gamma_{Is}$.
2757: For instance, at $T=0.25t$ the difference in the magnitude of 
2758: $\Gamma_{Is}({\bf q},0)$ between 
2759: ${\bf q}=(\pi,\pi)$ and $(\pi/4,\pi/4)$ is about $10t$,
2760: which is the same as in $\Gamma_s({\bf q},0)$.
2761: This is similar to the suppression of the screened Coulomb
2762: repulsion, which varies more slowly in frequency 
2763: compared to the phonon propagator, 
2764: in the usual phonon-mediated pairing.
2765: As a $d_{x^2-y^2}$-wave superconducting instability 
2766: is approached, 
2767: the expected behaviour for an infinite system is that 
2768: $\Gamma_s({\bf q}=(\pi,\pi),0) \rightarrow +\infty$
2769: while $\Gamma_s({\bf q}=(0,0),0) \rightarrow -\infty$.
2770: These QMC data show that such resonant scattering in the 
2771: $d_{x^2-y^2}$-wave channel is not taking place 
2772: at these temperatures.
2773: 
2774: \begin{figure}
2775: \centering
2776: \iffigure
2777: \mbox{
2778: \subfigure{
2779: \epsfysize=8cm
2780: \epsffile[100 150 480 610]{ch6-fig6a.ps}}
2781: \quad
2782: \subfigure{
2783: \epsfysize=8cm
2784: \epsffile[50 150 600 610]{ch6-fig6b.ps}}}
2785: \fi
2786: \caption{
2787: (a) Momentum and (b) the Matsubara-frequency dependence of the
2788: irreducible particle-particle scattering vertex in the singlet channel
2789: $\Gamma_{Is}({\bf q},i\omega_m)$. 
2790: In (a), $\Gamma_{Is}$ versus ${\bf q}={\bf p'}-{\bf p}$ 
2791: is plotted where ${\bf p}$ is kept fixed at $(\pi,0)$.
2792: In (b), $\Gamma_{Is}$ versus $\omega_m$ 
2793: is plotted for ${\bf q}=(\pi,\pi)$.
2794: Here, $\omega_m=\omega_{n'}-\omega_n$ and 
2795: $\omega_n$ is kept fixed at $\pi T$.
2796: These results are for $U=4t$ and $\langle n\rangle=0.87$
2797: on an $8\times 8$ lattice.
2798: }
2799: \label{6.6}
2800: \end{figure}
2801: 
2802: It is desirable to know how $\Gamma_{Is}$ 
2803: varies as $U/t$ increases.
2804: However, it has not been possible to obtain $\Gamma_{Is}$ for $U=8t$
2805: from the $t$-matrix equation because 
2806: of the larger error bars for this case.
2807: Nevertheless, in Fig.~6.5(b) it is seen that  
2808: $\Gamma_s({\bf q},0)$ for $U=8t$ and $T=0.5t$ exhibits 
2809: large variation of order $20t$ 
2810: between points $(\pi/4,\pi/4)$ and $(\pi,\pi)$.
2811: If at these temperatures the effect of the $t$-matrix 
2812: scattering is only to suppress the 
2813: ${\bf q}$ and $\omega_m$ independent background in 
2814: $\Gamma_{Is}$, then this means that $\Gamma_{Is}$ grows
2815: considerably as $U/t$ increases from 4 to 8.
2816: 
2817: \begin{figure}
2818: \centering
2819: \iffigure
2820: \mbox{
2821: \subfigure[]{
2822: \epsfysize=8cm
2823: \epsffile[100 150 480 610]{ch6-fig7a.ps}}
2824: \quad
2825: \subfigure[]{
2826: \epsfysize=8cm
2827: \epsffile[50 150 600 610]{ch6-fig7b.ps}}}
2828: \fi
2829: \caption{
2830: (a) Momentum and (b) the Matsubara-frequency dependence of 
2831: the magnetic susceptibility $\chi({\bf q},i\omega_m)$
2832: at various temperatures.
2833: In (a), $\chi({\bf q},i\omega_m)$ versus ${\bf q}$ is plotted
2834: for $\omega_m=0$.
2835: In (b), $\chi({\bf q}=(\pi,\pi),i\omega_m)$ versus $\omega_m$ 
2836: is plotted.
2837: These results are for $U=4t$ and $\langle n\rangle=0.87$
2838: on an $8\times 8$ lattice.
2839: }
2840: \label{6.7}
2841: \end{figure}
2842: 
2843: Next,
2844: the temperature evolution of 
2845: $\Gamma_{Is}({\bf q},i\omega_m)$ 
2846: for $U=4t$ is compared with 
2847: that of the magnetic susceptibility 
2848: $\chi({\bf q},i\omega_m)$.
2849: Figure 6.7(a) shows Monte Carlo data on $\chi({\bf q},0)$ 
2850: versus ${\bf q}$ at the same temperatures as in Fig.~6.6.
2851: The Matsubara frequency dependence of 
2852: $\chi({\bf q}=(\pi,\pi),i\omega_m)$ is shown in Fig.~6.7(b).
2853: Comparing Figs.~6.6 and 6.7, one sees that the temperature 
2854: evolution of $\Gamma_{Is}({\bf q},i\omega_m)$ 
2855: closely follows that of $\chi({\bf q},i\omega_m)$.
2856: Both of these quantities peak at ${\bf q}=(\pi,\pi)$, 
2857: and as $\omega_m$ increases $\Gamma_{Is}$ goes to 
2858: the bare $U$ value while $\chi$ decays to zero.
2859: The relation between $\Gamma_{Is}$ and the spin fluctuations
2860: will be studied in more detail in Section~6.4, 
2861: where the QMC results on $\Gamma_{Is}$ will be compared 
2862: with the single spin-fluctuation exchange interaction.
2863: 
2864: \begin{figure}
2865: \centering
2866: \iffigure
2867: \epsfysize=8cm
2868: \epsffile[100 200 550 560]{ch6-fig8.ps}
2869: \fi
2870: \caption{
2871: Real-space pattern of $\Gamma_{Is}({\bf R})$ for 
2872: $U=4t$, $T=0.25t$ and $\langle n\rangle=0.87$.
2873: Here, it is seen that for ${\bf R}=0$, 
2874: $\Gamma_{Is}({\bf R})$ is strongly repulsive while 
2875: being attractive when the singlet electron pair is 
2876: separated by one lattice spacing.
2877: }
2878: \label{6.8}
2879: \end{figure}
2880: 
2881: At this point, it should be noted that 
2882: the ${\bf q}$ dependence of the effective pairing 
2883: interaction for the cuprates was also studied 
2884: within the spin-bag approach where a dip at ${\bf q}\sim 0$
2885: is found in addition to the peak at ${\bf q}\sim (\pi,\pi)$
2886: [Kampf and Schrieffer 1990a].
2887: The dip is due to the cross-exchange of two AF spin 
2888: fluctuations and it is responsible for the effective attraction 
2889: between two spin bags.
2890: At $T=0.25t$, 
2891: which is the lowest temperature where the 
2892: QMC calculation of $\Gamma_{Is}$ can be carried out, 
2893: a dip in $\Gamma_{Is}$ at ${\bf q}\sim 0$ is not found.
2894: 
2895: In order to gain insight into the 
2896: real-space structure of the effective
2897: particle-particle interaction, 
2898: it is useful to consider the Fourier transform
2899: \begin{equation}
2900: \Gamma_{Is}({\bf R})=
2901: {1\over N^2} \, \sum_{{\bf p},{\bf p'}} \,
2902: e^{ i({\bf p}-{\bf p'})\cdot {\bf R}} \,
2903: \Gamma_{Is}({\bf p'},i\omega_{n'} | {\bf p},i\omega_n)
2904: \end{equation}
2905: for the lowest Matsubara frequencies 
2906: $\omega_n=\omega_{n'}=\pi T$.
2907: Figure~6.8 shows the Monte Carlo data on 
2908: $\Gamma_{Is}({\bf R})$ as a function of ${\bf R}$ 
2909: for $T=0.25t$.
2910: At ${\bf R}=0$,  
2911: $\Gamma_{Is}$ is strongly repulsive, as expected, 
2912: but for a singlet electron pair separated by one lattice
2913: spacing, $\Gamma_{Is}$ is attractive. 
2914: As the pair separation increases further,
2915: $\Gamma_{Is}$ oscillates in sign and its magnitude 
2916: decreases rapidly, reflecting the short-range nature of the interaction.
2917: Pairing correlations with the proper space-time structure 
2918: can avoid the large onsite repulsion while taking 
2919: advantage of the near-neighbour attraction. 
2920: Thus, the interaction $\Gamma_{Is}$ is attractive 
2921: in the $d_{x^2-y^2}$-wave channel.
2922: 
2923: In Fig.~6.8, it is seen that the effective 
2924: attractive interaction is $-0.5t$ at the nearest-neighbour site. 
2925: Here, one might argue that the long-range Coulomb 
2926: repulsion between the electrons, 
2927: which is not taken into account in the Hubbard model, 
2928: could overcome this attraction. 
2929: However, 
2930: one would expect
2931: the long-range Coulomb repulsion to have 
2932: weak $\omega_m$ dependence
2933: compared to that seen in Fig.~6.6(b), 
2934: and after the $t$-matrix scatterings 
2935: it should have weak influence on the strength of the 
2936: pairing in the $d_{x^2-y^2}$-wave channel.
2937: 
2938: \subsection{Bethe-Salpeter equation in the particle-particle channel}
2939: 
2940: In this section,  
2941: the particle-particle Bethe-Salpeter equation
2942: will be solved
2943: using the QMC data on the irreducible interaction $\Gamma_I$ 
2944: and the one-electron Green's function.
2945: This way one can determine the magnitude of the eigenvalues 
2946: and the $({\bf p},i\omega_n)$ structure of the leading pair-field
2947: eigenfunctions. 
2948: This approach is useful for studying the leading
2949: scattering channels in the $t$-matrix quantitatively.
2950: For instance, one 
2951: can determine how close 
2952: the system is to a 
2953: Kosterlitz-Thouless superconducting transition,
2954: and compare the strength of the pairing  
2955: in various channels.
2956: The particle-particle Bethe-Salpeter equation is 
2957: \begin{equation}
2958: \label{BS}
2959: \lambda_{\alpha} \phi_{\alpha}(p) = - {T\over N} 
2960: \sum_{p'} \,
2961: \Gamma_I(p|p') G_{\uparrow}(p') G_{\downarrow}(-p')
2962: \phi_{\alpha}(p'),
2963: \end{equation}
2964: where $\phi_{\alpha}(p)$ is the pair-field eigenfunction and 
2965: $\lambda_{\alpha}$ is the corresponding eigenvalue.
2966: The Feynman diagram representing Eq.~(\ref{BS}) is 
2967: shown in Fig.~6.9.
2968: When the leading eigenvalue reaches one,
2969: the superconducting transition takes place,
2970: and the Bethe-Salpeter equation for the corresponding 
2971: eigenfunction becomes equivalent to the superconducting gap 
2972: equation.
2973: 
2974: \begin{figure}[ht]
2975: \centering
2976: \iffigure
2977: \epsfig{file=ch6-fig9.ps,height=1.5cm}
2978: \fi
2979: \caption{
2980: Feynman diagram representing the Bethe-Salpeter equation.
2981: }
2982: \label{6.9}
2983: \end{figure}
2984: 
2985: Here, data from Refs.~[Bulut {\it et al.} 1993]
2986: will be shown.
2987: Both the triplet and the singlet solutions of the 
2988: Bethe-Salpeter equation will be considered.
2989: The momentum and the Matsubara-frequency structures of the 
2990: leading eigenfunctions will be presented. 
2991: In addition to the singlet $d_{x^2-y^2}$-wave eigenfunction 
2992: $\phi_d({\bf p},i\omega)$, 
2993: solutions which are odd in $\omega_n$ and have $s$ and $p$-wave
2994: symmetries are found. 
2995: The odd-frequency eigenfunctions with $s$-wave and 
2996: $p$-wave symmetries correspond to 
2997: triplet and singlet solutions, respectively.
2998: The finite-size scaling of the eigenvalues indicates
2999: that as $T$ is lowered the singlet $d_{x^2-y^2}$-wave 
3000: eigenvalue grows fastest,
3001: and at low $T$ the dominant singlet pairing channel has 
3002: the $d_{x^2-y^2}$-wave symmetry.
3003: In these calculations, the even-frequency extended $s$-wave channel 
3004: was not found to be one of the leading pairing channels.
3005: 
3006: The case of the odd-frequency pairing channel is interesting.
3007: The possibility of pairing in the triplet odd-frequency $s$-wave 
3008: channel was first studied within the context of $^3$He in 
3009: Ref.~[Berezinskii 1974].
3010: The singlet odd-frequency $p$-wave channel was studied 
3011: initially by using an effective interaction which is
3012: mediated by phonons [Balatsky and Abrahams 1992].
3013: The possibility that an effective attraction in 
3014: the singlet odd-frequency $p$-wave channel 
3015: could be generated by the spin fluctuations 
3016: was noted in Ref.~[Abrahams {\it et al.} 1993] 
3017: at about the same time the QMC calculations reviewed here 
3018: were carried out.
3019: The general properties of the odd-frequency superconductors 
3020: were discussed in Ref.~[Abrahams {\it et al.} 1995].
3021: 
3022: In addition, here 
3023: the effects of increasing $U/t$ 
3024: on the leading pairing channels will be discussed.
3025: As seen in the previous section, it was not possible to calculate
3026: $\Gamma_{Is}$ for $U=8t$.
3027: However, if $\Gamma$ is used rather than $\Gamma_I$ 
3028: in Eq.~(\ref{BS}), then the corresponding eigenvalues are given 
3029: by $\lambda_{\alpha}/(1-\lambda_{\alpha})$,
3030: where $\lambda_{\alpha}$ are the eigenvalues for the
3031: irreducible vertex.
3032: Hence, through this indirect way it is possible to study
3033: the leading Bethe-Salpeter eigenvalues for $U=8t$.
3034: It will be shown that as $U/t$ increases from 4 to 8,
3035: the leading eigenvalues grow, 
3036: since the effective particle-particle interaction gets stronger,
3037: but the momentum and the frequency structure of their
3038: eigenfunctions does not exhibit qualitative changes.
3039: 
3040: In general, the Bethe-Salpeter equation can have both singlet and
3041: triplet solutions corresponding to a pair-wave function that has 
3042: overall even or odd parity when 
3043: $p=({\bf p},i\omega_n)$ goes to $(-{\bf p},-i\omega_n)$.
3044: Here, the pair wave functions are characterized by 
3045: its symmetry in momentum and spin space. 
3046: The usual singlet $s$ and $d_{x^2-y^2}$-wave states are even in 
3047: frequency and even in momentum,
3048: $\phi({\bf p},-i\omega_n)=\phi({\bf p},i\omega_n)$ and 
3049: $\phi(-{\bf p},i\omega_n)=\phi({\bf p},i\omega_n)$,
3050: while the usual triplet $p_x$ or ($p_y$) state is even in frequency 
3051: and odd when $p_x$ goes to $-p_x$.
3052: There are also odd-frequency pair-wave functions. 
3053: In this case, one can have an odd-frequency $s$-wave 
3054: triplet for which 
3055: $\phi({\bf p},-i\omega_n)=-\phi({\bf p},i\omega_n)$ and 
3056: $\phi(-{\bf p},i\omega_n)=\phi({\bf p},i\omega_n)$, or an 
3057: odd-frequency $p_x$ (or $p_y$)-wave singlet with 
3058: $\phi({\bf p},-i\omega_n)=-\phi({\bf p},i\omega_n)$ and 
3059: $\phi(-{\bf p},i\omega_n)=-\phi({\bf p},i\omega_n)$. 
3060: In the regime of the Hubbard model that is being studied,
3061: the $s$-wave triplet, and the $p$ and $d_{x^2-y^2}$-wave 
3062: singlet solutions are dominant.
3063: 
3064: The temperature evolution of the 
3065: four largest eigenvalues for $U=4t$ and $\langle n\rangle=0.87$
3066: are given in Table~1.
3067: The momentum and the frequency dependence of the corresponding 
3068: pair-wave functions are shown in Figs.~6.10 and 6.11
3069: for $T=0.5t$.
3070: At this temperature, 
3071: an $s$-wave triplet state has the largest eigenvalue
3072: $\lambda_s \approx 0.23$.
3073: As seen in Figs.~6.10(a) and 6.11(a) (solid circles), 
3074: the pair-wave function $\phi_s({\bf p},i\omega_n)$ of the
3075: $s$-wave triplet state is even in ${\bf p}$ and odd in $\omega_n$. 
3076: The open circles in Figs.~6.10(a) and 6.11(a) represent the pair-wave 
3077: function $\phi_{s'}({\bf p},i\omega_n)$ which has the second 
3078: largest eigenvalue. 
3079: This is also an odd-frequency $s$-wave triplet state.
3080: The $d_{x^2-y^2}$-wave singlet state shown as the solid 
3081: circles in Figs.~6.10(b) and 6.11(b) has the third largest eigenvalue.
3082: The fourth largest eigenvalue corresponds to a state which is 
3083: odd in both $\omega_n$ and ${\bf p}$, having $p_y$ (or $p_x$)
3084: symmetry [open circles in Figs.~6.10(b) and 6.11(b)], 
3085: hence it is also a singlet. 
3086: 
3087: \begin{table}[ht]
3088: \centering
3089: \begin{tabular} {| c | c | c | c | c |} \hline
3090: $T/t$  & $\lambda_s$ & $\lambda_{s'}$ & $\lambda_p$ & $\lambda_d$ \\ \hline 
3091: $1.0$  & $0.261$     &  0.037        & $0.023$     & $0.022$ \\ 
3092: $0.50$ & $0.228$     &  0.090         & $0.066$     & $0.076$ \\ 
3093: $0.33$ & $0.251$     &  0.104         & $0.095$     & $0.130$ \\ 
3094: $0.25$ & $0.264\pm 0.007$ & $0.124\pm 0.006$ & $0.129\pm 0.007$ & 
3095: $0.182\pm 0.006$ \\ 
3096: \hline
3097: \end{tabular}
3098: \caption{Temperature dependence of the Bethe-Salpeter eigenvalues
3099: for $U=4t$ and $\langle n\rangle=0.87$ on an $8\times 8$ lattice.
3100: The error bars represent the uncertainty due to the 
3101: Monte Carlo sampling. 
3102: The error bars were not calculated at all temperatures, 
3103: since it requires considerably more computer time.
3104: When the error bars are not indicated, 
3105: they are estimated to be less than 10\%.
3106: }
3107: \end{table}
3108: 
3109: \begin{figure}
3110: \centering
3111: \iffigure
3112: \mbox{
3113: \subfigure{
3114: \epsfysize=8cm
3115: \epsffile[100 150 480 610]{ch6-fig10a.ps}}
3116: \quad
3117: \subfigure{
3118: \epsfysize=8cm
3119: \epsffile[50 150 600 610]{ch6-fig10b.ps}}}
3120: \fi
3121: \caption{
3122: Momentum dependence of the leading 
3123: Bethe-Salpeter eigenfunctions 
3124: $\phi_{\alpha}({\bf p},i\omega_n)$.
3125: These results are for $\omega_n=\pi T$,
3126: $U=4t$, $T=0.5t$ and $\langle n\rangle=0.87$
3127: on an $8\times 8$ lattice.
3128: In (a), two odd-frequency $s$-wave eigenfunctions are plotted, 
3129: and in (b) the $d_{x^2-y^2}$-wave (filled circles) and 
3130: $p_y$-wave (open circles) are plotted.
3131: }
3132: \label{6.10}
3133: \end{figure}
3134: 
3135: \begin{figure}
3136: \centering
3137: \iffigure
3138: \mbox{
3139: \subfigure{
3140: \epsfysize=8cm
3141: \epsffile[100 150 480 610]{ch6-fig11a.ps}}
3142: \quad
3143: \subfigure{
3144: \epsfysize=8cm
3145: \epsffile[50 150 600 610]{ch6-fig11b.ps}}}
3146: \fi
3147: \caption{
3148: Matsubara frequency dependence of the leading 
3149: Bethe-Salpeter eigenfunctions 
3150: $\phi_{\alpha}({\bf p},i\omega_n)$.
3151: These results are for
3152: $U=4t$, $T=0.5t$ and $\langle n\rangle=0.87$
3153: on an $8\times 8$ lattice.
3154: In (a), 
3155: $\phi({\bf p}=(\pi,0),i\omega_n)$ 
3156: is shown for the odd-$\omega_n$ $s$ and $s'$ channels.
3157: In (b), 
3158: $\phi({\bf p},i\omega_n)$
3159: is shown for the singlet $d_{x^2-y^2}$-wave channel
3160: at ${\bf p}=(\pi,0)$ and for the 
3161: singlet $p_y$-wave channel at ${\bf p}=(\pi/2,\pi/2)$.
3162: }
3163: \label{6.11}
3164: \end{figure}
3165: 
3166: \begin{table}
3167: \centering
3168: \begin{tabular} {| c | c | c | c |} \hline
3169: $L\times L$ & $\lambda_s$      & $\lambda_p$      & $\lambda_d$ \\ \hline
3170: $4\times 4$ & $0.296\pm 0.010$ & $0.204\pm 0.014$ & $0.184\pm 0.016$ \\ 
3171: $8\times 8$ & $0.264\pm 0.007$ & $0.129\pm 0.007$ & $0.182\pm 0.006$ \\ 
3172: \hline
3173: \end{tabular}
3174: \caption{Finite-size dependence of the Bethe-Salpeter eigenvalues
3175: for  $T=0.25t$, $U=4t$ and $\langle n\rangle=0.87$.}
3176: \end{table}
3177: 
3178: \begin{figure}
3179: \centering
3180: \iffigure
3181: \epsfysize=8cm
3182: \epsffile[100 150 550 610]{ch6-fig12.ps}
3183: \fi
3184: \caption{
3185: Comparison of the $d_{x^2-y^2}$-wave eigenfunction 
3186: $\phi_d({\bf p},i\pi T)$ 
3187: with the usual $d_{x^2-y^2}$-wave form  
3188: $\Delta_{\bf p} = (\Delta_0/2)(\cos{p_x} + \cos{p_y})$
3189: for $U=4t$, $T=0.25t$ and $\langle n\rangle=0.87$.
3190: Here, both 
3191: $\phi_d({\bf p},i\pi T)$ and 
3192: $\Delta_{\bf p}$ have been normalised to 1 at
3193: ${\bf p}=(\pi,0)$.
3194: }
3195: \label{6.12}
3196: \end{figure}
3197: 
3198: As $T$ is lowered from $1.0t$ down to $0.25t$, 
3199: the largest eigenvalue $\lambda_s$ stays nearly the same, 
3200: while $\lambda_d$ increases 
3201: by about a factor of seven. 
3202: This can be understood in terms of the temperature 
3203: dependence of $\Gamma_I(p'|p)$ which enters the Bethe-Salpeter 
3204: equation. 
3205: The pair-wave functions which are smooth in ${\bf p}$
3206: but odd in $\omega_n$ make optimum use of the 
3207: $(\omega_n,\omega_{n'})$ frequency structure of the 
3208: repulsive $\Gamma_I(p'|p)$ for pairing.
3209: However, as the temperature is lowered 
3210: and $\Gamma_I(p'|p)$ for ${\bf p'}-{\bf p}=(\pi,\pi)$ grows,
3211: the $d_{x^2-y^2}$ and $p$-wave solutions can make better 
3212: use of the momentum structure in $\Gamma_I$, 
3213: and their eigenvalues get enhanced.
3214: 
3215: Table~2 shows the finite-size effects on the leading 
3216: eigenvalues at $T=0.25t$.
3217: Here, it is seen that the finite-size effects 
3218: are especially large for the $p$-wave channel, 
3219: and $\lambda_p$ decreases as the system size grows
3220: from $4\times 4$ to $8\times 8$.
3221: The finite size effects for the $d_{x^2-y^2}$-wave case
3222: are small.
3223: Hence, these results show that as $T$ is lowered, 
3224: $\lambda_d$ grows fastest and at low $T$ 
3225: the dominant singlet pairing channel has the 
3226: $d_{x^2-y^2}$-wave symmetry.
3227: 
3228: It is useful to compare the momentum dependence of the 
3229: $d_{x^2-y^2}$-wave eigenvalue 
3230: $\phi_d({\bf p},i\omega_n)$ at $\omega_n=\pi T$
3231: with the usual $d_{x^2-y^2}$-wave gap form
3232: $\Delta_{\bf p}=(\Delta_0/2)(\cos{p_x} - \cos{p_y})$,
3233: since this is often used in modelling the 
3234: superconducting state of the cuprates.
3235: Figure 6.12
3236: shows that the QMC data on $\phi_d({\bf p},i\pi T)$
3237: taken at $T=0.25t$
3238: follow the usual $d_{x^2-y^2}$-wave form closely. 
3239: 
3240: Finally, 
3241: the filled circles in Fig.~6.13 show the growth of 
3242: $\lambda_d$ for $U=4t$ as the temperature is lowered,
3243: while the open circles represent $\lambda_d$ for 
3244: $U=8t$.
3245: The results on $\lambda_d$ for $U=8t$ were obtained 
3246: as described above.
3247: Here, it is seen that $\lambda_d$ increases with $U/t$.
3248: 
3249: \begin{figure}
3250: \centering
3251: \iffigure
3252: \epsfysize=8cm
3253: \epsffile[100 150 550 610]{ch6-fig13.ps}
3254: \fi
3255: \caption{
3256: Temperature dependence of the $d_{x^2-y^2}$-wave eigenvalue
3257: $\lambda_d$ for $U=4t$ and $8t$
3258: at $\langle n\rangle=0.87$.
3259: }
3260: \label{6.13}
3261: \end{figure}
3262: 
3263: At the lowest temperature where $\lambda_d$ can be calculated,
3264: the system is far from a 
3265: Kosterlitz-Thouless superconducting transition
3266: which would be signalled by $\lambda_d\rightarrow 1$.
3267: Hence, while these data imply that at the temperatures
3268: where the QMC simulations are carried out, 
3269: the singlet $d_{x^2-y^2}$-wave pairing correlations are becoming dominant, 
3270: it is not known whether $\lambda_d\rightarrow 1$ 
3271: at lower temperatures.
3272: Below in Section~8.1, these results on $\lambda_d$ will be 
3273: compared with the results of the FLEX calculations.
3274: 
3275: \begin{figure}
3276: \centering
3277: \iffigure
3278: \mbox{
3279: \subfigure{
3280: \epsfysize=8cm
3281: \epsffile[100 150 480 610]{ch6-fig14a.ps}}
3282: \quad
3283: \subfigure{
3284: \epsfysize=8cm
3285: \epsffile[50 150 600 610]{ch6-fig14b.ps}}}
3286: \fi
3287: \caption{
3288: Eigenvalues of the 
3289: particle-hole and the particle-particle 
3290: Bethe-Salpeter equations versus $T$
3291: for (a) $U=4t$ and (b) $U=8t$ for an $8\times 8$ 
3292: half-filled lattice.
3293: The solid points ($\bullet$) are for the leading eigenvalue 
3294: $\overline{\lambda}_1$ of the Bethe-Salpeter 
3295: equation in the AF particle-hole channel
3296: with the center-of-mass momentum ${\bf Q}=(\pi,\pi)$.
3297: The open symbols denote the 
3298: even-frequency $d_{x^2-y^2}$-wave ($\circ$), and 
3299: the odd-frequency $p$ ($\triangle$) and $s$-wave ($\Box$)
3300: eigenvalues of the 
3301: particle-particle Bethe-Salpeter equation.
3302: }
3303: \label{6.14}
3304: \end{figure}
3305: 
3306: The QMC calculation of $\lambda_d$ for the doped case cannot be 
3307: carried out at any lower temperatures.
3308: However, for half-filling it is possible to calculate the various 
3309: pairing eigenvalues at low $T$
3310: and compare them with the eigenvalues of 
3311: the Bethe-Salpeter equation in the 
3312: AF particle-hole channel,
3313: $\overline{\lambda}_{\alpha}$.
3314: The leading magnetic eigenvalue $\overline{\lambda}_1$
3315: occurs in an even-frequency $s$-wave channel
3316: with center-of-mass momentum ${\bf Q}=(\pi,\pi)$.
3317: In Fig.~6.14, $\overline{\lambda}_1$ is
3318: compared with the various pairing eigenvalues
3319: for $U=4t$ and $8t$ at half-filling.
3320: As expected, at low temperatures $\overline{\lambda}_1$ 
3321: reaches 1 asymptotically, signalling the phase transition to the 
3322: AF order state on the $8\times 8$ lattice.
3323: Here, it is also seen that as the AF correlations develop
3324: at half-filling,
3325: $\lambda_d$ becomes the leading pairing eigenvalue
3326: while always staying smaller than the magnetic eigenvalue 
3327: $\overline{\lambda}_1$.
3328: 
3329: \subsection{Comparison with the spin-fluctuation exchange approximation}
3330: 
3331: Various spin-fluctuation exchange theories have been used 
3332: for studying $d_{x^2-y^2}$-wave pairing for the cuprates
3333: [Bickers {\it et al.} 1987 and 1989,
3334: Moriya {\it et al.} 1990, 
3335: Monthoux {\it et al.} 1991].
3336: In this context, it is of interest to see to 
3337: what extent the Monte Carlo results for the irreducible 
3338: vertex can be modelled by a single 
3339: spin-fluctuation exchange interaction. 
3340: For this purpose, here we compare the Monte Carlo data 
3341: with the approximate form
3342: \begin{equation}
3343: \label{GSF}
3344: \Gamma_I^{SF}({\bf q},i\omega_m) = 
3345: U + {3\over 2} g^2U^2 \, 
3346: \chi({\bf q},i\omega_{m}).
3347: \end{equation}
3348: This form is motivated by the 
3349: single spin-fluctuation exchange interaction
3350: [Berk and Schrieffer 1966, Doniach and Engelsberg 1966], 
3351: which basically has this form with $g=1$ 
3352: near the antiferromagnetic instability. 
3353: The Feynman diagrams illustrating the 
3354: single spin-fluctuation exchange 
3355: interaction were shown in Fig.~2.1.
3356: The factor of $3/2$ arises from the two transverse 
3357: and one longitudinal spin fluctuations. 
3358: In calculating $\Gamma_{Is}^{SF}$ with 
3359: Eq.~(\ref{GSF}), we will use Monte Carlo results for 
3360: $\chi({\bf q},i\omega_m)$ and also set $g=0.8$.
3361: The corresponding value of $3.2t$ for the effective coupling 
3362: $gU$ is consistent with the results of the Monte Carlo
3363: calculations of the effective irreducible 
3364: vertex in the particle-hole channel,
3365: $\overline{U}({\bf q},0)$, 
3366: which were discussed in Section~3.4.
3367: Formally, Eq.~(\ref{GSF}) is analogous to the effective 
3368: interaction in the electron-phonon superconductor
3369: \begin{equation}
3370: V({\bf q},i\omega_m) = U + 
3371: \sum_{\lambda} \, |g_{{\bf q}\lambda}|^2 \, 
3372: D_{\lambda}({\bf q},i\omega_m),
3373: \end{equation}
3374: where $D_{\lambda}({\bf q},i\omega_m)$ is the dressed 
3375: phonon propagator and $|g_{{\bf q}\lambda}|^2$ 
3376: is the renormalised electron-phonon coupling.
3377: 
3378: \begin{figure}
3379: \centering
3380: \iffigure
3381: \mbox{
3382: \subfigure{
3383: \epsfysize=8cm
3384: \epsffile[100 150 480 610]{ch6-fig15a.ps}}
3385: \quad
3386: \subfigure{
3387: \epsfysize=8cm
3388: \epsffile[50 150 600 610]{ch6-fig15b.ps}}}
3389: \fi
3390: \caption{
3391: Single-spin-fluctuation exchange approximation for 
3392: the irreducible particle-particle scattering vertex 
3393: in the singlet channel
3394: $\Gamma_{Is}({\bf q},i\omega_m)$. 
3395: In (a), $\Gamma_{Is}$ versus ${\bf q}$ is plotted.
3396: In (b), $\Gamma_{Is}$ versus $\omega_m$ 
3397: is plotted for ${\bf q}=(\pi,\pi)$.
3398: These results have been obtained 
3399: for $U=4t$ and $\langle n\rangle=0.87$
3400: using the Monte Carlo data on $\chi({\bf q},i\omega_m)$.
3401: }
3402: \label{6.15}
3403: \end{figure}
3404: 
3405: Figure~6.15(a) shows the single spin-fluctuation interaction 
3406: in the singlet channel 
3407: $\Gamma_{Is}^{SF}({\bf q},i\omega_m)$ versus ${\bf q}$ 
3408: at various temperatures.
3409: These results compare well with 
3410: $\Gamma_{Is}({\bf q},i\omega_m=0)$ seen in Fig.~6.6(a).
3411: Likewise, 
3412: the comparison of Fig.~6.15(b) with Fig.~6.6(b) 
3413: shows that the frequency dependence of 
3414: $\Gamma_{Is}^{SF}$ is in agreement with the Monte Carlo data.
3415: Considering the simplicity of Eq.~(\ref{GSF}), this agreement 
3416: with the Monte Carlo data is quite good.
3417: These comparisons suggest that a properly renormalized 
3418: single-spin-fluctuation exchange interaction 
3419: is capable of reproducing the basic features 
3420: of the effective particle-particle interaction in the 
3421: weak-to-intermediate coupling Hubbard model
3422: at temperatures greater or of order $J/2$.
3423: 
3424: \subsection{Comparison with the perturbation theory}
3425: 
3426: \begin{figure}
3427: \centering
3428: \iffigure
3429: \epsfig{file=ch6-fig16.ps,height=10cm}
3430: \fi
3431: \caption{
3432: Feynman diagrams contributing to the irreducible particle-particle
3433: interaction through third order in $U$.
3434: }
3435: \label{6.16}
3436: \end{figure}
3437: 
3438: In this section, the perturbation theory results for 
3439: $\Gamma_{Is}$ through third order in $U$ 
3440: from Ref.~[Bulut {\it et al.} 1995] will be shown.
3441: These provide insight into the 
3442: various subprocesses contributing to $\Gamma_{Is}$.
3443: Figure 6.16 shows the diagrams contributing to $\Gamma_{Is}$ 
3444: up to third order in $U$.
3445: The dashed line in this figure represents the bare Coulomb repulsion. 
3446: The diagrams (a) and (h) represent the first two terms in an
3447: RPA series corresponding to the exchange of a longitudinal
3448: spin fluctuation.
3449: Similarly, 
3450: the low-order contributions arising from the exchange 
3451: of a transverse spin fluctuation are represented by the
3452: diagrams (b) and (e).
3453: The diagrams (c) and (d) can be considered 
3454: as corrections to diagram
3455: (b), where the bare particle-hole irreducible vertex is 
3456: renormalized through Kanamori type of particle-particle
3457: scatterings [Kanamori 1963].
3458: The diagrams (f) and (g) represent vertex corrections 
3459: to the bare interaction (a).
3460: 
3461: \begin{figure}
3462: \centering
3463: \iffigure
3464: \mbox{
3465: \subfigure[]{
3466: \epsfysize=8cm
3467: \epsffile[100 150 480 610]{ch6-fig17a.ps}}
3468: \quad
3469: \subfigure[]{
3470: \epsfysize=8cm
3471: \epsffile[50 150 600 610]{ch6-fig17b.ps}}}
3472: \fi
3473: \caption{
3474: Various diagrams contributing to the 
3475: irreducible particle-particle interaction 
3476: $\Gamma_{Is}({\bf q},i\omega_m)$ through third order in $U$.
3477: In (a) the momentum dependence is shown for 
3478: $\omega_m=0$, and in (b) the frequency dependence is
3479: shown for ${\bf q}=(\pi,\pi)$.
3480: Here, the filled circles represent the contribution of the bare $U$ 
3481: and the longitudinal spin-fluctuation exchange 
3482: (diagrams of type (a) and (h) in Fig.~6.16),
3483: the open circles represent the contribution of 
3484: the transverse spin fluctuations
3485: (Fig.~6.16(b) and (e)), 
3486: the open squares show the ordinary vertex corrections 
3487: (Fig.~6.16(f) and (g)), and the open triangles show the 
3488: Kanamori type of vertex corrections 
3489: (Fig.~6.16(c) and (d)).
3490: }
3491: \label{6.17}
3492: \end{figure}
3493: 
3494: \begin{figure}
3495: \centering
3496: \iffigure
3497: \mbox{
3498: \subfigure[]{
3499: \epsfysize=8cm
3500: \epsffile[100 150 480 610]{ch6-fig18a.ps}}
3501: \quad
3502: \subfigure[]{
3503: \epsfysize=8cm
3504: \epsffile[50 150 600 610]{ch6-fig18b.ps}}}
3505: \fi
3506: \caption{
3507: Comparison of the perturbation theory results on 
3508: $\Gamma_{Is}({\bf q},i\omega_m)$ through third order in $U$
3509: with the QMC data.
3510: In (a) the momentum dependence is shown for 
3511: $\omega_m=0$, and in (b) the frequency dependence is
3512: shown for ${\bf q}=(\pi,\pi)$.
3513: }
3514: \label{6.18}
3515: \end{figure}
3516: 
3517: These diagrams have been evaluated on an $8\times 8$ 
3518: lattice with $U=4t$, 
3519: and the results of the various contributions are shown as a function 
3520: of ${\bf q}$ for $\omega_m=0$ in Fig.~6.17(a).
3521: The frequency dependence for ${\bf q}=(\pi,\pi)$ 
3522: is shown similarly in Fig.~6.17(b).
3523: Here, the results on $\Gamma_{Is}$ are plotted 
3524: in the same way as in the previous section.
3525: In these figures, the filled circles represent the 
3526: diagrams (a) and (h) in Fig.~6.16, and the open circles 
3527: are for the diagrams (b) and (e).
3528: The triangles and the squares in Figs.~6.17(a) and (b) 
3529: represent the contributions of the (c)-(d) and 
3530: (f)-(g), respectively.
3531: The dominant contributions arise from the leading contributions 
3532: to the spin-fluctuation exchange processes (a), (b), (e) and (h),
3533: and they give contributions which peak at ${\bf q}=(\pi,\pi)$.
3534: For $\omega_m=0$, both the Kanamori renormalization graphs 
3535: and the vertex corrections act to reduce the strength of the 
3536: ${\bf q}=(\pi,\pi)$ interaction by about one-third.
3537: In Fig.~6.18, the results obtained by summing the diagrams shown 
3538: in Fig.~6.16 are compared with the Monte Carlo results for 
3539: $T=0.25t$.
3540: At this temperature, 
3541: the low-order graphs in Fig.~6.16 are not adequate to represent 
3542: the large momentum behaviour of the effective interaction. 
3543: 
3544: The importance of the vertex corrections to the single 
3545: spin-fluctuation exchange interaction was pointed out in 
3546: Ref.~[Schrieffer 1995].
3547: In particular, it was noted that for large values of the ratio 
3548: $\chi({\bf q}=(\pi,\pi),0)/\chi({\bf q}\rightarrow 0,0)$,
3549: the vertex corrections should suppress the spin-fluctuation 
3550: exchange interaction at ${\bf q}=(\pi,\pi)$ momentum transfer.
3551: The diagrams (f) and (g) in Fig.~6.16 are the lowest order vertex
3552: corrections, and, indeed, at this level they act to suppress 
3553: the ${\bf q}=(\pi,\pi)$ component of $\Gamma_{Is}$ 
3554: as seen in Fig.~6.17.
3555: However, the QMC data indicates that 
3556: the suppression of the peak at ${\bf q}=(\pi,\pi)$
3557: in $\Gamma_{Is}$ is partial.
3558: This is because the ratio
3559: $\chi({\bf q}=(\pi,\pi),0)/\chi({\bf q}\rightarrow 0,0)$
3560: is about 5 rather than being of order 100
3561: at the lowest temperature $\Gamma_{Is}$ 
3562: was calculated with QMC.
3563: Hence, a large AF correlation length 
3564: is not necessary, and simply weight in $\Gamma_{Is}$ at large momentum 
3565: transfers is sufficient to yield a sizeable attractive interaction in 
3566: the $d_{x^2-y^2}$-wave channel.
3567: 
3568: The results reviewed in this section constitute 
3569: what has been learned 
3570: about $d_{x^2-y^2}$-wave pairing in the 2D Hubbard model
3571: from the determinantal QMC simulations.
3572: These simulations are not carried out at lower temperatures
3573: because of the sign problem, and, hence, 
3574: it is not possible to know whether 
3575: superconducting long-range order develops at lower $T$
3576: [Loh {\it et al.} 1990].
3577: In Fig.~6.19(a), 
3578: the average sign of the fermion determinants,
3579: $\langle {\rm sign}\rangle$,
3580: which is defined in the Appendix, 
3581: is plotted as a function of $T$ 
3582: at $\langle n\rangle=0.87$ for $U=8t$ and $4t$.
3583: As the value of $\langle {\rm sign}\rangle$ decreases below 1,
3584: the statistical error in the QMC data grows rapidly requiring 
3585: exponentially long simulation times.
3586: In Fig.~6.19(b), the filling dependence of 
3587: $\langle {\rm sign}\rangle$ is shown for $U=8t$ at
3588: $T=0.5t$ and $0.33t$.
3589: These figures show the boundary of the parameter regime 
3590: of the Hubbard model which cannot be probed because of the 
3591: sign problem.
3592: However, the DMRG calculations 
3593: [White 1992] are carried out at zero temperature,
3594: and they have provided valuable information 
3595: about this regime in the 2-leg Hubbard ladder.
3596: The DMRG studies of the 2-leg Hubbard ladder 
3597: will be reviewed in the next section. 
3598: 
3599: \begin{figure}
3600: \centering
3601: \iffigure
3602: \mbox{
3603: \subfigure{
3604: \epsfysize=8cm
3605: \epsffile[100 150 480 610]{ch6-fig19a.ps}}
3606: \quad
3607: \subfigure{
3608: \epsfysize=8cm
3609: \epsffile[50 150 600 610]{ch6-fig19b.ps}}}
3610: \fi
3611: \caption{
3612: (a) Temperature and (b) the filling dependence of 
3613: $\langle {\rm sign}\rangle$.
3614: In (a), the results are shown for 
3615: $U=8t$ and $4t$ at $\langle n\rangle=0.87$,
3616: and in (b) for $U=8t$.
3617: }
3618: \label{6.19}
3619: \end{figure}
3620: 
3621: \setcounter{equation}{0}\setcounter{figure}{0}
3622: \section{2-leg Hubbard ladder}
3623: 
3624: The 2-leg Hubbard ladder 
3625: has been studied using various many-body techniques 
3626: such as the DMRG 
3627: [Noack {\it et al.} 1994, 1995, 1996 and 1997], 
3628: the QMC [Dahm and Scalapino 1997], 
3629: the exact diagonalization [Yamaji and Shimoi 1994], 
3630: and the weak-coupling renormalization group
3631: [Balents and Fisher 1996].
3632: The results of the DMRG and 
3633: the QMC calculations will be reviewed here.
3634: 
3635: At half-filling, the ground state of this system does not have AF
3636: long-range order, rather it has a spin gap.
3637: Near half-filling, this model exhibits short-range AF correlations and 
3638: power-law decaying $d_{x^2-y^2}$-like superconducting and 
3639: "$4{\bf k}_F$" CDW correlations
3640: [Noack {\it et al.} 1994 and 1996].
3641: The DMRG calculations have found that in the ground state of 
3642: this system the pair-field correlations for the interacting 
3643: system can get enhanced with respect to those of the noninteracting 
3644: ($U=0$) system for a range of the model parameters.
3645: This is the first time ever in an exact ground state calculation 
3646: for a bulk system that 
3647: by turning on an onsite Coulomb repulsion the superconducting 
3648: correlations get enhanced. 
3649: For this reason, the 2-leg Hubbard model is quite important.
3650: 
3651: Another reason for studying this model is that here it is possible to 
3652: understand the mechanism mediating the $d_{x^2-y^2}$-like
3653: superconducting correlations by comparing the DMRG results with the 
3654: QMC data obtained at relatively low temperatures.
3655: These comparisons indicate that it is the short-range AF fluctuations 
3656: which mediate the pairing.    
3657: Furthermore, 
3658: the pairing is strongest when the model parameters 
3659: are such that there is enhanced single-particle spectral weight 
3660: near the $(\pi,0)$ and the $(0,\pi)$ points of the Brillouin zone.
3661: For $U=8t$ and $\langle n\rangle =0.875$, 
3662: this occurs when $t_{\perp}/t \sim 1.5$.
3663: In this case, the irreducible particle-particle vertex peaks at 
3664: momentum transfers near $(\pi,\pi)$ creating optimum conditions for 
3665: $d_{x^2-y^2}$ pairing.
3666: 
3667: It is also interesting to study the 2-leg Hubbard ladder 
3668: because the half-filled insulating state is spin gapped while 
3669: in the 2D case there is long-range AF order. 
3670: In the cuprates, on the other hand, the undoped system has 
3671: long-range AF order, and a spin gapped phase lies between the 
3672: superconducting and the insulating phases. 
3673: The 2-leg Hubbard model is a system where the relation between the 
3674: spin gap and the superconducting correlations as well as the 
3675: density correlations can be studied exactly.
3676: 
3677: In section 7.1 below, the DMRG results on the pair-field correlation 
3678: function from Ref.~[Noack {\it et al.} 1997] will be shown for various 
3679: values of the model parameters.
3680: In order to understand these results better, 
3681: in Section~7.2 the QMC results 
3682: on the single-particle spectral weight 
3683: from Ref.~[Noack {\it et al.} 1997], 
3684: the irreducible particle-particle interaction 
3685: and the solution of the Bethe-Salpeter equation 
3686: from Ref.~[Dahm and Scalapino 1997] will be presented. 
3687: In Section 7.3, the results on the 2-leg Hubbard ladder will be 
3688: compared with those on the 2D case.
3689: Later, in Section 8.5.3, comparisons will be made with the
3690: superconducting correlations found in the 2-leg $t$-$J$ ladder.
3691: 
3692: \subsection{DMRG results}
3693: 
3694: \begin{figure}
3695: \centering
3696: \iffigure
3697: \epsfig{file=ch7-fig1.ps,height=6cm}
3698: \fi
3699: \caption{
3700: Schematic drawing of the pair-wave function
3701: showing the values of the off-diagonal matrix element 
3702: $\langle N_2 | ( 
3703: c^{\dagger}_{{\bf r}\uparrow} c^{\dagger}_{{\bf r'}\downarrow} 
3704: -
3705: c^{\dagger}_{{\bf r}\downarrow} c^{\dagger}_{{\bf r'}\uparrow} )
3706: | N_1\rangle$
3707: for creating a singlet pair between near-neighbor sites.
3708: Here, it is seen that 
3709: the matrix element for this process is negative
3710: when the singlet-pair is created along a chain,
3711: while it is positive across a rung. 
3712: This shows the $d_{x^2-y^2}$-like nature of the pairing correlations 
3713: in the 2-leg Hubbard ladder.
3714: }
3715: \label{7.1}
3716: \end{figure}
3717: 
3718: The DMRG calculations found that there are power-law decaying 
3719: $d_{x^2-y^2}$-wave pair-field correlations 
3720: in the ground state of the 2-leg
3721: Hubbard ladder.
3722: The $d_{x^2-y^2}$-like internal structure of the pairs
3723: can be seen by considering the pair-creation amplitude
3724: \begin{equation}
3725: \label{amplitude}
3726: \langle N_2 | ( 
3727: c^{\dagger}_{{\bf r}\uparrow} c^{\dagger}_{{\bf r'}\downarrow} 
3728: -
3729: c^{\dagger}_{{\bf r}\downarrow} c^{\dagger}_{{\bf r'}\uparrow} )
3730: | N_1\rangle
3731: \end{equation}
3732: for adding a singlet pair on near-neighbor sites along 
3733: and across the legs.
3734: Here, $|N_1\rangle$ is the ground state with four holes relative 
3735: to the half-filled band and $|N_2\rangle$ is the ground state 
3736: with two holes on a $2\times 16$  ladder.
3737: The results on the pair amplitude 
3738: are shown in Fig.~7.1
3739: for $U/t=8$ and $t_{\perp}/t=1.5$.
3740: Note the $d_{x^2-y^2}$-like change 
3741: in the sign of this matrix element. 
3742: 
3743: Using the DMRG method it is possible to calculate 
3744: the rung-rung correlation function
3745: \begin{equation}
3746: \label{D}
3747: D(i,j) = \langle \Delta(i) \Delta^{\dagger}(j) \rangle
3748: \end{equation}
3749: in the ground state of the 2-leg Hubbard ladder with
3750: the open boundary conditions.
3751: Here, 
3752: \begin{equation}
3753: \Delta^{\dagger}(i) = 
3754: c^{\dagger}_{i1\uparrow} c^{\dagger}_{i2\downarrow} -
3755: c^{\dagger}_{i1\downarrow} c^{\dagger}_{i2\uparrow} 
3756: \end{equation}
3757: creates a singlet pair across the $i$'th rung,
3758: and $c^{\dagger}_{ik\sigma}$ is the electron creation operator 
3759: with spin $\sigma$
3760: at the $i$'th site of the $k$'th leg of the 2-leg ladder.
3761: The fact that the matrix elements shown in Fig.~7.1 
3762: are finite means that this bare pair-creation operator,
3763: which is composed of the bare electron-creation 
3764: operators $c^{\dagger}_{ik\sigma}$,
3765: has finite overlap with the true pair-creation
3766: eigen-operator for this system.
3767: In order to minimize the effects of the boundaries, 
3768: here $D(i,j)$ is averaged over six $(i,j)$ pairs with 
3769: $\ell=|i-j|$ fixed. 
3770: This averaging starts with symmetrically placed $(i,j)$ 
3771: values and then proceeds to shift 
3772: these to the left and right of the center. 
3773: By comparing results obtained on different size lattices, 
3774: it is possible to control the finite size effects. 
3775: In the following, 
3776: $D(\ell)$ calculated on the $2\times 32$ lattice for 
3777: $\ell<20$ will be shown. 
3778: In this case, the finite size effects are negligible.
3779: 
3780: \begin{figure}
3781: \centering
3782: \iffigure
3783: \epsfig{file=ch7-fig2.ps,height=8cm}
3784: \fi
3785: \caption{
3786: Pair-field correlation function 
3787: $D(\ell)$ versus $\ell$ for various values of 
3788: $t_{\perp}/t$ with $U=8t$ and $\langle n\rangle =0.875$.
3789: }
3790: \label{7.2}
3791: \end{figure}
3792: 
3793: Figure ~7.2 shows $D(\ell)$ versus $\ell$ for various values of 
3794: $t_{\perp}/t$ with $U=8t$ and $\langle n\rangle =0.875$.
3795: The dashed and the dotted lines represent power-law decays of 
3796: $\ell^{-2}$ and $\ell^{-1}$, respectively. 
3797: In this figure, it is seen that 
3798: $D(\ell)$ exhibits a power-law decay for 
3799: $t_{\perp}/t< 1.6$.
3800: For $t_{\perp}/t=1.0$, $D(\ell)$ decays as $\ell^{-2}$.
3801: When $t_{\perp}/t$ is increased from
3802: 1.0 to 1.4, the strength of $D(\ell)$ 
3803: gets enhanced and it decays more slowly.
3804: For $t_{\perp}/t=1.6$, $D(\ell)$ is reduced and it decays faster.
3805: Below in Section 7.1, it will be seen that when $t_{\perp}/t > 1.4$,
3806: the antibonding single-particle band becomes unoccupied,
3807: and the decrease in the strength of the pairing correlations
3808: is due to this effect. 
3809: Using the data in Fig.~7.2 for 
3810: $1 < \ell <18$, $D(\ell)$ has been fitted to a form
3811: \begin{equation}
3812: D(\ell) = {1 \over {\ell^{\theta}} }
3813: \end{equation}
3814: with a linear least-squares approximation.
3815: The resulting $\theta$ values are plotted as a function of 
3816: $t_{\perp}/t$ for various fillings in Fig.~7.3.
3817: Here, it is seen that the minimum in $\theta$
3818: versus $t_{\perp}/t$ depends on the filling, but it occurs
3819: for $t_{\perp}/t\sim 1.4$ near half-filling. 
3820: For $t_{\perp}/t > 1.4$, the antibonding band is no longer 
3821: occupied and the pairing correlations decrease rapidly.
3822: Thus the pairing correlations are enhanced near the point 
3823: at which the antibonding band moves through the Fermi level
3824: [Noack {\it et al.} 1995 and 1997, Yamaji and Shimoi 1994].
3825: 
3826: \begin{figure}[ht]
3827: \centering
3828: \iffigure
3829: \epsfig{file=ch7-fig3.ps,height=8cm}
3830: \fi
3831: \caption{
3832: Exponent $\theta$ versus $t_{\perp}/t$ for 
3833: $U/t=8$ and various values of $\langle n\rangle$.
3834: }
3835: \label{7.3}
3836: \end{figure}
3837: 
3838: Another measure of the strength of the pair-field 
3839: correlations which can be used is the average of 
3840: $D(\ell)/D(1)$ for rung separations $\ell=8$ to 12:
3841: \begin{equation}
3842: \label{Dbar}
3843: \overline{D} = {1\over 5}
3844: \sum_{\ell=8}^{12} \,
3845: { D(\ell) \over D(1) }.
3846: \end{equation}
3847: Figure~7.4 shows $\overline{D}$ versus 
3848: $t_{\perp}/t$ for $U=8t$ at various fillings.
3849: This clearly shows how sensitively the pairing correlations 
3850: depend on the value of $t_{\perp}/t$.
3851: Next, 
3852: the variation of $\overline{D}$ with $U/t$ is shown 
3853: in Fig.~7.5.
3854: In this figure, the crosses represent the results for $U=0$.
3855: Hence, the onsite Coulomb repulsion can significantly
3856: enhance the pairing correlations for a range of $t_{\perp}/t$
3857: values.
3858: Here, one also observes that the enhancement of 
3859: $\overline{D}$ is strongest 
3860: for $U/t$ between 3 and 8, which is in
3861: the intermediate coupling regime.
3862: In addition, as $U/t$ increases the value of $t_{\perp}/t$ 
3863: at which the peak in $\overline{D}$ occurs shifts towards
3864: smaller values.
3865: Hence, the strength of the pairing correlations is a 
3866: sensitive function of $t_{\perp}/t$, $U/t$ and
3867: $\langle n\rangle$. 
3868: 
3869: \begin{figure}[ht]
3870: \centering
3871: \iffigure
3872: \epsfig{file=ch7-fig4.ps,height=8cm}
3873: \fi
3874: \caption{
3875: Averaged pair-field correlation function 
3876: $\overline{D}$ versus $t_{\perp}/t$
3877: for $U=8t$ and various values of 
3878: $\langle n\rangle$.
3879: }
3880: \label{7.4}
3881: \end{figure}
3882: 
3883: \begin{figure}
3884: \centering
3885: \iffigure
3886: \epsfig{file=ch7-fig5.ps,height=8cm}
3887: \fi
3888: \caption{
3889: Averaged pair-field correlation function 
3890: $\overline{D}$ versus $t_{\perp}/t$
3891: for $\langle n\rangle=0.9375$ and various values of $U/t$.
3892: }
3893: \label{7.5}
3894: \end{figure}
3895: 
3896: These DMRG results have a special place in many-body physics.
3897: They represent the first case in an exact ground-state 
3898: calculation for a bulk system 
3899: where the pair-field correlation function gets enhanced 
3900: by turning on an onsite Coulomb repulsion.
3901: These calculations were carried out on ladders with up to 32 rungs 
3902: resulting in negligible finite-size effects.
3903: As seen in Section 6.1, 
3904: in the 2D Hubbard model, 
3905: it was found that by turning 
3906: on the Coulomb repulsion the $d_{x^2-y^2}$-wave pair-field 
3907: susceptibility $P_d$ gets enhanced with respect to 
3908: $\overline{P}_d$.
3909: However, $P_d$ was always found to be suppressed with respect 
3910: to the pair-field susceptibility $P_d^0$ of the $U=0$ system.
3911: This meant that, at these temperatures, the effective attractive 
3912: interaction in the $d_{x^2-y^2}$-wave channel is not sufficiently 
3913: strong to overcome the suppression 
3914: of $P_d$
3915: induced by the single-particle self-energy effects.
3916: Here, it is seen that for the 2-leg Hubbard ladder 
3917: in the ground state, 
3918: $\overline{D}$ can get enhanced over the $U=0$ result for 
3919: a set of the model parameters.
3920: 
3921: \subsection{QMC results}
3922: 
3923: In order to understand these DMRG results better, in this section 
3924: QMC data on the 2-leg Hubbard ladder will be shown.
3925: 
3926: \subsubsection{Single-particle spectral weight}
3927: 
3928: The single-particle properties of the 2-leg Hubbard 
3929: ladder at half-filling were studied in 
3930: Ref.~[Endres {\it et al.} 1996].
3931: Here,
3932: the evolution of the 
3933: single-particle spectral weight $A({\bf k},\omega)$ 
3934: with $t_{\perp}/t$ will be shown for the doped case. 
3935: These data, 
3936: which were obtained for a $2\times 16$ ladder 
3937: with periodic boundary conditions along the chains, 
3938: are from Ref.~[Noack {\it et al.} 1997]. 
3939: The comparison of the results 
3940: on $A({\bf k},\omega)$ and $D(\ell)$ will show 
3941: that the pairing correlations are enhanced 
3942: for $t_{\perp}/t$ such that the bottom 
3943: of the antibonding band moves through the Fermi level.
3944: 
3945: \begin{figure}
3946: \centering
3947: \iffigure
3948: \mbox{
3949: \subfigure{
3950: \epsfig{file=ch7-fig6a.ps,height=5cm}}
3951: \quad
3952: \subfigure{
3953: \epsfig{file=ch7-fig6b.ps,height=5cm}}}
3954: \fi
3955: \iffigure
3956: \epsfig{file=ch7-fig6c.ps,height=5cm}
3957: \fi
3958: \caption{
3959: Distribution of the 
3960: single-particle spectral weight $A({\bf k},\omega)$ 
3961: in the ${\bf k}$ and $\omega$ plane.
3962: The intensity of the shading indicates the amount
3963: of the spectral weight.
3964: These results are for $U=2t$, 
3965: $\langle n\rangle=0.94$, $T=0.125t$ and 
3966: (a) $t_{\perp}/t=1.6$, (b) 1.8 and (c) 2.0. 
3967: }
3968: \label{7.6}
3969: \end{figure}
3970: 
3971: The results on $A({\bf k},\omega)$ were obtained by the maximum-entropy 
3972: analytic continuation of the Monte Carlo data.
3973: Figures 7.6 and 7.7 show data for $U/t=2$ and 4, respectively.
3974: In these figures $\langle n\rangle=0.94$ and 
3975: $T=0.125t$, and the results are given for various values of 
3976: $t_{\perp}/t$. 
3977: Here, $A({\bf k},\omega)$ is shown for both 
3978: $k_{\perp}=0$ (bonding) and $k_{\perp}=\pi$ (antibonding)
3979: bands as a contour plot in the 
3980: $\omega$-$k$ plane where the intensity of the 
3981: shading represents the magnitude of $A({\bf k},\omega)$ and
3982: $k$ is the momentum along the chains. 
3983: In the $U=0$ system, 
3984: the quasiparticle dispersion consists of the bonding and the 
3985: antibonding bands given by 
3986: \begin{equation}
3987: \varepsilon_{\bf k} = - 2t \cos{k} \pm 2 t_{\perp},
3988: \end{equation}
3989: where ${\bf k}=(k,k_{\perp})$.
3990: The dotted curves in Fig.~7.6 represent 
3991: $\varepsilon_{\bf k}$ for $U=0$.
3992: Here, 
3993: it is seen that the dispersion obtained for
3994: $U/t=2$ closely follows that of the $U=0$ system.
3995: Note also that for $t_{\perp}/t=1.8$ the bottom 
3996: of the antibonding band is located right at the Fermi level while 
3997: for $t_{\perp}/t=2.0$ the antibonding band becomes unoccupied.
3998: In Fig.~7.5, it was seen that the pairing correlations for $U/t=2$ 
3999: are strongest when $t_{\perp}/t=1.8$.
4000: This comparison suggests that it is the variation 
4001: in the single-particle spectral weight with $t_{\perp}/t$ 
4002: which controls the dependence 
4003: of the pairing correlations on $t_{\perp}/t$.
4004: 
4005: \begin{figure}
4006: \centering
4007: \centering
4008: \iffigure
4009: \mbox{
4010: \subfigure{
4011: \epsfig{file=ch7-fig7a.ps,height=5cm}}
4012: \quad
4013: \subfigure{
4014: \epsfig{file=ch7-fig7b.ps,height=5cm}}}
4015: \fi
4016: \iffigure
4017: \epsfig{file=ch7-fig7c.ps,height=5cm}
4018: \fi
4019: \caption{
4020: Distribution of the 
4021: single-particle spectral weight $A({\bf k},\omega)$ 
4022: in the ${\bf k}$ and $\omega$ plane.
4023: The intensity of the shading indicates the amount
4024: of the spectral weight.
4025: These results are for $U=4t$, 
4026: $\langle n\rangle=0.94$, $T=0.125t$ and 
4027: (a) $t_{\perp}/t=1.4$, (b) 1.6 and (c) 1.8. 
4028: }
4029: \label{7.7}
4030: \end{figure}
4031: 
4032: Figure~7.7 shows similar data on $A({\bf k},\omega)$ 
4033: for $U/t=4$ and $t_{\perp}/t=1.4$, 1.6 and 1.8.
4034: In this case, 
4035: there are more differences 
4036: between the QMC data and the $U=0$ results
4037: denoted by the dotted curves.
4038: Here,
4039: the antibonding band becomes unoccupied for a smaller value of 
4040: $t_{\perp}/t=1.8$.
4041: One also notices that the top of the bonding and the bottom of the 
4042: antibonding bands are flattened, 
4043: especially for $t_{\perp}/t=1.6$, 
4044: increasing the amount of 
4045: the single-particle spectral weight near the Fermi level.
4046: This behaviour is similar to the build up of spectral weight near
4047: $(\pi,0)$ in the 2D Hubbard model as discussed in 
4048: Section~5, and in the ARPES experiments on the 
4049: high-$T_c$ cuprates which are reviewed by [Shen and Dessau 1995].
4050: 
4051: In the next section, 
4052: it will be seen that for the 2-leg 
4053: Hubbard ladder the irreducible particle-particle 
4054: scattering vertex $\Gamma_I$ peaks at $(\pi,\pi)$ 
4055: momentum transfer as for the 2D Hubbard model.
4056: These results suggest that the peaking of $\Gamma_I$ 
4057: near $(\pi,\pi)$ momentum transfer along with the enhanced 
4058: single-particle spectral weight enhances the 
4059: pairing correlations, and this is the reason for the strong
4060: dependence of $\overline{D}$ on $t_{\perp}/t$ seen in 
4061: Fig~7.4.
4062: 
4063: \subsubsection{Irreducible particle-particle interaction}
4064: 
4065: In the previous sections, 
4066: it has been seen that the 2-leg 
4067: Hubbard ladder has power-law $d_{x^2-y^2}$-wave pairing correlations 
4068: in its ground state, and for a range of the model parameters 
4069: the pairing correlations can get strong. 
4070: It is useful to gain insight into the mechanism which leads
4071: to the $d_{x^2-y^2}$-wave pairing correlations in this model.
4072: For this reason, here the Monte Carlo results on the 
4073: irreducible particle-particle interaction $\Gamma_I$ of the
4074: 2-leg Hubbard ladder will be reviewed.
4075: These are results from [Dahm and Scalapino 1997].
4076: Comparisons with the magnetic susceptibility $\chi({\bf q},0)$ 
4077: will show that the short-range AF spin-fluctuations are responsible 
4078: for the momentum structure in $\Gamma_I$
4079: for ${\bf q}$ near $(\pi,\pi)$.
4080: Using the data on $\Gamma_I$ and the single-particle 
4081: Green's function $G$, the Bethe-Salpeter equation will be solved 
4082: in the particle-particle channel
4083: and the leading singlet pairing channel will be shown to have 
4084: $d_{x^2-y^2}$-like symmetry.
4085: 
4086: \begin{figure}
4087: \centering
4088: \iffigure
4089: \mbox{
4090: \subfigure[]{
4091: \epsfysize=8cm
4092: \epsffile[100 150 480 610]{ch7-fig8a.ps}}
4093: \quad
4094: \subfigure[]{
4095: \epsfysize=8cm
4096: \epsffile[50 150 600 610]{ch7-fig8b.ps}}}
4097: \fi
4098: \caption{
4099: Monte Carlo results on the momentum dependence of 
4100: (a) the irreducible particle-particle interaction 
4101: in the singlet channel
4102: $\Gamma_{Is}({\bf q},i\omega_m=0)$ and 
4103: (b) the magnetic susceptibility $\chi({\bf q},i\omega_m=0)$. 
4104: These results were obtained on a $2\times 16$ lattice  
4105: for $U=4t$, $T=0.25t$, $\langle n\rangle=0.87$
4106: and $t_{\perp}=1.5t$.
4107: }
4108: \label{7.8}
4109: \end{figure}
4110: 
4111: Using Monte Carlo simulations,
4112: the singlet irreducible vertex 
4113: $\Gamma_{Is}({\bf q},i\omega_m)$ has been calculated on a 
4114: $2\times 16$ lattice for $U/t=4$ and 
4115: $\langle n\rangle = 0.875$
4116: in the same way as discussed in Section 6.2.
4117: In addition, here $t_{\perp}/t=1.5$ was chosen so that the 
4118: system has strong pairing correlations in the ground state. 
4119: In Fig.~7.8(a), 
4120: $\Gamma_{Is}({\bf q},i\omega_m=0)$ is plotted 
4121: as a function of $q$ where ${\bf q}=(q,\pi)$
4122: at various temperatures.
4123: Here, as in Section~6, ${\bf q}={\bf p}-{\bf p'}$ is the momentum 
4124: transfer and ${\bf p'}$ is kept fixed at $(\pi,0)$ while ${\bf p}$ 
4125: is scanned.
4126: At high temperatures, 
4127: $\Gamma_{Is}({\bf q})$ is flat in momentum space with 
4128: a magnitude varying between $8t$ and $10t$,
4129: and as $T$ is lowered to $0.25t$, 
4130: $\Gamma_{Is}({\bf q},0)$ develops significant amount of 
4131: weight at ${\bf q}=(\pi,\pi)$ momentum transfer 
4132: becoming of order $20t$.
4133: This behaviour is similar to what has been seen in Section~6 for the 
4134: 2D Hubbard model.
4135: 
4136: Figure~7.8(b) shows the momentum dependence of the magnetic susceptibility 
4137: $\chi({\bf q},0)$ for the 2-leg Hubbard ladder 
4138: for the same model parameters.
4139: Here, 
4140: $\chi({\bf q},0)$ is also plotted as a function of $q$
4141: where ${\bf q}=(q,\pi)$.
4142: Comparing with Fig.~7.8(a), 
4143: one observes that the evolution of 
4144: $\Gamma_{Is}$ with temperature is 
4145: closely related to that of $\chi({\bf q},0)$,
4146: which implies that the short-range antiferromagnetic correlations 
4147: are responsible for the momentum structure in 
4148: $\Gamma_{Is}({\bf q},0)$
4149: for ${\bf q}$ near $(\pi,\pi)$.
4150: 
4151: \begin{figure}
4152: \centering
4153: \iffigure
4154: \epsfysize=8cm
4155: \epsffile[100 150 550 610]{ch7-fig9.ps}
4156: \fi
4157: \caption{
4158: The leading singlet eigenfunction 
4159: $\phi({\bf p},i\omega_n)$
4160: versus $p_x$ where ${\bf p}=(p_x,p_y)$ and $\omega_n=\pi T$.
4161: These results were obtained on a $2\times 16$ lattice 
4162: for $U=4t$, $T=0.25t$, $\langle n\rangle=0.87$ and
4163: $t_{\perp}=1.5t$.
4164: }
4165: \label{7.9}
4166: \end{figure}
4167: 
4168: \subsubsection{Bethe-Salpeter equation}
4169: 
4170: In order to see what type of pairing state is favored by the
4171: irreducible particle-particle vertex 
4172: seen in the previous section, 
4173: here results from the solution of the Bethe-Salpeter equation
4174: \begin{equation}
4175: \lambda_{\alpha}\phi_{\alpha}({\bf p},i\omega_n) = 
4176: - {T\over N} \, \sum_{\bf p'} \,
4177: \Gamma_I({\bf p}-{\bf p'},
4178: i\omega_n-i\omega_{n'})
4179: |G({\bf p'},i\omega_n{n'})|^2
4180: \phi_{\alpha}({\bf p'},i\omega_{n'})
4181: \end{equation}
4182: from Ref.~[Dahm and Scalapino 1997] will be shown.
4183: Figure~7.9 shows the momentum dependence of the leading
4184: eigenfunction $\phi({\bf p},i\omega_n)$ 
4185: at $\omega_n=\pi T$ for $T=0.25t$.
4186: Figures~7.10(a) and (b) show the dependence of the leading 
4187: singlet eigenvalue $\lambda_1$ on $T/t$ and $t_{\perp}/t$,
4188: respectively.
4189: Note that $\phi({\bf p},i\pi T)$ peaks 
4190: near $(\pi,0)$ and $(0,\pi)$, and it changes sign 
4191: between these two points. 
4192: Hence, in this sense it is $d_{x^2-y^2}$-like, 
4193: but it does not have a node since 
4194: $\phi({\bf p},i\pi T)$ does not vanish 
4195: near any of the four Fermi surface points.
4196: In the previous section, we have seen that for 
4197: $t_{\perp}/t\sim 1.5$, the bonding band has spectral weight 
4198: near the Fermi level for ${\bf p}\sim(\pi,0)$, and 
4199: the antibonding band has spectral weight near the Fermi level
4200: for ${\bf p}\sim (0,\pi)$.
4201: Hence, these Fermi points can be connected by scatterings 
4202: involving ${\bf q}=(\pi,\pi)$ momentum transfer.
4203: Since $\Gamma_{Is}$ is large and repulsive for 
4204: ${\bf q}\sim (\pi,\pi)$, the leading singlet eigenfunction 
4205: $\phi$ of the Bethe-Salpeter equation has opposite signs for 
4206: ${\bf p}$ near $(\pi,0)$ and 
4207: $(0,\pi)$.
4208: 
4209: \begin{figure}
4210: \centering
4211: \iffigure
4212: \mbox{
4213: \subfigure{
4214: \epsfysize=8cm
4215: \epsffile[100 150 480 610]{ch7-fig10a.ps}}
4216: \quad
4217: \subfigure{
4218: \epsfysize=8cm
4219: \epsffile[50 150 600 610]{ch7-fig10b.ps}}}
4220: \fi
4221: \caption{
4222: (a) Temperature and (b) $t_{\perp}/t$ 
4223: dependence of the leading singlet
4224: eigenvalue $\lambda_1$.
4225: These results were obtained on a $2\times 16$ lattice 
4226: for $U=4t$ and $\langle n\rangle=0.87$.
4227: In (a) $t_{\perp}/t=1.5$ and in (b) $T=0.25t$ were used.
4228: }
4229: \label{7.10}
4230: \end{figure}
4231: 
4232: In Fig.~7.10(a), it is seen that the eigenvalue of the 
4233: $d_{x^2-y^2}$-like eigenfunction increases by about an order of magnitude 
4234: as $T$ is lowered from $1t$ to $0.25t$.
4235: However, even at $T=0.25t$, which is the lowest temperature 
4236: where the Monte Carlo calculation of the Bethe-Salpeter 
4237: eigenvalues can be carried out, the leading singlet eigenvalue 
4238: is only about 0.3.
4239: 
4240: Figure~7.10(b) shows that at $T=0.25t$ the leading singlet 
4241: eigenvalue has in fact a weak dependence on $t_{\perp}/t$, 
4242: and only a broad peak at $t_{\perp}/t=1.6$ is seen.
4243: This is unexpected considering the strong dependence of 
4244: $\overline{D}$ on $t_{\perp}/t$
4245: seen in the previous section.
4246: It means that at $T=0.25t$ the pairing interaction has not reached
4247: its full strength and, in addition, the thermal smearing effects 
4248: significantly weaken the pairing correlations
4249: which are observed in the ground state by the DMRG technique.
4250: 
4251: \subsection{Comparison of the 2-leg and the 2D Hubbard models}
4252: 
4253: It is interesting to compare these results    
4254: on the 2-leg ladder with those on the
4255: 2D case seen in Section~6.
4256: In Fig.~7.10(b), one sees that at $T=0.25t$, $U/t=4$ 
4257: and $t_{\perp}/t=1.0$ 
4258: the leading singlet eigenvalue is 0.22, 
4259: which is slightly larger than
4260: $\lambda_d=0.18$ at the same $U/t$, $\langle n\rangle$ and 
4261: $T/t$ values for the 2D case.
4262: For $t_{\perp}/t=1.0$, the DMRG calculations find that 
4263: $D(\ell)$ varies approximately as $\ell^{-2}$ 
4264: in the ground state and, hence, 
4265: the pairing correlations are only as strong as in the $U=0$ case.
4266: If one just uses this comparison of the eigenvalues for the 
4267: 2-leg and the 2D cases, then one would expect weak
4268: pairing correlations in the ground state of the 2D Hubbard model.
4269: However, in 2D,
4270: $\lambda_d$ for $U=8t$ is found to be higher 
4271: than for $U=4t$ at $T=0.5t$.
4272: Furthermore, 
4273: from the DMRG calculations one knows that when 
4274: $t_{\perp}/t$ is tuned to the right value, in the 2-leg ladder
4275: strong pairing correlations can develop.
4276: A similar dependence on the model parameters,
4277: such as the second-nearest-neighbour hopping $t'$,
4278: could exist in the 2D Hubbard model. 
4279: 
4280: It has been seen above that the QMC results 
4281: on the irreducible particle-particle vertex 
4282: of the 2-leg ladder are similar to those on the 
4283: 2D Hubbard model.
4284: In the ground state of the 2-leg ladder, the DMRG calculations 
4285: find enhanced $d_{x^2-y^2}$-wave pairing correlations 
4286: in a certain range of the parameters.
4287: Both in the 2D and the 2-leg models, we have seen that, when doped
4288: with holes there are short-range AF correlations and
4289: that they strongly influence the low-energy 
4290: single-particle properties.
4291: In both cases, 
4292: short-range AF correlations cause 
4293: $\Gamma_I({\bf q},0)$ to peak at $(\pi,\pi)$ momentum transfer.
4294: It needs to be noted that the pairing correlations observed 
4295: in these two models do not require a particularly sharp peak in 
4296: $\Gamma_{I}({\bf q},0)$ at ${\bf q}=(\pi,\pi)$, 
4297: but rather simply weight at large momentum transfers.
4298: 
4299: Currently, it is not possible to determine whether
4300: the doped 2D Hubbard model has long-range 
4301: $d_{x^2-y^2}$-wave superconducting order in its ground state,
4302: or, if it did, 
4303: whether it would be sufficient to explain superconducting
4304: transition temperatures as high as those found in the cuprates.
4305: In spite of this, 
4306: the results discussed above show that effects which 
4307: increase the single-particle spectral weight 
4308: near the $(\pi,0)$ and $(0,\pi)$ points of the 
4309: Brillouin zone as well as effects which increase 
4310: the strength of the particle-particle interaction 
4311: at $(\pi,\pi)$ momentum transfer will act to enhance 
4312: $d_{x^2-y^2}$-wave pairing.
4313: 
4314: It is possible that the strength of the $d_{x^2-y^2}$
4315: pairing correlations in the ground state of the 2D Hubbard
4316: model depends sensitively on the model parameters such as
4317: the second-near-neighbour hopping $t'$ in a way similar
4318: to what is seen for the 2-leg ladder in Section 7.
4319: In the 2-leg case, the pairing correlations are 
4320: as weak as those of the $U=0$ system when $t_{\perp}=t$.
4321: However, 
4322: when $t_{\perp}/t$ is tuned so that the flat bands 
4323: are located near the Fermi level, the system exhibits 
4324: enhanced pairing correlations.
4325: It is possible that the 2D Hubbard model 
4326: with only near-neighbour hopping similarly exhibits 
4327: weak pairing correlations in the ground state, 
4328: but there can be enhanced pairing when an additional 
4329: parameter such as $t'$ is tuned.
4330: In the 2-leg ladder case,
4331: the value of $t_{\perp}/t$ for which 
4332: the flat bands are near the Fermi level is renormalized
4333: by the Coulomb repulsion. 
4334: Similarly,  
4335: in the 2D case, 
4336: the optimum value of $t'$ could be renormalized,
4337: and it is difficult to estimate it in advance.
4338: In a bilayer Hubbard model, 
4339: the bilayer coupling could also play a role.
4340: These are issues which need to be resolved 
4341: by exact techniques in the future.
4342: 
4343: Beyond these,
4344: one would expect that any additional contribution 
4345: to the irreducible electron-electron interaction 
4346: which is repulsive for ${\bf q}\sim (\pi,\pi)$
4347: momentum transfers or attractive for ${\bf q}\sim 0$ 
4348: would act to enhance the $d_{x^2-y^2}$ pairing, 
4349: when added to the 2D or the 2-leg Hubbard system. 
4350: For instance, 
4351: when a phonon mediated interaction which is most attractive for 
4352: ${\bf q}\sim 0$ momentum transfers 
4353: is added to an AF spin-fluctuation exchange interaction, 
4354: it is found within the $t$-matrix approximation 
4355: that the $d_{x^2-y^2}$ eigenvalue gets enhanced
4356: [Bulut and Scalapino 1996].
4357: However, 
4358: it is necessary to study such effects using exact techniques.
4359: 
4360: It could also be possible to design flat band dispersion
4361: near the Fermi level by controlling the lattice geometry
4362: and parameters [Imada and Kohno 2000].
4363: Imada and Kohno have carried out exact diagonalization calculations
4364: for a 1D 16-site $t$-$J$ model with additional
4365: three-site terms and longer range hoppings.
4366: By tuning the longer-range hopping parameters,
4367: they have created flat band dispersion near the Fermi level and,
4368: in this case, 
4369: they find an enhanced spin gap and an enhanced tendency 
4370: for pairing.
4371: They have also proposed various multiband models 
4372: which could exhibit flat bands near the Fermi level 
4373: and enhanced pairing.
4374: 
4375: Even though for the 2D case the low-doping 
4376: and the low-temperature regime where a spin gap could 
4377: exist is beyond the reach of the exact techniques,
4378: in the 2-leg case the spin gap 
4379: $\Delta_s$ can attain large values. 
4380: In the doped 2-leg ladder, 
4381: $\Delta_s$ is maximum when the bottom of the 
4382: antibonding band is near the Fermi level
4383: [Noack {\it et al.} 1996].
4384: For instance, 
4385: for $U=8t$, $\langle n\rangle=0.875$ and $t_{\perp}=1.5t$,
4386: the spin gap has the value of $0.06t$,
4387: which corresponds to $\approx 300K$
4388: for a $t$ of order 0.45~eV.
4389: 
4390: It is also useful to compare the density correlations 
4391: seen in the 2-leg and the 2D Hubbard models.
4392: With the DMRG method [Noack {\it et al.} 1996], 
4393: the following density-density correlation 
4394: function has been calculated for the $2\times 32$ Hubbard ladder,
4395: \begin{equation}
4396: S(i,j,\lambda) = \langle  n_{i\lambda} n_{j1} \rangle 
4397: -\langle n_{i\lambda} \rangle   \langle n_{j1}\rangle. 
4398: \end{equation}
4399: Here, $n_{i\lambda}$ is the electron occupation number at the 
4400: $i$'th site of chain $\lambda$.
4401: By Fourier transforming,
4402: $S({\bf q})$ has been obtained for 
4403: $U=8t$, $\langle n\rangle =0.875$ and various values of 
4404: $t_{\perp}/t$.
4405: No obvious feature is found in $S({\bf q})$ 
4406: at the "$2{\bf k}_F$" wave vector
4407: of the 2-leg Hubbard ladder, which is ${\bf q^*}=(q^*,\pi)$ 
4408: with $q^*=\pi \langle n\rangle$. 
4409: On the other hand, 
4410: a feature is observed at the "$4{\bf k}_F$" wave vector, 
4411: which corresponds to $(\pi/4,0)$ for 
4412: $\langle n\rangle=0.875$.
4413: Especially for $t_{\perp}/t = 1.5$, 
4414: this feature becomes more obvious. 
4415: In order to isolate the "$4{\bf k}_F$" component of the 
4416: density correlations, a correlation function $N({\bf q})$ 
4417: involving four density operators has been calculated.
4418: This correlation function exhibits a clear peak at $(\pi/4,0)$ 
4419: for both $t_{\perp}/t=1.0$ and 1.5, 
4420: and in real space it decays as power law,
4421: while $S(i,j,\lambda)$ decays exponentially.
4422: 
4423: The results on the 2-leg Hubbard ladder show that, when doped, 
4424: this model exhibits simultaneously short-range AF correlations 
4425: and power-law decaying $d_{x^2-y^2}$-like superconducting and 
4426: "$4{\bf k}_F$" density correlations.
4427: For $t_{\perp}/t \sim 1.5$, 
4428: the superconducting correlations decay more slowly than the 
4429: "$4{\bf k}_F$" density correlations.
4430: The 2D Hubbard model also exhibits short-range AF and 
4431: $d_{x^2-y^2}$ superconducting correlations. 
4432: In addition, the QMC data on the 2D case seen in Section~4 imply 
4433: that the features found in the density susceptibility 
4434: $\Pi({\bf q},i\omega_m=0)$ might be related to the "$4{\bf k}_F$" 
4435: wave vectors rather than "$2{\bf k}_F$" for large 
4436: $U/t$.
4437: Hence, the 2-leg and the 2D Hubbard models, 
4438: when doped, appear to 
4439: have similar magnetic, superconducting and density properties. 
4440: Furthermore, in both cases, the single-particle dispersion 
4441: near the $(\pi,0)$ and $(0,\pi)$ points get flattened 
4442: by the many-body effects.
4443: These flat bands also seem to play a key role 
4444: in determining the strength of the pairing correlations 
4445: in both models.
4446: At this point, it is necessary to note that in order to compare 
4447: the 2-leg Hubbard model with the 2D case, one should not use 
4448: isotropic hopping, $t_{\perp}/t=1.0$, since in this case 
4449: the Fermi surface points are not connected by scatterings involving 
4450: ${\bf Q}\sim (\pi,\pi)$ momentum transfers.
4451: Instead, anisotropic hopping, where the bonding and the antibonding 
4452: Fermi surface points can be connected by $(\pi,\pi)$ scatterings,
4453: need to be used.
4454: 
4455: Finally, 
4456: an important question is whether the 2D Hubbard model 
4457: exhibits spin-charge separation as in the 1D case 
4458: [Kivelson {\it et al.} 1987, Anderson {\it et al.} 1987].
4459: While the 2D case cannot be resolved currently, 
4460: in the 2-leg Hubbard ladder 
4461: no indications of spin-charge separation 
4462: as in the 1D Hubbard model have been found
4463: [Noack {\it et al.} 1996].
4464: 
4465: \setcounter{equation}{0}\setcounter{figure}{0}
4466: \section{Discussion}
4467: 
4468: The numerical results discussed above point out that 
4469: the short-range AF spin fluctuations are responsible for the 
4470: $d_{x^2-y^2}$-wave superconducting correlations.
4471: The relation of the AF spin fluctuations to the
4472: $d_{x^2-y^2}$-wave superconductivity has been studied 
4473: using various diagrammatic approaches.
4474: Perhaps, 
4475: the most commonly used of these approaches
4476: is the fluctuation exchange (FLEX) approximation 
4477: [Bickers {\it et al.} 1989], 
4478: which self-consistently treats the fluctuations in the 
4479: magnetic, density and the pairing channels.
4480: The FLEX technique
4481: has been used for obtaining possible phase diagrams and the 
4482: estimates of the superconducting transition temperatures.
4483: In Section~8.1 below, 
4484: a comparison of the QMC data with the results of the FLEX 
4485: approach to the 2D Hubbard model will be carried out for the 
4486: single-particle and the pairing properties.
4487: The purpose of this comparison is to have an idea for the range 
4488: of applicability of the FLEX approximation, and see how it 
4489: should be extended further.
4490: Another reason for carrying out such a comparison is because 
4491: the Eliashberg type of calculations of the $T_c$'s using the 
4492: spin fluctuations for the cuprates [Monthoux and Pines 1992]
4493: or the similar self-consistent spin-fluctuation exchange calculations 
4494: [Moriya {\it et al.} 1990] are basically 
4495: at the same level with the FLEX approximation.
4496: Hence, 
4497: it is of interest to compare with the exact QMC calculations.
4498: 
4499: It will be seen that for $U=4t$, 
4500: the FLEX provides results in excellent 
4501: agreement with the QMC data on the density of states $N(\omega)$ 
4502: and the Bethe-Salpeter eigenvalues in the pairing channel
4503: at the temperatures where the QMC calculations are carried out, 
4504: as found earlier [Bickers {\it et al.} 1989].
4505: For stronger coupling $U=8t$, there are differences between the 
4506: FLEX and the QMC results.
4507: The correlated metallic band which develops by doping the 
4508: AF Mott-Hubbard insulator is not obtained within FLEX 
4509: and this affects the strength of the $d_{x^2-y^2}$-wave 
4510: pairing correlations.
4511: 
4512: Following this, 
4513: in Section~8.2, 
4514: other types of Monte Carlo approaches to the Hubbard model,
4515: in particular the variational and the projector 
4516: Monte Carlo algorithms, will be briefly discussed.
4517: In Section 8.3, 
4518: the results of the recent dynamical-cluster approximation 
4519: and the one-loop RG calculations for $d_{x^2-y^2}$ pairing 
4520: in the 2D Hubbard model will be discussed.
4521: There is also much interest in understanding 
4522: the unusual normal state properties of the cuprates 
4523: in the pseudogap regime. 
4524: The origin and the implications of the pseudogap
4525: are important issues in this field.
4526: In Section 8.4, 
4527: various calculations on the low-doping regime
4528: of the 2D Hubbard model will be discussed briefly
4529: and their results will be compared with the pseudogap
4530: seen in the cuprates.
4531: 
4532: The $t$-$J$ model which is closely related to the Hubbard model 
4533: has also drawn broad attention.  
4534: It is useful to compare the QMC and the DMRG results 
4535: on the Hubbard model
4536: with the numerical studies of the $t$-$J$ model.
4537: Below, 
4538: in Section~8.5, such a comparison will be given.  
4539: Here, the attention will be concentrated on the density 
4540: and the pairing properties since that is where there are 
4541: unresolved issues.
4542: A question of interest is whether in the ground state of the doped
4543: 2D Hubbard model there are special density correlations
4544: as seen in the $t$-$J$ model,
4545: for instance, phase separation or stripes,  
4546: and, if so, whether they would suppress the $d_{x^2-y^2}$-wave 
4547: pairing correlations observed at higher temperatures with the 
4548: QMC simulations.
4549: 
4550: In Section~8.6, what these numerical studies of the Hubbard model 
4551: imply for the mechanism of the $d_{x^2-y^2}$-wave superconductivity 
4552: seen in the high-$T_c$ cuprates will be discussed.
4553: The issue which will be addressed here is whether 
4554: the 2D Hubbard model, or some variation of it, is sufficient 
4555: for explaining why the values of the $T_c$'s found
4556: in the cuprates are so high.
4557: 
4558: \subsection{Comparisons with the fluctuation-exchange approach to 
4559: the 2D Hubbard model}
4560: 
4561: The FLEX approximation was used first by Bickers {\it et al.} 
4562: to study $d_{x^2-y^2}$-wave superconductivity in the 
4563: 2D Hubbard model 
4564: [Bickers {\it et al.} 1989, Bickers and White 1991].
4565: This approach self-consistently incorporates the fluctuations 
4566: in the magnetic, density and the pairing channels.
4567: It is an approximation around the band limit, and it is conserving 
4568: in the sense that the microscopic conservation laws for the particle 
4569: number, energy, and momentum are obeyed. 
4570: Within FLEX, it is found that the 
4571: $d_{x^2-y^2}$ pairing correlations are mediated  
4572: by the AF spin fluctuations.
4573: The phase diagram of the $U=4t$ Hubbard model within the 
4574: FLEX approximation is shown in Fig.~8.1.
4575: The FLEX calculations find a finite mean-field Neel temperature 
4576: at half-filling and at small dopings up to 6\%.
4577: Neighboring the SDW phase is a $d_{x^2-y^2}$-wave 
4578: superconducting phase which is stabilized for dopings 
4579: between 6\% and 20\%.
4580: Note that in the 2D Hubbard model, a finite Neel 
4581: temperature is not possible and only a Kosterlitz-Thouless type
4582: of superconducting transition can occur. 
4583: Hence, the finite transition temperatures seen in Fig.~8.1 
4584: represent transitions to regimes 
4585: where the $d_{x^2-y^2}$ correlations have a 
4586: power law decay at finite $T$.
4587: In order to induce true long-range order
4588: at finite temperature, 
4589: three dimensional couplings would be required.
4590: 
4591: \begin{figure}[ht]
4592: \centering
4593: \iffigure
4594: \epsfysize=8cm
4595: \epsffile[100 150 550 610]{ch8-fig1.ps}
4596: \fi
4597: \caption{
4598: Sketch of the 
4599: phase diagram of the 2D Hubbard model 
4600: within the FLEX approximation for $U=4t$.
4601: From [Bickers {\it et al.} 1989, Bickers and White 1991].
4602: }
4603: \label{8.1}
4604: \end{figure}
4605: 
4606: In addition, using the FLEX approximation 
4607: the effects of the nearest-neighbor hopping $t'$ 
4608: has been investigated for the 2D Hubbard model
4609: [Monthoux and Scalapino 1994, Dahm and Tewordt 1995].
4610: This approach has been also used for studying a three-band 
4611: CuO$_2$ model which has nearest neighbor copper-oxygen hopping
4612: and an onsite Coulomb repulsion at the Cu sites
4613: [Luo and Bickers 1993].
4614: The solutions of the FLEX equations in the 
4615: $d_{x^2-y^2}$-wave superconducting state were also 
4616: obtained [Pao and Bickers 1994 and 1995, Monthoux and Scalapino 1994, 
4617: Dahm and Tewordt 1995].
4618: 
4619: In the following, 
4620: a comparison of the FLEX results with the QMC data will be 
4621: carried out using results from Ref.~[Dahm and Bulut 1996].
4622: First the single-particle properties and then the pairing correlations 
4623: will be discussed.
4624: Figure~8.2 compares the QMC data on 
4625: the density of states $N(\omega)$ with the 
4626: FLEX results for various values of $U/t$ at 
4627: $\langle n\rangle=0.87$.
4628: While for $U=4t$, 
4629: the results from the two different approaches 
4630: are similar at these temperatures, 
4631: there are qualitative differences 
4632: for $U=8t$ and $12t$.
4633: The Fermi level within FLEX moves slower with the doping
4634: at large $U/t$.
4635: The correlated metallic band at the Fermi level
4636: as well as the lower and the upper Hubbard bands
4637: and the Mott-Hubbard pseudogap are not obtained within the 
4638: FLEX approximation.
4639: For $U=4t$ and $T=0.33t$, the FLEX and the QMC results 
4640: have similar qualitative features.
4641: However, at lower temperatures $N(\omega)$ could 
4642: develop a pseudogap for $U=4t$ also.
4643: Figure~8.3 shows the temperature evolution of $N(\omega)$ 
4644: for $U=8t$ and $\langle n\rangle=0.87$ within the 
4645: FLEX approximation. 
4646: Comparing this with Fig.~5.3, 
4647: one sees that the differences 
4648: with the QMC data continue to exist as $T$ is lowered.
4649: In Fig.~8.4, the QMC and the FLEX results on $N(\omega)$ 
4650: are compared at half-filling for $U=8t$ and $T=0.5t$.
4651: Here, it is seen that the development of the Mott-Hubbard 
4652: gap at half-filling is not obtained within FLEX.
4653: 
4654: \begin{figure}[ht]
4655: \centerline{
4656: \epsfysize=6cm \epsffile[-207 164 367 598]{ch8-fig2a.ps}
4657: \epsfysize=6cm \epsffile[18 164 592 598]{ch8-fig2b.ps}
4658: \epsfysize=6cm \epsffile[243 164 817 598]{ch8-fig2c.ps}}
4659: \caption{
4660: Comparison of the FLEX and the QMC results on $N(\omega)$ versus 
4661: $\omega$ at $\langle n\rangle =0.875$ 
4662: for (a) $U=4t$, (b) $U=8t$, and (c) $U=12t$.
4663: In (a) and (b) $T=0.33t$ was used and in (c) $T=0.5t$.
4664: Here, the vertical long-dashed and 
4665: the short-dashed lines denote the chemical potential
4666: for the FLEX and the QMC calculations, respectively.
4667: }
4668: \label{8.2}
4669: \end{figure}
4670: 
4671: \begin{figure}
4672: \centering
4673: \iffigure
4674: \epsfysize=8cm
4675: \epsffile[100 150 550 610]{ch8-fig3.ps}
4676: \fi
4677: \caption{
4678: Temperature evolution of $N(\omega)$ versus $\omega$ 
4679: within the FLEX approximation for $U=8t$ and 
4680: $\langle n\rangle=0.875$.
4681: Here, the vertical long-dashed line denotes the chemical potential.
4682: }
4683: \label{8.3}
4684: \end{figure}
4685: 
4686: \begin{figure}[ht]
4687: \centering
4688: \iffigure
4689: \epsfysize=8cm
4690: \epsffile[100 150 550 610]{ch8-fig4.ps}
4691: \fi
4692: \caption{
4693: Comparison of the FLEX and the QMC results on $N(\omega)$ versus 
4694: $\omega$ at half-filling 
4695: for $U=8t$ and $T=0.5t$.
4696: }
4697: \label{8.4}
4698: \end{figure}
4699: 
4700: Next, the FLEX results on the $d_{x^2-y^2}$-wave eigenvalue 
4701: $\lambda_d$ of the Bethe-Salpeter equation in the 
4702: particle-particle channel will be shown.
4703: In Fig.~8.5, 
4704: $\lambda_d$ is plotted as a function of $T/t$ for 
4705: $U/t=4$ and 8 in the temperature regime where 
4706: the QMC data exist.
4707: Here, it is seen that as $U/t$ increases from 4 to 8, 
4708: $\lambda_d$ changes by a small amount. 
4709: For $U=4t$, the FLEX results on $\lambda_d$ versus $T/t$ are in good 
4710: agreement with the QMC data. 
4711: However, for $U=8t$ the FLEX approximation underestimates 
4712: $\lambda_d$ by about a factor of two.
4713: This is the major difference between the QMC and 
4714: the FLEX results on $\lambda_d$.
4715: Within FLEX the effective 
4716: pairing interaction is also attractive in the odd-frequency 
4717: $s$ and $p$-wave channels,
4718: in addition to the even-frequency $d$-wave channel.
4719: 
4720: \begin{figure}[ht]
4721: \centering
4722: \iffigure
4723: \epsfysize=8cm
4724: \epsffile[100 150 550 610]{ch8-fig5.ps}
4725: \fi
4726: \caption{
4727: Temperature dependence of the $d_{x^2-y^2}$-wave 
4728: eigenvalue within the FLEX approximation 
4729: at $\langle n\rangle=0.875$ for $U=4t$ and $8t$.
4730: }
4731: \label{8.5}
4732: \end{figure}
4733: 
4734: Comparing Fig.~5.3 with Fig.~8.3, one sees that 
4735: the density of states at the Fermi level is  also underestimated 
4736: by about a factor of two within FLEX.
4737: From the simple spin-fluctuation exchange form of 
4738: Eq.~(2.1), one expects that the $d_{x^2-y^2}$-wave 
4739: pairing interaction increases with $U/t$.
4740: As discussed in Section 6.2, 
4741: the QMC data also shows that the reducible vertex $\Gamma_s$ 
4742: gets enhanced as $U/t$ is increased from 4 to 8, even though the 
4743: irreducible vertex $\Gamma_{Is}$ could not be obtained for 
4744: this value of $U/t$.
4745: Hence, it must be that the FLEX approach underestimates the 
4746: amount of the single-particle spectral weight at the Fermi 
4747: level, and this causes the $d_{x^2-y^2}$-wave 
4748: eigenvalue to be smaller with respect to the QMC value.
4749: 
4750: Here, it is seen that the correlated metallic band 
4751: which forms upon doping the AF Mott-Hubbard insulator 
4752: is not obtained within FLEX,
4753: which appears to cause why the FLEX 
4754: underestimates $\lambda_d$ for $U=8t$.
4755: Whether similar effects could take place when 
4756: a CDW insulator is doped is an interesting problem.
4757:  
4758: \begin{figure}[ht]
4759: \centering
4760: \iffigure
4761: \epsfysize=8cm
4762: \epsffile[100 150 550 610]{ch8-fig6.ps}
4763: \fi
4764: \caption{
4765: Superconducting transition temperature $T_c$ versus 
4766: $U/t$ within the FLEX approximation for $\langle n\rangle=0.875$.
4767: From [Pao and Bickers 1994].
4768: }
4769: \label{8.6}
4770: \end{figure}
4771: 
4772: Above, it has been seen that within FLEX $\lambda_d$ has a weak 
4773: dependence on $U/t$ at temperatures between $0.25t$ and $1.0t$.
4774: Related to this is Fig.~8.6 which shows the transition
4775: temperature $T_c$ obtained within FLEX as a function of $U/t$.
4776: Here, it is seen that $T_c$ has a broad peak at $U\sim 6t$.
4777: In this section, 
4778: it has been seen that the FLEX 
4779: underestimates the strength of the $d_{x^2-y^2}$-wave pairing 
4780: when $U$ is of order $8t$.
4781: An important question raised by these comparisons is 
4782: whether for $U\sim 8t$ 
4783: the $T_c$ should be higher than $0.025t$.
4784: There is a chance that the maximum possible $T_c$ in the 
4785: 2D Hubbard model is higher than the FLEX estimate.
4786: However, note that these values for $T_c$ are 
4787: just estimates for a possible Kosterlitz-Thouless transition. 
4788: In fact, 
4789: the ground state of the doped 2D Hubbard model is not known, 
4790: and it is currently beyond the reach 
4791: of the exact many-body calculations.
4792: 
4793: \subsection{Other Monte Carlo results on the 2D Hubbard model}
4794: 
4795: The QMC data presented in the previous sections
4796: were obtained with the determinantal 
4797: Monte Carlo technique [White {\it et al.} 1989b].
4798: Variational and projector Monte Carlo algorithms 
4799: were also used for studying the 2D Hubbard model.
4800: However,
4801: these approaches have arrived at different conclusions
4802: about the possibility of the $d_{x^2-y^2}$-wave pairing.
4803: In this section, some of these calculations will be reviewed briefly. 
4804: 
4805: In order to see whether superconducting long-range 
4806: order develops in the ground state of the doped 2D Hubbard model, 
4807: zero temperature projector Monte Carlo calculations 
4808: have been carried out 
4809: [Imada 1991, Furukawa and Imada 1992].
4810: In these calculations, 
4811: an optimised initial state was used 
4812: as a guiding function, mitigating the sign problem 
4813: but making the calculation variational rather than exact.
4814: With this technique, 
4815: the equal-time pair-field correlation functions 
4816: were calculated for the $s$, extended $s$ and 
4817: the $d_{x^2-y^2}$-wave singlet pairing channels.
4818: However, no size dependence was found in the data,
4819: which implies that there is no long-range superconducting order
4820: with these symmetries in the ground state.
4821: 
4822: The constrained-path Monte Carlo (CPMC) 
4823: algorithm developed in Ref.~[Zhang {\it et al.} 1995]
4824: is a variational method which 
4825: starts from a trial wave function 
4826: $| \Psi_T \rangle$ and uses the same $| \Psi_T\rangle$
4827: as a constraining wave function as the simulation 
4828: is carried out in the space of the Slater determinants.
4829: In calculations on up to $16\times 16$ lattices,
4830: free-electron and unrestricted Hartree-Fock wave functions 
4831: were used as $|\Psi_T \rangle$, 
4832: however no $d_{x^2-y^2}$-wave superconducting long-range 
4833: order was found [Zhang {\it et al.} 1997].
4834: The CPMC method has been extended by using as 
4835: the constraining function a $d_{x^2-y^2}$-wave 
4836: BCS wave function [Guerrero {\it et al.} 1999].
4837: Superconducting long-range order was not found in this case either.
4838: An important issue is how well the constraining function describes the 
4839: correlations in the Hubbard model.
4840: In fact, a comparison of the CPMC and the DMRG results, which will
4841: be discussed later in Section~8.5.1, has been carried out 
4842: for the magnetic and the density correlations in the 
4843: $12 \times 3$ Hubbard model [Bonca {\it et al.} 2000].
4844: This comparison has found that the CPMC results depend sensitively 
4845: on the constraining wave function especially for large $U/t \sim 8$.
4846: In this respect, it would be useful to test how well the CPMC 
4847: method describes the $d_{x^2-y^2}$-like 
4848: superconducting correlations found 
4849: in the 2-leg Hubbard ladder. 
4850: For instance, it is known from the DMRG calculations that 
4851: in the 2-leg Hubbard ladder for $U/t=8$,
4852: $t_{\perp}/t \sim 1.5$ and near half-filling the pair-field
4853: correlation function $D(\ell)$ decays as $\ell^{-\theta}$,
4854: where $\theta \alt 1.0$.
4855: It would be useful to see 
4856: whether the CPMC method describes the 
4857: power-law pairing correlations
4858: which exist in the 2-leg Hubbard ladder. 
4859: 
4860: Another variational Monte Carlo approach uses 
4861: as the trial wave function 
4862: [Nakanishi {\it et al.} 1997, Yamaji {\it et al.} 1998]
4863: \begin{equation}
4864: \label{PsiT}
4865: |\Psi_T\rangle = P_N P_G |\Psi_{BCS}\rangle,
4866: \end{equation}
4867: where
4868: \begin{equation}
4869: |\Psi_{BCS}\rangle = \Pi_{\bf k} 
4870: ( u_{\bf k} + v_{\bf k} 
4871: c^{\dagger}_{{\bf k}\uparrow}c^{\dagger}_{{-\bf k}\downarrow} )
4872: |0 \rangle 
4873: \end{equation}
4874: is the usual $d_{x^2-y^2}$-wave BCS wave function, 
4875: \begin{equation}
4876: P_G = \Pi_i [ 1-(1-g) n_{i\uparrow}n_{i\downarrow} ],
4877: \end{equation}
4878: is the Gutzwiller projection operator, 
4879: and $P_N$ is an operator which projects out states 
4880: with electron number $N$.
4881: In this method, the magnitude of the 
4882: $d_{x^2-y^2}$-wave gap function entering the usual BCS coefficients
4883: $u_{\bf k}$ and $v_{\bf k}$, the parameter $g$, and the 
4884: chemical potential $\mu$ are used as variational parameters,
4885: and their values are determined by minimising the ground 
4886: state energy with a Monte Carlo simulation.
4887: With this method it is found that,
4888: for fillings between 0.84 and 0.68, 
4889: the $d_{x^2-y^2}$-wave state has the lowest energy, 
4890: and for $\langle n \rangle \agt 0.84$ an SDW state is 
4891: favored.
4892: In addition, 
4893: when a next-near-neighbor hopping term $t'$ is turned on, 
4894: the superconducting condensation energy gets enhanced.
4895: For instance, for $U=8t$ this enhancement is maximum 
4896: when $t' \approx -0.1t$.
4897: It should be noted that this method was also applied
4898: to the 2-leg Hubbard ladder [Koike {\it et al.} 2000].
4899: In this case, 
4900: the gap function was assumed to take momentum-independent
4901: values $\Delta_1$ and $\Delta_2$ on the bonding and 
4902: the antibonding bands.
4903: Treating $\Delta_1$ and $\Delta_2$ as variational parameters,
4904: it was found that 
4905: a superconducting state with $d_{x^2-y^2}$-like symmetry 
4906: is stabilized when the bottom of the antibonding band 
4907: is near the Fermi level,
4908: which is in agreement with the exact-diagonalization 
4909: [Yamaji and Shimoi 1994] 
4910: and the DMRG [Noack {\it et al.} 1995 and 1997] calculations.
4911: 
4912: Next, a projector Monte Carlo approach 
4913: [Husslein {\it et al.} 1996]
4914: will be discussed where the ground state wave function 
4915: of the 2D Hubbard model is estimated from 
4916: \begin{equation}
4917: |\Psi_g \rangle = e^{\theta H} |\Psi_0 \rangle .
4918: \end{equation}
4919: Here, $|\Psi_0 \rangle$ is the ground state of the 
4920: noninteracting electrons and 
4921: the parameter $\theta$ is taken to be 
4922: about 8 on a $12 \times 12$ lattice.
4923: This approach has been used to calculate the 
4924: $d_{x^2-y^2}$-wave pair-field correlation function 
4925: in the weak-coupling regime, $0.5t < U < 3t$,
4926: near half-filling.
4927: It is found that 
4928: the system has long-range $d_{x^2-y^2}$-wave pairing 
4929: correlations for negative values of $t'$, 
4930: for instance, at $t'= -0.3t$ for $U=2t$.
4931: 
4932: In this section, 
4933: it was seen that various approaches arrive at different 
4934: conclusions about $d_{x^2-y^2}$-wave pairing 
4935: in the 2D Hubbard model.
4936: This emphasises the importance of the choice of 
4937: the trial wave functions and points out 
4938: at the need for carrying out calculations 
4939: without introducing approximations.
4940: 
4941: \subsection{Dynamical cluster approximation and 
4942: RG calculations for $d_{x^2-y^2}$
4943: pairing in the 2D Hubbard model}
4944: 
4945: Recently, 
4946: the dynamical cluster approximation 
4947: [Maier {\it et al.} 2000] 
4948: and the one-loop renormalization-group method 
4949: employing a 2D Fermi surface
4950: [Zanchi and Schulz 2000, Halboth and Metzner 2000, 
4951: Honerkamp {\it et al.} 2001] 
4952: were used for studying $d_{x^2-y^2}$ pairing in the 2D Hubbard model.
4953: In this section, 
4954: the findings of these studies will be discussed briefly. 
4955: 
4956: When the infinite dimensions limit of the Hubbard model 
4957: is taken with proper scaling, 
4958: the many-body problem becomes local 
4959: [Metzner and Vollhardt 1989, M\"uller-Hartman 1989], 
4960: and it can be mapped to an Anderson impurity problem, 
4961: which can then be solved with various many-body techniques 
4962: [Pruschke {\it et al.} 1995, Georges {\it et al.} 1996].
4963: This is called the dynamical mean-field approximation (DMFA). 
4964: The DMFA is interesting because it is a strong coupling technique.
4965: For instance, 
4966: for large $U/t$ and off of half-filling, 
4967: the DMFA yields a narrow peak in 
4968: $N(\omega)$ near the Fermi level [Jarrell 1992], 
4969: which is also seen in the 2D QMC data.
4970: However, 
4971: the DMFA does not incorporate the non-local correlations, 
4972: and hence it is not 
4973: possible to study $d_{x^2-y^2}$ pairing with it. 
4974: The DCA incorporates the non-local corrections to DMFA by 
4975: mapping the lattice problem onto an embedded cluster of size
4976: $N_c$, rather than onto an impurity problem. 
4977: In DCA calculations, 
4978: the dynamical correlation length is restricted to 
4979: $L=\sqrt{N_c}$, 
4980: and the DCA would become exact for 
4981: $N_c \rightarrow \infty$, 
4982: while it reduces to DMFA for $N_c=1$.
4983: With this approach, 
4984: the mean-field $d_{x^2-y^2}$ superconducting $T_c$'s 
4985: were calculated for $N_c=4$ 
4986: by using the non-crossing approximation to 
4987: solve the cluster problem [Maier {\it et al.} 2000].
4988: For $U=12t$, 
4989: it is found that  $T_c$ has the maximum value of $\approx 0.05t$ 
4990: for about 20\% doping. 
4991: It is also found that $T_c$ increases for positive values of 
4992: the second-nearest-neighbour hopping $t'$, 
4993: and decreases for negative values of $t'$.
4994: This result agrees with the DMRG calculations 
4995: on the $t$-$J$ model which find $d_{x^2-y^2}$ pairing 
4996: for $t'>0$ 
4997: [White and Scalapino 1998a].
4998: With DCA, 
4999: the low doping regime of the 2D Hubbard model was also studied, 
5000: and these results will be discussed in Section 8.4.
5001: 
5002: After the discovery of the high $T_c$ cuprates, 
5003: the one-loop RG approach was extended from 
5004: 1D to 2D in order to study this problem
5005: [Dzyaloshinskii 1987, Schulz 1987, Lederer {\it et al.} 1987].
5006: The RG calculations are interesting,
5007: because with this technique the particle-particle 
5008: and the particle-hole channels are treated on equal footing.
5009: These studies focused on the scattering processes between 
5010: the Fermi surface regions near the van Hove singularities.
5011: For the 2D Hubbard model, 
5012: the AF SDW state at half-filling and 
5013: the $d_{x^2-y^2}$ superconducting state induced 
5014: by AF fluctuations away from half-filling were found
5015: [Schulz 1987, Lederer {\it et al.} 1987].
5016: The scatterings involving the full 2D Fermi surface were taken 
5017: into account with the work of Zanchi and Schulz, 
5018: who studied the RG flows using 
5019: a 32-patch discretization of the 2D Fermi surface for $t'=0$ 
5020: [Zanchi and Schulz 2000].
5021: These calculations found two different regimes, 
5022: one dominated by the AF correlations 
5023: and the other by the $d_{x^2-y^2}$ pairing. 
5024: The 2D RG calculations were also extended to the $t' \neq 0$ case
5025: [Halboth and Metzner 2000].
5026: In the calculations for $t'\neq 0$ by Honerkamp {\it et al.}, 
5027: an intermediate regime is found between 
5028: the two regimes dominated by the 
5029: AF correlations and $d_{x^2-y^2}$ pairing
5030: [Honerkamp {\it et al.} 2001].
5031: In this intermediate regime, 
5032: there are competing AF and $d_{x^2-y^2}$ 
5033: superconducting correlations.
5034: This regime exists only for sizeable $t'<0$ and 
5035: it exhibits features which are similar to those seen 
5036: in the pseudogap regime of the cuprates,
5037: which will be discussed in the following section.
5038: Honerkamp {\it et al.} extracted a temperature versus 
5039: doping phase diagram from the 
5040: 2D RG flows for the 2D Hubbard model with $t'<0$, 
5041: which is similar to that of the cuprates. 
5042: 
5043: In spite of these interesting results 
5044: it has to be kept in mind that 
5045: the dynamical correlation length in DCA 
5046: is cut-off by the size of the cluster 
5047: and the one-loop RG is a weak-coupling approach.
5048: In addition, 
5049: in RG calculations the single-particle self-energy corrections 
5050: are not included at the one-loop level.
5051: In Section 8.1, 
5052: it was seen that the single-particle self-energy corrections
5053: could play an important role 
5054: in determining the strength of pairing. 
5055: It should also be noted that, 
5056: while the RG finds that the regime with $t'<0$ is favored, 
5057: in the DCA calculations the mean-field $T_c$'s 
5058: are higher for $t'>0$.
5059: These are some of the reasons for why exact results are necessary 
5060: in order to reach a final conclusion. 
5061: 
5062: \subsection{Low-doping regime of the 2D Hubbard model}
5063: 
5064: In the normal state of the underdoped cuprates, 
5065: a pseudogap is seen in the excitation spectrum 
5066: below a temperature $T^*$ which depends on the doping.
5067: The nature of the pseudogap is an important problem 
5068: in this field and there exist a wide range 
5069: of ideas about its origin and its implications
5070: [M$^2$S-HTSC VI proceedings 2000].
5071: In this section, 
5072: various calculations which have been carried out for 
5073: exploring whether there is a pseudogap regime 
5074: in the 2D Hubbard model will be discussed. 
5075: In Ref.~[Jaklic and Prelovsek 2000], 
5076: a review of the numerical calculations 
5077: on the $t$-$J$ model at finite temperatures is given 
5078: and the results of these calculations are compared with the 
5079: anomalous normal-state properties of the cuprates 
5080: including the pseudogap.
5081: 
5082: The normal state pseudogap, 
5083: while exhibiting dependence on the material properties, 
5084: is seen in the uniform magnetic susceptibility,
5085: the low-frequency optical conductivity, the ARPES spectrum 
5086: and various other measurements of the electronic properties. 
5087: An interesting feature of the normal state 
5088: pseudogap observed in the ARPES 
5089: spectrum is that it is anisotropic on the Fermi surface: 
5090: the pseudogap has maximum amplitude near 
5091: the $(\pi,0)$ and $(0,\pi)$ points of the Brillouin zone 
5092: and it is minimum for wave vectors near $(\pi/2,\pi/2)$
5093: [Ding {\it et al.} 1996, Ronning {\it et al.} 1998].
5094: 
5095: Kampf and Schrieffer first discussed 
5096: the possibility of a pseudogap within 
5097: the spin-bag approach for the 2D Hubbard model 
5098: [Kampf and Schrieffer 1990a and 1990b]. 
5099: They showed how a pseudogap could develop because of the 
5100: AF fluctuations already at the one-loop level 
5101: for the single-particle self-energy 
5102: as the AF instability is approached. 
5103: In the FLEX calculations, 
5104: it was also found that a pseudogap opens in the 
5105: density of states when the strength 
5106: of the short-range AF fluctuations
5107: increases [Dahm and Tewordt 1995].
5108: 
5109: In spite of the sign problem, 
5110: there are QMC data on 
5111: $A({\bf p},\omega)$ at temperatures as low as $0.25t$ 
5112: in the underdoped regime
5113: [Preuss {\it et al.} 1997].
5114: These data show indications that a pseudogap opens 
5115: in $A({\bf p},\omega)$ as $T$ decreases. 
5116: In particular, 
5117: for $\langle n\rangle \simeq 0.95$, 
5118: it is seen that spectral weight 
5119: above the Fermi level 
5120: for ${\bf p}$ between $(\pi,0)$ and $(\pi,\pi)$ 
5121: decreases gradually as $T$ decreases from $0.5t$ to $0.25t$. 
5122: In these calculations, 
5123: the pseudogap has been attributed to the AF spin correlations, 
5124: which are becoming larger than a lattice spacing 
5125: for $T < 0.3t$ in the underdoped regime. 
5126: However, 
5127: in order to be able to make direct comparisons with 
5128: the experimental data on the pseudogap, 
5129: it would be necessary to reach temperatures below $0.1t$. 
5130: 
5131: In recent one-loop RG calculations 
5132: for the Hubbard model with $t'<0$,
5133: a saddle-point regime is found between 
5134: the AF ordered and the $d_{x^2-y^2}$ superconducting regimes 
5135: [Honerkamp {\it et al.} 2001].
5136: This regime exhibits features similar 
5137: to those seen in the pseudogap regime of the cuprates.
5138: Here, 
5139: the uniform magnetic and charge susceptibilities 
5140: flow to zero because of the 
5141: pairing and the umklapp-scattering processes, 
5142: respectively. 
5143: The possible existence of an umklapp-gapped 
5144: spin-liquid phase was suggested in earlier RG calculations 
5145: where a two-patch discretization 
5146: of the Brillouin zone was used  
5147: [Furukawa {\it et al.} 1998].
5148: The similarity of this regime to the spin-gapped phase 
5149: in the 2-leg Hubbard ladder was also noted. 
5150: Because of the umklapp scatterings, in this regime, 
5151: the sections of the Fermi surface which are 
5152: near the $(\pi,0)$ and $(0,\pi)$ points 
5153: are truncated 
5154: while at wave vectors near $(\pi/2,\pi/2)$ 
5155: gapless single-particle excitations remain. 
5156: Clearly, 
5157: these results are useful for interpreting a number of 
5158: experimental data on the underdoped cuprates 
5159: and, in particular, the ARPES data.
5160: 
5161: The DCA calculations have also found interesting results 
5162: about this subject
5163: [Huscroft {\it et al.} 2001].
5164: Using this method, 
5165: the low-doping and the low-temperature regime 
5166: of the 2D Hubbard model was studied.
5167: Here, 
5168: QMC simulations were used to solve 
5169: the embedded lattice problem 
5170: with $N_c=8$.
5171: It was found that, as $T$ is lowered, 
5172: an anisotropic pseudogap develops in 
5173: $A({\bf p},\omega)$ and, in addition, 
5174: the uniform magnetic susceptibility $\chi(T)$ gets suppressed.
5175: In particular, 
5176: for $\langle n\rangle =0.95$, $U=6t$ and $T=0.06t$,
5177: a pseudogap is found in $A({\bf p},\omega)$ 
5178: for ${\bf p}=(\pi,0)$ and $(0,\pi)$ 
5179: and not for $(\pi/2,\pi/2)$. 
5180: For these values of $U/t$ and $\langle n\rangle$, 
5181: a mean-field $d_{x^2-y^2}$ superconducting transition temperature of 
5182: $T_c\simeq 0.04t$ is obtained with the DCA. 
5183: In addition, 
5184: a sharp drop in $\chi(T)$ is observed near the temperature 
5185: where the pseudogap in $A({\bf p},\omega)$ opens. 
5186: In these calculations, 
5187: the downturn in $\chi(T)$ has been attributed to the development 
5188: of the AF correlations. 
5189: 
5190: These results support the idea that perhaps
5191: the normal state pseudogap seen in the cuprates 
5192: could be understood within a 2D Hubbard framework.
5193: However, 
5194: there still exist a wide range of ideas 
5195: about the origin of the pseudogap,
5196: and it is one of the important unresolved issues 
5197: in this field.
5198: 
5199: \subsection{Comparisons with the $t$-$J$ model} 
5200: 
5201: In the large $U/t$ limit, the Hubbard model reduces to 
5202: [Hirsch 1985]
5203: \begin{eqnarray}
5204: \label{tJ}
5205: H = -t \sum_{\langle i,j\rangle, \sigma}
5206: ( c^{\dagger}_{i,\sigma} c_{j,\sigma} + 
5207: c^{\dagger}_{j,\sigma} c_{i,\sigma} )
5208: + J \sum_{\langle i,j\rangle} 
5209: ( {\bf S_i}\cdot {\bf S_j} - {1\over 4} n_in_j ) \nonumber \\ 
5210: -{J\over 4} \sum_{i,\sigma} 
5211: \sum_{\delta\neq\delta'} 
5212: (   c^{\dagger}_{i+\delta,\sigma} c^{\dagger}_{i,-\sigma} 
5213: c_{i,-\sigma} c_{i+\delta',\sigma} 
5214: -
5215: c^{\dagger}_{i+\delta,-\sigma} c^{\dagger}_{i,\sigma} 
5216: c_{i,-\sigma} c_{i+\delta',\sigma}   ),
5217: \end{eqnarray}
5218: where $J=4t^2/U$, and the double occupancy of a site is not 
5219: allowed.
5220: In Eq.~(\ref{tJ}), 
5221: $i+\delta$ and $i+\delta'$ sum over the nearest neighbors
5222: of site $i$.
5223: The last sum in this expression, which involves 
5224: operators acting at three different sites, is dropped 
5225: and what is remaining is called the $t$-$J$ model.
5226: Clearly, the $t$-$J$ and the Hubbard models
5227: have differences.
5228: The numerical studies of the $t$-$J$ model 
5229: have produced valuable information about the magnetic, density
5230: and the superconducting correlations in this system.
5231: Reviews of these studies can be found in 
5232: Refs.~[Dagotto 1994, Jaklic and Prelovsek 2000]. 
5233: In this section, 
5234: the density and the pairing correlations 
5235: in the $t$-$J$ and the Hubbard models will be compared.
5236: 
5237: In Section 8.5.1, the results on the phase separation and 
5238: the density correlations in the 2D $t$-$J$ model obtained 
5239: with various numerical techniques will be discussed.
5240: These data will be compared with the CPMC and the DMRG
5241: calculations for the $12\times 3$ Hubbard lattice.
5242: In Section 8.5.2, the nature of the pairing correlations seen 
5243: in the 2D $t$-$J$ model will be compared with those 
5244: in the 2D Hubbard model.
5245: The results on the 2-leg $t$-$J$ and Hubbard ladders will be 
5246: compared in Section 8.5.3.
5247: 
5248: \subsubsection{Comparisons with the density correlations in the 
5249: 2D $t$-$J$ model} 
5250: 
5251: The QMC simulations [Moreo and Scalapino 1991] and 
5252: the exact-diagonalization [Dagotto {\it et al.} 1992b]
5253: calculations on the 2D Hubbard model
5254: did not find any indication of phase separation of the system 
5255: into hole-rich and hole-poor regions.
5256: The 2-leg Hubbard ladder does not show any evidence 
5257: for phase separation either
5258: [Noack {\it et al.} 1994].
5259: In contrast to the Hubbard model, 
5260: the 2D $t$-$J$ model phase 
5261: separates at any electron filling for sufficiently large
5262: values of the interaction strength $J/t$.
5263: Various techniques including the variational
5264: and the exact-diagonalization calculations [Emery {\it et al.} 1990],
5265: the high-temperature series expansions [Putikka {\it et al.} 1992],
5266: and the exact-diagonalization 
5267: calculations [Dagotto 1994]
5268: were used for studying phase separation in the 
5269: $t$-$J$ model. 
5270: Recently, 
5271: the Green's function Monte Carlo (GFMC) 
5272: [Hellberg and Manousakis 1997, 1999 and 2000, 
5273: Calandra {\it et al.} 1998]
5274: and 
5275: the DMRG [White and Scalapino 1998a, 
5276: Rommer {\it et al.} 2000]
5277: were used for obtaining the phase-separation boundary 
5278: in the $t$-$J$ model.
5279: Currently, 
5280: where the true phase-separation boundary lies is controversial,
5281: since the approaches noted above arrive at different conclusions. 
5282: 
5283: \begin{figure}
5284: \centering
5285: \iffigure
5286: \epsfysize=8cm
5287: \epsffile[100 150 550 610]{ch8-fig7.ps}
5288: \fi
5289: \caption{
5290: Sketch of the phase-separation boundary for the 2D $t$-$J$
5291: model obtained from various calculations.
5292: The solid line represents the phase-separation boundary 
5293: obtained by the high-temperature series expansion 
5294: [Putikka {\it et al.} 1992].
5295: The open circles represent the results 
5296: of the GFMC calculations 
5297: from Ref.~[Calandra {\it et al.} 1998].
5298: The dashed line is from Ref.~[Hellberg and Manousakis 1997],
5299: which was also obtained by the GFMC.
5300: The arrow indicates the DMRG result 
5301: for the critical value of 
5302: $J/t$ for phase separation on a six-leg $t$-$J$ ladder
5303: in the limit of zero doping
5304: [Rommer {\it et al.} 2000].
5305: }
5306: \label{8.7}
5307: \end{figure}
5308: 
5309: Emery {\it et al.} suggested that the 2D $t$-$J$ model phase separates 
5310: at all interaction strengths
5311: [Emery {\it et al.} 1990].
5312: Subsequent calculations found that the phase separation 
5313: occurs only for $J \agt t$, 
5314: as indicated by the solid curve in Fig.~8.7
5315: [Putikka {\it et al.} 1992].
5316: However, recently, 
5317: the GFMC calculations [Calandra {\it et al.} 1998] 
5318: suggested that the phase-separation boundary 
5319: might occur at lower $J/t$ values 
5320: near half-filling. 
5321: This is indicated by the empty circles in Fig.~8.7.
5322: In these calculations, 
5323: next to the phase-separation boundary lies the 
5324: regime where the doped holes form 
5325: $d_{x^2-y^2}$-wave pairs. 
5326: The recent GFMC calculations by Hellberg and Manousakis,
5327: on the other hand,
5328: find that the critical value of $J/t$ for phase separation 
5329: extrapolates to zero at low dopings 
5330: as shown by the dashed curve in Fig.~8.7
5331: [Hellberg and Manousakis 1997, 1999 and 2000].
5332: In these calculations, the $t$-$J$ model phase separates 
5333: in the parameter regime appropriate for the cuprates. 
5334: For instance, at 15\% hole doping the critical value of 
5335: $J/t$ is about 0.4.
5336: The DMRG calculations find that, 
5337: in the physically relevant regime,
5338: the ground state of the $t$-$J$ model
5339: is striped and not phase separated
5340: [White and Scalapino 1998a, 2000].
5341: The DMRG calculations are carried out on large lattices
5342: with periodic boundary conditions at the long edges and 
5343: open boundary conditions at the short edges.
5344: In these calculations the phase separation occurs for 
5345: $J > t$.
5346: For instance, in the six-leg Hubbard ladder, 
5347: the critical value of $J/t$ is $\sim 1.4$
5348: as the doping approaches zero,
5349: which is indicated by the arrow in Fig.~8.7.
5350: 
5351: The stripes observed in the DMRG calculations 
5352: are a domain wall of holes 
5353: across which there is a $\pi$-phase shift in the AF background.
5354: According to the DMRG calculations, 
5355: the stripe formation represents an instability of the 
5356: system where pairs of bound holes 
5357: combine to form domain walls.
5358: In particular, through the stripe correlations the system
5359: reduces the frustration of the AF background and lowers
5360: the kinetic energy of the holes.
5361: However, the issue of 
5362: the phase separation versus the stripe formation 
5363: in the 2D $t$-$J$ model remains controversial.
5364: The role of the open boundary conditions in producing 
5365: the static stripes was questioned
5366: [Hellberg and Manousakis 2000],
5367: and the need to have more than one pair of holes in order to 
5368: see the stripe correlations was noted
5369: [White and Scalapino 2000].
5370: The calculation of the dynamical spin and charge
5371: susceptibilities for sufficiently large $t$-$J$ systems would resolve 
5372: these issues.
5373: 
5374: The possibility of stripe formation in strongly 
5375: correlated systems were noted in mean-field calculations of the 2D 
5376: Hubbard model soon after the discovery of the cuprates
5377: [Zaanen and Gunnarson 1989, Poilblanc and Rice 1989, 
5378: Schulz 1989, Machida 1989].
5379: There has been a surge of interest in this field 
5380: since the observation of static stripe ordering 
5381: in the neutron scattering experiments on 
5382: La$_{1.6-x}$Nd$_{0.4}$Sr$_x$CuO$_4$ [Tranquada {\it et al.} 1995].
5383: Beyond mean-field, the nature of 
5384: the stripe correlations in the Hubbard model 
5385: were studied with the DMRG and 
5386: the CPMC techniques for the 3-leg Hubbard ladder 
5387: [Bonca {\it et al.} 2000].
5388: The findings of these calculations are important and 
5389: they will be briefly described here.
5390: 
5391: The CPMC is an approximate method which projects 
5392: the ground state from a trial state 
5393: as has been discussed in Section~8.2.
5394: Both the DMRG and the CPMC calculations find that for 
5395: $U \agt 6t$ there are static stripes in the ground state of
5396: a $12 \times 3$ Hubbard lattice doped with six holes 
5397: when open boundary conditions are used.
5398: In Figures 8.8 and 8.9, the rung magnetization 
5399: \begin{equation}
5400: S^z(i) = \sum_{j=1}^3 \langle S^z(i,j) \rangle
5401: \end{equation}
5402: and the rung hole density
5403: \begin{equation}
5404: \rho(i) = \sum_{j=1}^3 \langle \rho(i,j) \rangle
5405: \end{equation}
5406: are plotted as a function of the rung location $i$
5407: at $U=8t$.
5408: Here, $S^z(i,j)$ and $\rho(i,j)$ are the spin and hole-density 
5409: operators at the $j$'th site of the $i$'th rung. 
5410: In Fig.~8.8, it is seen that 
5411: between rungs 3 and 4 the spins are ferromagnetically coupled,
5412: causing a $\pi$-phase shift, and the magnetisation density is small.
5413: The same occurs at rungs 9 and 10.
5414: In Fig.~8.9, it is seen that at these sites the holes form domain walls.
5415: In these figures the results of the unrestricted Hartree-Fock (UHF)
5416: calculations are also shown.
5417: The UHF approximation produces results in agreement with the 
5418: CPMC and the DMRG calculations when the reduced value of 
5419: $U=3t$ is used.
5420: This behavior is similar in spirit to using a reduced 
5421: Coulomb repulsion within RPA for the 
5422: spin susceptibility as discussed in Section~3.3.
5423: It is important to note that the structure of the stripes seen in these
5424: calculations for $U \agt 6t$ is quite similar to those found in the 
5425: three-leg $t$-$J$ ladder [White and Scalapino 1998b].
5426: However, for $U< 6t$ these features disappear and no evidence 
5427: is found for static stripes in the ground state.
5428: In this regime, only some evidence for low-lying states 
5429: with stripes are found.
5430: At weak $U/t$, the density of virtual holes due 
5431: to the double occupancy increases,
5432: and, it has been noted that this might 
5433: weaken the stripe correlations.
5434: At this point, 
5435: it is of interest to explore whether there is a 
5436: relation between the stripe patterns 
5437: which are observed within the 
5438: presence of the open boundary conditions and 
5439: any possible "4${\bf k}_F$" CDW correlations.
5440: It should also be noted that the CPMC method does not 
5441: show evidence for stripe correlations 
5442: even at large $U/t$ when the free-electron 
5443: trial wave function is used instead of the UHF wave function.
5444: Hence, caution is necessary in choosing a trial wave function 
5445: which is suitable for the ground state. 
5446: 
5447: \begin{figure}
5448: \centering
5449: \iffigure
5450: \epsfig{file=ch8-fig8.eps,height=8cm,angle=-90}
5451: \fi
5452: \caption{
5453: Rung-magnetisation versus the 
5454: rung location for the $12 \times 3$ Hubbard lattice doped
5455: with six holes. 
5456: From [Bonca {\it et al.} 2000]. 
5457: }
5458: \label{8.8}
5459: \end{figure}
5460: 
5461: \begin{figure}
5462: \centering
5463: \iffigure
5464: \epsfig{file=ch8-fig9.eps,height=8cm,angle=-90}
5465: \fi
5466: \caption{
5467: Rung-hole density versus the 
5468: rung location for the $12 \times 3$ Hubbard lattice doped
5469: with six holes.
5470: From [Bonca {\it et al.} 2000]. 
5471: }
5472: \label{8.9}
5473: \end{figure}
5474: 
5475: These results were found using open boundary conditions.
5476: The DMRG calculations are not carried out with periodic 
5477: boundary conditions, but the CPMC calculations can be.
5478: When carried out with the periodic boundary 
5479: conditions on square lattices, 
5480: the CPMC calculations do not find 
5481: stripes in the hole and spin densities and in the hole and 
5482: spin correlation functions.
5483: This indicates that the static stripe pattern seen in the 
5484: ground state results from the open boundary conditions breaking 
5485: the translational symmetry.
5486: With the periodic boundary conditions, there might be low-lying 
5487: states with stripe patterns but it might be difficult to detect them 
5488: numerically.
5489: Hence, whether static stripes exist in the ground state 
5490: of the 2D Hubbard model with periodic boundary conditions 
5491: remains as an important open question.
5492: Here and in Section~4, it has been seen that there are similarities
5493: in the density correlations of the $t$-$J$ model and the Hubbard 
5494: model with large $U/t$.
5495: But, there are differences as well.
5496: 
5497: \subsubsection{Comparisons with the superconducting correlations in the 
5498: 2D $t$-$J$ model} 
5499: 
5500: When two-holes are doped into the half-filled $t$-$J$ model, 
5501: they form a $d_{x^2-y^2}$-wave bound pair for interaction strengths
5502: relevant to the high-$T_c$ cuprates,
5503: $J/t\sim 0.35$ [Poilblanc 1993, Dagotto 1994].
5504: However, much controversy exists over what happens when more 
5505: than one pair of holes is doped.
5506: The GFMC calculations by Calandra {\it et al.} find that they form
5507: a $d_{x^2-y^2}$-wave superconducting ground state 
5508: for $J/t \sim 0.35$ and dopings relevant to the 
5509: high-$T_c$ cuprates [Calandra {\it et al.} 1998].
5510: However,
5511: the GFMC calculations by Hellberg and Manousakis 
5512: find that this parameter regime lies right at the boundary for 
5513: the phase separation of holes
5514: [Hellberg and Manousakis 1997 and 1999].
5515: 
5516: The DMRG calculations find that in the same regime 
5517: the ground state has static stripes, and the system does not 
5518: have $d_{x^2-y^2}$-wave superconducting long-range order
5519: [White and Scalapino 1998a, 1999].
5520: The fact that the static stripes compete with the superconductivity 
5521: is an important result of these calculations.
5522: However, when the static stripe patterns are broken by a 
5523: second-near-neighbor hopping $t' > 0$, it is seen that 
5524: the system develops long-range 
5525: $d_{x^2-y^2}$-wave superconducting order,
5526: and in this regime the low-lying stripe
5527: correlations coexist with the long-range superconducting order.
5528: 
5529: Both the $t$-$J$ and the Hubbard models exhibit
5530: $d_{x^2-y^2}$-wave pairing correlations due 
5531: to the AF spin fluctuations.
5532: However, there are differences between these models, 
5533: one of them being the no-double-occupancy constraint
5534: in the $t$-$J$ model.
5535: It is possible that this constraint is more favorable 
5536: of the stripe correlations 
5537: and, because of this, the 
5538: one-band Hubbard or a three-band CuO$_2$ model
5539: might have weaker tendency for stripe formation
5540: than the $t$-$J$ model
5541: [Daul {\it et al.} 2000, Jeckelmann {\it et al.} 1998].
5542: 
5543: Whether the doped 2D Hubbard model develops 
5544: static stripe patterns in the ground state is 
5545: an important issue.
5546: In the previous section, we have seen from the QMC data that 
5547: the doped 2D Hubbard model has $d_{x^2-y^2}$-wave  
5548: pairing correlations which develop as the temperature is lowered 
5549: and they are not weak correlations.
5550: However, if indeed a static striped phase 
5551: is more favored in the ground state 
5552: then this could suppress the 
5553: growth of the superconducting correlations at a temperature
5554: lower than where the QMC simulations are carried out.
5555: In this respect, the comparison of the CPMC and 
5556: the DMRG calculations for the $12\times 3$ Hubbard 
5557: lattice are useful
5558: [Bonca {\it et al.} 2000].
5559: However, further study is necessary before reaching conclusions 
5560: about the ground state of the doped 2D Hubbard model.
5561: 
5562: The results shown here emphasize the interplay of the 
5563: magnetic, density and the pairing correlations 
5564: in the Hubbard and $t$-$J$ models. 
5565: In the cuprates, 
5566: the substitution of nonmagnetic impurities
5567: gives useful information about the interplay of these
5568: correlations in these materials. 
5569: It is known that,
5570: when substituted in place of planar Cu, 
5571: nonmagnetic impurities strongly suppress the superconducting $T_c$ 
5572: [Xiao {\it et al.} 1990] 
5573: and locally enhance the AF correlations
5574: [Mahajan {\it et al.} 1994 and 2000].
5575: To the extend that a nonmagnetic impurity can be 
5576: considered as a pure potential scatterer, 
5577: these experimental data give information about 
5578: the magnetic and pairing response of these materials 
5579: when perturbed in the density channel.
5580: 
5581: \subsubsection{Comparisons with the 2-leg $t$-$J$ ladder} 
5582: 
5583: The Lanczos calculations on the 2-leg $t$-$J$
5584: ladder showed that this system, 
5585: which has a spin gap in the insulating state, 
5586: can exhibit superconducting correlations upon doping
5587: [Dagotto {\it et al.} 1992a].
5588: The mean-field calculations noted the $d_{x^2-y^2}$-like 
5589: structure of the superconducting pairs 
5590: in the $t$-$J$ ladder [Gopalan {\it et al.} 1994].
5591: The DMRG calculations on long $t$-$J$ ladders 
5592: established that in the ground state there are power-law decaying 
5593: superconducting correlations 
5594: [Hayward {\it et al.} 1995].
5595: In Ref.~[Dagotto and Rice 1996],
5596: the properties of $n$-leg spin-$1/2$ ladders are reviewed 
5597: and comparisons are made with various ladder compounds.
5598: 
5599: There are differences in the nature of the $d_{x^2-y^2}$-like
5600: superconducting correlations seen in the $t$-$J$ and the Hubbard 
5601: ladders.
5602:  For instance, 
5603: in the Hubbard ladder for $t_{\perp}/t \sim 1.0$, 
5604: the superconducting correlations are weak,
5605: and the pair-field correlation function 
5606: $D(\ell)$ decays as $\ell^{-2}$, which is like the $U=0$ case
5607: [Noack {\it et al.} 1994].
5608: Only when the distribution of the single-particle spectral weight 
5609: is such that it can make use of the momentum structure 
5610: in $\Gamma_I$, does the system 
5611: exhibit strong pairing correlations.
5612: On the other hand, 
5613: in the isotropic $t$-$J$ ladder, 
5614: the system has strong pairing correlations.
5615: For instance,
5616: $D(\ell)$ decays approximately as $\ell^{-1}$
5617: for $\langle n\rangle=0.8$ and $J/t=1$
5618: [Hayward {\it et al.} 1995].
5619: In addition, 
5620: Schulz has shown that $D(\ell)$ decays as 
5621: $\ell^{-1/2}$ in the limit $\langle n\rangle \rightarrow 1$
5622: [Schulz 1999].
5623: Hence, for isotropic hopping the $t$-$J$ ladder has 
5624: stronger pairing correlations compared to the 
5625: Hubbard ladder.
5626: Probably, this is related to the fact that 
5627: in the $t$-$J$ model the exchange term $J$,
5628: which is an effective attractive interaction between the 
5629: nearest-neighbor antiparallel spins, 
5630: is introduced by hand.
5631: On the other hand, 
5632: in the Hubbard model, the effective attractive interaction 
5633: which is responsible for the pairing 
5634: is generated by the onsite Coulomb 
5635: repulsion through higher-order many-body processes
5636: only when the system has the suitable conditions.
5637: 
5638: \subsection{Implications for $d_{x^2-y^2}$-wave pairing in the cuprates} 
5639: 
5640: The presence of $d_{x^2-y^2}$ pairing correlations 
5641: in the Hubbard model and the nature of the effective 
5642: interaction mediating it are issues 
5643: which are of interest for studying high
5644: temperature superconductivity in the cuprates.
5645: This is so especially after it became clear that the superconducting 
5646: gap function in the high $T_c$ cuprates is 
5647: of the $d_{x^2-y^2}$-wave type.
5648: Within this context,
5649: enormous amount of research has been carried out on this model.
5650: 
5651: In this article, the numerical studies 
5652: of $d_{x^2-y^2}$-wave pairing 
5653: in the Hubbard model have been reviewed.
5654: The Monte Carlo simulations have shown the presence of the 
5655: $d_{x^2-y^2}$-wave pairing correlations in the Hubbard model,
5656: even though sufficiently low temperatures, where long-range 
5657: order could establish, have not been reached.
5658: For $U/t=4$ and at the temperatures 
5659: where the simulations are carried out,
5660: the values of the $d_{x^2-y^2}$-wave eigenvalues of the 
5661: particle-particle Bethe-Salpeter equation are in agreement 
5662: with the FLEX calculations. 
5663: When carried out at low temperatures,
5664: the FLEX calculations find $T_c$'s of order $0.025t$.
5665: Since $t$ is estimated to be about $0.45$~eV 
5666: for a single-band model of the cuprates
5667: [Hybertson {\it et al.} 1990], 
5668: this value of the $T_c$ corresponds to $\sim 130$~K.
5669: This is a high value reflecting the electronic 
5670: energy scales of the model.
5671: Furthermore, the comparison of the FLEX and the QMC data
5672: suggests that the pairing could be stronger for $U=8t$.
5673: However, these are only mean-field estimates 
5674: for a possible Kosterlitz-Thouless transition
5675: and, in fact, 
5676: the ground state of the doped 2D Hubbard model is not known. 
5677: It might be that an additional parameter needs to be tuned 
5678: in order to create optimum conditions for $d_{x^2-y^2}$-wave  
5679: superconductivity in two dimensions.
5680: For instance, 
5681: in the 2-leg ladder case, it was necessary to tune $t_{\perp}/t$.
5682: In this respect, 
5683: the second-nearest-neighbour hopping $t'$ may play a role
5684: for the enhancement of the $d_{x^2-y^2}$ pairing in the 2D system
5685: like $t_{\perp}$ does for the 2-leg ladder. 
5686: It is also possible that a three-band CuO$_2$ model 
5687: with an onsite Coulomb repulsion at the Cu sites offers 
5688: additional degrees of freedom for creating more favourable 
5689: conditions for pairing.
5690: 
5691: The QMC results reviewed here show that the short-range AF
5692: correlations are responsible for the $d_{x^2-y^2}$ pairing
5693: correlations in the 2D Hubbard model.
5694: These results also suggest that effects which enhance the 
5695: magnitude of 
5696: $\Gamma_{Is}({\bf q}\sim (\pi,\pi),i\omega_m=0)$ 
5697: and the 
5698: single-particle spectral weight near the $(\pi,0)$ and the 
5699: $(0,\pi)$ points of the Brillouin zone act to enhance 
5700: the $d_{x^2-y^2}$-wave superconducting correlations.
5701: Clearly, other effects which enhance 
5702: $d_{x^2-y^2}$-wave pairing can exist.
5703: Exploring these possibilities is an active research field.
5704: 
5705: \setcounter{equation}{0}\setcounter{figure}{0}
5706: \section{Summary and conclusions}
5707: 
5708: In this paper, the numerical studies of the 2D and 
5709: the 2-leg Hubbard models have been reviewed.
5710: For the 2D Hubbard model, 
5711: data from the QMC simulations have been shown.
5712: These data represent what has been obtained over the years 
5713: in the parameter regime allowed by the sign problem. 
5714: For the 2-leg Hubbard ladder, 
5715: the QMC results at finite temperatures 
5716: and the DMRG data on the ground state have been shown. 
5717: 
5718: Here, the emphasis has been placed on the 
5719: $d_{x^2-y^2}$-wave superconducting 
5720: correlations observed in the Hubbard model. 
5721: In order to develop an understanding of the 
5722: origin of these correlations, results on the spin, 
5723: charge and the single-particle excitations have been shown along 
5724: with the data 
5725: on the particle-particle and the particle-hole interactions. 
5726: The observation of the short-range AF fluctuations by the 
5727: NMR and the neutron scattering experiments, 
5728: and the unusual single-particle spectrum 
5729: seen in the ARPES data are properties
5730: which support using a Hubbard framework for studying 
5731: the pairing correlations of the cuprates.
5732: 
5733: In the 2D Hubbard model, 
5734: upon hole doping the long-range AF order is 
5735: destroyed and the system exhibits short-range AF correlations. 
5736: The maximum-entropy analysis of the QMC data shows that 
5737: the AF and the Coulomb correlations 
5738: strongly affect the single-particle properties. 
5739: In particular, it has been seen that, 
5740: as the strength of $U/t$ increases, significant amount of single-particle 
5741: weight remains pinned near the Fermi level at the $(\pi,0)$ and 
5742: $(0,\pi)$ points of the Brillouin zone. 
5743: These generate phase space for scatterings in the 
5744: $d_{x^2-y^2}$-wave BCS channel.
5745: The results on the particle-particle and the particle-hole 
5746: irreducible interactions have been also presented. 
5747: It has been seen that, for $U=4t$, 
5748: a properly-renormalized single spin-fluctuation 
5749: exchange interaction can describe the momentum, the Matsubara-frequency
5750: and the temperature dependence of the effective-particle-particle
5751: interaction. 
5752: This means that, for these values of $U/t$ and $T/t$, 
5753: the attraction in the $d_{x^2-y^2}$ channel is mediated 
5754: by the AF spin fluctuations. 
5755: 
5756: The DMRG results on the 2-leg ladder, 
5757: which have been presented in Section~7 are important. 
5758: These calculations represent the first example 
5759: where a purely repulsive
5760: onsite interaction leads to superconducting correlations which are 
5761: enhanced over the noninteracting ($U=0$) case in the ground state
5762: of a bulk system.
5763: For certain values of $t_{\perp}/t$ 
5764: and in the intermediate coupling regime, 
5765: this model exhibits enhanced 
5766: power-law $d_{x^2-y^2}$-like superconducting correlations. 
5767: The QMC simulations for this system show that 
5768: the effective particle-particle interaction $\Gamma_I$ peaks at 
5769: $(\pi,\pi)$ momentum transfer and, in addition, 
5770: the $d_{x^2-y^2}$ pairing correlations are strongest when there 
5771: is significant amount of single-particle spectral weight 
5772: near the Fermi level at the $(\pi,0)$ and $(0,\pi)$ 
5773: points in the Brillouin zone. 
5774: This way the system makes optimum use of the momentum structure 
5775: in $\Gamma_I$ for $d_{x^2-y^2}$-wave pairing. 
5776: 
5777: These data were also compared with various other approaches 
5778: to the Hubbard model, such as the 
5779: diagrammatic FLEX approximation, the 
5780: variational and the projector Monte Carlo simulations,
5781: the dynamical cluster approximation and the one-loop 
5782: RG calculations. 
5783: In addition, the similarities and the differences with the 
5784: $t$-$J$ model were briefly discussed. 
5785: The implications for the $d_{x^2-y^2}$-wave 
5786: superconductivity seen in the cuprates were also noted.
5787: 
5788: It is difficult to reach conclusions 
5789: about the strongly correlated systems. 
5790: Nevertheless, here, the following conclusions are given:
5791: 
5792: (1) An onsite Coulomb repulsion can lead to 
5793: superconducting correlations which are enhanced with respect to the 
5794: noninteracting ($U=0$) case in the ground state
5795: of a bulk system.
5796: This is proven for the case of the 2-leg Hubbard ladder.
5797: 
5798: (2) The 2D Hubbard model exhibits $d_{x^2-y^2}$-wave pairing 
5799: correlations which grow as the temperature is lowered in the parameter 
5800: regime where the QMC simulations are carried out. 
5801: The fastest growing pairing correlations occur in the 
5802: singlet $d_{x^2-y^2}$-wave channel.
5803: These correlations are not weak.
5804: 
5805: (3) The QMC simulations also find that the effective pairing interaction 
5806: in the 2D and the 2-leg Hubbard models 
5807: in the intermediate coupling regime and 
5808: at temperatures greater or of order $J/2$
5809: is consistent with the 
5810: spin-fluctuation exchange approximation. 
5811: 
5812: (4) Two factors which create optimum conditions for 
5813: $d_{x^2-y^2}$ pairing in the Hubbard model are enhanced 
5814: single-particle spectral weight at the Fermi level near the 
5815: $(\pi,0)$ and $(0,\pi)$ points in the Brillouin zone
5816: and large weight in the effective pairing interaction
5817: $\Gamma_I$ at large momentum transfers. 
5818: Clearly, 
5819: there can be other ways of enhancing 
5820: the $d_{x^2-y^2}$  pairing correlations.
5821: 
5822: \vskip 0.5 truecm
5823: {\bf Acknowledgments}
5824: 
5825: Most of the numerical data presented here are from
5826: various calculations carried out with 
5827: T. Dahm, R.M. Noack, D.J. Scalapino and S.R. White.
5828: The QMC data shown here were obtained 
5829: at the San Diego Supercomputer Center.
5830: 
5831: \setcounter{equation}{0}\setcounter{figure}{0}
5832: \section{Appendix}
5833: 
5834: The results shown in Sections 3 through 7 of this review 
5835: were obtained with the determinantal QMC and the DMRG techniques.
5836: Here, these two approaches to the many-body problem 
5837: will be described briefly. 
5838: 
5839: \subsection{Determinantal QMC technique}
5840: 
5841: The determinantal QMC algorithm used in obtaining 
5842: the QMC data shown here
5843: is described in Ref.~[White {\it et al.} 1989b].
5844: Reviews of this technique can be found 
5845: in Refs.~[Scalapino 1993, Muramatsu 1999].
5846: The basic idea of this approach 
5847: is due to Blankenbecler, Scalapino and Sugar 
5848: [Blankenbecler {\it et al.} 1981].
5849: Here, the purpose is to calculate the expectation value 
5850: of an operator $O$ at finite 
5851: temperature $T$ in the grand canonical ensemble, 
5852: \begin{equation}
5853: \langle O\rangle = {1\over Z} {\rm Tr}\, Oe^{-\beta H}
5854: \end{equation}
5855: where
5856: \begin{equation}
5857: Z={\rm Tr}\, e^{-\beta H}.
5858: \end{equation}
5859: It is convenient to take the Hubbard hamiltonian 
5860: as $H=K+V$ where
5861: \begin{equation}
5862: \label{K}
5863: K = -t \sum_{ \langle ij\rangle,\sigma} 
5864: ( c^{\dagger}_{i\sigma} c_{j\sigma} + h.c.) 
5865: -\mu \sum_i (n_{i\uparrow} + n_{i\downarrow})
5866: \end{equation}
5867: and
5868: \begin{equation}
5869: V=U\sum_i 
5870: (n_{i\uparrow}- {1\over 2}) (n_{i\downarrow}- {1\over 2}).
5871: \end{equation}
5872: After discretizing the inverse temperature into $L$ time slices, 
5873: $\beta=L\Delta\tau$, 
5874: the Trotter approximation is used to rewrite 
5875: the partition function as 
5876: \begin{equation}
5877: Z={\rm Tr}\, e^{-L\Delta\tau H} 
5878: \cong {\rm Tr}\, ( e^{-\Delta\tau V} e^{-\Delta\tau K})^L.
5879: \end{equation}
5880: Next,
5881: the interaction term $V$, 
5882: which is quartic in the fermion operators, 
5883: is reduced to a bilinear form with 
5884: the Hubbard-Stratonovich transformation 
5885: \begin{equation}
5886: e^{-\Delta\tau U (n_{i\uparrow}-1/2)(n_{i\downarrow}-1/2)} = 
5887: { e^{-\Delta\tau U/4} \over 2 }
5888: \sum_{S_{i\ell}=\pm 1} \,
5889: e^{-\Delta\tau S_{i\ell} 
5890: \lambda (n_{i\uparrow} - n_{i\downarrow})}
5891: \end{equation}
5892: where $\cosh({\lambda\Delta\tau}) = \exp({U\Delta\tau/2})$.
5893: Here, 
5894: at each lattice site $i$ and 
5895: for each time slice $\ell$, an Ising spin 
5896: $S_{i\ell}= \pm 1$ is used for the decoupling. 
5897: This form of the Hubbard-Stratonovich decoupling 
5898: was introduced by Hirsch
5899: [Hirsch 1985].
5900: After integrating out the fermion degrees of freedom, 
5901: the partition function is given by 
5902: \begin{equation} 
5903: Z = \sum_{\{S_{i\ell}\}} \, 
5904: {\rm det}\, M^{\uparrow}(\{ S_{i\ell} \})
5905: {\rm det}\, M^{\downarrow}(\{ S_{i\ell} \})
5906: \end{equation}
5907: where
5908: \begin{equation}
5909: M^{\sigma}(\{S_{i\ell}\}) = I + B^{\sigma}_L B^{\sigma}_{L-1}...B^{\sigma}_1
5910: \end{equation}
5911: and 
5912: \begin{equation}
5913: B^{\sigma}_{\ell} = e^{ -\sigma \Delta\tau \nu(\ell) } 
5914: e^{-\Delta\tau k}
5915: \end{equation}
5916: with electron spin $\sigma=\pm 1$.
5917: Here, 
5918: the summation is over all configurations
5919: $\{S_{i\ell}\}$ of the Ising fields, 
5920: $I$ is the $N\times N$ unit matrix, 
5921: $B_{\ell}^{\sigma}$'s are $N\times N$ matrices where
5922: $\nu(\ell)_{ij} = \delta_{ij} S_{i\ell}$ 
5923: and  $k$ is the matrix representation 
5924: of the kinetic energy operator, Eq.~(\ref{K}).
5925: Similar expressions can be obtained for the expectation value  
5926: $\langle O \rangle$.
5927: For instance, 
5928: the equal-time single-particle correlation function 
5929: $\langle c_{j\sigma} c_{j'\sigma}^{\dagger}\rangle$ 
5930: can be expressed as 
5931: \begin{equation} 
5932: \label{cc}
5933: \langle c_{j\sigma} c_{j'\sigma}^{\dagger} \rangle = 
5934: {1\over Z}
5935: \sum_{\{ S_{i\ell}\} } \,
5936: G_{\sigma} (j,j'; \{S_{i\ell}\}) 
5937: {\rm det}\,M^{\uparrow}(\{S_{i\ell}\})
5938: {\rm det}\,M^{\downarrow}(\{S_{i\ell}\})
5939: \end{equation}
5940: where
5941: \begin{equation}
5942: G_{\sigma}(j,j';\{S_{i\ell}\}) = 
5943: \left[ 
5944: (I + B^{\sigma}_L B^{\sigma}_{L-1}...B^{\sigma}_1)^{-1} 
5945: \right]_{jj'}.
5946: \end{equation}
5947: 
5948: The summation over the Ising spin variables 
5949: in Eq.~(\ref{cc}) is evaluated using Monte Carlo 
5950: sampling techniques.
5951: In this way,
5952: Eq.~(\ref{cc}) becomes
5953: \begin{equation}
5954: \langle  c_{j\sigma} c^{\dagger}_{j'\sigma} \rangle = 
5955: \langle G_{\sigma} (j,j';\{S_{i\ell}\}) 
5956: \rangle_{P}
5957: \end{equation}
5958: where $\langle ...\rangle_{P}$
5959: is evaluated by averaging over spin configurations 
5960: $\{ S_{i\ell}\}$ 
5961: generated with the probability distribution 
5962: \begin{equation}
5963: P(\{S_{i\ell}\}) = 
5964: {1\over Z} 
5965: {\rm det}\, M^{\uparrow}(\{S_{i\ell}\})
5966: {\rm det}\, M^{\downarrow}(\{S_{i\ell}\}).
5967: \end{equation}
5968: Similarly, 
5969: the time-dependent single-particle Green's function 
5970: $\langle c_{j\sigma}(\tau) c_{j'\sigma}(\tau') \rangle$
5971: is calculated by averaging 
5972: $G_{\sigma}(j,\tau; j',\tau'; \{S_{i\ell}\})$,
5973: which can also be expressed in terms of $B^{\sigma}_{\ell}$'s.
5974: 
5975: The Hubbard-Stratonovich transformation maps 
5976: the interacting fermion problem on the 2D lattice 
5977: to that of non-interacting fermions 
5978: in a random Ising field on a 3D lattice,
5979: the third dimension being the imaginary time axis $\tau$.
5980: At this stage, 
5981: for a given spin configuration $\{S_{i\ell}\}$, 
5982: the Wick's theorem applies 
5983: to the higher order correlation functions. 
5984: Hence,
5985: the higher-order correlation functions 
5986: can be obtained by averaging over the 
5987: products of the single-particle propagators
5988: in random Ising fields.
5989: For instance, 
5990: the two-particle correlation function 
5991: \begin{equation}
5992: \langle 
5993: T\,
5994: c_{\uparrow}(i_4,\tau_4) c_{\downarrow}(i_3,\tau_3)
5995: c^{\dagger}_{\downarrow}(i_2,\tau_2) c^{\dagger}_{\uparrow}(i_1,\tau_1)
5996: \rangle,
5997: \end{equation}
5998: which is used in extracting the particle-particle 
5999: reducible interaction $\Gamma$ in Section 6, 
6000: can be obtained from 
6001: \begin{equation}
6002: \langle 
6003: G_{\uparrow}  ( i_4,\tau_4;i_1,\tau_1;\{S_{i\ell}\} ) 
6004: G_{\downarrow}( i_3,\tau_3;i_2,\tau_2;\{S_{i\ell}\} ) 
6005: \rangle_P.
6006: \end{equation}
6007: 
6008: For $\langle n\rangle=1$, 
6009: the product of the fermion determinants,
6010: ${\rm det}\, M^{\uparrow}(\{S_{i\ell}\}) \,
6011:  {\rm det}\, M^{\downarrow}(\{S_{i\ell}\})$,
6012: is positive.
6013: However, 
6014: away from half-filling and for $U>0$, 
6015: this product is not positive for all configurations 
6016: $\{S_{i\ell}\}$.
6017: In this case, 
6018: one can use the probability distribution 
6019: \begin{equation}
6020: \tilde{P}(\{S_{i\ell}\}) = 
6021: {  
6022: |{\rm det}\, M^{\uparrow}(\{S_{i\ell}\}) \,
6023:  {\rm det}\, M^{\downarrow}(\{S_{i\ell}\})|
6024: \over 
6025: \sum_{ \{S_{i\ell}\} } 
6026: |{\rm det}\, M^{\uparrow}(\{S_{i\ell}\}) \,
6027:  {\rm det}\, M^{\downarrow}(\{S_{i\ell}\})|
6028: },
6029: \end{equation}
6030: in order to calculate the expectation value $\langle O\rangle$ with 
6031: \begin{equation}
6032: \label{O}
6033: \langle O\rangle = 
6034: { 
6035: \langle 
6036: O( \{S_{i\ell}\} ) {\rm sign}( \{S_{i\ell}\} ) 
6037: \rangle_{\tilde{P}}
6038: \over 
6039: \langle {\rm sign}( \{S_{i\ell}\} ) \rangle_{\tilde{P}}
6040: }.
6041: \end{equation}
6042: Here,
6043: ${\rm sign}(\{S_{i\ell}\})$ is the sign of 
6044: ${\rm det}\, M^{\uparrow}(\{S_{i\ell}\}) \, 
6045:  {\rm det}\, M^{\downarrow}(\{S_{i\ell}\})$
6046: and the average 
6047: $\langle ...\rangle_{\tilde{P}}$ 
6048: is calculated with the probability distribution $\tilde{P}$. 
6049: The denominator in Eq.~(\ref{O}),
6050: \begin{equation}
6051: \langle {\rm sign}\rangle \equiv 
6052: \langle {\rm sign}(\{S_{i\ell}\})\rangle_{\tilde{P}},
6053: \end{equation}
6054: becomes small when the number of configurations 
6055: with positive signs is close to that with the negative signs.
6056: It has been shown that away from half-filling 
6057: $\langle {\rm sign}\rangle$ 
6058: decreases exponentially as $T$ decreases
6059: [Loh {\it et al.} 1990].
6060: This causes large statistical errors in 
6061: $\langle O\rangle$ and it is the cause of the sign problem. 
6062: The temperature and the doping regime of the 2D Hubbard model 
6063: which cannot be studied with QMC 
6064: because of the sign problem is shown at the end of Section 6.
6065: In spite of the sign problem, 
6066: useful information has been extracted from the 2D 
6067: and the 2-leg Hubbard models with QMC simulations.
6068: Currently, the search for ways of removing 
6069: the sign problem is an important field of study.
6070: Improvements in the sign problem
6071: which decreases the temperatures accessible 
6072: to QMC simulations by even a factor of two would be valuable. 
6073: 
6074: \subsection{DMRG technique}
6075: 
6076: The DMRG technique which was invented by White 
6077: represents a significant development in the application 
6078: of the renormalization group 
6079: ideas to interacting lattice models
6080: [White 1992 and 1993].
6081: Currently, it is the numerical method of choice 
6082: for studying the ground state properties of 
6083: quasi-1D interacting systems. 
6084: The brief discussion of this technique given 
6085: here follows closely that of 
6086: Ref.~[White 1993]. 
6087: For simplicity, 
6088: the following discussion will be for a 1D lattice.
6089: 
6090: One of the differences between the DMRG and the standard real-space
6091: RG approach is in the treatment of the boundary conditions 
6092: in the basic blocking procedure.
6093: In the standard approach for a finite 1D system, 
6094: the chain is broken into finite blocks $B_{\ell}$ 
6095: of $\ell$ sites.  
6096: Here, first 
6097: the Hamiltonian for two identical blocks is diagonalized 
6098: and then the lowest eigenstates are kept 
6099: to form a new approximate Hamiltonian describing a
6100: larger block $B'_{2\ell}$ with $2\ell$ sites.
6101: This procedure is illustrated in Fig.~10.1(a).
6102: White and Noack have shown that this was of blocking introduces 
6103: large errors for a model of a free particle on a 1D lattice
6104: because of the way the boundary conditions are treated
6105: during the blocking
6106: [White and Noack 1992].
6107: In the standard RG scheme, 
6108: neglecting the connection of the two blocks 
6109: to the neighboring blocks corresponds to 
6110: setting the wave function of the particle to 0 outside of the blocks. 
6111: Consequently, 
6112: the low lying states from the previous iteration 
6113: cause a `kink` in the wave function in the middle of the enlarged 
6114: block $B'_{2\ell}$.
6115: Hence,
6116: in order to accurately represent the states of block $B'_{2\ell}$, 
6117: it is necessary to use 
6118: almost all of the states of block $B_{\ell}$, 
6119: and any truncation of these states introduces large errors
6120: in the RG process.
6121: White and Noack have noted that 
6122: using a combination of boundary conditions 
6123: during the blocking fixes this deficiency 
6124: for a free particle on a 1D lattice. 
6125: They have also shown that an alternative approach
6126: is to use a superblock 
6127: which is composed of more than two blocks. 
6128: In this way,
6129: two of the blocks in the superblock are used 
6130: to form a larger block for the next iteration 
6131: while the effect of the other blocks 
6132: is to apply a variety of boundary conditions. 
6133: The DMRG technique uses a superblock during the RG iteration,
6134: which is illustrated in Fig.~10.1(b)
6135: for a 1D lattice of $L$ sites.
6136: This superblock is composed of two blocks 
6137: with $\ell$ and $\ell'=L-\ell-2$ sites 
6138: and two additional sites in the middle, 
6139: which can be considered as additional blocks.
6140: The block $B_{\ell}$ and 
6141: its neighboring lattice site are called the `system',
6142: while the rest of the superblock is called the `environment'.
6143: With each iteration, 
6144: the system block gets enlarged by one lattice site. 
6145: 
6146: \begin{figure}
6147: \centering
6148: \iffigure
6149: \mbox{
6150: \subfigure[]{
6151: \epsfysize=5.5cm
6152: \epsffile[150 250 470 520]{ch10-fig1a.ps}}
6153: \quad
6154: \subfigure[]{
6155: \epsfysize=5.5cm
6156: \epsffile[70 250 600 520]{ch10-fig1b.ps}}}
6157: \fi
6158: \caption{
6159: (a) Standard blocking scheme used 
6160: for real-space RG for a 1D system.
6161: Here, the Hamiltonian for two blocks 
6162: $B_{\ell}$ each with $\ell$ sites is 
6163: diagonalized and truncated to form an approximate Hamiltonian 
6164: for the new block $B'_{2\ell}$ with $2\ell$ sites. 
6165: (b) DMRG blocking scheme for a finite 1D system using a superblock 
6166: which is composed of blocks $B_{\ell}$ and $B^R_{\ell'}$,
6167: and two additional sites in between. 
6168: Here, first the Hamiltonian for the superblock is diagonalized 
6169: and the density matrix is formed 
6170: for the enlarged block $B'_{\ell+1}$.
6171: The Hamiltonian for block $B'_{\ell+1}$ is then expressed 
6172: in a reduced basis composed of the 
6173: leading eigenstates of this density matrix.
6174: In the next iteration, 
6175: the block $B'_{\ell+1}$ replaces $B_{\ell}$.
6176: }
6177: \label{10.1}
6178: \end{figure}
6179: 
6180: Another novel feature of the DMRG technique is 
6181: the use of the density matrices 
6182: to choose the states which are to be 
6183: kept during the iteration [White 1992 and 1993].
6184: In the standard RG approach, 
6185: the lowest $m$ eigenstates of the Hamiltonian for two
6186: blocks is used in  forming the truncated Hamiltonian 
6187: for the larger block,
6188: as illustrated in Fig.~10.1(a).
6189: This would be a reasonable approximation for a model 
6190: where the coupling between the blocks is weak. 
6191: However, 
6192: for an interacting system such as the Hubbard model, 
6193: each block is strongly coupled to its environment. 
6194: In this case, 
6195: White has shown that it is better to use during the truncation 
6196: the eigenstates of the density matrix of the system,
6197: and, in fact, 
6198: the optimal states to be kept are the eigenvectors 
6199: of the density matrix of the system with the largest eigenvalues. 
6200: The use of the superblock 
6201: and the density matrix formulation are the 
6202: important new ideas used by the DMRG technique.
6203: 
6204: At this point, 
6205: the DMRG algorithm for a finite 1D lattice can be 
6206: summarized as follows:
6207: (1) The first step is to diagonalize the Hamiltonian 
6208: for the superblock, which is illustrated in Fig.~10.1(b),
6209: using exact diagonalization techniques
6210: in order to find the ground state $\psi_{ij}$. 
6211: Here, 
6212: the index $i$ refers to the system part of the lattice 
6213: and $j$ refers to the environment. 
6214: (2) Next, the reduced density matrix for block $B'_{\ell+1}$ 
6215: is calculated by tracing over the environment,
6216: \begin{equation}
6217: \rho_{ii'}=\sum_j \, \psi_{ij}\psi^*_{i'j},
6218: \end{equation}
6219: and then $\rho$ is diagonalized.
6220: (3) At this stage, the Hamiltonian for $B'_{\ell+1}$ can be 
6221: transformed to a reduced basis which consists of the $m$
6222: leading eigenstates of $\rho$.
6223: Typically a few hundred eigenstates of $\rho$ are kept. 
6224: (4) In the next iteration, this approximate Hamiltonian for 
6225: $B'_{\ell+1}$ is used in place of the Hamiltonian for $B_{\ell}$.
6226: 
6227: With this procedure, the system block is built up 
6228: by adding one lattice site at a time 
6229: as the iteration continues form one end 
6230: of the lattice to the other end. 
6231: With each sweep through the lattice, 
6232: a better approximation is obtained for 
6233: the Hamiltonian of the block $B_{\ell}$. 
6234: In addition, 
6235: during the sweeps, 
6236: the Hamiltonians of the system blocks from the previous sweep 
6237: are used as the Hamiltonians for the environment blocks.
6238: For 1D lattices, 
6239: reflections of $B_{\ell}$ can be used in place of $B^R_{\ell'}$ 
6240: in order to build up the superblock to the proper size 
6241: during the first sweep.
6242: In order to extend this algorithm to 2D, 
6243: a row of neighboring sites can be added 
6244: to the system block at each iteration, or
6245: a single site can be added
6246: using a connected 1D path through the 2D lattice. 
6247: For higher dimensional lattices, 
6248: an empty environment block
6249: $B^R_{\ell'}$ is used in the first sweep. 
6250: 
6251: The DMRG technique is more accurate 
6252: when the number of connections 
6253: between the system and the environment blocks is minimized. 
6254: Hence, 
6255: usually open boundary conditions are employed.
6256: When periodic boundary conditions are used, 
6257: the accuracy decreases. 
6258: For 2D lattices, the accuracy of DMRG also decreases 
6259: as a function of the width of the lattice. 
6260: However, 
6261: if a larger number of the eigenstates of $\rho$ are kept 
6262: during the truncation, the accuracy of the algorithm increases.
6263: Finally, 
6264: the DMRG is most accurate for large $U$ and least accurate for $U=0$.
6265: 
6266: With this technique, 
6267: various equal-time correlation functions for the 2-leg
6268: Hubbard ladder have been calculated 
6269: on lattices up to $2\times 32$ in size
6270: [Noack {\it et al.} 1996].
6271: For the 3-leg Hubbard ladder 
6272: there are DMRG data for a $3\times 12$ lattice
6273: [Bonca {\it et al.} 2000]. 
6274: For the $t$-$J$ model, 
6275: larger systems can be studied with this technique. 
6276: In this case, 
6277: 2D clusters up to $28\times 8$ in size have been studied 
6278: using cylindrical boundary conditions
6279: [White and Scalapino 1998a].
6280: Currently,
6281: the DMRG technique is used in a wide variety of fields, 
6282: and the efforts to develop new algorithms 
6283: to study larger clusters continues.
6284: In addition, 
6285: there is continuing work to extend the DMRG technique 
6286: in order to calculate the dynamical correlation functions. 
6287: 
6288: \newpage
6289: 
6290: \section{References}
6291: %\setlength{\parindent}{0pt}
6292: 
6293: \begin{list}{}{\setlength{\itemindent}{-1cm}\setlength{\itemsep}{0pt}
6294: \setlength{\parsep}{0pt}}
6295: \item $^*$ E-mail: nbulut@ku.edu.tr
6296: 
6297: \item ABRAHAMS, E., BALATSKY, A., SCHRIEFFER, J. R., and 
6298: ALLEN, P. B., 
6299: 1993,
6300: {\it Phys. Rev.} B, {\bf 47}, 513.
6301: 
6302: \item ABRAHAMS, E., BALATSKY, A., SCALAPINO, D. J., and
6303: SCHRIEFFER, J. R.,
6304: 1995,
6305: {\it Phys. Rev.} B, {\bf 52}, 1271. 
6306: 
6307: \item AEBI, P., OSTERWALDER, J., SCHWALLER, P., SCHLAPBACH, L., 
6308: SHIMODA, M., MOCHIKU, T., and KADOWAKI, K.,  
6309: 1994,
6310: {\it Phys. Rev. Lett.}, {\bf 72}, 2757.
6311: 
6312: \item ANDERSON, P. W., 
6313: 1987,
6314: {\it Science}, {\bf 235}, 1196.
6315: 
6316: \item ANDERSON, P. W., 
6317: 1997,
6318: {\it Adv. Phys.}, {\bf 46}, 3.
6319: 
6320: \item ANDERSON, P. W., BASKARAN, G., ZOU, Z., and HSU, T.,
6321: 1987,
6322: {\it Phys. Rev. Lett.}, {\bf 58}, 2790.
6323: 
6324: \item ANDERSON, P. W., and ZOU, Z.,
6325: 1988,
6326: {\it Phys. Rev. Lett.}, {\bf 60}, 132.
6327: 
6328: \item BALATSKY, A. V., and ABRAHAMS, E.,
6329: 1992,
6330: {\it Phys. Rev.} B, {\bf 45}, 13125.
6331: 
6332: \item BALENTS, L., and FISHER, M. P. A.,
6333: 1996,
6334: {\it Phys. Rev.} B, {\bf 53}, 12133.
6335: 
6336: \item BECCA, F., CAPONE, M., and SORELLA, S., 
6337: 2000,
6338: {\it Phys. Rev.} B, {\bf 62}, 12700.
6339: 
6340: \item BEDNORZ, J. G., and M\"ULLER, K. A., 
6341: 1986, 
6342: {\it Z. Phys.} B, {\bf 64}, 189.
6343: 
6344: \item BEENEN, J., and EDWARDS, D. M., 
6345: 1995,
6346: {\it J. Low Temp. Phys.}, {\bf 99}, 403;
6347: {\it Phys. Rev.} B, {\bf 52}, 13636.
6348: 
6349: \item BEREZINSKII, V. L., 
6350: 1974,
6351: {\it Pisma Zh. Eksp. Teor. Fiz.}, {\bf 20}, 628
6352: [1974, {\it JETP Lett.}, {\bf 20}, 287].
6353: 
6354: \item BERK, N. F., and SCHRIEFFER, J. R., 
6355: 1966, 
6356: {\it Phys. Rev. Lett.}, {\bf 17}, 433.
6357: 
6358: \item BICKERS, N. E., SCALAPINO, D. J., and SCALETTAR, R. T.,
6359: 1987, 
6360: {\it Int. J. Mod. Phys.} B, {\bf 1}, 687.
6361: 
6362: \item BICKERS, N. E., SCALAPINO, D. J., and WHITE, S. R.,
6363: 1989, 
6364: {\it Phys. Rev. Lett.}, {\bf 62}, 961.
6365: 
6366: \item BICKERS, N. E., and WHITE, S. R., 
6367: 1991, 
6368: {\it Phys. Rev.} B, {\bf 43}, 8044.
6369: 
6370: \item BIRGENEAU, R. J.,
6371: 1990,
6372: in {\it Physical Properties of High Temperature Superconductors II}, 
6373: edited by D. M. Ginsberg (World Scientific, Singapore).
6374: 
6375: \item BLANKENBECLER, R., SCALAPINO, D. J., and SUGAR, R. L., 
6376: 1981,
6377: {\it Phys. Rev.} D, {\bf 24}, 2278.
6378: 
6379: \item BONCA, J., GUBERNATIS, J. E., GUERRERO, M., 
6380: JECKELMANN, E., and WHITE, S. R.,
6381: 2000,
6382: {\it Phys. Rev.} B, {\bf 61}, 3251.
6383: 
6384: \item BULUT, N., HONE, D., SCALAPINO, D. J., and BICKERS, N. E., 
6385: 1990,
6386: {\it Phys. Rev.} B, {\bf 41}, 1797.
6387: 
6388: \item BULUT, N., 
6389: 1990,
6390: in {\it Dynamics of Magnetic Fluctuations in High-Temperature
6391: Superconductors},
6392: edited by G. Reiter, P. Horsch, and G.C. Psaltakis (Plenum).
6393: 
6394: \item BULUT, N., and SCALAPINO, D. J., 
6395: 1991, 
6396: {\it Phys. Rev. Lett.}, {\bf 67}, 2898.
6397: 
6398: \item BULUT, N., and SCALAPINO, D. J., 
6399: 1992, 
6400: {\it Phys. Rev. Lett.}, {\bf 68}, 706.
6401: 
6402: \item BULUT, N., SCALAPINO, D. J., and WHITE, S. R., 
6403: 1993, 
6404: {\it Phys. Rev.} B, {\bf 47}, 2742;
6405: {\it Phys. Rev.} B, {\bf 47}, 6157;
6406: {\it Phys. Rev.} B, {\bf 47}, 14599.
6407: 
6408: \item BULUT, N., SCALAPINO, D. J., and WHITE, S. R., 
6409: 1994a, 
6410: {\it Phys. Rev. Lett.}, {\bf 72}, 705;
6411: {\it Phys. Rev.} B, {\bf 50}, 7215;
6412: {\it Phys. Rev. Lett.}, {\bf 73}, 748.
6413: 
6414: \item BULUT, N., SCALAPINO, D. J., and WHITE, S. R., 
6415: 1994b, 
6416: {\it Phys. Rev.} B, {\bf 50}, 9623.
6417: 
6418: \item BULUT, N., SCALAPINO, D. J., and WHITE, S. R., 
6419: 1995, 
6420: {\it Physica} C, {\bf 246}, 85.
6421: 
6422: \item BULUT, N., and SCALAPINO, D. J., 
6423: 1995, 
6424: {\it J. Phys. Chem. Solids}, {\bf 56}, 1597.
6425: 
6426: \item BULUT, N.,
6427: 1996,
6428: {\it Tr. J. Phys.}, {\bf 20}, 548.
6429: 
6430: \item BULUT, N., and SCALAPINO, D. J., 
6431: 1996, 
6432: {\it Phys. Rev.} B, {\bf 54}, 14971.
6433: 
6434: \item CALANDRA, M., BECCA, F., and SORELLA, S.,
6435: 1998,
6436: {\it Phys. Rev. Lett.}, {\bf 81}, 5185.
6437: 
6438: \item CHEN, L., BOURBONNAIS, C., LI, T., and TREMBLAY, A.-M. S.
6439: 1991, 
6440: {\it Phys. Rev. Lett.}, {\bf 66}, 369.
6441: 
6442: \item CHEN, Y. C., MOREO, A., ORTOLANI, F., DAGOTTO, E., 
6443: and LEE, T. K.,
6444: 1994,
6445: {\it Phys. Rev.} B, {\bf 50}, 655. 
6446: 
6447: \item CYROT, M., 
6448: 1986,
6449: {\it Solid State Commun.}, {\bf 60}, 253.
6450: 
6451: \item DAGOTTO, E., ORTOLANI, F., and SCALAPINO, D. J., 
6452: 1991, 
6453: {\it Phys. Rev.} B, {\bf 46}, 3183;
6454: DAGOTTO, E., MOREO, A., ORTOLANI, F., RIERA, J., and SCALAPINO, D. J., 
6455: 1991, 
6456: {\it Phys. Rev. Lett.}, {\bf 67}, 1918.
6457: 
6458: \item DAGOTTO, E., RIERA, J., and SCALAPINO, D. J., 
6459: 1992a, 
6460: {\it Phys. Rev.} B, {\bf 45}, 5744.
6461: 
6462: \item DAGOTTO, E., MOREO, A., ORTOLANI, F., POILBLANC, D., and 
6463: RIERA, J., 
6464: 1992b, 
6465: {\it Phys. Rev.} B, {\bf 45}, 10741.
6466: 
6467: \item DAGOTTO, E., MOREO, A., ORTOLANI, F., RIERA, J., and 
6468: SCALAPINO, D. J., 
6469: 1992c, 
6470: {\it Phys. Rev.} B, {\bf 45}, 10107.
6471: 
6472: \item DAGOTTO, E., 
6473: 1994,
6474: {\it Rev. Mod. Phys.}, {\bf 66}, 763.
6475: 
6476: \item DAGOTTO, E., and NAZARENKO, A., and BONINSEGNI, M.,
6477: 1994,
6478: {\it Phys. Rev. Lett.}, {\bf 73}, 728.
6479: 
6480: \item DAGOTTO, E., and RICE, T. M., 
6481: 1996, 
6482: {\it Science}, {\bf 271}, 618.
6483: 
6484: \item DAHM, T., and TEWORDT, L., 
6485: 1995, 
6486: {\it Physica} C, {\bf 246}, 61;
6487: 1995, 
6488: {\it Phys. Rev.} B, {\bf 52}, 1297;
6489: 1995,
6490: {\it Phys. Rev. Lett.}, {\bf 74}, 793. 
6491: 
6492: \item DAHM, T., and BULUT, N.,
6493: 1996,
6494: unpublished.
6495: 
6496: \item DAHM, T., and SCALAPINO, D. J., 
6497: 1997, 
6498: {\it Physica} C, {\bf 288}, 33.
6499: 
6500: \item DAMASCELLI, A., LU, D. H., and SHEN, Z.-X., 
6501: 2001,
6502: {\it J. Electron Spectr. Relat. Phenom.}, {\bf 117-118}, 165.
6503: 
6504: \item DAUL, S., SCALAPINO, D. J., and WHITE, S. R., 
6505: 2000, 
6506: {\it Phys. Rev. Lett.}, {\bf 84}, 4188.
6507: 
6508: \item DESSAU, D., S., {\it et al.},
6509: 1993,
6510: {\it Phys. Rev. Lett.}, {\bf 71}, 2781.
6511: 
6512: \item DING, H. {\it et al.}, 
6513: 1996,
6514: {\it Nature}, {\bf 382}, 51.
6515: 
6516: \item DONIACH, S., and ENGELSBERG, S., 
6517: 1966, 
6518: {\it Phys. Rev. Lett.}, {\bf 17}, 750.
6519: 
6520: \item DOPF, G., MURAMATSU, A., and HANKE, W.,
6521: 1992a,
6522: {\it Phys. Rev. Lett.}, {\bf 68}, 353.
6523: 
6524: \item DOPF, G., WAGNER, J., DIETERICH, P., MURAMATSU, A., 
6525: and HANKE, W., 
6526: 1992b, 
6527: {\it Phys. Rev. Lett.}, {\bf 68}, 2082.
6528: 
6529: \item DORNEICH, A., ZACHER, M. G., GR\"OBER, C., and EDER, R., 
6530: 2000,
6531: {\it Phys. Rev.} B, {\bf 61}, 12816.
6532: 
6533: \item DUFFY, D., and MOREO, A., 
6534: 1995,
6535: {\it Phys. Rev.} B, {\bf 51}, 11882.
6536: 
6537: \item DZYALOSHINSKII, I., 
6538: 1987,
6539: {\it Zh. Eksp. Teor. Fiz.} {\bf 93} 1487
6540: [{\it Sov. Phys. JETP} {\bf 66}, 848 (1987)].
6541: 
6542: \item EMERY, V. J., 
6543: 1986, 
6544: {\it Synth. Metals}, {\bf 13}, 21.
6545: 
6546: \item EMERY, V. J., KIVELSON, S. A., and LIN, H. Q., 
6547: 1990, 
6548: {\it Phys. Rev. Lett.}, {\bf 64}, 475.
6549: 
6550: \item ENDRES, H., NOACK, R. M., HANKE, W., POILBLANC, D., and 
6551: SCALAPINO, D. J., 
6552: 1996,
6553: {\it Phys. Rev.} B, {\bf 53}, 5530.
6554: 
6555: \item FURUKAWA, N., and IMADA, M., 
6556: 1991, 
6557: {\it J. Phys. Soc. Jpn.}, {\bf 60}, 3604.
6558: 
6559: \item FURUKAWA, N., and IMADA, M., 
6560: 1992, 
6561: {\it J. Phys. Soc. Jpn.}, {\bf 61}, 3331.
6562: 
6563: \item FURUKAWA, N., RICE, T. M., and SALMHOFER, M., 
6564: 1998,
6565: {\it Phys. Rev. Lett.}, {\bf 81}, 3195.
6566: 
6567: \item GEORGES, A., KOTLIAR, G., KRAUTH, W., and ROZENBERG, M.J., 
6568: 1996,
6569: {\it Rev. Mod. Phys.} {\bf 68}, 13.
6570: 
6571: \item GOFRON, K., {\it et al.},
6572: 1993,
6573: {\it J. Phys. Chem. Solids}, {\bf 54}, 1193. 
6574: 
6575: \item GOPALAN, S., RICE, T. M., and SIGRIST, M.,
6576: 1994,
6577: {\it Phys. Rev.} B, {\bf 49}, 8901.
6578: 
6579: \item GR\"OBER, C., EDER, R., and HANKE, W.,
6580: 2000,
6581: {\it Phys. Rev.} B, {\bf 62}, 4336.
6582: 
6583: \item GUERRERO, M., ORTIZ, G., and GUBERNATIS, J. E., 
6584: 1999, 
6585: {\it Phys. Rev.} B, {\bf 59}, 1706. 
6586: 
6587: \item HALBOTH, C. J. and METZNER, W., 
6588: 2000,
6589: {\it Phys. Rev.} B, {\bf 61}, 7364.
6590: 
6591: \item HARDY, W. N., BONN, D. A., MORGAN, D. C., LIANG, R., and ZHANG, K., 
6592: 1993,
6593: {\it Phys. Rev. Lett.} {\bf 70}, 3999.
6594: 
6595: \item HAAS, S., MOREO, A., and DAGOTTO, E.,
6596: 1995,
6597: {\it Phys. Rev. Lett.}, {\bf 74}, 4281.
6598: 
6599: \item HAYWARD, C. A., POILBLANC, D., NOACK, R. M., 
6600: SCALAPINO, D. J., and HANKE, W.,
6601: 1995, 
6602: {\it Phys. Rev. Lett.}, {\bf 75}, 926. 
6603: 
6604: \item HELLBERG, C. S., and MANOUSAKIS, E., 
6605: 1997, 
6606: {\it Phys. Rev. Lett.}, {\bf 78}, 4609. 
6607: 
6608: \item HELLBERG, C. S., and MANOUSAKIS, E., 
6609: 1999, 
6610: {\it Phys. Rev. Lett.}, {\bf 83}, 132. 
6611: 
6612: \item HELLBERG, C. S., and MANOUSAKIS, E., 
6613: 2000, 
6614: {\it Phys. Rev.} B, {\bf 61}, 11787. 
6615: 
6616: \item HIRSCH, J. E., 
6617: 1985, 
6618: {\it Phys. Rev. Lett.}, {\bf 54}, 1317;
6619: 1985, 
6620: {\it Phys. Rev.} B, {\bf 31}, 4403. 
6621: 
6622: \item HIRSCH, J. E., and TANG, S.,
6623: 1989, 
6624: {\it Phys. Rev. Lett.}, {\bf 62}, 591. 
6625: 
6626: \item HONERKAMP, C. SALMHOFER, M., FURUKAWA, N., and RICE, T.M., 
6627: 2001,
6628: {\it Phys. Rev.} B, {\bf 63}, 035109.
6629: 
6630: \item HUBBARD, J.,
6631: 1963, 
6632: {\it Proc. Roy. Soc.} A, {\bf 276}, 238.
6633: 
6634: \item HUSCROFT, C., JARRELL, M., MAIER, Th., MOUKOURI, S., 
6635: and TAHVILDARZADEH, A. N., 
6636: 2001,
6637: {\it Phys. Rev. Lett.}, {\bf 86}, 139.
6638: 
6639: \item HUSSLEIN, T., MORGENSTERN, I., NEWNS, D. M., 
6640: PATTNAIK, P. C., SINGER, J. M., and MATUTTIS, H. G.,
6641: 1996,
6642: {\it Phys. Rev.} B, {\bf 54}, 16179.
6643: 
6644: \item HYBERTSEN, M. S., STECHEL, E. B., SCHL\"UTER, M., and
6645: JENNISON, D. R., 
6646: 1990, 
6647: {\it Phys. Rev.} B, {\bf 41}, 11068.
6648: 
6649: \item IMADA, M., 
6650: 1991,
6651: {\it J. Phys. Soc. Jpn.}, {\bf 60}, 2740.
6652: 
6653: \item IMADA, M., and KOHNO, M., 
6654: 2000,
6655: {\it Phys. Rev. Lett.}, {\bf 84}, 143.
6656: 
6657: \item IMAI, T., SLICHTER, C. P., YOSHIMURA, K., and KOSUGE, K., 
6658: 1993,
6659: {\it Phys. Rev. Lett.}, {\bf 70}, 1002.
6660: 
6661: \item ITOH, Y., YASUOKA, H., FUJIWARA, Y., UEDA, Y., 
6662: MACHI, T., TOMENO, I., TAI, K., KOSHIZUKA, N., and TANAKA, S.,
6663: 1992, 
6664: {\it J. Phys. Soc. Jpn.}, {\bf 61}, 1287.
6665: 
6666: \item JAKLIC, J., and PRELOVSEK, P., 
6667: 2000,
6668: {\it Adv. Phys.}, {\bf 49}, 1.
6669: 
6670: \item JARRELL, M., 
6671: 1992,
6672: {\it Phys. Rev. Lett.}, {\bf 69}, 168.
6673: 
6674: \item JARRELL, M., and GUBERNATIS, J. E., 
6675: 1996,
6676: {\it Phys. Rep.}, {\bf 269}, 134.
6677: 
6678: \item JECKELMANN, E., SCALAPINO, D. J., and WHITE, S. R.,  
6679: 1998,
6680: {\it Phys. Rev.} B, {\bf 58}, 9492.
6681: 
6682: \item KAMPF, A. P., and SCHRIEFFER, J. R., 
6683: 1990a, 
6684: {\it Phys. Rev.} B, {\bf 41}, 6399.
6685: 
6686: \item KAMPF, A. P., and SCHRIEFFER, J. R., 
6687: 1990b, 
6688: {\it Phys. Rev.} B, {\bf 42}, 7967.
6689: 
6690: \item KANAMORI, J.,
6691: 1963, 
6692: {\it Prog. Theor. Phys.}, {\bf 30}, 275.
6693: 
6694: \item KIVELSON, S. A., ROKHSAR, D. S., and SETHNA, J. P., 
6695: 1987,
6696: {\it Phys. Rev.} B, {\bf 35}, 8865.
6697: 
6698: \item KOIKE, S., YAMAJI, K., and YANAGISAWA, T.,
6699: 2000,
6700: {\it Physica} B, {\bf 284}, 417.
6701: 
6702: \item LEDERER, P. MONTHAMBAUX, G., and POILBLANC, D.,
6703: 1987,
6704: {\it J. Phys.} {\bf 48}, 1613. 
6705: 
6706: \item LEGGETT, A., J., 
6707: 1975,
6708: {\it Rev. Mod. Phys.}, {\bf 47}, 331.
6709: 
6710: \item LIU, Z., and MANOUSAKIS, E., 
6711: 1992,
6712: {\it Phys. Rev.} B, {\bf 44}, 2414.
6713: 
6714: \item LOH, E. Y, GUBERNATIS, J. E., SCALETTAR, R. T., WHITE, S. R., 
6715: SCALAPINO, D. J., and SUGAR, R. L., 
6716: 1990,
6717: {\it Phys. Rev.}, B, {\bf 41}, 9301.
6718: 
6719: \item LUO, J., and BICKERS, N. E.,
6720: 1993, 
6721: {\it Phys. Rev.} B, {\bf 47}, 12153.
6722: 
6723: \item M$^2$S-HTSC VI, 
6724: 2000,
6725: {\it Proceedings of the International Conference on Materials 
6726: and Mechanisms of Superconductivity 
6727: and High Temperature Superconductors},
6728: edited by K. Salama, W.K. Chu and P.W.C. Chu,
6729: {\it Physica} C, {\bf 341-348}.
6730: 
6731: \item MACHIDA, K.,
6732: 1989,
6733: {\it Physica} C, {\bf 158}, 192.
6734: 
6735: \item MAHAN, G. D.,
6736: 1981,
6737: ``Many-particle Physics", (Plenum, New York).
6738: 
6739: \item MAHAJAN, A. V., ALLOUL, H., COLLIN, G., MARUCCO, J. F.,
6740: 1994,
6741: {\it Phys. Rev. Lett.}, {\bf 72}, 3100.
6742: 
6743: \item MAHAJAN, A. V., ALLOUL, H., COLLIN, G., MARUCCO, J. F.,
6744: 2000,
6745: {\it Euro. Phys. J.} B, {\bf 13}, 457.
6746: 
6747: \item MAIER, Th., JARRELL, M., PRUSCHKE, T., and KELLER, J.,
6748: 2000,
6749: {\it Phys. Rev. Lett.},  {\bf 85}, 1524.
6750: 
6751: \item MARTINDALE, J. A., BARRETT, S. E., KLUG, C. A., O'HARA, K. E., 
6752: DESOTO, S. M., SLICHTER, C. P., FRIEDMAN, T. A., and GINSBERG, D. M., 
6753: 1992, 
6754: {\it Phys. Rev. Lett.}, {\bf 68}, 702. 
6755: 
6756: \item METZNER, W. and VOLLHARDT, D., 
6757: 1989,
6758: {\it Phys. Rev. Lett.}, {\bf 62}, 324.
6759: 
6760: \item MILA, F., and RICE, T. M., 
6761: 1989,  
6762: {\it Physica} C, {\bf 157}, 561.
6763: 
6764: \item MILLIS, A. J.,  MONIEN, H., and PINES, D., 
6765: 1990, 
6766: {\it Phys. Rev.} B, {\bf 42}, 167.
6767: 
6768: \item MIYAKE, K., SCHMITT-RINK, S., and VARMA, C. M., 
6769: 1986,
6770: {\it Phys. Rev.} B, {\bf 34}, 6554.
6771: 
6772: \item MONTHOUX, P., BALATSKY, A. V., and PINES, D., 
6773: 1991, 
6774: {\it Phys. Rev. Lett.}, {\bf 67}, 3448;
6775: {\it Phys. Rev.} B, {\bf 46}, 14803. 
6776: 
6777: \item MONTHOUX, P., and PINES, D., 
6778: 1992, 
6779: {\it Phys. Rev. Lett.}, {\bf 69}, 961;
6780: {\it Phys. Rev.} B, {\bf 50}, 16015. 
6781: 
6782: \item MONTHOUX, P., and SCALAPINO, D. J., 
6783: 1994, 
6784: {\it Phys. Rev. Lett.}, {\bf 72}, 1874. 
6785: 
6786: \item MOREO, A., and SCALAPINO, D. J.,
6787: 1991,
6788: {\it Phys. Rev.} B, {\bf 43}, 8211.
6789: 
6790: \item MOREO, A., and SCALAPINO, D. J., and DAGOTTO, E.,
6791: 1991,
6792: {\it Phys. Rev.} B, {\bf 43}, 11442.
6793: 
6794: \item MOREO, A.,
6795: 1992,
6796: {\it Phys. Rev.} B, {\bf 45}, 5059.
6797: 
6798: \item MOREO, A., HAAS, S., SANDVIK, A. W., and DAGOTTO, E., 
6799: 1995,
6800: {\it Phys. Rev.} B, {\bf 51}, 12045.
6801: 
6802: \item MORIYA, T., TAKAHASHI, Y., and UEDA, K., 
6803: 1990,
6804: {\it J. Phys. Soc. Jpn.}, {\bf 59}, 2905.
6805: 
6806: \item MORIYA, T., and UEDA, K., 
6807: 2000,
6808: {\it Adv. Phys.}, {\bf 49}, 555.
6809: 
6810: \item M\"ULLER-HARTMANN, E., 
6811: 1989,
6812: {\it Z. Phys.} B, {\bf 74}, 507.
6813: 
6814: \item MURAMATSU, A., 
6815: 1999,
6816: in {\it Quantum Monte Carlo Methods in Physics and Chemistry},
6817: edited by M.P. Nightingale and C.J. Umrigar (Kluwer Academic).
6818: 
6819: \item NAKANISHI, T., YAMAJI, K., and YANAGISAWA, T.,
6820: 1997, 
6821: {\it J. Phys. Soc. Jpn.}, {\bf 66}, 294.
6822: 
6823: \item NEWNS, D. M., TSUEI, C. C., HUEBENER, R. P., 
6824: VAN BENTUM, P. J. M., PATTNAIK, P. C., and CHI, C. C., 
6825: 1994, 
6826: {\it Phys. Rev. Lett.}, {\bf 73}, 1695.
6827: 
6828: \item NOACK, R. M., WHITE, S. R., and SCALAPINO, D. J., 
6829: 1994, 
6830: {\it Phys. Rev. Lett.}, {\bf 73}, 882.
6831: 
6832: \item NOACK, R. M., WHITE, S. R., and SCALAPINO, D. J., 
6833: 1995, 
6834: {\it Europhys. Lett.}, {\bf 30}, 163.
6835: 
6836: \item NOACK, R. M., WHITE, S. R., and SCALAPINO, D. J., 
6837: 1996, 
6838: {\it Physica} C, {\bf 270}, 281.
6839: 
6840: \item NOACK, R. M., BULUT, N., SCALAPINO, D. J., and ZACHER, M.G., 
6841: 1997, 
6842: {\it Phys. Rev.} B, {\bf 56}, 7162.
6843: 
6844: \item PAO, C. H., and BICKERS, N. E.,
6845: 1994, 
6846: {\it Phys. Rev. Lett.}, {\bf 72}, 1870.
6847: 
6848: \item PAO, C.-H., and BICKERS, N. E.,
6849: 1995, 
6850: {\it Phys. Rev.} B, {\bf 51}, 16310.
6851: 
6852: \item PENNINGTON, C. H., and SLICHTER, C. P.,
6853: 1990,
6854: in {\it Physical Properties of High Temperature Superconductors II}, 
6855: edited by D. M. Ginsberg (World Scientific, Singapore).
6856: 
6857: \item PENNINGTON, C. H., and SLICHTER, C. P.,
6858: 1991
6859: {\it Phys. Rev. Lett.}, {\bf 66}, 381.
6860: 
6861: \item POILBLANC, D. and RICE, T. M., 
6862: 1989,
6863: {\it Phys. Rev.} B, {\bf 39}, 9749.
6864: 
6865: \item POILBLANC, D., 
6866: 1993,
6867: {\it Phys. Rev.} B, {\bf 48}, 3368;
6868: 1994,
6869: {\it Phys. Rev.} B, {\bf 49}, 1477.
6870: 
6871: \item PREUSS, R., MURAMATSU, A., von der LINDEN, W.,
6872: DIETRICH, P., ASSAAD, F. F., and HANKE, W., 
6873: 1994,
6874: {\it Phys. Rev.  Lett.}, {\bf 73}, 732.
6875: 
6876: \item PREUSS, R., HANKE, W., and von der LINDEN, W., 
6877: 1995,
6878: {\it Phys. Rev.  Lett.}, {\bf 75}, 1344.
6879: 
6880: \item PREUSS, R., HANKE, W., GR\"OBER, C.,  and EVERTZ, H. G., 
6881: 1997,
6882: {\it Phys. Rev.  Lett.}, {\bf 79}, 1122.
6883: 
6884: \item PRUSCHKE, T., JARRELL, M., and FREERICKS, J. K.,
6885: 1995,
6886: {\it Adv. Phys.}, {\bf 42}, 187.
6887: 
6888: \item PUTIKKA, W. O., LUCHINI, M. U., and RICE, T. M., 
6889: 1992,
6890: {\it Phys. Rev. Lett.},  {\bf 68}, 538.
6891: 
6892: \item PUTIKKA, W. O., GLENISTER, R. L., SINGH, R. R. P., and
6893: TSUNETSUGU, H., 
6894: 1994, 
6895: {\it Phys. Rev. Lett.}, {\bf 73}, 170.
6896: 
6897: \item ROMMER, S., WHITE, S. R., and SCALAPINO, D. J., 
6898: 2000,
6899: {\it Phys. Rev.} B, {\bf 61}, 13424.
6900: 
6901: \item RONNING, F. {\it et al.},
6902: 1998,
6903: {\it Science}, {\bf  282}, 2067.
6904: 
6905: \item SCALAPINO, D. J.,
6906: 1993, 
6907: in {\it Proceedings of the Summer School on Modern Perspectives 
6908: in Many-Body Physics},
6909: Canberra (World Scientific, Singapore).
6910: 
6911: \item SCALAPINO, D. J.,
6912: 1995, 
6913: {\it Phys. Rep.}, {\bf 250}, 330.
6914: 
6915: \item SCALAPINO, D. J., LOH, E. Jr., and HIRSCH, J. E., 
6916: 1986,
6917: {\it Phys. Rev.} B, {\bf 34}, 8190.
6918: 
6919: \item SCALAPINO, D. J., WHITE, S. R., and ZHANG, S. C., 
6920: 1992,
6921: {\it Phys. Rev. Lett.}, {\bf 68}, 2830.
6922: 
6923: \item SCALAPINO, D. J., WHITE, S. R., and ZHANG, S. C., 
6924: 1993,
6925: {\it Phys. Rev.} B, {\bf 47}, 7995.
6926: 
6927: \item SCALETTAR, R. T., SCALAPINO, D. J., SUGAR, R. L., 
6928: and WHITE, S. R., 
6929: 1991,
6930: {\it Phys. Rev.} B, {\bf 44}, 770.
6931: 
6932: \item SCHRIEFFER, J. R., WEN, X. G., and ZHANG, S. C., 
6933: 1988,
6934: {\it Phys. Rev. Lett.}, {\bf 60}, 944.
6935: 
6936: \item SCHRIEFFER, J. R., WEN, X. G., and ZHANG, S. C., 
6937: 1989,
6938: {\it Phys. Rev.} B, {\bf 39}, 11663.
6939: 
6940: \item SCHRIEFFER, J. R., 
6941: 1994,
6942: {\it Solid State Commun.}, {\bf 92}, 129.
6943: 
6944: \item SCHRIEFFER, J. R., 
6945: 1995,
6946: {\it J. Low Temp. Phys.}, {\bf 99}, 397.
6947: 
6948: \item SCHULZ, H. J., 
6949: 1987,
6950: {\it Europhys. Lett.} {\bf 4}, 609.
6951: 
6952: \item SCHULZ, H. J., 
6953: 1989,
6954: {\it J. Physique}, {\bf 50}, 2833.
6955: 
6956: \item SCHULZ, H. J., 
6957: 1999,
6958: {\it Phys. Rev.} B, {\bf 59}, R2471.
6959: 
6960: \item SHEN, Z.-X., DESSAU, D. S., WELLS, B. O., KING, D. M., SPEICER, W. E.,
6961: ARKO, A. J., MARSHALL, D., LOMBARDO, L. W., KAPITULNIK, A.,
6962: DICKINSON, P., DONIACH, S., DICARLO, J., LOESER, T., and PARK, C. H., 
6963: 1993,
6964: {\it Phys. Rev. Lett.}, {\bf 70}, 1553.
6965: 
6966: \item SHEN, Z.-X., and DESSAU, D. S., 
6967: 1995, 
6968: {\it Phys. Rep.}, {\bf 253}, 1.
6969: 
6970: \item SILVER, R., N., SILVIA, D., S., and GUBERNATIS, J. E.,
6971: 1990,
6972: {\it Phys. Rev.} B, {\bf 41}, 2380.
6973: 
6974: \item TAKIGAWA, M.,
6975: 1990,
6976: in {\it Dynamics of Magnetic Fluctuations in High-Temperature
6977: Superconductors},
6978: edited by G. Reiter, P. Horsch, and G.C. Psaltakis (Plenum).
6979: 
6980: \item TAKIGAWA, M., SMITH, J. L., and HULTS, W. L., 
6981: 1991,
6982: {\it Physica} C, {\bf 185}, 1105;
6983: {\it Phys. Rev.} B, {\bf 44}, 7764.
6984: 
6985: \item TRANQUADA,  J. M., STERNLIEB, B. J., AXE, J. D., 
6986: NAKAMURA, Y., and UCHIDA, S., 
6987: 1995, 
6988: {\it Nature}, {\bf 375}, 561.
6989: 
6990: \item TSUEI, C. C., CHI, C. C., NEWNS, D. M., PATTNAIK, P. C., 
6991: and DAUMLING, M.,
6992: 1992, 
6993: {\it Phys. Rev. Lett.}, {\bf 69}, 2134.
6994: 
6995: \item TSUEI, C. C., KIRTLEY, J. R., CHI, C. C., YU-JAHNES, L. S., 
6996: GUPTA, A., SHAW, T., SUN, J. Z., and KETCHEN, M. B., 
6997: 1994, 
6998: {\it Phys. Rev. Lett.}, {\bf 73}, 593.
6999: 
7000: \item TSUEI, C. C., and KIRTLEY, 
7001: 2000, 
7002: {\it Rev. Mod. Phys.}, {\bf 72}, 969.
7003: 
7004: \item VAN HARLINGEN, D., 
7005: 1995, 
7006: {\it Rev. Mod. Phys.}, {\bf 67}, 515.
7007: 
7008: \item VARMA, C. M., LITTLEWOOD, P. B., SCHMITT-RINK, S., 
7009: ABRAHAMS, E., and RUCKENSTEIN, A. E., 
7010: 1989,
7011: {\it Phys. Rev. Lett.}, {\bf 63}, 1996.
7012: 
7013: \item VEKIC, M., and WHITE, S. R., 
7014: 1993,
7015: {\it Phys. Rev.} B, {\bf 47}, 1160.
7016: 
7017: \item WHITE,  S. R., SCALAPINO, D. J., SUGAR, R. L., 
7018: BICKERS, N. E., and SCALETTAR, R. T., 
7019: 1989a, 
7020: {\it Phys. Rev.} B, {\bf 39}, 839.
7021: 
7022: \item WHITE,  S. R., SCALAPINO, D. J., SUGAR, R. L., 
7023: LOH, E. Y., GUBERNATIS, J. E., 
7024: and SCALETTAR, R. T., 
7025: 1989b, 
7026: {\it Phys. Rev.} B, {\bf 40}, 506.
7027: 
7028: \item WHITE, S. R., SCALAPINO, D. J., SUGAR, R. L., and 
7029: BICKERS, N. E.,
7030: 1989c,
7031: {\it Phys. Rev. Lett.}, {\bf 63}, 1523.
7032: 
7033: \item WHITE, S. R., 
7034: 1991,
7035: {\it Phys. Rev.} B, {\bf 44}, 4670;
7036: 1992,
7037: {\it Phys. Rev.} B, {\bf 46}, 5678.
7038: 
7039: \item WHITE,  S. R., 
7040: 1992, 
7041: {\it Phys. Rev. Lett.}, {\bf 69}, 2863.
7042: 
7043: \item WHITE,  S. R., 
7044: 1993, 
7045: {\it Phys. Rev.} B, {\bf 48}, 10345.
7046: 
7047: \item WHITE,  S. R., and NOACK, R. M.,  
7048: 1992, 
7049: {\it Phys. Rev. Lett.}, {\bf 68}, 3487.
7050: 
7051: \item WHITE,  S. R., and SCALAPINO, D., 
7052: 1998a, 
7053: {\it Phys. Rev. Lett.}, {\bf 80}, 1272;
7054: {\it Phys. Rev. Lett.}, {\bf 81}, 3227.
7055: 
7056: \item WHITE,  S. R., and SCALAPINO, D., 
7057: 1998b, 
7058: {\it Phys. Rev.} B, {\bf 57}, 3031.
7059: 
7060: \item WHITE,  S. R., and SCALAPINO, D., 
7061: 1999, 
7062: {\it Phys. Rev.} B, {\bf 60}, 753.
7063: 
7064: \item WHITE,  S. R., and SCALAPINO, D., 
7065: 2000, 
7066: {\it Phys. Rev.} B, {\bf 61}, 6320.
7067: 
7068: \item WOLLMAN, D. A., VAN HARLINGEN, D. J., LEE, W. C., GINSBERG, D. M., 
7069: and LEGGETT, A. J., 
7070: 1993, 
7071: {\it Phys. Rev. Lett.}, {\bf 71}, 2134.
7072: 
7073: \item XIAO, G., CIEPLAK, M. Z., XIAO, J. Q., CHIEN, C. L., 
7074: 1990,
7075: {\it Phys. Rev.} B, {\bf 42}, 8752.
7076: 
7077: \item YAMAJI, K., and SHIMOI, Y., 
7078: 1994, 
7079: {\it Physica} C, {\bf 222}, 349.
7080: 
7081: \item YAMAJI, K., and YANAGISAWA, T., NAKANISHI, T., and KOIKE, S., 
7082: 1998, 
7083: {\it Physica} C, {\bf 340}, 225.
7084: 
7085: \item ZAANEN, J., and GUNNARSON, O., 
7086: 1989, 
7087: {\it Phys. Rev.} B, {\bf 40}, 7391.
7088: 
7089: \item ZACHER, M. G., ARRIGONI, E. HANKE, W., and
7090: SCHRIEFFER, J. R., 
7091: 1998,
7092: {\it Phys. Rev.} B, {\bf 57}, 6370.
7093: 
7094: \item ZANCHI, D. and SCHULZ, H.J., 
7095: 2000,
7096: {\it Phys. Rev.} B, {\bf 61}, 13609.
7097: 
7098: \item ZHANG, S., CARLSON, J., and GUBERNATIS, J. E.,
7099: 1995, 
7100: {\it Phys. Rev. Lett.}, {\bf 74}, 3652.
7101: 
7102: \item ZHANG, S., CARLSON, J., and GUBERNATIS, J. E.,
7103: 1997, 
7104: {\it Phys. Rev. Lett.}, {\bf 78}, 4486.
7105: 
7106: \end{list}
7107: \end{document}
7108: 
7109: 
7110: 
7111: 
7112: 
7113: 
7114: