cond-mat0410445/rmp.tex
1: % Group addresses by affiliation; use superscriptaddress for long
2: % author lists, or if there are many overlapping affiliations.
3: % For Phys. Rev. appearance, change preprint to twocolumn.
4: %  Add 'draft' option to mark overfull boxes with black boxes
5: %  Add 'showpacs' option to make PACS codes appear
6: %  Add 'showkeys' option to make keywords appear
7: %\documentclass[aps,preprint,groupedaddress]{revtex4}
8: %\documentclass[aps,preprint,superscriptaddress]{revtex4}
9: %\documentclass[aps,twocolumn,groupedaddress,showkeys,showpacs]{revtex4}
10: \documentclass[rmp,twocolumn,groupedaddress,showpacs]{revtex4}
11: %\documentclass[aps,preprint,groupedaddress,showpacs]{revtex4}
12: %\usepackage{showkeys}
13: 
14: %\documentclass{article}
15: %\input rmp.rty
16: 
17: 
18: %\setlength{\textheight}{9.7in}
19: 
20: \def\Ref#1{\citeauthor{#1}, \citeyear{#1}}
21: \def\cit#1{\citeauthor{#1}, \citeyear{#1}}
22: 
23: 
24: \begin{document}
25: % You should use BibTeX and apsrev.bst for references
26: %\bibliographystyle{apsrev}
27: \bibliographystyle{apsrmp}
28: 
29: % Use the \preprint command to place your local institutional report
30: % number on the title page in preprint mode.
31: % Multiple \preprint commands are allowed.
32: %\preprint{}
33: 
34: %Title of paper
35: \title{
36: Doping a Mott Insulator: Physics of High Temperature Superconductivity
37: }
38: % Optional argument for running titles on pages
39: %\title[]{}
40: 
41: % repeat the \author .. \affiliation  \etc. as needed
42: % \email, \thanks, \homepage, \altaffiliation all apply to the current
43: % author. Explanatory text should go in the []'s, actual e-mail
44: % address or url should go in the {}'s for \email and \homepage.
45: % Please use the appropriate macro for the type of information
46: 
47: % \affiliation command applies to all authors since the last
48: % \affiliation command. The \affiliation command should follow the
49: % other information
50: % \affiliation can be followed by \email, \homepage, \thanks as well.
51: \author{Patrick A. Lee$^a$}
52: \author{Naoto Nagaosa$^b$}
53: \author{Xiao-Gang Wen$^a$}
54: \affiliation{$^a$ Department of Physics, Massachusetts Institute of Technology,
55: Cambridge, Massachusetts 02139\\
56: $^b$ CREST, Department of Applied Physics,
57: University of Tokyo,
58: 7--3--1 Hongo, Bunkyo-ku, Tokyo 113, Japan}
59: %\email{Your e-mail address}
60: %\homepage{http://dao.mit.edu/~wen}
61: %\thanks{}
62: %\altaffiliation{}
63: 
64: %Collaboration name if desired (requires use of superscriptaddress
65: %option in \documentclass). \noaffiliation is required (may also be
66: %used with the \author command).
67: %\collaboration can be followed by \email, \homepage, \thanks as well.
68: %\collaboration{}
69: %\noaffiliation
70: 
71: \date{\today}
72: %\date{Dec. 2002}
73: 
74: \begin{abstract}
75: This article reviews the effort to understand the physics of high temperature
76: superconductors from the point of view of doping a Mott insulator.  The basic
77: electronic structure of the cuprates is reviewed, emphasizing the physics of
78: strong correlation and establishing the model of a doped Mott insulator as a
79: starting point.  A variety of experiments are discussed, focusing on the
80: region of the phase diagram close to the Mott insulator (the underdoped
81: region) where the behavior is most anomalous.  The normal state in this region
82: exhibits the pseudogap phenomenon.  In contrast, the quasiparticles in the
83: superconducting state are well defined and behave according to theory.  We
84: introduce Anderson's idea of the resonating valence bond (RVB) and argue that
85: it gives a qualitative account of the data.  The importance of phase
86: fluctuation is discussed, leading to a theory of the transition temperature
87: which is driven by phase fluctuation and thermal excitation of quasiparticles.
88: However, we argue that phase fluctuation can only explain the pseudogap
89: phenomenology over a limited temperature range, and some additional physics is
90: needed to explain the onset of singlet formation at very high temperatures.
91: We then describe the numerical method of projected wavefunction which turns
92: out to be a very useful technique to implement the strong correlation
93: constraint, and leads to a number of predictions which are in agreement with
94: experiments.  The remainder of the paper deals with an analytic treatment of
95: the $t$-$J$ model, with the goal of putting the RVB idea on a more formal
96: footing.  The slave-boson is introduced to enforce the constraint of no double
97: occupation.  The implementation of the local constraint leads naturally to
98: gauge theories.  We follow the historical order and first review the $U(1)$
99: formulation of the gauge theory.  Some inadequacies of this formulation for
100: underdoping are discussed, leading to the $SU(2)$ formulation.  Here we
101: digress with a rather thorough discussion of the role of gauge theory in
102: describing the spin liquid phase of the undoped Mott insulator.  We emphasize
103: the difference between the high energy gauge group in the formulation of the
104: problem versus the low energy gauge group which is an emergent phenomenon.
105: Several possible routes to deconfinement based on different emergent gauge
106: groups are discussed, which lead to the physics of fractionalization and 
107: spin-charge separation.  We next describe the extension of the $SU(2)$ formulation to
108: nonzero doping.  We focus on a part of the mean field phase diagram called the
109: staggered flux liquid phase.  We show that inclusion of gauge fluctuation
110: provides a reasonable description of the pseudogap phase.  We emphasize that
111: $d$-wave superconductivity can be considered as evolving from a stable $U(1)$
112: spin liquid.  We apply these ideas to the high $T_c$ cuprates, and discuss
113: their implications for the vortex structure and the phase diagram.  A possible
114: test of the topological structure of the pseudogap phase is discussed.
115: \end{abstract}
116: % insert suggested PACS numbers in braces on next line
117: \pacs{74.20.Mn, 71.27.+a}
118: % insert suggested keywords - APS authors don't need to do this
119: \keywords{High $T_c$ superconductivity, spin liquid, slave-boson theory}
120: 
121: %\maketitle must follow title, authors, abstract, \pacs, and \keywords
122: \maketitle
123: 
124: % body of paper here - Use proper section commands
125: % References should be done using the \cite, \ref, and \label commands
126: %\section{Introduction}
127: %\label{intro}
128: 
129: \tableofcontents
130: 
131: 
132: 
133: 
134: \section{Introduction}
135: The discovery of high temperature superconductivity in cuprates \cite{BM8689}
136: %(Bednorz and M\"{u}ller, 1986) 
137: and the rapid raising of the transition temperature to well
138: above the melting point of nitrogen \cite{WAT8708}
139: %(Wu {\em et al.}, 1987)
140: ushered in an era
141: of great excitement for the condensed matter physics community.  For decades
142: prior to this discovery, the highest $T_c$ had been stuck at 23~K.  Not only
143: was the old record $T_c$ shattered, but the fact that high $T_c$
144: superconductivity was discovered in a rather unexpected material, a transition
145: metal oxide, made it clear that some novel mechanism must be at work.  The
146: intervening years have seen great strides in high  $T_c$ research.  First and
147: foremost, the growth and characterization of cuprate single crystals and thin
148: films have advanced to the point where sample quality and reproducibility
149: problems which plagued the field in the early days are no longer issues.  At
150: the same time, basically all conceivable experimental tools have been applied
151: to the cuprates.  Indeed, the need for more and more refined data has spurred
152: the development of experimental techniques such as angle resolved
153: photoemission spectroscopy (ARPES) and low temperature scanning tunneling
154: microscopy (STM).  Today the cuprate is arguably the best studied material
155: outside of the semiconductor family and a great deal of facts are known.  It
156: is also clear that many of the physical properties are unusual, particularly
157: in the metallic state above the superconductor.  Superconductivity is only one
158: aspect of a rich phase diagram which must be understood in its totality.
159: \begin{figure}[t]
160: \centerline{
161: \includegraphics[width=3.4in]{fig1.1.eps}
162: %\psfig{figure=Fig.1.eps,width=3.25in}
163: }
164: %\vspace{0.5cm}
165: \caption{Schematic phase diagram of high $T_c$ superconductors showing hole doping (right side) and
166: electron doping (left side). From 
167: %Damascelli {\em et al.}, 2003
168: \Ref{DHS0373}. }
169: \label{Fig.1}
170: \end{figure}
171: 
172: While there are hundreds of high $T_c$ compounds, they all share a layered
173: structure made up of one or more copper-oxygen planes.  They all fit into a
174: ``universal'' phase diagram shown in Fig. 1.  We start with the so-called
175: ``parent compound,'' in this case La$_2$CuO$_4$.  There is now general
176: agreement that the parent compound is an insulator, and should be classified
177: as a Mott insulator.  The concept of Mott insulation was introduced many years
178: ago \cite{M4916}
179: %(Mott, 1949)
180: to describe a situation where a material should be metallic according to band
181: theory, but is insulating due to strong electron-electron repulsion.  In our
182: case, in the copper-oxygen layer there is an odd number of electrons per unit
183: cell.  More specifically, the copper ion is doubly ionized and is in a $d^9$
184: configuration, so that there is a single hole in the $d$ shell per unit cell.
185: According to band theory, the band is half-filled and must be metallic.
186: Nevertheless, there is a strong repulsive energy cost to put two electrons (or
187: holes) on the same ion, and when this energy (commonly called $U$) dominates
188: over the hopping energy $t$, the ground state is an insulator due to strong
189: correlation effects.  It also follows that the Mott insulator should be an
190: antiferromagnet (AF), because when neighboring spins are oppositely aligned,
191: one can gain an energy $4t^2/U$ by virtual hopping.  This is called the
192: exchange energy $J$.  The parent compound is indeed an antiferromagnetic
193: insulator.  The ordering temperature $T_N \approx 300$K shown in
194: Fig.~\ref{Fig.1} is in fact misleadingly low because it is governed by a small
195: interlayer coupling, which is furthermore frustrated in La$_2$CuO$_4$ (see
196: \Ref{KBS9897}).
197: %Kastner {\em et al.}, 1998).  
198: %\cite{Kastner98}.
199: The exchange energy $J$ is in fact extraordinarily high, of order 1500~K, and
200: the parent compound shows strong antiferromagnetic correlation much above
201: $T_N$.
202: 
203: 
204: The parent compound can be doped by substituting some of the trivalent La by
205: divalent Sr.  The result is that $x$ holes are added to the Cu-O plane in
206: La$_{2-x}$Sr$_x$CuO$_4$.  This is called hole doping.  In the compound
207: Nd$_{2-x}$Ce$_x$CuO$_4$, the reverse happens in that $x$ electrons are added
208: to the Cu-O plane.  This is called electron doping.  As we can see from Fig.
209: 1, on the hole doping side the AF order is rapidly suppressed and is gone by 3
210: to 5\% hole concentration.  Almost immediately after the suppression of AF,
211: superconductivity appears, ranging from $x = 6$\% to 25\%.  The dome-shaped
212: $T_c$ is characteristic of all hole-doped cuprates, even though the maximum
213: $T_c$ varies from about 40~K in the La$_{2-x}$Sr$_x$CuO$_4$ (LSCO) family to
214: 93~K and higher in other families such as YBa$_2$Cu$_3$O$_{6+y}$ (YBCO) and
215: Ba$_2$Sr$_2$CaCu$_2$O$_{8+y}$ (Bi-2212).  On the electron doped side, AF is
216: more robust and survives up to $x = 0.14$, beyond which a region of
217: superconductivity arises.  One view is that the carriers are more prone to be
218: localized on the electron doped side, so that electron doping to closer to
219: dilution by nonmagnetic ions, which is less effective in suppressing AF order
220: than itinerant carriers.  Another possibility is that
221: the next neighbor hopping term favors AF on the electron doped side
222: \cite{SG0214}.
223: It is as though a more robust AF region is covering
224: up the more interesting phase diagram revealed on the hole doped side.  In
225: this review we shall focus on the hole doped materials, even though we will
226: address the issue of the particle-hole asymmetry of the phase diagram from
227: time to time.
228: 
229: The region in the phase diagram with doping $x$ less than that of the maximum
230: $T_c$ is called the underdoped region.  The metallic state above $T_c$ has
231: been under intense study and exhibits many unusual properties not encountered
232: before in any other metal.  This region of the phase diagram has been called
233: the pseudogap phase.  It is not a well defined phase in that a definite finite
234: temperature phase boundary has never been found.  The line drawn in Fig.~1
235: should be regarded as a cross-over.  Since we view the high $T_c$ problem as
236: synonymous with that of doping a Mott insulator, the underdoped region is
237: where the battleground between Mott insulator and superconductivity is drawn
238: and this is where we shall concentrate on in this review.
239: 
240: The region of the normal state above the optimal $T_c$ also exhibits unusual
241: properties.  The resistivity is linear in $T$ and the Hall coefficient is
242: temperature dependent (see \Ref{CWO9188}).
243: %Chien, Wang and Ong, 1991).   
244: %\cite{Chien91}.
245: These were cited as examples of non-Fermi liquid behavior since the early days
246: of high $T_c$.  Beyond optimal doping (called the overdoped region), sanity
247: gradually returns.  The normal state behaves more normally in that the
248: temperature dependence of the resistivity resembles $T^2$ over a temperature
249: range which increases with further overdoping.  The anomalous region above
250: optimal doping is sometimes referred to as the ``strange metal'' region.  We
251: offer a qualitative description of this region in section IX, but in our mind
252: the understanding of the ``strange metal'' is even more rudimentary that
253: of the ``pseudogap.''  A popular notion is that the strange metal is
254: characterized by a quantum critical point lying under the superconducting dome
255: (\cit{V9754}; \cit{CCG9737}; \cit{TL0053})
256: %(Varma, 1997; Tallon and Loram, 2000).
257: %\cite{Varma97; Tallon00}.
258: In our view, unless the nature of the ordered side of a quantum critical point
259: is classified, the simple statement of quantum criticality does not teach us
260: too much about the behavior in the critical region.  For this reason, we
261: prefer to concentrate on the underdoped region and leave the strange metal
262: phase to future studies.
263: 
264: Contrary to the experimental situation, the development of high $T_c$ theory
265: follows a rather tortuous path and people often have the impression that the
266: field is highly contentious and without a clear direction  or consensus.  We
267: do not agree with this assessment and would like to clearly state our point of
268: view from the outset.  Our starting point is, as already stated, that the
269: physics of high $T_c$ superconductivity is the physics of doping a Mott
270: insulator.  Strong correlation is the driving force behind the phase diagram.
271: We believe that there is a general consensus on this starting point.  The
272: simplest model which captures the strong correlation physics is the Hubbard
273: model and its strong coupling limit, the $t$-$J$ model.  Our view is that one
274: should focus on understanding these simple models before adding various
275: elaborations.  For example, further neighbor hopping certainly is significant
276: and as we shall discuss, plays an important role in understanding the
277: particle-hole asymmetry of the phase diagram.  Electron-phonon coupling can
278: generally be expected to be strong in transition metal oxides, and we shall
279: discuss their role in affecting spectral line shape.  However, these
280: discussions must be made in the context of strong correlation.  The logical
281: step is to first understand whether simple models such as the $t$-$J$ model
282: contains enough physics to explain the appearance of superconductivity and
283: pseudogaps in the phase diagram.
284: 
285: The strong correlation viewpoint was put forward by 
286: \Ref{A8796},
287: %Anderson, 1987, 
288: who revived his earlier work on a possible spin liquid state in a frustrated
289: antiferromagnet.  This state, called the resonating valence band (RVB), has no
290: long range AF order and is a unique spin singlet ground state.  It has spin
291: 1/2 fermionic excitations which are called spinons.  The idea is that when
292: doped with holes the RVB is a singlet state with coherent mobile carriers, and
293: is indistinguishable in terms of symmetry from a singlet BCS superconductor.
294: The process of hole doping was further developed by \Ref{KRS8765}
295: %Kivelson, Rokhsar and Sethna, 1987 
296: who 
297: %introduced the term holon to describe 
298: argue that the combination of the
299: doped hole with the spinon form a bosonic excitation. This excitation,
300: called the holon, carries charge but no spin whereas the
301: spinon carries spin 1/2 but no charge, and the notion of spin-charge
302: separation was born.  Meanwhile, a slave-boson theory was formulated by
303: \Ref{BZA8773}.
304: %Baskaran, Zou and Anderson, 1987.  
305: %\cite{BZA8773}
306: Many authors contributed to the development of the mean field theory,
307: culminating in the paper by 
308: \Ref{KL8842}
309: %Kotliar and Liu, 1988 
310: who found that the
311: superconducting state should have $d$-symmetry and that a state with spin gap
312: properties should exist above the superconducting temperature in the
313: underdoped region.  The 
314: possibility of d-wave superconductivity has been discussed in terms of the
315: exchange of spin fluctuations
316: (\cit{SLH8690}, \citeyear{SLH8794}; \cit{MSV8654}; \cit{E8377},
317: \citeyear{E8621}; \cit{MP9369}).
318: These discussions are either based on phenomenological coupling between spins
319: and fermions or via RPA treatment for the Harbbard model which is basically a
320: weak coupling expansion. In contrast, the salve-boson
321: theory was developed in the limit of strong repulsion. Details of the
322: mean field theory will be discussed
323: in section VIII.
324: 
325: At about the same time, the proposal by 
326: \Ref{A8796}
327: %Anderson, 1987 
328: of using projected mean
329: field states as trial wavefunctions was implemented on the computer by 
330: %Gros, 1988a,b
331: \Ref{G8831}, \citeyear{G8853}.  
332: The idea is to remove by hand on a computer all components of the
333: mean field wavefunction with doubly occupied sites, and use this as a
334: variational wavefunction for the $t$-$J$ model.  
335: \Ref{G8831}, \citeyear{G8853} 
336: %Gros, 1988a,b  
337: concluded that
338: the projected $d$-wave superconductor is the variational ground state for the
339: $t$-$J$ model over a range of doping.  The projected wavefunction method
340: remains one of the best numerical tools to tackle the $t$-$J$ or Hubbard model
341: and is reviewed in section VI.
342: 
343: It was soon realized that inclusion of fluctuations about the mean field
344: invariably leads to gauge theory 
345: \cite{BA8880,IL8988,NL9050}.
346: %(Baskaran and Anderson, 1988; Ioffe and Larkin, 1989; Nagaosa and Lee, 1990).
347: The gauge field fluctuations can be
348: treated at a Gaussian level and these early developments together with some of
349: the difficulties are reviewed in section IX.
350: 
351: In hindsight, the slave-boson mean field theory and the projected wavefunction
352: studies contain many of the qualitative aspects of the hole doped phase
353: diagram.  It is indeed quite remarkable that the main tools of treating the
354: $t$-$J$ model, \ie projected trial wavefunction, slave-boson mean field, and
355: gauge theory, were in place a couple of years after the discovery of high
356: $T_c$.  In some way the theory was ahead of its time, because the majority
357: view in the early days was that the pairing symmetry was $s$-wave, and the
358: pseudogap phenomenology remains to be discovered.  (The first hint came from
359: Knight shift measurements in 1989 shown in Fig. \ref{chi}(a).)  Some of the
360: early history and recent extensions are reviewed by 
361: %Anderson {\em et al.} (2004)
362: \Ref{ARR0455}.
363: 
364: The gauge theory approach is a difficult one to pursue systematically because
365: it is a strong coupling problem.  One important development is the realization
366: that the original $U(1)$ gauge theory should be extended to $SU(2)$ in order
367: to make a smooth connection to the underdoped limit 
368: \cite{WLsu2}
369: %(Wen and Lee, 1996).  
370: This
371: is discussed in sections XI and XII.  More generally, it was gradually
372: realized that the concepts of confinement/deconfinement which are central to
373: QCD also play a key role here, except that the presence of matter field make
374: this problem even more complex.  Since gauge theories are not so familiar to
375: condensed matter physicists, these concepts are discussed in some detail in
376: section X.  One of the notable recent advances is that the notion of the spin
377: liquid and its relation to deconfinement in gauge theory has been greatly
378: clarified and several soluble models and candidates based on numerical exact
379: diagonalization have been proposed.  It remains true, however, that so far no
380: two-dimensional spin liquid has been convincingly realized experimentally.
381: 
382: Our overall philosophy is that the RVB idea of a spin liquid and its relation
383: to superconductivity contains the essence of the physics and gives a
384: qualitative description of the underdoped phase diagram.  The goal of our
385: research is to put these ideas on a more quantitative footing.  Given the
386: strong coupling nature of the problem, the only way progress can be made is
387: for theory to work in consort with experiment.  Our aim is to make as many
388: predictions as possible, beyond saying that the pseudogap is a RVB spin
389: liquid, and challenge the experimentalists to perform tests.  Ideas along
390: these lines are reviewed in section XII.
391: 
392: High $T_c$ research is an enormous field and we cannot hope to be complete in
393: our references.  Here we refer to a number of excellent review articles on
394: various aspects of the subject.  
395: %M. Imada, A. Fujimori, and Y. Tokura
396: \Ref{IFT9839} reviewed the general topic of metal-insulator transition.
397: %Orenstein and Millis, 2000 and Norman and Pepin, 2003 
398: \Ref{OM0068} and \Ref{NP0347}
399: have provided highly readable accounts of experiments and general
400: theoretical approaches.  
401: Early numerical work was reviewed by \Ref{D9463}.
402: %Kastner {\em et al.}, 1998
403: \Ref{KBS9897} summarized the earlier
404: optical and magnetic neutron scattering data mainly on
405: La$_{2-x}$Sr$_x$CuO$_4$.  Major reviews of angle resolved photoemission data
406: (ARPES) have been provided by 
407: %Campuzzano {\em et al.}, 2003 and Damascelli {\em et al.}, 2003
408: \Ref{CNR03} and \Ref{DHS0373}.  Optics measurements on underdoped materials are reviewed by
409: %Timusk and Statt, 1999
410: \Ref{TS9961}.  The volumes edited by 
411: %D.M. Ginzberg, 1989
412: \Ref{Gin89} contain
413: excellent reviews of early NMR work by C.P. Slichter and early transport
414: measurement by N.P. Ong among others.  Discussions of stripe physics are
415: recently given by 
416: %Carlson {\em et al.}, 2003 and Kivelson {\em et al.}, 2003
417: \Ref{CEK03} and \Ref{KBF0301}.
418: A discussion of spin liquid states 
419: is given by
420: %Sachdev, 2003
421: \Ref{S0313} with an emphasis on dimer order and by
422: \Ref{Wen04} with an emphasis on quantum order. 
423: For an account of experiments and early RVB theory, see the
424: book by 
425: %Anderson, 1997
426: \Ref{AndHtc97}.
427:  
428: \section{Basic electronic structure of the cuprates} 
429: 
430: It is generally agreed that the physics of high $T_c$ superconductivity is
431: that of the copper oxygen layer, as shown in Fig. \ref{fig.2}.  In the parent
432: compound such as La$_2$CuO$_4$, the formal valence of Cu is $2+$, which means
433: that its electronic state is in the $d^9$ configuration.  The copper is
434: surrounded by six oxygens in an octahedral environment (the apical oxygen
435: lying above and below Cu are not shown in Fig. 2).  The distortion from a
436: perfect octahedron due to the shift 
437: \begin{figure}
438: \centerline{
439: \includegraphics[width=3.5in]{fig2.2.eps}
440: %\psfig{figure=Fig.2.eps,width=4.5in}
441: }
442: \centerline{
443: \includegraphics[width=1.5in]{fig2b.eps}
444: %\psfig{figure=Fig.2.eps,width=4.5in}
445: }
446: %\vspace{0.5cm}
447: \caption{The two-dimensional copper-oxygen layer (left) is simplified to the one-band model (right). 
448: Bottom figure shows the copper $d$ and oxygen $p$ orbitals in the hole picture.  A single hole with  $S = 1/2$ occupies the copper $d$ orbital in the insulator.
449: }
450: \label{fig.2}
451: \end{figure}
452: \noindent of the apical oxygens splits the $e_g$ orbitals so that the highest
453: partially occupied $d$ orbital is $x^2 -y^2$.  The lobes of this orbital point
454: directly to the $p$ orbital of the neighboring oxygen, forming a strong
455: covalent bond with a large hopping integral $t_{pd}$.  As we shall see, the
456: strength of this covalent bonding is responsible for the unusually high energy
457: scale for the exchange interaction.  Thus the electronic state of the cuprates
458: can be described by the so-called 3 band model, where in each unit cell we
459: have the Cu $d_{x^2-y^2}$ orbital and two oxygen $p$ orbitals  
460: \cite{E8759,VSA8781}.
461: %(Emery, 1987; Varma {\em et al.}, 1987).
462: The Cu orbital is singly occupied while the $p$
463: orbitals are doubly occupied, but these are admixed by $t_{pd}$.  In addition,
464: admixtures between the oxygen orbitals may be included.  These tight-binding
465: parameters may be obtained by fits to band structure calculations 
466: \cite{M8728,YFX8735}.
467: %(Mattheiss, 1987; Yu {\em et al.}, 1987).  
468: However, the largest energy in the problem
469: is the correlation energy for doubly occupying the copper
470: % or oxygen
471: orbital.  To describe these correlation energies, it is more convenient to go
472: to the hole picture. 
473: %As shown in Fig.~3, 
474: The Cu $d^9$ configuration is represented by energy level $E_d$ occupied by a
475: single hole with $S={1\over 2}$.  The oxygen $p$ orbital is empty of holes and
476: lies at energy $E_p$ which is higher than $E_d$. The energy to doubly occupy
477: $E_d$ (leading to a $d^8$ configuration) is $U_d$, which is very large and can
478: be considered infinity. The lowest energy excitation is the charge transfer
479: excitation where the hole hops from $d$ to $p$ with amplitude $-t_{pd}$.  If
480: $E_p-E_d$ is sufficiently large compared with $t_{pd}$, the hole will form a
481: local moment on Cu.  This is referred to as a charge transfer insulator in the
482: scheme of 
483: %Zaanen, Allen, and Sawatzky (1985)
484: \Ref{ZSA8518}.  Essentially, $E_p-E_d$ plays
485: the role of the Hubbard $U$ in the one-band model of the Mott insulator.
486: Experimentally an energy gap of 2.0~eV is observed and interpreted as the
487: charge transfer excitation (see 
488: %Kastner {\em et al.}, 1998
489: \Ref{KBS9897}).
490: 
491: %\begin{figure}
492: %\centerline{
493: %\psfig{figure=Fig.3.eps,width=4.5in}}
494: %\vspace{0.5cm}
495: %\caption{}
496: %\end{figure}
497: 
498: 
499: 
500: 
501: Just as in the one-band Mott-Hubbard insulator, where virtual hopping to
502: doubly occupied states leads to an exchange interaction $J{\v{S}}_1\cdot
503: {\v{S}}_2$ where $J=4t^2/U$, in the charge-transfer insulator, the local
504: moments on nearest neighbor Cu prefer antiferromagnetic alignment because both
505: spins can virtually hop to the $E_p$ orbital.  Ignoring the $U_p$ for doubly
506: occupying the $p$ orbital with holes, the exchange integral is given by
507: \be
508: J = {t^4_{pd}\over (E_p-E_d)^3} .
509: \label{Eq.1}
510: \en
511: The relatively small size of the charge transfer gap means that we are not
512: deep in the insulating phase and the exchange term is expected to be large.
513: Indeed experimentally the insulator is found to be in an antiferromagnetic
514: ground state.  By fitting Raman scattering to two magnon excitations 
515: \cite{SFL9025},
516: %(Sulewsky {\em et al.}, 1990), 
517: the exchange energy is found to be $J = 0.13$~eV.   This
518: is one of the largest exchange energies known and is exceeded only by that of
519: the ladder compounds which involve the same Cu-O bonding.  This value of $J$
520: is confirmed by fitting spin wave energy to theory,  where an additional ring
521: exchange terms is found 
522: \cite{CHA0177}.
523: %(Coldea {\em et al.}, 2001).
524: 
525: By substituting divalent Sr for trivalent La, the electron count on the Cu-O
526: layer can be changed in a process called doping. For example, in
527: La$_{2-x}$Sr$_x$CuO$_4$, $x$ holes per Cu is added to the layer. As seen in
528: Fig.~2, due to the large $U_d$, the hole will reside on the oxygen $p$
529: orbital.  The hole can hop via $t_{pd}$  and due to translational symmetry,
530: the holes are mobile and form a metal, unless localization due to disorder or
531: some other phase transition intervenes.  The full description of the hole
532: hopping in the three-band model is complicated, and a number of theories
533: consider this essential to the understanding of high T$_c$ superconductivity
534: \cite{E8759,VSA8781}.
535: %(Emery, 1987;  Varma {\em et al.}, 1987).
536: On the other hand, there is strong
537: evidence that the low energy physics (on a scale small compared with $t_{pd}$
538: and $E_p-E_d$) can be understood in terms of an effective one-band model, and
539: we shall follow this route in this review.   The essential insight is that
540: the doped hole resonates on the four oxygen sites surrounding a Cu and the
541: spin of the doped hole combines with the spin on the Cu to form a spin
542: singlet.  This is known as the Zhang-Rice singlet \cite{ZR8859}.
543: This state is split off by an energy of order $t^2_{pd}/(E_p-E_d)$ because the
544: singlet gains energy by virtual hopping.  On the other hand, the Zhang-Rice
545: singlet can hop from site to site.  Since the hopping is a two step process,
546: the effective hopping integral $t$ is also of order $t^2_{pd}/(E_p-E_d)$.
547: Since $t$ is the same parametrically as the binding energy of the singlet, the
548: justification of this point of view relies on a large  numerical factor for
549: the binding energy which is obtained by studying small clusters.
550: 
551: By focusing on the low lying singlet, the hole doped three-band model
552: simplifies to a one-band tight binding model on the square lattice, with an
553: effective nearest neighbor hopping integral $t$ given earlier 
554: %and an effective on-site repulsion $U$ which is of order 
555: and with $E_p-E_d$ playing a role analogous to $U$.  In the large $E_p-E_d$ 
556: limit this maps onto the $t$-$J$ model
557: \be
558: H = &P&  \left(
559: -\sum_{\langle \v i\v j \rangle,\sigma} t_{\v i\v j}c^\dagger_{\v i\sigma}c_{\v i\sigma}  \right.  \nonumber  \\ 
560: &+& J\sum_{\langle \v i\v j \rangle}
561: \left. \rule{0in}{.3in}
562: \left(
563: \v{S}_{\v i} \cdot \v{S}_{\v j} - {1\over 2}n_{\v i}n_{\v j} \right)
564: \right) P
565: \label{Eq.2}
566: \en
567: where the projection operator $P$ restricts the Hilbert space to one which
568: excludes double occupation of any site.  $J$ is given by $4t^2/U$ and we can
569: see that it is the same functional form as that of the three-band model
570: described earlier.  It is also possible to dope with electrons rather than
571: holes. The typical electron doped system is Nd$_{2-x}$Ce$_x$CuO$_{4+\delta}$
572: (NCCO). The added electron corresponds to removal of a hole from the copper
573: site in the hole picture (Fig. \ref{fig.2}), \ie the Cu ion is in the
574: $d^{10}$ configuration.  This vacancy can hop with a $t_\text{eff}$ and the mapping
575: to the one-band model is more direct than the hole doped case.  Note that in
576: the full three-band model the object which is hopping is the Zhang-Rice
577: singlet for hole doping and the Cu $d^{10}$ configuration for electron doping.
578: These have rather different spatial structure and are physically quite
579: distinct.  For example, the strength of their coupling to lattice distortions
580: may be quite different.  When mapped to the one-band model, the nearest
581: neighbor hopping $t$ has the same parametric dependence, but could have a
582: different numerical constant.  As we shall see, the value of $t$ derived from
583: cluster calculations turn out to be surprisingly similar for electron and hole
584: doping.  For a bi-partate lattice, the $t$-$J$ model with nearest neighbor $t$
585: has particle-hole symmetry because the sign of $t$ can be absorbed by changing
586: the sign of the orbital on one sublattice.  Experimentally the phase diagram
587: exhibits strong particle-hole asymmetry. On the electron doped side, 
588: the antiferromagnetic insulator
589: survives up to much higher doping concentration (up to $x \approx 0.2$) and
590: the superconducting transition temperature is quite low (about 30~K).  Many of
591: the properties  of the superconductor resemble that of the overdoped region of
592: the hole doped side and the pseudogap phenomenon, which is so prominent in the
593: underdoped region, is not observed with electron doping.  It is as though the
594: greater stability of the antiferromagnet has covered up any anomalous regime
595: that might exist otherwise.  Precisely why is not clear at  the moment.  One
596: possibility is that polaron effects may be stronger on the electron doped
597: side, leading to carrier localization over a broader range of doping.  There
598: has been some success in modeling the contrast in the single hole spectrum by
599: introducing further neighbor coupling into the one-band model which breaks
600: the particle-hole symmetry 
601: %(Shih {\em et al.}, 2004)
602: \cite{SLE0402}.  This will be discussed
603: further below.
604: 
605: We conclude that
606: the electron correlation is strong enough to produce a Mott insulator at half
607: filling. Furthermore, the one band $t$-$J$
608: %that the one-band Hubbard 
609: model captures the essence of the low
610: energy electronic excitations of the cuprates.  
611: %The Hubbard $U$ is large
612: %enough to place the system on the Mott insulator side of the model at half
613: %filling.  In that case, we may further simplify the model to the $t$-$J$
614: %model, which is the strong coupling limit of the Hubbard model.  
615: Particle-hole
616: asymmetry may be accounted for by including further neighbor hopping
617: $t^\prime$. 
618: %The parameters of this model is known.  $J= 0.13$~eV from experiment and $t
619: %\approx 3J$ from cluster and LDA band calculations (Schluter, O. Anderson).
620: %and $t^\prime = t^{\prime\prime} = ...........$ where $t^\prime$ is the
621: %hopping across the diagonal while $t^{\prime\prime}$ is the hopping to the
622: %next-nearest neighbor along a straight line. (references ????)  In
623: %particular, Anderson {\em et. al.} (1996) (J. Low Temp. Phys. {\bf 105}, 285)
624: %have shown that the low energy band structure can be parameterized  by a
625: %nearest-neighbor tight binding model involving Cu $d$ and oxygen $p$ orbitals
626: %plus the Cu $4s$ orbital.  The latter can give rise to 
627: This point of view has been tested extensively by 
628: %Hybertsen {\em et al.} (1990), 
629: \Ref{HSS9068}
630: who used {\em ab initio} local density functional theory to generate
631: input parameters for the three-band Hubbard model and then solve the spectra
632: exactly on finite clusters.  The results are compared with the low energy
633: spectra of the one-band Hubbard model and the $t-t^\prime-J$ model.  They found
634: an excellent overlap of the low lying wavefunctions for both the one-band
635: Hubbard and the $t-t^\prime-J$ model and were able to extract effective
636: parameters.  They found $J$ to be $128 \pm 5$~meV, in excellent agreement with
637: experimental values.  Furthermore they found $t \approx 0.41$~eV and $0.44$~eV
638: for electron and hole doping, respectively.  The near particle-hole symmetry
639: in $t$ is surprising because the underlying electronic states are very
640: different in the two cases, as already discussed.  Based on their results, the
641: commonly used parameter $J/t$ for the $t$-$J$ model is 1/3.  They also found a
642: significant next nearest neighbor $t^\prime$ term, again almost the same for
643: electron and hole doping.
644: 
645: \begin{figure}[t]
646: \centerline{
647: \includegraphics[width=3in]{fig4.eps}
648: %\psfig{figure=Fig.4.eps,width=3.5in}
649: }
650: %\vspace{0.5cm}
651: \caption{%Fig.4 
652: The Knight shift for YB$_2$Cu$_4$O$_8$.  It is an underdoped material with
653: $T_c = 79$K.  From 
654: %Curro {\it et al}. (1997)
655: \Ref{CIS9777}.}
656: \label{Knight}
657: \end{figure}
658: 
659: More recently, 
660: %Andersen {\em et al.}, 1996
661: \Ref{Ao9685} have pointed out that in addition
662: to the three-band model, an additional Cu $4s$ orbital has a strong influence
663: on further neighbor hopping $t^\prime$ and $t^{\prime\prime}$, where
664: $t^\prime$ is the hopping across the diagonal and $t^{\prime\prime}$ is
665: hopping to the next-nearest neighbor along a straight line.  Recently 
666: %Pavarini {\em et. al.}, 2001 
667: \Ref{PDS0103}
668: emphasized the importance of the apical oxygen in
669: modulating the energy of the Cu $4s$ orbital and found a sensitive dependence
670: of $t^\prime/t$ on the apical oxygen distance.  They also pointed out an
671: empirical correlation between optimal $T_c$ and $t^\prime/t$.  As we will
672: discuss in sections VI.D and VII, $t^\prime$ may play an important role in
673: determining $T_c$ and in explaining the difference between electron and hole
674: doping.  However, in view of the fact that on-site repulsion is the largest
675: energy scale in the problem, it would make sense to begin our modeling of the
676: cuprates with the $t$-$J$ model and ask to what extent the phase diagram can
677: be accounted for.  As we shall see, even this is not a simple task and will
678: constitute the major thrust of this review.
679: 
680: \begin{figure}[t]
681: \centerline{
682: \includegraphics[width=3in]{fig5a_old.eps}
683: }
684: \centerline{
685: \includegraphics[width=3in]{fig5b_old.eps}
686: %\psfig{figure=Fig.5a.eps,width=3.25in}
687: %\psfig{figure=Fig.5b.eps,width=3.0in}
688: }
689: \vspace{0.5cm}
690: \caption{%Fig.5 
691: (a) Knight shift data of YBCO for a variety of doping (from 
692: %Alloul {\em et al.}, 1989
693: \Ref{AOM8900}).  The zero reference level for the spin  contribution
694: is indicated by the arrow and the dashed line represents the prediction of the 2D $S = {1\over 2}$ Heisenberg model for
695: $J = 0.13$~eV. (b)~Uniform magnetic susceptibility for LSCO (from 
696: %Nakano {\em et al.}, 1994
697: \Ref{NOM9400}). The orbital contribution $\chi_0$ is shown (see
698: text) and the solid line represents the Heisenberg model prediction.}
699: \label{chi}
700: \end{figure}
701: 
702: 
703: \begin{figure}[t]
704: \centerline{
705: \includegraphics[width=2.75in]{fig6a-fixed.eps}
706: }
707: \centerline{
708: \includegraphics[width=3.1in]{fig6b-fixed.eps}
709: }
710: \vspace{0.5cm}
711: \caption{ 
712: The specific heat coefficient $\gamma$ for YBa$_2$Cu$_3$O$_{6+y}$ (top) and
713: La$_{2-x}$Sr$_x$CuO$_4$ (bottom).  Curves are labeled by the oxygen content
714: $y$ in the top figure and by the hole concentration $x$ in the bottom figure.
715: Optimal and overdoped samples are shown in the inset.  The jump in $\gamma$
716: indicates the superconducting transition.  Note the reduction of the jump size
717: with underdoping.  (From 
718: %Loram {\em et al.}, 1993, 2001
719: \Ref{LMC9340} and \Ref{LLC0159}).
720: }
721: \label{gamma}
722: \end{figure}
723: 
724: 
725: 
726: \section{Phenomenology of the underdoped cuprates} The essence of the problem
727: of doping into a Mott insulator is readily seen from Fig. \ref{fig.2}.  When a
728: vacancy is introduced into an antiferromagnetic spin background, it would like
729: to hop with amplitude $t$ to lower its kinetic energy.  However, after one hop
730: its neighboring spin finds itself in a ferromagnetic environment, at an energy
731: cost of ${3\over 2} J$ if the spins are treated as classical $S = {1\over 2}$.
732: It is clear that the holes are very effective in destroying the
733: antiferromagnetic background.  This is particularly so when $t \gg J$ when the
734: hole is strongly delocalized.  The basic physics is the competition between
735: the exchange $J$ and the kinetic energy which is of order $t$ per hole or $xt$
736: per unit area.  When $xt \gg J$ we expect the kinetic energy to win and the
737: system should be a Fermi liquid metal with weak residual antiferromagnetic
738: correlation.  When $xt \leq J$, however, the outcome is much less clear
739: because the system would like to maintain the antiferromagnetic correlation
740: while allowing the hole to move as freely as possible.  Experimentally we know
741: that N\'{e}el order is destroyed with 3\% hole doping, after which $d$-wave
742: superconducting state emerges as the ground state up to 30\% doping. Exactly how and
743: why superconductivity emerges as the best compromise is the centerpiece of the
744: high T$_c$ puzzle but we already see that the simple competition between $J$
745: and $xt$ sets the correct scale $x = J/t = {1\over 3}$ for the appearance of
746: nontrivial ground states. We shall focus our attention on the so-called
747: underdoped region, where this competition rages most fiercely.  Indeed it is
748: known experimentally that the ``normal'' state above the superconducting T$_c$
749: behaves differently from any other metallic state that we have known about up
750: to now.  Essentially an energy gap appears in some properties and not others.
751: This region of the phase diagram is referred to as the pseudogap region and is
752: well documented experimentally.  We review below some of the key properties.
753: 
754: 
755: 
756: 
757: 
758: 
759: \subsection{The pseudogap phenomenon in the normal state}
760: 
761: As seen in Fig. \ref{Knight} Knight shift measurement in the YBCO 124 compound
762: shows that while the spin susceptibility $\chi_s$ is almost temperature
763: independent between 700~K and 300~K, as in an ordinary metal, it decreases
764: below 300~K and by the time the T$_c$ of 80~K is reached, the system has lost
765: 80\% of the spin susceptibility 
766: %(Curro {\em et al.}, 1997)
767: \cite{CIS9777}.  To emphasize the
768: universality of this phenomenon, we reproduce in Fig. \ref{chi} some old data
769: on YBCO and LSCO.  Figure \ref{chi}(a) shows the Knight shift data from 
770: %Alloul {\it et al}. (1989)
771: \Ref{AOM8900}. We have subtracted the orbital contribution, which is
772: generally agreed to be 150~ppm 
773: %(Takigawa {\em et al.}, 1993)
774: \cite{THS9350}, and drawn in the
775: zero line to highlight the spin contribution to the Knight shift which is
776: proportional to $\chi_s$.  The proportionality constant is known, which allows
777: us to draw in the Knight shift which corresponds to the 2D square $S = {1\over
778: 2}$ Heisenberg antiferromagnet with $J = 0.13$~eV 
779: %(Ding {\em et al.}, 1991; Sandvik {\em et al.}, 1997)
780: \cite{DM9162,SDS9701}.   The point of this exercise is to show that in the
781: underdoped region, the spin susceptibility drops {\it below} that of the
782: Heisenberg model at low temperatures before the onset of superconductivity.
783: This trend continues even in the severely underdoped limit (O$_{0.53}$ to
784: O$_{0.41}$), showing that the $\chi_s$ reduction cannot simply be understood
785: as fluctuations towards the antiferromagnet.  Note that the discrepancy is
786: worse if $J$ were replaced by a smaller $J_\text{eff}$ due to doping, since
787: $\chi_s \sim J_\text{eff}^{-1}$.  The data seen in this light strongly point
788: to singlet formation as the origin of the pseudogap seen in the uniform spin
789: susceptibility.
790: 
791: It is worth noting that the trend shown in Fig. \ref{chi}(a) is not so
792: apparent if one looks at the measured spin susceptibility directly 
793: %(Tranquada {\em et al.}, 1988)
794: \cite{TMG8877}. This is because the van~Vleck part of the spin
795: susceptibility is doping dependent, due to the changing chain contribution.
796: This problem does not arise for LSCO, and in Fig. 4(b) we show the uniform
797: susceptibility data 
798: %(Nakano {\em et al.}, 1994)
799: \cite{NOM9400}.  The zero of the spin part is
800: determined by comparing susceptibility measurements to $^{17}$O Knight shift
801: data 
802: %{Ishida {\em et al.}, 1991}
803: \cite{IKZ9116}.  
804: %Nakano {\em et al.}, 1994
805: \Ref{NOM9400}  find an
806: excellent fit for the $x = 0.15$ sample (see Fig. 9 of this reference) and
807: determine the orbital contribution for this sample to be $\chi_0 \sim 0.4
808: \times 10^{-7}$~emu/g.  This again allows us to plot the theoretical
809: prediction for the Heisenberg model.  Just as for YBCO, $\chi_s$ for the
810: underdoped samples ($x = $0.1 and 0.08) drops below that of the Heisenberg
811: model.  In fact, the behavior of $\chi_s$ for the two systems is remarkably
812: similar, especially in the underdoped region.\footnote{ We note that a
813: comparison of $\chi_s$ for YBCO and LSCO was made by 
814: %Millis and Monien, 1993
815: \Ref{MM9310}.
816: Their YBCO analysis is similar to ours.  However, for LSCO they find  a rather
817: different  $\chi_0$ by matching the measurement above 600~K to that of the
818: Heisenberg model.  Consequently, their $\chi_s$ looks different for YBCO and
819: LSCO.  We believe their procedure is not really justified.}
820: 
821: 
822: 
823: A second indication of the pseudogap comes from the linear $T$ coefficient of
824: the specific heat, which shows a marked decrease below room temperature
825: %\cite{N9310} 
826: (see Fig. \ref{gamma}).  
827: %\begin{figure}
828: %\centerline{
829: %\includegraphics[width=3in]{Dummy_Fig5.eps} }
830: %\vspace{0.5cm}
831: %\caption{ This is Dummy Fig.5.}
832: %\end{figure}
833: Furthermore, the specific heat jump at T$_c$ is greatly reduced with
834: decreasing doping.  It is apparent that the spins are forming into singlets
835: and the spin entropy is gradually lost.  On the other hand, as shown in Fig.
836: \ref{sigma} the frequency dependent conductivity behaves very differently
837: depending on whether the electric field is in the $ab$ plane $(\sigma_{ab})$
838: or perpendicular to it $(\sigma_c)$.  
839: 
840: 
841: \begin{figure}
842: \centerline{
843: \includegraphics[width=3in]{uchida_s.eps}
844: }
845: %\vspace{0.5cm}
846: \caption{ 
847: The frequency dependent conductivity with electric field parallel to the plane ($\sigma_a(\omega)$, top figure) and perpendicular to the plane ($\sigma_c(\omega)$ bottom figure) in an underdoped YBCO crystal.  From 
848: %Uchida (1997)
849: \Ref{U9712}.
850: }
851: \label{sigma}
852: \end{figure}
853: 
854: \begin{figure}
855: \centerline{
856: \includegraphics[width=3.1in]{fig7a_s.eps}
857: }
858: \vskip 0.05in
859: \centerline{
860: \includegraphics[width=2.3in]{fig7b.eps}
861: }
862: %\vspace{0.5cm}
863: \caption{%Fig.8 
864: (a--c)~Spectra from underdoped Bi-2212 ($T_c = 85$K) taken at different $k$ points along the Fermi surface shown in (d).  Note the pullback of the spectrum from the Fermi surface as determined by the Pt reference (red lines) for $T > T_c$. (e)~Temperature dependence of the leading-edge midpoints. (from 
865: %Norman {\em et al.}, 1998
866: \Ref{NDR9857})  Bottom figure shows the temperature $T^\ast$
867: where the pseudogap determined from the leading edge first appears plotted as a function of doping for
868: Bi-2212 samples.  Triangles are determined from data such as shown in Fig.~7(a) and squares are lower
869: bound estimates. Circles show the energy gap $\Delta$ measured at $(0,\pi)$ at low temperatures. (from
870: %Campuzzano {\it et al}., 2003)
871: \Ref{CNR03}).}
872: \label{edge}
873: \end{figure}
874: 
875: At low frequencies (below 500~cm$^{-1}$) $\sigma_{ab}$ shows a typical
876: Drude-like behavior for a metal with a width which decreases with temperature,
877: but an area (spectral weight) which is independent of temperature
878: %(Santander-Syro {\em et al.}, 2002)
879: \cite{SLB0205}. Thus there is no sign of the pseudogap in
880: the spectral weight.  This is surprising because in other examples where an
881: energy gap appears in a metal, such as the onset of charge or spin density
882: waves, there is a redistribution of the spectral weight from the Drude part to
883: higher frequencies.  An important observation concerning the spectral weight
884: is that the integrated area under the Drude peak is found to be proportional
885: to $x$ 
886: %(Orenstein {\em et al.}, 1990; Cooper {\em et al.}, (1993); Uchida {\em
887: %et al.}, 1991; Padilla {\em et al.}, 2004)
888: (\cit{OTM9042}; \cit{CRK9333}; \cit{UIT9142}; \cit{PLD04}).  In the superconducting state this
889: weight collapses to form the delta function peak, with the result that the
890: superfluid density $n_s/m$ is also proportional to $x$.  It is as though only
891: the doped holes contribute to charge transport in the plane.  In contrast,
892: angle-resolved photoemission shows a Fermi surface at optimal doping very
893: similar to that predicted by band theory, with an area corresponding to
894: $(1-x)$ electrons (see Fig. \ref{edge}(d)).  With underdoping, this Fermi
895: surface is partially gapped in an unusual manner which we shall next discuss.
896: 
897: 
898: In contrast to the metallic behavior of $\sigma_{ab}$, it was discovered by
899: %Homes {\em et al.} (1993)
900: \Ref{HTL9310} that  below 300~K $\sigma_c(\omega)$ is gradually
901: reduced for frequencies below 500~cm$^{-1}$ and a deep hole is carved out of
902: $\sigma_c(\omega)$ by the time T$_c$ is reached.
903: This is clearly seen in the lower panel of Fig.
904: \ref{sigma}.
905: 
906: 
907: 
908: 
909: Finally, angle-resolved photoemission shows that an energy gap (in the form of
910: a pulling back of the leading edge of the electronic spectrum from the Fermi
911: energy) is observed near momentum $(0,\pi)$. Note that the lineshape is
912: extremely broad and completely incoherent. The onset of superconductivity is
913: marked by the appearance of a small coherent peak at this gap edge (Fig.
914: \ref{edge}).  The size of the pull back of the leading edge is the same as the
915: energy gap of the superconducting state as measured by the location of the
916: coherence peak.  As shown in Fig. \ref{edge} this gap energy increases with
917: decreasing doping, while the superconducting T$_c$ decreases. This trend is
918: also seen in tunneling data.
919: 
920: 
921: 
922: It  is possible to map out the Fermi surface by tracking the momentum of the
923: minimum excitation energy in the superconducting state for each momentum
924: direction.  Along the Fermi surface the energy gap does exactly what is
925: expected for a $d$-wave superconductor.  It is maximal near $(0,\pi)$ and
926: vanishes along the line connection $(0,0)$ and $(\pi,\pi)$ where the
927: excitation is often referred to as nodal quasiparticles.  Above T$_c$ the
928: gapless region expands to cover a finite region near the nodal point, beyond
929: which the pseudogap gradually opens as one moves towards $(0,\pi)$. This
930: unusual behavior is sometimes referred to as the Fermi arc 
931: %(Loeser {\em et al.}, 1996; Marshall {\em et al.}, 1996; 
932: %Ding {\em et al.}, 1996)
933: (\cit{LSD9625}; \cit{MDL9641}; \cit{DYC9651}).  It is
934: worth noting that unlike the anti-nodal direction (near $(0,\pi)$) the
935: lineshape is relatively sharp along the nodal direction even above T$_c$.
936: From the width in  momentum space, a lifetime which is linear  in temperature
937: has been extracted for a sample near optimal doping 
938: %(Valla {\em et al.}, 1999)
939: \cite{VFJ9910}.  
940: A narrow lineshape in the nodel direction has also been observed in LSCO
941: \cite{Yo0301} and in $Na$ doped $Ca_2CuO_2Cl_2$
942: \cite{RSK0301}. So the notion of relatively well defined nodal excitations in
943: the normal state is most likely a universal feature.
944: 
945: \begin{figure}[t]
946: \centerline{
947: \includegraphics[width=3.5in]{fig9.eps}
948: %\psfig{figure=Fig.9.eps,width=4.75in}}
949: }
950: %\vspace{0.5cm}
951: \caption{%Fig.9 
952: (A) Doping dependence of the ARPES spectra at $(0,\pi)$ at $T \ll T_c$ for
953: overdoped (OD), optimally doped (OP), and underdoped (UD) materials labeled by
954: their $T_c$'s.  (B)~The spectral weight of the coherent peak in Fig.~8(a)
955: normalized to the background is plotted vs. doping $x$.  From 
956: %Feng {\it et al}. (2000)
957: \Ref{Fo0077}.}
958: \label{ARPES}
959: \end{figure}
960: 
961: 
962: As mentioned earlier, the onset of superconductivity is marked by the
963: appearance of a sharp coherence peak near $(0,\pi)$.  The spectral weight of
964: this peak is small and gets even smaller with decreasing doping, as shown in
965: Fig. \ref{ARPES}(b).  Note that this behavior is totally different from
966: conventional superconductors.  There the quasiparticles are well defined in
967: the normal state and according to BCS theory, the sharp peak pulls back from
968: the Fermi energy and opens an energy gap in the superconducting state.
969: 
970: 
971: \begin{figure}[t]
972: \centerline{
973: \includegraphics[width=3.5in]{STM_s.eps}
974: %\psfig{figure=Fig.10.eps,width=4.75in}}
975: }
976: %\vspace{0.5cm}
977: \caption{%Fig.9
978: From 
979: %McElroy {\em et al.}, 2004
980: \Ref{MLH0405}.  STM images showing the spatial distribution of energy gaps for
981: a variety of samples which are progressively more underdoped from A to E.
982: Panel F shows the average spectrum for a given energy gap.
983: }
984: \label{STM}
985: \end{figure} 
986: 
987: In the past few years, low temperature STM data have become available, mainly
988: on Bi-2212 samples.  STM provides a measurement of the local density of states
989: $\rho(E, \v{r})$ with atomic resolution.  It is complementary to ARPES in that
990: it provides real space information but no direct momentum space information.
991: One important outcome is that STM reveals spatial inhomogeneity of the Bi-2212
992: on roughly 50 to 100~\AA \hspace{.002in} length scale, which becomes more and
993: more significant with underdoping.  As shown in Fig.~\ref{STM}(f) spectra with
994: different energy gaps are associated with different patches and with
995: progressively more underdoping, patches with large gaps become more and more
996: pre-dominant.  Since ARPES is measuring the same surface, it becomes necessary
997: to reinterpret the ARPES data with inhomogeneity in mind.  In particular, the
998: decrease of the weight of the coherent peak shown in Fig.~\ref{ARPES}(b) may
999: simply be due to a reduction of the fraction of the sample which has sharp
1000: coherent peaks. 
1001: 
1002: 
1003: A second remarkable observation by STM is that the low lying density of states
1004: ($\rho (E, \v{r})$ for $E \lesssim 10$ to 15 meV) is remarkably homogeneous.
1005: This is clearly seen in Fig. 9(f).  It is reasonable to associate this low
1006: energy excitation with the quasiparticles near the nodes.  Indeed, the low
1007: lying quasiparticles exhibit interference effects due to scattering by
1008: impurities, which is direct evidence for their spatial coherence over long
1009: distances.  Then the combined STM and ARPES data suggest a kind of phase
1010: separation in momentum space, \ie the spectra in the anti-nodal region (near
1011: $0,\pi$) is highly inhomogeneous in space whereas the quasiparticles near the
1012: nodal region are homogeneous and coherent.  The nodal quasiparticles must be
1013: extended and capable of averaging over the spatial homogeneity, while the
1014: anti-nodal quasiparticles appear more localized.  In this picture the
1015: pseudogap phenomenon mainly has to do with the anti-nodal region.
1016: 
1017: 
1018: %McElroy {\em et al.}, 2004 
1019: \Ref{MLH0405} argued that there is a limiting spectrum (the
1020: broadest curve in Fig. \ref{STM}(f)) which characterizes the extreme
1021: underdoped region at zero temperature.  It has no coherent peak at all, but
1022: shows a reduction of spectral weight up to a very high energy of 100 to 200
1023: meV.  Very recently, 
1024: %Hanaguri {\em et al.}, 2004
1025: \Ref{HLK04} provided support of this
1026: point of view in their study of Na doped Ca$_2$CuO$_2$Cl$_2$.  In this
1027: material the apical oxygen in the CuO$_4$ cage is replaced by Cl and the
1028: crystal cleaves easily.  For Na doping ranging from $x = 0.08$ to $0.12$, a
1029: tunneling spectrum very similar to the limiting spectrum for Bi-2212 is
1030: observed.  This material appears free of the inhomogeneity which plagues the
1031: Bi-2212 surface.  ARPES experiments on these crystals are becoming available
1032: \cite{RSK0301}
1033: and the combination of STM and ARPES should yield much information on the real
1034: and momentum space dependence of  the electron spectrum.  There is much
1035: excitement concerning the discovery of a static $4 \times 4$ pattern in this
1036: material, and their relation to the incommensurate pattern seen in the vortex
1037: core of Bi-2212 
1038: %(Hoffman {\em et al.}, 2002)
1039: \cite{HHL0266} and reported also in the absence
1040: of magnetic field, albeit in a much weaker form 
1041: %(Howard {\em et al.}, 2003; Vershinin {\em et al.}, 2004)
1042: \cite{HEK0333,VSO0495}.  How this spatial modulation is related to the
1043: pseudogap spectrum is a topic of current debate.
1044: 
1045: In the literature, the pseudogap behavior is often associated with anomalous
1046: behavior of the nuclear spin relaxation rate ${1\over T_1}$.  In normal metals
1047: the nuclear spin relaxes by exciting low energy particle-hole excitations,
1048: leading to the Koringa behavior, \ie ${1\over T_{1}T}$ is temperature
1049: independent.  In high T$_c$ materials, it is rather ${1\over T_1}$ which is
1050: temperature independent, and the enhanced relaxation (relative to Koringa) as
1051: the temperature is reduced is ascribed to antiferromagnetic spin fluctuations.
1052: It was found that in underdoped YBCO, the nuclear spin relaxation rate at the
1053: copper site reaches a peak at a temperature $T^\ast_1$ and decreases rapidly
1054: below this temperature 
1055: %(Warren {\em et al.}, 1989; Takigawa {\em et al.}, 1991)
1056: (\cit{WWB8993}; \cit{YIS8954}; \cit{TRH9147}).  
1057: The resistivity also shows a decrease below $T^\ast_1$.  In some
1058: literature $T^\ast_1$ is referred to as the pseudogap scale.  However, we note
1059: that $T^\ast_1$ is lower than the energy scale we have been discussing so far,
1060: especially compared with that for the uniform spin susceptibility and the
1061: $c$-axis conductivity.  Furthermore, the gap in ${1\over T_1}$ is not
1062: universally observed in cuprates. It is  not seen in LSCO.  In
1063: YBa$_2$Cu$_4$O$_8$, which is naturally underdoped, the gap in ${1\over T_1T}$
1064: is wiped out by 1\% Zn doping, while the Knight shift remains unaffected
1065: %(Zheng {\em et al.}, 2003)
1066: \cite{ZOM0391}.  It is known from neutron scattering that the low
1067: lying spin excitations near $(\pi,\pi)$ is sensitive to disorder.  Since
1068: ${1\over T_1}$ at the copper site is dominated by these fluctuations, it is
1069: reasonable that ${1\over T_1}$ is sensitive as well.  In contrast, the
1070: gap-like behavior we described thus far in a variety of physical properties is
1071: universally observed across different families of cuprates (wherever data
1072: exist) and are robust.  Thus we prefer not to consider $T^\ast_1$ as the
1073: pseudogap temperature scale.
1074: 
1075: \subsection{Neutron scattering, resonance and stripes} Neutron scattering
1076: provides a direct measure of the spin excitation spectrum.  Early work (see
1077: %Kastner {\em et al.}, 1998
1078: \Ref{KBS9897}) has shown that the long range N\'{e}el order gives
1079: way to short range order with progressively shorter correlation length with
1080: doping, so that at optimal doping, the static spin correlation length is no
1081: more than 2 or 3 lattice spacings.  Much of the early work was focused on the
1082: La$_{2-x}$Sr$_x$CuO$_4$ family, because of the availability of large single
1083: crystals.  It was found that there is enhanced spin scattering at low
1084: energies, centered around the incommensurate positions $\v{q}_0 = \left( \pm
1085: {\pi \over 2}, \pm \delta  \right)$, 
1086: %(Cheong {\em et al.}, 1991)
1087: \cite{CAM9191}.  
1088: %Yamada {\em et al.}, 1998 
1089: \Ref{Yo9865}
1090: found that $\delta$ increases systematically with doping, as
1091: shown in Fig. \ref{incomm}.  
1092: \begin{figure}[t]
1093: \centerline{
1094: \includegraphics[width=3in]{incomm.eps}
1095: %\psfig{figure=Fig.11.eps,width=4.75in}}
1096: }
1097: %\vspace{0.5cm}
1098: \caption{%Fig.10 
1099: From 
1100: %Matsuda {\em et al.}, 2000
1101: \Ref{MFY0048}.  Plot of the incommensurability $\delta$ vs.
1102: hole concentration $x$.  In the superconducting state, the open circles denote
1103: the position of the fluctuating spin density wave observed by neutron
1104: scattering.  (Data from 
1105: %Yamada {\em et al.}, 1998.
1106: \Ref{Yo9865}.)  In the insulator the spin density wave becomes static at low
1107: temperatures and its orientation is rotated by 45$^\circ$.  The dashed line
1108: $(\delta = x)$ is the prediction of the stripe model which assumes a fixed
1109: density of holes along the stripe.
1110: }
1111: \label{incomm}
1112: \end{figure} 
1113: Meanwhile it was noted that in the La$_2$CuO$_4$ family, there is a marked
1114: suppression of $T_c$ near $x = {1 \over 8}$.  This suppression is particularly
1115: strong with Ba doping, and $T_c$ is completely destroyed if some Nd is
1116: substituted for La, as in La$_{1.6-x}$Nd$_{0.4}$Sr$_x$CuO$_4$ for $x= {1 \over
1117: 8}$.  
1118: %Tranquada {\em et al.}, 1995
1119: \Ref{TSA9561}
1120: discovered static spin density wave and
1121: charge density wave order in this system, which onsets below about 50~K.  The
1122: period of the spin and charge density waves are 8 and 4 lattice constants,
1123: respectively.  The static order is modeled by a stripe picture where the holes
1124: are concentrated in period 4 charge stripes separated by spin ordered regions
1125: with anti-phase domain walls.  Recently, the same kind of stripe order was
1126: observed in La$_{1.875}$Ba$_{0.125}$CuO$_4$ 
1127: %(Fujita {\em et al.}, 2004)
1128: \cite{FGY0496}.  Note
1129: that in this model there is one hole per two sites along the charge stripe.
1130: It is tempting to interpret the low energy spin density wave observed in LSCO
1131: as a slowly fluctuating form of stripe order, even though the associated
1132: charge order (presumably dynamical also) has not yet been seen.  The most
1133: convincing argument for this interpretation comes from the observation that
1134: over a range of doping $x = 0.06$ to $x = 0.125$, the observed
1135: incommensurability $\delta$ is given precisely by the stripe picture, \ie
1136: $\delta = x$, while $\delta$ saturates at approximately ${1 \over 8}$ for $x
1137: \gtrsim 0.125$ (see Fig.\ref{incomm}).  However, it must be noted that in this
1138: interpretation, the charge stripe must be incompressible, \ie they behave as
1139: charge insulators.  Upon changing $x$, it is energetically more favorable to
1140: add or remove stripes and change the average stripe spacing, rather than
1141: changing the hole density on each stripe, which is pinned at ${1 \over 4}$
1142: filling.  It is difficult to reconcile this picture with the fact that LSCO is
1143: metallic and superconducting in the same doping range.  An alternative
1144: interpretation of the incommensurate spin scattering is that it is due to
1145: Fermi surface nesting 
1146: %(Littlewood {\em et al.}, 1993; Si {\em et al.}, 1993)
1147: \cite{LZA9387,SZL9355,TKF9355}.
1148: However, in this case the $x$ dependence of $\delta$ requires some fine
1149: tuning.  Regardless of interpretation, it is clear that in the LSCO family,
1150: there are low lying spin density wave fluctuations which are almost ready to
1151: condense.  At low temperatures, static SDW order is stabilized by Zn doping
1152: %(Kimura {\em et al.}, 1999)
1153: \cite{KHY9917}, near $x = {1 \over 8}$ 
1154: %(Wakimoto {\em et al.}, 1999)
1155: \cite{WSE9969}, and in oxygen doped systems 
1156: %(Lee {\em et al.}, 1999)
1157: \cite{LBK9943}.  However, in the
1158: latter case, there is evidence from $\mu$SR 
1159: %(Savici {\em et al.}, 2002)
1160: \cite{SFG0224} that
1161: there may be microscopic phase separations in this material (not too
1162: surprising in view of the STM data on Bi-2202).  It was also found that SDW
1163: order is stabilized in the vicinity of vortex cores 
1164: %(Kitano {\em et al.}, 2000; Lake {\em et al.}, 2001; 
1165: %Khaykovich {\em et al.}, 2002)
1166: (\cit{KYS0077}; \cit{LAC0159}; \cit{KLE0228}).
1167: 
1168: \begin{figure}[t]
1169: \centerline{
1170: \includegraphics[width=3in]{hourglass.eps}
1171: %\psfig{figure=Fig.11.eps,width=4.75in}}
1172: }
1173: %\vspace{0.5cm}
1174: \caption{%Fig.10 
1175: Neutron scattering from YBCO$_{6.5}$.  This sample has $T_c = 59$~K and the
1176: experiment was performed at 6~K (from 
1177: %Stock {\em et al.}, 2004b
1178: \Ref{SBC0471}).  Top panel
1179: refers to in-phase fluctuations between the bilayer which shows a resonance
1180: located at $(\pi,\pi)$ ($q = 0$ in the figure) and at energy 33~meV.
1181: Incommensurate peaks disperse down from the resonance.  Broad peaks also
1182: disperse upward from the resonance, forming the hourglass pattern.  Solid line
1183: is the spin wave spectrum of the insulating parent compound.  Bottom panel
1184: denotes out-of-phase fluctuations between the bilayers.
1185:  }
1186:  \label{hourglass}
1187: \end{figure} 
1188: 
1189: 
1190: 
1191: The key question is then whether the fluctuating stripe picture is special to
1192: the LSCO family or plays a significant role in all the cuprates.  Outside of
1193: the LSCO family, the spin response is dominated by a narrow resonance at
1194: $(\pi,\pi)$.  The resonance was first discovered at 41~meV for optimally doped
1195: YBCO 
1196: %(Rossat-Mignon {\em et al.}, 1991; Mook {\em et al.}, 1993)
1197: (\cit{RRV9186}; \cit{MYA9390}).  Careful subtraction of an accidentally degenerate
1198: phonon line reveals that the resonance appears only below $T_c$ at optimal
1199: doping 
1200: %(Fong {\em et al.}, 1995)
1201: \cite{FKA9516}.  Now it is known that with underdoping, the resonance moves
1202: down in energy and survives into the pseudogap regime above $T_c$.  The
1203: resonance moves smoothly to almost zero energy at the edge of the transition
1204: to N\'{e}el order in YBa$_2$Cu$_3$O$_{6.35}$ 
1205: %(Buyers {\em et al.}, 2004)
1206: \cite{Bo04} and
1207: clearly plays the role of a soft mode at that transition.
1208: 
1209: The resonance was interpreted as a spin triplet exciton bound below
1210: $2\Delta_0$ 
1211: %(Fong {\em et al.}, 1995)
1212: \cite{FKA9516}.  This idea was elaborated upon by a number of RPA calculations 
1213: %(Liu {\em et al.}, 1995; Bulurt and Scalapino, 1995; Norman, 2000, 2001; Kao
1214: %{\em et al.}, 2000; Onufrieva and Pfeuty, 2002; Brinckmann and Lee, 1999,
1215: %2002; Abanov {\em et al.}, 2002)
1216: (\cit{LZL9530}; \cit{BS9649}; \cit{N0051}; \cit{N0109}; \cit{KSL0098};
1217: \cit{OP0215}; \cit{BL9915}; \cit{BL0202}; \cit{ACE0202}).  An alternative
1218: picture making use of the particle-particle channel was proposed
1219: %(Demler and Zhang, 1995)
1220: \cite{DZ9526}.  However, as explained by 
1221: %Tchernyshyov {\em et al.} (2001)
1222: \Ref{TNC0107} and by 
1223: %Norman and Pepin (2003)
1224: \Ref{NP0347}, this theory predicts an anti-bound resonance above the
1225: two-particle continuum, which is not in accord with experiments.
1226: 
1227: Further support of the triplet exciton idea comes from the observation that
1228: incommensurate branches extend below the resonance energy 
1229: %(Bourges {\em et al.}, 2000)
1230: \cite{BSF0034}.  This behavior is predicted by RPA-type theories 
1231: %(Norman, 2000; Onufrieva and Pfuety, 2002; Brinckmann and Lee, 2002)
1232: (\cit{N0051}; \cit{OP0215}; \cit{BL0202})
1233:  in that the gap in the
1234: particle-hole continuum extends over a region near $(\pi,\pi)$, where the
1235: resonance can be formed.  With further underdoping this incommensurate branch
1236: extends to lower energies (see Fig.~\ref{hourglass}).  Now it becomes clear
1237: that the low energy incommensurate scattering previously reported for
1238: underdoped YBCO 
1239: %(Mook {\em et al.}, 2000)
1240: \cite{MDD0004} is part of this downward dispersing
1241: branch 
1242: %(Stock {\em et al.} 2004a; Pailhes {\em et al.}, 2004)
1243: (\cit{SBL0402}; \cit{PSB0409}).  
1244: 
1245: It should be
1246: noted that while the resonance is prominent due to its sharpness, its spectral
1247: weight is actually quite small, of order 2\% of the total spin moment sum rule
1248: for optimal doping and increasing somewhat with underdoping.  There is thus
1249: considerable controversy over its significance in terms of its contribution to
1250: the electron self-energy and towards pairing (see 
1251: %Norman and Pepin, 2003
1252: \Ref{NP0347}).
1253: The transfer of this spectral weight from above to below $T_c$ has been
1254: studied in detail by 
1255: %Stock {\em et al.} (2004a)
1256: \Ref{SBL0402}.  These authors emphasized
1257: that in the pseudogap state above $T_c$ in YBa$_2$Cu$_3$O$_{6.5}$, the
1258: scattering below the resonance is gapless and in fact increases in strength
1259: with decreasing temperature.  This is in contrast with the sharp drop seen in
1260: ${1 \over T_1T}$ below 150~K.  Either a gap open up at very low energy (below
1261: 4~meV) or the $(\pi,\pi)$ spins fluctuating seen by neutrons are not the
1262: dominant contribution to the nuclear spin relaxation, \ie the latter may be
1263: due to excitations which are smeared out in momentum space and undetected by
1264: neutrons.  We note that a similar discrepancy between neutron scattering
1265: spectral weight and ${1\over T_1T}$ was noted for LSCO 
1266: %(Aeppli {\em et al.}, 1995)
1267: \cite{AMH9511}.  This reinforces our view that the decrease in ${1\over T_1T}$ should
1268: not be considered a signature of the pseudogap.  We also note that an enhanced
1269: $(\pi,\pi)$ scattering together with singlet formation is just what is
1270: predicted by the $SU(2)$ theory in section XI.D.
1271: 
1272: Recently, neutron scattering has been extended to energies much above the
1273: resonance.  It is found that very broad features disperse upward from the
1274: resonance, resulting in the ``hourglass'' structure shown in
1275: Fig.~\ref{hourglass} which was first proposed by 
1276: %Bourges {\em et al.}, 2000
1277: \Ref{BSF0034}
1278: %(Hayden {\em et al.}, 2004; Stock {\em et al.}, 2004a)
1279: \cite{HMD0431,SBL0402}.  Interestingly, there
1280: has also been a significant evolution of the understanding of the neutron
1281: scattering in the LSCO family.  For a long time it has been thought that the
1282: LSCO family does not exhibit the resonance which shows up prominently below
1283: $T_c$ in YBCO and other compounds.  However, neutron scattering does show a
1284: broad peak around 50~meV which is temperature independent.  
1285: %Tranquada {\em et al.}, (2004)
1286: \Ref{TWP0434} studied La$_{1.875}$Ba$_{0.125}$CuO$_4$ which exhibits static
1287: charge and spin stripes below 50~K, and a greatly suppressed $T_c$.  Their
1288: data also exhibits an ``hourglass''-type dispersion, remarkably similar to
1289: that of underdoped YBCO.  In particular, the incommensurate scattering which
1290: was previously believed to be dispersionless now exhibits downward dispersion
1291: %(Fujita {\em et al.}, 2004)
1292: \cite{FGY0496}.  The same phenomenon is also seen in optimally
1293: doped La$_{2-x}$Sr$_x$CuO$_4$ 
1294: %(Christensen {\em et al.}, 2004)
1295: \cite{CMR0439}.  It is
1296: remarkable that in these materials known to have static or dynamic stripes,
1297: the incommensurate low energy excitations are not spin waves emanating  from $
1298: \left( {\pi \over 2} \pm \delta \right) $ as one might have expected, but
1299: instead are connected to the peak at $(\pi, \pi)$ in the hourglass fashion.
1300: %Tranquada {\em et al.} (2004)
1301: \Ref{TWP0434} fit the $\v{k}$ integrated intensity to a model
1302: of a two-leg ladder.  It is not clear how unique this fit is because one may
1303: expect high energy excitations to be relatively insensitive to details of the
1304: model.  What is emerging though is a picture of a universal hourglass shaped
1305: spectrum which is common to LSCO and YBCO families.  The high energy
1306: excitations appear common while the major difference seems to be in the
1307: re-arrangement of spectral weight at low energy.  In LBCO, significant weight
1308: has been transferred to the low energy incommensurate scattering, as shown in
1309: Fig.~\ref{tranquada}, and is associated with stripes.  In our view the
1310: universality supports the picture that 
1311: \begin{figure}[t]
1312: \centerline{
1313: \includegraphics[width=3.4in]{tranquada_s.eps}
1314: %\psfig{figure=Fig.11.eps,width=4.75in}}
1315: }
1316: %\vspace{0.5cm}
1317: \caption{%Fig.10 
1318: Neutron scattering from La$_{1.875}$Ba$_{0.125}$CuO$_4$ at 12~K $(> T_c)$
1319: (from 
1320: %Tranquada {\em et al.}, 2004)
1321: \Ref{TWP0434}).  Right panel shows the hourglass pattern
1322: of the excitation spectrum (cf Fig. \ref{hourglass}).  Solid line is a fit to
1323: a two-leg ladder spin model.  Left panel shows the momentum integrated
1324: scattering intensity.  Dashed line is a Lorenztian fit to the rising intensity
1325: at the incommensurate positions. Sharp peak at 40~meV could be a phonon.
1326:  }
1327:  \label{tranquada}
1328: \end{figure} 
1329: all the cuprates share the same short distance and high energy physics, which
1330: include the pseudogap behavior.  Stripe formation is a competing state which
1331: becomes prominent in the LSCO family, especially near $x = {1 \over 8}$, and
1332: may dominate the low energy and low temperature (below 50~K) physics.  There
1333: is a school of thought which holds the opposite view (see 
1334: %Carlson {\em et al.}, 2003)
1335: \Ref{CEK03}), that fluctuating stripes are responsible for the pseudogap
1336: behavior and the appearance of superconductivity.  From this point of view the
1337: same data have been interpreted as an indication that stripe fluctuations are
1338: also important in the YBCO family 
1339: %(Tranquada {\em et al.}, 2004)
1340: \cite{TWP0434}.  Clearly,
1341: this is a topic of much current debate.
1342: 
1343: 
1344: 
1345: 
1346: 
1347: \subsection{Quasiparticles in the superconducting state} In contrast with the
1348: anomalous properties of the normal state, the low temperature properties of
1349: the superconductor seem relatively normal.  There are two major differences
1350: with conventional BCS superconductors, however.  First, due to the proximity
1351: to the Mott insulator, the superfluid density of the superconductor is small,
1352: and vanishes with decreasing hole concentration.  Second, because the pairing
1353: is $d$-wave, the gap vanishes on four points on the Fermi surface (called gap
1354: nodes), so that the quasiparticle excitations are gapless and affect the
1355: physical properties even at the lowest temperatures.  We will focus on these
1356: nodal quasiparticles in this sub-section.
1357: 
1358: The nodal quasiparticles clearly contribute to the thermal dynamical
1359: quantities such as the specific heat.  Because their density of states  vanish
1360: linearly in energy, they give rise to a $T^2$ term which dominates the low
1361: temperature specific heat.  In practice, disorder rounds off the linear
1362: density of states, giving instead an  $\alpha T + \beta T^3$  behavior.  An
1363: interesting effect in the presence of a magnetic field
1364: was proposed by 
1365: %Volovik (1993)
1366: \Ref{V9369}.  He argued that in the presence of a vector potential or superfluid flow,
1367: the quasiparticle dispersion $E(\v{k}) = \sqrt{(\epsilon_{\v{k}}-\mu)^2 + \Delta_{\v k}^2}$ is shifted by
1368: \be
1369: E_{\v{A}}(\v{k}) = E(\v{k}) + \left(
1370: {1\over 2e} \v{\nabla}\theta - \v{A}
1371: \right) \cdot \v{j}_{\v k}
1372: \label{Eq.3}
1373: \en
1374: where $\v{j}_{\v k}$ is the current carried by ``normal state'' quasiparticles with
1375: momentum $\v{k}$ and is usually taken to be $-e {\partial \epsilon_{\v k} \over
1376: \partial\v{k}}$.  Note that since the BCS quasiparticle is a superposition of
1377: a particle and a hole, the charge is not a good quantum number.  However, the
1378: particle component with momentum $\v{k}$ and the hole component with momentum
1379: $-\v{k}$ each carry the same electrical current $\v{j}_{\v k} = -e {\partial
1380: \epsilon_{\v k} \over \partial\v{k}}$ and it makes sense to consider this to be the
1381: current carried by the quasiparticle.  Note that $\v{j}_{\v{k}}/e$ is very
1382: different from the group velocity ${\partial E(\v{k}) \over \partial\v{k}}$.
1383: 
1384: In a magnetic field which exceeds $H_{c1}$, vortices enter the sample. The
1385: superfluid flow $\v{\nabla}\theta \sim {2\pi \over r}$ where $r$ is the
1386: distance to the vortex core.  On average, ${1 \over 2}|\v{\nabla}\theta|
1387: \approx \pi/R$ where $R = (\phi_0/H)^{1/2}$ is the average spacing between
1388: vortices and $\phi_0 =  hc/2e$ is the flux quantum.  Volovik then predicts a
1389: shift of the quasiparticle spectra by $\approx ev_F (H/\phi_0)^{1/2}$ which in
1390: turn gives a contribution to the specific heat proportional to $\sqrt{H}$.
1391: This contribution has been observed experimentally
1392: %(Moler {\em et al.}, 1994)
1393: \cite{MBU9444}.
1394: 
1395: The quasiparticles contribute to the low temperature transport properties as
1396: well. 
1397: %Lee (1993)
1398: \Ref{L9387} considered the frequency-dependent conductivity
1399: $\sigma(\omega)$ due to quasiparticle excitations.  In the low temperature
1400: limit, he found that the low frequency limit of the conductivity is universal
1401: in the sense that it does not depend on impurity strength, but only on the
1402: ratio $v_F/v_\Delta$ where $v_\Delta$ is the velocity of the nodal
1403: quasiparticle in the direction of the maximum gap $\Delta_0$, \ie 
1404: $\sigma(\omega \rightarrow 0) = {e^2 \pi v_F \over h v_\Delta}$, if $v_F \gg
1405: v_\Delta$.  This result was derived within the self-consistent $t$-matrix
1406: approximation and can easily be understood as follows.  In the presence of
1407: impurity scattering, the density of states at zero energy becomes finite.  At
1408: the same time, the scattering rate is proportional to the self-consistent
1409: density of states.  Since the conductivity is proportional to the density of
1410: states and inversely to the scattering rate, the impurity dependence cancels.
1411: 
1412: 
1413: \begin{figure}[t]
1414: \centerline{
1415: \includegraphics[width=3.5in]{fig10_new_s.eps}
1416: %\psfig{figure=Fig.11.eps,width=4.75in}}
1417: }
1418: %\vspace{0.5cm}
1419: \caption{%Fig.10 
1420: Figure from 
1421: %Sutherland {\em et al.}, 2003
1422: \Ref{SHH0320}. Doping dependence of the
1423: superconducting gap $\Delta_0$ obtained from the quasiparticle velocity
1424: $v_\Delta$ using eq. (\ref{Eq.4}) (filled symbols).  Here we assume $\Delta = \Delta_0
1425: \cos2\phi$, so that $\Delta_0 = \hbar k_Fv_\Delta/2$, and we plot data for
1426: YBCO alongside Bi-2212 
1427: %(Chiao {\em et al.}, 2000)
1428: \cite{CHL0054} and Tl-2201 
1429: %(Proust {\em et al.}, 2003)
1430: \cite{PBH0203}.  For comparison, a BCS gap of the form $\Delta_{\bf BCS} = 2.14
1431: k_BT_c$ is also plotted.  The value of the energy gap in Bi-2212, as
1432: determined by ARPES, is shown as measured in the superconducting state
1433: %(Campuzzano {\em et al.}, 1999)
1434: \cite{CDN9909} and the normal state 
1435: %(Norman {\em et al.}, 1998)
1436: \cite{NDR9857} (open symbols).  The thick dashed line is a guide to the eye.
1437:  }
1438:  \label{gap}
1439: \end{figure} 
1440: 
1441: The frequency-dependent $\sigma(\omega)$ is difficult to measure and it was
1442: realized that thermal conductivity $\kappa$ may provide a better test of the
1443: theory because according to the Wiedemann-Franz law, $\kappa/T$ is
1444: proportional to the conductivity and should be universal.  Unlike
1445: $\sigma(\omega)$, thermal conductivity does not have a superfluid contribution
1446: and can be measured at DC.  More detailed considerations by 
1447: %Durst and Lee (2000)
1448: \Ref{DL0070}
1449:  show that $\sigma(\omega)$ has two non-universal corrections: one due
1450: to backscattering effects, which distinguishes the transport rate from the
1451: impurity rate which enters the density of state; and a second one due to Fermi
1452: liquid corrections.  On the other hand, these corrections do not exist for
1453: thermal conductivity.  Consequently, the  Wiedemann-Franz law is violated, but
1454: the thermal conductivity per layer is truly universal and is given by
1455: \be
1456: {\kappa \over T} = {k^2_B \over 3\hbar c}  \left(
1457: {v_F \over v_\Delta }+ {v_\Delta \over v_F}
1458: \right)
1459: \label{Eq.4}
1460: \en
1461: We note that this result is obtained within the self-consistent $t$-matrix
1462: approximation which is expected to break down if the impurity scattering is
1463: strong, leading to localization effects.  The localization of nodal
1464: quasiparticles is a complex subject.  Due to particle-hole mixing in the
1465: superconductor, zero energy is a special point and quasiparticle localization
1466: belongs to a different universality class 
1467: %(Senthil \& Fisher 1999)
1468: \cite{SF9993}  from the standard ones.  Senthil and Fisher also pointed out
1469: that since quasiparticles carry well defined spin, the Wiedemann-Franz law for
1470: spin conductivity should hold and spin conductivity should be universal.  We
1471: note that 
1472: %Durst and Lee (2000) 
1473: \Ref{DL0070} argued that Fermi liquid corrections enter the spin conductivity,
1474: but we now believe their argument on this point is faulty.
1475: 
1476: Thermal conductivity has been measured to mK temperatures in a variety of YBCO
1477: and BCCSO samples.  The universal nature of $\kappa/T$ has been demonstrated
1478: by studying samples with different Zn  doping and showing that $\kappa/T$
1479: extrapolates to the same constant at low temperatures 
1480: %(Taillefer {\em et al.}, 1997)  
1481: \cite{TLG9783}
1482: A magnetic field dependence analogous to the Volovik effect for the
1483: specific heat has also been observed 
1484: %(Chiao, {\em et al.}, 2000)
1485: \cite{CHL0054}  Using
1486: eq.~(\ref{Eq.4}), the experimental data can be used to extract the ratio
1487: $v_F/v_\Delta$.  In the case of BCCSO where photoemission data for $v_F$ and
1488: the energy gap is available, the extracted ratio $v_F/v_\Delta$ is in
1489: excellent agreement with ARPES results, assuming a simple $d$-wave
1490: extrapolation of the energy gap from the node to the maximum gap $\Delta_0$.
1491: In particular, the trend that $\Delta_0$ increases with decreasing doping $x$
1492: is directly observed as a decrease of $v_F/v_\Delta$ extracted from
1493: $\kappa/T$.  A summary of the data is shown in Fig.~\ref{gap} 
1494: %(Sutherland {\em et al.}, 2003)
1495: \cite{SHH0320}.    Results of such systematic studies strongly support the
1496: notion that in clean samples the nodal quasiparticles behave exactly as one
1497: expects for well defined quasiparticles in a $d$-wave superconductor.  We
1498: should add that in LSCO the ratio $v_F/v_\Delta$ extracted from $\kappa/T$
1499: seems anomalously small, suggesting that strong disorder may be playing a role
1500: here to invalidate eq.~(\ref{Eq.4}).
1501: 
1502: \begin{figure}[tb]
1503: \centerline{
1504: \includegraphics[width=3in]{fig10.eps}
1505: %\psfig{figure=Fig.11.eps,width=4.75in}}
1506: }
1507: %\vspace{0.5cm}
1508: \caption{%Fig.11 
1509: The London penetration depth measured in a series of YBCO film 
1510: %displayed in three different ways.  The top inset shows the raw data
1511: %$\lambda^{-2}$ plotted vs. temperature.  These date are converted to
1512: %$\rho_s(T)$ in absolute scale and $\rho_s(T) - \rho_s(0)$ is plotted in the
1513: %bottom inset to highlight the temperature dependent part.  The main plot
1514: %shows the data normalized to $\rho_s(T=0)$ vs. $T/T_c$.  From Stacjic {\it et
1515: %al}., 2003.  
1516: with different oxygen concentration and T$_c$'s.  The plot shows
1517: $\lambda^{-2}$ plotted vs.  temperature.  Data provided by T.R. Lemberger and
1518: published in 
1519: %Boyce {\it et al.} (2000)
1520: \Ref{BSL0061}. 
1521: }
1522: \label{Lemberger}
1523: \end{figure} 
1524: 
1525: 
1526: %Lee and Wen (1997) 
1527: \Ref{LW9711}
1528: pointed out that the nodal quasiparticles also manifest
1529: themselves in the linear $T$ dependence of the superfluid density.  They
1530: showed that by treating them as well defined quasiparticles in  the sense of
1531: Landau, a general expression of the linear $T$ coefficient can be written
1532: down, independent of the microscopic origin of the superconductivity.  We have
1533: \be
1534: {n_s(T) \over m}  = {n_s(0) \over m} - {2\ln 2 \over \pi} \alpha^2 \left( {v_F  \over v_\Delta}
1535: \right) T
1536: \label{Eq.5}
1537: \en
1538: The only assumption made is that the quasiparticles carry an electric current
1539: \be
1540: \v{j}(\v{k}) = -e \alpha \v{v}_F
1541: \label{Eq.6}
1542: \en
1543: where $\alpha$ is a phenomenological Landau parameter which was left out in
1544: the original Lee-Wen paper but added in by 
1545: %Millis {\it et al} (1998)
1546: \Ref{MGI9842}.  While
1547: the linear $T$ dependence is well known in the conventional BCS theory of a
1548: $d$-wave superconductor, the same theory gives $n_s/m$ of order unity.  It is
1549: therefore useful to write $n_s$ in this phenomenological way, and choose
1550: $n_s(T=0)$ to be of order $x$ as we discussed in section III.A.  The key
1551: question raised by eq.~(\ref{Eq.6}) is whether $\alpha$ depends on $x$ or not.  There
1552: is experimental evidence that the linear $T$ coefficient of
1553: $n_s(T)/m$ which is directly related to London penetration depth measurements,
1554: is almost independent of $x$ for $x$ less than optimal doping.
1555: Figure~\ref{Lemberger} shows data obtained for a series of thin films of YBCO
1556: %(Boyce {\em et al.}, 2000; Stajic {\em et al.}, 2003).   
1557: \cite{BSL0061,SIL0320}
1558: %The raw data with $\lambda$ measured in absolute unit is shown in the upper
1559: %inset.  The lower inset shows the temperature dependent part of $n_s$, again
1560: %in absolute unit, and shows a remarkable insensitivity to the doping $x$.  We
1561: %emphasize that in order to make this comparison, it is essential to obtain an
1562: %absolute determination of $\lambda(T)$, which turns out to be a difficult
1563: %experimental problem for bulk samples. 
1564: The thin film data are in full agreement with earlier  but less extensive data
1565: on bulk crystals 
1566: %(Bonn, 1996)
1567: \cite{Bo9695}.
1568: However, we note that very recent data on severely underdoped YBCO crystal
1569: ($T_c<20$K)  show that $\frac{d(n_s/m)}{dT}$ is roughly linear in $T_c$
1570: \cite{BTO04}.
1571: 
1572: Since $v_F/v_\Delta$ is known to go to a constant for small $x$ (and, indeed,
1573: decreases with decreasing $x$), the independence of the linear $T$ term in
1574: $n_s/m$ on $x$ means that $\alpha$ approaches a constant for small $x$.  By
1575: combining with $v_F/v_\Delta$ extrapolated from thermal conductivity,
1576: $\alpha^2$ has been estimated to be 0.5 (see 
1577: %Ioffe and Millis (2002)
1578: \Ref{IM0259} for an
1579: excellent summary).  This is an important result because it states that
1580: despite the proximity to the Mott insulator, the nodal quasiparticles carry a
1581: current which is similar to that of the tight-binding Fermi liquid band.  We
1582: note that the simplest microscopic theory which gives correctly $n_s(T=0)$ to
1583: be proportional to $x$ is the slave-boson  mean-field theory to be discussed
1584: in section IX.B.  That theory predicts $\alpha$ to be proportional to $x$ and
1585: the resulting $x^2T$ term is in strong disagreement with experiment.  The
1586: search for a microscopic theory which gives correctly both $n_s(T=0)$ and the
1587: linear $T$ term is one of the open problems that faces us today.
1588: 
1589: 
1590: 
1591: The unusual combination of a small $n_s(T=0)$ and a large linear $T$ reduction
1592: due to quasiparticles has a number of immediate  consequences.  Simply by
1593: extrapolating the linear $T$ dependence, we can conclude that $n_s$ vanishes
1594: at the temperature scale proportional to $x$ and T$_c$ must be bound by it.
1595: Furthermore, at $T_c$ the number of quasiparticles which are thermally excited
1596: is still small, and not sufficient to close the gap as in standard BCS theory.
1597: Thus the transition must not be thought of as a gap-closing transition, and
1598: the effect of an energy gap must persist considerably above $T_c$.  This can
1599: potentially explain at least part of the pseudogap phenomenon.  As we shall
1600: see in the next section, when combined with phase fluctuations, the
1601: quasiparticle excitations explain the magnitude of $T_c$ in the underdoped
1602: cuprates and account for a wide phase fluctuation region above $T_c$, but not
1603: the full pseudogap phenomenon.
1604: 
1605: The disconnect between the gap energy $\Delta_0$ and $kT_c$ introduces two
1606: length scales, $\xi_0 = \hbar v_F/\Delta_0$ and $R_2 = \hbar v_F/kT_c$, where
1607: $kT_c$ is proportional to $x$.  Around a vortex, the supercurrent induces a
1608: population of quasiparticles by the Volovik effect, and in analogy to eq.
1609: (\ref{Eq.5})
1610: causes a reduction in $n_s$.  
1611: %Lee and Wen (1997) 
1612: \Ref{LW9711} show that at a radius of
1613: $R_2$ the circulating supercurrent exceeds the critical current and inside
1614: that radius the superconductor looses its phase stiffness.  They suggest that
1615: the system becomes normal once the large core radius $R_2$ overlaps and
1616: $H_{c2} \approx \phi_0/R_2^2$, in contrast with $H_{c2}^\ast \approx
1617: \phi_0/\xi_0^2$ as in conventional BCS theory.  Note that $H_{c2}$ decreases
1618: while $H_{c2}^\ast$ increases with underdoping.  Experimentally the resistive
1619: transition to the normal state indeed takes place at an $H_{c2}$ which
1620: decreases with decreasing $T_c$.  However, there are signs that vortices
1621: survive above this magnetic field up to $H_{c2}^\ast$, as will be discussed in
1622: section. V.B.
1623: 
1624: Finally, we comment on suggestions in the literature that classical
1625: fluctuations of the superconducting phase can lead to a linear reduction of
1626: $n_s$ at low temperatures 
1627: %(Carlson {\em et al}, 1999)
1628: \cite{CKE9912}.  Just as in the case of lattice displacements, such
1629: fluctuations must be treated quantum mechanically at low temperatures (as
1630: phonons in that case) to avoid the $3 k_B$ low temperature limit for the
1631: specific heat.  In the case of phonons, the characteristic temperature scale
1632: is the phonon frequency.  In the case of the superconductor, the phase mode is
1633: pushed up to the plasma frequency by long-range  Coulomb interaction.
1634: Nevertheless, due to the coupling to the low-lying particle-hole excitations,
1635: the cross-over from classical to quantum fluctuations must be treated with
1636: some care.  
1637: %Paramekanti {\em et al}, 2000, 2002 and Benfatto {\em et al.}, 2001
1638: \Ref{PRR0086}, \citeyear{P0221} and \Ref{BCC0113}
1639:  have calculated that the cross-over happens at quite a high temperature
1640: scale and we believe the low-temperature linear reduction of $n_s$ is entirely
1641: due to thermal excitations of quasiparticles.
1642:  
1643: \section{Introduction to RVB and a simple explanation of the pseudogap} 
1644: 
1645: We explained in the last section that the N\'{e}el spin order is incompatible
1646: with hole hopping.  The question is whether there is another arrangement of
1647: the spin which achieves a better compromise between exchange energy and the
1648: kinetic energy of the hole.  For $S={1\over 2}$ it appears possible to take
1649: advantage of the special stability of the singlet state. The ground state of
1650: two spins $S$ coupled with antiferromagnetic Heisenberg exchange is a spin
1651: singlet with energy $-S(S+1)J$.  Compared with the classical large  spin
1652: limit, we see that quantum mechanics provides an additional stability in the
1653: term unity in $(S+1)$ and this contribution is strongest for $S={1\over 2}$.
1654: Let us consider a one-dimensional spin chain.  A N\'{e}el ground state with
1655: $S_z = \pm {1\over 2}$ gives an energy of $-{1\over 4}J$ per site.  On the
1656: other hand, a simple trial wavefunction of singlet dimers already gives a
1657: lower energy of $-{3\over 8}J$ per site.  This trial wavefunction breaks
1658: translational symmetry and the exact ground state can be considered to be a
1659: linear superposition of singlet pairs which are not limited to nearest
1660: neighbors, resulting in a ground state energy of 0.443~J.  In a square and
1661: cubic lattice the N\'{e}el energy is $-{1\over 2}J$ and $-\frac34J$ per site,
1662: respectively, while the dimer variational energy stays at $-{3\over 8}J$.  It
1663: is clear that in a 3D cubic lattice, the N\'{e}el state is a far superior
1664: starting  point, and in two dimensions the singlet state may present a serious
1665: competition.  Historically, the notion of a linear superposition of spin
1666: singlet pairs spanning different ranges, called the resonating valence bond
1667: (RVB), was introduced by 
1668: %Anderson (1973)
1669: \Ref{A7353} and 
1670: %Fazekas and Anderson (1974)
1671: \Ref{FA7432} as a
1672: possible ground state for the $S={1\over 2}$ antiferromagnetic Heisenberg
1673: model on a triangular lattice.  The triangular lattice is of special interest
1674: because an Ising-like ordering of the spins is frustrated.  Subsequently, it
1675: was decided that the ground state forms a $\sqrt{3} \times \sqrt{3}$
1676: superlattice where the moments lie on the same plane and form $120^\circ$
1677: angles between neighboring sites 
1678: %(Huse and Elser, 1988)
1679: \cite{HE8831}.  Up to now there is no known spin Hamiltonian with full $S(U2)$
1680: spin rotational symmetry outside of one dimension which is known to have an
1681: RVB ground state.  However, see section X.H for examples which either violate
1682: spin rotation or which permit charge fluctuations.
1683: 
1684: The N\'{e}el state has long range order of the staggered magnetization and an
1685: infinite degeneracy of ground states leading to Goldstone modes which are
1686: magnons.  In contrast, the RVB state is a unique singlet ground state with
1687: either short range or power law decay of antiferromagnetic order.  This state
1688: of affairs is sometimes referred to as a spin liquid.  However, the term spin
1689: liquid is often used more generally to denote any kind of short range or power
1690: law decay, \ie the absence of long range order, even when the unit cell is
1691: doubled, either spontaneously or explicitly. For example, the ladder system
1692: has two states per unit cell and in the limit of strong coupling across the
1693: rung, the ground state is naturally a spin singlet with short range
1694: antiferromagnetic order.  Another example is the spontaneously dimerized
1695: ground state for the frustrated spin chains when the next-nearest neighbors
1696: exchange $J^\prime$ is sufficiently large.  This kind of ground state is more
1697: properly called a valence band solid and is smoothly connected to spin
1698: singlet ground states often observed for systems with an even number of
1699: electrons per unit cell, the extreme example being Si.  Thus we think it is
1700: better to reserve the term spin liquid to cases where there is an odd number
1701: of electrons per unit cell.
1702: 
1703: Soon after the discovery of high T$_c$ superconductors, 
1704: %Anderson (1987)
1705: \Ref{A8796}
1706: revived the RVB idea and proposed that with the introduction of holes the
1707: N\'{e}el state is destroyed and the spins form a superposition of singlets.
1708: The vacancy can hop in the background of what he envisioned as a liquid of
1709: singlets and a better compromise between the hole kinetic energy and the spin
1710: exchange energy may be achieved.  Many elaborations of this idea followed,
1711: but here we argue that the basic physical picture described above gives a
1712: simple account of the pseudogap phenomenon.  The singlet formation explains
1713: the decrease of the uniform spin susceptibility and the reduction of the
1714: specific heat $\gamma$. The vacancies are responsible for transport in the
1715: plane.  The conductivity spectral weight in the $ab$ plane is given by the
1716: hole concentration $x$ and is unaffected by the singlet formation.  On the
1717: other hand, for $c$-axis conductivity, an electron is transported between
1718: planes.  Since an electron  carries spin ${1\over 2}$, it is necessary to
1719: break a singlet.  This explains the gap formation in $\sigma_c(\omega)$ and
1720: the energy scale of this gap should be correlated with that of the uniform
1721: susceptibility.  In photoemission, an electron leaves the solid and reaches
1722: the detector, the pull back of the leading edge simply reflects the energy
1723: cost to break a singlet.
1724: 
1725: A second concept associated with the RVB idea is the notion of spinons and
1726: holons, and spin charge separations.  Anderson  postulated that the spin
1727: excitations in an RVB state are $S={1\over 2}$ fermions which he called
1728: spinons.  This is in contrast with excitations in a N\'{e}el state which are
1729: $S = 1$ magnons or $S = 0$ gapped singlet excitations.
1730: 
1731:  
1732:  \begin{figure}[t]
1733: \centerline{
1734: \includegraphics[width=3in]{fig11.eps}
1735: %\psfig{figure=Fig.11.eps,width=4.75in}}
1736: }
1737: %\vspace{0.5cm}
1738: \caption{%Fig.11 
1739: A cartoon representation of the RVB liquid or singlets.  Solid bond represents a spin singlet
1740: configuration and circle represents a vacancy.  In (b) an electron is removed from the plane in
1741: photoemission or $c$-axis conductivity experiment.  This necessitates the breaking of a singlet. }
1742:  \label{RVB}
1743: \end{figure}  
1744: 
1745:  
1746: Initially the spinons are  suggested to form a Fermi surface, with Fermi
1747: volume equal to that of $1-x$ fermions.  Later it was proposed that the Fermi
1748: surface is gapped to form $d$-wave type structure, with maximum gap near
1749: $(0,\pi)$.  This $\v{k}$ dependence of the energy gap is needed to explain the
1750: momentum dependence observed in photoemission.
1751: 
1752: The concept of spinons is a familiar one in one-dimensional spin chains where
1753: they are well understood to be domain walls.  In two dimensions the concept is
1754: a novel one  which does not involve domain walls.  Instead, a rough physical
1755: picture is as follows.  If we assume a background of short range singlet
1756: bonds, forming the so-called short-range RVB state, a cartoon of the spinon is
1757: shown in Fig. \ref{RVB}.  If the singlet bonds are ``liquid,'' two $S={1\over
1758: 2}$ formed by breaking a single bond can drift apart, with the liquid of
1759: singlet bonds filling in the space between them.  They behave as free
1760: particles and are called spinons.  The concept of holons follows naturally
1761: %(Kivelson, Rokhsar, and Sethna 1988) 
1762: \cite{KRS8765}
1763: as the vacancy left over by removing a
1764: spinon.  A holon carries charge $e$ but no spin.
1765: 
1766: 
1767: \section{Phase fluctuation vs. competing order} 
1768: 
1769: One of the hallmarks of doping
1770: a Mott insulator is that the spectral weight of the frequency dependent
1771: conductivity $\sigma(\omega)$ should go to zero in the limit of small doping.
1772: Indeed, $\sigma(\omega)$ shows a Drude-like peak at low frequencies and its
1773: area was shown to be proportional to the hole concentration 
1774: %(Orenstein {\em et al.}, 1990; Cooper {\em et al.}, 1993; 
1775: %Uchida {\em et al.}, 1991; Padilla {\em et al.}, 2004)
1776: (\cit{OTM9042}; \cit{CRK9333}; \cit{UIT9142}; \cit{PLD04}).
1777: Results from exact diagonalization of small samples  are
1778: consistent with a Drude weight of order $xt$ 
1779: %(Dagotto {\em et al.}, 1992)
1780: \cite{DMO9241}.
1781: When the metal becomes superconducting, all the spectral weight collapses into
1782: a $\delta$-function if the sample is in the clean limit. The London
1783: penetration depth for field penetration perpendicular to the $ab$ plane is
1784: given by
1785: \be
1786: \lambda_\perp^{-2}
1787: %(T=0) 
1788: = {4\pi n_s^{3d}e^2\over m^\ast c^2} ,
1789: \label{Eq.7}
1790: \en
1791: where $n_s^{3d}/m^\ast$ is the spectral weight and $n_s^{3d}$ is the $3d$
1792: superfluid charge density.  As an example, if we take $\lambda_\perp =
1793: 1600$~Angstrom for YBa$_2$Cu$_3$O$_{6.9}$,  and take $n_s^{3d}$ to be the hole
1794: density, we find from eq. (\ref{Eq.7}) $m^\ast \approx 2m_e$ which corresponds
1795: to an effective hopping $t^\ast = {1\over 3} t$.  The notion that
1796: $\lambda_\perp^{-2}$ is proportional to $xt$ is also predicted by slave-boson
1797: theory, as will be discussed in section IX.B.
1798: 
1799:  
1800:  \begin{figure}[t]
1801: \centerline{
1802: \includegraphics[width=3.3in]{fig12.eps}
1803: %\psfig{figure=Fig.12.eps,width=4.75in}}
1804: }
1805: %\vspace{0.5cm}
1806: \caption{%Fig.12 
1807: The phase stiffness $T_\theta$ measured at different frequencies $(T_\theta =
1808: \hbar^2 n_s/m^\ast)$.  The solid dots give the bare stiffness obtained by
1809: extrapolation to infinite frequency.  $T_c$ of this sample is 74~K.This is
1810: where the phase stiffness measured at low frequency would vanish according to
1811: BKT theory.  Note the linear decrease of the bare stiffness with $T$ which
1812: extends considerably above $T_c$.  This decrease is due to thermal excitations
1813: of nodal quasiparticles.  Inset shows the time scale of the phase fluctuation.
1814: Hatched region denotes ${\hbar \over \tau} = kT_c$. From 
1815: %Corson {\it et al}. (1999)
1816: \Ref{CMO9921}. 
1817: }
1818: \label{Corson}
1819: \end{figure} 
1820:  
1821:  
1822: %Uemura {\em et al.}, 1989
1823: \Ref{ULS8917}
1824: discovered empirically a linear relation between
1825: $\lambda_\perp^{-2}$ measured by $\mu$SR and the superconducting T$_c$.   He
1826: interpreted this relation as indicative of Bose condensation of holes, since in
1827: two dimensions the Bose-Einstein condensation temeprature is proportional to
1828: the areal density.  Since $\lambda_{\perp}^{-2}$ is proportional to the 3d
1829: density, in principle, some adjustment for the layer spacing should be made.
1830: Furthermore, $\lambda_{\perp}^{-2}$ is highly sensitive to disorder, and it is
1831: now known that in many systems, not all the spectral weight collapses to the
1832: $\delta$-function, \ie  some residual normal conductivity is left, presumably
1833: due to inhomogeneity 
1834: %(Basov {\em et al.}, 1994; Corson {\em et al.}, 2000)
1835: \cite{BPH9465,COO0069}.
1836: Thus the Uemura plot should be viewed as providing a qualitative trend, rather
1837: than a quantitative relation.  Nevertheless, it is important in that it draws
1838: a relationship between T$_c$ and carrier density.
1839: 
1840: 
1841: \begin{figure}[t]
1842: \centerline{
1843: \includegraphics[width=2.5in]{123layer.eps}
1844: }
1845: \vspace{0.5cm}
1846: \caption{ 
1847: Schematic plot of the phase stiffness $T_\theta = \hbar^2 n_s/m^\ast$ for superconductors with $N$ coupled layers.  The linear decrease with temperature is due to the thermal excitation of quasiparticles.  The transition temperatures $T_{cN}, N=1, 2, 3$ are estimated by the interception with the BKT line $T_\theta = 8 T/\pi$.
1848: }
1849: \label{Tc}
1850: \end{figure}
1851:  
1852:  
1853: \subsection{A theory of $T_c$} 
1854: 
1855: The next important step was taken by 
1856: %Emery and Kivelson (1995)
1857: \Ref{EK9534}, who noted that it is the superfluid density which controls
1858: the phase stiffness of the superconducting order parameter $\Delta =
1859: |\Delta|e^\theta$, \ie the energy density cost of a phase twist is
1860: \be
1861: H = {1\over 2} K_s^0 (\nabla\theta)^2
1862: \label{Eq.8}
1863: \en
1864: Here the superscript on $K^0_s$ denotes the bare stiffness on a short distance
1865: scale.  For two-dimensional layers the stiffness $K_s =
1866: \hbar^2(n_s/2)/2m^\ast$, \ie the kinetic energy of Cooper pairs.  The
1867: spectral weight $n_s/m^\ast$ and the stiffness are given by
1868: \be
1869: K_s = {1 \over 4}{\hbar^2n_s \over m^\ast}
1870: = {1\over 4} {\hbar^2n_s^{3d}c_0 \over m^\ast} 
1871: \label{Eq.9}
1872: \en
1873: where $c_0$ is the spacing between the layers 
1874: %Combining Eq. (5.1) and (5.3) $\rho_s^0$ can  be written
1875: and using Eqs. (\ref{Eq.7}) and (\ref{Eq.9}), can be directly measured in terms of
1876: $\lambda_\perp$.  If $K_s^0$ is small due to the proximity to the Mott
1877: insulator, then phase fluctuation is strong and the T$_c$ in the underdoped
1878: cuprates may be governed by phase fluctuations.  The theory of phase
1879: fluctuations in two dimensions is well understood due to the work of
1880: %Berezinskii (1971)
1881: \Ref{B7144}
1882: and 
1883: %Kosterlitz and Thouless (1973)
1884: \Ref{KT7381}.  The BKT transition is
1885: described by the thermal unbinding of vortex anti-vortex pairs.  The energy of
1886: a single vortex is given by
1887: \be
1888: E_\text{vortex} = E_c + 2\pi K_s^0 \ln (L/\xi_0)
1889: \label{Eq.10}
1890: \en
1891: where $L$ is the sample size, $\xi_0$ is the BCS coherence length which serves
1892: as a short distance cut-off, and $E_c$ is the core energy.  For vortex
1893: anti-vortex pairs, the sample size $L$ is replaced by the separation of the
1894: pairs.  The vortex unbinding transition is driven by the balance between this
1895: energy and the entropy which also scales logarithmically with the vortex
1896: separation.  At T$_c$, $K_s$ is predicted to jump between zero and a finite
1897: value $K_s(T_c)$ given by a universal relation 
1898: \be
1899: kT_c = (\pi/2)K_s(T_c) =
1900: {\pi \over 8} {\hbar^2 n_s \over m^\ast}
1901: \label{Eq.11}
1902: \en
1903: %(Nelson and Kosterlitz, 1977)
1904: \cite{NK7701}.  The precise value of T$_c$ depends on $K_s(T=0)$ and weakly on
1905: the core energy.  In  the limit of very large core energy, $kT_c = 1.5 K_s^0$,
1906: whereas for an  XY model on a square lattice 
1907: %whose long distance behavior is also governed by Eq. (5.2), 
1908: $E_c$ is basically zero if $\xi_0$ in eq. (\ref{Eq.10}) is replaced by the lattice
1909: constant and $T_c = 0.95 K_s^0$.  Thus $K_s^0$ should give a reasonable guide
1910: to T$_c$ in the phase fluctuation scenario.  Emery and Kivelson estimated
1911: $K^0_s$ from $\lambda_\perp$ data for a variety of materials and concluded that
1912: $K_s^0$ is indeed on the scale of T$_c$.  However, they assumed that each
1913: layer is fluctuating independently, even for systems with strongly Josephson
1914: coupled bi-layers. Subsequent work using microwave conductivity has confirmed
1915: the BKT nature of the phase transition, but concluded that in BSCCO, it is the
1916: bi-layer which should be considered as a unit, \ie the superconducting phase
1917: is strongly correlated between the two layers of a bi-layer 
1918: %(Corson {\it et al}. 1999)
1919: \cite{CMO9921}.  This increases the $K_s^0$ estimate by a factor of 2.  For
1920: example, for $\lambda_\perp = 1600$ Angstroms, Emery and Kivelson quoted
1921: $K_s^0$ to be 145~K for YBCO.  This should really be replaced by 290~K, a
1922: factor of 3 higher than T$_c$.
1923: %On the other hand, it is realized that thermal excitation of nodal
1924: %quasiparticles reduces the superfluid stiffness in a linear fashion with
1925: %increasing temperature (Lee \& Wen) and $K_s^0$ should include this
1926: %reduction, which is easily a factor of 2.  Indeed, as discussed in section
1927: %III.B, Lee and Wen pointed out that in a Landau quasiparticle treatment, the
1928: %linear $T$  term is independent of doping concentration and the extrapolation
1929: %of $\rho_s^0$ to zero gives a reasonable estimate of T$_c$.  This can be seen
1930: %in Figs. 10 and 12.  There is now good evidence that T$_c$ for underdoped
1931: %cuprates can be  understood and even quantitative fits to the temperature
1932: %dependence of $\lambda_\perp^{-2}$ have been attempted along these lines.
1933: %(G. William, 2003)
1934: 
1935: 
1936: 
1937: 
1938: We can get around this difficulty by realizing that $K_s$ is reduced by
1939: thermal excitation of quasiparticles and the bare $K_s^0$ in the BKT  theory
1940: should include this effect.  In Section III.B we showed empirical evidence
1941: that the linear $T$ coefficient of $n_s(T)$ is relatively independent of $x$.
1942: The bare $K_s^0$ is measured as the high frequency limit in a microwave
1943: experiment 
1944: %(Corson {\it et al}., 1999)
1945: \cite{CMO9921}.  As seen in Fig. \ref{Corson}, the
1946: bare phase stiffness $T_\theta^0 \equiv \hbar^2 n_s^0/m^\ast$ continues to
1947: decrease linearly with $T$ above T$_c$ = 74~K.  Given the universal relation
1948: eq.~(\ref{Eq.11}), an estimate of T$_c$ can be obtained by the interception of
1949: the straight line $T_\theta = (8/\pi)kT$ with the bare stiffness.  This yields
1950: an estimate of the BKT transition temperature of $\approx 60$~K.  The somewhat
1951: higher actual T$_c$ of 74~K is due to three dimensional ordering effects
1952: between bi-layers.  Now we can extend this procedure to a multi-layer
1953: superconductor.  In Fig.\ref{Tc} we show schematically $T_\theta^0 = \hbar^2
1954: n_s^0/m^\ast$ plot of single-layer, bi-layer and tri-layer systems $(N=1,2,3)$
1955: assuming that the layers are identical.  We expect $n_s^0(T=0)$, which is the
1956: areal density per $N$ layers, to scale linearly with $N$.  On the other hand,
1957: the linear $T$ slope also scales with $N$, because the number of thermally
1958: excited quasiparticles per area scale with $N$.  The extrapolated ``T$_c$'s''
1959: are therefore the same.  Now we may estimate $T_c(N)$ from the interception of
1960: the line $T_\theta^0 = (8/\pi)kT$.  We see that $T_c$ increases monotonically
1961: with $N$, but much slower than linear.  This trend is in agreement with what
1962: is seen experimentally, notably in the Tl and Hg compounds.  As $N$ increases
1963: further, the assumption that the layers are identical breaks down as the
1964: charge density of each layer begins to differ.  We therefore conclude that the
1965: combination of phase fluctuations and the thermal excitation of $d$-wave
1966: quasiparticles can account for $T_c$ in underdoped cuprates, including the
1967: qualitative trend as a function of the number of layers within a unit cell.
1968: 
1969:  
1970:  
1971:  
1972:  This theory of $T_c$ receives confirmation from measurement of the oxygen
1973: isotope effect of $T_c$ and on the penetration depth.  It is found that there
1974: is substantial isotope effect on the $n_s/m^\ast$ for both underdoped and
1975: optimally doped YBCO films.
1976: %, which the experimentalists interpret as an effect on $n_s/m^\ast$.  
1977: On the other hand, there is significant isotope effect on
1978: $T_c$ in underdoped YBCO 
1979: %(Khasanov {\em et al,}, 2003)
1980: \cite{KSC0317}, but no effect on
1981: optimally doped samples 
1982: %(Khasanov {\em et al.}, 2004)
1983: \cite{Ko0402}.  Setting aside the
1984: origin of the isotope effect on $n/m^\ast$, the remarkable doping dependence
1985: of the isotope effect on $T_c$ is readily explained in our theory, since $T_c$
1986: is controlled by $n_s/m^\ast$ in the underdoped but not in the overdoped
1987: region.  In fact, a more detailed examination of the data for two underdoped
1988: samples show that $n_s(T)/m^\ast$ appears to be shifted down by a constant
1989: when O$^{16}$ is replaced by O$^{18}$.  This suggests that there is no isotope
1990: effect on the temperature dependent term in eq.~\ref{Eq.5} which depends on
1991: $v_F$.  This is consistent with direct ARPES measurements 
1992: %(Gweon {\em et al.}, 2004)
1993: \cite{GSZ0487}.  Thus the data is consistent with an isotope effect only on
1994: the zero temperature spectral weight $n_s(0)/m^\ast$.  The latter is a
1995: complicated many body property of the ground state which is not simply related
1996: to the effective mass of the quasiparticles in the naive manner.
1997: 
1998:  
1999: \subsection{Cheap vortices and the Nernst effect} 
2000: 
2001: %Emery and Kivelson, 1995
2002: \Ref{EK9534},
2003: also suggested that the notion of strong phase fluctuations may provide an
2004: explanation of the pseudogap phenomenon.  They proposed that the pairing
2005: amplitude is formed at a temperature $T_{MF}$ which is much higher than T$_c$
2006: and the region between $T_{MF}$ and T$_c$ is characterized by robust pairing
2007: amplitude and energy gap.  
2008: 
2009: This leaves open the microscopic origin of the robust pairing amplitude and
2010: high $T_{MF}$ but we shall argue that even as phenomenology, phase
2011: fluctuations alone cannot be the full explanation of the pseudogap.  Since
2012: T$_c$ is driven by the unbinding of vortices, let us examine the vortex energy
2013: more carefully.  As an extreme example, let us suppose $T_{MF}$ is described
2014: by the standard BCS theory.  The vortex core energy in BCS theory is estimated
2015: as $E_c \approx {\Delta_0^2 \over E_Fa^2} \xi_0^2$ where $\Delta_0$ is the
2016: energy gap, $\Delta_0^2/(E_Fa^2)$ is the condensation energy per area, and
2017: $\xi_0^2$ is the core size.  Using $\xi_o = {v_F \over \Delta_0}$, we conclude
2018: that $E_c \approx E_F$ in BCS theory, an enormous energy compared with T$_c$.
2019: Even if we assume $E_c$ to be of order of the exchange energy $J$ or the mean
2020: field energy $T_{MF}$, it is still much larger than T$_c$.  We already note
2021: that in BKT theory, T$_c$ is relatively insensitive to the core energy. Now we
2022: emphasize that despite the insensitivity of $T_c$ to $E_c$, the physical
2023: properties above $T_c$ are very sensitive to the core energy. This is because
2024: BKT theory is an asymptotic long-distance theory which becomes simple in the
2025: limit of dilute vortex or large $E_c$.  The typical vortex spacing which is
2026: $n_v^{-{1\over 2}}$ where the vortex density $n_v$ goes as $e^{-E_c/kT}$.
2027: Vortex unbinding happens on a renormalized length scale, \ie the typical
2028: spacing between {\em free} vortices, which is much larger than $n_V^{-{1\over
2029: 2}}$.  As a result, the physics of the system above T$_c$ is very sensitive to
2030: $E_c$.  If $E_c \gg kT_c$, vortices are dilute and the system will behave like
2031: a superconductor for all measurements performed on a reasonable spatial or
2032: temporal scale.  However, except for the close vicinity of  T$_c$, the
2033: pseudogap region is not characterized by strong superconducting fluctuations,
2034: but rather behaves like a metal.  Thus a large vortex core energy can be ruled
2035: out.  The core energy must be small, of the order T$_c$, \ie it is comparable
2036: to the second term in eq.~(\ref{Eq.10}).  The notion of ``cheap'' vortices has
2037: two important consequences.  First, it is clear that the amplitude fluctuation
2038: and phase fluctuation are controlled by the same energy scale, $kT_c$.  This
2039: is because the vortex core is a region where the pairing amplitude vanishes
2040: and, in addition, the phase $\theta$ winds by $2\pi$.  If we do away with the
2041: phase winding and retain the amplitude fluctuation, this should cost even less
2042: energy.  Thus the temperature scale where vortices proliferate is also the
2043: scale where amplitude fluctuation proliferates.  Then the notion of strong
2044: phase fluctuations is applicable only on a temperature scale of say 2~T$_c$
2045: and this scale must become small as $x$ becomes small.  Thus phase fluctuation
2046: cannot explain a pseudogap phenomenon which extends to finite $T$ in the small
2047: $x$ limit.
2048: 
2049: 
2050: Second, the notion of a cheap vortex means that there is a non-superconducting
2051: state which is very close in energy.  In an ordinary superconductor, the core
2052: can be thought of as a  patch of normal metals with a finite density of states
2053: at the Fermi level.  The reason the core energy is large is because the energy
2054: gained by opening up an energy gap is lost.  In underdoped and in slightly
2055: overdoped cuprates there is experimental evidence from STM tunneling into the
2056: core that the energy gap is retained inside the core 
2057: %(Maggio-April {\em et al.}, 1995; Pan {\em et al.}, 2000)
2058: \cite{MRE9554,PHG0036}.  The large peak in the density state predicted for
2059: $d$-wave BCS theory 
2060: %(Wang and MacDonald, 1995)
2061: \cite{WM9576} is simply not there.  The nature of the state in the core, which
2062: one can think of as a competing state to the superconductor, is highly
2063: nontrivial and is a topic of current debate.
2064: 
2065: The above discussion is summarized by a schematic phase diagram shown in Fig.
2066: \ref{Nernst}.  A temperature scale of about 2~T$_c$ in the underdoped region
2067: marks the range of phase fluctuation.  This is the region where the picture
2068: envisioned by 
2069: %Emery and Kivelson (1995)
2070: \Ref{EK9534} may be valid.  Here the phase is
2071: locally well defined and vortices are identifiable objects.  Indeed, this is
2072: the region where a large Nernst effect has been measured 
2073: %(Wang {\em et al.}, 2001, 2002, 2003)
2074: \cite{WYK0119,WOX0203,WOO0386}.
2075: The Nernst effect is the voltage transverse to a thermal
2076: gradient in the presence of a magnetic field perpendicular to the plane.  It
2077: is exquisitely sensitive to the presence of vortices, because vortices drift
2078: along the thermal gradient and produce the phase winding which supports a
2079: transverse voltage by the Josephson effect.  A large Nernst signal has been
2080: taken to be strong evidence for the presence of well-defined vortices above
2081: T$_c$ 
2082: %(Wang {\em et al.}, 2001, 2002, 2003)
2083: \cite{WYK0119,WOX0203,WOO0386}.   At higher temperatures,
2084: vortices overlap and the Nernst signal smoothly crosses over to that
2085: describable by Gaussian fluctuation of superconducting amplitude and phase
2086: %(Usshishkin {\em et al.}, 2002)
2087: \cite{USH0201}.  Very recently, the identification of the
2088: Nernst region with fluctuating superconductivity was confirmed by the
2089: observation of diamagnetic fluctuations which persist up to the same
2090: temperature as the onset of the Nernst signal 
2091: %(Wang {\em et al.}, 2004)
2092: \cite{WLN04}.
2093: 
2094: \begin{figure}
2095: \centerline{
2096: \includegraphics[scale=1.2]{fig13.eps}
2097: %\psfig{figure=Fig.13.eps,width=3.25in}}
2098: }
2099: %\vspace{0.5cm}
2100: \caption{%Fig.13 
2101: Schematic phase diagram showing the phase fluctuation regime where the Nernst effect is large. 
2102: Note that this regime is a small part of the pseudogap region for small doping. }
2103: \label{Nernst}
2104: \end{figure} 
2105: 
2106: 
2107: It remains necessary to explain  why the resistivity looks metallic-like in
2108: this temperature range and does not show the strong magnetic field dependence
2109: one ordinarily expects for flux flow resistivity in the presence of thermally
2110: excited vortices. The explanation may lie in the breakdown of the standard
2111: Bardeen-Stephen model of flux flow resistivity.  Here the vortices have
2112: anomalously low dissipation because in contrast to BCS superconductors, there
2113: are no states inside the core to dissipate.  
2114: %Ioffe and Millis, 2002b 
2115: \Ref{IM0213}
2116: proposed
2117: that the vortices are fast and yield a large flux flow resistivity.  In the
2118: two fluid model, the conductivity is the sum of the flux flow conductivity
2119: (the superfluid part) and the quasiparticle conductivity (the normal part).
2120: The small flux flow conductivity is quickly shorted out by the nodal
2121: quasiparticle contributions, and the system behaves like a metal, but with
2122: carriers only in the nodal region.  This is also reminiscent of the Fermi arc
2123: picture.  Unfortunately, a more detailed modeling requires an understanding
2124: of the state inside the large core radius $R_2$ introduced in section III.C
2125: which is not available up to now.
2126: 
2127: Instead of generating vortices thermally, one can also generate them by
2128: applying a magnetic field.  
2129: %Wang {\em et al.}, 2003
2130: \Ref{WOO0386} have applied fields up to
2131: 45~T and found evidence that the Nernst signal remains large beyond that field
2132: in the underdoped samples.  They estimate that the field needed to suppress
2133: the Nernst signal to be of order $H^\ast_{c2} \approx \phi_0/\xi_0^2$ where
2134: $\phi_0 \approx \hbar v_F/\Delta_0$.  This is the core size consistent with
2135: what is reported by STM tunneling experiments.  At the same time, the field
2136: needed for a resistive transition is much lower.  Recently 
2137: %Sutherland~{\em et al.}, 2004 
2138: \Ref{SHH04}
2139: showed that in YBCO$_{6.35}$ superconductivity is destroyed 
2140: by annealing or by applying a modest magnetic field.
2141: Beyond this point the material is a thermal metal, with a
2142: thermal conductivity which is unchanged from the superconducting side, where
2143: it is presumably due to nodal quasiparticles and described by
2144: eq.~(\ref{Eq.4}). Thus this field induced metal may be coexisting with pairing
2145: amplitude and may be a very interesting new metallic state.
2146: 
2147: What is the nature of the gapped state inside the vortex core as revealed by
2148: STM tunneling and how is it related to the pseudogap region?  A popular
2149: notion is that the vortex core state is characterized by a competing order.  A
2150: variety of competing order has been proposed in the literature.  An early
2151: suggestion was that the core has antiferromagnetic order and an explicit model
2152: was constructed based on the $SU(5)$ model of 
2153: %Zhang (1997)
2154: \Ref{Z9789}
2155: % (Arovas {\em et al.}, 1997)
2156: \cite{ABK9771}.  However, this particular version has been criticized for its
2157: failure to take into account the strong Coulomb repulsion and the proximity to
2158: the Mott insulator 
2159: %(Greiter, 1997; Baskaran and Anderson, 1998)
2160: (\cit{G9798}; \cit{BA9880}).  Recently,
2161: more phenomenological version based on Landau theory have been proposed
2162: %(Demler, Sachdev and Zhang, 2000; Chen {\em et al.}, 2004)
2163: (\cit{DSZ0102}; \cit{CCA0416}) where the
2164: antiferromagnetism may be incommensurate.  As the temperature is raised into
2165: the pseudogap regime, vortices proliferate and their cores overlap and,
2166: according to this view, the pseudogap is characterized by fluctuating
2167: competing order.  The dynamic stripe picture 
2168: %(Carlson {\em et al.}, 2003)
2169: \cite{CEK03} is an example of this point of view.  Another proposal for
2170: competing order is for orbital currents 
2171: %(Varma, 1997; Chakravarty {\em et al.}, 2001)
2172: (\cit{V9754}; \cit{CLM0203}).  In this case the competing order is proposed to persist
2173: in the pseudogap region but is ``hidden'' from detection because of
2174: difficulties of coupling to the order.  Finally, as discussed earlier, the
2175: recent observation of checkerboard patterns in the vortex core and in some
2176: underdoped cuprates has inspired various proposals of charge density ordering. 
2177: 
2178: Most of the proposals for competing order are phenomenological in nature.  For
2179: example, the proximity of $d$-wave superconductivity to antiferromagnetism is
2180: simply assumed as an experimental fact.  However, from a microscopic point of
2181: view, the surprise is that $d$-wave superconductivity turns out to be the
2182: winner of this competition.  Our goal is a microscopic explanation of both the
2183: superconducting and the pseudogap states.  We shall give a detailed proposal
2184: for the vortex core in section XII.C.  Here we mention that while our proposal
2185: also calls for slowly varying orbital currents in the core, this fluctuating
2186: order is simply one manifestation of a quantum state.  For example, enhanced
2187: antiferromagnetic fluctuation is another manifestation.  As discussed in
2188: section VI.C, this picture is fundamentally different from competing states
2189: described by Landau theory.  In the pseudogap phase, vortices proliferate and
2190: overlap and all orders become very sort range.  Apart from characterizing this
2191: state as a spin liquid (or RVB), the only possibility of order is a subtle
2192: one, called topological/quantum order.  These concepts are described in
2193: section X and a possible experimental consequence is described in section
2194: XII.E.
2195: 
2196: 
2197: 
2198: 
2199: 
2200: \subsection{Two kinds of pseudogaps} Since the pseudogap is fundamentally a
2201: cross-over phenomenon, there is a lot of confusion about the size of the
2202: pseudogap and the temperature scale where it is observed.  Upon surveying the
2203: experimental literature, it seems to us that we should distinguish between two
2204: kinds of pseudogaps.  The first is clearly due to superconducting
2205: fluctuations.  The energy scale of the pseudogap 
2206: is the same as the low temperature superconducting gap
2207: and it extends over a surprisingly large range of temperatures above $T_c$.
2208: This is what we called the Nernst region in the last section.  This kind of
2209: pseudogap has been observed in STM tunneling, where it is found that a
2210: reduction of the density of states persists above $T_c$ even in overdoped
2211: samples 
2212: %(Kugler {\em et al.}, 2001)
2213: \cite{KFR0111}.  We believe the pull-back of the leading
2214: edge observed in ARPES shown in Fig.~\ref{edge}(a) should be understood along these
2215: lines.  There is another kind of pseudogap which is associated with singlet
2216: formation.  A clear signature of this phenomenon is the downturn in uniform
2217: spin susceptibility shown in Figs. \ref{Knight} and \ref{chi}.  The
2218: temperature scale of the onset is high and increases up to 300 to 400~K with
2219: underdoping.  The energy scale associated with this pseudogap is also very
2220: large, and can extend up to 100~meV or beyond.  For example, the onset of the
2221: reduction of the $c$-axis conductivity (which one may interpret as twice the
2222: gap) has been reported to exceed 1000~cm$^{-1}$.  This is also the energy
2223: scale one associates with the limiting STM tunneling spectrum observed in
2224: highly underdoped Bi-2212 (Fig. \ref{STM}(f)) and in Na doped
2225: Ca$_2$CuO$_2$Cl$_2$ 
2226: %(Hanaguri {\em et al.}, 2004)
2227: \cite{HLK04}.  The gap in these spectra
2228: is very broad and ill defined.  In the ARPES literature it is described as the
2229: ``high-energy pseudogap'' (see 
2230: %Damascelli {\em et al.}, 2003
2231: \Ref{DHS0373}) or the ``hump''
2232: energy.  These spectra evolve smoothly into that  of the insulating parent.
2233: This is most clearly demonstrated in Na-doped Ca$_2$CuO$_2$Cl$_2$ and the
2234: ARPES spectrum near the antinodal point looks remarkably similar to that seen
2235: by STM 
2236: %(Ronning {\em et al.}, 2003)
2237: \cite{RSK0301}.  Examples of this kind of a spectrum can
2238: be seen in the samples UD46 and UD30 shown in Fig. \ref{ARPES}(a).  In
2239: contrast to the low energy pseudogap, a coherent quasiparticle peak is never
2240: seen at these very high energies when the system enters the superconducting
2241: state.  Instead, weak peaks may appear at lower energies, but judging from the
2242: STM data, these may be associated with inhomogeneity.
2243: %Examples of this kind of spectra can be seen in the samples UD46 and UD30
2244: %shown in Fig. \ref{ARPES}(a).  
2245: In this connection, we point out that the often quoted $T^\ast$ line shown in
2246: Fig. \ref{edge} is actually a combination of the two kinds of pseudogaps. The
2247: solid triangles marking the onset of the leading edge refer to the fluctuating
2248: superconductor gap, while the solid squares are lower bounds based on the
2249: observation of the ``hump.''  Another example of this difference is that in
2250: LSCO, the superconducting gap is believed to be much smaller and the pull back
2251: of the leading edge is not seen by ARPES.  On the other hand, the singlet
2252: formation is clearly seen in Fig. \ref{Knight}(b) and the broad hump-like
2253: spectra is also seen by ARPES 
2254: %(Zhou {\em et al.}, 2004)
2255: \cite{ZYL0481}.
2256: 
2257: We note that in contrast to superconducting fluctuations which extend across
2258: the entire doping range but are substantially reduced for overdoped samples,
2259: the onset of singlet formation seems to end rather abruptly near optimal
2260: doping.  The Knight shift is basically temperature independent just above
2261: $T_c$ in optimally doped and certainly in slightly overdoped samples 
2262: %(Takigawa {\em et al.}, 1993; Horvatic {\em et al.}, 1993)
2263: (\cit{THS9350}; \cit{HBB9348}).  For this reason, we propose
2264: that the pseudogap line and the Nernst line may cross in the vicinity of
2265: optimal doping, as sketched in Fig. \ref{Nernst}.  In this connection it is
2266: interesting to note that the pseudogap has also been seen inside the vortex
2267: core 
2268: %(Maggio-Aprile {\em et al.}, 1995; Pan {\em et al.}, 2000)
2269: \cite{MRE9554,PHG0036}.  By
2270: definition, this is where the superconducting amplitude is suppressed to zero
2271: and the gap is surely not associated with the pairing amplitude.  We have
2272: argued that the gap offers a glimpse of the state which lies behind the
2273: pseudogap associated with singlet formation.  It is interesting to note that
2274: the gap in the vortex core has been reported in a somewhat overdoped sample
2275: %(Hoogenboom {\em et al.}, 2001)
2276: \cite{HKR0101}.  It is as though at zero temperature the
2277: state with a gap in the core is energetically favorable compared with the
2278: normal metallic state up to quite high doping.  It will be interesting to
2279: extend these measurements to even more highly overdoped samples to see when
2280: the gap in the vortex core finally fills in.  At the same time, it will be
2281: interesting to extend the tunneling into the vortex core in overdoped samples
2282: to higher temperatures, to see if the gap will fill in at some temperature
2283: below $T_c$.
2284: 
2285:  
2286: 
2287: \section{Projected trial wavefunctions and other numerical results} 
2288: 
2289: In the original RVB article, \Ref{A8796} 
2290: %(1987) 
2291: proposed a projected trial wavefunction as a description of the RVB state.
2292: \be
2293: \Psi = P_G|\psi_0\rangle
2294: \label{Eq.12}
2295: \en
2296: where $P_G = \prod_{\v i}(1-n_{\v i\up}n_{\v i\down})$ is called the Gutzwiller
2297: projection operator.  It has the effect of suppressing all amplitudes in
2298: $|\psi_0\>$ with double occupation, thereby enforcing the constant of the $t$-$J$
2299: model exactly.  The unprojected wavefunction contains variational parameters
2300: and its choice is guided by mean-field theory (see section XIII).  The full
2301: motivation for the choice of $|\psi_0\>$ becomes clear only after the discussion
2302: of mean-field theory, but we discuss the projected wavefunction first because
2303: the results are concrete and the concepts are simple. The projection operator
2304: is too complicated to be treated analytically, but properties of the trial
2305: wavefunction can be evaluated using Monte Carlo sampling.
2306: 
2307:  
2308:   
2309:  
2310: \subsection{The half-filled case} 
2311: 
2312: We shall first discuss the half-filled case,
2313: where the problem reduces to the Heisenberg model.  While the original
2314: proposal was for $|\psi_0\>$ to be the $s$-wave BCS wavefunction, it was soon
2315: found that the $d$-wave BCS state is a better trial wavefunction, \ie
2316: consider
2317: \begin{align}
2318: H_d & = 
2319: -\sum_{\langle \v i\v j \rangle ,\sigma}  \left(
2320: \chi_{\v i\v j} f_{\v i\sigma}^\dagger f_{\v i\sigma} + c.c.  \right)-
2321: \sum_{\v i,\sigma} 
2322: \mu f_{\v i\sigma}^\dagger f_{\v i\sigma} +
2323: \nonumber\\
2324:  & +\sum_{\langle \v i\v j \rangle} 
2325: \left[\Delta_{\v i\v j} \left( f_{\v i\up}^\dagger f_{\v j\down}^\dagger -
2326: f_{\v i\down}^\dagger f_{\v j\up}^\dagger \right) + c.c.\right]
2327: \label{Eq.13}
2328: \end{align}
2329: where $\chi_{\v i\v j} = \chi_0$ for nearest neighbors, and $\Delta_{\v i\v j} = \Delta_0$
2330: for $\v j =\v i + \hat{x}$ and $-\Delta_0$ for $j = i + \hat{y}$.  The eigenvalues
2331: are the well known BCS spectrum
2332: \be
2333: E_{\v k} = \sqrt{(\epsilon_{\v{k}}-\mu)^2 + \Delta_{\v{k}}^2}
2334: \label{Eq.14} 
2335: \en
2336: where 
2337: \be
2338: \epsilon_{\v{k}} = -2 \chi_0 \left(
2339: \cos k_x + \cos k_y
2340: \right)
2341: \label{Eq.15}
2342: \en
2343: \be
2344: \Delta_{\v{k}} = 2 \Delta_0 \left(
2345: \cos k_x - \cos k_y
2346: \right)
2347: \label{Eq.16}
2348: \en
2349: At half filling, $\mu = 0$ and $|\psi_0\>$ is the usual BCS wavefunction 
2350: $|\psi_0\>
2351: = \prod_{\v{k}} \left( u_{\v{k}} + v_{\v{k}} f_{\v{k}\up}^\dagger
2352: f_{-\v{k}\down}^\dagger \right) |0\rangle.  $
2353: 
2354: 
2355: A variety of mean-field wavefunctions were soon discovered which give
2356: identical energy and dispersion.  Notable among these is the staggered flux
2357: state 
2358: %(Affleck and Marston, 1988)
2359: \cite{AM8874}.  In this state the hopping $\chi_{\v i\v j}$ is
2360: complex, $\chi_{\v i\v j} = \chi_0 \exp \left( i (-1)^{i_x+j_y} \Phi_0 \right) $,
2361: and the phase is arranged in such a way that it describes free fermion hopping
2362: on a lattice with a fictitious flux $\pm 4 \Phi_0$ threading alternative
2363: plaquettes.  Remarkably, the eigenvalues of this problem are identical to that
2364: of the $d$-wave superconductor given by eq.~(\ref{Eq.14}), with
2365: \be
2366: \tan \Phi_0 = {\Delta_0 \over \chi_0}  .
2367: \label{Eq.17}
2368: \en
2369: The case $\Phi_0 = \pi/4$, called the $\pi$ flux phase, is special in that it
2370: does not break the lattice translation symmetry.  As we can see from
2371: eq.~(\ref{Eq.17}),
2372: the corresponding $d$-wave problem has a very large energy gap and its
2373: dispersion is shown in Fig.~\ref{flux}.  The key feature is that the energy
2374: gap vanishes at the nodal points located at $\left( \pm{\pi\over 2},
2375: \pm{\pi\over 2} \right)$.  Around the nodal points the dispersion rises
2376: linearly, forming a cone which resembles the massless Dirac spectrum.  For the
2377: $\pi$ flux state the dispersion around the node is isotropic.  For $\Phi_0$
2378: less than $\pi /4$ the gap is smaller and the Dirac cone becomes progressively
2379: anisotropic.    The anisotropy can be characterized by two velocities, $v_F$
2380: in the direction towards $(\pi,\pi)$ and $v_\Delta$ in the direction towards
2381: the maximum gap at $(0,\pi)$.
2382: 
2383: 
2384: 
2385: The reason various mean-field theories have the same energy was explained by
2386: %Affleck {\it et al}. ,1988 and Dagotto {\em et al.}, 1988
2387: \Ref{AZH8845} and \Ref{DFM8826}  as being due to a
2388: certain $SU(2)$ symmetry.  We defer a full discussion of the $SU(2)$ symmetry to
2389: section X but we only mention here that it corresponds to the following
2390: particle-hole transformation
2391: \be
2392: f_{\v i\up}^\dagger &\rightarrow& \alpha_{\v i} f_{\v i\up}^\dagger + \beta_{\v i} f_{\v i\down} \label{Eq.18} \\ \nonumber
2393: f_{\v i\down} &\rightarrow& -\beta_{\v i}^\ast f_{\v i\up}^\dagger + \alpha_{\v i}^\ast f_{\v i\down}  .
2394: \en
2395: Note that the spin quantum number is conserved.  It describes the physical
2396: idea that adding a spin-up fermion or removing a spin-down fermion are the
2397: same state after projection to the subspace of singly occupied fermions.  It
2398: is then not a surprise to learn that the Gutzwiller projection of the $d$-wave
2399: superconductor and that of the staggered flux state gives the same trial
2400: wavefunction, up to a trivial overall phase factor, provided $\mu = 0$ and
2401: eq.~(\ref{Eq.17}) is satisfied.  A simple proof of this is given by 
2402: %Zhang {\it et al}, 1988 
2403: \Ref{ZGR8836}.  The energy of this state is quite good.  The best estimate for the
2404: ground state energy of the square lattice Heisenberg antiferromagnet which is
2405: a N\'{e}el ordered state is $\langle S_{\v i} \cdot S_{\v j} \rangle = -0.3346$~J
2406: %(Trivedi and Ceperley, 1989; Runge, 1992)
2407: (\cit{TC8937}; \cit{R9292}).  The projected $\pi$ flux state
2408: %(Gros, 1988; Yokoyama and Ogata, 1996)
2409: \cite{G8831,YO9615} gives $-0.319$J, which is excellent
2410: considering that there is no variational parameter.  
2411:  
2412: \begin{figure}[t]
2413: \centerline{
2414: \includegraphics[width=3in]{fig14.eps}
2415: %\psfig{figure=Fig.14.eps,width=4.0in}}
2416: }
2417: %\vspace{0.5cm}
2418: \caption{%Fig.14 
2419: The energy dispersion of the $\pi$ flux phase.  Note the massless Dirac spectrum at the nodal
2420: points at $\left( \pm {\pi \over 2}, \pm{\pi \over 2} \right)$.}
2421: \label{flux}
2422: \end{figure}
2423: 
2424:  
2425:  
2426: We note that the projected $d$-wave state has power law decay for the
2427: spin-spin correlation function.  The equal time spin-spin correlater decays as
2428: $r^{-\alpha}$ where $\alpha$ has been estimated to be 1.5 \cite{I00,PRT0453}.
2429: This projection has considerably enhanced the spin correlation
2430: compared with the exponent of 4 for the unprojected state.  One might expect a
2431: better trial wavefunction by introducing a sublattice magnetization in the
2432: mean-field Hamiltonian.  A projection of this state gives an energy which is
2433: marginally better than the projected flux state, $-0.3206$J.  It also has a
2434: sublattice magnetization of 84\% which is too classical.  The best trial
2435: wavefunction is one which combines staggered flux and sublattice magnetization
2436: before projection 
2437: %(Gros, 1989a,b; Lee and Feng, 1988)
2438: \cite{G8831,G8853,LF8809}.  
2439: It gives an energy of $-0.332$~J and a sublattice magnetization of about 70\%,
2440: both in excellent agreement with the best estimates.
2441: 
2442: 
2443: \begin{figure}[t]
2444: \centerline{
2445: \includegraphics[width=3in]{fig15a.eps}
2446: }
2447: %\centerline{
2448: %\includegraphics[width=2.5in]{fig15b.eps}
2449: %%\psfig{figure=Fig.15a.eps,width=3.0in}
2450: %%\psfig{figure=Fig.15b.eps,width=2.75in}}
2451: %%\vspace{0.5cm}
2452: %}
2453: \caption{%Fig.15 
2454: (a) Comparison of the energy of various projected trial wavefunctions. From 
2455: %Ivanov (2003)
2456: \Ref{I0365}.
2457: (b)~The condensation energy estimated from the difference of the projected $d$-wave superconductor and
2458: the projected staggered flux state. From 
2459: %Ivanov and Lee (2003)
2460: \Ref{IL0301}.}
2461: \label{Ivanov}
2462: \end{figure} 
2463: 
2464: \subsection{Doped case} 
2465: 
2466: In the presence of a hole, the projected wavefunction
2467: eq.~(\ref{Eq.12}) has been studied for a variety of mean-field states $\psi_0$.  Here
2468: $P_G$ stands for a double projection: the amplitudes with double occupied
2469: sites are projected out and only amplitudes with the desired number of holes
2470: are kept.  The ratio $\Delta_0/\chi_0$, $\mu/\chi_0$ and $h_s/\chi_0$, where
2471: $h_s$ is the field conjugate to the sublattice magnetization, are the
2472: variational parameters.  It was found that the best state is a projected
2473: $d$-wave superconductor and the sublattice magnetization is nonzero for $x  <
2474: x_c$, where $x_c = 0.11$ for $t/J = 3$. 
2475: %(Yokoyama and Ogata, 1996)
2476: \cite{YO9615}  The
2477: energetics of various state are shown in Fig.~(\ref{Ivanov}(a)).  It is
2478: interesting to note that  the projected staggered-flux state always lies above
2479: the projected $d$-wave superconductor, but the energy difference is small and
2480: vanishes as $x$ goes to zero, as expected.  The staggered-flux state also
2481: prefers antiferromagnetic order for small $x$, and the critical $x_c^{SF}$ is
2482: now 0.08, considerably less than that for the projected $d$-wave
2483: superconductor.  The energy difference between the projected flux state and
2484: projected $d$ superconductor (with antiferromagnetic order) is shown in
2485: Fig.~(\ref{Ivanov}(b)).  As we can see from Fig.~(\ref{Ivanov}(a)), inclusion
2486: of AF will only give a small enhancement of the energy difference for small $x
2487: \leq 0.05$.  The projected staggered flux state is the lowest energy
2488: non-superconducting state that has been constructed so far.  For $x > 0.18$,
2489: the flux $\Phi_0$ vanishes and this state connects smoothly to the projected
2490: Fermi sea, which one ordinarily thinks of as the normal state.  It is then
2491: tempting to think of the projected staggered flux state as the ``normal''
2492: state in the underdoped region $(x < 0.18)$ and interpret the energy
2493: difference shown in Fig.~(\ref{Ivanov}(b)) as the condensation energy.  Such a
2494: state may serve as the ``competing'' state that we have argued must live
2495: inside the vortex core.  The fact that the energy difference vanishes at $x =
2496: 0$ guarantees that it is small for small $x$.
2497: 
2498: %Ivanov (2003)
2499: \Ref{I0365} pointed out that the concave nature of the energy curves shown
2500: in Fig.~\ref{Ivanov}(a) for small $x$ indicate that the system is prone to
2501: phase separation.  Such a phase separation may be suppressed by long-range
2502: Coulomb interaction and the energy curves are indeed sensitive to
2503: nearest-neighbor repulsion.  Thus we believe that Fig.~\ref{Ivanov}(a) still
2504: provides a useful comparison of different trial wavefunctions.
2505: 
2506: 
2507: 
2508: 
2509: 
2510: 
2511: \subsection{Properties of projected wavefunctions} It is interesting to put
2512: aside the question of energetics and study the nature of the projected
2513: $d$-wave superconductor.  A thorough study by 
2514: %Paramekanti, Randeria and Trivedi (2001, 2003)
2515: \Ref{PRT0102}, \citeyear{PRT0404}
2516: showed that it correctly captures many of the properties
2517: of the cuprate superconductors.  For example, the superfluid density vanishes
2518: linearly in $x$ for small $x$.  This is to be expected since the projection
2519: operator is designed to yield an insulator at half-filling.  The momentum
2520: distribution exhibits a jump near the noninteracting Fermi surface. The size
2521: of the jump is interpreted as the quasiparticle weight $z$ according to Fermi
2522: liquid theory and again goes to zero smoothly as $x \rightarrow 0$.  Using the
2523: sum rule and assuming Ferm liquid behavior for the nodal quasiparticles, the
2524: Fermi velocity is estimated and found to be insensitive to doping, in
2525: agreement with photoemission experiments.
2526: 
2527: A distinctive feature of the projected staggered flux state is that it breaks
2528: translational symmetry and orbital currents circulate the plaquettes in a
2529: staggered fashion as soon as $x \neq 0$.  Motivated by the $SU(2)$ symmetry
2530: which predicts a close relationship between the projected $d$-wave
2531: superconductor and the projected staggered flux states, 
2532: %Ivanov, Lee and Wen (2000) 
2533: \Ref{ILW0053}
2534: examined whether there are signs of the orbital current in the
2535: projected $d$-wave superconductor.  Since this state does not break
2536: translation or time-reversal symmetry, there is no static current.  However,
2537: the current-current correlation
2538: \be
2539: G_{\v j} = \langle
2540: j(\alpha) j(\beta)
2541: \rangle
2542: \label{Eq.19}
2543: \en
2544: where $j(\alpha)$ is the current on the $\alpha$ bond, shows a power law-type
2545: decay and its magnitude is much larger than the naive expectation that it
2546: should scale as $x^2$.  Note that before projection the $d$-wave
2547: superconductor shows no hint of the staggered current correlation.  The
2548: correlation that emerges is entirely a consequence of the projection.  We
2549: believe the emergence of orbital current fluctuations provides strong support
2550: for the importance of $SU(2)$ symmetry near half filling.  Orbital current
2551: fluctuations of similar magnitude were found in the exact ground sate
2552: wavefunction of the $t$-$J$ model on a small lattice, two holes on 32 sites.
2553: %(Leung, 2001)  Lee and Sha (2003)
2554: (\cit{L0012}; \cit{LS0371}) showed that the orbital current correlation
2555: has the same power law decay as the hole-chirality correlation,
2556: \be
2557: G_{\chi_h} = \langle
2558: \chi_h(i) \chi_h(j) 
2559: \rangle \nonumber
2560: \en
2561:  where $\chi_h$ is defined on a plaquette $i$ as $n_h(4) \v{S}_1 \cdot
2562: (\v{S}_2 \times \v{S}_3)$ where 1 to 4 labels the sites around the plaquette
2563: and $n_h(\v i) = 1 - c_{\v i\sigma}^\dagger c_{\v i\sigma}$ is the hole density
2564: operator.  This is in agreement with the notion  that a hole moving around the
2565: plaque experiences a Berry's phase due to the non-colinearity of the spin
2566: quantization axis of the instantaneous spin configurations.  For $S = {1\over
2567: 2}$ the Berry's phase is given by ${1\over 2} \phi$ where $\phi$ is the solid
2568: angle subtended by the instantaneous spin orientations $\v{S}_1$, $\v{S}_2$
2569: and $\v{S}_3$. 
2570: %(Wen, Wilczek and Zee, 1989; E. Fradkin, 1991)
2571: (\cit{WWZcsp}; \cit{Fra91}) This solid angle
2572: is related to the spin chirality $\v{S}_1 \cdot (\v{S}_2 \times \v{S}_3)$.
2573: This phase drives the hole in a clockwise or anti-clockwise direction
2574: depending on its sign, just as a magnetic flux through the center of the
2575: plaquette would.  Thus the flux $\Phi_0$ of the staggered flux state has its
2576: physical origin in the coupling between the hole kinetic energy and the spin
2577: chirality.
2578: 
2579: 
2580: %Insert A
2581: It is important to emphasize that the projected $d$-wave state possesses long
2582: range superconducting pairing order, while at the same time exhibiting power
2583: law correlation in antiferromagnetic order and staggered orbital current.  On
2584: the other hand, projection of a staggered flux phase at finite doping will
2585: possess long range orbital current order, but short range pairing and
2586: antiferromagnetic order.  A useful analogy is to think of these projected
2587: states as a person with a variety of personalities.  He may be courteous and
2588: friendly at one time, and aggressive and even belligerent at another, depending
2589: on his environment.  Thus different versions of projected states shown in
2590: Fig.~\ref{Ivanov}(a) all have the same kinds of fluctuations; it is just
2591: that one kind of order may dominate over the others.  Then it is easy to
2592: imagine that the system may shift from one state to another in different
2593: environments.  For instance, in section XII.C we will argue that the pairing
2594: state will switch to a projected staggered flux state inside the vortex core.
2595: Note that this is a different picture from the traditional Landau picture of
2596: competing states as advocated by 
2597: %Chakravarty {\em et al.}, 2001
2598: \Ref{CLM0203}, for instance.
2599: These authors suggested on phenomenological grounds that the pseudogap region
2600: is characterized by staggered orbital current order, which they call
2601: $d$-density waves (DDW).  The symmetry of this order is indistinguishable from
2602: the doped staggered flux phase 
2603: %(Hsu {\em et al.}, 1991; Lee, 2002)
2604: \cite{HMA9166,L0249}.  According
2605: to Landau theory, the competition between DDW and superconducting order will
2606: result in either a first order transition or a region of co-existing phase at
2607: low temperatures.  This view of competing order is very different from the one
2608: proposed here, where a single quantum state possesses a variety of fluctuating
2609: orders.
2610: 
2611: \subsection{Improvement of projected wavefunctions, effect of $t^\prime$, and
2612: the Gutzwiller approximation} 
2613: 
2614: The projected wavefunction is the starting point
2615: for various schemes to further improve the trial wavefunction.  Indeed, the
2616: variational energy can be lowered and 
2617: %Sorella {\it et al}. (2002)
2618: \Ref{SMB0202}
2619:  provide
2620: strong evidence that  a $d$-wave superconducting state may be the ground state
2621: of the $t$-$J$ model.  On the other hand, other workers 
2622: %(Heeb and Rice, 1993; Shih {\it et al}., 1998)
2623: \cite{HR9373,SCL9894}
2624:  found that the superconducting tendency decreases
2625: with the improvement of the trial wavefunctions.  Studies based  on  other
2626: methods such as DMRG 
2627: %(White and Scalapino, 1999)
2628: \cite{WS9953} found that next-nearest
2629: neighbor hopping $t^\prime$ with $t^\prime/t > 0$ is needed to stabilize the
2630: $d$-wave superconductor.  Otherwise the holes are segregated into strip-like
2631: structures.  All these computational schemes suffer from some form of
2632: approximation and cannot give definitive answers.  What is clear is that the
2633: $d$-wave superconductor is a highly competitive candidate for the ground state
2634: of the $t$-$J$ model.
2635: 
2636: 
2637: Recently 
2638: %Shih {\em et al.} (2004)
2639: \Ref{SLE0402} have examined the pairing correlation in
2640: projected wavefunctions including the effect of $t^\prime$.  They find that
2641: for moderate doping ($x \gtrsim 0.1$) $t^\prime/t$ with a negative sign
2642: greatly enhances the pairing correlation.  The effect increases with
2643: increasing $t^\prime$ and is maximal around $t^\prime/t \approx -0.4$.  Their
2644: result contradicts expectations based on earlier DMRG work 
2645: %(White and Scalapino, 1999)
2646: \cite{WS9953} which found a suppression of superconductivity with negative
2647: $t^\prime/t$.  However, Shih {\em et al.} pointed out that the earlier work
2648: was limited to very low doping and is really not in disagreement with their
2649: finding for $x \gtrsim 0.1$.  This result should be confirmed by improving the
2650: wavefunction but the pair correlation with $t^\prime$ is so robust that the
2651: controversy surrounding the $t^\prime = 0$ case may well be avoided.  It
2652: should be noted that a negative $t^\prime$ is what band theory predicts.
2653: Furthermore, 
2654: %Pavarini {\em et al.} (2001)
2655: \Ref{PDS0103} have noted a correlation of $T_c$
2656: with $| t^\prime |$ and shown that the Hg and Tl compounds which have the
2657: highest $T_C$ have $t^\prime/t$ in the range $-0.3$ to $-0.4$.  Thus the role
2658: of $t^\prime$ may well explain the variation of $T_c$ among different families
2659: of cuprates.
2660: 
2661: The Gutzwiller projection is a rather cumbersome machinery to implement and a
2662: simple approximate scheme has been proposed, called the Gutzwiller
2663: approximation
2664: %(Zhang {\it et al}. 1988; Hsu 1990)
2665: (\cit{ZGR8836}; \cit{H9079}).  The essential step is to construct an effective Hamiltonian
2666: \be
2667: H_\text{eff} = -g_t t\sum_{\langle \v i\v j \rangle \sigma} c_{\v i\sigma}^\dagger c_{\v i\sigma} + g_J \sum_{\langle \v i\v j
2668: \rangle} \v{S}_{\v i} \cdot \v{S}_{\v j}
2669: \label{Eq.20}
2670: \en
2671: and treat this in the Hartree-Fock-BCS approximation. The projection operator
2672: in the original $t$-$J$ model is eliminated in favor of the reduction factors
2673: $g_t = 2x/(1 + x)$ and $g_J = 4/(1 + x)^2$, which are estimated by assuming
2674: statistical independence of the population of the sites
2675: %(Vollhardt, 1984)
2676: \cite{V8499}.
2677: The important point is that $g_t = 2x/(1 + x)$ reduces the kinetic energy to
2678: zero in the $x \rightarrow 0$ limit, in an attempt to capture the physics of
2679: the approach to the Mott insulator.  The Gutzwiller approximation bears a
2680: strong resemblance to the slave-boson mean-field theory and is just as easy to
2681: handle analytically.  It has the advantage that the energetics compare well
2682: with the Monte Carlo projection results.  The Gutzwiller approximation has
2683: been applied to more complicated problems such as impurity and vortex
2684: structure 
2685: %(Tsuchiura {\em et al.}, 2000, 2003)
2686: \cite{TTO0005,TOT0309} with good results.
2687: 
2688: %\section{The single hole problem}
2689: %
2690: %\section{Slave boson formulation of the $t$-$J$ model and mean-field theory}
2691: %
2692: %\section{$U(1)$ gauge fields}
2693: %
2694: %\section{$SU(2)$ formulation}
2695: %
2696: %\section{Spin-charge separation, confinement and fractionalization}
2697: %
2698: %\section{Summary and Future Outlook}
2699: %
2700: 
2701: 
2702: %\newcommand{\non}{\nonumber} \newcommand{\noi}{\noindent}
2703: %\newcommand{\e}{\varepsilon} \newcommand{\ii}{{i}}
2704: %\newcommand{\be}{\begin{eqnarray}}
2705: %\newcommand{\en}{\end{eqnarray}}
2706: %\newcommand{\dis}{\displaystyle}
2707: %\newcommand{\up}{\uparrow}
2708: %\newcommand{\down}{\downarrow}
2709: %\newcommand{\dia}{\noi$\diamond\,\,\,$}
2710: %\newcommand{\r}{\bbox{r}}
2711: %\newcommand{\k}{\bbox{k}}
2712: %\newcommand{\q}{\bbox{q}}
2713: %\newcommand{\bigQ}{\bbox{q}}
2714: %\newcommand{\bv}{\bbox{v}}
2715: %\newcommand{\bb}{\bbox}
2716: %\newcommand{\ga}{{\bbox{a}^3}}
2717: %\newcommand{\bbi}{\r_i}
2718: %\newcommand{\bbx}{\hat{\bf{e}}_x}
2719: %\newcommand{\bby}{\hat{\bf{e}}_y}
2720: 
2721: \section{The single hole problem}
2722: 
2723: The motion of a single hole doped into the antiferromagnet is a most
2724: fundamental issue to start with.  The $t$-$J$ type model is again the
2725: canonical Hamiltonian to study this problem.  The key physics of the problem
2726: is the competition between the antiferromagnetic (AF) correlation/long range
2727: ordering and the kinetic energy of the hole. The motion of the single hole
2728: distort the AF ordering when it hops between different sublattices.
2729: %Shraiman-Siggia
2730: \Ref{SS8867} studied this distortion in a semiclassical way,
2731: and found the new coupling between the spin current of the hole and the
2732: magnetization current of the background. This coupling leads to the long
2733: range dipolar distortion of the staggered magnetization and the minimum of the
2734: hole dispersion at $k=(\pi/2,\pi/2)$.  This position of the energy minimum is
2735: interpreted as follows.  Even if we start with the pure $t$-$J$ model, the
2736: direct hopping between nearest neighbor sites is suppressed, while the second
2737: order processes in $t$ leads to the effective hopping between the sites
2738: belonging to the same sublattice.  This effective $t'$ and $t''$ has the
2739: negative sign and hence lower the energy of $k=(\pi/2,\pi/2)$ compared with
2740: $k=(\pi,0),(0,\pi)$.
2741: 
2742: The dynamics of the single hole, \ie the spectral function of the Green's
2743: function is also studied by analytic method.  When the spin excitation is
2744: approximated by the magnon (spin wave), the Hamiltonian for the signle hole is
2745: given by \cite{KLR8980}
2746: \be
2747: H = { t \over N} \sum_{\v k ,q} M_{\v k,\v q} [ h^\dagger_{\v k} h_{\v k - \v q}
2748: \alpha_{\v q}
2749: + h.c.] + \sum_{q} \Omega_{\v q} \alpha_{\v q}^\dagger \alpha_{\v q}
2750: \label{Eq.21}
2751: \en
2752: where 
2753: \be
2754: \Omega_{\v q} =  2J \sqrt{ 1 - \gamma_{\v q}^2}
2755: \label{Eq.22}
2756: \en
2757: with $\gamma_{\v q} = ( \cos q_x + \cos q_y )/2$ and
2758: \be
2759: M(\v k,\v q) = 4(u_{\v q} \gamma_{\v k-\v q} + v_{\v q} \gamma_{\v k})
2760: \label{Eq.23}
2761: \en
2762: with $u_{\v k} = \sqrt{(1 + \nu_{\v k})/(2 \nu_{\v k})}$, $v_{\v k} = - \text{sign}(\gamma_{\v k})
2763: \sqrt{(1 - \nu_{\v k})/(2 \nu_{\v k})}$, and $\nu_{\v k} = \sqrt{1 -\gamma_{\v k}^2}$.  This
2764: Hamiltonian dictates that the magnon is emitted or absorbed every time the
2765: hole hops.  The most widely accepted method to study this model is the
2766: self-consistent Born approximation (SCBA) initiated by Kane, Lee and
2767: Read where the Feynman diagrams with the crossing magnon
2768: propagators are neglected. This leads to the self-consistent equation for the
2769: hole propagator:
2770: \be
2771: G(\v k,\omega) = [ \omega - \sum_{\v q} g(\v k,\v q)^2 G(\v k-\v q,\omega-\Omega_{\v q}) ]^{-1}.
2772: \label{Eq.24}
2773: \en
2774: The result is that there are two components of the spectral function $A(\v
2775: k,\omega) = - (1/\pi) \Im G^R(\v k,\omega)$: One is the coherent sharp peak
2776: corresponding to the quasi-particle and the other is the incoherent
2777: background. The former has the lowest energy at $k=(\pi/2,\pi/2)$ at the
2778: energy $\sim -t$ and disperses of the order of $J$, while the latter does not
2779: depend on the momentum $k$ so much and extends over the energy of the order of
2780: $t$.  Intuitively the hole has to wait for the spins to flip to hop, which
2781: takes a time of the order of $J^{-1}$.  Therefore the bandwidth is reduced
2782: from $\sim t$ to $\sim J$.  This mass enhancement leads to the reduced weight
2783: $z \sim J/t$ for the quasi-particle peak.  Later, more detailed studies have
2784: been done in SCBA\cite{LM9225}.  The conclusions obtained are the followings:
2785: (i) At $k=(\pi/2,\pi/2)$, there appear two additional two peaks at $E_{2,3}$
2786: in addition to the ground state delta-functional peak at $E_1$.
2787: 
2788: These energies are given for $J/t<0.4$ by
2789: \be
2790: E_n/t = - b + a_n (J/t)^{2/3}
2791: \label{Eq.25}
2792: \en
2793: where $a_1=2.16,a_2=5.46,a_3=7.81$, and $b=3.28$.  (ii) The spectral weight
2794: $z$ at $k=(\pi/2,\pi/2)$ scales as $z = 0.65(J/t)^{2/3}$.
2795: 
2796: These can be understood as the "string" excitation of the hole moving in the
2797: linear confining potential due to the AF background.  It has also been
2798: interpreted in terms of the confining interaction between spinon and holon
2799: %(Laughlin, 1997)
2800: \cite{L9726}.  Exact diagonalization studies have reached consistent results
2801: with SCBA.  Experimentally angle-resolved-photoemission spectroscopy
2802: (ARPES)(\cit{WSM9564}; \cit{Ro9867}) in undoped cuprates has revealed the spectral function of
2803: the single doped hole.  The energy dispersion of the hole looks like that of
2804: the $\pi$-flux state shifted by the Mott gap to the low energy 
2805: %(Laughlin, 1997)
2806: \cite{L9726}.  However, in real materials the second ($t'$) and third ($t"$)
2807: nearest neighbor hoppings are important.  The calculated energy dispersion is
2808: found to be sensitive to $t^\prime$ and $t^{\prime\prime}$.  For $t^\prime =
2809: t^{\prime\prime} = 0$, the dispersion is flat between $(\pi/2, \pi/2)$ and
2810: $(0,\pi)$ and does not agree with the data.  It turns out that the data is
2811: well fitted by $J/t=0.3$, $t'/t=-0.3$, $t"/t=0.2$.  On the other hand, ARPES
2812: in slightly electron-doped Ne$_{2-x}$Ce$_x$CuO$_2$ showed that the electron is
2813: doped into the point $k=(\pi,0)$ and $(0,\pi)$ \cite{Ao0103}. This difference
2814: will be discussed below.
2815: 
2816: 
2817: The variational wavefunction approach to the antiferromagnet and single hole
2818: problem has been pursued by several authors \cite{LHN0301,LLH0301}.  A good
2819: ground state variational wavefunction (vwf) at half-filling is 
2820: \be
2821: | \Psi_0\> = P_G
2822: \biggl[ \sum_{\v k} ( A_{\v k} a_{\v k \uparrow}^\dagger a_{-\v k \downarrow}^\dagger
2823: + B_{\v k} b_{\v k \uparrow}^\dagger b_{-\v k \downarrow}^\dagger \biggr]^{N/2}
2824: | 0\>
2825: \label{Eq.26}
2826: \en
2827: with $N$ being the number of atoms.  The operators $a_{\v k \sigma}^\dagger$,
2828: $b_{\v k \sigma}^\dagger$ are those for the upper and lower bands split by SDW
2829: with the energy $\pm \xi_{\v k}$, respectively, and $A_{\v k} = (E_{\v k} + \xi_{\v k})/\Delta_{\v k}$,
2830: $B_{\v k} =  (-E_{\v k} + \xi_{\v k})/\Delta_{\v k}$ with $E_{\v k} = \sqrt{ \xi_{\v k}^2+ \Delta_{\v k}^2}$ and
2831: $\Delta_{\v k} = (3/8)J \Delta( \cos k_x - \cos k_y)$.  The picture here is that in
2832: addition to the SDW, the RVB singlet formation represented by $\Delta$ is
2833: taken into account.  As mentioned in the last section, this vwf gives much
2834: better energy compared to  that with $\Delta=0$. Hence the ground state is far
2835: from the classical N\"{e}el state and includes strong quantum fluctuations.
2836: Next the vwf in the case of single doped hole with momentum $q$ and $S^z =
2837: 1/2$ is
2838: \be
2839: | \Psi_{\v q}\> = P_G c^\dagger_{q \uparrow}
2840: \biggl[ \sum_{\v k (\ne q)} ( A_{\v k} a_{\v k \uparrow}^\dagger 
2841: a_{-\v k \downarrow}^\dagger
2842: + B_{\v k} b_{\v k \uparrow}^\dagger b_{-\v k \downarrow}^\dagger \biggr]^{N/2-1}
2843: | 0\>
2844: \label{Eq.27}
2845: \en
2846: 
2847: This vwf does not contain the information of $t'$,$t''$ except the very small
2848: dependence of $A_{\v k}$, and  $B_{\v k}$.  The robustness of this vwf is the
2849: consequence of the large quantum fluctuation already present in the
2850: half-filled case, so that the hole motion is possible even without disturbing
2851: the spin liquid-like state.  Although the vwf does not depends on the
2852: parameters $t'$,$t''$, the energy dispersion $E(\v k)$ is given by the
2853: expectation value as
2854: \begin{equation}
2855: E(\v k) = \< \Psi_{\v k} | H_{t-J} + H_{t'-t''}| \Psi_{\v k} \>,
2856: \label{Eq.28}
2857: \end{equation}
2858: and depends on these parameters.  This expression gives a reasonable agreement
2859: with the experiments both in undoped material \cite{Ro9867} and electron-doped
2860: material \cite{Ao0103}.  Here an important question is the relation
2861: between the hole- and electron-doped cases.  There is a particle-hole symmetry
2862: operation which relates the $t$-$t'$-$t''$-$J$ model for a hole to that for an
2863: electron.  The conclusion is that the sign change of $t'$, and $t''$ together
2864: with the shift in the momentum by $(\pi,\pi)$ gives the mapping between the
2865: two cases.  Using this transformation, one can explain the difference between
2866: hole- and electron-doped cases in terms of the common vwf eq.~(\ref{Eq.27}).
2867: The former has the minimum at $k=(\pi/2,\pi/2)$ while the latter at
2868: $k=(\pi,0),(0,\pi)$.
2869: 
2870: Exact diagonalization study \cite{T009} has shown that the electronic state is
2871: very different between $k = (\pi/2,\pi/2)$ and $k =(\pi,0)$ for the
2872: appropriate values of $t'$ and $t''$ for hole doped case.
2873: The spectral weight becomes very small at $(\pi,0)$ and the hole is surrounded
2874: by anti-parallel spins sitting on the same sublattice.  Both these features
2875: are captured by a trial wavefunction which differs from eq.~(\ref{Eq.27}) in
2876: that the momentum $\v{q}$ of the broken pair is different from the momentum
2877: of the inserted electron.  This can also be interpreted as the decay of the
2878: quasiparticle state via the emission of a spin wave \cite{LLH0301}.  There are
2879: thus two types of wf's with qualitatively different nature, {\em i.e.}, one
2880: describes the quasi-particle state and another  which is highly incoherent and
2881: may be realized as a spin-charge separated state.
2882: 
2883: One important discrepancy between experiment and theory is the line-shape of
2884: the spectral function. Namely the experiments show broad peak with the width
2885: of the order of $\sim 0.3$eV in contrast to the delta-functional peak expected
2886: for the ground state at $k=(\pi/2,\pi/2)$. One may attribute this large width
2887: to the disorder effect in the sample. However the ARPES in the overdoped region
2888: shows even sharper peak at the Fermi energy even though the doping introduces
2889: further disorder.  Therefore the disorder effect is unlikely to explain this
2890: discrepancy.  Recently the electron-phonon coupling to the single hole in
2891: $t$-$J$ model has been studied using quantum Monte Carlo simulation
2892: %(Mishchenko {\em et al.}, 2004)
2893: \cite{MN0402}.  It is found that the small polaron 
2894: formation
2895: in the presence of strong
2896: correlation 
2897: reduces the dispersion and the weight of the zero-phonon line, while the center of
2898: mass of the spectral weight for the originally "quasi-particle" peak remain
2899: the same as the pure $t$-$J$ model, even though the shape is broadened.
2900: Therefore the polaron effect is a promising scenario to explain the spectral
2901: shape.
2902: 
2903: Recently, 
2904: %Shen {\em et al.} (2004)
2905: \Ref{SRL0402} pointed out that the polaron picture also
2906: explains a long standing puzzle regarding the location of the chemical
2907: potential with doping.  Naive expectation based on doping a Hubbard model
2908: predicts that the chemical potential should lie at the top of the valence
2909: band, whereas experimentally in Na-doped Ca$_2$CuO$_2$Cl$_2$ it was found that
2910: the chemical potential appears to lie somewhere in mid-gap, \ie with a small
2911: but finite density of holes, the chemical potential is several tenths of eV
2912: higher than the energy of the peak position of the one-hole spectrum.  This is
2913: naturally explained if the one-hole spectrum has been shifted down by polaron
2914: effects, so that the top of the valence band should be at the zero-phonon
2915: line, rather than the center of mass of the one-hole spectrum.
2916: 
2917: 
2918: \section{Slave boson formulation of $t$-$J$ model and mean field theory}
2919: 
2920: As has been discussed in II, it is widely believed that the low energy physics
2921: of high-Tc cuprates is described in terms of $t$-$J$ type model, which is
2922: given by \cite{LN9221}
2923: \be
2924: H= \sum_{\<\v i\v j\>} J\left(
2925: {\v{S}}_{\v i}\cdot {\v{S}}_{\v j}-{1\over
2926: 4}n_{\v i} n_{\v j} \right)
2927: -\sum_{\v i\v j}t_{\v i\v j}
2928: \left(c_{\v i\sigma}^\dagger
2929: c_{\v j\sigma}+{\rm H.c.}\right).
2930: \label{t-J}
2931: \en
2932: where $t_{\v i\v j}=t$, $t'$, $t''$ for the nearest, second nearest and 3rd nearest
2933: neighbor pairs, respectively. The effect of the strong Coulomb repulsion is
2934: represented by the fact that the electron operators $c^\dagger_{\v i\sigma}$ and
2935: $c_{\v i\sigma}$ are the projected ones, where the double occupation is
2936: forbidden.  This is written as the inequality
2937: \be
2938: \sum_{\sigma} c^\dagger_{\v i\sigma} c_{\v i \sigma} \le 1,
2939: \label{Eq.30}
2940: \en
2941: which is very difficult to handle.  A powerful method to treat this constraint
2942: is so called the slave-boson method 
2943: %(Barnes, 1976; Coleman, 1984)
2944: \cite{B7675,C8435}.  In most
2945: general form, the electron operator is represented as
2946: \be
2947: c^\dagger_{\v i\sigma} = f_{\v i\sigma}^\dagger b_{\v i} + 
2948: \epsilon_{\sigma \sigma'} f_{\v i \sigma'} d_{\v i}^\dagger
2949: \label{Eq.31}
2950: \en
2951: where $\epsilon_{\uparrow \downarrow} = 
2952: - \epsilon_{\downarrow \uparrow} = 1$ is the antisymmetric tensor.
2953:   $f_{\v i\sigma}^\dagger$, $f_{\v i \sigma}$ are the fermion operators, while
2954: $b_{\v i}$, $d^\dagger_{\v i}$ are the slave-boson operators.  This representation
2955: together with the constraint 
2956: \be
2957: f_{\v i\uparrow}^\dagger f_{\v i\uparrow}
2958: + f_{\v i\downarrow}^\dagger f_{\v i\downarrow} + b^\dagger_{\v i} b_{\v i}
2959: + d^\dagger_{\v i} d_{\v i} = 1
2960: \label{Eq.32}
2961: \en
2962: reproduces all the algebra of the electron (fermion) operators.  From eqs.
2963: (\ref{Eq.31}) and (\ref{Eq.32}), the physical meaning of these operators is
2964: clear. Namely, there are 4 states per site and $b^\dagger$, $b$ corresponds to
2965: the vacant state, $d^\dagger$,$d$ to double occupancy, and $f^\dagger_\sigma$,
2966: $f_{\sigma}$ to the single electron with spin $\sigma$.  With this formalism
2967: it is easy to exclude the double occupancy just by deleting $d^\dagger$, $d$
2968: from the above equations (\ref{Eq.31}) and (\ref{Eq.32}).  Then the projected
2969: electron operator is written as
2970: \be
2971: c_{\v i\sigma}^\dagger = f_{\v i\sigma}^\dagger b_{\v i}
2972: \label{Eq.33}
2973: \en
2974: with the condition
2975: \be
2976: f_{\v i\uparrow}^\dagger f_{\v i\uparrow}
2977: + f_{\v i\downarrow}^\dagger f_{\v i\downarrow} + b^\dagger_{\v i} b_{\v i} = 1.
2978: \label{Eq.34}
2979: \en
2980: This constraint can be enforced with a Lagrangian multiplier $\lambda_{\v i}$.
2981: Note that unlike eq. (\ref{Eq.31}), eq. (\ref{Eq.33}) is not an operator
2982: identity and the R.H.S. does not satisfy the fermion commutation relation.
2983: Rather, the requirement is that both sides have the correct matrix elements in
2984: the reduced Hilbert space with no doubly occupied states.  For example, the
2985: Heisenberg exchange term is written in terms of $f^\dagger_{\v i\sigma}$,
2986: $f_{\v i\sigma}$ only \cite{BZA8773}
2987: \be
2988: {\v{S}}_{\v i}\cdot {\v{S}}_{\v j} &=& -{1\over 4} f_{\v i\sigma}^\dagger f_{\v j\sigma}
2989: f_{\v j\beta}^\dagger f_{\v i\beta}  \nonumber  \label{Eq.35} \\
2990: &-& {1\over 4} \left(
2991: f_{\v i\up}^\dagger f_{\v j\down}^\dagger - f_{\v i\down}^\dagger f_{\v j\up}^\dagger
2992: \right) \left(
2993: f_{\v j\down} f_{\v i\up} - f_{\v j\up} f_{\v i\down}
2994: \right)  \nonumber \\ 
2995: &+& {1\over 4} \left( f_{\v i\alpha}^\dagger f_{\v i\alpha}  \right) .
2996: \en
2997: We write
2998: \be
2999: n_{\v i}n_{\v j} = (1 - b_{\v i}^\dagger b_{\v i} ) (1 - b_{\v j}^\dagger b_{\v j} )  .
3000: \label{Eq.36}
3001: \en
3002: Then ${\v{S}}_{\v i}\cdot {\v{S}}_{\v j} - {1\over 4}n_{\v i}n_{\v j}$ can be written in terms of
3003: the first two terms of eq. (\ref{Eq.35})  plus quadratic terms, provided we
3004: ignore the nearest-neighbor hole-hole interaction ${1\over 4}b_{\v i}^\dagger
3005: b_{\v i}b_{\v j}^\dagger b_{\v j} $.  We then decouple the exchange term in both the
3006: particle-hole and particle-particle channels via the Hubbard-Stratonovich (HS)
3007: transformation.  
3008: %For this purpose the $SU(2)$ doublets defined below are useful
3009: %\cite{AZH8845,DFM8826}
3010: 
3011: 
3012: Then the partition function is written in the 
3013: %compact 
3014: form
3015: \be
3016: Z = \int{D f D f^\dagger Db D\lambda D\chi D\Delta} \exp \left(
3017: -\int^\beta _0 d\tau L_1
3018: \right)
3019: \label{Eq.38}
3020: \en
3021: where
3022: \be
3023: \label{Eq.39}
3024: %L_1 &=& {\tilde{J}\over 2} \sum_{\<\v i\v j\>}\Tr [U_{\v i\v j}^\dagger U_{\v i\v j}] +
3025: %{\tilde{J}\over 2} \sum_{\<\v i\v j\>,\sigma} \left (
3026: %\Phi_{\v i\sigma}^\dagger U_{\v i\v j} \Phi_{\v j\sigma} + c.c.
3027: %\right) \nonumber \\
3028: L_1 &=& \tilde{J} \sum_{\<\v i\v j\>} \left(
3029: |\chi_{\v i\v j}|^2 + | \Delta_{\v i\v j} |^2 \right)
3030: +\sum_{\v i\sigma}f_{\v i\sigma}^\dagger (\partial_\tau - i\lambda_{\v i})
3031: f_{\v i\sigma} 
3032: \nonumber\\
3033: &-& \tilde{J} 
3034: \left[
3035:  \sum_{\<\v i\v j \>} \chi_{\v i\v j}^\ast \left( \sum_{\sigma} f_{\v i\sigma}^\dagger
3036: f_{\v j\sigma} \right) + c.c. \right]  
3037: \\
3038: &+& \tilde{J} \left[ \sum_{\<\v i\v j\>} \Delta_{\v i\v j} \left(
3039: f^\dagger_{\v i\up}f^\dagger_{\v j\down} - f^\dagger_{\v i\down} f^\dagger_{\v j\up} \right) + c.c. \right] \nonumber \\
3040: &+&\sum_{\v i} b_{\v i}^\ast(\partial_\tau - i\lambda_{\v i} + \mu_B ) b_{\v
3041: i} 
3042: - \sum_{\v i\v j}{t}_{\v i\v j}b_{\v i}b_{\v j}^\ast f_{\v i\sigma}^\dagger f_{\v j\sigma} ,
3043: \nonumber 
3044: \en
3045: with $\chi_{\v i\v j}$ representing fermion hopping and $\Delta_{\v i\v j}$
3046: representing fermion pairing corresponding to the two ways of representing the
3047: exchange interaction in terms of the fermion operators. From eqs.
3048: (\ref{Eq.35}) and (\ref{Eq.39})  it is concluded that $\tilde{J}=J/4$, but in
3049: practice the choice of $\tilde{J}_{\v i\v j}$ is not so trivial, namely one
3050: would like to study the saddle point approximation (SPA) and the Gaussian
3051: fluctuation around it, and requires SPA to reproduce the mean field theory.
3052: %It is usually the case 
3053: The latter requirement is satisfied when only one HS variable is relevant, but
3054: not for the multicomponent HS variables 
3055: %(Orland and Negele; Ubbens and Lee, 1992)
3056: \cite{NegO87,UL9234}.  In the latter case, it is better to chose the
3057: parameters in the Lagrangian to reproduce the mean field theory. In the
3058: present case, $\tilde{J} = 3J/8$ reproduces the mean field self-consistent
3059: equation which is obtained by the Feynman variational principle \cite{BL0102}.
3060: 
3061: We note that $L_1$ in \Eq{Eq.39} is invariant under a local $U(1)$
3062: transformation
3063: \begin{align}
3064: \label{U1gaugetrans}
3065:  f_{\v i} &\to e^{i\th_{\v i}} f_{\v i} \nonumber\\
3066:  b_{\v i} &\to e^{i\th_{\v i}} b_{\v i} \nonumber\\
3067:  \chi_{\v i\v j} &\to e^{-i\th_{\v i}} \chi_{\v i\v j} e^{i\th_{\v j}} \nonumber\\
3068:  \Del_{\v i\v j} &\to e^{i\th_{\v i}} \Del_{\v i\v j} e^{i\th_{\v j}} \nonumber\\
3069:  \la_{\v i} &\to  \la_{\v i} +\prt_\tau \th_{\v i}
3070: \end{align}
3071: which is called $U(1)$ gauge transformation.
3072: Due to such a $U(1)$ gauge invariance, the phase fluctuations of $\chi_{\v i\v j}$ and $\la_{\v i}$
3073: have a dynamics of $U(1)$ gauge field (see section IX).
3074: 
3075: Now we describe the various mean field theory corresponding to the saddle
3076: point solution to the functional integral.  The mean field conditions are
3077: \begin{align}
3078: \label{Eq.39a}
3079: \chi_{\v i\v j} &= \sum_\sigma \langle
3080: f^\dagger_{\v i\sigma} f_{\v j\sigma} \rangle \\
3081: \label{Eq.40} 
3082: \Delta_{\v i\v j} &= \langle  f_{\v i\up}f_{\v j\down} - f_{\v i\down}f_{\v i\up}  \rangle
3083: \end{align}
3084: 
3085: Let us first consider the $t$-$J$ model in the undoped case,
3086: \ie the half-filled case.  There are no bosons in this case, and the theory
3087: is purely that of fermions.  The original one, \ie uniform RVB state,
3088: proposed by Baskaran-Zou-Anderson \cite{BZA8773} is given by
3089: \be
3090: \chi_{\v i\v j} = \chi = \text{real}
3091: \label{Eq.41}
3092: \en
3093: for all the bond and $\Delta_{\v i\v j}=0$.  The fermion spectrum is that of the
3094: tight binding model
3095: \be
3096: H_\text{uRVB} = - \sum_{\v k \sigma} 2 {\tilde J}
3097: \chi ( \cos k_x + \cos k_y )
3098: f^\dagger_{\v k \sigma} f_{\v k \sigma}, 
3099: \label{Eq.42}
3100: \en
3101: with the saddle point value to the Lagrange multiplier $\lambda_{\v i}=0$.  The so
3102: called ``spinon Fermi surface'' is large, \ie it is given by the condition
3103: $k_x \pm k_y = \pm \pi$ with a diverging density of states (van Hove
3104: singularity) at the Fermi energy. 
3105: %Affleck-Marston 
3106: Soon after, many authors 
3107: %studied 
3108: found lower energy states than the uniform RVB state. One can easily
3109: understand that lower energy states 
3110: %are there 
3111: exist because the Fermi surface is perfectly nested with the nesting
3112: wavevector ${\vec Q} = (\pi,\pi)$ and the various instabilities with $\vec Q$
3113: are expected. Of particular importance are the $d$-wave state [see (Eq.
3114: \ref{Eq.13})] and the staggered flux state [see (Eq. \ref{Eq.17})] which give
3115: identical energy dispersion.  This was explained as being due to a local
3116: $SU(2)$ symmetry when the spin problem is formulated in terms of fermions
3117: \cite{AZH8845,DFM8826}.  We write
3118: \be
3119: \Phi_{\v i\up} =
3120: \bpm f_{\v i\up} \\
3121: f_{\v i\down}^\dagger \epm \,\, , \,\,\,\,
3122: \Phi_{\v i\down} =
3123: \bpm f_{\v i\down} \\
3124: -f_{\v i\up}^\dagger \epm \,\, ,
3125: \label{Eq.37}
3126: \en
3127: Then eq.~(\ref{Eq.39}) can be written in the more compact form
3128: \be
3129: L_1 &=& {\tilde{J}\over 2} \sum_{\<\v i\v j\>}\Tr [U_{\v i\v j}^\dagger U_{\v i\v j}] +
3130: {\tilde{J}\over 2} \sum_{\<\v i\v j\>,\sigma} \left (
3131: \Phi_{\v i\sigma}^\dagger U_{\v i\v j} \Phi_{\v j\sigma} + c.c.
3132: \right) \nonumber \\
3133: &+&\sum_{\v i\sigma}f_{\v i\sigma}^\dagger (\partial_\tau - i\lambda_{\v i})
3134: f_{\v i\sigma} \nonumber \\
3135: &+&\sum_{\v i} b_{\v i}^\ast(\partial_\tau - i\lambda_{\v i} + \mu_B ) b_{\v i} \nonumber \\
3136: &-& \sum_{\v i\v j}{t}_{\v i\v j}b_{\v i}b_{\v j}^\ast f_{\v i\sigma}^\dagger f_{\v j\sigma} ,
3137: \label{Eq.44}
3138: \en
3139: where
3140: \be
3141: U_{\v i\v j}=
3142: \bpm -\chi_{\v i\v j}^\ast&\Delta_{\v i\v j}\\
3143: \Delta_{\v i\v j}^\ast&\chi_{\v i\v j} \epm.
3144: \label{Eq.45} 
3145: \en
3146: At half filling $b = \mu_B =0$ and the mean field solution corresponds to
3147: $\lambda_{\v i} = 0$.  The Lagrangian is invariant under
3148: \be
3149: \Phi_{\v i\si} &\rightarrow&  W_{\v i}\Phi_{\v i\si} \\
3150: U_{\v i\v j} &\rightarrow& W_{\v i} U_{\v i\v j} W_{\v j}^\dagger
3151: \en
3152: where $W_{\v i}$ is an $SU(2)$ matrix [see eq.~(\ref{Eq.18})].  We reserve a fuller discussion of the $SU(2)$ gauge symmetry to Section X, but here we just give a simple example.
3153: %\be
3154: %U_{\v i\v j}=
3155: %\bpm -\chi_{\v i\v j}^\ast&\Delta_{\v i\v j}\\
3156: %\Delta_{\v i\v j}^\ast&\chi_{\v i\v j} \epm.
3157: %\label{Eq.40}
3158: %\en
3159: %Of particular important is the $\pi$-flux state which is given by 
3160: %\be 
3161: %\chi_{\v i i+x} = \chi
3162: %\nonumber \\
3163: %\chi_{\v i i+y} = \chi (-1)^{\v i_x + i_y}
3164: %\label{Eq.43}
3165: %\en
3166: %with $\Delta_{\v i\v j} = 0$.
3167: %When one pick up one plaquette and consider the 
3168: %product of the hopping integral around its perimeter,
3169: %\be 
3170: %\chi_{ i i+x} \chi_{\v i +x i+x+y} \chi_{\v i+x+y i+y} \chi_{\v i+y i} 
3171: %= - \chi^4
3172: %\label{Eq.44}
3173: %\en
3174: %This product is invariant with respect to the local gauge
3175: %transformation 
3176: %\be
3177: %f_{\v i \sigma} \to e^{\v i \varphi_{\v i}} f_{\v i \sigma}
3178: %\label{Eq.45}
3179: %\en
3180: %and plays the essential role in the gauge theory described later.
3181: %Writing the product as $\chi^4 e^{ i \Phi}$ the "flux" $\Phi$
3182: %penetrating the plaquette is $\pi$ in the present case, and hence
3183: %this state is called "$\pi$-flux state". 
3184: %The fermion spectrum is given by 
3185: %\be
3186: %& & H_{\pi-flux}
3187: %\nonumber \\
3188: %&=& \sum_{\v k:half BZ, \sigma} {\tilde J} \chi
3189: %\sqrt{ \cos^2 k_x + \cos^2 k_y}
3190: %(f^\dagger_{1 k \sigma} f_{1 k \sigma} - 
3191: %f^\dagger_{2 k \sigma} f_{2 k \sigma} )
3192: %\label{Eq.46}
3193: %\en
3194: %where the summation over $k$ is done with the reduced half-Brillouin 
3195: %zone shown in Fig. . 
3196: %The fermion operator $f_{1 k \sigma}$, $f_{2 k \sigma}$
3197: %are given by a linear combination of $f_{\v k \sigma}$ and $f_{\v k+ Q \sigma}$
3198: %unitary transformation 
3199: %Up to now, we consider the case without the pairing $\Delta_{\v i\v j}$.
3200: %The "superconducting" state of fermions are pursued first by 
3201: %Kotliar-Liu \cite{KL8842}, who found the d-wave pairing s\v tate is the most 
3202: %stable. This d-wave RVB state is given by 
3203: %\be 
3204: %\chi_{\v i\v j} = \chi
3205: %\nonumber \\
3206: %\Delta_{\v i i+x} = - \Delta_{\v i i+y} = \Delta.
3207: %\label{Eq.47}
3208: %\en
3209: %The self-consistent equation results in $\chi=\Delta$ 
3210: %at half-filing. The mean field energy is the same as 
3211: %the $\pi$-flux state, \ie $E_G = -  {\tilde J}$.
3212: %The fermion spectrum is the same as that of the 
3213: %BCS theory:
3214: %\be
3215: %H_{d-RVB} = \sum_{\v k} {\tilde J} \chi
3216: %\sqrt{ \cos^2 k_x + \cos^2 k_y}
3217: %(f^\dagger_{+ k } f_{+ k } - 
3218: %f^\dagger_{- k } f_{- k } )
3219: %\label{Eq.48}
3220: %\en
3221: %Again one case see the similarity between the 
3222: %fermion spectrum of the $\pi$-flux and
3223: %d-RVB states, although the former breaks the 
3224: %periodicity of the lattice while the latter 
3225: %mixes the particle and hole. 
3226: %Here some caution is needed. 
3227: %According to the gauge transformation eq. ( ),
3228: %the momentum of the fermion can be changed. Therefore
3229: %it is possible that the solution with periodicity doubling 
3230: %is equivalent to that without it via a gauge transformation.
3231: %However one need one step further to relate the $\pi$-flux state and 
3232: %d-RVB state by generalizing the gauge transformation to mix
3233: %particle and hole.
3234: %Affleck {\it et al.}\cite{AZH8845} 
3235: %found exactly this $SU(2)$ local gauge symmetry
3236: %for the $t$-$J$ model at half-filling, and showed the
3237: %equivalence between $\pi$-flux state and d-RVB state.
3238: %The $SU(2)$ doublet in Eq. (5) expresses the
3239: %physical idea that a physical up-spin can be represented by
3240: %the presence of an up-spin fermion, or the absence
3241: %of a down-spin fermion, once the constraint is imposed.
3242: %This symmetry has far more physical consequences even in the
3243: %doped case as will be discussed in section  , but here we restrict to 
3244: %the mean field solution at half-filling.
3245: In terms of the link variable $U_{\v i\v j}$, the $\pi$-flux and $d$-RVB states are
3246: represented as
3247: \be  
3248: U_{\v i\v j}^{\pi\text{-flux}} = -\chi ( \tau^3 - i(-1) ^{i_x+j_y} ),
3249: \label{Eq.49}
3250: \en
3251: and
3252: \be
3253: U_{\v i,i+\mu}^d = -\chi ( \tau^3 + \eta_\mu \tau^1 ),
3254: \label{Eq.50}
3255: \en
3256: respectively.  These two are related by
3257: \be
3258: U_{\v i\v j}^{SF} = W_{\v i}^\dagger U_{\v i\v j}^d W_{\v j}
3259: \label{Eq.51}
3260: \en
3261: where
3262: \be
3263: W_{\v j} = \exp \left[
3264: i(-1)^{j_x+j_y} {\pi\over 4} \tau^1 \right].
3265: \label{Eq.52}
3266: \en
3267: Therefore the $SU(2)$ transformation of the fermion variable
3268: \be
3269: \Phi'_{\v i} = W_{\v i} \Phi_{\v i}
3270: \label{Eq.53}
3271: \en
3272: relates the $\pi$-flux and d-RVB states.  Here some remarks are in order.
3273: First it should be noted that we are discussing the Mott insulating state and
3274: its spin dynamics. The charge transport is completely suppressed by the
3275: constraint eq. (\ref{Eq.34}). This will be discussed in sec. X where the mean
3276: field theory is elaborated into gauge theory.  Secondly, it is now established
3277: that the ground state of the two-dimensional antiferromagnetic Heisenberg
3278: model shows the antiferromagnetic long range ordering (AFLRO).  This
3279: corresponds to the third (and most naive) way of decoupling the exchange
3280: interaction, \ie
3281: \be
3282: {\v{S}}_{\v i}\cdot {\v{S}}_{\v j} = 
3283: {1\over 4} f_{\v i\alpha}^\dagger \sigma^{\mu}_{\alpha \beta} f_{\v i \beta}
3284: f_{\v j \gamma}^\dagger \sigma^{\mu}_{\gamma \delta} f_{\v j\delta}
3285: \label{Eq.54}
3286: \en
3287: However even with the AFLRO, the singlet formation represented by $\chi_{\v i\v j}$
3288: and $\Delta_{\v i\v j}$ dominates and AFLRO occurs on top of it. This view has been
3289: stressed by Hsu \cite{H9079,HMA9166} generalizing the $\pi$-flux state to
3290: include the AFRLO, and is in accord with the energetics of the projected
3291: wavefunctions, as discussed in section VI.A.
3292: 
3293: 
3294: \begin{figure}[t]
3295: \centerline{
3296: \includegraphics[width=2.5in]{fig21.eps}
3297: }
3298: \vspace{0.5cm}
3299: \caption{ 
3300: Schematic phase diagram of the $U(1)$ mean field theory.  The solid line
3301: denotes the onset of the uniform RVB state $(\chi \neq 0)$.  The dashed line
3302: denotes the onset of fermion pairing $(\Delta \neq 0)$ and the dotted line
3303: denotes mean field Bose condensation $(b \neq 0)$.  The four regions are
3304: (I)~Fermi liquid $\chi \neq 0$, $b \neq 0$; (II)~spin gap $\chi \neq 0$,
3305: $\Delta \neq 0$; (III)~$d$-wave superconductor $\chi \neq 0$, $\Delta \neq 0$,
3306: $b \neq 0$; and (IV)~ strange metal $\chi \neq 0$ \cite{LN9221}.
3307: }
3308: \label{U1}
3309: \end{figure}
3310: 
3311: 
3312: Now we turn to the doped case, \ie $x \ne 0$.  Then the behavior of the
3313: bosons are crucial for the charge dynamics. At the mean field theory, the
3314: bosons are free and condensed at $T_{BE}$.  In three-dimensional system,
3315: $T_{BE}$ is finite while $T_{BE} = 0$ for purely two-dimensional system.
3316: Theories assume weak three dimensional hopping between layers, and obtain the
3317: finite $T_{BE}$ roughly proportional to the boson density $x$
3318: \cite{KL8842,SHF8868}.  This materializes the original idea by Anderson
3319: \cite{A8796} that the preformed spin superconductivity (RVB) turns into the
3320: real superconductivity via the Bose condensation of holons.  
3321: %Kotliar and Liu, 1988 
3322: \Ref{KL8842}
3323: and 
3324: %Suzumura {\em et al.}, 1988
3325: \Ref{SHF8868} found the d-wave superconductivity in the
3326: slave-boson mean field theory presented above, and the schematic phase diagram
3327: is given in Fig.~\ref{U1}.  There are 5 phases classified by the order
3328: parameters $\chi$, $\Delta$, and $b=\<b_{\v i}\>$ for the Bose condensation.  In
3329: the incoherent state at high temperature, all the order parameters are zero.
3330: In the uniform RVB state (IV in Fig.~\ref{U1}), only $\chi$ is finite. In the
3331: spin gap state (II), $\Delta$ and $\chi$ are nonzero while $b=0$. This
3332: corresponds to the spin singlet ``superconductivity'' with the incoherent
3333: charge motion, and can be viewed as the precursor phase of the
3334: superconductivity. 
3335: %However as mentioned in sec. and also as will be discussed in sec. , this
3336: %interpretation is too naive. 
3337: This state has been interpreted as the pseudogap phase \cite{F9287}.
3338: We note that at the mean field level, the $SU(2)$ symmetry is broken by the
3339: nonzero $\mu_B$ in eq.~(\ref{Eq.44}) and the $d$-wave pairing state is chosen
3340: because it has lower energy than the staggered flux state.  We shall return to
3341: this point in Section X.  In the Fermi liquid state (I), both $\chi$ and $b$
3342: are nonzero while $\Delta=0$. This state is similar to the slave-boson
3343: description of heavy fermion state.  Lastly when all the order parameter is
3344: nonzero, we obtain the $d$-wave superconducting state (III). This mean field
3345: theory, in spite of its simplicity, captures rather well the experimental
3346: features as described in sections III and IV.
3347: 
3348: 
3349: Before closing this section, we mention the slave fermion method and its mean
3350: field theory (\cit{AA8816}; \cit{Y8916}; \cit{CRK9019}).  One can exchange the
3351: statistics of fermion and boson in eqs.  (\ref{Eq.31}) and (\ref{Eq.33}). Then
3352: the bosons has the spin index, \ie $b_{\v i \sigma}$ while the fermion
3353: becomes spinless, \ie $f_{\v i}$. This boson is called Schwinger boson, and
3354: is suitable to describe the AFLRO state. The large $N$-limit of Schwinger
3355: boson theory gives the AFLRO state for $S=1/2$.  The holes are represented by
3356: the spinless fermion forming a small hole pockets around $k=(\pi/2,\pi/2)$.
3357: The size of the hole pocket is twice as large as the usual doped SDW state due
3358: to the absence of the spin index. Therefore the slave fermion method violates
3359: the Luttinger theorem.  Finally we mention that by introducing a phase-string
3360: in the slave fermion approach, one obtains a phase-string formulation of high
3361: $T_c$ superconductivity (\cit{WST0067}; \cit{W0361}).  In such an approach both
3362: spin-1/2 neutral particles and spin-0 charged particles are bosons with a
3363: non-trivial mutual statistics between them.
3364: 
3365: 
3366: \section{$U(1)$ gauge theory of the uRVB state}
3367: 
3368: 
3369: %Insert A1
3370: The mean field theory only enforces the constraint on the average.
3371: Furthermore, the fermions and bosons introduce redundancy in representing the
3372: original electron, which results in an extra gauge degree of freedom.  The
3373: fermions and bosons are not gauge invariant and should not be thought of as
3374: physical particles.  To include these effects we need to consider fluctuations
3375: around the mean field saddle points, which immediately become gauge theories,
3376: as first pointed out by 
3377: %Baskaran and Anderson, 1988
3378: \Ref{BA8880}.  Here, we review the
3379: early work on the $U(1)$ gauge theory, which treats gauge fluctuations on the
3380: Gaussian level 
3381: %(Ioffe and Larkin, 1989; Nagaosa and Lee, 1990; Lee and Nagaosa, 1992; 
3382: %Ioffe and Kotliar, 1990)
3383: (\cit{IL8988}; \cit{NL9050}; \cit{LN9221}; \cit{IK9048}).  The theory can be
3384: worked out in some detail, leading to a nontrivial recipe for obtaining
3385: physical response functions in terms of the fermion and boson ones, called the
3386: Ioffe-Larkin composition rule.  It highlights the importance of calculating
3387: gauge invariant quantities and the fact that the fermion and bosons only enter
3388: as useful intermediate steps.  The Gaussian $U(1)$ gauge theory was mainly
3389: designed for the high temperature limit of the optimally doped cuprate, \ie
3390: the so-called strange metal phase in Fig.~\ref{U1}.  We will describe its
3391: failure in the underdoped region, which leads to the $SU(2)$ formulation of
3392: the next two sections.  The Gaussian theory also misses the confinement
3393: physics which is important for the ground state.
3394: 
3395: \subsection{Effective gauge action and non-Fermi-liquid behavior}
3396: 
3397: %Gauge structure often appears in the quantum field theory of constrained
3398: %system.  can be intuitively understood by the following analogy.  Gauge
3399: %theory is closely related to the differential geometry of the curved space.
3400: %When the vector is constrained on a tangential space to the curved surface,
3401: %for example, the metric for it is different from that in the three
3402: %dimensional free space. Particular interesting is the parallel transport of
3403: %the vector, which is defined so that the change of the vector has no
3404: %components within the tangential plane to the surface at each point. When one
3405: %continues this parallel transport along a closed loop and takesit back to the
3406: %original point, the resultant vector is different from the initial one.  This
3407: %phenomenon is called holonomy, and the difference of the final and initial
3408: %vectors are related to the curvature integrated over the surface enclosed by
3409: %the loop. Now in the case of quantum mechanics, one can consider the Hilbert
3410: %space and vector, which corresponds to the wavefunction. Then when the
3411: %wavefunction is constrained within the subspace of the Hilbert space, the
3412: %similar situation occurs.  Namely the parallel transport of the wavefunction
3413: %results in the quantum holonomy, and the curvature corresponds to the gauge
3414: %field.  
3415: 
3416: 
3417: As has been discussed in section III, the phenomenology of the optimally doped
3418: Mott insulator is required to describe the two seemingly contradicting
3419: features, \ie the doped insulator with small hole carrier concentration and
3420: the electrons forming the large Fermi surface. The former is supported various
3421: transport and optical properties, representatively the Drude weight
3422: proportional to $x$, while the latter by the angle resolved photoemission
3423: spectra (ARPES) in the normal state of optimal doped samples. In the
3424: conventional single-particle picture, the reduction of the 1st Brillouin zone
3425: due to the antiferromagnetic long range ordering (AFLRO) distinguishes these
3426: two. Namely small hole pockets with area $x$ are formed in the reduced
3427: 1st BZ in the AFLRO state, while the large metallic Fermi surface of area
3428: $1-x$ appears otherwise. The challenge for the theory of the optimally doped
3429: case is that aspects of the doped insulator appear in some experiments even
3430: with the large Fermi surface. Also it is noted that the ARPES shows that there
3431: is no sharp peak corresponding to the quasi-particle in the normal state,
3432: especially at the anti-nodal region near ${\v k} = (\pi,0)$. The fermi
3433: surface is defined by a rather broad peak dispersing near the Fermi energy.
3434: These strongly suggests that the normal state of high temperature
3435: superconductors is not described in terms of the usual Landau Fermi liquid
3436: picture. 
3437: 
3438: A promising theoretical framework to describe this dilemma is the slave-boson
3439: formalism introduced above. It has the two species of particles, \ie
3440: fermions and bosons, due to the strong correlation, and the electron is
3441: ``fractionalized'' into these two particles.  However, one must not take
3442: naively this conclusion, because the fermions and bosons cannot be regarded as
3443: ``physical'' particles in that they are not gauge invariant as explained
3444: below.  Furthermore, they are not noninteracting particles; they are strongly
3445: coupled to the gauge field. This arises from the fact that the conservation of
3446: the gauge charge $Q_{\v i} = \sum_\sigma f^\dagger_{\v i \sigma} f_{\v i \sigma} +
3447: b^\dagger_{\v i} b_{\v i}$ can be derived by the Noether theorem starting from the local
3448: $U(1)$ gauge transformation
3449: \be
3450: f_{\v i \sigma} \to e^{\v i \varphi_{\v i}} f_{\v i \sigma}
3451: \nonumber  \\
3452: b_{\v i} \to e^{i \varphi_{\v i}} b_{\v i \sigma}.
3453: \label{Eq.55}
3454: \en
3455: Therefore the constraint  eq.~(\ref{Eq.34})  is equivalent to a local gauge
3456: symmetry.  The Green's functions for fermions and bosons $G_F(\v i,\v j;\tau) = -
3457: \<T_\tau f_{\v i \sigma}(\tau) f^\dagger_{\v j \sigma} \>$ and $G_B(\v i,\v j;\tau) = -
3458: \<T_\tau b_{\v i}(\tau) b^\dagger_{\v j} \>$ transforms as
3459: \be
3460: G_F(\v i,\v j;\tau) \to e^{i (\varphi_{\v i} - \varphi_{\v j})} G_F(\v i,\v j;\tau)
3461: \nonumber \\
3462: G_B(\v i,\v j;\tau) \to e^{i (\varphi_{\v i} - \varphi_{\v j})} G_B(\v i,\v j;\tau).
3463: \label{Eq.56}
3464: \en
3465: Therefore these fermions and bosons are not gauge invariant and should be
3466: regarded as only the particles which are useful in the intermediate step of
3467: the theory to calculate the physical (gauge invariant) quantities as will be
3468: done in the next section. 
3469: %Here it is noted that this gauge symmetry is not the ``symmetry'' of the
3470: %system. It is rather a symmetry to impose the constraint, and can be
3471: %different from $U(1)$ as discussed in sec.X.  Namely it could be $Z_2$ or
3472: %$SU(2)$.  In this section, we will focus on the $U(1)$ formulation which was
3473: %originally proposed and reveal its limitations ans drawbacks.
3474: 
3475: At the mean field level, the constraint was replaced by the averaged one
3476: $\<Q_{\v i}\> = 1$. This average is controlled by the saddle point value of the
3477: Lagrange multiplier field $\<\lambda_{\v i}\> = \lambda$. Originally $\lambda_{\v i}$ is
3478: the functional integral variable and is a function of (imaginary) time. When
3479: this integration is done exactly, the constraint is imposed. Therefore we have
3480: to go beyond the mean field theory and take into account the fluctuation
3481: around it.  In other words, the local gauge symmetry is restored by the gauge
3482: fields which transform as
3483: \be
3484: a_{\v i\v j} \to a_{\v i\v j} + \varphi_{\v i} - \varphi_{\v j}
3485: \nonumber \\
3486: a_0 (\v i)  \to  a_0(\v i) + { {\partial \varphi_{\v i}(\tau)} \over {\partial \tau}},
3487: \label{Eq.57}
3488: \en
3489: corresponding to eq.~(\ref{Eq.55}).  The fields satisfying this condition are
3490: already in the Lagrangian eq.~(\ref{Eq.39}). Namely the phase of the HS
3491: variable $\chi_{\v i\v j}$ and the fluctuation part of the Lagrange multiplier
3492: $a_0(\v i) = \lambda_{\v i}$ are these fields. 
3493: 
3494: Let us study this $U(1)$ gauge theory for the uRVB state in the phase diagram
3495: Fig.~\ref{U1}. This state is expected to describe the normal state of the
3496: optimally doped cuprates, where the $SU(2)$ particle-hole symmetry 
3497: described by eq.~(\ref{Eq.37}) is not so important.   Here we neglect
3498: $\Delta$-field, and consider $\chi$ and $\lambda$ field. There are amplitude
3499: and phase fluctuations of $\chi$-field, but the former one is massive and does
3500: not play important roles in the low energy limit.  Therefore the relevant
3501: Lagrangian to start with is
3502: \be
3503: L_1 &=& \sum_{ i, \sigma} f^*_{\v i \sigma} 
3504: \biggl( {\partial \over {\partial \tau}} - \mu_F + i a_0({\v r}_{\v i})
3505: \biggr) f_{\v i \sigma} 
3506: \nonumber \\
3507:  &+& \sum_{ i} b^*_{\v i} 
3508: \biggl( {\partial \over {\partial \tau}} - \mu_B + i a_0({\v r}_{\v i})
3509: \biggr) b_{\v i} 
3510: \nonumber \\
3511: &-& {\tilde J}\chi \sum_{\<\v i\v j\> \sigma} 
3512: ( e^{i a_{\v i\v j}} f^*_{\v i \sigma} f_{\v j \sigma} + h.c.) \nonumber \\
3513: &-& t \eta \sum_{\<\v i\v j\>} 
3514: ( e^{i a_{\v i\v j}} b^*_{\v i} b_{\v j} + h.c.)
3515: \label{Eq.58}
3516: \en
3517: where $\eta$ is the saddle point value of another HS variable to decouple the
3518: hopping term.  We can take $\eta = \chi$ using eq.~(\ref{Eq.39a}). Here the
3519: lattice structure and the periodicity with respect to $a_{\v i\v j} \to a_{\v
3520: i\v j} + 2 \pi$ are evident, and the problem is that of the lattice gauge
3521: theory coupled to the fermions and bosons. It is also noted here that there is
3522: no dynamics of the gauge field at this starting Lagrangian. Namely the
3523: coupling constant of the gauge field is infinity, and the system is in the
3524: strong coupling limit.  This is because the gauge field represents the
3525: constraint; by integrating over the gauge field we obtain the original problem
3526: with the constraint.  This raises the issue of confinement as will be
3527: discussed in section  IX.D and XI.F.  Here we exchange the order of the
3528: integration between the gauge field ($a_{\v i\v j},a_0$) and the matter fields
3529: (fermions and bosons). Namely the matter fields are integrated over first, and
3530: we obtain the effective action for the gauge field.
3531: \be
3532: e^{- S_{\rm eff.} (a)} = 
3533: \int D f^* Df D b^* Db e^{ - \int_0^\beta L_1 }
3534: \label{Eq.59}
3535: \en
3536: However this integration can not be done exactly, and the approximation is
3537: introduced here. The most standard one is the Gaussian approximation or RPA,
3538: where the effective action is obtained by perturbation theory up to the
3539: quadratic order in $a$.  For this purpose we introduce here the continuum
3540: approximation to the Lagrangian $L_1$ in eq.~(\ref{Eq.58}).
3541: \be
3542: L &=& \int d^2 {\v r} 
3543: \biggl[  \sum_\sigma f^*_{ \sigma}({\v r}) 
3544: \biggl( {\partial \over {\partial \tau}} - \mu_F + i a_0({\v r})
3545: \biggr) f_{ \sigma}({\v r}) 
3546: \nonumber \\
3547:  &+& b^*({\v r}) 
3548: \biggl( {\partial \over {\partial \tau}} - \mu_B + i a_0({\v r})
3549: \biggr) b({\v r}) 
3550: \nonumber \\
3551:  &-& {1 \over {2 m_F}} 
3552: \sum_{\sigma, j = x,y} 
3553:  f^*_{ \sigma}({\v r}) 
3554: \biggl( {\partial \over {\partial x_j} } + i a_j
3555: \biggr)^2 f_{ \sigma}({\v r}) 
3556: \nonumber \\
3557:  &-& {1 \over {2 m_B}} 
3558: \sum_{j = x,y} 
3559: b^*({\v r}) 
3560: \biggl( {\partial \over {\partial x_j} } + i a_j
3561: \biggr)^2 b({\v r}) 
3562: \biggr],
3563: \label{Eq.60}
3564: \en
3565: where the vector field ${\v a}$ is introduced by $a_{\v i\v j} = ({\v r}_{\v i} - {\v
3566: r}_{\v j}) \cdot {\v a}[ ({\v r}_{\v i} + {\v r}_{\v j})/2]$.  Note $1/m_F \approx J$ and
3567: $1/m_B \approx t$.  The coupling between the matter fields and gauge field is
3568: given by
3569: \be
3570: L_{\rm int.} = \int d^2 {\v r} (j^F_\mu + j^B_\mu) a_\mu
3571: \label{Eq.61}
3572: \en
3573: where $j^F_\mu$ ($j^B_\mu$) is the fermion (boson) current density. 
3574: 
3575: Note that integration over $a_0$ recovers the constraint eq.~(\ref{Eq.34}) and
3576: integration over the vector potential $\v{a}$ yields the constraint
3577: \be
3578: \v{j}_F + \v{j}_B = 0 ,
3579: \label{Eq.61A}
3580: \en
3581: \ie the fermion and boson can move only by exchanging places.  Thus the
3582: Gaussian approximation apparently enforces the local constraint exactly
3583: \cite{L0094}.  We must caution that this is true only in the continuum limit,
3584: and an important lattice effect related to the $2\pi$ periodicity of the phase
3585: variable has been ignored.  These latter effects lead to instantons and
3586: confinement, as will be discussed later in section IX.D.  Thus it is not
3587: surprising that the ``exact'' treatment of D.H. Lee yields the same
3588: Ioffe-Larkin composition rule which is derived based on the Gaussian theory
3589: (see section IX.C).
3590: 
3591: We now proceed to reverse the order of integration.  We integrate out the
3592: fermion and boson fields to obtain an effective action for $a_\mu$.  We then
3593: consider the coupling of the fermions and bosons to the gauge fluctuations
3594: which are controlled by the effective action.  To avoid double counting, it
3595: may be useful to consider this procedure in the renormalization group sense,
3596: \ie we integrate out the high energy fermion and boson fields to produce an
3597: effective action of the gauge field which in turn modifies the low energy
3598: matter field. This way we convert the initial problem of infinite coupling to
3599: one of finite coupling.  The coupling is of order unity but may be formally
3600: organized as a $1/N$ expansion by artificially introducing $N$ species of
3601: fermions.  Alternatively, we can think of this as an RPA approximation, \ie
3602: a sum of fermion and boson bubbles.  The effective action for $a_\mu$ is given
3603: by  the following 
3604: \be
3605: S_{\rm eff.}^\text{RPA} (a) = 
3606: (\Pi^F_{\mu \nu} (q) + \Pi^B_{\mu \nu} (q) ) a_\mu(q) a_\nu(-q)
3607: \label{Eq.62}
3608: \en
3609: where $q= ({\v q},\omega_n)$ is a three dimensional vector.  The
3610: current-current correlation function $\Pi^F_{\mu \nu}(q)$ ($\Pi^B_{\mu
3611: \nu}(q)$) of the fermions (bosons) is given  by
3612: \be
3613: \Pi^\alpha_{\mu\nu}(q) = \< j_\mu^\alpha(q) j_\nu^\alpha(-q)\>
3614: \label{Eq.63}
3615: \en
3616: with $\alpha=F,B$.  Taking the transverse gauge by imposing the gauge fixing
3617: condition
3618: \be
3619: \nabla \cdot {\v a} = 0
3620: \label{Eq.64}
3621: \en
3622: the scalar ($\mu=0$) and vector parts of the gauge field dynamics are
3623: decoupled. The scalar part $\Pi^\alpha_{00}(q)$ corresponds to the
3624: density-density response function and does not show any singular behavior in
3625: the low energy/momentum limit. On the other hand, the transverse
3626: current-current response function shows singular behavior for small $\v q$ and
3627: $\om$.  Explicitly the fermion correlation function is given by
3628: \be
3629: \Pi^F_T(q) = i \omega \sigma^T_{F1}({\v q},\omega) - \chi_F 
3630: {\v q}^2
3631: \label{Eq.65}
3632: \en
3633: where $\chi_F = 1/(24 \pi m_F)$ is the fermion Landau diamagnetic
3634: susceptibililty.  The first term describes the dissipation and the static limit
3635: of $\sigma^T_{F1}$ (real part of the fermion conductivity) for $\omega <
3636: \gamma_{\v q}$ is $\sigma^T_{F1}({\v q},\omega) = \rho_F/(m_F \gamma_{\v
3637: q})$ where $\rho_F$ is the fermion density and
3638: \be
3639: \gamma_{\v q} &=& \tau_{\rm tr}^{-1} \ \ \ {\rm for}\
3640: |{\v q}| < (v_F  \tau_{\rm tr})^{-1}
3641: \nonumber \\
3642:  &=& v_F |{\v q}|/2  \ \ \ {\rm for}\ 
3643: |{\v q}| > (v_F  \tau_{\rm tr})^{-1}
3644: \label{Eq.66}
3645: \en
3646: where $\tau_{\rm tr}$ is the transport lifetime due to the scatterings by the
3647: disorder and/or the gauge field.  A similar expression is obtained for the
3648: bosonic contribution as
3649: \be
3650: \Pi^B_T(q) = i \omega \sigma^T_{B1}({\v q},\omega) - \chi_B 
3651: {\v q}^2
3652: \label{Eq.67}
3653: \en
3654: where $\chi_B = n(0)/(48 \pi m_B)$ where $n(\epsilon) $ is the Bose occupation
3655: factor.  $\chi_B$ diverges at the Bose condensation temperature $T_{BE}^{(0)}
3656: = 2 \pi x/ m_B$ when we assume a weak 3D transfer of the bosons. Assuming that
3657: the temperature is higher than $T_{BE}^{(0)}$, the boson conductivity is
3658: estimated as
3659: \be
3660: \sigma^T_{B1}
3661:  \cong x^{1/2}/|{\v q}|
3662: \label{Eq.68}
3663: \en
3664: for $|{\v q}|> \ell_B^{-1}$, where $\ell_B$ is the mean free path of the
3665: bosons. It can be seen from eqs. (\ref{Eq.66}) and (\ref{Eq.68}),
3666: $\sigma^T_{B1} \ll \sigma^T_{F1}$.
3667: 
3668: Summarizing, the propagator of the transverse gauge field is given by
3669: \be
3670: \<a_\alpha(q) a_{\beta}(-q)\> = ( \delta_{\alpha \beta}
3671: - q_\alpha q_\beta / |{\v q}|^2 ) D_T(q)
3672: \label{Eq.69}
3673: \en
3674: \be
3675: D_T(q) = [\Pi^F_T(q) + \Pi^B_T(q) ]^{-1}
3676: \cong [ i \omega \sigma ({\v q}) - \chi_d {\v q}^2 ]^{-1}.
3677: \label{Eq.70}
3678: \en
3679: Here  
3680: \be
3681: \sigma({\v q}) &\cong& k_0/|{\v q}| \ \ \ {\rm for}\ |{\v q}|\ell>1
3682: \nonumber \\
3683:  &\cong& k_0 \ell \ \ \ {\rm for}\ |{\v q}|\ell<1
3684: \label{Eq.71}
3685: \en
3686: where $\ell$ is the fermion mean free path and $k_0$ is of the order $k_F$ of
3687: the fermions.
3688: 
3689: This gauge field is coupled to the fermions and bosons and leads to their
3690: inelastic scatterings.  By estimating the lowest order self-energies of the
3691: fermion and boson propagators, it is found that these are diverging at any
3692: finite temperature.  It is because of the singular behavior of $D_T(q)$ for
3693: small $|{\v q}|$ and $\omega$.  This kind of singularity was first noted by
3694: %Reizer, 1989
3695: \Ref{R8902} for the problem of electrons coupled to a transverse
3696: electromagnetic field, even though related effects such as non-Fermi liquid
3697: corrections for the specific heat have been noted earlier 
3698: %(Holstein {\em et al.}, 1973)
3699: \cite{HNP7349}.  However this does not cause any trouble since the propagators
3700: of fermions and bosons are not the gauge invariant quantity and hence is not
3701: physical as discussed above.  As the representative of gauge invariant
3702: quantities, we consider the conductivity of fermions and bosons.  (Note that
3703: these are not still ``physical'' because one must combine these to obtain the
3704: physical conductivity as discussed in the next section.) For example the
3705: integral for the (inverse of) transport life-time $\tau_{\rm tr}$ contains the
3706: factor $1 - \cos \theta$ where $\theta$ is the angle between the initial and
3707: final momentum for the scattering. This factor scales with $|{\v q}|^2$ for
3708: small ${\v q}$, and gets rid of the divergence.  The explicit estimate gives
3709: \be
3710: { 1 \over {\tau^F_{\rm tr}} } &\cong& \xi_{\v k}^{4/3} \ \ \ {\rm for} \ \ 
3711: \xi_{\v k} > kT
3712: \nonumber \\
3713:  &\cong& T^{4/3} \ \ \ {\rm for} \ \ 
3714: \xi_{\v k} < kT
3715: \label{Eq. 72}
3716: \en
3717: for the fermions while
3718: \be
3719: { 1 \over {\tau^B_{\rm tr}} } &\cong& 
3720: {{ k T} \over { m_B \chi_d} }
3721: \label{Eq.73}
3722: \en
3723: for bosons.  These results are interpreted as the scattering by the
3724: fluctuating gauge flux 
3725: %enclosed 
3726: whose propagator is given by the loop representing the particle-hole
3727: propagator for the two-particle current-current correlation function.
3728:  
3729: Now some words on the physical meaning of the gauge field are in order. For
3730: simplicity let us consider the three sites, and that the electron is moving
3731: around these.  The quantum mechanical amplitude for this process is
3732: \be
3733: P_{123} = \<\chi_{12} \chi_{23} \chi_{34}\>
3734: =\< f^\dagger_{1 \alpha} f_{2 \alpha}  
3735:  f^\dagger_{2 \beta} f_{3 \beta}  
3736:  f^\dagger_{3 \gamma} f_{1 \gamma}\>.  
3737: \label{Eq.74}
3738: \en
3739: One can prove that
3740: \be
3741: (P_{123} - P_{132})/(4i)  =  {\v S}_1 \cdot
3742: ( {\v S}_2 \times {\v S}_3)
3743: \label{Eq.75}
3744: \en
3745: and the righthand side of the above equation corresponds to the solid angle
3746: subtended by the three vectors ${\v S}_1,{\v S}_2, {\v S}_3$, and is called
3747: spin chirality \cite{WWZcsp}.  Therefore the gauge field fluctuation is
3748: regarded as that of the spin chirality. Recently it is discussed that the spin
3749: chirality will produce the anomalous Hall effect in some ferromagnets such as
3750: manganites and pyrochlore oxides, where the non-coplanar spin configurations
3751: are realized by thermal excitation of the Skymion or the strong spin anisotropy
3752: in the ground state 
3753: %(Ye {\em et al.}, 1999; Taguchi {\em et al.}, 2001)
3754: (\cit{YKM9937}; \cit{TOY0173}).  This
3755: phenomenon can be interpreted as the static limit of the gauge field, while
3756: the gauge field discussed here has both quantum and thermal fluctuations.
3757: 
3758: \subsection{Ioffe-Larkin composition rule}
3759: 
3760: In order to discuss the physical properties of the total system, we have to
3761: combine the information obtained for fermions and bosons.  This has been first
3762: discussed by 
3763: %Ioffe and Larkin (1989)
3764: \Ref{IL8988}.  Let us start with the physical
3765: conductivity $\sigma$, which is given by
3766: \be
3767: \sigma^{-1} = \sigma_F^{-1} + \sigma_B^{-1}
3768: \label{Eq.76}
3769: \en
3770: in terms of the conductivities of fermions ($\sigma_F$) and bosons
3771: ($\sigma_B$).  This formula corresponds to the sequential circuit (not
3772: parallel) of the two resistance, and is intuitively understood from the fact
3773: that both fermions and bosons have to move subject to the constraint.  This
3774: formula can be derived in terms of the shift of the gauge field $\v{a}$, and
3775: resultant backflow effect.  In the presence of the external electric field
3776: ${\v E}$, the gauge field $\v{a}$ and hence the internal electric field ${\v
3777: e}$ is induced.  Let us assume that the external electric field ${\v E}$ is
3778: coupled to the fermions. Then the effective electric field 
3779: %which
3780: seen by the fermions 
3781: %are feeling 
3782: is
3783: \be
3784: {\v e}_F = {\v E} + {\v e}
3785: \label{Eq.77}
3786: \en
3787: while that for the boson is
3788: \be
3789: {\v e}_B = {\v e}.
3790: \label{Eq.78}
3791: \en
3792: The fermion current ${\v j}_F$ and boson current ${\v j}_B$ are induced,
3793: respectively as
3794: \be
3795: {\v j}_F = \sigma_F {\v e}_F, \ \ \ {\v j}_B = \sigma_F {\v e}_B.
3796: \label{Eq.79}
3797: \en
3798: The constraint ${\v j}_F + {\v j}_B = {\v 0}$ given by eq.~(\ref{Eq.61A})
3799: leads to the relation
3800: \be
3801: {\v e} = - { {\sigma_F} \over { \sigma_F + \sigma_B} } {\v E} .
3802: \label{Eq.81}
3803: \en
3804: The physical current ${\v j}$ given by
3805: \be
3806: {\v j} = {\v j}_F = - {\v j}_B = 
3807: { {\sigma_F \sigma_B } \over { \sigma_F + \sigma_B} } {\v E}
3808: \label{Eq.82}
3809: \en
3810: leading to the expression for the physical conductivity $\sigma$ in
3811: eq.~(\ref{Eq.76}).  It is also noted here that the same result 
3812: %can be 
3813: is obtained 
3814: %when 
3815: if instead we couple the e.m. field to bosons. 
3816: %in the starting.
3817: In this case the internal electric field ${\v e}$ is different, but ${\v
3818: e}_F$ and ${\v e}_B$ remain unchanged.  Therefore it is not a physical
3819: question which particle is charged, \ie fermion or boson. This is related to
3820: the fact that both fermions and bosons are not physical particles as
3821: repeatedly stated.  Note that $\sigma_F \gg \sigma_B$ in the uRVB state, we
3822: conclude that $\sigma \cong \sigma_B = x \tau_{\rm tr}^B/m_B$ which is
3823: inversely proportional to the temperature $T$. Furthermore the Drude weight
3824: of the optical conductivity is determined by $x/m_B$ as is observed
3825: experimentally. It remains true that the superfluidity density $\rho_S$ in the
3826: superconducting state is given by the missing oscillator strength below the
3827: gap, this also means that $\rho_s \propto x$.
3828: 
3829: A more formal way of deriving the physical electromagnetic response follows.
3830: We can generalize the discussion of the effective action $S_{\rm eff.}(\v{a})$
3831: for the gauge field to include the external e.m. field $A_\mu$. Let us couple
3832: $A_\mu$ again to the fermions. Then the effective action becomes instead of
3833: eq.~(\ref{Eq.62})
3834: \be
3835: S_{\rm eff.}^\text{RPA} (a,A) &=& 
3836: \Pi^F_{\mu \nu} (q) (a_\mu(q)+ A_\mu(q))( a_\nu(-q) + A_\nu(-q))
3837: \nonumber \\
3838:  &+& \Pi^B_{\mu \nu} (q)  a_\mu(q) a_\nu(-q).
3839: \label{Eq.83}
3840: \en
3841: Then after integrating over the gauge field $a_\mu$, we end up with the
3842: effective action for $A_\mu$ only as
3843: \be
3844: S_{\rm eff.}^\text{RPA} (A) = 
3845: \Pi_{\mu \nu} (q) A_\mu(q)A_\nu(-q)
3846: \label{Eq.84}
3847: \en
3848: with the physical e.m. response function 
3849: \be
3850: \Pi_\alpha(q)^{-1} = (\Pi_\alpha^F(q))^{-1}+(\Pi_\alpha^B(q))^{-1}
3851: \label{Eq.85}
3852: \en
3853: where $\alpha = 0$ or $T$ stands for the longitudinal and the transverse
3854: parts.  Then the physical diamagnetic susceptibility ${\tilde \chi}$ is given
3855: by ${\tilde \chi}^{-1} = \chi_F^{-1}+ \chi_B^{-1}$.  Again in the
3856: superconducting state, $\Pi_T^F \propto \rho^F_s$ and $\Pi_T^B \propto
3857: \rho^B_s$, where $\rho^F_s$ and $\rho_s^B$ are superfluidity density of the
3858: fermion pairing and boson condensation.  This leads to the composition rule
3859: for $\rho_s$ as $\rho_s^{-1} = (\rho^F_s)^{-1} + (\rho_s^B)^{-1} \cong
3860: (\rho^B_s)^{-1} \propto x^{-1}$ with $\rho^F_s \gg \rho_s^B $, reproducing the
3861: same result as suggested from the Drude weight.  
3862: On the other hand the temperature dependence of $\rho_s^F$ is of the form
3863: $\rho_s^F(T)=\rho_s^F(0)(1-aT)$ where $a$ is given by the nodal fermion
3864: dispersion, while the temperature dependence of $\rho_s^B$ is expected to be 
3865: higher power in $T$ and negligible. The Ioffe-Larkin composition rule then
3866: predicts that
3867: \begin{align}
3868: \label{rhosT}
3869:  \rho_s(T) &\approx \rho_s^B(1-\frac{\rho_s^B}{\rho_s^F}) \nonumber\\
3870: &\approx \rho_s^B(0) - \frac{(\rho_s^B(0))^2}{\rho_s^F(0)} aT .
3871: \end{align} 
3872: Since $\rho_s^B(0)\sim x$, this predicts that the temperature dependence of
3873: the superfluid density is proportional to $x^2$.
3874: Comparison with \Eq{Eq.5} implies that $\al\sim x$ in the slave-boson theory.
3875: As shown in Fig. \ref{Lemberger}, this prediction does not agree with
3876: experiment and is probably an indication of the breakdown of gaussian
3877: fluctuations which underlines the Ioffe-Larkin rule.
3878: 
3879: We conclude this section by remarking that the Ioffe-Larkin rule can be extended to
3880: various other physical quantities.  For example the Hall constant $R_H$ is
3881: given by
3882: \be
3883: R_H = { {R_H^F \chi_B + R_H^F \chi_F} \over
3884: {\chi_B + \chi_F} }
3885: \label{Eq.86}
3886: \en
3887: while the thermopower $S = S_B + S_F$ and the electronic thermal conductivity
3888: $\kappa = \kappa_B + \kappa_F$ are sum of the bosonic and fermionic
3889: contributions.
3890: 
3891: Compared with the two-particle correlation functions discussed above, the single
3892: particle Green's function is more complicated.  At the mean field level, the
3893: electron Green's function is given by the product of those of fermions and
3894: boson in the $({\v r},\tau)$ space. Therefore in the momentum-frequency
3895: space, it is given by the convolution. The spectral function is composed of
3896: the two contributions, one is the quasi-particle peak with the weight $\sim x$
3897: while the other is the incoherent background. Even the former one is broadened
3898: due to the momentum distribution of the noncondensed bosons, \ie there is no
3899: quasi-particle peak in the strict sense.  This absence of the delta-functional
3900: peak occurs also in the $SU(2)$ theory in sec. XI indicating that the fermions 
3901: %is 
3902: are not free and hence can not be regarded as the quasi-particle.  On the
3903: other hand, the dispersion of this ``quasi-particle'' peak is determined by
3904: that of fermions, and hence its locus of zero energy constitutes the large
3905: Fermi surface enclosing the area $1-x$.  However this simple calculation does
3906: not reproduce some of the novel features in the ARPES experiments such as the
3907: ``Fermi arc'' in underdoped samples, which will be discussed later in section
3908: XI.
3909: 
3910: Combined with the discussion on the transport properties and the electron
3911: Green's function, the present uniform RVB state in the $U(1)$ formulation
3912: offers an explanation on the dichotomy between the doped Mott insulator and
3913: the metal with large Fermi surface.  In particular, the conclusion that the
3914: conductivity is dominated by the boson conductivity $\sigma \approx \sigma_B
3915: \approx x \tau_{tr}^B/m_B \approx xt T$ explains the linear $T$ resistivity
3916: which has been taken as a sign of non-Fermi liquid behavior from the beginning
3917: of high $T_c$ research.  However, we must caution that this conclusion was
3918: reached for $T > T_{BE}^{(0)}$ while in the experiment the linear $T$ behavior
3919: persists to much lower temperature near optimal doping.  It is possible that
3920: gauge fluctuations suppress the effective Bose condensation.  
3921: %D. Lee, D. Kim and P. Lee, 1996 
3922: \Ref{LKL9601}
3923: attempted  to include the effect of strong gauge fluctuations
3924: on the boson conductivity by assuming a quasi-static gauge fluctuation and
3925: treating the problem by quantum Monte Carlo.  The picture is that the boson
3926: tends to make self-retracing paths to cancel out the effect of the gauge
3927: field \cite{NL9133}.  They indeed find that the boson conductivity remains linear in $T$
3928: down to much lower temperature than $T_{BE}^{(0)}$.                                                                     
3929: 
3930: 
3931: 
3932: \subsection{Ginzburg-Landau theory and vortex structure}
3933:  
3934: Up to now, we have focused on the uRVB state where the pairing amplitude
3935: $\Delta$ of the fermions is zero. In this subsection we review the
3936: phenomenological Ginzburg-Landau theory to treat this pairing field.  The free
3937: energy for a single CuO$_2$ layer is given by
3938: \be
3939: F = F_F[\psi, {\v a}, {\v A}] + 
3940: F_B[\phi, {\v a}]
3941: + F_{\rm gauge} [ {\v a}]
3942: \label{Eq.87}
3943: \en
3944: with
3945: \be
3946: F_F[\psi, {\v a},{\v A}]
3947: &=& {{H_{cF}^2} \over {8 \pi}}
3948: \int d^2 r 
3949: \biggl[ 2 \xi_F^2 |( \nabla - 2 i {\v a} - i { {2e} \over c}
3950: {\v A}) \psi |^2
3951: \nonumber \\
3952: &+& 2 \text{sign}(T-T_D^{(0)}) |\psi|^2 + |\psi|^4 \biggr],
3953: \label{Eq.88}
3954: \en
3955: \be F_B[\psi, {\v a}] &=& {{H_{cB}^2} \over {8 \pi}} \int d^2 r \biggl[ 2
3956: \xi_B^2 |( \nabla - i {\v a}) \phi |^2 \nonumber \\ &+& 2
3957: \text{sign}(T-T_{BE}^{(0)}) |\phi|^2 + |\phi|^4 \biggr], \label{Eq.89} \en and
3958: \be
3959: F_{\rm gauge}[{\v a}]
3960: = \int d^2 r 
3961: \biggl[ \chi_F [ \nabla \times ({\v a}+(e/c){\v A})]^2
3962: + \chi_B (\nabla \times {\v a})^2
3963: \biggr]
3964: \label{Eq.90}
3965: \en
3966: where ${\v A}$ is the e.m. vector potential, $c$ is the velocity of light,
3967: and $\hbar$ is put to be unity. In the above equations, the optimal value of
3968: the order parameter is scaled to be unity, and hence the correlation lengths
3969: $\xi_B,\xi_F$ and the thermodynamic critical fields $H_{cF},H_{cB}$ are
3970: temperature dependent both for fermion pairing and Bose condensation. It is
3971: noted that the penetration length of the fermion pairing (boson condensation)
3972: $\lambda_F$ ($\lambda_B$) is related to $H_{cF}$ ($H_{cB}$) as $H_{cF} =
3973: \phi_0/(2 \sqrt{2} \pi \xi_F \lambda_F)$ ($H_{cB} = \phi_0/( \sqrt{2} \pi
3974: \xi_B \lambda_B)$).  We take the lattice constant as the unit of length.  Then
3975: $\xi_F(0) \sim J/\Delta$, $\xi_B \sim x^{-1/2}$, and the condensation energy
3976: per unit area is given by $H_{cF}(0)^2/(8 \pi) \sim \Delta^2/J$, and
3977: $H_{cB}(0)^2/(8 \pi) \sim t x^2$.
3978: 
3979: Now we consider the consequences derived from this GL free energy. One is on
3980: the interplay between the Berezinskii-Kosterlitz-Thouless (BKT) transitions
3981: for the fermion pairing and boson condensation. We consider the type II limit,
3982: and neglect ${\v A}$ for the moment. As is well known, the binding-unbinding
3983: of the topological vortex excitations leads to the novel phase transition (BKT
3984: transition) in 2D.  This is due to the logarithmic divergence of the vortex
3985: energy with respect to the sample size.  This energy is competing with the
3986: entropy term which is also logarithmically diverging. Above some critical
3987: temperature the entropy dominates, and the free vortex excitations are
3988: liberated resulting in the exponential decay of the order parameter. However
3989: this logarithmic divergence is cut-off when the order parameter is coupled to
3990: the massless gauge field ${\v a}$. Namely the gauge field screens the vortex
3991: current, and $| (\nabla - i{\v a}) \phi|$ and $|(\nabla - 2 i{\v a})\psi|$
3992: decays exponentially beyond some penetration lengths. This means that the BKT
3993: transition for the fermion pairing and boson condensation disappear when the
3994: gauge field ${\v a}$ is massless. In other words, these two order parameters
3995: are coupled through the gauge field, and the BKT transition occurs only
3996: simultaneously where the gauge field becomes massive due to the Higgs
3997: mechanism.  Therefore the phase transition lines for fermion pairing and boson
3998: condensation in the phase diagram Fig.~\ref{U1}  become the crossover
3999: lines and only the superconducting transition remains to be the real BKT
4000: transition. 
4001: 
4002: Now we turn to the vortex structures in the superconducting state. The most
4003: intriguing issue here is the quantization of the magnetic flux. Because the
4004: boson has charge $e$ while the fermion pairing $-2e$, the question is whether
4005: the $hc/e$ vortex may be more stable than the conventional $hc/2e$ vortex. To
4006: study this issue, we compare the energy cost of the two types of vortex
4007: structure, \ie (i) type A: the fermion pairing order parameter $\psi$
4008: vanished at the core, with its phase winding around it. The boson condensation
4009: does not vanish and the vortex core state is the Fermi liquid. The flux
4010: quantization is $hc/2e$. (ii) type B: the Bose condensation is destroyed at the
4011: core and the fermion pairing remains finite. Then the vortex core state is the
4012: spin gap state. The flux quantization is $hc/e$ in this case.  The energy of
4013: each vortex is estimated as follows. First the Ioffe-Larkin composition rule
4014: results in the penetration length $\lambda$ of the magnetic field as 
4015: \be
4016: \lambda^2 = \lambda_F^2 + \lambda_B^2,
4017: \label{Eq.91}
4018: \en
4019: which is equivalent to $\rho_s^{-1} = (\rho^F_s)^{-1} + (\rho_s^B)^{-1}$
4020: derived in the previous subsection.  The contribution from the region where
4021: the distance from the core is larger than $\xi_F,\xi_B$ is estimated similarly
4022: to the usual case.
4023: \be
4024: E_0 = \biggl[ { {\phi_0} \over { 4 \pi \lambda} }\biggr]^2
4025: \ln \biggl[ { {\lambda} \over { {\rm max}(\xi_F, \xi_B)} } \biggr]
4026: \label{Eq.92}
4027: \en
4028: for the type A, and $4E_0$ for the type B because the quantized flux is
4029: doubled in the latter case.  The core energy $E_c$ is given by the
4030: condensation energy per area times the area of the core.  For type A vortex
4031: \be
4032: E_c^{(A)} \approx H_{cF}^2 \xi_F^2 \approx J
4033: \label{Eq.93}
4034: \en
4035: while for type B vortex
4036: \be
4037: E_c^{(B)} \approx H_{cB}^2 \xi_B^2 \approx tx .
4038: \label{Eq.94}
4039: \en
4040: Then the vortex energies are estimated to be $E^{(A)} \approx E_0 + E_c^{(A)}$
4041: and $E^{(B)} = 4E_0 + E_c^{(B)}$, respectively.  Note that $E_0$ is
4042: proportional to $\lambda^{-2}$ which is dominated by  $\lambda_B^{-2} = x$ and
4043: hence $E_0$, $E_c^{(B)}$ are proportional to $x$ while $E_c^{(A)}$ is a
4044: constant of order $J$.  The latter energy is in agreement with the estimate of
4045: the vortex in the BCS theory discussed in Section V.B and is the dominant
4046: energy for sufficiently small $x$.  We come to the conclusion that type B
4047: vortex (with $hc/e$ flux quantization) will be more stable in the underdoped
4048: region. This conclusion was reached by 
4049: %Sachdev, 1992 
4050: \Ref{S9289}
4051: and by
4052: %Nagaosa and Lee, 1992 
4053: \Ref{NL9266} and appears to be a general feature of the $U(1)$ gauge
4054: theory.  Unfortunately, the experimental search for stable $hc/e$ vortices
4055: have so far come up negative 
4056: %(Wynn {\em et al.}, 2001)
4057: \cite{WBG0102}.   In section XII.C we
4058: will describe how this problem is fixed by the $SU(2)$ gauge theory, which is
4059: designed to be more accurate for small doping. 
4060: 
4061: 
4062: \subsection{Confinement-deconfinement problem}
4063: 
4064: Despite the qualitative success of the mean field and $U(1)$ gauge field theory,
4065: there are several difficulties with this picture. One is that the gauge
4066: fluctuations are strong and one can not have a well controlled small expansion
4067: parameter, except rather formal ones such as the large $N$ expansion.  This
4068: issue is closely related to the confinement problem in lattice gauge theory,
4069: and will be discussed  below and also in section X.H and XI.F.
4070: 
4071: The coupling constant of the gauge field is defined as the inverse of the
4072: coefficient of $f_{\mu \nu}^2$ in the Lagrangian. It is well-known that the
4073: strong coupling gauge field leads to confinement.  In the confining phase,
4074: only the gauge singlet particles appear in the physical spectrum, which
4075: corresponds for example to the physical electron and magnon in the present
4076: context. Below we give a brief introduction to this issue.
4077: 
4078: Up to now the discussion is at the Gaussian fluctuation level where the
4079: effective action for the gauge field has been truncated at the quadratic order
4080: in the continuum approximation.  However we are starting from the
4081: infinite-coupling limit, and even if the finite coupling is produced by
4082: integrating over the matter field, the strong coupling effect must be
4083: considered seriously. In the original problem the gauge field is defined on
4084: the lattice and the periodicity with respect to $a_{\v i\v j} \to a_{\v i\v j} + 2 \pi$
4085: must be taken into account. Namely the relevant model is that of the {\em
4086: compact} lattice gauge theory.  Let us first consider the most fundamental
4087: model without the matter field;
4088: \be
4089: S_{\rm gauge}  =  - { 1 \over g} \sum_{\rm plaquette}
4090: ( 1 - \cos f_{\mu \nu})
4091: \label{Eq.95}
4092: \en
4093: where 
4094: \be
4095: f_{\mu \nu} =
4096:   a_{\v i, \v i+\v \mu} 
4097: + a_{\v i+\v \mu, \v i+ \v \mu + \v \nu}
4098: - a_{\v i+ \v \nu, \v i + \v \mu + \v \nu} 
4099: - a_{\v i, \v i +\v \nu}  
4100: \label{Eq.96}
4101: \en
4102: is the flux penetrating through the plaquette in the (d+1)-dimensional space,
4103: and $\mu,\nu=x,y,\cdots$.
4104: Now $S_{\rm gauge}$ is a periodic function of $f_{\mu\nu}$ with period $2\pi$
4105: and one can consider tunneling between different potential minima.  This leads
4106: to the ``Bloch state'' of $f_{\mu \nu}$ when the potential barrier height
4107: $1/g$ is low enough, while it is ``localized'' near one minimum when $1/g$ is
4108: high.  The former corresponds to the quantum disordered $f_{\mu \nu}$, and
4109: leads to the linear confining force as shown below (confining state).  On the
4110: other hand, in the latter case, one can neglect the compact nature of the
4111: gauge field, and the analysis in previous sections are justified (deconfining
4112: state).  For this confinement-deconfinement transition, one can define the
4113: following order parameter, \ie the Wilson loop:
4114: \be 
4115: W( C )= \< \exp [ i q \oint_C d {x}_\mu a_\mu (x) ]\>  
4116: \label{Eq.97}
4117: \en
4118: where the loop $C$ consists of the paths of length $T$ along the time
4119: direction and those of length $R$ along the spatial direction. It is related
4120: to the gauge potential $V(R)$ between the two static gauge charges $\pm q$
4121: with opposite sign put at the distance $R$ as
4122: \be
4123: W( C ) = \exp[ - V(R)T ].
4124: \label{Eq.98}
4125: \en
4126: There are two types of behavior of $W(C)$, \ie (i) area law: $W( C ) \sim
4127: e^{- \alpha RT }$, and (ii) perimeter law: $W( C ) \sim e^{- \beta (R+T) }$,
4128: where $\alpha,\beta$ are constants.
4129: In the first case (i), the potential $V(R)$ is increasing linearly 
4130: in $R$, and hence the two gauge charges can never be free. Therefore it
4131: corresponds to confinement, while the other case (ii) to 
4132: deconfinement.
4133: 
4134: It is known that the compact QED (pure gauge model) 
4135: in (2+1)D is always confining however small the coupling 
4136: constant is \cite{Pol87}.
4137: The argument is based on the instanton configuration,
4138: which is enabled by the compactness of the gauge field.
4139: This instanton is the source of the flux with the field distribution 
4140: \be 
4141: {\v b}({\v x}) = { { {\v x}} \over { 2|{\v x}|^3} }.
4142: \label{Eq.99}
4143: \en
4144: where ${\v x} = ({\v r},\tau)$ is the (2+1)D coordinates in the imaginary time
4145: formalism, and ${\v b}({\v x}) = ( e_y({\v x}), - e_x({\v x}), b({\v x}))$ is
4146: the combination of the ``electric field'' $e_\alpha({\v x})$ and ``magnetic
4147: field'' $b({\v x})$.  This corresponds to the tunneling phenomenon of the flux
4148: because the total flux slightly above (future) or below (past) of the
4149: instanton differs by $2 \pi$. The anti-instanton corresponds to the sink of
4150: the flux.  This instanton/anti-instanton corresponds to the singular
4151: configuration in the continuous approximation, but is allowed in the compact
4152: model on a lattice.  Therefore (anti)instantons take into account the compact
4153: nature of the original model in the continuum approximation.  It is also clear
4154: from eq. (\ref{Eq.99}) that the (anti)instanton behaves as the (negative)
4155: positive magnetic charge. Then it is evident that when we plug in the
4156: (anti)instanton configurations into the action
4157: \be
4158: S= \int d^3 {\v x}{ 1 \over {2g}}  [{\v b}({\v x})]^2
4159: \label{Eq.100}
4160: \en
4161: ( $g$ is the coupling constant), we obtain the Coulomb $1/|{\v
4162: x}|$-interaction between the (anti)instantons as
4163: \be
4164: S_{\rm inst} =  \sum_{i<j} {{q_i q_j} \over {| {\v x}_i - {\v x}_j |} }.
4165: \label{101}
4166: \en
4167: where $q_i$ is the magnetic charge, which is $\sqrt{g}/2$ for instanton and $
4168: -\sqrt{g}/2$ for anti-instanton.
4169: 
4170: Now it is well-known that the Coulomb gas in 3D is always in the screening
4171: phase, namely the long range Coulomb interaction is screened to be the short
4172: range one due to the cloud of the opposite charges surrounding the charge.
4173: Therefore the creation energy of the (anti)instanton is finite and the free
4174: magnetic charges are liberated. This free magnetic charges disorder the gauge
4175: field and  makes the Wilson loop show the area law, \ie confinement. 
4176: 
4177: The discussion up to now is for the pure gauge model without matter field.
4178: With matter field the confinement issue becomes very subtle since the Wilson
4179: loop does not work as the order parameter any more. Furthermore the
4180: confinement disappears above some transition temperature even in the pure
4181: gauge model. In the presence of matter field, the confinement-deconfinement
4182: transition at finite temperature is replaced by the gradual crossover to the
4183: plasma phase in the high temperature limit (\cit{P7877}; \cit{S7910}; \cit{S861}). 
4184: Therefore we can expect that the
4185: slave-boson theory without  confinement describes the physics of the
4186: intermediate energy scale even though the ground state is the confining state.
4187: Indeed, within the $U(1)$ gauge theory, the ground states are either
4188: antiferromagnetic, superconductor or Fermi-liquid and are all confining.
4189: Nevertheless, the pseudogap region which exists only at finite temperatures
4190: may be considered ``deconfined'' and describable by fermions and bosons
4191: coupled to noncompact gauge fields.  We emphasize once again that in this
4192: scenario the fermions and bosons are not to be considered free physical
4193: objects.  Their interaction with gauge fields are important and physical
4194: gauge-invariant quantities are governed by the Ioffe-Larkin rule within the
4195: Gaussian approximation.
4196: 
4197: It is of great interest to ask the question of whether a deconfined ground
4198: state is possible in a $U(1)$ gauge theory in the presence of matter field. This
4199: issue was first addressed in  a seminal paper by 
4200: %Fradkin and Shenker (1979)
4201: \Ref{FS7982}
4202: who considered a boson field coupled to a compact $U(1)$ gauge field.  The
4203: following bosonic action is added to $S_{\rm gauge}$ :
4204: \be
4205: S_B = t \sum_{\v i} \cos \left(
4206: \Delta_\mu \theta(\v r_{\v i}) - q a_\mu (\v r_{\v i})
4207: \right)
4208: \label{Eq.102}
4209: \en
4210: Here the Bose field is represented by phase fluctuation only, $\Delta_\mu$ is
4211: the lattice derivative and $a_\mu(\v r_{\v i}) = a_{\v i,\v i+\v \mu}$ 
4212: is the gauge field on
4213: the link $\v i, \v i+\v \mu$ and $q$ is an integer.  It is interesting to consider the
4214: phase diagram in the $t, g$ plane.  Along the $t=0$ line, we have pure gauge
4215: theory which is always confining in $2+1$ dimension.  For $g \ll 1$, gauge
4216: fluctuations are weak and $S_B$ reduces to the XY model weakly coupled to a
4217: $U(1)$ gauge field, which exhibits an ordered phase called the Higgs phase at
4218: zero temperature.  Note that in the Higgs phase, the gauge field is gapped by
4219: the Anderson-Higgs mechanism. On the other hand,  it is also gapped in the
4220: confinement phase due to the screening of magnetic charges described earlier.
4221: There is no easy way to distinguish between these two phases and the central
4222: result of Fradkin and Shenker is that for $q = 1$ the Higgs phase and the
4223: confinement phases are smoothly connected to each other.  Indeed, it was
4224: argued by 
4225: %Nagaosa and Lee, 2000
4226: \Ref{NL0066} that for the 1+2D case the entire $t$-$g$ plane is covered by
4227: the Higgs-confinement phase, with the exception of the line $g = 0$, which
4228: contains the XY transition.
4229: 
4230: The situation is dramatically different for $q = 2$, \ie if the boson field
4231: corresponds to a pairing field.  Then it is possible to distinguish between
4232: the Higgs phase and the confinement phase by asking whether two $q = \pm 1$
4233: have a linear confinement potential between them or not.  In this case there
4234: is a phase boundary between the confined and the Higgs phase, and the Higgs
4235: phase (the pairing phase) is deconfined.  One way of understanding this
4236: deconfinement is that the paired phase has a residual $Z_2$ gauge symmetry,
4237: \ie the pairing order parameter is invariant under a sign change of the
4238: underlying $q = 1$ fields which make up the pair.  Furthermore, it is known
4239: that the $Z_2$ gauge theory has a confinement-deconfinement transition in $2 +
4240: 1$ dimensions.  Thus the conclusion is that a compact $U(1)$ gauge theory
4241: coupled to a pair field can have a deconfined phase.  This is indeed the route
4242: to a deconfined ground state proposed by \Ref{RS9173} and \Ref{Wsrvb}.  
4243: In the context
4244: of the $U(1)$ gauge theory, the fermion pair field $\Delta$ plays the role of
4245: the $q = 2$ boson field in eq.~(\ref{Eq.102}). In such a phase,
4246: the spinon and holons are deconfined, leading to the phenomenon of spin and
4247: charge fractionalization.  A third elementary excitation in this theory is the
4248: $Z_2$ vortex, which is gapped.
4249: 
4250: \Ref{SF0050} pointed out that the square root of $\Delta$ carried unit gauge
4251: charge and one can combine this with the fermion to form a gauge invariant
4252: spinon and with the boson to form a gauge invariant ``chargon''.  The spinon
4253: and chargon only carry $Z_2$ gauge charges and can be considered almost free,
4254: They propose an experiment to look for the gapped $Z_2$ vortex but the results
4255: have so far been negative.  The connection between the $U(1)$ slave-boson
4256: theory and their $Z_2$ gauge theory was clarified by 
4257: %Senthil and Fisher, 2001a
4258: \Ref{SF0119}.  
4259: 
4260: There is yet another route to a deconfined ground state, and that is a
4261: coupling of a compact $U(1)$ gauge field to gapless fermions.  
4262: %Nagaosa, 1993
4263: \Ref{N9310}
4264: suggested that dissipation due to gapless excitations lead to  deconfinement.
4265: The special case of coupling to gapless Dirac fermions is of special interest.
4266: This route (called the $U(1)$ spin liquid) appears naturally in the $SU(2)$
4267: formulation and will be discussed in detail in section XI.F.  The hope
4268: expressed in section XII is to use the proximity to this deconfined state to
4269: understand the pseudogap state.  This is a more attractive scenario compared
4270: with the reliance purely on finite temperature to see deconfinement effects as
4271: described earlier in this section.
4272: 
4273: 
4274: 
4275: In the literature there have been some confusing discussions of the role of
4276: confinement in the gauge theory approach to strong correlation.  In
4277: particular, %Nayak
4278: \Ref{N0078}, \citeyear{N0193} has claimed that slave particles are
4279: always confined in $U(1)$ gauge theories.  His argument is based on the fact
4280: that since these gauge fields are introduced to enforce constraint, they
4281: do not have restoring force and the coupling constant is infinite.  What he
4282: overlooks is the possibility that partially integrating out the matter fields
4283: will generate restoring forces, which brings the problem to one of strong but
4284: finite coupling, and then sweeping conclusions can no longer be made.
4285: Comments by 
4286: %Ichinose and Matsui
4287: \Ref{IM0142}, 
4288: %Ichinose {\em et al.}
4289: \Ref{IMO0116}, and by 
4290: %Oshikawa\
4291: \Ref{O0301} have clarified the issues and
4292: in our opinion adequately answered Nayak's objections.  For example, 
4293: %Ichinose and Matsui (2001)
4294: \Ref{IM0142} pointed out that $3+1$ dimensional $SU(3)$ gauge theory
4295: coupled to $N$ fermions is in the deconfined phase even at infinite coupling
4296: for $N > 7$.  Another counter example is found by
4297: \Ref{Wqoslpub}, \Ref{RWspin} and 
4298: %Hermele {\em et al.} (2004)
4299: \Ref{HSF0451}
4300: who showed that the $U(1)$ gauge theory coupled to massless Dirac fermions is
4301: in a gapless phase (or the deconfined phase) for sufficiently large $N$ (see
4302: section XI.F).  There is also numerical evidence from Monte Carlo studies that
4303: the $SU(N)$ Hubbard-Heisenberg model at $N=4$ exhibits a gapless spin liquid
4304: phase, \ie a Mott insulator with power law spin correlation without breaking
4305: of lattice translation symmetry \cite{A0474}.  This spin liquid state is
4306: strongly suggestive of the stability of a deconfined phase with $U(1)$ gauge
4307: field coupled to Dirac fermions.
4308: 
4309: \subsection{Limitations of the $U(1)$ gauge theory} 
4310: 
4311: The $U(1)$ gauge theory, which only includes Gaussian fluctuations about mean
4312: field theory, suffers from several limitations which are especially serious in
4313: the underdoped regime.  Apart from the confinement issue discussed in the last
4314: section, we first mention a difficulty with the linear $T$ coefficient of the
4315: superfluid density.  As long as the gauge fluctuation is treated as Gaussian,
4316: the Ioffe-Larkin law holds and one predicts that the superfluid density
4317: $\rho_s(T)$ behaves as $\rho_s(T) \approx ax - bx^2T$.  The $ax$ term agrees
4318: with experiment while the $-bx^2T$ term does not (\cit{LW9711}; \cit{IM0109})
4319: as already explained in section V.A.  This failure is traced to the fact that
4320: in the Gaussian approximation, the current carried by the quasiparticles in
4321: the superconducting state is proportional to $x v_F$.  We believe this failure
4322: is a sign that nonperturbative effects again become important and confinement
4323: takes place, so that the low energy quasiparticles near the nodes behave like
4324: BCS quasiparticles which carry the full current $v_F$.  This is certainly
4325: beyond the Gaussian fluctuation treatment described here.
4326: 
4327: A second difficulty is that experimentally it is known from neutron scattering
4328: that spin correlations at $(\pi,\pi)$ are enhanced in the underdoped regime.
4329: This happens at the same time while a spin gap is forming in the pseudogap
4330: regime.  The $U(1)$ mean field theory explains the existence of the spin gap
4331: as due to fermion pairing.  However, this reduces the fermion density of
4332: states and it is not clear how one can get an enhancement of the spin
4333: correlation unless one introduces phenomenologically RPA interactions
4334: \cite{BL0102}.  The problem is more serious because the gauge field is gapped
4335: in the fermion paired state and one cannot use gauge fluctuation to enhance
4336: the spin correlation.  The gapping of the gauge field also tends to suppress
4337: fermion pairing self-consistently 
4338: %(Ubbens and Lee, 1994).
4339: \cite{UL9453}.  We shall see that both these difficulties are resolved by the
4340: $SU(2)$ formulation.
4341: 
4342: A third difficulty has to do with the structure of the vortex core in the
4343: underdoped limit.\cite{WLsu2}  As mentioned in section IX.C, the $U(1)$ gauge
4344: theory predicts the stability of $hc/e$ vortices, which has not been observed.
4345: This is a serious issue especially because the STM experiments show that the
4346: pseudogap remains in the vortex core. Therefore it should be type B in the
4347: $U(1)$ theory, which carries $hc/e$ flux.  On the other hand, the $hc/2e$ vortex
4348: is not ``cheap'' because the pairing amplitude vanishes and one has to pay the
4349: pairing energy at the core.  These difficulties arise because in the $U(1)$
4350: theory, the fermions becomes ``strong'' superconductor at low temperature in
4351: the underdoped region. However this contradicts with the fact that at
4352: half-filling the d-wave RVB state is equivalent to the $\pi$-flux state, which
4353: is not ``superconducting''.  In short, the $U(1)$ theory misses the important
4354: low lying fluctuation related to the $SU(2)$ particle-hole symmetry at
4355: half-filling.  By incorporating this symmetry to the gauge field even at
4356: finite doping, we will be lead to the $SU(2)$ gauge theory of high Tc
4357: superconductors, which we will next discuss.
4358: 
4359: 
4360: 
4361: 
4362: 
4363: \section{$SU(2)$ slave-boson representation for spin liquids}
4364: 
4365: In this section we are going to develop $SU(2)$ slave-boson theory for spin
4366: liquids and underdoped high $T_c$ superconductors.  The $SU(2)$ slave-boson
4367: theory is equivalent to the $U(1)$ slave-boson theory discussed in the last
4368: section. However, the $SU(2)$ formalism makes more symmetries of the
4369: slave-boson theory explicit. This makes it easier to see the low energy
4370: collective modes in the $SU(2)$ formalism, which in turn allows us to resolve
4371: some difficulties of the $U(1)$ slave-boson theory.\footnote{We would like to
4372: point out that those difficulties are not because the $U(1)$ slave-boson
4373: theory is incorrect. The difficulties are results of incorrect treatment of
4374: the $U(1)$ slave-boson theory, for example, overlooking some low energy soft
4375: modes.} To develop the $SU(2)$ slave-boson theory, let us first describe
4376: another way to understand the $U(1)$ gauge fluctuations in the slave-boson
4377: theory. In this section we will concentrate on undoped case where the model is
4378: just a pure spin system.  Even though the theory involves only fermionic
4379: representation of the spin in the underdoped case, we continue to refer to
4380: the theory as slave-boson theory in anticipation of the doped case.  We
4381: generalize the $SU(2)$  salve-boson theory to doped model in the next section.
4382: 
4383: \subsection{Where does the gauge structure come from?}
4384: 
4385: According to the $U(1)$ slave-boson mean-field theory, the fluctuations around
4386: the mean-field ground state are described by gauge fields and fermion fields.
4387: Remember that the original model is just a interacting spin model which is a
4388: purely bosonic model. How can a purely bosonic model contain excitations
4389: described by gauge fields and fermion fields?  Should we believe the result?
4390: 
4391: Let us examine how the results are obtained. We first split the bosonic spin
4392: operator into a product of two fermionic operators $\v S_{\v i}=\frac12
4393: f^\dag_{\v i}\v \si f_{\v i}$.  We then introduce a gauge field to glue the
4394: fermions back into a bosonic spin.  From this point of view it appears that
4395: the gauge bosons and the fermions are fake and their appearance is just a
4396: mathematical artifact. The appearance of the fermion field and gauge field in
4397: a purely bosonic model seems only indicates that the slave-boson theory is
4398: incorrect.
4399: %Their appearance seems entirely due to the mean-field approximation. 
4400: 
4401: However, we should not discard the slave-boson mean-field theory too quickly.
4402: It is actually capable of producing pictures that agree with the common sense:
4403: the excitations in a bosonic spin system are bosonic excitations corresponding
4404: to spin flips, provided that the gauge field is in a confining phase.
4405: %(although through a long detour).  It actually can reproduce a picture that
4406: %agrees with common sense: a spin model only contain bosonic spin-1
4407: %excitations, provided that the gauge field is in a confining phase. 
4408: In the confining phase of the $U(1)$ gauge theory, the fermions interact with
4409: each other through a linear potential and can never appear as quasiparticles
4410: at low energies. The gauge bosons have a large energy gap in the confining
4411: phase and are absent from the low energy spectrum.  The only low energy
4412: excitations are bound state of two fermions which carry spin-1 and are bosons.
4413: So the mean-field theory plus the gauge fluctuations, may not be very useful,
4414: but is not wrong. 
4415: 
4416: On the other hand, the slave-boson mean-field theory (plus gauge fluctuations)
4417: is also capable of producing pictures that defy the common senses, if the
4418: gauge field is in a deconfined phase. In this case the fermions and gauge
4419: bosons 
4420: %will 
4421: may appear as well defined quasiparticles. The question is do we believe the
4422: picture of deconfined phase? Do we believe the possibility of emergent gauge
4423: bosons and fermions from a purely bosonic model?  Clearly, the slave-boson
4424: construction outlined above is far too formal to convince most people to
4425: believe such drastic results.  However, recently, it was realized that some
4426: models \cite{K032,LWsta,Wqoem} can be solved by the slave-boson theory exactly
4427: \cite{Wqoexct}.  Those models are in deconfined phases and confirm the
4428: striking results of emergence of gauge bosons and fermions
4429: from the slave-boson theory.  
4430: 
4431: %It turns
4432: %out that a deconfined state correspond to a many-body correlated state that
4433: %contains a new kind of order.  The striking results from the slave-boson
4434: %construction are results of new kind of many-body correlation in the ground
4435: %state which correspond to a new kind of order - topological/quantum
4436: %order \cite{Wtoprev,Wqoslpub}.  
4437: 
4438: To have an intuitive picture of the correlated ground state
4439: which leads to emergent gauge bosons and fermions, let us try to understand
4440: how a mean-field ansatz $\chi_{\v i\v j}$ is connected to a physical spin wave
4441: function.
4442: % which has exactly one fermion per site.  
4443: We know that the ground
4444: state, $|\Psi_\text{mean}^{(\chi_{\v i \v j})}\>$, of the mean-field Hamiltonian
4445: \begin{align}
4446: \label{HmeanU1}
4447: H_\text{mean}=\tilde J \sum (\chi_{\v i\v j}f^\dag_{\v i}f_{\v j}+h.c.) 
4448: +\sum a_0(f^\dag_{\v i}f_{\v i}-1),
4449: \end{align}
4450: is not a valid wavefunction for
4451: the spin system, since it may not have one fermion per site.  To connect to
4452: physical spin wavefunction, we need to include fluctuations of $a_0$ to
4453: enforce the one-fermion-per-site constraint. With this understanding, we may
4454: obtain a valid wave-function of the spin system $\Psi_\text{spin}(\{\al_{\v
4455: i}\})$ by projecting the mean-field state to the subspace of
4456: one-fermion-per-site:
4457: \begin{equation}
4458: \Psi_\text{spin}^{(\chi_{\v i \v j})}(\{\al_{\v i}\}) 
4459: = \<0_f|\prod_{\v i} f_{\al_{\v i}\v i} |
4460: \Psi_\text{mean}^{(\chi_{\v i \v j})}\> .
4461: \label{PsiPsichi}
4462: \end{equation}
4463: where $|0_f\>$ is the state with no $f$-fermions: $f_{\al\v i}|0_f\>=0$.
4464: Eq. (\ref{PsiPsichi}) connects the mean-field ansatz to 
4465: physical spin wavefunction.
4466: It allows us to understand the physical meaning of the mean-field ansatz and
4467: mean-field fluctuations.
4468: 
4469: For example, the projection \Eq{PsiPsichi} give the gauge transformation
4470: \Eq{U1gaugetrans} a physical meaning.  Usually, for different choices of
4471: $\chi_{\v i\v j}$, the ground states of $H_\text{mean}$ \Eq{HmeanU1} correspond to
4472: different mean-field wavefunctions $|\Psi_\text{mean}^{(\chi_{\v i \v j})}\>$.
4473: After projection they lead to different physical spin wavefunctions
4474: $\Psi_\text{spin}^{(\t \chi_{\v i \v j})}(\{\al_{\v i}\})$. Thus we can regard
4475: $\chi_{\v i\v j}$ as labels that label different physical spin states.
4476: However, two mean-field ansatz $\chi_{\v i \v j}$ and $\t \chi_{\v i \v j}$
4477: related by a gauge transformation
4478: \begin{equation}
4479: \label{tchichiga}
4480:  \t \chi_{\v i \v j} = e^{i\th_{\v i}}  \chi_{\v i \v j}e^{-i\th_{\v j}}
4481: \end{equation}
4482: give rise to the same physical spin state after the projection
4483: \begin{equation}
4484: \label{tchichi}
4485:  \Psi_\text{spin}^{(\t \chi_{\v i \v j})}(\{\al_{\v i}\})
4486: = e^{i\sum_{\v i} \th_{\v i}}
4487: \Psi_\text{spin}^{(\chi_{\v i \v j})}(\{\al_{\v i}\})
4488: \end{equation}
4489: %Thus there is a many to one correspondence between the mean-field ansatz
4490: %$\chi_{\v i \v j}$ and the physical spin wavefunction obtained by the
4491: %projection.
4492: Thus $\chi_{\v i \v j}$ is not a one-to-one label, but a many-to-one label.
4493: This property is important for us to understand the unusual dynamical
4494: properties of $\chi_{\v i\v j}$ fluctuations.  Using many labels to label the
4495: same physical state also make our theory a gauge theory.
4496: 
4497: Let us consider how the many-to-one property or the gauge structure of
4498: $\chi_{\v i\v j}$ affect its dynamical properties.  If $\chi_{\v i\v j}$ was
4499: an one-to-one label of physical states, then $\chi_{\v i\v j}$ would be like
4500: the condensed boson amplitude $\<\phi(\v x,t)\>$ in boson superfluid or the
4501: condensed spin moment $\<\v S_{\v i}(t)\>$ in SDW state. The fluctuations of
4502: $\chi_{\v i\v j}$ would correspond to a bosonic mode similar to sound mode or
4503: spin-wave mode.\footnote{More precisely, the sound mode and spin-wave mode are
4504: so called scaler bosons. The fluctuations of local order parameters always
4505: give rise to scaler bosons.} However, $\chi_{\v i\v j}$ does not behave like
4506: local order parameters, such as $\<\phi(\v x,t)\>$ and $\<\v S_{\v i}(t)\>$,
4507: which label physical states without redundancy.  $\chi_{\v i\v j}$ is a
4508: many-to-one label as discussed above.  The many-to-one label creates a
4509: interesting situation when we consider the fluctuations of $\chi_{\v i \v j}$
4510: -- some fluctuations of $\chi_{\v i \v j}$ do not change the physical state
4511: and are unphysical. Those fluctuations are called the pure gauge fluctuations.
4512: %For a generic fluctuation $\del \chi_{\v i\v j}$, part of it is physical and
4513: %the other part is unphysical.  
4514: The effective theory for $\chi_{\v i\v j}$ must be gauge invariant: for
4515: example, the energy for the ansatz $\chi_{\v i\v j}$ satisfies
4516: \begin{equation*}
4517:  E(\chi_{\v i\v j})=E(e^{i\th_{\v i}}  \chi_{\v i \v j}e^{-i\th_{\v j}}).
4518: \end{equation*}
4519: If we consider the phase fluctuations of $\chi_{\v i\v j}=\bar\chi_{\v i\v
4520: j}e^{ia_{\v i\v j}}$, then the energy for the fluctuations $a_{\v i\v j}$
4521: satisfies
4522: \begin{equation*}
4523:  E(a_{\v i\v j})=E(a_{\v i \v j}+\th_{\v i}-\th_{\v j}).
4524: \end{equation*}
4525: This gauge invariant property of the energy (or more precisely, the action)
4526: drastically change the dynamical properties of the fluctuations.  It is this
4527: property that makes fluctuations of $a_{\v i \v j}$ behave like gauge bosons,
4528: which are very different from sound mode and spin-wave mode.\footnote{In the
4529: continuum limit, the gauge bosons are vector bosons -- bosons described by
4530: vector fields.}
4531: %We know that the gauge field $a_{\v i\v j}$ is not physical. Because of this,
4532: %it is hard to picture the ``shape'' of a gauge fluctuation.  The similar
4533: %thing also happen to the fermions since the fermion operator $f_{\v i\al}$ is
4534: %also not gauge invariant.  However, in 
4535: 
4536: If we believe that gauge bosons and fermions do appear as low energy
4537: excitations in the deconfined phase, then a natural question will be what do
4538: those excitations looks like?  The slave-boson construction \Eq{PsiPsichi}
4539: allows us to construct an explicit physical spin wavefunction that
4540: corresponds to a gauge fluctuation $a_{\v i\v j}$
4541: \begin{equation*}
4542: \Psi_\text{spin}^{(a_{\v i\v j})}
4543: = \<0_f|\prod_{\v i} f_{\al_{\v i}\v i} 
4544: |\Psi_\text{mean}^{(\bar \chi_{\v i \v j}e^{ia_{\v i\v j}})}\> .
4545: \end{equation*}
4546: We would like to mention that the gauge fluctuations affect the average
4547: \begin{equation*}
4548:  P_{123} = \<\chi_{12} \chi_{23} \chi_{31}\>=
4549:   \<\bar\chi_{12} \bar\chi_{23} \bar\chi_{31}\> e^{i(a_{12}+a_{23}+a_{31})}
4550: \end{equation*}
4551: Thus the $U(1)$ gauge fluctuations $a_{\v i\v j}$, or more precisely the flux
4552: of $U(1)$ gauge fluctuations $a_{12}+a_{23}+a_{31}$, correspond to the
4553: fluctuations of the spin chirality ${\v S}_1 \cdot ( {\v S}_2 \times {\v
4554: S}_3)=\frac{P_{123} - P_{132}}{4i}$ as pointed out in the last section.
4555: 
4556: Similarly, the slave-boson construction also allows us to construct a physical
4557: spin wavefunction that corresponds to a pair of the fermion excitations.  We
4558: start with the mean-field ground state with a pair of particle-hole
4559: excitations.  After the projection \Eq{PsiPsichi}, we obtain the physical spin
4560: wavefunctions that contain a pair of fermions:
4561: \begin{equation*}
4562: \Psi_\text{spin}^\text{ferm}(\v i_1,\la_1; \v i_2,\la_2)
4563: = \<0| (\prod_{\v i} f_{\al_{\v i}\v i}) 
4564: f^\dag_{\la_1\v i_1} f_{\la_2\v i_2} 
4565: |\Psi_\text{mean}^{(\bar \chi_{\v i \v j})}\> .
4566: \end{equation*}
4567: We see that the gauge fluctuation $a_{\v i\v j}$ and fermion excitation do
4568: have a physical ``shape'' given by the spin wavefunction
4569: $\Psi_\text{spin}^{(a_{\v i\v j})}$ and $\Psi_\text{spin}^\text{ferm}$,
4570: although the shape is too complicated to picture. 
4571: %Naively, one may say that the slave-boson construction produces a correlated
4572: %ground state $\Psi_\text{spin}^{0} = \<0|\prod_{\v i} f_{\al_{\v i}\v i}
4573: %|\Psi_\text{mean}^{(\chi^{(mean)}_{\v i \v j})}\>$. We may view the
4574: %complicated correlation in the ground state as a pattern of entanglement.
4575: %Such a pattern corresponds to the new order -- quantum order. The gauge
4576: %fluctuation described by $\Psi_\text{spin}^{(a_{\v i\v j})}$ can be viewed as
4577: %a fluctuation of the entanglement and the fermion excitations can be viewed
4578: %as topological defects in the entanglement. 
4579: 
4580: Certainly, the two types of excitations, the gauge fluctuations and the
4581: fermion excitations, interact with each other. The form of the interaction is
4582: determined by the fact that the fermions carry unit charge of the $U(1)$ gauge
4583: field. The low energy effective theory is given by \Eq{Eq.39} with $\Delta_{\v
4584: i\v j} = 0$ and $b_{\v i} = 0$.
4585: %Eq. 37
4586: 
4587: %This is as far as I can go without using the picture of string-net
4588: %condensation.
4589: 
4590: %If you are not satisfied by the physical picture presented and do not believe
4591: %that projective construction can produce emergent gauge bosons and fermions,
4592: %you may go directly to chapter \ref{sec:string}.  In chapter \ref{sec:string}
4593: %we construct several spin models which can be solved exactly (or
4594: %quasi-exactly) by the projective construction. Those models have emergent
4595: %$Z_2$/$U(1)$ gauge structure and fermions. The exact soluble models reveal
4596: %the string-net origin of the emergent gauge bosons and fermions.  The rest of
4597: %this chapter is for believers.  We assume projective construction make sense
4598: %and study its consequences. 
4599: 
4600: \subsection{What determines the gauge group?}
4601: 
4602: We have mentioned that the collective fluctuations around the a slave-boson
4603: mean-field ground state are described by $U(1)$ gauge field. Here we would
4604: like to ask why the gauge group is $U(1)$?  The reason for the gauge group to
4605: be $U(1)$ is that the fermion Hamiltonian and the mean-field Hamiltonian are
4606: invariant under the local $U(1)$ transformation
4607: \begin{equation*}
4608: \label{lclU1}
4609:  f_{\v i} \to e^{i\th_{\v i}} f_{\v i},\ \ \ \ \ \ \
4610:  \chi_{\v i \v j} \to e^{i\th_{\v i}}  \chi_{\v i \v j}e^{-i\th_{\v j}}
4611: \end{equation*}
4612: The reason that the fermion Hamiltonian is invariant under the local $U(1)$
4613: transformation is that the fermion Hamiltonian is a function of spin operator
4614: $\v S_{\v i}$ and the spin operator $\v S_{\v i}=\frac12 f_{\v i}^\dag \v \si
4615: f_{\v i}$ is invariant under the local $U(1)$ transformation.  So the gauge
4616: group is simply the group formed all the transformations between $f_{\up\v i}$
4617: and $f_{\down\v i}$ that leave the physical spin operator invariant.
4618: 
4619: \subsection{From $U(1)$ to $SU(2)$}
4620: 
4621: This deeper understanding of gauge transformation allows us to realize that
4622: $U(1)$ is only part of the gauge group. The full gauge group is actually
4623: $SU(2)$. To see the gauge group to be $SU(2)$ let us introduce
4624: \begin{equation*}
4625:  \psi_{1\v i}=f_{\up\v i},\ \ \ \ \ \
4626:  \psi_{2\v i}=f^\dag_{\down\v i}
4627: \end{equation*}
4628: We find
4629: \begin{align*}
4630:  S^+_{\v i}=& f_{\v i}^\dag  \si^+ f_{\v i} = 
4631: \frac12 (
4632: \psi_{1\v i}^\dag \psi_{2\v i}^\dag -\psi_{2\v i}^\dag \psi_{1\v i}^\dag )
4633: \nonumber\\
4634:  S^z_{\v i} = & \frac12 f_{\v i}^\dag  \si^z f_{\v i} = 
4635: \frac12 (\psi_{1\v i}^\dag \psi_{1\v i} 
4636: + \psi_{2\v i}^\dag \psi_{2\v i} -1)
4637: \end{align*}
4638: Now it is clear that $\v S_{\v i}$ and any Hamiltonian expressed in terms of
4639: $\v S_{\v i}$ are invariant under local $SU(2)$ gauge transformation:
4640: \begin{equation*}
4641: \bpm \psi_{1\v i}\\ \psi_{2\v i} \epm
4642: \to
4643: W_{\v i}\bpm \psi_{1\v i}\\ \psi_{2\v i} \epm,
4644: \ \ \ \ \ \ W_{\v i} \in SU(2)
4645: \end{equation*}
4646: The local $SU(2)$ invariance of the spin Hamiltonian implies that the
4647: mean-field Hamiltonian not only should have the $U(1)$ gauge invariance, it
4648: should also have the $SU(2)$ gauge invariance.
4649: 
4650: To write down the mean-field theory with explicit $SU(2)$ gauge invariance, we
4651: start with the mean-field ansatz that includes pairing correlation:
4652: %  The
4653: %generalized mean-field ansatz is described by $\chi_{\v i\v j}$ and $\Del_{\v
4654: %i\v j}$:
4655: \begin{align}
4656: \label{sl2a.4}
4657: \chi_{\v i \v j}\del_{\al\bt}&= 2 \< f_{\v i\al}^{\dag} f_{\v j \bt} \>,
4658: & \chi_{\v i \v j} &= \chi_{\v j \v i}^* .
4659: \nonumber \\
4660: \Del_{\v i \v j} \eps_{\al\bt}&=  2 \< f_{\v i \alpha}~f_{\v j \bt} \> ,
4661: & \Del_{\v i \v j} &=\Del_{\v j \v i} ,
4662: \end{align}
4663: After replacing fermion bi-linears with $\chi_{\v i\v j}$ and $\Del_{\v i\v
4664: j}$ in \Eq{Eq.35}, we obtain the following mean-field Hamiltonian with pairing
4665: \begin{align*}
4666: H_\text{mean}= &
4667: \ \ \sum_{\<\v i\v j\>} -\frac{3}{8} J_{\v i\v j} 
4668: \left[(\chi_{\v j \v i}f^\dag_{\v i\al}f_{\v j\al}  - \Del_{\v i \v j}
4669: ~f^\dag_{ i \alpha} f^\dag_{ j \bt}\eps_{\alpha \bt}) + h.c
4670: \right. \nonumber\\
4671: & \left.
4672:  - |\chi_{\v i \v j}|^2  -|\Del_{\v i \v j}|^2 \right]
4673: \end{align*}
4674: 
4675: However, the above mean-field Hamiltonian is incomplete.  We know that the
4676: physical Hilbert space is formed by states with one $f$-fermion per site. Such
4677: states correspond to states with even $\psi$-fermion per site.  The states
4678: with even $\psi$-fermion per site are $SU(2)$ singlet one every site. The
4679: operators $\psi_{\v i}^\dag \v \tau \psi_{\v i}$ that generate local $SU(2)$
4680: transformations vanishes within the physical Hilbert space, where $\v
4681: \tau=(\tau^1,\tau^2, \tau^3)$ are the Pauli matrices.  In the mean-field
4682: theory, we replace the constraint $\psi_{\v i}^\dag \v \tau \psi_{\v i}=0$ by
4683: its average
4684: \begin{equation*}
4685: \< \psi_{\v i}^\dag \v \tau \psi_{\v i} \>=0 .
4686: \end{equation*}
4687: The averaged constraint can be enforced by including Lagrangian multiplier
4688: $\sum_{\v i} a^l_0(\v i)\psi_{\v i}^\dag \tau^l \psi_{\v i}$ in the mean-field
4689: Hamiltonian. This way we obtain the mean-field Hamiltonian of $SU(2)$
4690: slave-boson theory \cite{AZH8845,DFM8826}:
4691: \begin{align}
4692: \label{sl2.6su2}
4693: &H_\text{mean} \\
4694: =&
4695: \ \ \sum_{\<\v i\v j\>} -\frac{3}{8} J_{\v i\v j} 
4696: \Big[(\chi_{\v j \v i}f^\dag_{\v i\al}f_{\v j\al}  - \Del_{\v i \v j}
4697: ~f^\dag_{ i \alpha} f^\dag_{ j \bt}\eps_{\alpha \bt}) + h.c
4698: \nonumber\\
4699: &\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
4700:  - |\chi_{\v i \v j}|^2  -|\Del_{\v i \v j}|^2 \Big]
4701: \nonumber \\
4702: & 
4703: \ \  +\sum_{\v i} \left[ a_0^3(f_{\v i\al}^\dag f_{\v i\al} -1)+[(a_0^1+i a_0^2)
4704:  f_{\v i\al} f_{\v i\bt}\eps_{\al\bt}+h.c.]\right]
4705: \nonumber 
4706: \end{align}
4707: So the mean-field ansatz that describes a $SU(2)$ slave-boson mean-field state
4708: is really given by $\chi_{\v i\v j}$, $\Del_{\v i\v j}$, and $\v a_0$.  We
4709: note that $\chi_{\v i\v j}$, $\Del_{\v i\v j}$, and $\v a_0$ are invariant
4710: under spin rotation. Thus the mean-field ground state of $H_\text{mean}$ is a spin
4711: singlet. Such a state describes a spin liquid state.
4712: 
4713: The $SU(2)$ mean-field Hamiltonian \Eq{sl2.6su2} is invariant under local
4714: $SU(2)$ gauge transformation. To see such an invariance explicitly, we need to
4715: rewrite \Eq{sl2.6su2} in terms of $\psi$:
4716: \begin{align}
4717: \label{sl2a.8}
4718: H_\text{mean} =& \sum_{\<\v i \v j\>} \frac{3}{8}
4719:  J_{\v i\v j} \left[ 
4720: \frac12 \Tr (U_{\v i \v j}^\dag~U_{\v i \v j}) +(\psi^\dag_ i
4721: U_{\v i \v j}
4722: \psi_ j +~h.c.)\right]  \nonumber\\
4723: &
4724: +\sum_{\v i}  a_0^l \psi_{\v i}^\dag \tau^l \psi_{\v i} 
4725: \end{align}
4726: where
4727: \begin{equation}
4728: U_{\v i \v j }= \bpm - \chi_{\v i \v j}^\ast &\Del_{\v i \v j} \\
4729:                \Del_{\v i \v j}^\ast & \chi_{\v i \v j}
4730:                                         \epm
4731: =U_{\v j \v i}^\dag
4732: \label{sl2a.7}
4733: \end{equation}
4734: Note that $\det (U) < 0$, so that $U_{\v j \v k}$ is not a member of $SU(2)$,
4735: but $iU_{\v j \v k}$ is a member up to a normalization constant.  From
4736: \Eq{sl2a.8} we now can see clearly that the mean-field Hamiltonian is
4737: invariant under a local $SU(2)$ transformation $W_{\v i}$:
4738: \begin{align}
4739: \psi_{\v i} \to & W_{\v i}\psi_{\v i}  \nonumber \\
4740: U_{\v i \v j} \to &W_{\v i}~U_{\v i \v j}~W^\dag_{\v j}    
4741: \label{sl2a.9}
4742: \end{align}
4743: 
4744: 
4745: We note that in contrast to $\Phi_{i\up}$ and $\Phi_{i\down}$ introduced in
4746: eq.~(\ref{Eq.37}), the doublet $\psi_{\v i}$ does not carry a spin index.  Thus the
4747: redundancy in the $\Phi_{i\sigma}$ representation is avoided, which accounts
4748: for a factor of 2 difference in front of the bilinear $\psi_{\v i}$ term in
4749: \Eq{sl2a.8} vs eq.~(\ref{Eq.44}).  However, the spin-rotation symmetry is not
4750: explicit in our formalism and it is hard to tell if \Eq{sl2a.8} describes a
4751: spin-rotation invariant state or not. In fact, for a general $U_{\v i \v j}$
4752: satisfying $U_{\v i \v j}=U_{\v j \v i}^\dag$, \Eq{sl2a.8} may not describe a
4753: spin-rotation invariant state.  But, if $U_{\v i \v j}$ has a form
4754: \begin{align}
4755: \label{UuW}
4756: U_{\v i \v j} =& \chi^\mu_{\v i\v j}\tau^\mu,\ \ \ \ \ \ \mu=0,1,2,3,\nonumber\\
4757: \chi^0_{\v i\v j}=& \text{imaginary}, \ \ \ \ \ \ 
4758: \chi^l_{\v i\v j} = \text{real},\ \ \ \ \ \ l=1,2,3,
4759: \end{align}
4760: then \Eq{sl2a.8} will describe a spin-rotation invariant state.  This is
4761: because the above $U_{\v i \v j}$ has the form of \Eq{sl2a.7}.  In this case
4762: \Eq{sl2a.8} can be rewritten as \Eq{sl2.6su2} where the spin-rotation
4763: invariance is explicit.  In \Eq{UuW}, $\tau^0$ is the identity matrix.
4764: % and $\tau^{1,2,3}$ are the Pauli matrices.  
4765: 
4766: Now the mean-field ansatz can be more compactly represented by $(U_{\v i\v j},
4767: \v a_0(\v i))$. Again the mean-field ansatz $(U_{\v i\v j}, \v a_0(\v i))$ can
4768: be viewed as a many-to-one label of physical spin states. The physical spin
4769: state labeled by $(U_{\v i\v j}, \v a(\v i))$ is given by
4770: \begin{equation*}
4771:  |\Psi^{(U_{\v i\v j},\v a_0(\v i))}_\text{spin}\>
4772: =\cP |\Psi^{(U_{\v i\v j},\v a_0(\v i))}_\text{mean}\>
4773: \end{equation*}
4774: where $|\Psi^{(U_{\v i\v j},\v a_0(\v i))}_\text{mean}\>$ is the ground state of
4775: the mean-field Hamiltonian \Eq{sl2a.8} and $\cP$ is the projection that
4776: project into the subspace with even numbers of $\psi$-fermion per site.  From
4777: the relation between the $f$-fermion and the $\psi$-fermion, we note that the
4778: state with zero $\psi$-fermion correspond to the spin-down state and the state
4779: with two $\psi$-fermions correspond to the spin-up state.  Since the states
4780: with even numbers of $\psi$-fermion per site are $SU(2)$ singlet on every
4781: site, we find that two mean-field ansatz $(U_{\v i\v j}, \v a_0(\v i))$ and
4782: $(\t U_{\v i\v j}, \t{\v a}(\v i))$ related by a local $SU(2)$ gauge
4783: transformation
4784: \begin{equation*}
4785: \t U_{\v i \v j} = W_{\v i}~U_{\v i \v j}~W^\dag_{\v j},\ \ \ \ \ \ \ \ 
4786: \t {\v a}_0(\v i)\cdot \v\tau  
4787: = W_{\v i}~\v a_0(\v i)\cdot \v\tau ~W^\dag_{\v i} .
4788: \end{equation*}
4789: label the same physical spin state
4790: \begin{equation*}
4791: \cP |\Psi^{(U_{\v i\v j},\v a_0(\v i))}_\text{mean}\>
4792: =\cP |\Psi^{(\t U_{\v i\v j},\t{\v a}_0(\v i))}_\text{mean}\>
4793: \end{equation*}
4794: This relation represents the physical meaning of the $SU(2)$ gauge structure.
4795: 
4796: Just like the $U(1)$ slave-boson theory, the fluctuations of the mean-field
4797: ansatz correspond to collective excitations.  In particular, the ``phase''
4798: fluctuations of $U_{\v i\v j}$ represent the potential gapless excitations.
4799: However, unlike the $U(1)$ slave-boson theory, the ``phase'' of $U_{\v i\v j}$
4800: is described by a two by two hermitian matrix $a^l_{\v i\v j}\tau^l$,
4801: $l=1,2,3$, on each link.  If $(\bar U_{\v i\v j},\bar{\v a}(\v i))$ is the
4802: ansatz that describe the mean-field ground state, then the potential gapless
4803: fluctuations are described by
4804: \begin{equation*}
4805:  U_{\v i\v j}= \bar U_{\v i\v j}e^{ia^l_{\v i\v j}\tau^l},\ \ \ \ \
4806: \v a_0(\v i)= \bar{\v a}_0(\v i)+ \del \v a_0(\v i).
4807: \end{equation*}
4808: Since $(U_{\v i\v j},\v a_0(\v i))$ is a many-to-one labeling, the
4809: fluctuations $(\v a_{\v i\v j}, \del \v a_0(\v i))$ correspond to $SU(2)$
4810: gauge fluctuations rather than usual bosonic collective modes such as phonon
4811: modes and spin waves.
4812: 
4813: 
4814: \subsection{A few mean-field ansatz for symmetric spin liquids}
4815: 
4816: After a general discussion of the $SU(2)$ slave-boson theory, let us discuss a
4817: few mean-field ansatz that have spin rotation, translation $T_{x,y}$, and
4818: parity $P_{x,y,xy}$ symmetries.  We will call such a spin state symmetric spin
4819: liquid.  Here $T_x$ and $T_y$ are translation in $x$- and $y$-directions, and
4820: $P_x$, $P_y$ and $P_{xy}$ are parity transformations $(x, y)\to (-x,y)$, $(x,
4821: y)\to (x,-y)$, and $(x, y)\to (y,x)$ respectively.  We note that  $P_{x,y,xy}$
4822: parity symmetries imply the $90^\circ$ rotation symmetry.  
4823: 
4824: %After finding some mean-field ansatz of symmetric spin liquids that have low
4825: %energies, we can then study the physical properties of those symmetric spin
4826: %liquid states and finally determine which spin liquids describe the under
4827: %doped high $T_c$ superconductors by comparing the theoretical results with
4828: %experiments.
4829: 
4830: We will concentrate on three simple mean-field ansatz that describe symmetric
4831: spin liquids:\\
4832: 
4833: \noindent
4834: $\pi$-flux liquid ($\pi$fL) state\footnote{
4835: This state was called $\pi$-flux ($\pi$F) state in literature.}
4836: \cite{AM8874}
4837: \begin{align}
4838: \label{piF}
4839: %U_{\v i, \v i + \v x} & = - \tau^3 \chi - i (-)^{\v i} \chi ,
4840: %\nonumber\\
4841: %U_{\v i, \v i + \v y} & = - \tau^3 \chi + i (-)^{\v i} \chi ,
4842: U_{\v i, \v i + \v x} & = - i(-)^{i_y} \chi ,
4843: \nonumber\\
4844: U_{\v i, \v i + \v y} & = - i\chi ,
4845: \end{align}
4846: staggered flux liquid (sfL)
4847: \footnote{In
4848: \Ref{WLsu2} and \Ref{LNNWsu2}, this phase was called simply the staggered
4849: flux (sF) state.  In this paper we reserve sF to denote the $U(1)$ mean field
4850: state which explicitly breaks translational symmetry and which exhibits
4851: staggered orbital currents, as originally described by 
4852: %Hsu {\em et al.}, 1991
4853: \Ref{HMA9166}.  This latter state is also called $d$-density wave (ddw),
4854: following 
4855: %Chakravarty {\em et al.}, 2001.
4856: \Ref{CLM0203}.}
4857: state \cite{AM8874}
4858: \begin{align}
4859: \label{sF}
4860: U_{\v i, \v i + \v x} & = - \tau^3 \chi - i (-)^{\v i} \Delta ,
4861: \nonumber\\
4862: U_{\v i, \v i + \v y} & = - \tau^3 \chi + i (-)^{\v i} \Delta ,
4863: \end{align}
4864: $Z_2$-gapped state \cite{Wsrvb}
4865: \begin{align}
4866: U_{\v i,\v i+\v x }=&  U_{ \v i,\v i+\v y }= -\chi \tau^3 \nonumber \\
4867: U_{\v i,\v i+ \v x + \v y} =&  \eta\tau^1 +\la\tau^2 \nonumber \\
4868: U_{\v i,\v i-\v x + \v y} =&  \eta\tau^1 -\la\tau^2 \nonumber\\
4869: a_0^{2,3} =&  0, \ \ \ a_0^1 \neq 0
4870: \label{Z2gA}
4871: \end{align}
4872: where $(-)^{\v i}\equiv (-)^{i_x+i_y}$.  Note that the $Z_2$ mean-field state
4873: has pairing along the diagonal bond.
4874: 
4875: At first sight, those mean-field ansatz appear not to have all the symmetries.
4876: For example, the $Z_2$-gapped ansatz are not invariant under the $P_x$ and
4877: $P_y$ parity transformations and the $\pi$fL and sfL ansatz are not invariant
4878: under translation in the $y$-direction. However, those ansatz do describe spin
4879: states that have all the symmetries. This is because the mean-field ansatz are
4880: many-to-one labels of the physical spin state, the non-invariance of the
4881: ansatz does not imply the non-invariance of the corresponding physical spin
4882: state after the projection.  We only require the mean-field ansatz to be
4883: invariant up to a $SU(2)$ gauge transformation in order for the projected
4884: physical spin state to have a symmetry. For example, a $P_{xy}$  parity
4885: transformation changes the sfL ansatz to
4886: \begin{align*}
4887: U_{\v i, \v i + \v x} & = - \tau^3 \chi + i (-)^{\v i} \Delta ,
4888: \nonumber\\
4889: U_{\v i, \v i + \v y} & = - \tau^3 \chi - i (-)^{\v i} \Delta ,
4890: \end{align*}
4891: The reflected ansatz can be transformed into the original ansatz via a $SU(2)$
4892: gauge transformation $W_{\v i}=(-)^{\v i}i\tau^1$.  Therefore, after the
4893: projection, the sfL ansatz describes a $P_{xy}$ parity symmetric spin state.
4894: Using the similar consideration, one can show that the above three ansatz
4895: are invariant under translation $T_{x,y}$ and parity $P_{x,y,xy}$ symmetry
4896: transformations followed by corresponding $SU(2)$ gauge transformations
4897: $G_{T_x,T_y}$ and $G_{P_x,P_y,P_{xy}}$, respectively. Thus the three ansatz all
4898: describe symmetric spin liquids.  In the following, we list the corresponding
4899: gauge transformations $G_{T_x,T_y}$ and $G_{P_x,P_y,P_{xy}}$ for the above
4900: three ansatz:\\
4901: 
4902: \noindent
4903: $\pi$fL state
4904: \begin{align}
4905: \label{GssF}
4906: %G_{T_x}(\v i) =& G_{T_y}(\v i) = i(-)^{\v i}\tau^1,  
4907: %&
4908: %G_{P_x}(\v i) =& G_{P_y}(\v i) = \tau^0, \nonumber\\
4909: %G_0(\v i) =& e^{i(-)^{\v i} (\th^1\tau^1+\th^2\tau^2)+i\th^3\tau^3}
4910: %& 
4911: %G_{P_{xy}}(\v i) =&  i(-)^{\v i}\tau^1, 
4912: G_{T_x}(\v i) =& (-)^{i_x}G_{T_y}(\v i) = \tau^0,  
4913: &
4914: G_{P_{xy}}(\v i) =&  (-)^{i_xi_y} \tau^0, 
4915: \nonumber\\
4916: (-)^{i_x}G_{P_x}(\v i) =& (-)^{i_y}G_{P_y}(\v i) = \tau^0, 
4917: & 
4918: G_0(\v i) =& e^{i\th^l\tau^l}
4919: \end{align}
4920: sfL state
4921: \begin{align}
4922: \label{GspiF}
4923: G_{T_x}(\v i) =& G_{T_y}(\v i) = i(-)^{\v i}\tau^1,  
4924: &
4925: G_{P_{xy}}(\v i) =&  i(-)^{\v i}\tau^1, 
4926: \nonumber\\
4927: G_{P_x}(\v i) =& G_{P_y}(\v i) = \tau^0, 
4928: & 
4929: G_0(\v i) =& e^{i\th\tau^3}
4930: \end{align}
4931: $Z_2$-gapped state
4932: \begin{align}
4933: \label{GsZ2gA}
4934: G_{T_x}(\v i) =& G_{T_y}(\v i) = i\tau^0,  
4935: &
4936: G_{P_{xy}}(\v i) =&  \tau^0, 
4937: \nonumber\\
4938: G_{P_x}(\v i) =& G_{P_y}(\v i) = (-)^{\v i}\tau^1, 
4939: & 
4940: G_0(\v i) =& -\tau^0
4941: \end{align}
4942: In the above we also list the pure gauge transformation $G_0(\v i)$ that leave
4943: the ansatz invariant.
4944: 
4945: \subsection{Physical properties of the symmetric spin liquids at mean-field
4946: level}
4947: 
4948: To understand the physical properties of the above three symmetric spin
4949: liquids, let us first ignore the mean-field fluctuations of $U_{\v i\v j}$ and
4950: consider the excitations at mean-field level.
4951: 
4952: At mean-field level, the excitations are spin-1/2 fermions $\psi$ (or $f$).
4953: Their spectrum is determined by the mean-field Hamiltonian \Eq{sl2a.8} (or
4954: \Eq{sl2.6su2}).  In the $\pi$fL state, the fermions has a dispersion
4955: \begin{equation*}
4956:  E_{\v k}=\pm \frac34 J|\chi| \sqrt{\sin^2 k_x + \sin^2 k_y}
4957: \end{equation*}
4958: In the sfL state, the dispersion is given by
4959: \begin{equation*}
4960:  E_{\v k}=\pm \frac34 J\sqrt{\chi^2 (\cos k_x+\cos k_y)^2 
4961:  + \Del^2 (\cos k_x - \cos k_y)^2 }
4962: \end{equation*}
4963: In the $Z_2$-gapped state, we have
4964: \begin{align*}
4965: E_{\v k}=&\pm \sqrt{\eps_{\v k}^2+\Del^2_{1\v k}+\Del^2_{2\v k}}
4966: \nonumber\\
4967: \eps_{\v k} = & -\frac34 J \chi (\cos k_x + \cos k_y) 
4968: \nonumber\\
4969: \Delta_{1\v k} = & \frac34  \eta J' \cos (k_x + k_y) +a^1_0
4970: \nonumber\\
4971: \Delta_{2\v k} = & \frac34 \la J' \cos (-k_x + k_y)
4972: \end{align*}
4973: where $J$ is the nearest-neighbor spin coupling and $J'$ is the
4974: next-nearest-neighbor spin coupling.  We find that the $\pi$fL state and the
4975: sfL
4976: state, at the mean-field level, have gapless spin-1/2 fermion excitations,
4977: while the $Z_2$-gapped state has gapped spin-1/2 fermion excitations.
4978: 
4979: %%Usually 
4980: %Hopefully the results from mean-field theory are at least qualitatively
4981: %correct and we can use the mean-field results to qualitatively understand the
4982: %properties of real physical systems.  If we accept this picture, we would
4983: %conclude that the above three spin liquids contain spin-1/2 fermionic
4984: %excitations.  This result is very striking.  Usually, the excitations in a
4985: %spin system are spin-1 bosons corresponding to spin flips. 
4986: 
4987: Should we trust the mean-field results from the slave-boson theory?  The
4988: answer is that it depends on the importance of the gauge fluctuations.  Unlike
4989: usual mean-field theory, the fluctuations in the slave-boson theory include
4990: gauge fluctuations which can generate confining interactions between the
4991: fermions. In this case the gauge interactions represent relevant perturbations
4992: and the mean-field state is said to be unstable.  The mean-field results from
4993: an unstable mean-field ansatz cannot be trusted and cannot be applied to real
4994: physical spin state. In particular, the spin-1/2 fermionic excitations in the
4995: mean-field theory in this case will not appear in the physical spectrum of
4996: real spin state.
4997: 
4998: If the dynamics of the gauge fluctuations is such that the gauge interaction
4999: is short ranged, then  the gauge interactions represent irrelevant
5000: perturbations and can be ignored.  In this case the mean-field state is said
5001: to be stable and the mean-field results can be applied to the real physical
5002: spin liquids.  In particular, the corresponding physical spin state contain
5003: fractionalized spin-1/2 fermionic excitations.
5004: 
5005: 
5006: 
5007: \subsection{Classical dynamics of the $SU(2)$ gauge fluctuations}
5008: 
5009: We have seen that the key to understand the physical properties of a spin
5010: liquid described by a mean-field ansatz $(U_{\v i\v j}, a^l_0)$ is to
5011: understand the dynamics of the $SU(2)$ gauge fluctuations.  To gain some
5012: intuitive understanding, let us treat the mean-field ansatz $(U_{\v i\v j}, \v
5013: a_0(\v i))$ as classical fields and study classical dynamics of their
5014: fluctuations.  The  dynamics of the fluctuations is determined by the
5015: effective Lagrangian $L_\text{eff}(U_{\v i\v j}(t), \v a_0(\v i)(t))$.  To obtained
5016: the effective Lagrangian, we start with the Lagrangian representation of the
5017: mean-field Hamiltonian
5018: \begin{align*}
5019:  L(\psi_{\v i}, U_{\v i\v j}, \v a_0)
5020: =&\sum_{\v i} i\psi^\dag_{\v i}\prt_t\psi_{\v i} -H_\text{mean}
5021: \end{align*}
5022: where $H_\text{mean}$ is given in \Eq{sl2a.8}. The effective Lagrangian $L_\text{eff}$
5023: is then obtained by integrating out $\psi$:
5024: \begin{equation*}
5025:  e^{i\int dt L_\text{eff}(U_{\v i\v j},\v a_0)}
5026: =\int \cD\psi\cD\psi^\dag e^{i\int dt L(\psi,U_{\v i\v j},\v a_0)}
5027: \end{equation*}
5028: We note that $L$ describes a system of fermions $\psi_{\v i}$ and $SU(2)$
5029: gauge fluctuations $U_{\v i\v j}$.  Thus the effective Lagrangian is invariant
5030: under the $SU(2)$ gauge transformation
5031: \begin{align}
5032: \label{Egauge}
5033: & L_\text{eff}(\t U_{\v i\v j}, \t{\v a}_0)=L_\text{eff}(U_{\v i\v j}, \v a_0),
5034: \nonumber\\
5035: & 
5036: \t U_{\v i\v j}= W_{\v i}(U_{\v i\v j})W_{\v j},\ \ \ \ 
5037: \t a^l_0(\v i)\tau^l= W_{\v i}a^l_0(\v i)\tau^lW_{\v i},
5038: \nonumber\\
5039: & W_{\v i}\in SU(2)
5040: \end{align}
5041: The classical equation of motion obtained from $L_\text{eff}(U_{\v i\v j},\v
5042: a_0)$ determines the classical dynamics of the fluctuations.
5043: 
5044: To see if the collective fluctuations are gapless, we would like to examine if
5045: the frequencies of the collective fluctuations are bound from below. We know
5046: that the time independent saddle point of $L_\text{eff}(U_{\v i\v j},\v
5047: a_0)$, $(\bar U_{\v i\v j},\bar {\v a}_0)$, corresponds to mean-field ground
5048: state ansatz, and $-L_\text{eff}(\bar U_{\v i\v j},\bar {\v a}_0)$ is the
5049: mean-field ground state energy. If we expand $-L_\text{eff}(\bar U_{\v i\v
5050: j}e^{ia^l_{\v i\v j}\tau^l},\bar {\v a}_0)$ to the second order in the
5051: fluctuation $a_{\v i\v j}$, then the presence or the absence of the mass term
5052: $a_{\v i\v j}^2$ will determine if the collective $SU(2)$ gauge fluctuations
5053: have an energy gap or not.
5054: 
5055: 
5056: %To see how $SU(2)$ gauge bosons commit suicide, we consider a lattice $SU(2)$
5057: %gauge theory. The lattice $SU(2)$ gauge field is given by link variables
5058: %$U_{\v i\v j} \in SU(2)$. The first-order mean-field theory \Eq{sl2a.12} is a
5059: %$SU(2)$ lattice gauge theory. The energy of a configuration is a function of
5060: %$U_{\v i\v j}$: $E(U_{\v i\v j})$.  The energy is invariant under $SU(2)$
5061: %gauge transformation \begin{align}
5062: %\label{Egauge}
5063: % E(\t U_{\v i\v j})=E(U_{\v i\v j}),\ \ \ \ \ \ 
5064: % \t U_{\v i\v j}= W_{\v i}(U_{\v i\v j})W_{\v j},\ \ \ \ \ \  W_{\v i}\in SU(2)
5065: %\end{align}
5066: %To understand the dynamic of the lattice $SU(2)$ gauge fluctuations, we write
5067: %$U_{\v i\v j}=U_{\v i\v j}^{(mean)}e^{ia_{\v i\v j}^l\tau^l}$ where two by two
5068: %matrices $a_{\v i\v j}$ on the links describe the gauge fluctuations. The
5069: %energy now can be written as $E(U_{\v i\v j}^{(mean)},e^{ia_{\v i\v
5070: %j}^l\tau^l})$. To see if the $SU(2)$ gauge fluctuations gain an energy gap or
5071: %not, we need to examine if $ E(U_{\v i\v j}^{(mean)},e^{ia_{\v i\v
5072: %j}^l\tau^l})$ contains a mass term $(a^l_{\v i\v j})^2$ or not, in the small
5073: %$a^l_{\v i\v j}$ limit.
5074: 
5075: To understand how the mean-field ansatz $\bar U_{\v i\v j}$ affect the
5076: dynamics of the gauge fluctuations, it is convenient to introduce the loop
5077: variable of the mean-field solution \index{flux!$SU(2)$}
5078: \begin{equation}
5079: \label{sl2a.15}
5080: P(C_{\v i})= (i\bar{U}_{\v i\v j}) (i\bar{U}_{\v j\v k})...  (i\bar{U}_{\v k\v i})
5081: \end{equation}
5082: Following the comment after \Eq{sl2a.7} $P(C_{\v i})$ belongs to $SU(2)$ and we can
5083: write $P(C_{\v i})$ as $P(C_{\v i})=e^{i\Phi(C_{\v i})}$, where $\Phi$ is the
5084: $SU(2)$ flux through the loop $C_{\v i}$: $\v i\to \v j\to \v k\to ..\to \v
5085: l\to \v i$ with base point $\v i$.
5086: %We will also call it $SU(2)$-flux operator.  \index{SU(2)-flux
5087: %operator@$SU(2)$-flux operator}
5088: The $SU(2)$ flux correspond to gauge field strength in the continuum limit.
5089: Compare with the $U(1)$ flux, the $SU(2)$ flux has two new features.  First
5090: the flux $\Phi$ is a two-by-two traceless Hermitian matrix.  If we expand
5091: $\Phi$ as $\Phi=\Phi^l\tau^l$, $l=1,2,3$ we can say that the flux is
5092: represented by a vector $\Phi^l$ in the $\tau^l$ space.  Second, the flux is
5093: not gauge invariant.  Under the gauge transformations, $\Phi(C_{\v i})$
5094: transforms as
5095: \begin{equation}
5096: \label{PWPW}
5097:  \Phi(C_{\v i}) \to W_{\v i} \Phi(C_{\v i}) W_{\v i}
5098: \end{equation}
5099: Such a transformation rotates the direction of the vector $\Phi^l$.
5100: %We note that the $SU(2)$ flux has a form $P(C)=\chi^0(C)\tau^0+i
5101: %\chi^l(C)\tau^l$.  Thus when $\chi^l\neq 0$, the $SU(2)$ flux has a sense of
5102: %direction in the $SU(2)$ space as indicated by $\chi^l$.  From \Eq{PWPW}, we
5103: %see that the local $SU(2)$  gauge transformations rotate the direction of the
5104: %$SU(2)$ flux.  
5105: Since the direction of the $SU(2)$ flux for loops with different base point
5106: can be rotated independently by the local $SU(2)$ gauge transformations, it is
5107: meaningless to directly compare the directions of $SU(2)$ flux for different
5108: base points.  However, it is quite meaningful to compare the directions of
5109: $SU(2)$ flux for loops with the same base point. We can divide different
5110: $SU(2)$ flux configurations into three classes based  on the $SU(2)$ flux
5111: through loops with the \emph{same} base point: (a) trivial $SU(2)$ flux where
5112: all $P(C)\propto \tau^0$, (b) collinear $SU(2)$ flux where all the $SU(2)$
5113: fluxes point in the same direction, and (c) non-collinear $SU(2)$ flux where
5114: $SU(2)$ flux for loops with the same base point are in different directions.
5115: \index{SU2 flux@$SU(2)$ flux!trivial} \index{SU2 flux@$SU(2)$ flux!collinear}
5116: \index{SU2 flux@$SU(2)$ flux!non-collinear} We will show below that different
5117: $SU(2)$ flux can lead to different dynamics for the gauge field \cite{Wsrvb}.
5118: 
5119: \subsubsection{Trivial $SU(2)$ flux}
5120: 
5121: First let us consider an ansatz $\bar U_{\v i\v j}$ with trivial $SU(2)$ flux
5122: $\Phi(C)=0$ for all the loops (such as the $\pi$fL ansatz in \Eq{piF}). We will
5123: call the state described by such an ansatz the $SU(2)$ state. 
5124: %The $SU(2)$ flux is invariant under the $SU(2)$ gauge transformation. We can
5125: %choose a mean-field ansatz (by
5126: We can perform a $SU(2)$ gauge transformations to transform the ansatz into a
5127: form where all $\bar U_{\v i\v j} \propto \tau^0$.
5128: %\footnote{This is hard to prove mathematically. Here we will just accept it
5129: %as a mathematical fact.}
5130: In this case, the gauge invariance of the effective Lagrangian implies that
5131: \begin{equation} 
5132: L_\text{eff}(\bar U_{\v i\v j}e^{ia_{\v i\v j}^l\tau^l} )=
5133: L_\text{eff}(\bar U_{\v i\v j}e^{i\th^l_{\v i}\tau^l} e^{ia_{\v i\v j}^l\tau^l}
5134: e^{-i\th^l_{\v j}\tau^l} ).
5135: \end{equation}
5136: Under gauge transformation $e^{i\th^1_{\v i}\tau^1}$, $a_{\v i\v j}^1$
5137: transform as $a_{\v i\v j}^1=a_{\v i\v j}^1+\th^1_{\v i}-\th^1_{\v j}$.  The
5138: mass term $(a^1_{\v i\v j})^2$ is not invariant under such a transformation
5139: and is thus not allowed.  Similarly, we can show that none of the mass terms
5140: $(a^1_{\v i\v j})^2$, $(a^2_{\v i\v j})^2$, and $(a^3_{\v i\v j})^2$ are
5141: allowed in the expansion of $L_\text{eff}$.  Thus the $SU(2)$ gauge fluctuations
5142: are gapless and appear at low energies.  
5143: 
5144: We note that all the pure gauge transformations $G_0(\v i)$ that leave the
5145: ansatz invariant form a group. We will call such a group invariant gauge group
5146: (IGG).  For the ansatz $\bar U_{\v i\v j} \propto \tau^0$,  the IGG is an
5147: $SU(2)$ group formed by uniform $SU(2)$ gauge transformation $G_0(\v
5148: i)=e^{i\th^l\tau^l}$. 
5149: %It is interesting to see that the gauge group 
5150: We recall from the last paragraph that the (classical) gapless gauge
5151: fluctuations 
5152: %is just the IGG of the ansatz.  Such a
5153: is also $SU(2)$.  Such a relation between the IGG and the gauge group of the
5154: gapless classical gauge fluctuations is general and applies to all the ansatz
5155: \cite{Wqoslpub}.  
5156: 
5157: To understand the dynamics of the gapless gauge fluctuations beyond the
5158: classical level, we need to treat two cases separately. In the first case, the
5159: fermions have a finite energy gap. Those fermions will generate the following
5160: low energy effective Lagrangian for the gauge fluctuations
5161: \begin{equation*}
5162:  \cL = \frac{g}{8\pi} \Tr f_{\mu\nu}f^{\mu\nu}
5163: \end{equation*}
5164: where $f_{\mu\nu}$ is a 2 by 2 matrix representing the field strength of the
5165: $SU(2)$ gauge field $a_{\v i\v j}$ in the continuum limit.  At classical
5166: level, such an effective Lagrangian leads to $\sim g\log (r)$ interaction
5167: between $SU(2)$ charges in two spatial dimensions.  So the gauge interaction
5168: at classical level is not confining (\ie not described by a linear potential).  However, if we go beyond the classical
5169: level (\ie beyond the quadratic approximation) and include the interactions
5170: between gauge fluctuations, the picture is changed completely.  In 1+2D, the
5171: interactions between gauge fluctuations change the $g\log (r)$ interaction to
5172: a linear confining interaction, regardless the value of the coupling constant
5173: $g$. So the $SU(2)$ mean-field states with gapped fermions  are not stable.
5174: The mean-field results from such ansatz cannot be trusted.
5175: 
5176: In the second case, the fermions are gapless and have a linear dispersion.  In
5177: the continuum limit, those fermions correspond to massless Dirac fermions.
5178: Those fermions will generate a non-local low energy effective Lagrangian for
5179: the gauge fluctuations, which roughly has a form, $ \cL = \frac{g}{8\pi} \Tr
5180: f_{\mu\nu}\frac{1}{\sqrt{-\prt^2}}f^{\mu\nu}$.  Due to the screening of
5181: massless fermions the interaction potential between $SU(2)$ charges becomes
5182: $\sim g/r$ at classical level.  Such an interaction represents a marginal
5183: perturbation.  It is a quite complicated matter to determine if the $SU(2)$
5184: states with gapless Dirac fermions are stable or not beyond the quadratic
5185: approximation.
5186: 
5187: 
5188: \subsubsection{Collinear $SU(2)$ flux}
5189: 
5190: Second, let us assume the $SU(2)$ flux is collinear.  This means the $SU(2)$
5191: flux for different loops with the same base point all point in the same
5192: direction. 
5193: However, the $SU(2)$ flux for loops with different base points may still point
5194: in different directions (even for the collinear $SU(2)$ flux).  Using the
5195: local $SU(2)$ gauge transformation we can always rotate the $SU(2)$ flux for
5196: different base points into the same direction, and we can pick this direction
5197: to be $\tau^3$ direction.  In this case all the $SU(2)$ flux have a form
5198: $P(C)\propto \chi^0(C)+ i\chi^3(C)\tau^3$.  We can choose a gauge such that
5199: the mean-field ansatz  have a form $\bar U_{\v i\v j}=ie^{i\phi_{\v i\v j}
5200: \tau^3}$.  
5201: %Since the ansatz is invariant under the global $U(1)$ gauge transformation
5202: %$e^{i\th\tau^3}$ but not $e^{i\th\tau^{1,2}}$, we say the $SU(2)$ gauge
5203: %structure is broken down to a $U(1)$ gauge structure.  
5204: The gauge invariance of the energy implies that
5205: \begin{equation} 
5206: L_\text{eff}(\bar U_{\v i\v j}e^{ia_{\v i\v j}^l\tau^l} )=
5207: L_\text{eff}(\bar U_{\v i\v j}e^{i\th_{\v i}\tau^3} e^{ia_{\v i\v j}^l\tau^l}
5208: e^{-i\th_{\v j}\tau^3} ).
5209: \end{equation}
5210: %In order for the gauge bosons to obtain a gap we need all the three mass terms
5211: %$ (a^l)^2$ to appear in the expansion of the energy.  However 
5212: When $a^{1,2}_{\v i\v j}=0$, The above reduces to
5213: \begin{equation} 
5214: L_\text{eff}(\bar U_{\v i\v j}e^{ia_{\v i\v j}^3\tau^3} )=
5215: L_\text{eff}(\bar U_{\v i\v j}e^{i(a_{\v i\v j}^3+\th_{\v i}-\th_{\v j})\tau^3}
5216: ).
5217: \label{sl2a.16}
5218: \end{equation}
5219: We find that the mass term $(a^3_{\v i\v j})^2$ is incompatible with
5220: \Eq{sl2a.16}.  Therefore at least the gauge field $a^3_{\v i\v j}$ is 
5221: gapless.  How
5222: about $a^1_{\v i\v j}$ and $a^2_{\v i\v j}$ gauge fields?  Let $P_A(\v i)$ be
5223: the $SU(2)$ flux through a loop with base point $\v i$.  If we assume all the
5224: gauge invariant terms that can appear in the effective Lagrangian do appear,
5225: then $L_\text{eff}(U_{\v i\v j})$ will contain the following term
5226: \begin{equation} 
5227: L_\text{eff}=a\Tr[P_A(\v i)iU_{\v i,\v i+\v x}P_A(\v i+\v x)iU_{\v i+\v x, \v i}]+...
5228: \label{sl3a.6c}
5229: \end{equation}
5230: If we write $iU_{\v i, \v i+\v x}$ as $\chi e^{i\phi_{\v i\v j}\tau^3}e^{i
5231: a_x^l\tau^l}$, using the fact $U_{\v i, \v i+\v x}=U^\dag_{\v i+\v x,\v i}$
5232: (see \Eq{sl2a.7}), and expand to $(a_x^l)^2$ order, \Eq{sl3a.6c} becomes
5233: \begin{equation} 
5234: L_\text{eff}=-\frac{1}{2} a\chi^2 \Tr ([P_A, a_x^l\tau^l]^2) +...
5235: \label{sl3a.6d}
5236: \end{equation}
5237: We see from \Eq{sl3a.6d} that the mass term for $a^1_{\v i\v j}$ and $a^2_{\v
5238: i\v j}$ are generated if $P_A\propto \tau^3$.  
5239: 
5240: To summarize, we find that if the $SU(2)$ flux is collinear, then the ansatz
5241: is invariant only under a $U(1)$ rotation $e^{i\th \v n\cdot\tau}$ where $\v
5242: n$ is the direction of the $SU(2)$ flux.  Thus the IGG=$U(1)$.  The collinear
5243: $SU(2)$ flux also break the $SU(2)$ gauge structure down to a $U(1)$ gauge
5244: structure, \ie the low lying gauge fluctuations are described by a $U(1)$
5245: gauge field.  Again we see that the IGG of the ansatz is the gauge group of
5246: the (classical) gapless gauge fluctuations.
5247: %The corresponding mean-field state will has  gapless $U(1)$ gauge
5248: %fluctuations.
5249: We will call the states with collinear $SU(2)$ flux the $U(1)$ states.  The
5250: sfL ansatz in \Eq{sF} is an example of collinear $SU(2)$ flux.
5251: 
5252: For the $U(1)$ states with gapped fermions, the fermions will generate the
5253: following effective Lagrangian for the gauge fluctuations
5254: \begin{equation*}
5255:  \cL = \frac{g}{8\pi} (\v e^2 - b^2)
5256: \end{equation*}
5257: where $\v e$ is the ``electric'' field and $b$ is the ``magnetic'' field of
5258: the $U(1)$ gauge field.  Again at classical level, the effective Lagrangian
5259: leads to $\sim g\log (r)$ interaction between $U(1)$ charges and the gauge
5260: interaction at classical level is not confining.  If we go beyond the
5261: classical level and include the interactions between gauge fluctuations
5262: induced by the space-time monopoles,  the $g\log (r)$ interaction will be
5263: changed to a linear confining interaction, regardless the value of the
5264: coupling constant $g$ \cite{P7729}. So the $U(1)$ mean-field states with
5265: gapped fermions  are not stable.  
5266: 
5267: If the fermions in the $U(1)$ state are gapless and are described by massless
5268: Dirac fermions (such as those in the sfL state), those fermions will generate a non-local low energy effective
5269: Lagrangian, which, at quadratic level, has a form
5270: \begin{equation}
5271: \label{LaNL}
5272:  \cL = \frac{g}{8\pi} f_{\mu\nu}\frac{1}{\sqrt{-\prt^2}}f^{\mu\nu}
5273: \end{equation}
5274: Again the screening of massless fermions change the $g\log (r)$ interaction to
5275: $g/r$ interactions between $U(1)$ charge, at lease at classical level.  Such
5276: an interaction represents a marginal perturbation.  Beyond the classical
5277: level, we will show in subsections \ref{asl} and \ref{screen} that, when there
5278: are many Dirac fermions, the $U(1)$ gauge interactions with Dirac fermions are
5279: exact marginal perturbations.  So the $U(1)$ states with enough gapless Dirac
5280: fermions are not unstable.  The mean-field theory can give us a good starting
5281: point to study the properties of the corresponding physical spin state (see
5282: subsection \ref{asl}).
5283: 
5284: \subsubsection{Non-collinear $SU(2)$ flux}
5285: 
5286: Third, we consider the situation where the $SU(2)$ flux is non-collinear.  In
5287: the above, we have shown that an $SU(2)$ flux $P_A$ can induce a mass term of
5288: form $\Tr([P_A, a_x^l\tau^l]^2)$. For a non-collinear $SU(2)$ flux
5289: configuration, we can have in \Eq{sl3a.6c} another $SU(2)$ flux, 
5290: %$P_A$ and 
5291: $P_B$, pointing in a different direction from $P_A$. The mass term will 
5292: %also 
5293: contain in addition to \Eq{sl3a.6d} a term $\Tr([P_B, a_x^l\tau^l]^2)$. In
5294: this case, the mass terms for all the $SU(2)$ gauge fields $(a^1_{\v i\v
5295: j})^2$, $(a^2_{\v i\v j})^2$, and $(a^3_{\v i\v j})^2$ will be generated.  All
5296: $SU(2)$ gauge bosons will gain an energy gap.  
5297: %Those gauge fluctuation can only generate short range interaction between
5298: %fermions.  The gauge interactions are irrelevant and the mean-field state
5299: %described by the ansatz will be a stable state.  
5300: 
5301: We note that ansatz $U_{\v i\v j}$ is always invariant under the global $Z_2$
5302: gauge transformation $-\tau^0$. So the IGG always contains a $Z_2$ subgroup
5303: and the $Z_2$ gauge structure is unbroken at low energies.  The global $Z_2$
5304: gauge transformation is the only invariance for the non-collinear ansatz. Thus
5305: IGG=$Z_2$ 
5306: %and the low energy gauge group is exactly $Z_2$.
5307: %Since the ansatz is invariant under a $Z_2$ transformation $W_{\v i}=-\tau^0$
5308: %but not other more general global $SU(2)$ gauge transformations (\ie
5309: %IGG=$Z_2$), the non-collinear $SU(2)$ flux break the $SU(2)$ gauge structure
5310: %down to a $Z_2$ gauge structure.  
5311: and the low energy effective theory is a $Z_2$ gauge theory.  
5312: %In sections \ref{trnANS} and \ref{sec:splq-3-5} we will study the low energy
5313: %properties of states with non-collinear $SU(2)$ flux. 
5314: We can show that the low energy properties of non-collinear states, such as
5315: the existence of $Z_2$ vortex and ground state degeneracy, are indeed
5316: identical to those of a $Z_2$ gauge theory. So we will call the state with
5317: non-collinear $SU(2)$ flux a $Z_2$ state.
5318: 
5319: In a $Z_2$ state, all the gauge fluctuations are gapped. Those fluctuations
5320: can only mediate short range interactions between fermions. The low energy
5321: fermions interact weakly and behave like free fermions. Therefore, including
5322: mean-field fluctuations does not qualitatively change the properties of the
5323: mean-field state. The gauge interactions are irrelevant and the $Z_2$
5324: mean-field state is stable at low energies.
5325: 
5326: A stable mean-field spin liquid state implies the existence of a real physical
5327: spin liquid. The physical properties of the stable mean-field state apply to
5328: the physical spin liquid. If we believe these two statements, then we can
5329: study the properties of a physical spin liquid by studying its corresponding
5330: stable mean-field state. Since the fermions are not confined in mean-field
5331: $Z_2$ states, the physical spin liquid derived from a mean-field $Z_2$ state
5332: contain neutral spin-1/2 fermions as its excitation. 
5333: %This is very striking result, after realizing that the spin model is a purely
5334: %bosonic model.
5335: 
5336: The $Z_2$-gapped ansatz in \Eq{Z2gA} is an example where the $SU(2)$ flux is
5337: non-collinear. To see this, let us consider the $SU(2)$ flux through two
5338: triangular loops $(\v i,\v i+\v y,\v i-\v x)$ and $(\v i,\v i+\v x,\v i+\v y)$
5339: with the same base point $\v i$:
5340: \begin{align*}
5341: U_{\v i,\v i+\v y} ~U_{\v i+\v y,\v i-\v x} ~U_{\v i-\v x,\v i} = 
5342:  -\chi^2 ( \eta\tau^1 +\la\tau^2 ),
5343: \nonumber\\ 
5344: U_{\v i,\v i+\v x} ~U_{\v i+\v x,\v i+\v y} ~U_{\v i+\v y,\v i} = 
5345:  -\chi^2 ( \eta\tau^1 -\la\tau^2 ).
5346: \end{align*}
5347: We see that when $\eta$ and $\la$ are non-zero, the $SU(2)$ flux is not
5348: collinear. Therefore, after projection, the $Z_2$-gapped ansatz give rise to a
5349: real physical spin liquid which contains fractionalized spin-1/2 neutral
5350: fermionic excitations \cite{Wsrvb}. The spin liquid also contains a $Z_2$
5351: vortex excitation.  The bound state of a spin-1/2 fermionic excitation and a
5352: $Z_2$ vortex give us a spin-1/2 bosonic excitation
5353: %Read and Chakraborty, 1989
5354: \cite{RC8933,Wsrvb}.
5355: 
5356: 
5357: %In this case $\Del_{\v i\v j}$ must be non-zero. Once the $SU(2)$ gauge
5358: %structure is broken, the infrared problems are well under control due to the
5359: %finite mass term.  In this case the low energy properties of the mean-field
5360: %theory and the corresponding spin liquid state can be reliably derived.
5361: 
5362: %In all the cases discussed above the infrared problem of the fermions with
5363: %the $SU(2)$ gauge interaction is resolved by opening energy gaps in gauge
5364: %fluctuations.  Due to those energy gaps the infrared behavior of the
5365: %mean-field theory is well under control. We can obtain the low energy
5366: %properties of the mean-field theory quite reliably. Because of the infrared
5367: %stability of the theory, those low energy properties are expected to be
5368: %robust against small perturbations.  Therefore it is reasonable to assume
5369: %that the low energy properties of the mean-field theory qualitatively
5370: %describe the low energy properties of the spin liquid state.  In the next
5371: %section, as an example, we will study a $T$ and $P$ symmetric spin liquid
5372: %state corresponding to the case B).
5373: 
5374: %Besides examining the $SU(2)$ flux, the gauge group of the (classical)
5375: %gapless gauge fluctuations can be determined by another more general method.
5376: %In the new method, we first calculate the invariant gauge group (IGG).  IGG
5377: %is a group formed by all the pure $SU(2)$ gauge transformations that leave
5378: %the mean-field ansatz $(U_{\v i\v j}, \v a_0(\v i))$ invariant.  It turns out
5379: %that %the gauge group of the gapless gauge fluctuations is nothing but the
5380: %IGG \cite{Wqoslpub}.  For ansatz with trivial, collinear, and non-collinear
5381: %$SU(2)$ flux, the IGG can be found to be $SU(2)$, $U(1)$, and $Z_2$
5382: %respectively. 
5383: 
5384: \subsection{The relation between different versions of slave-boson theory}
5385: 
5386: We have discussed two version of the slave-boson theories, the $U(1)$
5387: slave-boson theory and the $SU(2)$ slave-boson theory.  In \Ref{SF0050}, a
5388: third slave-boson theory -- $Z_2$ slave-boson theory -- was also proposed.
5389: Here we would like to point out that all the three version of the slave-boson
5390: theory are \emph{equivalent} description of the same spin-1/2 Heisenberg model
5391: on square lattice, if we treat the $SU(2)$, $U(1)$ or $Z_2$ gauge fluctuations
5392: exactly.
5393: 
5394: To understand the relation between the three version of the slave-boson
5395: theory, we would like to point out that the $SU(2)$, $U(1)$ or $Z_2$ gauge
5396: structures are introduced to project the fermion Hilbert space (which has four
5397: states per site) to the smaller spin-1/2 Hilbert space (which has two states
5398: per site).  In the $SU(2)$ slave-boson theory, we regard the two fermions
5399: $\psi_{1\v i}$ and $\psi_{2\v i}$ as an $SU(2)$ doublet.  Among the four
5400: fermion-states on each site, $|0\>$, $\psi^\dag_{1\v i}|0\>$, $\psi^\dag_{2\v
5401: i}|0\>$, and $\psi^\dag_{1\v i}\psi^\dag_{2\v i}|0\>$, only the $SU(2)$
5402: invariant state correspond to the physical spin state.  There are only two
5403: $SU(2)$ invariant states on each site: $|0\>$ and $\psi^\dag_{1\v
5404: i}\psi^\dag_{2\v i}|0\>$ which correspond to the spin-up and the spin-down
5405: states.  So the spin-1/2 Hilbert space is obtained from the fermion Hilbert
5406: space by projecting onto the local $SU(2)$ singlet subspace.
5407: 
5408: In the $U(1)$ slave-boson theory, we regard $\psi_{1\v i}$ as a charge $+1$
5409: fermion and $\psi_{2\v i}$ as a charge $-1$ fermion.  The spin-1/2 Hilbert
5410: space is obtained from the fermion Hilbert space by projecting onto the local
5411: charge neutral subspace.  Among the four fermion-states on each site, only two
5412: states $|0\>$ and $\psi^\dag_{1\v i}\psi^\dag_{2\v i}|0\>$ are charge neutral.
5413: % which correspond to the spin-up and the spin-down states.
5414: 
5415: In the $Z_2$ slave-boson theory, we regard $\psi_{a\v i}$ as a fermion that
5416: carries a unit $Z_2$-charge.  The spin-1/2 Hilbert space is obtained from the
5417: fermion Hilbert space by projecting onto the local $Z_2$-charge neutral
5418: subspace.  Again the two states $|0\>$ and $\psi^\dag_{1\v i}\psi^\dag_{2\v
5419: i}|0\>$ are the only $Z_2$-charge neutral states.
5420: 
5421: %So far 
5422: In the last subsection we discussed  $Z_2$, $U(1)$, and $SU(2)$ spin liquid
5423: states.  
5424: %We also discussed 
5425: These must not be confused with $Z_2$, $U(1)$, and $SU(2)$ slave-boson
5426: theories.  We would like to stress that $Z_2$, $U(1)$, and $SU(2)$ in the
5427: $Z_2$, $U(1)$, and $SU(2)$ spin liquid states  are gauge groups that appear in
5428: the low energy effective theories of those spin liquids. We will call those
5429: gauge group low energy gauge group.  They should not be confused with the
5430: $Z_2$, $U(1)$, and $SU(2)$ gauge groups in the $Z_2$, $U(1)$, and $SU(2)$
5431: slave-boson theories.  
5432: %The later are so called  
5433: We will call the latter high energy gauge groups.
5434: %In addition to the above $Z_2$ spin liquids, in this paper we will also study
5435: %many other spin liquids with different low energy gauge structures, such as
5436: %$U(1)$ and $SU(2)$ gauge structures. We will use the terms $Z_2$ spin
5437: %liquids, $U(1)$ spin liquids, and $SU(2)$ spin liquids to describe them. We
5438: %would like to stress that $Z_2$, $U(1)$, and $SU(2)$ here are gauge groups
5439: %that appear in the low energy effective theories of those spin liquids. They
5440: %should not be confused with the $Z_2$, $U(1)$, and $SU(2)$ gauge group in
5441: %slave-boson approach or other theories of the projective construction. The
5442: %latter are high energy gauge groups. 
5443: The high energy gauge groups have nothing to do with the low energy gauge
5444: groups.
5445: %that appear in the low energy effective theories.  
5446: A high energy $Z_2$ gauge theory (or a $Z_2$ slave-boson approach) can have a
5447: low energy effective theory that contains $SU(2)$, $U(1)$ or $Z_2$ gauge
5448: fluctuations.  Even the Heisenberg model, which has no gauge structure at
5449: lattice scale, can have a low energy effective theory that contains $SU(2)$,
5450: $U(1)$ or $Z_2$ gauge fluctuations.  The spin liquids studied in this paper
5451: all contain some kind of low energy gauge fluctuations. Despite their
5452: different low energy gauge groups, all those spin liquids can be constructed
5453: from any one of $SU(2)$, $U(1)$, or $Z_2$ slave-boson approaches.  After all,
5454: all those slave-boson approaches describe the same Heisenberg model and are
5455: equivalent.
5456: 
5457: The high energy gauge group is related to the way in which we write down the
5458: Hamiltonian. We can write Hamiltonian of the Heisenberg model in many
5459: different ways which can contain arbitrary high energy gauge group of our
5460: choice. We just need to split the spin into two, four, six, or some other even
5461: numbers of fermions.  While the low energy gauge group is a property of ground
5462: state of the spin model. It has nothing to do with how are we going to write
5463: down the Hamiltonian.  
5464: %The Hamiltonian of a given system can be written in several forms, and
5465: %different forms may have different high energy gauge groups.
5466: Thus we should not regard $Z_2$ spin liquids as the spin liquids constructed
5467: using $Z_2$ slave-boson approach.  A $Z_2$ spin liquid can be constructed and
5468: was first constructed within the $U(1)$ or $SU(2)$ slave-boson/slave-fermion
5469: approaches.  
5470: %A precise mathematical definition of the low energy gauge group will be given
5471: %in section \ref{psg}.
5472: 	      However, when we study a particular spin liquid state, a certain
5473: version of the slave-boson theory may be more convenient than other versions.
5474: Although a spin liquid can be described by all the versions of the slave-boson
5475: theory, sometimes a particular version may have the weakest fluctuations. 
5476: 
5477: \subsection{The emergence of gauge bosons and fermions in condensed matter
5478: systems}
5479: 
5480: In the early days, it was believed that a pure boson system can never generate
5481: gauge bosons and fermions. Rather, the gauge bosons and the fermions were
5482: regarded as fundamental.   The spin liquids discussed in this paper suggest
5483: that gauge bosons (or gauge structures) and fermions are not fundamental and
5484: can emerge from local bosonic model. Here we will discuss how those ideas were
5485: developed historically.
5486: 
5487: Let us first consider gauge bosons.  In the standard picture of gauge theory,
5488: the gauge potential $a_\mu$ is viewed as a geometrical object -- a connection
5489: of a fibre bundle.  However, there is another point of view about the gauge
5490: theory.  Many thinkers in theoretical physics were not happy with the
5491: redundancy of the gauge potential $a_\mu$.  It was realized in the early
5492: 1970's that one can use gauge invariant loop operators to characterize
5493: different phases of a gauge theory 
5494: (\cit{W7159}; \cit{W7445}; \cit{KS7595}).  It was later
5495: found that one can formulate the entire gauge theory using closed strings
5496: (\cit{BMK7793}; \cit{GRV7978}; \cit{M7991}; \cit{P7947}; \cit{F7987}; \cit{S8053}). Those studies reveal the
5497: intimate relation between gauge theories and closed-string theories --- a
5498: point of view very different from the geometrical notion of vector potential.
5499: 
5500: In a related development, it was found that gauge fields can emerge from a
5501: local bosonic model, if the bosonic model is in certain quantum phases.  This
5502: phenomenon is also called dynamical generation of gauge fields.  The emergence
5503: of gauge fields from local bosonic models has a long and complicated history.
5504: Emergent $U(1)$ gauge \emph{field} has been introduced in quantum disordered
5505: phase of 1+1D $CP^N$ model \cite{DDL7863,W7985}. In condensed matter physics,
5506: the $U(1)$ gauge \emph{field} have been found in the slave-boson approach to
5507: spin liquid states 
5508: (\cit{BA8880}; \cit{AM8874}).  
5509: The slave-boson approach not only has a $U(1)$ gauge
5510: field, it also has gapless fermion \emph{fields}.
5511: 
5512: %Due to instanton effects, 
5513: It
5514: is well known that the compact $U(1)$ gauge theory is confining in $1+1$ and
5515: $1+2$D \cite{P7582}.  
5516: %The effect of the matter field (the fermions) was not clear, but 
5517: The concern about
5518: confinement  led to an opinion that the $U(1)$ gauge field and the gapless
5519: fermion fields are just a unphysical artifact of the ``unreliable''
5520: slave-boson approach.  Thus the key to find emergent gauge bosons and emergent
5521: fermions is not to write down a Lagrangian that contain \emph{gauge fields}
5522: and \emph{Fermi fields}, but to show that gauge \emph{bosons} and
5523: \emph{fermions} actually appear in the physical low energy spectrum.  
5524: % In fact, for any given physical system,  
5525: %we can use many different form of Lagrangian to describe it.  In particular,
5526: %we can alway design a Lagrangian which contains a gauge field of an arbitrary
5527: %choice to describes the system.  However, a gauge field in a Lagrangian may
5528: %not give rise to a gauge boson that appear as a low energy
5529: %\emph{quasiparticle}.  
5530: However, only when the dynamics of gauge field is such that the gauge field is
5531: in the deconfined phase can the gauge boson appear as a low energy
5532: quasiparticle.  Thus after the initial finding of \Ref{DDL7863}; \Ref{W7985};
5533: \Ref{BA8880}; \Ref{AM8874}, many researches have been trying to find the
5534: deconfined phase of the gauge field.
5535: 
5536: %In high energy physics, a 3+1D local bosonic model with emergent deconfined
5537: %$U(1)$ gauge bosons was constructed in \Ref{FNN8035}.  It was suggested that
5538: %light in nature may be emergent.  In condensed matter physics, 
5539: One way to obtain deconfined phase is to give gauge boson a mass.  In 1988, it
5540: was shown that if we break the time reversal symmetry in a 2D spin-1/2 model,
5541: then the $U(1)$ gauge field from the slave-boson approach can be in a
5542: deconfined phase due to the appearance of Chern-Simons term (\cit{WWZcsp};
5543: \cit{KW8983}). The deconfined phase correspond to a spin liquid state of the
5544: spin-1/2 model \cite{KL8795} which is called chiral spin liquid.  The chiral
5545: spin state contains neutral spin-1/2 excitations that carry fractional
5546: statistics.  A second deconfined phase was found by breaking the $U(1)$ or
5547: $SU(2)$ gauge structure down to a $Z_2$ gauge structure. Such a phase contains
5548: a deconfined $Z_2$ gauge theory \cite{RS9173,Wsrvb} 
5549: %The $Z_2$ gauge theory also describe a spin liquid state which 
5550: and is called $Z_2$ spin liquid (or short ranged RVB state).\footnote{The
5551: $Z_2$ state obtained in \Ref{RS9173} breaks the $90^\circ$ rotation symmetry
5552: while the $Z_2$ state in \Ref{Wsrvb} has all the lattice symmetries.} The
5553: $Z_2$ spin state also contains neutral spin-1/2 excitations.  But now the
5554: spin-1/2 excitations are fermions and bosons.  
5555: 
5556: The above $Z_2$ spin liquids have a finite energy gap for their neutral
5557: spin-1/2 excitations. In \Ref{BFN9833}, a spin liquid with gapless spin-1/2
5558: excitations was constructed by studying quantum disordered $d$-wave
5559: superconductor.  Such a spin liquid was identified as a $Z_2$ spin liquid
5560: using a $Z_2$ slave-boson theory \cite{SF0050}.  The mean-field ansatz is
5561: given by  
5562: %(Ivanov \& Senthil) 
5563: \begin{align}
5564: \label{Z2lCs}
5565: U_{\v i,\v i+{\v x}} &= \chi \tau^3 +\eta \tau^1, & 
5566: U_{\v i,\v i+{\v y}} &= \chi \tau^3 -\eta \tau^1,
5567: \\
5568: a_0^3 \neq &0,  \ \ \ \ a^{1,2}_0=0,  
5569: &
5570: U_{\v i,\v i+{\v x}+\v y} &= 
5571: U_{\v i,\v i+{\v x}-\v y} = \ga \tau^3.
5572: \nonumber 
5573: \end{align}
5574: The diagonal hopping breaks particle-hole symmetry and breaks the $U(1)$
5575: symmetry of the $a^3_0=0$ $d$-wave pairing ansatz down to $Z_2$.  We will call
5576: such an ansatz $Z_2$-gapless ansatz.
5577: %(which was first written down in \Ref{KL8842}).  
5578: The ansatz describes a symmetric spin liquid, since it is invariant under the
5579: combined transformations $(G_{T_x}T_x, G_{T_y}T_{y},
5580: G_{P_x}P_x, G_{P_y}P_y, G_{P_{xy}}P_{xy}, G_0) $ with
5581: \begin{align}
5582: \label{GsZ2lCs}
5583:  G_{T_x}=&\tau^0, &
5584:  G_{T_y}=&\tau^0, &
5585:  G_0=&-\tau^0, \nonumber\\
5586:  G_{P_x}=&\tau^0, &
5587:  G_{P_y}=&\tau^0, &
5588:  G_{P_{xy}}=&i\tau^3 .
5589: \end{align}
5590: The fermion excitations are gapless only at four $\v k$ points with a linear
5591: dispersion.  
5592: 
5593: The $Z_2$-gapped state and the $Z_2$-gapless state are just two $Z_2$ states
5594: among over 100 $Z_2$ states that can be constructed within the $SU(2)$
5595: slave-boson theory \cite{Wqoslpub}.  The chiral spin liquid and the $Z_2$ spin
5596: liquids provide examples of emergent gauge structure and emergent fermions (or
5597: anyons).  However, those results were obtained using slave-boson theory, which
5598: is not very convincing to many people. 
5599: 
5600: In 1997, an exact soluble spin-1/2 model \cite{K032}
5601: \begin{equation*}
5602:  H_\text{exact}=16g \sum_{\v i} S^y_{\v i} S^x_{\v i+\hat{\v x}} S^y_{\v
5603: i+\hat{\v x}+\hat{\v y}} S^x_{\v i+\hat{\v y}}
5604: \end{equation*}
5605: was found. The $SU(2)$ slave-boson theory turns out to be exact for such a model
5606: \cite{Wqoexct}. That is by choosing a proper $SU(2)$ mean-field ansatz, the
5607: corresponding mean-field state give rise to an exact eigenstate of $H_\text{exact}$
5608: after the projection. In fact all the eigenstates of $H_\text{exact}$ can be
5609: obtained this way by choosing different mean-field ansatz.  The exact solution
5610: allows us to show the excitations of $H_\text{exact}$ to be fermions and $Z_2$
5611: vortices. This confirms the results obtained from the slave-boson theory.
5612: 
5613: %which confirms the results obtained form the salve-boson theory
5614: %\cite{Wqoexct}.  
5615: %Since the spin-1/2 model does not have the spin rotation
5616: %symmetry, the corresponding $SU(2)$ slave-boson mean-field Hamiltonian has a
5617: %more general form
5618: %\begin{equation}
5619: %\label{HmeanExct}
5620: % H_\text{mean}=
5621: %\sum_{\<\v i\v j\>} \left(
5622: %\psi^\dag_{a,\v i} u_{\v i\v j}^{ab} \psi_{b,\v j}
5623: %+\psi^\dag_{a,\v i} w_{\v i\v j}^{ab} \psi^\dag_{b,\v j}
5624: %+h.c. \right)
5625: %\end{equation}
5626: %where $a,b=1,2$.  We will use $u_{\v i\v j}$ and $w_{\v i\v j}$ to
5627: %denote the $2\times 2$ complex matrices whose elements are $u_{\v i\v
5628: %j}^{ab}$ and $w_{\v i\v j}^{ab}$. 
5629: %If we choose the mean-field ansatz to be
5630: %\begin{align}
5631: %\label{suw}
5632: %-w_{\v i,\v i+\hat{\v x}} =&  u_{\v i,\v i+\hat{\v x}}
5633: %=-is_{\v i,\v i+\hat{\v x}} (1+\tau^3), 
5634: %\nonumber\\
5635: %-w_{\v i,\v i+\hat{\v y}} =&  u_{\v i,\v i+\hat{\v y}}
5636: %=-is_{\v i,\v i+\hat{\v y}} (1-\tau^3)
5637: %\end{align}
5638: %where $s_{\v i\v j}=\pm 1$, then the ground state of $H_\text{mean}$ will give rise
5639: %to an exact eigenstate of $H_\text{exact}$ after the projection. The eigenvalue is
5640: %$g\sum s_{\v i,\v i+\v x} s_{\v i+\v x,\v i+\v x+\v y} s_{\v i+\v x+\v y,\v
5641: %i+\v y} s_{\v i+\v y,\v i} $. By choosing different signs of $s_{\v i\v j}$ on
5642: %different links we can obtain all the exact eigenstates of $H_\text{exact}$.  The
5643: %ground state of $H_\text{exact}$ is a $Z_2$ state since $s_{\v i\v j}$ is a
5644: %many-to-one label of the physical spin states. Two ansatz $s_{\v i\v j}$ and
5645: %$\t s_{\v i\v j}$ differ by a $Z_2$ gauge transformation $\t s_{\v i\v
5646: %j}=G_{\v i} s_{\v i\v j}G_{\v j}$, $G_{\v i}=\pm 1$, give rise to the same
5647: %spin state after the projection. 
5648: 
5649: More exactly soluble or quasi-exactly soluble models were find for dimmer
5650: model \cite{MS0181}, spin-1/2 model on Kagome lattice \cite{BFG0212}, boson
5651: model on square lattice \cite{SM0204}, and Josephson junction array
5652: \cite{IFI0203}.  A model of electrons coupled to pairing fluctuations, with a
5653: local constraint which results in a Mott insulator  that obeys the spin
5654: $SU(2)$, symmetry was also constructed \cite{MS0204}.  Those models realize
5655: the $Z_2$ states.  A boson model that realize $Z_3$ gauge structure
5656: \cite{M0308} and $U(1)$ gauge structure \cite{SM0204,Walight} were also found.
5657: 15 years after the slave-boson approach to the spin liquids, now it is easy to
5658: construct (quasi-)exactly soluble spin/boson models that have emergent gauge
5659: bosons and fermions.
5660: 
5661: We would like to point out that the spin liquids are not the first example of
5662: emergent fermions from local bosonic models.  The first example of emergent
5663: fermions, or more generally, emergent anyons is given by the FQH states.
5664: Although \Ref{ASW8422} only discussed how anyons can emerge from a fermion
5665: system in magnetic field, the same argument can be easily generalized to show
5666: how fermions and anyons  can emerge from a boson system in magnetic field.
5667: Also in 1987, in a study of resonating-valence-bond (RVB) states, emergent
5668: fermions (the spinons) were proposed in a nearest neighbor dimer model on
5669: square lattice (\cit{KRS8765}; \cit{RK8876}; \cit{RC8933}). But, according to
5670: the deconfinement picture, the results in \Ref{KRS8765} and \Ref{RK8876} 
5671: are valid only
5672: when the ground state of the dimer model is in the $Z_2$ deconfined phase. It
5673: appears that the dimer liquid on square lattice with only nearest neighbor
5674: dimers is not a deconfined state (\cit{RK8876}; \cit{RC8933}), and thus it is
5675: not clear if the nearest neighbor dimer model on square lattice \cite{RK8876}
5676: has the deconfined quasiparticles or not \cite{RC8933}.  However, on
5677: triangular lattice, the dimer liquid is indeed a $Z_2$ deconfined state
5678: \cite{MS0181}.  Therefore, the results in \Ref{KRS8765} and \Ref{RK8876} are valid for
5679: the triangular-lattice dimer model and deconfined quasiparticles do emerge in
5680: a dimer liquid on triangular lattice.  
5681: 
5682: All the above models with emergent fermions are 2D models, where the emergent
5683: fermions can be understood from binding flux to a charged particle
5684: \cite{ASW8422}. Recently, it was pointed out in \Ref{LWsta} that the key to
5685: emergent fermions is a string structure. Fermions can generally appear as ends
5686: of open strings in any dimensions, if the ground state has a condensation of
5687: closed strings. The string picture allows a construction of a 3D local bosonic
5688: model that has emergent fermions \cite{LWsta}.  According to this picture, all
5689: the models with emergent fermions contain closed-string condensation in their
5690: ground states.  Since the fluctuations of condensed closed strings are gauge
5691: fluctuations \cite{BMK7793,S8053,Walight}, this explains why the model with
5692: emergent fermions also have emergent gauge structures. Since the gauge charges
5693: are ends of open strings, this also explains why the emergent fermions always
5694: carry gauge charges.
5695: 
5696: The second way to obtain deconfined phase is to simply go to higher
5697: dimensions.  In 3+1 dimension, the gapless $U(1)$ fluctuations do not generate
5698: confining interactions. In 4+1 dimensions and above, even non-Abelian gauge
5699: theory can be in a deconfined phase.  So it is not surprising that one can
5700: construct bosonic models on cubic lattice that have emergent gapless photons
5701: ($U(1)$ gauge bosons) (\cit{Wlight}; \cit{MS0204}; \cit{Walight}).  
5702: 
5703: The third way to obtain deconfined phase is to include  gapless excitations
5704: which carry gauge charge. The charged gapless excitations can screen the gauge
5705: interaction to make it less confining.  We would like to remark that the
5706: deconfinement in this case has a different behavior than the previous two
5707: cases.  In the previous two cases, the charged particles in the deconfined
5708: phases become non-interacting quasiparticles at low energies.  In the present
5709: case, the deconfinement only means that those gapless charged particles remain
5710: to be gapless.  Those particles may not become non-interacting quasiparticles
5711: at low energies.  The spin liquids obtained from the sfL ansatz and the uRVB
5712: ansatz (given by \Eq{sF} with $\Del=0$) belong to this case. Those spin liquid
5713: are gapless.  But the gapless excitations are not described by free fermionic
5714: quasiparticles or free bosonic quasiparticles at low energies.  The uRVB state
5715: (upon doping) leads to strange metal states \cite{LN9221} with large Fermi
5716: surface. We will discuss the spin liquid obtained from the sfL ansatz in
5717: subsections \ref{asl} and \ref{screen}.
5718: 
5719: Finally, we remark that what is common among these three ways to get
5720: deconfinement is that instantons are irrelevant and a certain gauge flux is a
5721: conserved quantity.  We shall exploit this property in section XII.E.
5722: 
5723: \subsection{The projective symmetry group and quantum order}
5724: 
5725: The $Z_2$-gapped ansatz \Eq{Z2gA} and the $Z_2$-gapless ansatz \Eq{Z2lCs},
5726: after the projection, give rise to two spin liquid states. The two states have
5727: exactly the same symmetry. The question here is that whether there is a
5728: way to classify these as distinct phases.   According to Landau's symmetry
5729: breaking theory, two states with the same symmetry belong to the same phase.
5730: However, after the discovery of fractional quantum Hall states, we now know
5731: that Landau's symmetry breaking theory does not describe all the phases.
5732: Different quantum Hall states have the same symmetry, but yet they can belong
5733: to different phases since they contain different topological orders
5734: \cite{Wtoprev}. So it is possible that the two $Z_2$ spin liquids contain
5735: different orders that cannot be characterized by symmetry breaking and local
5736: order parameters. The issue here is to find a new set of universal quantum
5737: numbers that characterize the new orders.
5738: 
5739: To find a new set of universal quantum numbers, we note that although 
5740: the projected wavefunctions of the two
5741: $Z_2$ spin liquids have the same symmetry, their ansatz are invariant under
5742: the same set of symmetry transformations but followed by different gauge transformations (see
5743: \Eq{GsZ2gA} and \Eq{GsZ2lCs}). So the invariant group of the mean-field ansatz
5744: for the two spin liquids are different. The invariant group is called the
5745: Projective Symmetry Group (PSG).  The PSG is generated by the combined
5746: transformations $(G_{T_x}T_x, G_{T_y}T_y, G_{P_x}P_x, G_{P_y}P_y,
5747: G_{P_{xy}}P_{xy}) $ and $G_0$.  We note that the PSG is the symmetry group of
5748: the mean-field Hamiltonian.  Since the mean-field fluctuations in the $Z_2$
5749: states are weak and perturbative in nature, those fluctuations cannot change
5750: the symmetry group of the mean-field theory. Therefore, the PSG of an ansatz
5751: is a universal property, at least against perturbative fluctuations.  The PSG
5752: can be used to characterize the new order in the two $Z_2$ spin liquids
5753: \cite{Wqoslpub}. Such order is called the quantum order.  The two $Z_2$ spin
5754: liquids belong to two different phases since they have different PSG's and
5755: hence different quantum orders.
5756: 
5757: We know that the symmetry characterization of phases (or orders) have some
5758: important applications. It allows us to classify all the 230 crystal orders in
5759: three dimensions. The symmetry also produces and protects gapless collective
5760: excitations -- the Nambu-Goldstone bosons.  The PSG characterization of
5761: quantum orders has similar applications.  Using PSG, we can classify over 100
5762: different 2D $Z_2$ spin liquids that all have the same symmetry
5763: \cite{Wqoslpub}.  Just like the symmetry group, PSG can also produce and
5764: protect gapless excitations.  However, unlike the symmetry group, PSG can
5765: produce and protects gapless gauge bosons and fermions
5766: \cite{Wqoslpub,Wlight,WZqoind}. 
5767: 
5768: \section{ $SU(2)$ slave-boson theory of doped Mott insulators}
5769: 
5770: %The $SU(2)$ slave-boson theory can explain many highly unusual properties of
5771: %high $T_c$ superconductors.  However, 
5772: In order to apply the $SU(2)$ slave-boson theory to high $T_c$ superconductors
5773: we need to first generalize the $SU(2)$ slave-boson to the case with finite
5774: doping.  Then we will discuss how to use the $SU(2)$ slave-boson theory to
5775: explain some of those properties in detail.
5776: 
5777: \subsection{$SU(2)$ slave-boson theory at finite doping}
5778: 
5779: The $SU(2)$ slave-boson theory can be generalized to describe doped spin
5780: liquids (\cit{WLsu2}; \cit{LNNWsu2}).  The generalized  $SU(2)$ slave-boson
5781: theory involves two $SU(2)$ doublets $\psi_{\v i}$ and $h_{\v i} = \left
5782: ({b_{1\v i}\atop b_{2\v i}}\right) $.  Here $b_{1\v i}$ and $b_{2\v i}$ are
5783: two spin-0 boson fields.  The additional boson fields allow us to form $SU(2)$
5784: singlet to represent the electron operator $c_{\v i}$:
5785: \begin{align}
5786: \label{cpsib}
5787: c_{\up\v i} 
5788: =& {1\over \sqrt 2} h^\dagger_{\v i}\psi_{\v i}
5789: = {1\over \sqrt 2} \left (b^\dagger_{1\v i} f_{\up\v i}
5790: + b^\dagger_{2\v i}f^\dagger_{\down\v i}\right ) 
5791: \nonumber\\
5792: c_{\down\v i}  
5793: =& {1\over \sqrt 2} h^\dagger_{\v i} \bar \psi_{\v i}
5794: = {1\over \sqrt 2} \left ( b^\dagger_{1\v i}f_{\down\v i}
5795: - b^\dagger_{2\v i}f^\dagger_{\up\v i}\right ) 
5796: \end{align}
5797: where $\bar \psi = i\tau^2 \psi^*$ which is also an $SU(2)$ doublet.  The
5798: $t$-$J$ Hamiltonian 
5799: \begin{align*}
5800: H_{tJ}=\sum_{\<\v i\v j\>} \left [ J \left (\v S_{\v i}\cdot \v S_{\v j}
5801: - {1\over 4} n_{\v i} n_{\v j} \right ) -t(c^\dagger_{\alpha \v i} c_{\alpha \v j} + h.c.)
5802: \right ]
5803: \end{align*}
5804: can now be written in terms of our fermion-boson fields.  The Hilbert space of
5805: the fermion-boson system is larger than that of the $t$-$J$ model.  However,
5806: the local $SU(2)$ singlets satisfying $\left ( \psi^\dagger_{\v i} \v\tau
5807: \psi_{\v i} + h^\dagger_{\v i} \v\tau h_{\v i} \right ) |{\rm phys} \rangle =
5808: 0$ form a subspace that is identical to the Hilbert space of the $t$-$J$
5809: model.  On a given site, there are only three states that satisfy the above
5810: constraint.  They are $f^\dagger_1 |0\rangle$, $f^\dagger_2 |0\rangle$, and
5811: ${1\over \sqrt 2} \left (b^\dagger_1 + b^\dagger_2 f^\dagger_2 f^\dagger_1
5812: \right ) |0\rangle$ corresponding to a spin up and down electron, and a
5813: vacancy respectively.  Furthermore, the fermion-boson Hamiltonian $H_{tJ}$, as
5814: a $SU(2)$ singlet operator, acts within the subspace, and has same matrix
5815: elements as the $t$-$J$ Hamiltonian.
5816: 
5817:  
5818:  We note that just as in eq.~(\ref{Eq.36}), our treatment of the ${1\over 4}
5819: n_{\v i}n_{\v j}$ term introduces a nearest neighbor boson attraction term which we
5820: shall ignore from now on.\footnote{
5821: %Lee and Salk (2001)
5822: \Ref{LS0101} have introduced a
5823: slightly different formulation where the combination $\left( \v{S}_{\v i}\cdot
5824: \v{S}_{\v j} - {1\over 4} n_{\v i}n_{\v j} \right)$ is written as $
5825: - {1 \over 2} \left | \left( f_{\v i\up}^\dagger f_{\v j\down}^\dagger -
5826:   f_{\v i\down}^\dagger f_{\v j\up}^\dagger \right) \right |^2 (1 - h_{\v i}^\dagger h_{\v i})
5827: (1 - h_{\v j}^\dagger h_{\v j}) $.  The last two factors are the boson projections which
5828: are needed to take care of the case when both sites $i$ and $j$ are occupied
5829: by holes.  While the formulations are equivalent, the mean field phase diagram
5830: is a bit different in that a nearest-neighbor attraction term may lead to
5831: boson pairing.  The competition between boson condensation and boson pairing
5832: needs further studies but we will proceed without the boson interaction term.}
5833: %From Lee & Nagaosa insert A
5834: Now the partition function $Z$ is given by
5835: \begin{equation*}
5836: Z = \int D\psi D\psi^\dagger DhDa^1_0Da^2_0Da^3_0D U\exp
5837: \left(
5838: -\int^\beta_0 d\tau L_2
5839: \right)
5840: \end{equation*}
5841: with the Lagrangian taking the form
5842: \be
5843: \label{128}
5844: L_2 &=& \tilde{J} \sum_{\<\v i\v j\>} \Tr \left[ U_{\v i\v j}^\dagger
5845: U_{\v i\v j} \right] + \tilde{J} 
5846:  \sum_{<\v i\v j>} \left( \psi_{\v i}^\dagger
5847: U_{\v i\v j} \psi_{\v j} +
5848: c.c. \right) \nonumber \\
5849: &+& \sum_{\v i} \psi_{\v i}^\dagger \left(
5850: \partial_\tau - ia_{0\v i}^\ell \tau^\ell \right) \psi_{\v i} \nonumber \\
5851: &+& \sum_{\v i} h_{\v i}^\dagger \left(
5852: \partial_\tau - ia_{0\v i}^\ell \tau^\ell + \mu \right) h_{\v i} \nonumber \\
5853: &-& {1\over 2}\sum_{\<\v i\v j\>} t_{\v i\v j} 
5854: \left(
5855: \psi_{\v i}^\dagger
5856: h_{\v i}h_{\v j}^\dagger
5857:  \psi_{\v j} + c.c.
5858: \right)
5859: \,\,\, . \nonumber \\
5860: \en
5861: %end Lee & Nagaosa insert A
5862: 
5863: Following the standard approach with the choice $\tilde{J} = {3\over 8} J$, we
5864: obtain the following mean-field Hamiltonian
5865: %\cite{UL9453}  
5866: for the fermion-boson system, which is an extension of \Eq{sl2a.8} to the
5867: doped case:
5868: \begin{align}
5869: \label{5}
5870: H_\text{mean} =& \sum_{\<\v i \v j\>} \frac{3}{8}
5871:  J \left[ 
5872: \frac12 \Tr (U_{\v i \v j}^\dag~U_{\v i \v j}) +(\psi^\dag_ i
5873: U_{\v i \v j}
5874: \psi_ j +~h.c.)
5875: \right]  
5876: \nonumber\\
5877: &-{1\over 2}\sum_{\<\v i \v j\>} t (h^\dagger_{\v i} U_{\v i\v j} h_{\v j} + {\rm h.c.} )
5878: \nonumber\\
5879: &
5880: - \mu \sum_{\v i} h^\dagger_{\v i}h_{\v i}  
5881: +\sum_{\v i}  a_0^l (
5882: \psi_{\v i}^\dag \tau^l \psi_{\v i} 
5883: +h_{\v i}^\dag \tau^l h_{\v i} )
5884: \end{align}
5885: The value of the chemical potential $\mu$ is chosen such that the total boson
5886: density (which is also the density of the holes in the $t$-$J$ model) is
5887: \begin{equation*}
5888: \langle h^\dagger_{\v i} h_{\v i}\rangle =
5889: \langle b^\dagger_{1\v i} b_{1\v i} +  b^\dagger_{2\v i} b_{2\v i}\rangle = x.  
5890: \end{equation*}
5891: The values of $a^l_{0}(\v i)$ are chosen such that 
5892: \begin{equation*}
5893: \langle \psi^\dagger_{\v i} \tau^l \psi_{\v i} +
5894: h^\dagger_{\v i} \tau^l h_{\v i}\rangle = 0.
5895: \end{equation*}
5896: For $l = 3$ we have
5897: \begin{equation}
5898: \langle f^\dagger_{\alpha\v i} f_{\alpha\v i} +
5899: b^\dagger_{1\v i} b_{1\v i} - b^\dagger_{2\v i} b_{2\v i}\rangle = 1 \label{6}
5900: \end{equation}
5901: We see that unlike the $U(1)$ slave-boson theory, the density of the fermion
5902: $\langle f^\dagger_{\alpha \v i} f_{\alpha \v i}\rangle$ is not necessarily
5903: equal $1 - x$.  This is because a vacancy in the $t$-$J$ model may be
5904: represented by an empty site with a $b_1$ boson, or a doubly occupied site
5905: with a $b_2$ boson.  
5906: 
5907: %We also notice that the mean-field  Hamiltonian is invariant under local
5908: %$SU(2)$ transformations, $W_{\v i} \in SU(2)$: $\psi_{\alpha \v i} \to W_{\v
5909: %i} \psi_{\alpha \v i}$, $b_{\v i} \to W_{\v i}b_{\v i}$, $U_{\v i\v j} \to W_{\v i}
5910: %U_{\v i\v j}W^\dagger_{\v j}$, and $a^l_{0\v i}\tau^l \to W_{\v i} a^l_{0\v i} \tau^l
5911: %W_{\v i}^\dagger$.
5912: 
5913: \subsection{ The mean-field phase diagram}
5914: 
5915: To obtain the mean-field phase diagram, we have searched the minima of the
5916: mean-field  free energy for the mean-field ansatz with translation, lattice
5917: and spin rotation symmetries.  We find a phase diagram with six different
5918: phases (see Fig. \ref{su2phase}) \cite{WLsu2}.
5919: 
5920: {(1)}  The $d$-wave superconducting (SC) phase is described by the following
5921: mean-field ansatz
5922: \begin{align}
5923: \label{dSC}
5924: U_{\v i, \v i+ \hat x} = & -\chi\tau^3 + \Delta\tau^1, \nonumber\\
5925: U_{\v i, \v i+ \hat y} = & -\chi\tau^3 - \Delta\tau^1,
5926: \nonumber\\
5927: a^3_{0}  \neq & 0,\ \ \ \ \ \
5928: a^{1,2}_{0}  =  0,
5929: \nonumber\\
5930: \langle b_1\rangle \ne & 0 ,\ \ \ \ \ \ 
5931: \langle b_2\rangle = 0.
5932: \end{align}
5933: Notice that the boson condenses in the SC phase despite the fact that in our
5934: mean-field  theory the interactions between the bosons are ignored.  
5935: %The SC
5936: %mean-field  solution provides an interesting example of finite-temperature
5937: %free boson condensation in two dimensions.
5938: %%To understand this phenomenon, 
5939: %To see this, notice that the $a^3_0$ term in the mean-field Hamiltonian
5940: %%\begin{equation}
5941: %$\sum_{\v i} a^3_0 \left ( f^\dagger_{\alpha \v i} f_{\alpha \v i} +
5942: %b^\dagger_{1\v i} b_{1\v i} - b^\dagger_{2\v i} b_{2\v i} - 1 \right ) $
5943: %%\label{11} \end{equation} Thus $a^3_0$ is 
5944: %makes $a^3_0$ behaves like the chemical potential of the fermions.  The
5945: %fermions contribute a term $-(a^3_0 )^2$ to the free energy and favor a
5946: %non-zero $a^3_0$.  Let us assume $a^3_0 > 0$, which makes $\langle
5947: %f^\dagger_{\alpha \v i} f_{\alpha \v i}\rangle < 1$.  
5948: %%From \ref{11} we also see that 
5949: %A positive $a^3_0$ also makes the $b_1$-band bottom to be higher than that of
5950: %$b_2$, and the thermally excited bosons satisfy $ \langle b^\dagger_1
5951: %b_1\rangle_{therm} < \langle b^\dagger_2 b_2\rangle_{therm}$.  Thus the
5952: %thermally excited bosons alone cannot satisfy the constraint in  eq.~(\ref{6}).
5953: %The $b_1$ boson are forced to condense at the bottom of the $b_1$ band to
5954: %satisfy the constraint, in the same way that ordinary boson condense to
5955: %satisfy the density constraint.
5956: %%Despite the fact that the bosons condense into a higher energy state,
5957: %Due to the fermion contribution, the total free energy can still be lowered by
5958: %generating a finite $a^3_0$ at low temperatures.  
5959: In the SC phase, the fermion and boson dispersion are given by $\pm E_f$ and
5960: $\pm E_b-\mu$, where
5961: \begin{align}
5962: \label{EfEb}
5963: E_f = & \sqrt{(\epsilon_f+a_0^3)^2 + \eta^{2}_f},  \nonumber\\
5964: \epsilon_f  =& -{3J\over 4} (\cos k_x + \cos k_y ) \chi,  \nonumber\\ 
5965: \eta_f  =& -{3J\over 4} (\cos k_x - \cos k_y ) \Delta,  \nonumber\\
5966: E_b = & \sqrt{(\epsilon_b+a_0^3)^2 + \eta^{2}_b},  \nonumber\\
5967: \epsilon_b  =& -2t (\cos k_x + \cos k_y ) \chi,  \nonumber\\ 
5968: \eta_b  =& -2t (\cos k_x - \cos k_y ) \Delta.
5969: \end{align} 
5970: 
5971: 
5972: \begin{figure}
5973: \centerline{
5974: \includegraphics[width=3in]{su2phase.eps}
5975: }
5976: \caption{
5977: $SU(2)$ mean-field  phase diagram for $t/J=1$. The phase diagram for $t/J=2$
5978: is quantitatively very similar to the $t/J=1$ phase diagram, when
5979: plotted in terms of the scaled variable $xt/J$, except the
5980: $\pi$fL phase disappears at a lower scaled doping concentration.
5981: We also plotted the Fermi surface, the Fermi arcs, or the Fermi points in some
5982: phases. \cite{WLsu2} 
5983: }
5984: \label{su2phase}
5985: \end{figure}
5986: 
5987: {(2)}  The Fermi liquid (FL) phase is similar to the SC phase except that
5988: there is no fermion pairing ($\Delta = 0$).
5989: 
5990: {(3)} Staggered flux liquid (sfL) phase:
5991: \begin{align}
5992: \label{sFans}
5993: U_{\v i, i + \hat x} & = - \tau^3 \chi - i (-)^{\v i} \Delta ,
5994: \nonumber\\
5995: U_{\v i, i + \hat y} & = - \tau^3 \chi + i (-)^{\v i} \Delta ,
5996: \nonumber\\
5997: a^l_{0}  = & 0, \ \ \ \ \ \ 
5998: \langle b_{1,2}\rangle = 0.
5999: \end{align}
6000: %The flux through each plaquette $\Pi \chi_{\v i\v j} = e^{i\Phi}$ is $\Phi = \pm
6001: %4\arctg \left ( {\Delta\over\chi} \right )$.
6002: The $U$ matrix is the same as that of the staggered flux phase in the $U(1)$
6003: slave-boson theory, which breaks transition symmetry.  Here the breaking of
6004: translational invariance is a gauge artifact.  In fact, a site dependent
6005: $SU(2)$ gauge transformation $W_{\v i} = e^{-i\pi\tau^1/4} e^{-i\pi (i_x + i_y
6006: ) (\tau^1/2+1)}$ maps the sfL ansatz to the $d$-wave pairing ansatz:
6007: \begin{align}
6008: \label{8}
6009: U_{\v i, \v i+ \hat x} = & -\chi\tau^3 + \Delta\tau^1, \nonumber\\
6010: U_{\v i, \v i+ \hat y} = & -\chi\tau^3 - \Delta\tau^1,
6011: \nonumber\\
6012: a^l_{0}  = & 0, \ \ \ \ \ \ 
6013: \langle b_{1,2}\rangle = 0.
6014: \end{align}
6015: which is explicitly translation invariant.  However, the staggered flux
6016: representation of eq.~(\ref{sFans}) is more convenient because the gauge
6017: symmetry is immediately apparent.  Since this $U$ matrix commutes with
6018: $\tau^3$, it is clearly invariant under $\tau^3$ rotation, but not $\tau^1$
6019: and $\tau^2$, and the gauge symmetry has been broken from $SU(2)$ down to
6020: $U(1)$, following the discussion in section X.F.  For this reason we shall
6021: refer to this state as the staggered flux liquid (sfL).
6022: 
6023: In the sfL phase, the fermion and boson dispersion are still given by $\pm
6024: E_f$ and $\pm E_b-\mu$ with $E_f$ and $E_b$ in \Eq{EfEb}, but now $a_0^3=0$.
6025: Since $a^3_0 = 0$ we have $\langle f^\dagger_{\alpha \v i} f_{\alpha \v i}
6026: \rangle = 1$ and $\langle b^\dagger_1b_1\rangle = \langle
6027: b^\dagger_2b_2\rangle = {x\over 2}$. 
6028: 
6029: {(4)}  The $\pi$-flux liquid ($\pi$fL) phase is the same as the sfL phase
6030: except here $\chi = \Delta$.
6031: 
6032: {(5)}  The uniform RVB (uRVB) phase is described by eq. (\ref{sFans}) with
6033: $\Delta = 0$.
6034: 
6035: {(6)}  A localized spin (LS) phase has $U_{\v i\v j} = 0$ and $a^l_{0\v i} = 0$, where
6036: the fermions cannot hop.
6037: % and have zero band width.
6038: 
6039: 
6040: 
6041: %\begin{figure}
6042: %\centerline{
6043: %\includegraphics[width=3in]{su2f2.eps}
6044: %}
6045: %\caption{
6046: %(a) The electron spectral function Im$G_U$ in the sF phase for
6047: % $t/J=2$, $x=0.041$, $T/J=0.13$, where
6048: %$\chi=0.57$ and $\Delta=0.22$.
6049: %The sharp peaks near $\omega=0$ are quasi-particle peaks.
6050: %The insert shows a quarter of the Brillouin
6051: %zone.\hfil\break
6052: %(b)
6053: %The spectral functions Im$G_U$
6054: %for three linear scans along the line A, B, and C in the insert.
6055: %}
6056: %\label{fig2}
6057: %\end{figure}
6058: 
6059: \subsection{Simple properties of the mean-field  phases}
6060: 
6061: %In the following we would like to discuss some simple physical properties of
6062: %the mean-field  phases.  
6063: Note that the topology of the phase diagram is similar to that of $U(1)$ mean
6064: field theory shown in Fig.~\ref{U1}.  The uRVB, sfL, $\pi$fL and LS phases
6065: contain no boson condensation and correspond to unusual metallic states.  
6066: %Since $a^3_0 = 0$, the area of the fermion Fermi sea in the uRVB phase is
6067: %pinned at ${1\over 2}$ of the Brillouin zone.  
6068: As temperature is lowered, the uRVB phase changes into the sfL or $\pi$fL
6069: phases.  A gap is opened at the Fermi surface near $(\pi,0)$ which reduces the
6070: low energy spin excitations.  Thus the sfL and $\pi$fL phases correspond to
6071: the pseudo-gap phase.  
6072: 
6073: The FL phase contains boson condensation.  In this case the electron Green's
6074: function $\<c^\dag c\>=\<(\psi^\dag h)(h^\dag \psi)\>$ is proportional to the
6075: fermion Green's function $\<\psi^\dag\psi\>$.  Thus the electron spectral
6076: function contain $\del$-function peak in the FL phase.  Therefore, the low
6077: energy excitations in the FL phase are described by electron-like
6078: quasiparticles and the FL phase corresponds to a Fermi liquid phase of
6079: electrons. 
6080: 
6081: The SC phase contains both the boson and the fermion-pair condensations and
6082: corresponds to a $d$-wave superconducting state of the electrons.  Just like
6083: the $U(1)$ slave-boson theory, the superfluid density is given by
6084: $\rho_s=\frac{\rho_s^b \rho_s^f}{ \rho_s^b +\rho_s^f}$ where $\rho_s^b$ and
6085: $\rho_s^f$ are the superfluid density of the bosons and the condensed
6086: fermion-pairs, respectively. We see that in the low doping limit, $\rho_s \sim
6087: x$ and one need the condensation of both the bosons and the fermion-pairs to
6088: get a superconducting state.  
6089: 
6090: We would like to point out that the different mean-field  phases contain
6091: different gapless gauge fluctuations at classical level. \ie the gauge groups
6092: for gapless gauge fluctuations are different in different mean-field phases.
6093: The uRVB and the $\pi$fL phases have trivial $SU(2)$ flux and the gapless
6094: gauge fluctuations are $SU(2)$ gauge fluctuations.
6095: %  We can choose a gauge such that $U_{\v i\v j}$ is
6096: %proportional to an identity $2 \times 2$ matrix.  
6097: In the sfL phase, the collinear $SU(2)$ flux break the $SU(2)$ gauge structure
6098: to a $U(1)$ gauge structure.  In this case the gapless gauge fluctuations are
6099: $U(1)$ gauge fluctuations.  In the SC and FL phases, $\langle b_a\rangle \neq
6100: 0$. Since $b_a$ transform as a $SU(2)$ doublet, there is no pure $SU(2)$ gauge
6101: transformation that leave mean-field ansatz $(U_{\v i\v j}, a^l_0, b_a)$
6102: invariant. Thus the invariant gauge group (IGG) is trivial.  As a result, the
6103: $SU(2)$ gauge structure is completely broken and there is no low energy gauge
6104: fluctuations.
6105: 
6106: 
6107: \subsection{Effect of gauge fluctuations: enhanced $(\pi,\pi)$ spin
6108: fluctuations in pseudo-gap phase} \label{asl}
6109: 
6110: The pseudo-gap phase has a very puzzling property which seems hard to explain.
6111: As the doping is lowered, it was found experimentally that both the pseudo-gap
6112: and the antiferromagnetic (AF) spin correlation in the normal state increase.
6113: Naively, one expects the pseudo-gap and the AF correlations to work against
6114: each other. That is the larger the pseudo-gap, the lower the single particle
6115: density of states, the fewer the low energy spin excitations, and the weaker
6116: the AF correlations.
6117: 
6118: It turns out that the gapless $U(1)$ gauge fluctuations present in the sfL
6119: phase play a key role in resolving the above puzzle \cite{KL9930,RWspin}.  Due
6120: to the $U(1)$ gauge fluctuations, the AF spin fluctuations in the sfL phase
6121: are as strong as those of a nested Fermi surface, despite the presence of the
6122: pseudo-gap. 
6123: 
6124: To see how the $U(1)$ gauge fluctuation in the sfL phase enhance the AF spin
6125: fluctuations, we map the lattice effective theory for the sfL state onto a
6126: continuum theory.  In the low doping limit, the bosons do not affect the spin
6127: fluctuations much. So we will ignore the bosons and effectively consider the
6128: undoped case. In the sfL phase, the low energy fermions only appear near $\v
6129: k=(\pm \frac{\pi}{2},\pm \frac{\pi}{2})$ Since the fermion dispersion is
6130: linear near $\v k=(\pm \frac{\pi}{2},\pm \frac{\pi}{2})$, those fermions are
6131: described by massless Dirac fermions in the continuum limit:
6132: \begin{eqnarray}
6133: \label{Dirac}
6134: S&=&\int d^{3}x \sum_\mu \sum_{\al=1}^N\bar{\Psi}_{\al}
6135: v_{\al,\mu} \partial_{\mu}\gamma_{\mu}\Psi_{\al}
6136: \end{eqnarray}
6137: where $v_{\al, 0}=1$ and $N=2$, but in the following we will treat $N$ as
6138: an arbitrary integer, which gives us a large $N$ limit of the sfL state.  In
6139: general $v_{\al, 1}\neq v_{\al,2}$. However, for simplicity we will
6140: assume $v_{\al,i}=1$ here.  The Fermi field $\Psi_{\al}$ is a $4\times
6141: 1$ spinor which describes lattice fermions $f_{\v i}$ with momenta near $(\pm
6142: \pi/2, \pm \pi/2)$.  The $4\times4$ $\gamma_{\mu}$ matrices form a
6143: representation of the
6144: $\{\gamma_{\mu},\gamma_{\nu}\}=2\delta_{\mu\nu}$ ($\mu,\nu = 0,1,2$) and are
6145: taken to be
6146: \begin{eqnarray}
6147: \gamma_{0}&=&\bpm \sigma_{3}&0\\ 0&-\sigma_{3} \epm, \quad
6148: \gamma_{1}=\bpm \sigma_{2}&0\\ 0&-\sigma_{2}   \epm, \\
6149: \gamma_{2}&=&\bpm\sigma_{1}&0\\  0&-\sigma_{1}\epm
6150: \end{eqnarray}
6151: with $\sigma_{\mu}$ the Pauli matrices.  Finally note that
6152: $\bar{\Psi}_{\sigma} \equiv \Psi^{\dag}_{\sigma} \gamma_{0}$.  
6153: 
6154: 
6155: The fermion field $\Psi$ couples to the $U(1)$ gauge field in the sfL phase.
6156: To determine the form of the coupling, we note that the $U(1)$ gauge
6157: transformation takes the following form
6158: \begin{equation*}
6159:  f_{\v i} \to e^{i\th_{\v i}} f_{\v i}
6160: \end{equation*}
6161: if we choose the  ansatz \Eq{sFans} to describe the sfL phase.  By requiring
6162: the $U(1)$ gauge invariance of the continuum model, we find the continuum
6163: Euclidean action to be
6164: \begin{eqnarray}
6165: \label{QED3a}
6166: S&=&\int d^{3}x \sum_\mu \sum_{\sigma=1}^N\bar{\Psi}_{\sigma}
6167: v_{\sigma,\mu} (\partial_{\mu}-ia_{\mu})\gamma_{\mu}\Psi_{\sigma}
6168: \end{eqnarray}
6169: 
6170: The dynamics for the $U(1)$ gauge field arises solely due to the screening by
6171: bosons and fermions, both of which carry gauge charge. In the low doping
6172: limit, however, we will only include the screening by the fermion fields.
6173: After integrating out $\Psi$ in \Eq{QED3a}, we obtain the following effective
6174: action for the $U(1)$ gauge field \cite{KL9930}
6175: \begin{eqnarray}
6176: \cal Z&=&\int Da_{\mu}\exp\Big( -\frac{1}{2}\int\frac{d^3q}{(2\pi)^3}a_{\mu}
6177: (\v{q})\Pi_{\mu\nu}a_{\nu}(-\v{q})\Big) \nonumber \\
6178: \Pi_{\mu\nu}&=&\frac{N}{8}\sqrt{\v{q}^2}\Big(\delta_{\mu\nu}
6179: - \frac{q_{\mu}q_{\nu}}{\v{q}^{2}}\Big)
6180: \label{Pi}
6181: \end{eqnarray}
6182: By simple power counting we can see that the above polarizability makes the
6183: gauge coupling $a_\mu j^\mu$ a marginal perturbation at the free fermion fixed
6184: point.  
6185: %Importantly however we should note that 
6186: Since the conserved current $j^\mu$ cannot have any anomalous dimension, this
6187: interaction is an \emph{exact} marginal perturbation protected by current
6188: conservation.
6189: % which on the quantum level is implemented via the Ward identities.
6190: 
6191: For $N=2$, the spin operator with momenta near $\v q=(0,0)$, $(\pi,\pi)$, and
6192: $(\pi,0)$ has different form when expressed in terms of $\Psi_\al$.  Near $\v
6193: q=(0,0)$ 
6194: \begin{equation*}
6195:  \v S_u(\v x)=\frac12 \bar \Psi_\al \ga^0 \v \si_{\al\bt} \Psi_\bt
6196: \end{equation*}
6197: Near $\v q=(\pi,\pi)$ 
6198: \begin{equation*}
6199:  \v S_s(\v x)=\frac12 \bar \Psi_\al \v \si_{\al\bt} \Psi_\bt
6200: \end{equation*}
6201: Near $\v q=(\pi,0)$ 
6202: \begin{equation*}
6203:  \v S_{(\pi,0)}(\v x)=\frac12 \bar \Psi_\al 
6204: \bpm 0 & \sigma_{1}\\ \sigma_{1} & 0 \epm
6205: \v \si_{\al\bt} \Psi_\bt
6206: \end{equation*}
6207: 
6208: \begin{figure}[tb]
6209: %\epsfxsize=65mm
6210: %\centerline{ \epsfbox{Pol1N.eps} }
6211: \begin{center}
6212: \includegraphics[width=3in]{Fig1.eps}
6213: \caption{Non-zero leading $1/N$ corrections to the staggered spin
6214: correlation function. The ${\v x}$ denotes the vertex which is the $4 \times
6215: 4$ unit matrix in  the case of interest.}
6216: \label{Pol1N}
6217: \end{center}
6218: \end{figure}
6219: 
6220: At the mean-field level, all the above three spin operators have algebraic
6221: correlations $1/r^4$ with decay exponent $4$.  The effect of gauge
6222: fluctuations can be included at $\frac{1}{N}$ order by calculating the
6223: diagrams in Fig. \ref{Pol1N}.  We find that (\cit{RWspin}; \cit{FPS0308}) these three spin
6224: correlaters still have algebraic decays, indicating that the gauge interaction
6225: is indeed marginal.  The decay exponents of the spin correlation near $\v
6226: q=(0,0)$ and $\v q=(\pi,0)$ are not changed and remain to be $4$.  This result
6227: is expected for the spin correlation near $\v q=(0,0)$ since $S_u(\v x)$ is
6228: proportional to the conserved density operator that couple to the $U(1)$ gauge
6229: field.  Therefore $S_u(\v x)$ cannot have anomalous dimension.  
6230: %It is a surprise that
6231: $ \v S_{(\pi,0)}(\v x)$ does not have any anomalous dimension either (at $1/N$
6232: order).  In fact, this result holds to all orders in $1/N$ for the case of
6233: isotropic velocities, due to an $SU(4)$ symmetry 
6234: %(Hermele and Senthil, 2004)
6235: \cite{HS04}.
6236: Thus the spin fluctuations near $(\pi,0)$ is also not enhanced by the gauge
6237: interaction. This may explain why it is so hard to observe any spin
6238: fluctuations near $(\pi,0)$ in experiments.
6239: 
6240: 
6241: $S_s(\v x)$ is found to have a non-zero anomalous dimension.  The spin
6242: correlation near $\v q=(\pi,\pi)$ is found to be $1/r^{4-2\al}$ with
6243: \begin{equation}
6244: \label{eqnu}
6245: \al=\frac{32}{3\pi^2 N}
6246: \end{equation}
6247: In the $\om$-$\v k$ space, the imaginary part of the spin susceptibility near
6248: $(\pi,\pi)$ is given by
6249: \begin{align}
6250: \label{sspin}
6251: & \Im\chi(\omega,\mathbf{q})
6252: \equiv\Im\langle S^{+}(\omega,\mathbf{q}+\v Q)
6253: S^{-}(-\omega,-\mathbf{q}+\v Q) \rangle \nonumber \\ 
6254: =&
6255: \frac{C_s}{2}\sin(2\al\pi)\Gamma(2\al-2)
6256: \Theta(\omega^2-\mathbf{q}^2)\left(\omega^2-\mathbf{q}^2\right)^{1/2-\al} 
6257: \end{align}
6258: where $C_{s}$ is a constant depending on the physics at the lattice scale.
6259: 
6260: \begin{figure}[tb]
6261: \begin{center}
6262: \includegraphics[width=70mm]{Fig2.eps}
6263: \caption{Imaginary part of the spin susceptibility at $(\pi,\pi)$. 
6264: Note the divergence at small $\omega$. (from \Ref{RWspin}}
6265: \label{chi_Q_ASL}
6266: \end{center}
6267: \end{figure}
6268: 
6269: From \Eq{sspin} it is clear that the gauge fluctuations have reduced the
6270: mean-field exponent. If we boldly set $N=2$ which is the physically relevant
6271: case we find $\al =0.54 > 1/2$ which signals the divergence of $\chi(\omega = 0,
6272: q = 0)$.  Thus, after including the gauge fluctuations, the $(\pi,\pi)$ spin
6273: fluctuations are enhanced in the sfL phase despite the pseudo-gap.  In Fig.
6274: \ref{chi_Q_ASL}, we plot the imaginary part of the spin susceptibility at
6275: $(\pi,\pi)$. The $\om$ dependence of the spin susceptibility at $(\pi,\pi)$ is
6276: similar to the one from a nested Fermi surface.
6277: 
6278: The enhancement of the staggered spin correlation follows the trend found in
6279: Gutzwiller projection of the staggered flux (or equivalently the $d$-wave
6280: pairing) state.  
6281: %Ivanov (2000) 
6282: \Ref{I00} and 
6283: %Paramekanti {\em et al.} (2004)
6284: \Ref{PRT0404} reports a
6285: power law decay of the equal time staggered spin correlation function as
6286: $r^{-\nu}$ where $\nu = 1.5$ for the undoped case and $\nu = 2.5$ for
6287: 5\% doping, which are considerably slower than the $r^{-4}$ behavior before
6288: projection.  
6289: %Furthermore the result agrees with the first order term in the 1/N expansion
6290: %of the non-perturbative result obtained in the context of spontaneous chiral
6291: %symmetry breaking in Ref.\onlinecite{Gusynin}. Expanding their result to
6292: %$1/N^2$ allows us to at least estimate the sign and order of magnitude that
6293: %higher order corrections will play and shows in effect that $1/N^2$
6294: %corrections have the appropriate sign to increase the real space exponent
6295: %even further.  bringing the analytic result closer in line with the numerical
6296: %data.  
6297: 
6298: %Let us also remark that although we do not know the exact value of $\nu$,
6299: %many results discussed in this paper remain valid since they are not
6300: %sensitive to the value of $\nu$. Those results mainly depend on two things:
6301: %(A) $\nu$ is close to or bigger than $1/2$, a large change from the
6302: %mean-field exponent , and (B) $\nu$ is irrational. Condition (B) ensures a
6303: %branch-cut in the spin-spin correlation function, which is needed to explain
6304: %experimental data.
6305: 
6306: We remark that with doping, Lorentz invariance is broken by the presence of
6307: bosons. In this case the Fermi velocity receives an logarithmic correction
6308: which enhances the specific heat $\ga$  coefficient and the uniform
6309: susceptibility \cite{KLW9709}.
6310: 
6311: \subsection{Electron spectral function} \label{specG} 
6312: 
6313: 
6314: One of the striking properties of the high $T_c$ superconductor is the
6315: appearance of the pseudo-gap in electron spectral function for underdoped
6316: samples, even in the non-superconducting state. To understand this property
6317: within the $SU(2)$ slave-boson theory, we like to calculate the physical
6318: electron Green function.  Since the non-superconducting state for small $x$ is
6319: described by the sfL phase in the $SU(2)$ slave-boson theory, so we need to
6320: calculate the electron Green function in the sfL phase.  
6321: 
6322: 
6323: \subsubsection{Single
6324: hole spectrum}
6325: 
6326: The electron Green's function is given by
6327: \begin{equation*}
6328:  G_e(\v x)=\<h^\dag(\v x)\psi(\v x) h(0)\psi^\dag(0)\>
6329: \end{equation*}
6330: If we ignore the gauge interactions between the bosons and the fermions, the
6331: electron Green's function can be written as
6332: \begin{equation*}
6333:  G_{e0}=\<h^\dag h\>_0 \<\psi \psi^\dag\>_0
6334: \end{equation*}
6335: where the subscript $0$ indicates that we ignore the gauge fluctuations when
6336: calculating $\<...\>_0$.
6337: 
6338: \begin{figure}[t]
6339: \begin{center}
6340: \includegraphics[scale=0.5]{bsnfrm.eps}
6341: \end{center}
6342: \caption{
6343: The thick line represents the boson world line,
6344: the thin line represents the fermion world line, and the dash line represent
6345: the gauge interaction. The dash-dot line is the straight return path.
6346: The $U(1)$ gauge interaction is caused by the extra phase term
6347: $e^{i\oint d\v x\cdot \v a}$ due to the $U(1)$ flux through the loop
6348: formed by the boson and the fermion world lines. 
6349: Such flux can be approximated by
6350: the flux through the loop
6351: formed by the fermion world line and the straight return path.
6352: }
6353: \label{bsnfrm}
6354: \end{figure}
6355: 
6356: The effect of the $U(1)$ gauge fluctuations is an extra phase term $e^{i\oint
6357: d\v x\cdot \v a}$ determined by the $U(1)$ flux through the loop formed by the
6358: boson and the fermion world lines (see Fig. \ref{bsnfrm}).  Since the fermion
6359: has a linear dispersion relation, the area between the boson and the fermion
6360: world lines is of order $|\v x|^2$, where $|\v x|$ is the separation between
6361: the two points of the Green's function.  Such an area is about the same as the
6362: area between the fermion world line and the straight return path (see Fig.
6363: \ref{bsnfrm}).  So we may approximate the effect of $U(1)$ gauge
6364: fluctuations as the effect caused by the $U(1)$ flux through the fermion world
6365: line and the straight return path \cite{RW0140}. This corresponds to
6366: approximate the electron Green's function as
6367: \begin{equation*}
6368:  G_e(\v x)=\<h^\dag(\v x) h(0)\>_0 
6369: \<\psi(\v x) \psi^\dag(0) e^{i\int_0^{\v x} d\v x \cdot \v a}\>
6370: \end{equation*}
6371: where $\int_0^{\v x}d\v x$ is the integration along the straight return path
6372: and $\<...\>$ includes integrating out the gauge fluctuations.
6373: 
6374: First, let us consider the fermion Green's function.  At the leading
6375: order of a large-$N$ approximation, it was found that \cite{RW0171,RW0140}
6376: \footnote{ 
6377: Note that the usual fermion Green's function $\<\psi(\v x)
6378: \psi^\dag(0)\>$ is not gauge invariant.  As a result, the Green's function is
6379: not well defined and depends on the choices of gauge-fixing conditions
6380: \cite{FT0103,FTV0235,K0206,Y0217}.  If one incorrectly identifies $\<\psi(\v
6381: x) \psi^\dag(0)\>$ as the the electron Green's function, then the electron
6382: Green's function will have different decay exponents for different
6383: gauge-fixing conditions.  In contrast, the combination $\<\psi(\v x)
6384: \psi^\dag(0) e^{i\int_0^{\v x} d\v x \cdot \v a}\>$ is gauge invariant and
6385: well defined.  The resulting electron Green's function does not depend on
6386: gauge-fixing conditions.
6387: }
6388: \begin{equation}
6389: \label{fermG}
6390:  \<\psi(\v x) \psi^\dag(0) e^{i\int_0^{\v x} d\v x \cdot \v a}\>
6391: \propto (\v x^2)^{-(2-\al)/2}  .
6392: \end{equation}
6393: where $\al$ is given in \Eq{eqnu}.  
6394: %If we boldly set $N = 2$, we have
6395: %$\alpha = 0.54$.  
6396: We note that the above becomes the Green's function for free
6397: massless Dirac fermion when $\al=0$.  The finite $\al$ is the effect of gauge
6398: fluctuations.
6399: 
6400: \begin{figure}[tb]
6401: \begin{center}
6402: \includegraphics[scale=0.65]{Fig23.eps}
6403: \end{center}
6404: \caption{
6405: The single hole spectral function at $\left(  {\pi \over 2},{\pi \over 2}    \right)$.  Increasing $\alpha$ corresponds to increasing attraction between fermion and boson due to gauge field fluctuations. (from \Ref{RW0140})
6406: }
6407: \label{specG2}
6408: \end{figure}
6409: 
6410: For a single hole, the boson Green's function is simply that of a classical
6411: particle.  The electron Green's function $G_e(\v{r},\tau)$ is readily
6412: calculated using \Eq{fermG} and its Fourier transform yields the electron
6413: spectral function.  The result at the nodal position $\left( {\pi \over 2},
6414: {\pi \over 2}   \right)$ is shown in Fig.~(\ref{specG2}).  The $\alpha = 0$
6415: curve is the result without gauge fluctuation.  It is the convolution of the
6416: fermion and Bose spectra and is extremely broad.  The gauge field leads to an
6417: effective attraction between the fermion and boson in order to minimize the
6418: gauge flux enclosed by the fermion on boson vortex lines as shown in
6419: Fig.~(\ref{bsnfrm}).  The result is a piling up of a spectral weight at low
6420: energy with increasing $\alpha$.  Still, the one-hole spectrum remains
6421: incoherent, as is appropriate for a deconfined $U(1)$ spin liquid state.  This
6422: calculation can be extended to finite hole density, which requires making
6423: certain assumptions about the boson Green's function 
6424: (\cit{RW0140}; \cit{FT0103}).
6425: Under certain  conditions they obtain power-like type spectral functions
6426: similar to those of the Luttinger liquid.
6427: %(see Fig. \ref{specG2}).
6428: 
6429: \subsubsection{Finite hole density: pseudo-gap and Fermi arcs}
6430: \label{specmean}
6431: 
6432: Here we will consider the mean-field electron Green's function $G_0$
6433: at finite doping.
6434: Using the expression
6435: of $c_\alpha$ in \Eq{cpsib},
6436: % in terms of the fermion $f_\alpha$ and the boson
6437: %$b$ fields,
6438: the mean-field electron Green's function is given by the product of the
6439: fermion and boson Green functions.  So the electron spectral function is a
6440: convolution of the boson spectral function and the fermion spectral function.
6441: 
6442: %\begin{eqnarray}
6443: %\label{Gmean}
6444: %G_0(\v r,\tau) & = & -\frac12 \langle  T_\tau h^\dagger \psi
6445: %(\v r,\tau) \psi^\dagger(0,0) h\rangle \nonumber  \\
6446: %         % & \approx & \frac12 \langle  T_\tau \left(
6447: %         %   \right) \rangle G_F (\v{r}, \tau)
6448: %          &= & \frac12 \sum_{ab} G_B^{ba}(-\v r,-\tau) G_F^{ab} (\v r, \tau)
6449: %\end{eqnarray}
6450: %where
6451: %\begin{eqnarray}
6452: %G_B^{ab}(\v r, \tau) & = &  \langle T_\tau 
6453: %\left( b_a(\v r, \tau)b_b^{\dagger} (0,0) \right) \rangle \nonumber \\
6454: %G_F^{ab}(\v r, \tau) & = & -\langle T_\tau \left(
6455: %\psi_a (\v r, \tau) \psi_b^\dagger(0,0) \right) \rangle
6456: %\end{eqnarray}
6457: %
6458: %
6459: %Let us first calculate the electron Green function in the SC state, which is a
6460: %simpler problem.  The boson Green function contains two parts.  Note that at
6461: %temperature $T$ most bosons are in states which have energies of order $T$
6462: %from the bottom of the boson band.  Thus at high energies the boson Green
6463: %function is given by the single-boson Green function $G_{b}^{s}$ as if no
6464: %other bosons are present.  The imaginary part of this part of boson Green
6465: %function extends the whole band width of the boson band.  
6466: %
6467: %The first part comes from the condensed bosons and is constant.  The second
6468: %part comes from the unoccupied state at higher energies.  This part is given
6469: %by the single-boson Green function $G_B^{s}$ as if no other bosons are
6470: %present.  Thus the boson Green function has a form $x e^{i\v Q_b\cdot \v r} +
6471: %G_B^{s}$ where $\v Q_b$ is the momentum of the bottom of the boson band.  The
6472: %mean field electron Green function in SC phase has a form
6473: %\begin{align}
6474: %\label{G0SC}
6475: %G_0 =& \frac{x}{2} e^{i\v Q_b\cdot \v r}
6476: % G_F^{11} + G_{in} \nonumber\\
6477: %=& {x\over 2} \left ( {u_{\v k}^2\over \omega - E_f} 
6478: %+ {v_{\v k}^2\over \omega + E_f}\right ) + G_{in}
6479: %\end{align}
6480: %where $u$ and $v$ are the coherent factors: 
6481: %\begin{align*}
6482: %u_{\v k} = &\sqrt{{E_f + \epsilon_f\over 2E_f}} {\rm sgn} (\eta_f ),
6483: %\nonumber\\
6484: %v_{\v k} = &\sqrt{{E_f - \epsilon_f\over 2E_f}} .  
6485: %\end{align*}
6486: %The second term $G_{in}$ gives rise to a broad background in the electron
6487: %The second term comes from the convolution of $G_{B}^{s}$ and $G_{F}$ and is
6488: %the incoherent part of the Green function.  The first term is the coherent
6489: %part since its imaginary part is given by discrete $\delta$-functions.  It is
6490: %this coherent part that gives rise to the quasiparticle peaks in the SC phase
6491: %observed in photoemission experiments. 
6492: %We plot the above mean-field
6493: %electron spectral function in Fig. \ref{su2longf1}.  
6494: %The quasiparticle dispersion is given by $\pm E_f$.
6495: %The peak in the electron
6496: %spectral function crosses zero energy at four points at $\v k=(\pm
6497: %(\frac{\pi}{2}-\del), \pm (\frac{\pi}{2}-\del))$.  Thus the SC phase has four
6498: %Fermi points as represented in Fig. \ref{su2phase}.
6499: 
6500: %\begin{figure}
6501: %\includegraphics[width=3in]{su2longf1.eps}
6502: %\caption{
6503: %The electron spectral function for, from top down,
6504: % (a) $k=(-\pi/4, \pi/4) \to (\pi/4, 3\pi/4)$,
6505: %(b) $k=(-\pi/8, \pi/8) \to (3\pi/8, 5\pi/8)$, (c) 
6506: %$k=(0, 0) \to (\pi/2, \pi/2)$, and 
6507: %(d) $k=(0, \pi/2) \to (0,0)$.
6508: %We have chosen $J=1$ and
6509: %$\frac{\Delta}{\chi} = 0.2$.
6510: %}
6511: %\label{su2longf1}
6512: %\end{figure} 
6513: 
6514: %If the boson spectral
6515: %function was a $\del$-function in $\om$ and $\v k$, then $G_0$ would be
6516: %proportional to the fermion Green function $G_F$, In this case the electron
6517: %and the fermion would have the same dispersion given by $\pm E_f(\v k)$.  
6518: 
6519: Let us consider a region of the pseudogap above $T_c$ but at a temperature
6520: which is not too high.  The boson can be considered nearly condensed. The
6521: boson spectral function contain a sharp peak at $\om=0$ and $\v k=0$ and $\v
6522: k=(\pi,\pi)$.  The weight of the peak is of order $x$ and the width is of
6523: order $T$.  At high energies, the boson Green function is given by the
6524: single-boson Green function $G_{b}^{s}$ as if no other bosons are present.  So
6525: the boson spectral function also contain a broad background which extends the
6526: whole band width of the boson band.  The resulting  mean-field electron Green
6527: function has a form (\cit{WLsu2}; \cit{LNNWsu2})
6528: \begin{align}
6529: \label{G0SC}
6530: G_0 
6531: %= & \frac{x}{2} e^{i\v Q_b\cdot \v r}
6532: % G_F^{11} + G_{in} \nonumber\\
6533: =& {x\over 2} \left ( {u_{\v k}^2\over \omega - E_f} 
6534: + {v_{\v k}^2\over \omega + E_f}\right ) + G_{in}
6535: \end{align}
6536: where $u$ and $v$ are the coherent factors: 
6537: \begin{align*}
6538: u_{\v k} = &\sqrt{{E_f + \epsilon_f\over 2E_f}} {\rm sgn} (\eta_f ),
6539: \nonumber\\
6540: v_{\v k} = &\sqrt{{E_f - \epsilon_f\over 2E_f}} .  
6541: \end{align*}
6542: The second term $G_{in}$ gives rise to a broad background in the electron. It
6543: comes from the convolution of the background part of the boson spectral
6544: function  and the fermion spectral function.  The first term is the coherent
6545: part since its imaginary part is a peak of width $T$, which is approximated by
6546: a $\delta$-function here.  The quasiparticle dispersion is given by $\pm E_f$.
6547: The peak in the electron spectral function crosses zero energy at four points
6548: at $\v k=(\pm \frac{\pi}{2}, \pm \frac{\pi}{2})$.  Thus the mean-field sfL
6549: phase has four Fermi points.  Also, in the sfL phase, Im$G_{in}$ is non-zero
6550: only for $\omega<0$ and contributes $1/2$ to a total spectral weight which is
6551: $(1+x)/2$.
6552: 
6553: 
6554: 
6555: 
6556: %The boson spectral function also contain a broad background of width $\sim
6557: %8t$.  The total weight of the background is of order 1.  The electron
6558: %spectral function, as a convolution of the boson spectral function and the
6559: %fermion spectral function, has a sharp peak (of weight $\sim x$ and width
6560: %$\sim T$) at the fermion energy $\om=E_f(\v k)$, as a result of the sharp
6561: %peak in the boson spectral function.  The electron spectral function also
6562: %contain a broad background (of weight $\sim 1$ and width $\sim 8t$) as a
6563: %result of the broad back ground in the boson spectral function.  The detailed
6564: %expression of $G_0$ is lengthy, but it can be approximated at low
6565: %temperatures by \begin{equation}
6566: %\label{13}
6567: %G_0 (k, \omega ) = {x\over 2} \left ( {u_{\v k}^2\over \omega - E_f} 
6568: %+ {v_{\v k}^2\over \omega + E_f}\right ) + G_{in}
6569: %\end{equation}
6570: %The first term gives rise to the sharp peak in the electron spectral function
6571: %and describes the coherent motion of electrons with the fermion dispersion.
6572: %Here $u$ and $v$ are the coherent factors: 
6573: %\begin{align*}
6574: %u_{\v k} = &\sqrt{{E_f + \epsilon_f\over 2E_f}} {\rm sgn} (\eta_f ),
6575: %\nonumber\\
6576: %v_{\v k} = &\sqrt{{E_f - \epsilon_f\over 2E_f}} .  
6577: %\end{align*}
6578: %The second term $G_{in}$ gives rise to a broad background in the electron
6579: %spectral function which mainly reflects the boson density of states.
6580: %Im$G_{in}$ is non-zero only for $\omega<0$ and contributes $1/2$ to a
6581: %total spectral weight which is $(1+x)/2$.
6582: 
6583: %The peak in the electron spectral function from the coherent part of $G_0$
6584: %correspond to electron-like quasiparticles.  
6585: 
6586: From the dispersion relation of the peak $\om=E_f(\v k)$ and the fact that
6587: Im$G_{in}\approx 0$ when $\omega<-E_f(\v k)$, we find that the electron
6588: spectral function contain the gap of order $\Del$ at $(0,\pi)$ and $(\pi,0)$
6589: even in the non-superconducting state.  So the mean-field electron spectral
6590: function of the $SU(2)$ slave-boson theory can explain the pseudo-gap in the
6591: underdoped samples.  However, if we examine the mean-field electron spectral
6592: function more closely, we see that the Fermi surface of the quasiparticles is
6593: just four isolated points $(\pm \pi/2,\pm \pi/2)$.  This property does not
6594: agree with experiments. 
6595: 
6596: \begin{figure}
6597: \includegraphics[width=1.5in]{su2longf2.eps}
6598: %\centerline{\epsfile{file=su2longf2.eps,width=100mm,height=30mm}}
6599: \caption{
6600: A diagram for renormalized electron Green function.
6601: The solid (dash) line is the fermion (boson) propagator.
6602: }
6603: \label{su2longf2}
6604: \end{figure}
6605: 
6606: 
6607: In reality there is a strong attraction between the boson and the fermions due
6608: to the fluctuation around the mean-field state.  The dominant effect comes
6609: from the gauge fluctuations which attempt to bind the bosons and the fermions
6610: into electrons.  This corresponds to an effective attraction between the
6611: bosons and the fermions.  In the case of a single hole, the interaction with
6612: gauge fields can be treated as discussed in the last section.  Here we proceed
6613: more phenomenologically.  One way to include this effect is to use the diagram
6614: in Fig. \ref{su2longf2} to approximate the electron Green function, which
6615: corresponds to an effective short range interaction of form
6616: \begin{equation}
6617: -\frac{V}{2} (\psi^\dagger h) (h^\dagger \psi )=-V  c^\dagger c
6618: \end{equation}
6619: with $V < 0$.  We get
6620: \begin{equation}
6621: G = \frac{1}{\left( G_0 \right)^{-1} + V}
6622: \label{GV}
6623: \end{equation}
6624: The first contribution to $V$ come from the fluctuations of $a_{0}^{\ell}$
6625: which induces the following interaction between the fermions and the bosons:
6626: \begin{equation}
6627:  \psi^{\dagger} \v{\tau} \psi \cdot h^{\dagger}
6628: \v{\tau}h
6629: \end{equation}
6630: The second one (whose importance was pointed out by 
6631: %Laughlin 
6632: \Ref{L9527}) is
6633: the fluctuations of $|\chi_{\v i\v j}|$ which induces
6634: \begin{equation}
6635: -t (\psi^\dagger h)_{\v j} (h^\dagger \psi )_{\v i} =-2t c^\dagger_{\v j} c_{\v i}
6636: \end{equation}
6637: %The numerical calculations indicate that the simple formula eq. \ref{GV} fits
6638: %well with more complicated calculations if we choose
6639: This is nothing but the original hopping term. We expect the coefficient $t$
6640: to be reduced due to screening, but in the following we adopt the form
6641: \begin{equation}
6642: V(\v k)=U + 2t(\cos k_x +\cos k_y)
6643: \end{equation}
6644: for $V$ in eq. (\ref{GV}). 
6645: %Here the first and the second term comes from the first and the second kind
6646: %of fluctuations.  
6647: In Fig. \ref{su2longf3} and \ref{su2longf3a} we plot the electron spectral
6648: function calculated from eq. (\ref{GV}) (\cit{WLsu2}; \cit{LNNWsu2}).
6649: 
6650: 
6651: \begin{figure}
6652: \includegraphics[width=3in]{su2longf3.eps}
6653: %\centerline{\epsfile{file=su2longf3.eps,width=100mm,height=100mm}}
6654: \caption{
6655: The electron spectral function for, from top down, 
6656: (a) $k=(-\pi/4, \pi/4) \to (\pi/4, 3\pi/4)$,
6657: (b) $k=(-\pi/8, \pi/8) \to (3\pi/8, 5\pi/8)$, 
6658: (c) $k=(0, 0) \to (\pi/2, \pi/2)$, and
6659: (d) $k=(0, \pi) \to (0,0)$.
6660: We have chosen $J=1$.
6661: The paths of the four momentum scans are shown in Fig. \ref{su2longf7}.
6662: }
6663: \label{su2longf3}
6664: \end{figure} 
6665:  
6666: \begin{figure}
6667: \includegraphics[width=2in]{su2longf7.eps}
6668: \caption{
6669: The solid line a, b, c, and d are paths of the four momentum scans
6670: in Fig. \ref{su2longf3}.
6671: The solid curves are schematic representation of
6672: the Fermi segments where the quasiparticle peak
6673: crosses the zero energy.
6674: }
6675: \label{su2longf7}
6676: \end{figure}  
6677: 
6678: \begin{figure}
6679: \includegraphics[width=3in]{su2longf3a.eps}
6680: \caption{
6681: The points describe the dispersion of the quasi-particle peaks for the
6682: s-flux liquid phase in Fig. \ref{su2longf3a}.  The vertical bars are proportional
6683: to the peak values of Im$G_U$ which are proportional to
6684: the quasi-particle weight.
6685: }
6686: \label{su2longf3a}
6687: \end{figure} 
6688: 
6689: 
6690: 
6691: We have chosen 
6692: %  $\tilde J=J/2$, 
6693: $t=2J$, $\chi = 1$, $\Delta/\chi = 0.4$, $x = 0.1$, and $T = 0.1 J$. 
6694: %Here we choose $\frac{\Delta}{\chi} = 0.4$ so that the renormalized gap near
6695: %$(0, \pi)$ is about $0.4 J$.
6696: The value of $U$ is determined from requiring the renormalized electron Green
6697: function to satisfies the sum rule
6698: \begin{equation}
6699: \int_0^\infty \frac{d\omega }{2\pi}\int \frac{d^2k}{(2\pi)^2}
6700:  \Im G= x
6701: \end{equation}
6702: %Note that the mean field electron Green function does not satisfy this sum
6703: %rule, and instead obeys \begin{equation}
6704: %\int_0^\infty \frac{d\omega }{2\pi}\int \frac{d^2k}{(2\pi)^2}
6705:  % \Im G_0= x/4
6706: %\end{equation}
6707: 
6708: We find that the gap near $(0, \pm \pi)$ and $(\pm \pi,0)$ survives the
6709: binding potential $V(\v k)$.
6710: %inclusion of gauge and $|\chi_{\v i\v j}|$ fluctuations.  
6711: However spectral functions near $(\pm \frac{\pi}{2}, \pm \frac{\pi}{2})$ are
6712: modified.  The Fermi point at $(\frac{\pi}{2}, \frac{\pi}{2})$ for the  mean
6713: field electron Green function $G_0$ is stretched into a Fermi segment as shown
6714: in Fig.  \ref{su2longf7}.  As we approach the uRBV phase, $\Delta$ decreases
6715: and the Fermi arcs are elongated.  Eventually the arcs join together to form a
6716: large closed Fermi surface.
6717: 
6718: While the phenomenological binding picture successfully produces Fermi arcs,
6719: the results are not as satisfactory for the anti-nodal points.  While an
6720: energy gap is produced near $(0,\pi)$, the theory gives a rather sharp
6721: structure at the gap and we see from Fig. \ref{su2longf3a} that the gap above
6722: and below the Fermi energy is not symmetric.  
6723: 
6724: This exposed a serious weakness of the slave-boson gauge theory approach. With
6725: finite hole density, the bosons tend to condense at the mean field level. In
6726: reality the holes are strongly coupled to gauge fluctuations which tend to
6727: suppress the Bose condensation. While the fermions are also coupled to gauge
6728: fields, the Fermi statistics allow us to approach the problem perturbatively
6729: by introducing artificial expansion parameters such as $\frac{1}{N}$ . In
6730: contrast, the problem of bosons coupled to gauge fields is much less
6731: understood. Furthermore the gauge fields mediate strong attraction between
6732: fermions and bosons and, in the case of $SU(2)$ theory, between $b_1$ and $b_2$
6733: bosons which carry opposite gauge charges. In the phenomenological approach
6734: outlined above, the bosons are treated as almost condensed (\ie a narrow peak
6735: in the spectral function is assumed) and bind with a fermion. The assumption
6736: of ``almost Bose condensation'' leads to sharp hole spectra at both the nodal
6737: and anti-nodal points and the latter disagrees with experiment. 
6738: Furthermore it can be shown that the assumption of Bose condensation leads to
6739: a decoupling of the electron to the electromagnetic field, and as a result, the
6740: current carried by the quasiparticles $j = e\frac{dE_{\v k}(\v A)}{d\v A}$ 
6741: is strongly reduced from $ev_F$ which disagrees with experiments
6742: (see IX.B). 
6743: 
6744: \Ref{WL9893} took a first step towards addressing this problem by assuming
6745: that the binding between the bosons and the fermions and/or between the
6746: $b_1$ and $b_2$ bosons prevents single-boson
6747: condensation.  The superconducting state characterized by $\< cc\>\neq 0$
6748: contains only boson pair condensation, \ie $\< b_1b_2 \> \neq 0$ while
6749: $\<b\>=0$. They show that with this assumption the quasiparticle current can
6750: be a finite fraction of $ev_F$ , \ie the $\al$ parameter in \Eq{Eq.6} does
6751: not have to go as $x$. The competition between fermion-boson binding,
6752: boson-boson binding and Bose condensation is a complicated problem which is
6753: still poorly understood at present.
6754: 
6755: STM experiment reveals a rather
6756: broad structure for both particle and hole excitations 
6757: %(Hanaguri {\em et al.}, 2004) 
6758: \cite{HLK04}
6759: and ARPES measurements, which can measure only the occupied states, show
6760: a reduction of the density of states over a broad energy range 
6761: %(Ronning {\em et al.}, 2003)
6762: \cite{RSK0301}.  These lineshapes are more reminiscent of those shown in Fig.
6763: \ref{specG2} for intermediate $\alpha$.  It appears that the assumption of
6764: almost Bose condensation and simple binding via a
6765: short range potential do not capture the subtlety of gauge fluctuation
6766: effects near $(0,\pi)$ where fermions and bosons appear to be closer to being
6767: deconfined.  This dichotomy between nodal and anti-nodal electronic structure
6768: is an important issue which remains open for further theoretical work.
6769:  
6770: 
6771: %We would like to point out that the electron Green function obtained here
6772: %does not show any ``shadow band'' at $\omega=0$, {\it i.e.} Im$G_{e}(0, k)$
6773: %does not have any peak outside of the $(0,\pi)$ -- $(\pi,0)$ line as the
6774: %mirror image of the peaks that appear inside of the  $(0,\pi)$ -- $(\pi,0)$
6775: %line.
6776: %
6777: %In the limit of a single hole, this attraction can lead to a bound state with
6778: %the quantum number of a electron, as emphasized in Ref.  \cite{14}. In the
6779: %case of finite hole concentration, we expect that fermion particle-hole pairs
6780: %may be spontaneously excited out of the mean-field ground state so that the
6781: %$b_2$ ($b_1$) bosons can  bind to fermions (anti-fermions).  The result would
6782: %be low lying physical electron excitations which may resemble a Fermi
6783: %surface. In order to capture this physics, even at a very crude level, we
6784: %assume that after screening, the $a_0^l$ fluctuations induce the following
6785: %short range interaction $- \displaystyle\sum_i U \psi^\dagger_{\alpha i}
6786: %\tau^l \psi_{\alpha i}\ b^\dagger_i\tau^l b_i$, where $U < 0$.  We have
6787: %calculated $G_U$ which included this attraction in the Bethe-Salpeter
6788: %approximation. The results are shown in Fig. \ref{fig2} and \ref{fig3}.  To
6789: %interpret these results, we consider instead $G'_U=(G_0^{-1}+U)^{-1}$ which
6790: %we have found numerically to be nearly identical to $G_U$. It is easy to see
6791: %that $G'_U$ also has the form $G'_U=G'_{coh} +G'_{in}$. In the case when
6792: %either $|u|$ or $|v|\approx 1$, we find that $G'_{coh}={Z\over \omega \pm
6793: %E_f+\mu_f}$ with $\mu_f={xU\over 2 (1 + U G_{in})}$ and $Z={x/2\over ( 1 +
6794: %UG_{in} )^2}$ We see that a negative $U$ generates a negative $\mu_f$ which
6795: %produces small hole pockets.  A negative $U$ also enhance the spectral weight
6796: %of unoccupied quasi-particles. This allows us to choose $U$ by requiring that
6797: %the occupied spectral weight in $G_U$ is $1-x$. For general $u$ and $v$, we
6798: %find that the coherent part of $G'_U$ produce pocket-like Fermi surfaces near
6799: %$\Sigma$ (see Fig \ref{fig2}) which is determined by
6800: %$2E_f(1+UG_{in})=Ux(u^2-v^2)$. The quasi-particle weight at the Fermi surface
6801: %$Z = {2E^{2}_f\over xU^2}$ vanishes at $\Sigma$ and is very small on the
6802: %outer edge of the pocket, making it hard to detect.  
6803: %
6804: 
6805: %, with low lying excitations near $M$ and a shape which resembles the
6806: %experiment.\cite{1} A more detailed discussion on how gauge fluctuation
6807: %affect the quasiparticle line shape can be found in
6808: %\Ref{RW0171,FT0103,RW0140}. 
6809: 
6810: 
6811: 
6812: 
6813: \subsection{Stability of algebraic spin liquids} \label{screen}
6814: 
6815: %The calculation in the last subsection implies that 
6816: The sfL mean-field ansatz leads to an gapless spin liquid. We will call this
6817: the $U(1)$ spin liquid, and it is an example of a class which we call
6818: algebraic spin liquid (ASL) since all the spin correlations have algebraic
6819: decay.  We would like to stress that the ASL is a phase of matter, not a
6820: critical point at a phase transition between two phases.
6821: 
6822: ASL has a striking property: its low energy excitations interact with each
6823: other even down to zero energy. This can be seen from the correlation
6824: functions at low energies which always contain branch cut without any poles.
6825: The lack of poles implies that we cannot use free bosonic or free fermionic
6826: quasiparticles to describe the low energy excitations.  For all other commonly
6827: know gapless states, such as solids, superfluids, Fermi liquids, \etc, the
6828: gapless excitations are always described by free bosons or free fermions. The
6829: only exception is the 1D Luttinger liquid.  Thus the ASL can be view as an
6830: example of Luttinger liquids beyond one dimension.
6831: 
6832: We know that interactions tend to open up energy gaps.  From this point of
6833: view, one might have thought that the only self consistent gapless excitations
6834: are the ones  described by free quasiparticles. Knowing the gapless
6835: excitations in the ASL interact down to zero energy, we may wonder does ASL
6836: really exist?  Have we overlooked some effects which open up energy gap and
6837: make ASL unstable?
6838: 
6839: Indeed, in the above calculation, we have overlooked two effects. Both of them
6840: can potentially destabilize the ASL.  First, the self-energy in Fig.
6841: \ref{Pol1N}A,B, contains a cut-off dependent term which gives the fermion
6842: $\Psi$ a cut-off dependent mass $m(\La)\bar \Psi\Psi$. In the above
6843: calculation, we have dropped such a term. If such a cut-off-dependent term was
6844: kept, the fermions would gain a mass which would destabilize the ASL. 
6845: 
6846: Second, we have overlooked the effects of instantons described by the
6847: space-time monopoles of the $U(1)$ gauge field.  After integrating out the
6848: massless fermions, the effective action of the $U(1)$ gauge field has a form
6849: \Eq{LaNL}.  Unlike the Maxwell term discussed in section IX.D, which produced
6850: a $1/r$ potential, in this case the interaction of the  space-time monopoles
6851: is described by a $\log (r)$ potential.  That is the action of the pair of
6852: space-time monopoles separated by a distance $r$ is given by $ C\log (r)$.
6853: Just like the Coulomb gas in 2D, if the coefficient $C$ is larger than 6, then
6854: the instanton effect is an irrelevant perturbation and the inclusion of the
6855: instantons will not destabilize the ASL \cite{IL8988}. If the coefficient $C$
6856: is less then 6, then the instanton effect is a relevant perturbation and the
6857: inclusion of the instantons will destabilize the ASL.
6858: 
6859: Recently, it was argued in \Ref{HS0301} and \Ref{HSS0310} that the instanton
6860: effect always represent a relevant perturbation due to a screening effect of
6861: 3D Coulomb gas, regardless the value of $C$.  This led to a conclusion in
6862: \Ref{HS0301} that the ASL described by the sfL state does not exist.  The easiest way
6863: to understand the screening effect of the 3D Coulomb gas is to note that the
6864: partition function of the Coulomb gas can be written as a path integral
6865: \begin{align}
6866:  &\int \prod d^3\v x_i e^{-C \sum q_i q_j\log |\v x_i-\v x_j|}
6867: \nonumber\\
6868: =&\int \cD\phi e^{-\int d^3\v x \frac{2\pi}{C}
6869: \prt \phi \sqrt{-\prt^2} \prt \phi -g \cos(\phi)}
6870: \label{monopole}
6871: \end{align}
6872: If we integrate out short distance fluctuations of $\phi$, a counter term
6873: $K(\prt \phi)^2$ can be generated. The counter term changes the long distance
6874: interaction of the space-time monopoles from $\log (r)$ to $1/r$.  The
6875: space-time monopoles with $1/r$ interaction always represent a relevant
6876: perturbation, which will destabilize the ASL.  Physically, the change of the
6877: interaction from $\log (r)$ to $1/r$ is due to the screening effect of
6878: monopole-anti-monopole pairs. Thus the counter term $K(\prt \phi)^2$
6879: represents the screening effect.
6880: 
6881: 
6882: The issue of the stability of the ASL has been examined by \Ref{Wqoslpub},
6883: \Ref{RWspin} and more carefully by 
6884: %Hermele {\em et al.}, 2004.
6885: \Ref{HSF0451} using an argument based on PSG. They came to the conclusion that the $U(1)$ spin liquid is
6886: stable for large enough $N$, if the $SU(2)$ spin symmetry is generalized to
6887: $SU(N)$.  They showed that there is no relevant operator which can destabilize
6888: the deconfined fixed point which consists of $2N$ two-component Dirac fermions
6889: coupled to {\em noncompact} $U(1)$ gauge fields, for $N$ sufficiently large.
6890: \Ref{HSF0451} also pointed out the fallacy of the monopole screening argument.
6891: We summarize some of the salient points below.
6892: 
6893: The operators which perturb the noncompact fixed point can be classified into
6894: two types, those which preserve the flux and those which change the flux by
6895: $2\pi$.  The latter are instanton creation operators which restore the
6896: compactness of the gauge field.  Among the first type there are four fermion
6897: terms which are readily seen to be irrelevant, but as mentioned earlier, the
6898: dangerous term is the quadratic fermion mass term.  The important point is
6899: that the mass terms are forbidden by the special symmetry described by PSG.
6900: The discrete symmetry (such as translation and rotation) of the sfL state
6901: defined on the lattice imposes certain symmetry on the continuum Dirac field
6902: which forbids the mass term.  Another way of seeing this is that after
6903: integrating out the short distance fluctuations, if a mass term is generated
6904: it can be described in the lattice model as a deformation of the mean-field
6905: ansatz $\del U_{\v i\v j}$. Since the short distance fluctuations are
6906: perturbative in nature, the deformation $\del U_{\v i\v j}$ cannot change the
6907: symmetry of the ansatz $\bar  U_{\v i\v j}$ that describe the ground state,
6908: \ie if $\bar  U_{\v i\v j}$ is invariant under a PSG, $\del U_{\v i\v j}$ must
6909: be invariant under the same PSG. One can show that for all the possible
6910: deformations that are invariant under the sfL PSG described by \Eq{GssF}, none
6911: of them can generate the mass term for the fermions. Thus the masslessness of
6912: the fermions are protected by the sfL PSG.
6913: 
6914: As for the second type of operators which change the flux, 
6915: %Hermele {\em et al.}, 2004 
6916: \Ref{HSF0451} appeal to a result in conformal field theory which relates the
6917: scaling dimension of such operators to the eigenvalues of states on a sphere
6918: with a magnetic flux through the surface 
6919: %(Borokhov {\em et al.}, 2002)
6920: \cite{BKW0249}.  This
6921: is easily bound by the ground state energy of $2N$ component Dirac fermions on
6922: the sphere which clearly scale as $N$.  Thus the creation of instantons is
6923: also irrelevant for sufficiently large $N$.
6924: 
6925: As far as the monopole screening argument goes, the fallacy is that in that
6926: argument the fermions are first integrated out completely in order to derive
6927: an effective action for the field $\phi$ shown in eq.~\ref{monopole}.  Then
6928: renormalization group arguments generate a $K(\partial \phi)^2$ term.  However
6929: implicit in this procedure is the assumption that the fermions are rapidly
6930: varying variables compared with the monopoles.  The fact that the fermions are
6931: gapless makes this procedure unreliable.  (One could say that the screening
6932: argument implicitly assumes mass generation for the fermions.)  A better
6933: approach is to renormalize the monopoles and the fermions on the same footing,
6934: \ie let the infrared cutoff length scale for the fermion $(L_f)$ and the
6935: monopoles $(L_m)$ to approach infinity with a fixed ratio, e.g.  $L_f/L_m =
6936: 1$.  In this case integrating out the fermions down to scale $L_f$
6937: will produce an effective action for the $U(1)$ gauge field of the form
6938: 
6939: \begin{equation*}
6940: {g(L_f) \over 16 \pi} f_{\mu v}f^{\mu v}
6941: \end{equation*}
6942: where the running coupling constant $g(L_f) \sim L_f$.  This in turn generates an
6943: interaction between two monopoles separated by a distance $r$ which is of
6944: order $g(L_f)/r$.  To calculate such an interaction, we should integrate out all
6945: the fermions with wavelength less than $r$.  We find the interaction to be
6946: ${g(r) \over r} \sim r^0$, indicating a logarithmic interaction between
6947: monopoles.   Thus the logarithmic interaction is constantly being rejuvenated
6948: and cannot be screened.  This can be cast into a normalization group language
6949: and we can see that the flow equation for the coupling constant $g$ is
6950: modified from the form used by \cite{HS0301}.  The extra term leads to the
6951: conclusion that the instanton fugacity scales to zero and the instanton
6952: becomes irrelevant for $N$ larger than a certain critical value.  
6953: 
6954: To summarize, the ASL derived from the sfL ansatz contain a quantum order
6955: characterized by the sfL PSG \Eq{GssF}. The sfL PSG forbids the mass term of the
6956: fermions. To capture such an effect, we \emph{must} drop the mass term in the
6957: self-energy in our calculation in the continuum model \cite{RWspin}. Ignoring
6958: the mass term is a way to include the effects of PSG into the continuum model.
6959: Similarly,  we must ignore the screening effect described by the $K(\prt
6960: \phi)^2$ term when we consider instantons.  We are then assured that instanton
6961: effects are irrelevant in the large $N$ limits.  So the ASL exists and is
6962: stable at least in the large $N$ limit.  The interacting gapless excitations
6963: in the ASL are protected by the sfL PSG.  It is well known that the symmetry
6964: can protect gapless Nambu-Goldstone modes.  The above example shows that the
6965: PSG and the associated quantum order can also protect gapless excitations
6966: \cite{Wqoslpub,WZqoind,RWspin}.
6967: 
6968: 
6969: %\subsection{Temperature dependence of the superfluid density}
6970: 
6971: \section{Application of gauge theory to the high $T_c$ superconductivity
6972: problem}
6973: 
6974: 
6975: Now we summarize how the gauge theory concepts we have described may be
6976: applied to the high $T_c$ problem.  The central observation is that high $T_c$
6977: superconductivity emerges upon doping a Mott insulator.  The antiferromagnetic
6978: order of the Mott insulator disappears rather rapidly and is replaced by the
6979: superconducting ground state.  The ``normal'' state above the superconducting
6980: transition temperature exhibits many unusual properties which we refer to as
6981: pseudogap behavior.  How does one describe the simultaneous suppression of
6982: N\'{e}el order and the emergence of the pseudogap and the superconductor from
6983: the Mott insulator? The approach we take is to first understand the nature of
6984: a possible nonmagnetic Mott state at zero doping, the spin liquid state, which
6985: naturally becomes a singlet superconductor when doped.  This is the central
6986: idea behind the RVB proposal \cite{A8796} and is summarized in Fig.
6987: \ref{sldsc1}.  The idea is that doping effectively frustrates the N\'{e}el
6988: order so that the system is pushed across the transition where the N\'{e}el
6989: order is lost.  In the real system the loss of N\'{e}el order may proceed
6990: through complicated states, such as incommensurate charge and spin order,
6991: stripes or inhomogeneous charge segregation 
6992: %(Carlson { \it et al.}, 2003)
6993: \cite{CEK03}.
6994: However, in this direct approach the connection with superconductivity is not
6995: al all clear.  Instead it is conceptually useful to arrive at the
6996: superconducting state via a different path, starting from a spin liquid state.
6997: Recently, 
6998: %Senthil and Lee (2004)
6999: \Ref{SL0466} have elaborated upon this point of view which
7000: we summarize below.
7001: 
7002: %Insert from Senthil and Lee
7003: \begin{figure}
7004: %\includegraphics[width=2.3in]{sldsc_a.eps}
7005: %\vskip 0.1in
7006: %\includegraphics[width=2.3in]{sldsc_b.eps}
7007: \includegraphics[width=2in]{zeroTphs.eps}
7008: \caption{
7009: a) Schematic zero temperature phase diagram showing the route between the
7010: antiferromagnetic Mott insulator and the $d$-wave superconductor.  The
7011: vertical axis is labeled by a parameter $g$ which may be taken as a measure
7012: of the frustration in the interaction between the spins in the Mott insulator.
7013: AF represents the antiferromagnetically ordered state. SL is a spin liquid
7014: insulator that could potentially be reached by increasing the frustration.
7015: The path taken by the cuprate materials as a function of doping $x$ is shown
7016: in a thick  dashed-dot line. The question marks represent regions where the
7017: physics is not clear at present. Doping the spin liquid naturally leads to the
7018: dSC state. The idea behind the spin liquid approach is to regard the
7019: superconducting system at non-zero $x$  as resulting from doping the spin
7020: liquid as shown in the solid line, though this is not the path actually taken
7021: by the material.  b) Same as in Fig. \protect\ref{sldsc1}(a) but as a function
7022: of chemical potential rather than hole doping.
7023: }
7024: \label{sldsc1}
7025: \end{figure}  
7026: 
7027: 
7028: \subsection{Spin liquid, quantum critical point and the pseudogap} It is
7029: instructive to consider the phase diagram as a function of the chemical
7030: potential rather than the hole doping as shown in Fig. \ref{sldsc1}(b). 
7031: 
7032: Consider any spin liquid Mott state that when doped leads to a $d$-wave
7033: superconductor. As a function of chemical potential, there will then be a 
7034: zero temperature phase transition where the holes first enter the system. For
7035: concreteness we will simply refer to this 
7036: as the Mott transition.
7037: % - the corresponding phase boundary is marked in red in Fig \ref{sldsc1}(b). 
7038: The associated quantum critical fixed point will control 
7039: the physics in a finite non-zero range of parameters. The various crossovers
7040: expected near such transitions are well-known and are shown in Fig.
7041: \ref{mqc}.
7042: 
7043: Sufficiently close to this zero temperature critical point many aspects of the
7044: physics will be universal. The regime in which such universal behavior 
7045: is observed will be limited by `cut-offs' determined by microscopic
7046: parameters. In particular we may expect that the cutoff scale is provided by
7047: an energy of 
7048: a fraction of $J$ (the exchange energy for the spins in the Mott insulator).
7049: We note that this corresponds to a reasonably high temperature scale. 
7050: 
7051: 
7052: 
7053: Now consider an underdoped cuprate material at fixed doping $x$. Upon
7054: increasing the temperature this will follow a path in Fig.\ref{mqc} that is
7055: shown schematically. 
7056: The properties of the system along this path may be usefully discussed in
7057: terms of the various crossover regimes.  In particular it is clear that the
7058: `normal' state above the superconducting transition is to be understood
7059: directly as the finite temperature `quantum critical'
7060: region associated with the Mott transition. Empirically this region
7061: corresponds to the pseudogap regime. Thus our assertion is that the pseudogap
7062: regime 
7063: is controlled by the unstable zero temperature fixed point associated with the
7064: (Mott) transition to a Mott insulator.
7065: 
7066: 
7067: 
7068: 
7069: 
7070: \begin{figure}
7071: \includegraphics[width=2.4in]{mqc.eps}
7072: \caption{Schematic phase diagram for a doping induced Mott transition between
7073: a spin liquid insulator and a $d$-wave superconductor.  The bold dot-dashed
7074: line is the path taken by a system at hole density $x$ that has a
7075: superconducting ground state. The region marked FS represents the fluctuation
7076: regime of the superconducting transition. The region marked QC is the quantum
7077: critical region associated with the Mott critical point. This region may be
7078: identified with the high temperature pseudogap phase in the experiments. }
7079: \label{mqc}
7080: \end{figure}   
7081: 
7082: %end of insert from Senthil and Lee
7083: 
7084: What are the candidates for the spin liquid phase?  There have been several
7085: proposals in the literature.  One proposal is the dimer phase  
7086: %(Sachdev, 2003)
7087: \cite{S0313}.  Strictly speaking, this is a valence bond solid and not a spin liquid:
7088: it is a singlet state which breaks translational symmetry.  It has been shown
7089: by 
7090: %Read and Sachdev (1990)
7091: \Ref{RS9068} that within the large $N$ Schwinger boson approach
7092: the dimer phase emerges upon disordering the N\'{e}el state.  Sachdev and
7093: collaborators have shown that doping the dimer state produces a $d$-wave
7094: superconductor \cite{VS9916}.  However, such a superconductor also inherits
7095: the dimer order and has a full gap to spin excitations, at least for low
7096: doping.  As we have seen in this review, there are strong empirical evidence
7097: for gapless nodal quasiparticles in the superconducting state.  In our view,
7098: it is more natural to start with translation invariant spin liquid states
7099: which produce $d$-wave superconductors with nodal quasiparticles when doped.
7100: 
7101: We see from Section X that the spin liquid states are rather exotic beasts in
7102: that their excitations are conveniently described in terms of fractionalized
7103: spin 1/2 ``spinon'' degrees of freedom.  We discussed in Section X.G that spin
7104: liquids are characterized by their low energy gauge group.  Among spin liquids
7105: with nodal fermionic spinons, two versions, the $Z_2$ and the $U(1)$ spin
7106: liquids have bee proposed.  The $Z_2$ gauge theory was advocated by
7107: \cite{SF0050}.  It can be considered as growing out of the fermion pairing
7108: phase of the $U(1)$ mean field phase diagram shown in Fig.~\ref{U1}.  The
7109: pairing of fermions $\Delta_{\v i\v j} = \langle   f_{\v i\up} f_{\v i\down} - f_{\v i\down}
7110: f_{\v i\up}     \rangle$ breaks the $U(1)$ gauge symmetry down to $Z_2$, \ie
7111: only $f \rightarrow -f$ remains unbroken.  One feature of this theory is that
7112: in the superconducting state $hc/e$ vortices tend to have lower energy than
7113: $hc/2e$ vortices, particularly at low doping.  We saw in section IX.C that
7114: $hc/2e$ vortices involve suppression of the pairing amplitude $|\Delta_{\v i\v j}|$
7115: at the center and cost a large energy of order $J$.  On the other hand, one
7116: can form an $hc/e$ vortice by winding the boson phase by $2\pi$, leaving the
7117: fermion pairing intact inside the core.  Another way of describing this from
7118: the point of view of $Z_2$ gauge theory is that the $hc/2e$ vortex necessarily
7119: involves the presence of a $Z_2$ gauge flux (called a vison by Senthil and
7120: Fisher) in its core.  The finite energy cost of the $Z_2$ flux dominates in
7121: the low doping limit and raises the energy of the $hc/2e$ vortices.
7122: Experimental proposals were made 
7123: %(Senthil and Fisher)
7124: \cite{SF0192} to provide for a
7125: critical test of such a theory by detecting the vison excitation or by
7126: indirectly looking for signatures of stable $hc/e$ vortices.  To date, all
7127: such experiments have yielded negative results and provided fairly tight
7128: bounds on the vison energy 
7129: %(Bonn {\em et al.}, 2001)
7130: \cite{BWG0187}.   
7131: 
7132: We are then left with the $U(1)$ spin liquid as the final candidate.  The mean
7133: field basis of this state is the staggered flux liquid state of the $SU(2)$
7134: mean field phase diagram (Fig. \ref{su2phase}).  The low energy theory of this
7135: state consists of fermions with massless Dirac spectra (nodal quasiparticles)
7136: interacting with a $U(1)$ gauge field.  Note that this $U(1)$ gauge field
7137: refers to the low energy gauge group and is not to be confused with the $U(1)$
7138: gauge theory in section IX, which refers to the high energy gauge group, in
7139: the nomenclature of section X.G.  This theory was treated in some detail in
7140: Section XI.  This state has enhanced ($\pi,\pi$) spin fluctuations but no long
7141: range N\'{e}el order, and the ground states becomes a $d$-wave superconductor
7142: when doped with holes.  As we shall see, a low energy $hc/2e$ vortex can be
7143: constructed, thus overcoming a key difficulty of the $Z_2$ gauge theory.
7144: Furthermore, an objection in the literature about the stability of the $U(1)$
7145: spin liquid has been overcome, at least for sufficiently large $N$ (see
7146: section IX.F)  It has also been argued by Senthil and Lee, 2004 that even if
7147: the physical spin 1/2 case does not possess a stable $U(1)$ liquid phase, it
7148: can exist as a critical state separating the N\'{e}el phase from a $Z_2$ spin
7149: liquid and may still have the desired property of dominating the physics of
7150: the pseudogap and the superconducting states.
7151: An example of deconfinement appearing at the critical point between two ordered
7152: phases is recently pointed out by \cite{SVB0490}.
7153: 
7154: In the next section we shall further explore the properties of the $U(1)$ spin
7155: liquid upon doping.  We approach the problem from the low temperature limit
7156: and work our way up in temperature.  This regime is conveniently described by
7157: a nonlinear $\sigma$-model effective theory.
7158: 
7159: \subsection{$\sigma$-model effective theory and new collective modes in the
7160: superconducting state}
7161: 
7162: Here we attempt to reduce the large number of degrees of freedom in the
7163: partition function in \Eq{128} to the few which dominate the low energy
7164: physics.  We shall ignore the amplitude fluctuations in the fermionic degree
7165: of freedom which are gapped on the scale of $J$.  The bosons tend to Bose
7166: condense.  We shall ignore the amplitude fluctuation and assume that its phase
7167: is slowly varying on the fermionic scale, which is given by $\xi =
7168: \epsilon_F/\Delta$ in space.  In this case we can have an effective field
7169: theory ($\sigma$-model) description where the local boson phases are the slow
7170: variables and the fermionic degrees of freedom are assumed to follow them.  We
7171: begin by picking a mean field representation $U_{\v i\v j}^{(0)}$.  The choice of
7172: the staggered flux state $U_{\v i\v j}^{SF}$ given by \Eq{EfEb} is most convenient
7173: because $U_{\v i\v j}^{SF}$ commutes with $\tau^3$, making explicit the residual
7174: $U(1)$ gauge symmetry which corresponds to a $\tau^3$ rotation.  Thus we
7175: choose $ U_{\v i\v j}^{(0)} = U_{\v i\v j}^{SF} e^{ia_{\v i\v j}^3\tau^3} $ and replace the
7176: integral over $U_{\v i\v j}$ by an integral over the gauge field $a_{\v i\v j}^3$.  It
7177: should be noted that any $U_{\v i\v j}^{(0)}$ which are related by $SU(2)$ gauge
7178: transformation will give the same result.  At the mean field level, the bosons
7179: form a band with minima at $Q_0$.  Writing $h = \tilde{h} e^{i\v{Q}_0\cdot
7180: \v{r}}$, we expect $\tilde{h}$ to be slowly  varying in space and time.  We
7181: transform to the radial gauge, 
7182: %from Lee and Nagaosa, insert B
7183: \ie we write
7184: \be
7185: \tilde{h}_{\v i} = g_{\v i}
7186: \bpm b_{\v i} \\
7187: 0 
7188: \epm 
7189: \; , 
7190: \label{eq.154}
7191: \en
7192: where $b_{\v i}$ can be taken as real and positive and $g_{\v i}$ is an $SU(2)$ matrix
7193: parametrized by
7194: \be
7195: g_{\v i}=
7196: \bpm z_{\v i1} & -z_{\v i2}^\ast \\
7197: z_{\v i2} & z_{\v i1}^\ast 
7198: \epm
7199: \; 
7200: \label{eq.155}
7201: \en
7202: where
7203: \be
7204: z_{\v i1}  = e^{i\alpha_{\v i}} e^{-i{\phi_{\v i}\over 2}} \cos {\theta_{\v i} \over 2}
7205: \label{eq.156}
7206: \en
7207: and
7208: \be
7209: z_{\v i2}  = e^{i\alpha_{\v i}} e^{i{\phi_{\v i}\over 2}} \sin {\theta_{\v i} \over 2} \,\,\, .
7210: \label{eq.157}
7211: \en
7212: We ignore the boson amplitude fluctuation and replace $b_{\v i}$ by a constant
7213: $b_0$.
7214: 
7215: An important feature of eq. (\ref{128}) is that $L_2$ is invariant under the 
7216: $SU(2)$ gauge transformation
7217: \be
7218: \tilde{h}_{\v i} &=& g_{\v i}^\dagger h_{\v i}   \label{eq.158} \\ 
7219: \tilde{\psi}_{\v i} &=& g_{\v i}^\dagger \psi_{\v i}  \label{eq.159}\\
7220: \tilde{U}_{\v i\v j} &=& g_{\v i}^\dagger U_{\v i\v j}^{(0)} g_{\v j}
7221: \label{eq.160}
7222: \en
7223: and
7224: \be
7225: \tilde{a}_{0\v i}^\ell \tau^\ell = g^\dagger a_{0\v i}^\ell \tau^\ell g - g \left(
7226: \partial _\tau g^\dagger \right) \,\,\, .
7227: \label{eq.161}
7228: \en
7229: Starting from eq. (\ref{128}) and making the above gauge transformation, the
7230: partition function is
7231: integrated over $g_{\v i}$ instead of $h_{\v i}$ and the Lagrangian
7232: takes the form
7233: \begin{align}
7234: L_2^\prime &= {\tilde{J} \over 2} \sum_{<\v i\v j>} \Tr \left(
7235: \tilde{U}_{\v i\v j}^\dagger \tilde{U}_{\v i\v j} \right) +
7236: \tilde{J} \sum_{<\v i\v j>} \psi_{\v i}^\dagger
7237: \tilde{U}_{\v i\v j}\psi_{\v j} + c.c. \nonumber \\
7238: %&+&  \sum_{\v i} \psi_{\v i}^\dagger 
7239: &+ \sum_{\v i} \psi_{\v i}^\dagger  \left( \partial_\tau - ia_{0\v i}^\ell \tau^\ell  \right)
7240:  \psi_{\v i}\nonumber + \sum_{\v i}  
7241:  \left(- ia_{0\v i}^3 + \mu_B \right)
7242:  b_0^2 \nonumber \\
7243:   &- \sum_{\v i\v j,\sigma} \tilde{t}_{\v i\v j} b_0^2 f_{\v j\sigma}^\dagger f_{\v i\sigma}
7244: \label{eq.162} 
7245: \end{align}
7246: We have removed the tilde from $\tilde{\psi}_{\v i\sigma}$,
7247: $\tilde{f}_{\v i\sigma}$, $\tilde{a}_0^\ell$ because these are integration
7248: variables and $\tilde{t}_{\v i\v j} = t_{\v i\v j}/2$.  %end Lee and Nagaosa, insert B
7249: Note that $g_{\v i}$ appears only in $\tilde{U}_{\v i\v j}$.  For every configuration
7250: $\{g_{\v i}(\tau),  a^3_{\v i\v j}(\tau)\} $ we can, in principle, integrate out the
7251: fermions and $a_0^\ell$ to obtain an energy functional.  This will constitute
7252: the $\sigma$-model description.  In practice, we can make the slowly varying
7253: $g_{\v i}$ approximation and solve the local mean field equation for
7254: $a_{0\v i}^\ell$.  This is the approach taken by 
7255: %Lee, Nagaosa, Ng and Wen (1998)
7256: \Ref{LNNWsu2}.
7257: Note that since $\{g_{\v i}\}$ appears only in the fermionic Lagrangian via
7258: $\tilde{U}_{\v i\v j}$ in eq. (\ref{eq.162}), the resulting energy functional is
7259: entirely fermionic in origin, and no longer has any bosonic contribution.
7260: 
7261: 
7262: The $\si$-model depends on $\{ g_{\v i}(\tau), a^3_{\v i\v j}(\tau) \}$,
7263: \ie it is characterized by $\al_{\v i}$, $\th_{\v i}$, $\phi_{\v i}$ and the gauge field $a_{\v i\v
7264: j}^3$. $\al_{\v i}$ is the familiar overall phase of the electron operator which
7265: becomes half of the pairing phase in the superconducting state. To help
7266: visualize the remaining dependence of freedom, 
7267: %To get a physical picture for this approach, 
7268: it is useful to introduce the local quantization axis
7269: \be \v{I}_{\v i} = z_{\v i}^\dagger \v{\tau} z_{\v i} = ( \sin \theta_{\v i} \cos \phi_{\v i}, \sin
7270: \theta_{\v i} \sin \phi_{\v i}, \cos \theta_{\v i}) \label{eq.163} 
7271: \en 
7272: Note that $\v{I}_{\v i}$ is
7273: independent of the overall phase $\alpha_{\v i}$, which is the phase of the
7274: physical electron operator.  Then different orientations of $\v{I}$ represent
7275: different mean field states in the $U(1)$ mean field theory.  This is shown in
7276: Fig. \ref{fig.26}.  For example, $\v{I}$ pointing to the north pole
7277: corresponds to $g_{\v i} = I$ and the staggered flux state.  This state has $a_0^3
7278: \neq 0$, $a_0^1 = a_0^2 = 0$ and has small Fermi pockets.  It also has orbital
7279: staggered currents around the plaquettes.  $\v{I}$ pointing to the south pole
7280: corresponds to the degenerate staggered flux state whose staggered pattern is
7281: shifted by one unit cell.  On the other hand, when $\v{I}$ is in the equator,
7282: it corresponds to a $d$-wave superconductor.  Note that the angle $\phi$ is a
7283: gauge degree of freedom and states with different $\phi$ anywhere along the
7284: equator are gauge equivalent.  A general orientation of $\v{I}$ corresponds to
7285: some combination of $d$-SC and $s$-flux.
7286: 
7287: \begin{figure}
7288: \includegraphics[width=2.0in]{fig26.eps}
7289: \caption{The quantization axis $\v{I}$ in the $SU(2)$ gauge theory.  The north
7290: and south poles correspond to the staggered flux phases with shifted orbital
7291: current patterns.  All points on the equators are equivalent and correspond to
7292: the $d$-wave superconductor.  In the superconducting state one particular
7293: direction is chosen on the equator.  There are two important collective modes.
7294: The $\theta$ modes correspond to fluctuations in the polar angle
7295: $\delta\theta$ and the $\phi$ gauge mode to a spatially varying fluctuation in
7296: $\delta\phi$.}
7297:  \label{fig.26}
7298: \end{figure}   
7299: 
7300: 
7301: At zero doping, all orientations of $\v{I}$ are energetically the same.  This
7302: symmetry is broken by doping, and the $\v{I}$ vector has a small preference to
7303: lie on the equator.  At low temperature, there is a phase transition to a
7304: state where $\v{I}$ lies on the equator, \ie the $d$-SC ground state.  It is
7305: possible to carry out a small expansion about this state and work out
7306: explicitly the collective modes 
7307: %(Lee \& Nagaosa, 2003)
7308: \cite{LN0316}.  In an ordinary
7309: superconductor, there is a single complex order parameter $\Delta$ and we
7310: expect an amplitude mode and a phase mode.  For a charged superconductor the
7311: phase mode is pushed up to the plasma frequency and one is left with the
7312: amplitude mode only.  In the gauge theory we have in addition to $\Delta_{\v i\v j}$
7313: the order parameter $\chi_{\v i\v j}$.  Thus it is natural to expect additional
7314: collective modes.  From Fig. \ref{fig.26} we see that two modes are of special
7315: interest corresponding to small $\theta$ and $\phi$ fluctuations.  Physically
7316: the $\theta$ mode corresponds to local fluctuations of the $s$-flux states
7317: which generate local orbital current fluctuations.  These currents generate a
7318: small magnetic field (estimated to be $\sim10$ gauss) which couples to
7319: neutrons. 
7320: %Lee and Nagaosa (2003)
7321: \Ref{LN0316} predict a peak in the neutron scattering
7322: cross-section at $(\pi,\pi)$, at energy just below $2\Delta_0$, where
7323: $\Delta_0$ is the maximum $d$-wave gap.  This is {\em in addition} to the
7324: resonance mode discussed in section III.B which is purely spin fluctuation in
7325: origin.  The orbital origin of this mode can be distinguished from the spin
7326: fluctuation by its distinct form factor 
7327: %(Hsu, Marston and Affleck, 1991; Chakravarty, Kee and Nayak, 2002).
7328: (\cit{HMA9166}; \cit{CKN0240})
7329: 
7330: 
7331: The $\phi$ mode is more subtle because 
7332: $\phi$ is the phase of a Higgs field, \ie it is part of
7333: the gauge degree of freedom.  It
7334: turns out to correspond to a relative oscillation of the {\em amplitudes} of
7335: $\chi_{\v i\v j}$ and $\Delta_{\v i\v j}$ and is again most prominent at $(\pi,\pi)$.
7336: Since $|\chi_{\v i\v j}|$ couples to the bond density fluctuation, inelastic Raman
7337: scattering is the tool of choice to study this mode, once the technology
7338: reaches the requisite 10~meV energy resolution.  
7339: %Lee and Nagaosa, 2003
7340: \Ref{LN0316} point
7341: out that due to the special nature of the buckled layers in LSCO, this mode
7342: couples to photons and may show up as a transfer of spectral weight from a
7343: buckling phonon to a higher frequency peak.  Such a peak was reported
7344: experimentally \cite{KTM0304}, but it is apparently not unique to LSCO as the
7345: theory would predict, and hence its interpretation remains unclear at this
7346: point.
7347: 
7348: From Fig. \ref{fig.26} it is clear that the $\sigma$-model representation of
7349: the $SU(2)$ gauge theory is a useful way of parameterizing the myriad $U(1)$
7350: mean field states which become almost degenerate for small doping.  The low
7351: temperature $d-SC$ phase is the ordered phase of the $\sigma$-model, while in
7352: the high temperature limit we expect the $\v{I}$ vector to be disordered in
7353: space and time, to the point where the $\sigma$-mode approach fails and one
7354: crosses over to the $SU(2)$ mean field description.  The disordered phase of
7355: the $\sigma$-model then corresponds to the pseudogap phase.  How does this
7356: phase transition take place?  It turns out that the destruction of
7357: superconducting order proceeds via the usual route of BKT proliferation of
7358: vortices.  To see how this comes about in the $\sigma$-model description, we
7359: have to first understand the structure of vortices. 
7360: 
7361: \subsection{Vortex structure} 
7362: 
7363: The $\sigma$-model picture leads to a natural
7364: model for a low energy $hc/2e$ vortex 
7365: %(Lee and Wen, 2001)
7366: \Ref{LW0117}.  It takes advantage
7367: of the existence of two kinds of bosons $b_1$ and $b_2$ with opposite gauge
7368: charges but the same coupling to electromagnetic fields.  Far away from the
7369: vortex core, $|b_1| = |b_2|$ and $b_1$ has constant phase while $b_2$ winds
7370: its phase by $2\pi$ around the vortex.  As the core is approached $|b_2)$ must
7371: vanish in order to avoid a divergent kinetic energy, as shown in Fig.
7372: \ref{vortex}(top).
7373: The quantization axis $\v{I}$ provides a nice way to visualize this structure
7374: [Fig. \ref{vortex}(bottom)].  It smoothly rotates to the north pole at the vortex core,
7375: indicating that at this level of approximation, the core consists of the
7376: staggered flux state.  The azimuthal angle winds by $2\pi$ as we go around the
7377: vortex. It is important to remember that $\v{I}$ parameterizes  only the
7378: internal gauge degrees of freedom $\theta$ and $\phi$ and the winding of
7379: $\phi$ by $2\pi$ is different from the usual winding of the overall phase
7380: $\alpha$ by $\pi$ in an $hc/2e$ vortex.  To better understand the phase
7381: winding we write down the following continuum model for the phase $\theta_1,
7382: \theta_2$ of $b_1$ and $b_2$, valid far away from the core.
7383: \be
7384: D = \int d^2x {K\over 2} 
7385: \left[ \rule{0in}{.15in}
7386: \left( \v{\nabla} \theta_1 -\v{a} - \v{A} \right) \right.
7387: &+&
7388: \left.
7389: \left(
7390: \nabla \theta_2 + \v{a} - \v{A}
7391: \right)^2
7392: \right] \nonumber \\
7393: && + \cdots
7394: \label{eq.164}
7395: \en
7396: where $\v{a}$ stands for the continuum version of $a^3_{\v i\v j}$ in the last
7397: section, and $\v{A}$ is the electromagnetic field ($e/c$ has been set to be
7398: unity).  We now see that the $hc/2e$ vortex must contain a half integer vortex
7399: of the $\v{a}$ gauge flux with an opposite sign.  Then $\theta_1$ sees zero
7400: flux while $\theta_2$ sees $2\pi$ flux, consistent with the windings chosen in
7401: Fig.\ref{vortex}.  This vortex structure has low energy for small $x$ because
7402: the fermion degrees of freedom remain gapped in the core and one does not pay
7403: the fermionic energy of order $J$ as in the $U(1)$ gauge theory.  Physically,
7404: the above description takes advantage of the states with almost degenerate
7405: energies (in this case the staggered flux state) which is guaranteed by the
7406: $SU(2)$ symmetry near half filling.  There is direct evidence from STM
7407: tunneling that the energy gap is preserved in the core 
7408: %(Maggio-Aprile {\em et al.}, 1995; Pan {\em et al.}, 2000)
7409: \cite{MRE9554,PHG0036}.  This is in contrast to theoretical
7410: expectations for conventional $d$-wave vortex cores, where a large resonance
7411: is expected to fill in the gap in the tunneling spectra 
7412: %(Wang and McDonald, 1995)
7413: \cite{WM9576}.
7414: 
7415: \begin{figure}
7416: \includegraphics[width=2.5in]{fig27.eps}
7417: \caption{Structure of the superconducting vortex.  Top: $b_1$ is constant
7418: while $b_2$ vanishes at the center and its phase winds by $2\pi$.  Bottom: The
7419: isospin quantization axis points to the north pole at the center and rotates
7420: towards the equatorial plane as one moves out radially.  The pattern is
7421: rotationally symmetric around the $\hat{z}$ axis.}
7422:  \label{vortex}
7423: \end{figure} 
7424: 
7425: 
7426: We can clearly reverse the roles of $b_1$ and $b_2$ to produce another vortex
7427: configuration which is degenerate in energy.  In this case $\v{I}$ in Fig.
7428: \ref{vortex}
7429: points to the south pole.  These configurations are sometimes referred to
7430: merons (half of a hedgehog) and the two halves can tunnel to each other via
7431: the appearance of instantons in space-time.  The time scale of the tunneling
7432: event is difficult to estimate, but should be considerably less than $J$.
7433: Depending on the time scale, the orbital current of the staggered flux state
7434: in the core generates a physical staggered magnetic field which may be
7435: experimentally observable by NMR (almost static), $\mu$SR (intermediate time
7436: scale) and neutron (short time scale).  The experiment must be performed in a
7437: large magnetic field so that a significant fraction of the area consists of
7438: vortices and the signal of the staggered field should be proportional to $H$.
7439: A $\mu$SR experiment on underdoped YBCO has detected such a field dependent
7440: signal with a local field of $\pm$18 gauss 
7441: %(Miller {\em {et al.}, 2002})
7442: \cite{MKB0202}.
7443: However $\mu$SR is not able to determine whether the field has an orbital or
7444: spin origin and this experiment is only suggestive, but by no means
7445: definitive, proof of orbital currents in the vortex core.  In principle,
7446: neutron scattering is a more definitive probe, because one can use the form
7447: factor to distinguish between orbital and spin effects.  However, due to the
7448: small expected intensity, neutron scattering has so far not yielded any
7449: definite results.
7450: 
7451: As discussed in section XI.E, we expect enhanced $(\pi, \pi)$ fluctuations to
7452: be associated with the staggered flux liquid phase.  Indeed, the $s$-flux
7453: liquid state is our route to N\'{e}el order and if gauge fluctuations are
7454: large, we may expect to have quasi-static N\'{e}el order inside the vortex
7455: core.  Experimentally, there are reports of enhanced spin fluctuations in the
7456: vortex core by NMR experiments 
7457: (\cit{CMH0073}; \cit{MSB0105}; \cit{MSH0303}; \cit{KKM0203}).  There
7458: are also reports of static incommensurate spin order forming a halo around the
7459: vortex in the LSCO family 
7460: %(Katano {\em et al.}, 2000; Lake {\em et al.}, 2001, 2002; 
7461: %Khaykovich {\em et al.}, 2002)
7462: (\cit{KYS0077}; \cit{LAC0159}; \cit{LRC0299}; \cit{KLE0228}.  One possibility is that these halos are
7463: the condensation of pre-existing soft incommensurate modes known to exist in
7464: LSCO, driven by quasi-static N\'{e}el order inside the core.  We emphasize the
7465: $s$-flux liquid state is our way of producing antiferromagnetic order starting
7466: from microscopies and hence is fully consistent with the appearance of static
7467: or dynamical antiferromagnetism in the vortex core.  Our hope is that gauge
7468: fluctuations (including instanton effects) are sufficiently reduced in doped
7469: systems to permit a glimpse of the staggered orbital current.  The detection
7470: of such currently fluctuations will be a strong confirmation of our approach.
7471: 
7472: Finally, we note that orbital current does not show up directly in STM
7473: experiments, which are sensitive to the local density of states.  However,
7474: %Kishine, Lee and Wen, 2002 
7475: \Ref{KLW0226} have considered the possibility of interference
7476: between Wannier orbitals on neighboring lattice sites, which could lead to
7477: modulations of STM signals {\em between} lattice positions.  STM experiments
7478: have detected $4 \times 4$ modulated patterns in the vortex core region and
7479: also in certain underdoped regions.  Such patterns appear to require density
7480: modulations which are in addition to our vortex model.
7481: 
7482: \subsection{Phase diagram} 
7483: 
7484: We can now construct a phase diagram of the
7485: underdoped cuprates starting from the $d$-wave superconductor ground state at
7486: low temperatures.  The vortex structure allows us to unify the $\sigma$-model
7487: picture with the conventional picture of the destruction of superconducting
7488: order in two dimensions, ie., the BKT transition via the unbinding of
7489: vortices.  The $\sigma$-model 
7490: contains in addition to the pairing phase $2\al$, the phases
7491: $\th$ and $\phi$. However, we saw in section XII.C that a particular
7492: configuration of $\th$ and $\phi$ is favored in side the vortex core.
7493: %the low energy fluctuations are those of the
7494: %pairing phase $\alpha$ and the phase $\theta$.  We saw in section XII.C that
7495: %the $\theta$ fluctuations are tied to the vortex core.  
7496: The $SU(2)$ gauge
7497: theory provides a  mechanism for cheap vortices which are necessary for a BKT
7498: description, as discussed in section V.B.  If the core energy is too large,
7499: the system will behave like a superconductor on any reasonable length scale
7500: above  $T_{\rm BKT}$, which is not in accord with experiment.  On the other
7501: hand, if the core energy is small compared with $T_c$, vortices will
7502: proliferate rapidly. They overlap and lose their identity.  As discussed
7503: section V.B, there is strong experimental evidence that vortices survive over
7504: a considerable temperature range above $T_c$.  Taken as a whole, these
7505: experiments require the vortex core energy to be cheap, but not too cheap,
7506: \ie of the order of $T_c$.  
7507: %Honerkamp and Lee, 2004
7508: \Ref{HL0401} have attempted a
7509: microscopic modeling of the proliferation of vortices.  They assume an
7510: $s$-flux core and estimate the energy from projected wavefunction
7511: calculations.  They indeed found that there is a large range of temperature
7512: above the BKT transition where vortices grow in number but still maintain
7513: their identity.  This forms a region in the phase diagram which may be called
7514: the Nernst region shown in Fig.~\ref{Nernst}.  The corresponding picture of
7515: the $\v{I}$ vector fluctuation is shown in Fig.~\ref{isospin}.  Above the
7516: Nernst region the $\v{I}$ vector is strongly fluctuating and is almost
7517: isotopic.  This is the strongly disordered phase of the $\sigma$-model. The
7518: vortices have lost their identity and indeed the $\sigma$-model description
7519: which assumes well defined phases of $b_1$ and $b_2$ begin to break down.
7520: Nevertheless, the energy gap associated with the fermions remains.  This is
7521: the pseudogap part of the phase diagram in Fig.~\ref{Nernst}.  In the $SU(2)$
7522: gauge theory this is understood as the $U(1)$ spin liquid.  There is no order
7523: parameter in the usual sense associated with this phase, as all fluctuations
7524: including staggered orbital currents and $d$-wave pairing become short range.
7525: Is there a way to characterize this state of affairs other than the term spin
7526: liquid?  This question is addressed in the next section.
7527: 
7528: \begin{figure}
7529: \includegraphics[width=2.5in]{isospin.eps}
7530: \caption{
7531: Schematic picture of the quantization axis $\v{I}$ in different parts of the
7532: phase diagram shown in Fig.~\ref{Nernst}. (a)~In the superconducting phase
7533: $\v{I}$ is ordered in the $x$-$y$ plane.  (b)~In the Nernst phase, $\v{I}$
7534: points to the north or south pole inside the vortex core. (c)~The pseudogap
7535: corresponds to a completely disordered arrangement of $\v{I}$. ($\v{I}$ is a
7536: three dimensional vector and only a two dimensional projection is shown.)
7537: }
7538: \label{isospin}
7539: \end{figure} 
7540: 
7541: 
7542: \subsection{Signature of the spin liquid}
7543: 
7544: %Senthil and Lee (2004) 
7545: \Ref{SL0466}
7546: pointed out that if the pseudogap region is controlled
7547: by the $U(1)$ spin liquid fixed point, it is possible to characterize this
7548: region in a certain precise way.  The spin liquid is a de-confined state,
7549: meaning that instantons are irrelevant.  Then the $U(1)$ gauge flux is a
7550: conserved quantity.  Unfortunately, it is not clear how to couple to this
7551: gauge flux using conventional probes.  We note that the flux associated with
7552: the $\v{a}^3$ gauge field is {\em different} from the $U(1)$ gauge flux
7553: considered in section IX, which had the meaning of spin chirality.  In the
7554: case where the bosons are locally condensed and their local phase well
7555: defined, it is possible to identify the gauge flux in terms of the local 
7556: phase variables.
7557: The gauge magnetic field $\cal{B}$ is given by
7558: \be
7559: {\cal B} & =& (\v \nabla \times \v{a}^3)_z \nonumber \\
7560: & = & \frac{1}{2} \hat{n} \cdot \partial_x \hat{n} \times \partial_y \hat{n}
7561: \label{eq.165}
7562: \en
7563: where
7564: \be
7565: \hat{n} = (\sin \theta \cos \alpha, \sin \theta \sin \alpha, \cos \theta) \nonumber \;\; .
7566: \en
7567: with $\th$ and $\al$ defined in \Eq{eq.156}.
7568: Note that the azimuthal angle associated with $\hat{n}$ is now the pairing
7569: phase $\alpha$, in contrast with the vector $\v{I}$ we considered earlier.
7570: The gauge flux is thus related to the local pairing and $s$-flux order as
7571: \be
7572: {\cal B} = \frac{1}{2} 
7573: \left( \v \nabla  \hat{n}_z  \times \v \nabla \alpha \right)_z
7574: \label{eq.166}
7575: \en
7576: and it is easily checked that the vortex structure described in section XII.C
7577: contains a half integer gauge flux.
7578: 
7579: In the superconducting state the gauge flux is localized in the vortex core
7580: and fluctuations between $\pm$ half integer vortices are possible via
7581: instantons, because the instanton action is finite.  The superconductor is in
7582: a confined phase as far as the $U(1)$ gauge field is concerned.  As the
7583: temperature is raised towards the pseudogap phase this gauge field leaks out
7584: of the vortex cores and begins to fluctuate more and more homogeneously.
7585: 
7586: %Senthil and Lee, insert B
7587: 
7588: The asymptotic conservation of the gauge flux at the Mott transition fixed
7589: point
7590: potentially provides some possibilities for its detection. At non-zero
7591: temperatures in the non-superconducting regions, the flux 
7592: conservation is only approximate (as the instanton fugacity is small but
7593: non-zero). Nevertheless at low enough temperature the conserved flux will 
7594: propagate diffusively over a long range of length and time scales. Thus there
7595: should be an extra diffusive mode that is present at low temperatures in the 
7596: non-superconducting state. It is however not clear how to design a probe that
7597: will couple to this diffusive mode at present.
7598: 
7599: 
7600: 
7601: 
7602: 
7603: Alternately the vortex structure described above provides a useful way to
7604: create and then detect the gauge flux in the non-superconducting
7605: normal state. We will first describe this by ignoring the
7606: instantons completely in the normal state. The effects of
7607: instantons will then be discussed.
7608: 
7609: Consider first a large disc of
7610: cuprate material which is such that the doping level changes as a function of
7611: the radial distance from the center as shown in Fig. \ref{u1ft1}.
7612: The outermost annulus has the largest doping $x_1$. The inner annulus has a
7613: lower doping level $x_2$. The rest of the sample is at a doping level
7614: $x_3< x_2 < x_1$. The corresponding transition temperatures $T_{c1,2,3}$
7615: will be such that $T_{c3} < T_{c2} < T_{c1}$. We also imagine that
7616: the thickness $\Delta R_o, \Delta R_i$ of the outer and inner  annuli
7617: are both much smaller than the penetration
7618: depth for the physical vector potential $A$. The penetration depth
7619: of the internal gauge field $a$ is expected to be small and we
7620: expect it will be smaller than $\Delta R_o, \Delta R_i$.  
7621: We also imagine that the radius of this inner annulus $R_i$ is a substantial
7622: fraction of the radius $R_o$
7623: of the outer annulus.
7624: 
7625: 
7626: \begin{figure}[tb]
7627: \centering
7628: \includegraphics[width=2.0in]{u1ft1.eps}
7629: \caption{Structure of the sample needed for the proposed experiment. The outer annulus (in dark blue) has the highest $T_c$. The inner annulus 
7630: (in light blue) has a smaller $T_c$. The rest of the sample (in brown) has even smaller $T_c$. 
7631:  } 
7632: \label{u1ft1}
7633: \end{figure} 
7634: 
7635:  Now consider the following set of operations on such a sample.
7636: 
7637: (i) First cool in a magnetic field to a temperature $T_{in}$ such that
7638: $T_{c2} < T_{in} < T_{c1}$. The outer ring will then go superconducting while
7639: the rest of the sample stays normal.
7640: In the presence of the field the outer ring will condense into a state in
7641: which there is a net vorticity on going around the ring.
7642: We will be interested in the case where this net vorticity is an odd multiple
7643: of the basic $hc/2e$ vortex. If as assumed the
7644: physical penetration depth is much bigger than the thickness $\Delta R_o$ then
7645: the physical magnetic flux enclosed by the ring will not be quantized.
7646: 
7647: 
7648: (ii) Now consider turning off
7649: the external magnetic field. The vortex present in the outer superconducting
7650: ring will stay (manifested as a small circulating persistent current)
7651: and will give rise to a small magnetic field. As explained above if the
7652: vorticity is odd, then it must be associated with a flux of the
7653: internal gauge field that is $\pm \pi$. This internal gauge flux must
7654: essentially all be in the inner `normal' region of the sample with
7655: very small penetration into the outer superconducting ring. It will spread out
7656: essentially evenly over the full inner region.
7657: 
7658: We have thus managed to create a configuration with a non-zero  internal gauge
7659: flux
7660: in the non-superconducting state.
7661: 
7662: 
7663: 
7664: (iii) How do we detect the presence of this internal gauge flux? For that
7665: imagine now cooling the sample further to a temperature
7666: $T_{fin}$ such that $T_{c3} < T_{fin} < T_{c2}$. Then the inner ring will also
7667: go superconducting. This is to be understood as the condensation
7668: of the two boson species $b_{1,2}$. But this condensation occurs in the
7669: presence of some internal gauge flux. When the bosons $b_{1,2}$ condense in
7670: the inner ring,
7671: they will do so in a manner that quantizes the internal gauge flux enclosed by
7672: this inner ring into an integer multiple of $\pi$. If as
7673: assumed the inner radius is a substantial fraction of the outer radius then
7674: the net internal gauge flux will prefer the quantized values $\pm \pi$ rather
7675: than
7676: be zero (see below). However configurations of the inner ring that enclose
7677: quantized internal gauge flux of $\pm \pi$ also necessarily
7678: contain a physical vortex that is an odd multiple of $hc/2e$. With the
7679: thickness of the inner ring being smaller than the physical penetration depth, 
7680: most of the physical magnetic flux will escape. There will still be a small
7681: residual physical flux due to the current in the inner ring associated with
7682: the 
7683: induced vortex. This residual physical magnetic flux can then be detected.
7684: 
7685: 
7686: 
7687: Note that the sign of the induced physical flux is independent of the sign of
7688: the initial magnetic field. Furthermore
7689: the effect obtains only if the initial vorticity in the outer ring is odd. If
7690: on the other hand the initial vorticity is
7691: even the associated internal gauge flux is zero, and there will be no induced
7692: physical flux when the inner ring goes superconducting.
7693: 
7694: The preceding discussion ignores any effects of instantons. In
7695: contrast to a bulk vortex in the superconducting state the
7696: vortices in the set-up above have macroscopic cores. The internal
7697: gauge flux is therefore distributed over a region of macroscopic
7698: size. Consequently if instantons are irrelevant at long scales in
7699: the normal state, their rate may be expected to be small. At any
7700: non-zero temperature (as in the proposed experiment) there will be
7701: a non-zero instanton rate which will be small for small
7702: temperature.
7703: 
7704: When such instantons are allowed then the internal gauge flux
7705: created in the sample after step (ii) will fluctuate between the
7706: values $+\pi$ and $-\pi$. However so long as the time required to
7707: form the physical vortex in step (iii), which we expect to be short electronic
7708: time scale, is much shorter than the
7709: inverse of the instanton rate we expect that the effect will be seen.
7710: Since the cooling is assumed slow enough that the system always stays in
7711: equilibrium, the outcome of the experiment is determined by thermodynamic
7712: considerations.  
7713: %Senthil and Lee (2004)
7714: \Ref{SL0466} estimated the energies of the various
7715: stages of the operation and concluded that for sample diameters under a micron
7716: and sufficiently low temperatures (= 10 K), such an experiment may be
7717: feasible.
7718: 
7719: 
7720: \section{Summary and outlook}
7721: 
7722: In this review we have summarized a large body of work which views high
7723: temperature superconductivity as the problem of doping of a Mott insulator.
7724: We have argued that the $t$-$J$ model, supplemented by $t^\prime$ terms,
7725: contains the essence of the physics.  We offer as evidence numerical work
7726: based on the projected trial wavefunctions, which correctly predicts the
7727: $d$-wave pairing ground state and a host of properties such as the superfluid
7728: density and the quasiparticle spectral weight and dispersion.  Analytic theory
7729: hinges on the treatment of the constraint of no double occupation.  The
7730: redundancy in the representations used to enforce the constraint naturally
7731: leads to various gauge theories.  We argue that with doping, the gauge theory
7732: may be in a deconfined phase, in which case the slave-boson and fermion
7733: degrees of freedom, which were introduced as mathematical devices, take on a
7734: physical meaning in that they are sensible starting points to describe
7735: physical phenomena.  However, even in the deconfined phase, the coupling to
7736: gauge fluctuations is still of order unity and approximation schemes (such as
7737: large $N$ expansion) are needed to calculate physical properties such as spin
7738: correlation and electron spectral function.  These results qualitatively
7739: capture the physics of the pseudogap phase, but certainly not at a
7740: quantitative level.  Nevertheless, our picture of the vortex structure and how
7741: they proliferate gives us a reasonable account of the phase diagram and the
7742: onset of $T_c$.
7743: 
7744: One direction of future research is to refine the treatment of the low energy
7745: effective model, \ie fermions and bosons coupled to gauge fields, and
7746: attempt more detailed comparison with experiments such as photoemission
7747: lineshapes, \etc.  On the other hand, it is worthwhile to step back and take a
7748: broader perspective.  What is really new and striking about the high
7749: temperature superconductors is the strange ``normal'' metallic state for
7750: underdoped samples.  The carrier density is small and the Fermi surface is
7751: broken up by the appearance of a pseudogap near $(0,\pi)$ and $(\pi, 0)$,
7752: leaving a ``Fermi arc'' near the nodal points.  All this happens without
7753: doubling of the unit cell via breaking translation or spin rotation symmetry.
7754: How this state comes into being in a lightly doped Mott insulator is the crux
7755: of the problem.  We can distinguish between two classes of answers.  The
7756: first, perhaps the more conventional one, postulates the existence of a 
7757: symmetry-breaking state
7758: which gaps the Fermi surface, and further assumes that thermal fluctuation
7759: prevents this state from ordering.  A natural candidate for the state is the
7760: superconducting state itself.  However, it now appears that phase fluctuations
7761: of a superconductor can explain the pseudogap phenomenon only over a
7762: relatively narrow temperature range, which we called the Nernst regime.
7763: Alternatively, a variety of competing states which have nothing to do with
7764: superconductivity have been proposed, often on a phenomenological level, to
7765: produce the pseudogap.  We shall refer to this class of theory as ``thermal''
7766: explanation of the pseudogap.
7767: 
7768: A second class of answer, which we may dub the ``quantum'' explanation,
7769: proposes that the pseudogap is connected with a fundamentally new quantum
7770: state.  Thus, despite its appearance at high temperatures, it is argued that
7771: it is a high frequency phenomenon which is best understood quantum
7772: mechanically.  The gauge theory reviewed here belongs to this class, and views
7773: the pseudogap state as derived from a new state of matter, the quantum spin
7774: liquid state.  The spin liquid state is connected to the N\'{e}el state at
7775: half filling by confinement.  At the same time, with doping a $d$-wave
7776: superconducting ground state is naturally produced.  We argue that rather than
7777: following the route taken by the cuprate in the laboratory of evolving
7778: directly from the antiferromagnet to the superconductor, it is better
7779: conceptually to start from the spin liquid state and consider how AF and
7780: superconductivity develop from it.  In this view the pseudogap is the
7781: closest we can get to obtaining a glimpse of the spin liquid which up to now
7782: is unstable in the square lattice $t$-$J$ model. 
7783: 
7784: Is there a ``smoking gun'' signature to prove or disprove the validity of this
7785: line of theory?  Our approach is to make specific predictions as much as
7786: possible in the hope of stimulating experimental work.  This is the reason we
7787: make special emphasis on the staggered flux liquid with its orbital current
7788: fluctuations, because it is a unique signature which may be experimentally
7789: detectable.  Our predictions range from new collective modes in the
7790: superconducting state, to quasi-static order in the vortex core.
7791: Unfortunately the physical manifestation of the orbital current is a very weak
7792: magnetic field, which is difficult to detect, and to date we have not found
7793: experimental verification.  Besides orbital current, we also propose an
7794: experiment involving flux generation in a special geometry. This experiment
7795: addresses the fundamental issue of the quantum spin liquid as the origin of
7796: the pseudogap phase.
7797: 
7798: The pseudogap metallic state is so strange that at the beginning, it is not
7799: clear if a microscopic description is even possible. So the microscopic
7800: description provided by the $SU(2)$ slave-boson theory, although still
7801: relatively qualitative, represents  important progress and leads to some deep
7802: insights.  A key finding is that the parent spin liquid is a new state of
7803: matter that cannot be described by Landau's symmetry breaking theory.  The
7804: description of the parent spin liquid, such as the $SU(2)$ slave-boson theory,
7805: must involve gauge theory.  Even if one starts with an ordered phase and later
7806: uses quantum fluctuations to restore the symmetry, the resulting description
7807: of the symmetry restored  state, if found, appears to always contain gauge
7808: fields \cite{WMW9846}.  Thus the appearance of the gauge field in the quantum
7809: description of the  pseudogap metal is not a mathematical artifact of the
7810: slave-boson theory. It is a consequence of a new type of correlations in those
7811: states. The new type of correlations represents a new type of order
7812: \cite{Wqoslpub}, which make those states different from the familiar states
7813: described by Landau's symmetry breaking theory.
7814: 
7815: From this perspective, the study of high temperature superconductivity may
7816: have a much broader and deeper impact than merely understanding high
7817: temperature superconductivity. Such a study is actually a study of new states
7818: of matter. It represents our entry into a new exciting world that lies beyond
7819: Landau's world of symmetry breaking.  Hopefully the new states of matter may
7820: be discovered in some materials other than high temperature superconductors.
7821: The slave-boson theory and the resulting gauge theory developed for high
7822: temperature superconductivity may be useful for these new states of matter
7823: once they are discovered in experiments. [Examples of these new states of
7824: matter have already been discovered in many theoretical toy models
7825: (\cit{K032}; \cit{MS0181}; \cit{BFG0212}; \cit{Wqoexct}).] At the moment,
7826: gauge theory is the only known language to describe this new state of
7827: affairs.  We believe the introduction of this subject to condensed matter
7828: physics has enriched the field and will lead to may interesting further
7829: developments.
7830: 
7831: 
7832: 
7833: \begin{acknowledgments}
7834: P.A.L. acknowledges support by NSF grant number
7835: DMR-0201069.
7836: X.G.W. acknowledges support by NSF Grant No. DMR--01--23156,
7837: NSF-MRSEC Grant No. DMR--02--13282, and NFSC no. 10228408.
7838: \end{acknowledgments}
7839: 
7840: % Create the reference section using BibTeX:
7841: %\bibliography{/home/wen/bib/wencross,/home/wen/bib/all,/home/wen/bib/rmp,/home/wen/bib/publst} 
7842: %\end{document}
7843: 
7844: 
7845: 
7846: \begin{thebibliography}{338}
7847: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
7848: \expandafter\ifx\csname bibnamefont\endcsname\relax
7849:   \def\bibnamefont#1{#1}\fi
7850: \expandafter\ifx\csname bibfnamefont\endcsname\relax
7851:   \def\bibfnamefont#1{#1}\fi
7852: \expandafter\ifx\csname citenamefont\endcsname\relax
7853:   \def\citenamefont#1{#1}\fi
7854: \expandafter\ifx\csname url\endcsname\relax
7855:   \def\url#1{\texttt{#1}}\fi
7856: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
7857: \providecommand{\bibinfo}[2]{#2}
7858: \providecommand{\eprint}[2][]{\url{#2}}
7859: 
7860: \bibitem[{\citenamefont{Abanov} \emph{et~al.}(2002)\citenamefont{Abanov,
7861:   Chubukov, Esehig, Norman, and Schmalian}}]{ACE0202}
7862: \bibinfo{author}{\bibnamefont{Abanov}, \bibfnamefont{A.}},
7863:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Chubukov}},
7864:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Esehig}},
7865:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Norman}}, and
7866:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Schmalian}},
7867:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. Lett.}
7868:   \textbf{\bibinfo{volume}{89}}, \bibinfo{pages}{177002}.
7869: 
7870: \bibitem[{\citenamefont{Aeppli} \emph{et~al.}(1995)\citenamefont{Aeppli, Mason,
7871:   Hayden, and Mook}}]{AMH9511}
7872: \bibinfo{author}{\bibnamefont{Aeppli}, \bibfnamefont{G.}},
7873:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Mason}},
7874:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Hayden}}, and
7875:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Mook}}, \bibinfo{year}{1995},
7876:   \bibinfo{journal}{J. Phys. Chem. Solids} \textbf{\bibinfo{volume}{56}},
7877:   \bibinfo{pages}{1911}.
7878: 
7879: \bibitem[{\citenamefont{Affleck and Marston}(1988)}]{AM8874}
7880: \bibinfo{author}{\bibnamefont{Affleck}, \bibfnamefont{I.}}, and
7881:   \bibinfo{author}{\bibfnamefont{J.~B.} \bibnamefont{Marston}},
7882:   \bibinfo{year}{1988}, \bibinfo{journal}{Phys. Rev. B}
7883:   \textbf{\bibinfo{volume}{37}}, \bibinfo{pages}{3774}.
7884: 
7885: \bibitem[{\citenamefont{Affleck} \emph{et~al.}(1988)\citenamefont{Affleck, Zou,
7886:   Hsu, and Anderson}}]{AZH8845}
7887: \bibinfo{author}{\bibnamefont{Affleck}, \bibfnamefont{I.}},
7888:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Zou}},
7889:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Hsu}}, and
7890:   \bibinfo{author}{\bibfnamefont{P.~W.} \bibnamefont{Anderson}},
7891:   \bibinfo{year}{1988}, \bibinfo{journal}{Phys. Rev. B}
7892:   \textbf{\bibinfo{volume}{38}}, \bibinfo{pages}{745}.
7893: 
7894: \bibitem[{\citenamefont{Alloul} \emph{et~al.}(1989)\citenamefont{Alloul, Ohno,
7895:   and Mendels}}]{AOM8900}
7896: \bibinfo{author}{\bibnamefont{Alloul}, \bibfnamefont{H.}},
7897:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Ohno}}, and
7898:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Mendels}},
7899:   \bibinfo{year}{1989}, \bibinfo{journal}{Phys. Rev. Lett.}
7900:   \textbf{\bibinfo{volume}{63}}, \bibinfo{pages}{1700}.
7901: 
7902: \bibitem[{Andersen \emph{et~al.}(1996)\citenamefont{Andersen}
7903:   \emph{et~al.}}]{Ao9685}
7904: \bibinfo{author}{\bibnamefont{Andersen}, \bibfnamefont{O.~K.}}, \emph{et~al.},
7905:   \bibinfo{year}{1996}, \bibinfo{journal}{J. Low Temp. Phys.}
7906:   \textbf{\bibinfo{volume}{105}}, \bibinfo{pages}{285}.
7907: 
7908: \bibitem[{\citenamefont{Anderson}(1973)}]{A7353}
7909: \bibinfo{author}{\bibnamefont{Anderson}, \bibfnamefont{P.~W.}},
7910:   \bibinfo{year}{1973}, \bibinfo{journal}{Mat. Res. Bull.}
7911:   \textbf{\bibinfo{volume}{8}}, \bibinfo{pages}{153}.
7912: 
7913: \bibitem[{\citenamefont{Anderson}(1987)}]{A8796}
7914: \bibinfo{author}{\bibnamefont{Anderson}, \bibfnamefont{P.~W.}},
7915:   \bibinfo{year}{1987}, \bibinfo{journal}{Science}
7916:   \textbf{\bibinfo{volume}{235}}, \bibinfo{pages}{1196}.
7917: 
7918: \bibitem[{\citenamefont{Anderson}(1997)}]{AndHtc97}
7919: \bibinfo{author}{\bibnamefont{Anderson}, \bibfnamefont{P.~W.}},
7920:   \bibinfo{year}{1997}, \emph{\bibinfo{title}{The Theory of Superconductivity
7921:   in the High Tc Cuprates}} (\bibinfo{publisher}{Princeton University Press},
7922:   \bibinfo{address}{Princeto}).
7923: 
7924: \bibitem[{\citenamefont{Anderson} \emph{et~al.}(2004)\citenamefont{Anderson,
7925:   Randeria, Rice, Trivedi, and Zhang}}]{ARR0455}
7926: \bibinfo{author}{\bibnamefont{Anderson}, \bibfnamefont{P.~W.}},
7927:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Randeria}},
7928:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Rice}},
7929:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Trivedi}}, and
7930:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Zhang}},
7931:   \bibinfo{year}{2004}, \bibinfo{journal}{J. Phys. Cond. Matter}
7932:   \textbf{\bibinfo{volume}{16}}, \bibinfo{pages}{R755}.
7933: 
7934: \bibitem[{Armitage \emph{et~al.}(2001)\citenamefont{Armitage}
7935:   \emph{et~al.}}]{Ao0103}
7936: \bibinfo{author}{\bibnamefont{Armitage}, \bibfnamefont{N.~P.}}, \emph{et~al.},
7937:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. Lett.}
7938:   \textbf{\bibinfo{volume}{87}}, \bibinfo{pages}{147003}.
7939: 
7940: \bibitem[{\citenamefont{Arovas} \emph{et~al.}(1984)\citenamefont{Arovas,
7941:   Schrieffer, and Wilczek}}]{ASW8422}
7942: \bibinfo{author}{\bibnamefont{Arovas}, \bibfnamefont{D.}},
7943:   \bibinfo{author}{\bibfnamefont{J.~R.} \bibnamefont{Schrieffer}}, and
7944:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Wilczek}},
7945:   \bibinfo{year}{1984}, \bibinfo{journal}{Phys. Rev. Lett.}
7946:   \textbf{\bibinfo{volume}{53}}, \bibinfo{pages}{722}.
7947: 
7948: \bibitem[{\citenamefont{Arovas and Auerbach}(1988)}]{AA8816}
7949: \bibinfo{author}{\bibnamefont{Arovas}, \bibfnamefont{D.~P.}}, and
7950:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Auerbach}},
7951:   \bibinfo{year}{1988}, \bibinfo{journal}{Phys. Rev. B}
7952:   \textbf{\bibinfo{volume}{38}}, \bibinfo{pages}{316}.
7953: 
7954: \bibitem[{\citenamefont{Arovas} \emph{et~al.}(1997)\citenamefont{Arovas,
7955:   Berlinsky, Kallin, and Zhang}}]{ABK9771}
7956: \bibinfo{author}{\bibnamefont{Arovas}, \bibfnamefont{D.~P.}},
7957:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Berlinsky}},
7958:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Kallin}}, and
7959:   \bibinfo{author}{\bibfnamefont{S.-C.} \bibnamefont{Zhang}},
7960:   \bibinfo{year}{1997}, \bibinfo{journal}{Phys. Rev. Lett.}
7961:   \textbf{\bibinfo{volume}{79}}, \bibinfo{pages}{2871}.
7962: 
7963: \bibitem[{\citenamefont{Assaad}(2004)}]{A0474}
7964: \bibinfo{author}{\bibnamefont{Assaad}, \bibfnamefont{F.}},
7965:   \bibinfo{year}{2004}, \bibinfo{journal}{cond-mat/0406074}.
7966: 
7967: \bibitem[{\citenamefont{Balents} \emph{et~al.}(2002)\citenamefont{Balents,
7968:   Fisher, and Girvin}}]{BFG0212}
7969: \bibinfo{author}{\bibnamefont{Balents}, \bibfnamefont{L.}},
7970:   \bibinfo{author}{\bibfnamefont{M.~P.~A.} \bibnamefont{Fisher}}, and
7971:   \bibinfo{author}{\bibfnamefont{S.~M.} \bibnamefont{Girvin}},
7972:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. B}
7973:   \textbf{\bibinfo{volume}{65}}, \bibinfo{pages}{224412}.
7974: 
7975: \bibitem[{\citenamefont{Balents} \emph{et~al.}(1998)\citenamefont{Balents,
7976:   Fisher, and Nayak}}]{BFN9833}
7977: \bibinfo{author}{\bibnamefont{Balents}, \bibfnamefont{L.}},
7978:   \bibinfo{author}{\bibfnamefont{M.~P.~A.} \bibnamefont{Fisher}}, and
7979:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Nayak}},
7980:   \bibinfo{year}{1998}, \bibinfo{journal}{Int. J. Mod. Phys. B}
7981:   \textbf{\bibinfo{volume}{12}}, \bibinfo{pages}{1033}.
7982: 
7983: \bibitem[{\citenamefont{Banks} \emph{et~al.}(1977)\citenamefont{Banks, Myerson,
7984:   and Kogut}}]{BMK7793}
7985: \bibinfo{author}{\bibnamefont{Banks}, \bibfnamefont{T.}},
7986:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Myerson}}, and
7987:   \bibinfo{author}{\bibfnamefont{J.~B.} \bibnamefont{Kogut}},
7988:   \bibinfo{year}{1977}, \bibinfo{journal}{Nucl. Phys. B}
7989:   \textbf{\bibinfo{volume}{129}}, \bibinfo{pages}{493}.
7990: 
7991: \bibitem[{\citenamefont{Barnes}(1976)}]{B7675}
7992: \bibinfo{author}{\bibnamefont{Barnes}, \bibfnamefont{S.~E.}},
7993:   \bibinfo{year}{1976}, \bibinfo{journal}{J. Phys. F}
7994:   \textbf{\bibinfo{volume}{6}}, \bibinfo{pages}{1375}.
7995: 
7996: \bibitem[{\citenamefont{Baskaran and Anderson}(1998)}]{BA9880}
7997: \bibinfo{author}{\bibnamefont{Baskaran}, \bibfnamefont{G.}}, and
7998:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Anderson}},
7999:   \bibinfo{year}{1998}, \bibinfo{journal}{J. Phys. Chem. Solids}
8000:   \textbf{\bibinfo{volume}{59}}, \bibinfo{pages}{1780}.
8001: 
8002: \bibitem[{\citenamefont{Baskaran and Anderson}(1988)}]{BA8880}
8003: \bibinfo{author}{\bibnamefont{Baskaran}, \bibfnamefont{G.}}, and
8004:   \bibinfo{author}{\bibfnamefont{P.~W.} \bibnamefont{Anderson}},
8005:   \bibinfo{year}{1988}, \bibinfo{journal}{Phys. Rev. B}
8006:   \textbf{\bibinfo{volume}{37}}, \bibinfo{pages}{580}.
8007: 
8008: \bibitem[{\citenamefont{Baskaran} \emph{et~al.}(1987)\citenamefont{Baskaran,
8009:   Zou, and Anderson}}]{BZA8773}
8010: \bibinfo{author}{\bibnamefont{Baskaran}, \bibfnamefont{G.}},
8011:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Zou}}, and
8012:   \bibinfo{author}{\bibfnamefont{P.~W.} \bibnamefont{Anderson}},
8013:   \bibinfo{year}{1987}, \bibinfo{journal}{Solid State Comm.}
8014:   \textbf{\bibinfo{volume}{63}}, \bibinfo{pages}{973}.
8015: 
8016: \bibitem[{\citenamefont{Basov} \emph{et~al.}(1994)\citenamefont{Basov, Puchkov,
8017:   Hughes, Strach, Preston, Timusk, Bonn, Liang, and Hardy}}]{BPH9465}
8018: \bibinfo{author}{\bibnamefont{Basov}, \bibfnamefont{D.~N.}},
8019:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Puchkov}},
8020:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Hughes}},
8021:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Strach}},
8022:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Preston}},
8023:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Timusk}},
8024:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Bonn}},
8025:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Liang}}, and
8026:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Hardy}},
8027:   \bibinfo{year}{1994}, \bibinfo{journal}{Phys. Rev. B}
8028:   \textbf{\bibinfo{volume}{49}}, \bibinfo{pages}{12165}.
8029: 
8030: \bibitem[{\citenamefont{Bednorz and Mueller}(1986)}]{BM8689}
8031: \bibinfo{author}{\bibnamefont{Bednorz}, \bibfnamefont{J.~G.}}, and
8032:   \bibinfo{author}{\bibfnamefont{K.~A.} \bibnamefont{Mueller}},
8033:   \bibinfo{year}{1986}, \bibinfo{journal}{Z. Phys. B}
8034:   \textbf{\bibinfo{volume}{64}}, \bibinfo{pages}{189}.
8035: 
8036: \bibitem[{\citenamefont{Benfatto} \emph{et~al.}(2001)\citenamefont{Benfatto,
8037:   Capara, Castellani, Paramekanti, and Randeria}}]{BCC0113}
8038: \bibinfo{author}{\bibnamefont{Benfatto}, \bibfnamefont{L.}},
8039:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Capara}},
8040:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Castellani}},
8041:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Paramekanti}}, and
8042:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Randeria}},
8043:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. B.}
8044:   \textbf{\bibinfo{volume}{63}}, \bibinfo{pages}{174513}.
8045: 
8046: \bibitem[{\citenamefont{Berezinski}(1971)}]{B7144}
8047: \bibinfo{author}{\bibnamefont{Berezinski}, \bibfnamefont{V.~L.}},
8048:   \bibinfo{year}{1971}, \bibinfo{journal}{Zh. Eksp. Teor. Fiz.}
8049:   \textbf{\bibinfo{volume}{61}}, \bibinfo{pages}{1144}.
8050: 
8051: \bibitem[{\citenamefont{Bonn} \emph{et~al.}(2001)\citenamefont{Bonn, Wynn, b.W.
8052:   Gardner, Liang, Hardy, Kiteley, and Moler}}]{BWG0187}
8053: \bibinfo{author}{\bibnamefont{Bonn}, \bibfnamefont{D.~A.}},
8054:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Wynn}},
8055:   \bibinfo{author}{\bibnamefont{b.W. Gardner}},
8056:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Liang}},
8057:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Hardy}},
8058:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kiteley}}, and
8059:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Moler}},
8060:   \bibinfo{year}{2001}, \bibinfo{journal}{Nature}
8061:   \textbf{\bibinfo{volume}{414}}, \bibinfo{pages}{887}.
8062: 
8063: \bibitem[{Bonn \emph{et~al.}(1996)\citenamefont{Bonn} \emph{et~al.}}]{Bo9695}
8064: \bibinfo{author}{\bibnamefont{Bonn}, \bibfnamefont{D.~A.}}, \emph{et~al.},
8065:   \bibinfo{year}{1996}, \bibinfo{journal}{Czech. J. Phys.}
8066:   \textbf{\bibinfo{volume}{46}}, \bibinfo{pages}{3195}.
8067: 
8068: \bibitem[{\citenamefont{Borokhov} \emph{et~al.}(2002)\citenamefont{Borokhov,
8069:   Kapustin, and Wu}}]{BKW0249}
8070: \bibinfo{author}{\bibnamefont{Borokhov}, \bibfnamefont{V.}},
8071:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kapustin}}, and
8072:   \bibinfo{author}{\bibfnamefont{X.}~\bibnamefont{Wu}}, \bibinfo{year}{2002},
8073:   \bibinfo{journal}{J. High Energy Phys.} \textbf{\bibinfo{volume}{11}},
8074:   \bibinfo{pages}{49}.
8075: 
8076: \bibitem[{\citenamefont{Bourges} \emph{et~al.}(2000)\citenamefont{Bourges,
8077:   Sidis, Fong, Regnault, Bossy, Ivanov, and B.Keimer}}]{BSF0034}
8078: \bibinfo{author}{\bibnamefont{Bourges}, \bibfnamefont{P.}},
8079:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Sidis}},
8080:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Fong}},
8081:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Regnault}},
8082:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Bossy}},
8083:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Ivanov}}, and
8084:   \bibinfo{author}{\bibnamefont{B.Keimer}}, \bibinfo{year}{2000},
8085:   \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{288}},
8086:   \bibinfo{pages}{1234}.
8087: 
8088: \bibitem[{\citenamefont{Boyce} \emph{et~al.}(2000)\citenamefont{Boyce, Skinta,
8089:   and Lemberger}}]{BSL0061}
8090: \bibinfo{author}{\bibnamefont{Boyce}, \bibfnamefont{B.~R.}},
8091:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Skinta}}, and
8092:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Lemberger}},
8093:   \bibinfo{year}{2000}, \bibinfo{journal}{Physica C}
8094:   \textbf{\bibinfo{volume}{341-348}}, \bibinfo{pages}{561}.
8095: 
8096: \bibitem[{\citenamefont{Brinckmann and Lee}(2001)}]{BL0102}
8097: \bibinfo{author}{\bibnamefont{Brinckmann}, \bibfnamefont{J.}}, and
8098:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Lee}}, \bibinfo{year}{2001},
8099:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{65}},
8100:   \bibinfo{pages}{014502}.
8101: 
8102: \bibitem[{\citenamefont{Brinckmann and Lee}(1999)}]{BL9915}
8103: \bibinfo{author}{\bibnamefont{Brinckmann}, \bibfnamefont{J.}}, and
8104:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
8105:   \bibinfo{year}{1999}, \bibinfo{journal}{Phys. Rev. Lett.}
8106:   \textbf{\bibinfo{volume}{82}}, \bibinfo{pages}{2915}.
8107: 
8108: \bibitem[{\citenamefont{Brinckmann and Lee}(2002)}]{BL0202}
8109: \bibinfo{author}{\bibnamefont{Brinckmann}, \bibfnamefont{J.}}, and
8110:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
8111:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. B}
8112:   \textbf{\bibinfo{volume}{65}}, \bibinfo{pages}{014502}.
8113: 
8114: \bibitem[{\citenamefont{Brown} \emph{et~al.}(2004)\citenamefont{Brown, Turner,
8115:   Ozcan, Movgan, Liang, Huang, and Bonn}}]{BTO04}
8116: \bibinfo{author}{\bibnamefont{Brown}, \bibfnamefont{D.~M.}},
8117:   \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Turner}},
8118:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Ozcan}},
8119:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Movgan}},
8120:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Liang}},
8121:   \bibinfo{author}{\bibfnamefont{W.~N.} \bibnamefont{Huang}}, and
8122:   \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Bonn}},
8123:   \bibinfo{year}{2004}, \bibinfo{journal}{to be published} .
8124: 
8125: \bibitem[{\citenamefont{Bulut and Scalapino}(1996)}]{BS9649}
8126: \bibinfo{author}{\bibnamefont{Bulut}, \bibfnamefont{N.}}, and
8127:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Scalapino}},
8128:   \bibinfo{year}{1996}, \bibinfo{journal}{Phys. Rev. B}
8129:   \textbf{\bibinfo{volume}{53}}, \bibinfo{pages}{5149}.
8130: 
8131: \bibitem[{Buyers \emph{et~al.}(2004)\citenamefont{Buyers} \emph{et~al.}}]{Bo04}
8132: \bibinfo{author}{\bibnamefont{Buyers}, \bibfnamefont{W.~J.~L.}}, \emph{et~al.},
8133:   \bibinfo{year}{2004}, \bibinfo{journal}{SNS2004 Conference Proceedings, to be
8134:   published} .
8135: 
8136: \bibitem[{\citenamefont{Campuzano} \emph{et~al.}(1999)\citenamefont{Campuzano,
8137:   Ding, Norman, Fretwell, Randeria, Kaminski, Mesot, Takeuchi, Sato, Yokoya,
8138:   Takahashi, Mochiku} \emph{et~al.}}]{CDN9909}
8139: \bibinfo{author}{\bibnamefont{Campuzano}, \bibfnamefont{J.~C.}},
8140:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ding}},
8141:   \bibinfo{author}{\bibfnamefont{M.~R.} \bibnamefont{Norman}},
8142:   \bibinfo{author}{\bibfnamefont{H.~M.} \bibnamefont{Fretwell}},
8143:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Randeria}},
8144:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kaminski}},
8145:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Mesot}},
8146:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Takeuchi}},
8147:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Sato}},
8148:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Yokoya}},
8149:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Takahashi}},
8150:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Mochiku}}, \emph{et~al.},
8151:   \bibinfo{year}{1999}, \bibinfo{journal}{Phys. Rev. Lett.}
8152:   \textbf{\bibinfo{volume}{83}}, \bibinfo{pages}{3709}.
8153: 
8154: \bibitem[{\citenamefont{Campuzano} \emph{et~al.}(2003)\citenamefont{Campuzano,
8155:   Norman, and Randeria}}]{CNR03}
8156: \bibinfo{author}{\bibnamefont{Campuzano}, \bibfnamefont{J.~C.}},
8157:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Norman}}, and
8158:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Randeria}},
8159:   \bibinfo{year}{2003}, \bibinfo{journal}{in {\em Physics of Conventional and
8160:   Unconventional Superconductors}, edited by K.H. Bennemann and J.B. Ketterson
8161:   (Springer, Berlin)} .
8162: 
8163: \bibitem[{\citenamefont{Carlson} \emph{et~al.}(2003)\citenamefont{Carlson,
8164:   Emery, Kivelson, and Orgad}}]{CEK03}
8165: \bibinfo{author}{\bibnamefont{Carlson}, \bibfnamefont{E.~W.}},
8166:   \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Emery}},
8167:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Kivelson}}, and
8168:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Orgad}},
8169:   \bibinfo{year}{2003}, \bibinfo{journal}{in {\em Physics of Conventional and
8170:   Unconventional Superconductors}, edited by K.H. Bennemann and J.B. Ketterson
8171:   (Springer, Berlin).} .
8172: 
8173: \bibitem[{\citenamefont{Carlson} \emph{et~al.}(1999)\citenamefont{Carlson,
8174:   Kivelson, Emery, and Manousakis}}]{CKE9912}
8175: \bibinfo{author}{\bibnamefont{Carlson}, \bibfnamefont{E.~W.}},
8176:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Kivelson}},
8177:   \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Emery}}, and
8178:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Manousakis}},
8179:   \bibinfo{year}{1999}, \bibinfo{journal}{Phys. Rev. Lett.}
8180:   \textbf{\bibinfo{volume}{83}}, \bibinfo{pages}{612}.
8181: 
8182: \bibitem[{\citenamefont{Castellani}
8183:   \emph{et~al.}(1997)\citenamefont{Castellani, Castro, and Grilli}}]{CCG9737}
8184: \bibinfo{author}{\bibnamefont{Castellani}, \bibfnamefont{C.}},
8185:   \bibinfo{author}{\bibfnamefont{C.~D.} \bibnamefont{Castro}}, and
8186:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Grilli}},
8187:   \bibinfo{year}{1997}, \bibinfo{journal}{Z. Phys.}
8188:   \textbf{\bibinfo{volume}{103}}, \bibinfo{pages}{137}.
8189: 
8190: \bibitem[{\citenamefont{Chakraborty}
8191:   \emph{et~al.}(1990)\citenamefont{Chakraborty, Read, Kane, and Lee}}]{CRK9019}
8192: \bibinfo{author}{\bibnamefont{Chakraborty}, \bibfnamefont{B.}},
8193:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Read}},
8194:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Kane}}, and
8195:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
8196:   \bibinfo{year}{1990}, \bibinfo{journal}{Phys. Rev. B}
8197:   \textbf{\bibinfo{volume}{42}}, \bibinfo{pages}{4819}.
8198: 
8199: \bibitem[{\citenamefont{Chakravarty}
8200:   \emph{et~al.}(2002{\natexlab{a}})\citenamefont{Chakravarty, Kee, and
8201:   Nayak}}]{CKN0240}
8202: \bibinfo{author}{\bibnamefont{Chakravarty}, \bibfnamefont{S.}},
8203:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Kee}}, and
8204:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Nayak}},
8205:   \bibinfo{year}{2002}{\natexlab{a}}, \bibinfo{journal}{Int. J. Mod. Phys. B}
8206:   \textbf{\bibinfo{volume}{16}}, \bibinfo{pages}{3140}.
8207: 
8208: \bibitem[{\citenamefont{Chakravarty}
8209:   \emph{et~al.}(2002{\natexlab{b}})\citenamefont{Chakravarty, Laughlin, Morr,
8210:   and Nayak}}]{CLM0203}
8211: \bibinfo{author}{\bibnamefont{Chakravarty}, \bibfnamefont{S.}},
8212:   \bibinfo{author}{\bibfnamefont{R.~B.} \bibnamefont{Laughlin}},
8213:   \bibinfo{author}{\bibfnamefont{D.~K.} \bibnamefont{Morr}}, and
8214:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Nayak}},
8215:   \bibinfo{year}{2002}{\natexlab{b}}, \bibinfo{journal}{Phys. Rev. B}
8216:   \textbf{\bibinfo{volume}{64}}, \bibinfo{pages}{094503}.
8217: 
8218: \bibitem[{\citenamefont{Chen} \emph{et~al.}(2004)\citenamefont{Chen, Caponi,
8219:   Alet, and Zhang}}]{CCA0416}
8220: \bibinfo{author}{\bibnamefont{Chen}, \bibfnamefont{H.-D.}},
8221:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Caponi}},
8222:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Alet}}, and
8223:   \bibinfo{author}{\bibfnamefont{S.-C.} \bibnamefont{Zhang}},
8224:   \bibinfo{year}{2004}, \bibinfo{journal}{Phys. Rev. B}
8225:   \textbf{\bibinfo{volume}{70}}, \bibinfo{pages}{024516}.
8226: 
8227: \bibitem[{\citenamefont{Cheong} \emph{et~al.}(1991)\citenamefont{Cheong,
8228:   Aeppli, Mason, Mook, Hayden, Canfield, Fisk, Clausen, and
8229:   Martinez}}]{CAM9191}
8230: \bibinfo{author}{\bibnamefont{Cheong}, \bibfnamefont{S.-W.}},
8231:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Aeppli}},
8232:   \bibinfo{author}{\bibfnamefont{T.~E.} \bibnamefont{Mason}},
8233:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Mook}},
8234:   \bibinfo{author}{\bibfnamefont{S.~M.} \bibnamefont{Hayden}},
8235:   \bibinfo{author}{\bibfnamefont{P.~C.} \bibnamefont{Canfield}},
8236:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Fisk}},
8237:   \bibinfo{author}{\bibfnamefont{K.~N.} \bibnamefont{Clausen}}, and
8238:   \bibinfo{author}{\bibfnamefont{J.~L.} \bibnamefont{Martinez}},
8239:   \bibinfo{year}{1991}, \bibinfo{journal}{Phys. Rev. Lett.}
8240:   \textbf{\bibinfo{volume}{67}}, \bibinfo{pages}{1791}.
8241: 
8242: \bibitem[{\citenamefont{Chiao} \emph{et~al.}(2000)\citenamefont{Chiao, Hill,
8243:   Lupien, Taillefer, Lambert, Gagnon, and Fourier}}]{CHL0054}
8244: \bibinfo{author}{\bibnamefont{Chiao}, \bibfnamefont{M.}},
8245:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Hill}},
8246:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Lupien}},
8247:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Taillefer}},
8248:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Lambert}},
8249:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Gagnon}}, and
8250:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Fourier}},
8251:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. B}
8252:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{3554}.
8253: 
8254: \bibitem[{\citenamefont{Chien} \emph{et~al.}(1991)\citenamefont{Chien, Wang,
8255:   and Ong}}]{CWO9188}
8256: \bibinfo{author}{\bibnamefont{Chien}, \bibfnamefont{T.~R.}},
8257:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Wang}}, and
8258:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Ong}}, \bibinfo{year}{1991},
8259:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{67}},
8260:   \bibinfo{pages}{2088}.
8261: 
8262: \bibitem[{\citenamefont{Christensen}
8263:   \emph{et~al.}(2004)\citenamefont{Christensen, McMorrow, Ronnow, Lake, Hayden,
8264:   Aeppli, Perring, Mangkorntong, Nohara, and Tagaki}}]{CMR0439}
8265: \bibinfo{author}{\bibnamefont{Christensen}, \bibfnamefont{N.~B.}},
8266:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{McMorrow}},
8267:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ronnow}},
8268:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Lake}},
8269:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Hayden}},
8270:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Aeppli}},
8271:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Perring}},
8272:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Mangkorntong}},
8273:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Nohara}}, and
8274:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Tagaki}},
8275:   \bibinfo{year}{2004}, \bibinfo{journal}{cond-mat/0403439}.
8276: 
8277: \bibitem[{\citenamefont{Coldea} \emph{et~al.}(2001)\citenamefont{Coldea,
8278:   Hayden, Aeppli, Perrig, Frost, Mason, Cheong, and Fisk}}]{CHA0177}
8279: \bibinfo{author}{\bibnamefont{Coldea}, \bibfnamefont{R.}},
8280:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Hayden}},
8281:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Aeppli}},
8282:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Perrig}},
8283:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Frost}},
8284:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Mason}},
8285:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Cheong}}, and
8286:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Fisk}}, \bibinfo{year}{2001},
8287:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{86}},
8288:   \bibinfo{pages}{5377}.
8289: 
8290: \bibitem[{\citenamefont{Coleman}(1984)}]{C8435}
8291: \bibinfo{author}{\bibnamefont{Coleman}, \bibfnamefont{P.}},
8292:   \bibinfo{year}{1984}, \bibinfo{journal}{Phys. Rev. B}
8293:   \textbf{\bibinfo{volume}{29}}, \bibinfo{pages}{3035}.
8294: 
8295: \bibitem[{\citenamefont{Cooper} \emph{et~al.}(1993)\citenamefont{Cooper,
8296:   Reznik, Kotz, Karlow, Liu, Klein, Lee, Giapintzakis, Ginsberg, Veal, and
8297:   Paulikas}}]{CRK9333}
8298: \bibinfo{author}{\bibnamefont{Cooper}, \bibfnamefont{S.~L.}},
8299:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Reznik}},
8300:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kotz}},
8301:   \bibinfo{author}{\bibfnamefont{M.~A.} \bibnamefont{Karlow}},
8302:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Liu}},
8303:   \bibinfo{author}{\bibfnamefont{M.~V.} \bibnamefont{Klein}},
8304:   \bibinfo{author}{\bibfnamefont{W.~C.} \bibnamefont{Lee}},
8305:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Giapintzakis}},
8306:   \bibinfo{author}{\bibfnamefont{D.~M.} \bibnamefont{Ginsberg}},
8307:   \bibinfo{author}{\bibfnamefont{B.~W.} \bibnamefont{Veal}}, and
8308:   \bibinfo{author}{\bibfnamefont{A.~P.} \bibnamefont{Paulikas}},
8309:   \bibinfo{year}{1993}, \bibinfo{journal}{Phys. Rev. B}
8310:   \textbf{\bibinfo{volume}{47}}, \bibinfo{pages}{8233}.
8311: 
8312: \bibitem[{\citenamefont{Corson} \emph{et~al.}(1999)\citenamefont{Corson,
8313:   Mallozzi, Orenstein, Eckstein, and Bozovic}}]{CMO9921}
8314: \bibinfo{author}{\bibnamefont{Corson}, \bibfnamefont{J.}},
8315:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Mallozzi}},
8316:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Orenstein}},
8317:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Eckstein}}, and
8318:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Bozovic}},
8319:   \bibinfo{year}{1999}, \bibinfo{journal}{Nature}
8320:   \textbf{\bibinfo{volume}{398}}, \bibinfo{pages}{221}.
8321: 
8322: \bibitem[{\citenamefont{Corson} \emph{et~al.}(2000)\citenamefont{Corson,
8323:   Orenstein, Oh, O'Donnell, and Eckstein}}]{COO0069}
8324: \bibinfo{author}{\bibnamefont{Corson}, \bibfnamefont{J.}},
8325:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Orenstein}},
8326:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Oh}},
8327:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{O'Donnell}}, and
8328:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Eckstein}},
8329:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. Lett.}
8330:   \textbf{\bibinfo{volume}{85}}, \bibinfo{pages}{2569}.
8331: 
8332: \bibitem[{\citenamefont{Curro} \emph{et~al.}(1997)\citenamefont{Curro, Imai,
8333:   Slichter, and Dabrowski}}]{CIS9777}
8334: \bibinfo{author}{\bibnamefont{Curro}, \bibfnamefont{N.~J.}},
8335:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Imai}},
8336:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Slichter}}, and
8337:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Dabrowski}},
8338:   \bibinfo{year}{1997}, \bibinfo{journal}{Phys. Rev. B}
8339:   \textbf{\bibinfo{volume}{56}}, \bibinfo{pages}{877}.
8340: 
8341: \bibitem[{\citenamefont{Curro} \emph{et~al.}(2000)\citenamefont{Curro, Milling,
8342:   Haase, and Slichter}}]{CMH0073}
8343: \bibinfo{author}{\bibnamefont{Curro}, \bibfnamefont{N.~J.}},
8344:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Milling}},
8345:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Haase}}, and
8346:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Slichter}},
8347:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. B}
8348:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{3473}.
8349: 
8350: \bibitem[{\citenamefont{D'Adda} \emph{et~al.}(1978)\citenamefont{D'Adda,
8351:   Vecchia, and L{\"u}scher}}]{DDL7863}
8352: \bibinfo{author}{\bibnamefont{D'Adda}, \bibfnamefont{A.}},
8353:   \bibinfo{author}{\bibfnamefont{P.~D.} \bibnamefont{Vecchia}}, and
8354:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{L{\"u}scher}},
8355:   \bibinfo{year}{1978}, \bibinfo{journal}{Nucl. Phys. B}
8356:   \textbf{\bibinfo{volume}{146}}, \bibinfo{pages}{63}.
8357: 
8358: \bibitem[{\citenamefont{Dagotto}(1994)}]{D9463}
8359: \bibinfo{author}{\bibnamefont{Dagotto}, \bibfnamefont{E.}},
8360:   \bibinfo{year}{1994}, \bibinfo{journal}{Rev. Mod. Phys.}
8361:   \textbf{\bibinfo{volume}{66}}, \bibinfo{pages}{763}.
8362: 
8363: \bibitem[{\citenamefont{Dagotto} \emph{et~al.}(1988)\citenamefont{Dagotto,
8364:   Fradkin, and Moreo}}]{DFM8826}
8365: \bibinfo{author}{\bibnamefont{Dagotto}, \bibfnamefont{E.}},
8366:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Fradkin}}, and
8367:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Moreo}},
8368:   \bibinfo{year}{1988}, \bibinfo{journal}{Phys. Rev. B}
8369:   \textbf{\bibinfo{volume}{38}}, \bibinfo{pages}{2926}.
8370: 
8371: \bibitem[{\citenamefont{Dagotto} \emph{et~al.}(1992)\citenamefont{Dagotto,
8372:   Moreo, Ortolani, Poilblanc, and Riera}}]{DMO9241}
8373: \bibinfo{author}{\bibnamefont{Dagotto}, \bibfnamefont{E.}},
8374:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Moreo}},
8375:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Ortolani}},
8376:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Poilblanc}}, and
8377:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Riera}},
8378:   \bibinfo{year}{1992}, \bibinfo{journal}{Phys. Rev. B}
8379:   \textbf{\bibinfo{volume}{45}}, \bibinfo{pages}{10741}.
8380: 
8381: \bibitem[{\citenamefont{Damascelli}
8382:   \emph{et~al.}(2003)\citenamefont{Damascelli, Hussin, and Shen}}]{DHS0373}
8383: \bibinfo{author}{\bibnamefont{Damascelli}, \bibfnamefont{A.}},
8384:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Hussin}}, and
8385:   \bibinfo{author}{\bibfnamefont{Z.-X.} \bibnamefont{Shen}},
8386:   \bibinfo{year}{2003}, \bibinfo{journal}{Rev. Mod. Phys.}
8387:   \textbf{\bibinfo{volume}{75}}, \bibinfo{pages}{473}.
8388: 
8389: \bibitem[{\citenamefont{Demler} \emph{et~al.}(2001)\citenamefont{Demler,
8390:   Sachdev, and Zhang}}]{DSZ0102}
8391: \bibinfo{author}{\bibnamefont{Demler}, \bibfnamefont{E.}},
8392:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Sachdev}}, and
8393:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Zhang}},
8394:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. Lett.}
8395:   \textbf{\bibinfo{volume}{87}}, \bibinfo{pages}{067202}.
8396: 
8397: \bibitem[{\citenamefont{Demler and Zhang}(1995)}]{DZ9526}
8398: \bibinfo{author}{\bibnamefont{Demler}, \bibfnamefont{E.}}, and
8399:   \bibinfo{author}{\bibfnamefont{S.-C.} \bibnamefont{Zhang}},
8400:   \bibinfo{year}{1995}, \bibinfo{journal}{Phys. Rev. Lett.}
8401:   \textbf{\bibinfo{volume}{75}}, \bibinfo{pages}{4126}.
8402: 
8403: \bibitem[{\citenamefont{Ding} \emph{et~al.}(1996)\citenamefont{Ding, Yokoya,
8404:   Campuzano, Takahashi, Randeria, Norman, Mochiku, and Giapintzakis}}]{DYC9651}
8405: \bibinfo{author}{\bibnamefont{Ding}, \bibfnamefont{H.}},
8406:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Yokoya}},
8407:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Campuzano}},
8408:   \bibinfo{author}{\bibfnamefont{T.~T.} \bibnamefont{Takahashi}},
8409:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Randeria}},
8410:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Norman}},
8411:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Mochiku}}, and
8412:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Giapintzakis}},
8413:   \bibinfo{year}{1996}, \bibinfo{journal}{Nature}
8414:   \textbf{\bibinfo{volume}{382}}, \bibinfo{pages}{51}.
8415: 
8416: \bibitem[{\citenamefont{Ding and Makivic}(1991)}]{DM9162}
8417: \bibinfo{author}{\bibnamefont{Ding}, \bibfnamefont{H.~O.}}, and
8418:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Makivic}},
8419:   \bibinfo{year}{1991}, \bibinfo{journal}{Phys. Rev. B}
8420:   \textbf{\bibinfo{volume}{43}}, \bibinfo{pages}{3562}.
8421: 
8422: \bibitem[{\citenamefont{Durst and Lee}(2000)}]{DL0070}
8423: \bibinfo{author}{\bibnamefont{Durst}, \bibfnamefont{A.~C.}}, and
8424:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
8425:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. B}
8426:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{1270}.
8427: 
8428: \bibitem[{\citenamefont{Emery}(1983)}]{E8377}
8429: \bibinfo{author}{\bibnamefont{Emery}, \bibfnamefont{V.~J.}},
8430:   \bibinfo{year}{1983}, \bibinfo{journal}{J. Phys. (Paris) Colloq}
8431:   \textbf{\bibinfo{volume}{44}}, \bibinfo{pages}{C3}.
8432: 
8433: \bibitem[{\citenamefont{Emery}(1986)}]{E8621}
8434: \bibinfo{author}{\bibnamefont{Emery}, \bibfnamefont{V.~J.}},
8435:   \bibinfo{year}{1986}, \bibinfo{journal}{Synth. Met.}
8436:   \textbf{\bibinfo{volume}{13}}, \bibinfo{pages}{21}.
8437: 
8438: \bibitem[{\citenamefont{Emery}(1987)}]{E8759}
8439: \bibinfo{author}{\bibnamefont{Emery}, \bibfnamefont{V.~J.}},
8440:   \bibinfo{year}{1987}, \bibinfo{journal}{Phys. Rev. Lett.}
8441:   \textbf{\bibinfo{volume}{58}}, \bibinfo{pages}{3759}.
8442: 
8443: \bibitem[{\citenamefont{Emery and Kivelson}(1995)}]{EK9534}
8444: \bibinfo{author}{\bibnamefont{Emery}, \bibfnamefont{V.~J.}}, and
8445:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Kivelson}},
8446:   \bibinfo{year}{1995}, \bibinfo{journal}{Nature}
8447:   \textbf{\bibinfo{volume}{374}}, \bibinfo{pages}{434}.
8448: 
8449: \bibitem[{\citenamefont{Fazekas and Anderson}(1974)}]{FA7432}
8450: \bibinfo{author}{\bibnamefont{Fazekas}, \bibfnamefont{P.}}, and
8451:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Anderson}},
8452:   \bibinfo{year}{1974}, \bibinfo{journal}{Philos. Mag.}
8453:   \textbf{\bibinfo{volume}{30}}, \bibinfo{pages}{432}.
8454: 
8455: \bibitem[{Feng \emph{et~al.}(2000)\citenamefont{Feng} \emph{et~al.}}]{Fo0077}
8456: \bibinfo{author}{\bibnamefont{Feng}, \bibfnamefont{D.~L.}}, \emph{et~al.},
8457:   \bibinfo{year}{2000}, \bibinfo{journal}{Science}
8458:   \textbf{\bibinfo{volume}{289}}, \bibinfo{pages}{277}.
8459: 
8460: \bibitem[{\citenamefont{Foerster}(1979)}]{F7987}
8461: \bibinfo{author}{\bibnamefont{Foerster}, \bibfnamefont{D.}},
8462:   \bibinfo{year}{1979}, \bibinfo{journal}{Physics Letters B}
8463:   \textbf{\bibinfo{volume}{87}}, \bibinfo{pages}{87}.
8464: 
8465: \bibitem[{\citenamefont{Fong} \emph{et~al.}(1995)\citenamefont{Fong, Keiman,
8466:   Anderson, Reznik, Dogan, and Aksay}}]{FKA9516}
8467: \bibinfo{author}{\bibnamefont{Fong}, \bibfnamefont{H.~F.}},
8468:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Keiman}},
8469:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Anderson}},
8470:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Reznik}},
8471:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Dogan}}, and
8472:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Aksay}},
8473:   \bibinfo{year}{1995}, \bibinfo{journal}{Phys. Rev. Lett.}
8474:   \textbf{\bibinfo{volume}{75}}, \bibinfo{pages}{316}.
8475: 
8476: \bibitem[{\citenamefont{Fradkin}(1991)}]{Fra91}
8477: \bibinfo{author}{\bibnamefont{Fradkin}, \bibfnamefont{E.}},
8478:   \bibinfo{year}{1991}, \emph{\bibinfo{title}{Field Theories of Condensed
8479:   Matter Systems}} (\bibinfo{publisher}{Addison-Wesley}).
8480: 
8481: \bibitem[{\citenamefont{Fradkin and Shenker}(1979)}]{FS7982}
8482: \bibinfo{author}{\bibnamefont{Fradkin}, \bibfnamefont{E.}}, and
8483:   \bibinfo{author}{\bibfnamefont{S.~H.} \bibnamefont{Shenker}},
8484:   \bibinfo{year}{1979}, \bibinfo{journal}{Phys. Rev. D}
8485:   \textbf{\bibinfo{volume}{19}}, \bibinfo{pages}{3682}.
8486: 
8487: \bibitem[{\citenamefont{Franz} \emph{et~al.}(2003)\citenamefont{Franz,
8488:   Pereg-Barnea, Sheehy, and Tesanovic}}]{FPS0308}
8489: \bibinfo{author}{\bibnamefont{Franz}, \bibfnamefont{M.}},
8490:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Pereg-Barnea}},
8491:   \bibinfo{author}{\bibfnamefont{D.~E.} \bibnamefont{Sheehy}}, and
8492:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Tesanovic}},
8493:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. B}
8494:   \textbf{\bibinfo{volume}{68}}, \bibinfo{pages}{024508}.
8495: 
8496: \bibitem[{\citenamefont{Franz and Tesanovic}(2001)}]{FT0103}
8497: \bibinfo{author}{\bibnamefont{Franz}, \bibfnamefont{M.}}, and
8498:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Tesanovic}},
8499:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. Lett.}
8500:   \textbf{\bibinfo{volume}{87}}, \bibinfo{pages}{257003}.
8501: 
8502: \bibitem[{\citenamefont{Franz} \emph{et~al.}(2002)\citenamefont{Franz,
8503:   Tesanovic, and Vafek}}]{FTV0235}
8504: \bibinfo{author}{\bibnamefont{Franz}, \bibfnamefont{M.}},
8505:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Tesanovic}}, and
8506:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Vafek}},
8507:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. B}
8508:   \textbf{\bibinfo{volume}{66}}, \bibinfo{pages}{054535}.
8509: 
8510: \bibitem[{\citenamefont{Fujita} \emph{et~al.}(2004)\citenamefont{Fujita, Goka,
8511:   Yamada, Tranquada, and Regnault}}]{FGY0496}
8512: \bibinfo{author}{\bibnamefont{Fujita}, \bibfnamefont{M.}},
8513:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Goka}},
8514:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Yamada}},
8515:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Tranquada}}, and
8516:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Regnault}},
8517:   \bibinfo{year}{2004}, \bibinfo{journal}{cond-mat/0403396}.
8518: 
8519: \bibitem[{\citenamefont{Fukuyama}(1992)}]{F9287}
8520: \bibinfo{author}{\bibnamefont{Fukuyama}, \bibfnamefont{H.}},
8521:   \bibinfo{year}{1992}, \bibinfo{journal}{Prog. Theo. Phys. Suppl.}
8522:   \textbf{\bibinfo{volume}{108}}, \bibinfo{pages}{287}.
8523: 
8524: \bibitem[{\citenamefont{Ginzberg}(1989)}]{Gin89}
8525: \bibinfo{author}{\bibnamefont{Ginzberg}, \bibfnamefont{D.~M.}},
8526:   \bibinfo{year}{1989}, \emph{\bibinfo{title}{Physical Properties of High
8527:   Temperature Superconductors}} (\bibinfo{publisher}{World Scientific},
8528:   \bibinfo{address}{Singapore}).
8529: 
8530: \bibitem[{\citenamefont{Gliozzi} \emph{et~al.}(1979)\citenamefont{Gliozzi,
8531:   Regge, and Virasoro}}]{GRV7978}
8532: \bibinfo{author}{\bibnamefont{Gliozzi}, \bibfnamefont{F.}},
8533:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Regge}}, and
8534:   \bibinfo{author}{\bibfnamefont{M.~A.} \bibnamefont{Virasoro}},
8535:   \bibinfo{year}{1979}, \bibinfo{journal}{Physics Letters B}
8536:   \textbf{\bibinfo{volume}{81}}, \bibinfo{pages}{178}.
8537: 
8538: \bibitem[{\citenamefont{Greiter}(1997)}]{G9798}
8539: \bibinfo{author}{\bibnamefont{Greiter}, \bibfnamefont{M.}},
8540:   \bibinfo{year}{1997}, \bibinfo{journal}{Phys. Rev. Lett.}
8541:   \textbf{\bibinfo{volume}{79}}, \bibinfo{pages}{4898}.
8542: 
8543: \bibitem[{\citenamefont{Gros}(1988{\natexlab{a}})}]{G8831}
8544: \bibinfo{author}{\bibnamefont{Gros}, \bibfnamefont{C.}},
8545:   \bibinfo{year}{1988}{\natexlab{a}}, \bibinfo{journal}{Phys. Rev. B}
8546:   \textbf{\bibinfo{volume}{38}}, \bibinfo{pages}{931}.
8547: 
8548: \bibitem[{\citenamefont{Gros}(1988{\natexlab{b}})}]{G8853}
8549: \bibinfo{author}{\bibnamefont{Gros}, \bibfnamefont{C.}},
8550:   \bibinfo{year}{1988}{\natexlab{b}}, \bibinfo{journal}{Annals of Phys.}
8551:   \textbf{\bibinfo{volume}{189}}, \bibinfo{pages}{53}.
8552: 
8553: \bibitem[{\citenamefont{Gweon} \emph{et~al.}(2004)\citenamefont{Gweon,
8554:   Sasagawa, Zhou, Graf, Takagi, Lee, and Lanzara}}]{GSZ0487}
8555: \bibinfo{author}{\bibnamefont{Gweon}, \bibfnamefont{G.-H.}},
8556:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Sasagawa}},
8557:   \bibinfo{author}{\bibfnamefont{S.~Y.} \bibnamefont{Zhou}},
8558:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Graf}},
8559:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Takagi}},
8560:   \bibinfo{author}{\bibfnamefont{D.-H.} \bibnamefont{Lee}}, and
8561:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Lanzara}},
8562:   \bibinfo{year}{2004}, \bibinfo{journal}{Nature}
8563:   \textbf{\bibinfo{volume}{430}}, \bibinfo{pages}{187}.
8564: 
8565: \bibitem[{\citenamefont{H.~Yasuoka}(1989)}]{YIS8954}
8566: \bibinfo{author}{\bibnamefont{H.~Yasuoka}, \bibfnamefont{T.~S., T.~Imai}},
8567:   \bibinfo{year}{1989}, in \emph{\bibinfo{booktitle}{Strong Correlation and
8568:   Superconductivity}}, edited by \bibinfo{editor}{\bibfnamefont{S.~M.}
8569:   \bibnamefont{H.~Fukuyama}} and \bibinfo{editor}{\bibfnamefont{A.~P.}
8570:   \bibnamefont{Malozemoff}} (\bibinfo{publisher}{Springer-Verlag}), p.
8571:   \bibinfo{pages}{254}.
8572: 
8573: \bibitem[{\citenamefont{Hanaguri} \emph{et~al.}(2004)\citenamefont{Hanaguri,
8574:   Lupien, Kohsaka, Lee, Azuma, Takano, Takagi, and Davis}}]{HLK04}
8575: \bibinfo{author}{\bibnamefont{Hanaguri}, \bibfnamefont{T.}},
8576:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Lupien}},
8577:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Kohsaka}},
8578:   \bibinfo{author}{\bibfnamefont{D.-H.} \bibnamefont{Lee}},
8579:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Azuma}},
8580:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Takano}},
8581:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Takagi}}, and
8582:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Davis}},
8583:   \bibinfo{year}{2004}, \bibinfo{journal}{to appear} .
8584: 
8585: \bibitem[{\citenamefont{Hayden} \emph{et~al.}(2004)\citenamefont{Hayden, Mook,
8586:   Dau, Perrig, and Dogan}}]{HMD0431}
8587: \bibinfo{author}{\bibnamefont{Hayden}, \bibfnamefont{S.~M.}},
8588:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Mook}},
8589:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Dau}},
8590:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Perrig}}, and
8591:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Dogan}},
8592:   \bibinfo{year}{2004}, \bibinfo{journal}{Nature}
8593:   \textbf{\bibinfo{volume}{429}}, \bibinfo{pages}{531}.
8594: 
8595: \bibitem[{\citenamefont{Heeb and Rice}(1993)}]{HR9373}
8596: \bibinfo{author}{\bibnamefont{Heeb}, \bibfnamefont{E.~S.}}, and
8597:   \bibinfo{author}{\bibfnamefont{T.~M.} \bibnamefont{Rice}},
8598:   \bibinfo{year}{1993}, \bibinfo{journal}{Z. Phys. B}
8599:   \textbf{\bibinfo{volume}{90}}, \bibinfo{pages}{73}.
8600: 
8601: \bibitem[{\citenamefont{Herbut and Seradjah}(2003)}]{HS0301}
8602: \bibinfo{author}{\bibnamefont{Herbut}, \bibfnamefont{I.~F.}}, and
8603:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Seradjah}},
8604:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. Lett.}
8605:   \textbf{\bibinfo{volume}{91}}, \bibinfo{pages}{171601}.
8606: 
8607: \bibitem[{\citenamefont{Herbut} \emph{et~al.}(2003)\citenamefont{Herbut,
8608:   Seradjeh, Sachdev, and Murthy}}]{HSS0310}
8609: \bibinfo{author}{\bibnamefont{Herbut}, \bibfnamefont{I.~F.}},
8610:   \bibinfo{author}{\bibfnamefont{B.~H.} \bibnamefont{Seradjeh}},
8611:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Sachdev}}, and
8612:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Murthy}},
8613:   \bibinfo{year}{2003}, \bibinfo{journal}{Physical Review B}
8614:   \textbf{\bibinfo{volume}{63}}, \bibinfo{pages}{195110}.
8615: 
8616: \bibitem[{\citenamefont{Hermele and Senthil}(2004)}]{HS04}
8617: \bibinfo{author}{\bibnamefont{Hermele}, \bibfnamefont{M.}}, and
8618:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Senthil}},
8619:   \bibinfo{year}{2004}, \bibinfo{journal}{to be published} .
8620: 
8621: \bibitem[{\citenamefont{Hermele} \emph{et~al.}(2004)\citenamefont{Hermele,
8622:   Senthil, Fisher, Lee, Nagaosa, and Wen}}]{HSF0451}
8623: \bibinfo{author}{\bibnamefont{Hermele}, \bibfnamefont{M.}},
8624:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Senthil}},
8625:   \bibinfo{author}{\bibfnamefont{M.~P.~A.} \bibnamefont{Fisher}},
8626:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
8627:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Nagaosa}}, and
8628:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
8629:   \bibinfo{year}{2004}, \bibinfo{journal}{cond-mat/0404751}.
8630: 
8631: \bibitem[{\citenamefont{Hoffman} \emph{et~al.}(2002)\citenamefont{Hoffman,
8632:   Hudson, Lang, Madhaven, Eisaki, Uchida, and Davis}}]{HHL0266}
8633: \bibinfo{author}{\bibnamefont{Hoffman}, \bibfnamefont{J.~E.}},
8634:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Hudson}},
8635:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Lang}},
8636:   \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Madhaven}},
8637:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Eisaki}},
8638:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Uchida}}, and
8639:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Davis}},
8640:   \bibinfo{year}{2002}, \bibinfo{journal}{Science}
8641:   \textbf{\bibinfo{volume}{195}}, \bibinfo{pages}{466}.
8642: 
8643: \bibitem[{\citenamefont{Holstein} \emph{et~al.}(1973)\citenamefont{Holstein,
8644:   Norton, and Pincus}}]{HNP7349}
8645: \bibinfo{author}{\bibnamefont{Holstein}, \bibfnamefont{T.}},
8646:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Norton}}, and
8647:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Pincus}},
8648:   \bibinfo{year}{1973}, \bibinfo{journal}{Phys. Rev. B}
8649:   \textbf{\bibinfo{volume}{8}}, \bibinfo{pages}{2649}.
8650: 
8651: \bibitem[{\citenamefont{Homes} \emph{et~al.}(1993)\citenamefont{Homes, Timusk,
8652:   Liang, Bonn, and Hardy}}]{HTL9310}
8653: \bibinfo{author}{\bibnamefont{Homes}, \bibfnamefont{C.~C.}},
8654:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Timusk}},
8655:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Liang}},
8656:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Bonn}}, and
8657:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Hardy}},
8658:   \bibinfo{year}{1993}, \bibinfo{journal}{Phys. Rev. Lett.}
8659:   \textbf{\bibinfo{volume}{71}}, \bibinfo{pages}{4210}.
8660: 
8661: \bibitem[{\citenamefont{Honerkamp and Lee}(2004)}]{HL0401}
8662: \bibinfo{author}{\bibnamefont{Honerkamp}, \bibfnamefont{C.}}, and
8663:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
8664:   \bibinfo{year}{2004}, \bibinfo{journal}{Phys. Rev. Lett.}
8665:   \textbf{\bibinfo{volume}{39}}, \bibinfo{pages}{1201}.
8666: 
8667: \bibitem[{\citenamefont{Hoogenboom}
8668:   \emph{et~al.}(2001)\citenamefont{Hoogenboom, Kadowaki, Revaz, Li, Renner, and
8669:   Fischer}}]{HKR0101}
8670: \bibinfo{author}{\bibnamefont{Hoogenboom}, \bibfnamefont{C.}},
8671:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Kadowaki}},
8672:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Revaz}},
8673:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Li}},
8674:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Renner}}, and
8675:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Fischer}},
8676:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. Lett.}
8677:   \textbf{\bibinfo{volume}{87}}, \bibinfo{pages}{267001}.
8678: 
8679: \bibitem[{\citenamefont{Horvatic} \emph{et~al.}(1993)\citenamefont{Horvatic,
8680:   Berthier, Berthier, Segramsan, Butaud, Clark, Gillet, and Henry}}]{HBB9348}
8681: \bibinfo{author}{\bibnamefont{Horvatic}, \bibfnamefont{M.}},
8682:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Berthier}},
8683:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Berthier}},
8684:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Segramsan}},
8685:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Butaud}},
8686:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Clark}},
8687:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Gillet}}, and
8688:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Henry}},
8689:   \bibinfo{year}{1993}, \bibinfo{journal}{Phys. Rev. B}
8690:   \textbf{\bibinfo{volume}{48}}, \bibinfo{pages}{13848}.
8691: 
8692: \bibitem[{\citenamefont{Howland} \emph{et~al.}(2003)\citenamefont{Howland,
8693:   Eisaki, Kaneko, Greven, and Kapitulnik}}]{HEK0333}
8694: \bibinfo{author}{\bibnamefont{Howland}, \bibfnamefont{C.}},
8695:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Eisaki}},
8696:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Kaneko}},
8697:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Greven}}, and
8698:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kapitulnik}},
8699:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. B}
8700:   \textbf{\bibinfo{volume}{67}}, \bibinfo{pages}{014533}.
8701: 
8702: \bibitem[{\citenamefont{Hsu} \emph{et~al.}(1991)\citenamefont{Hsu, Marston, and
8703:   Affleck}}]{HMA9166}
8704: \bibinfo{author}{\bibnamefont{Hsu}, \bibfnamefont{T.}},
8705:   \bibinfo{author}{\bibfnamefont{J.~B.} \bibnamefont{Marston}}, and
8706:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Affleck}},
8707:   \bibinfo{year}{1991}, \bibinfo{journal}{Phys.Rev. B}
8708:   \textbf{\bibinfo{volume}{43}}, \bibinfo{pages}{2866}.
8709: 
8710: \bibitem[{\citenamefont{Hsu}(1990)}]{H9079}
8711: \bibinfo{author}{\bibnamefont{Hsu}, \bibfnamefont{T.~C.}},
8712:   \bibinfo{year}{1990}, \bibinfo{journal}{Phys. Rev. B}
8713:   \textbf{\bibinfo{volume}{41}}, \bibinfo{pages}{11379}.
8714: 
8715: \bibitem[{\citenamefont{Huse and Elser}(1988)}]{HE8831}
8716: \bibinfo{author}{\bibnamefont{Huse}, \bibfnamefont{D.~A.}}, and
8717:   \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Elser}},
8718:   \bibinfo{year}{1988}, \bibinfo{journal}{Phys. Rev. Lett.}
8719:   \textbf{\bibinfo{volume}{60}}, \bibinfo{pages}{2531}.
8720: 
8721: \bibitem[{\citenamefont{Hybertson} \emph{et~al.}(1990)\citenamefont{Hybertson,
8722:   Stechel, Schuter, and Jennison}}]{HSS9068}
8723: \bibinfo{author}{\bibnamefont{Hybertson}, \bibfnamefont{M.~S.}},
8724:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Stechel}},
8725:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Schuter}}, and
8726:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Jennison}},
8727:   \bibinfo{year}{1990}, \bibinfo{journal}{Phys. Rev. B}
8728:   \textbf{\bibinfo{volume}{41}}, \bibinfo{pages}{11068}.
8729: 
8730: \bibitem[{\citenamefont{Ichinose and Matsui}(2001)}]{IM0142}
8731: \bibinfo{author}{\bibnamefont{Ichinose}, \bibfnamefont{I.}}, and
8732:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Matsui}},
8733:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. Lett.}
8734:   \textbf{\bibinfo{volume}{86}}, \bibinfo{pages}{942}.
8735: 
8736: \bibitem[{\citenamefont{Ichinose} \emph{et~al.}(2001)\citenamefont{Ichinose,
8737:   Matsui, and Onoda}}]{IMO0116}
8738: \bibinfo{author}{\bibnamefont{Ichinose}, \bibfnamefont{I.}},
8739:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Matsui}}, and
8740:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Onoda}},
8741:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. B}
8742:   \textbf{\bibinfo{volume}{64}}, \bibinfo{pages}{104516}.
8743: 
8744: \bibitem[{\citenamefont{Imada} \emph{et~al.}(1998)\citenamefont{Imada,
8745:   Fujimori, and Tokura}}]{IFT9839}
8746: \bibinfo{author}{\bibnamefont{Imada}, \bibfnamefont{M.}},
8747:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Fujimori}}, and
8748:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Tokura}},
8749:   \bibinfo{year}{1998}, \bibinfo{journal}{Rev. Mod. Phys.}
8750:   \textbf{\bibinfo{volume}{70}}, \bibinfo{pages}{1039}.
8751: 
8752: \bibitem[{\citenamefont{Ioffe and Larkin}(1989)}]{IL8988}
8753: \bibinfo{author}{\bibnamefont{Ioffe}, \bibfnamefont{L.}}, and
8754:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Larkin}},
8755:   \bibinfo{year}{1989}, \bibinfo{journal}{Phys. Rev. B}
8756:   \textbf{\bibinfo{volume}{39}}, \bibinfo{pages}{8988}.
8757: 
8758: \bibitem[{\citenamefont{Ioffe and Millis}(2001)}]{IM0109}
8759: \bibinfo{author}{\bibnamefont{Ioffe}, \bibfnamefont{L.}}, and
8760:   \bibinfo{author}{\bibfnamefont{A.~J.} \bibnamefont{Millis}},
8761:   \bibinfo{year}{2001}, \bibinfo{journal}{comd-mat} , \bibinfo{pages}{0112509}.
8762: 
8763: \bibitem[{\citenamefont{Ioffe} \emph{et~al.}(2002)\citenamefont{Ioffe,
8764:   Feigel'man, Ioselevich, Ivanov, Troyer, and Blatter}}]{IFI0203}
8765: \bibinfo{author}{\bibnamefont{Ioffe}, \bibfnamefont{L.~B.}},
8766:   \bibinfo{author}{\bibfnamefont{M.~V.} \bibnamefont{Feigel'man}},
8767:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Ioselevich}},
8768:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Ivanov}},
8769:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Troyer}}, and
8770:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Blatter}},
8771:   \bibinfo{year}{2002}, \bibinfo{journal}{Nature}
8772:   \textbf{\bibinfo{volume}{415}}, \bibinfo{pages}{503}.
8773: 
8774: \bibitem[{\citenamefont{Ioffe and Kotliar}(1990)}]{IK9048}
8775: \bibinfo{author}{\bibnamefont{Ioffe}, \bibfnamefont{L.~B.}}, and
8776:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Kotliar}},
8777:   \bibinfo{year}{1990}, \bibinfo{journal}{Phys. Rev. B}
8778:   \textbf{\bibinfo{volume}{42}}, \bibinfo{pages}{10348}.
8779: 
8780: \bibitem[{\citenamefont{Ioffe and Millis}(2002{\natexlab{a}})}]{IM0259}
8781: \bibinfo{author}{\bibnamefont{Ioffe}, \bibfnamefont{L.~B.}}, and
8782:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Millis}},
8783:   \bibinfo{year}{2002}{\natexlab{a}}, \bibinfo{journal}{J. Phys. Chem. Solids}
8784:   \textbf{\bibinfo{volume}{63}}, \bibinfo{pages}{2259}.
8785: 
8786: \bibitem[{\citenamefont{Ioffe and Millis}(2002{\natexlab{b}})}]{IM0213}
8787: \bibinfo{author}{\bibnamefont{Ioffe}, \bibfnamefont{L.~B.}}, and
8788:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Millis}},
8789:   \bibinfo{year}{2002}{\natexlab{b}}, \bibinfo{journal}{Phys. Rev. B}
8790:   \textbf{\bibinfo{volume}{66}}, \bibinfo{pages}{094513}.
8791: 
8792: \bibitem[{\citenamefont{Ishida} \emph{et~al.}(1991)\citenamefont{Ishida,
8793:   Kitaoka, Zhang, and Asayama}}]{IKZ9116}
8794: \bibinfo{author}{\bibnamefont{Ishida}, \bibfnamefont{K.}},
8795:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Kitaoka}},
8796:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Zhang}}, and
8797:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Asayama}},
8798:   \bibinfo{year}{1991}, \bibinfo{journal}{J. Phys. Soc. Jpn.}
8799:   \textbf{\bibinfo{volume}{60}}, \bibinfo{pages}{3516}.
8800: 
8801: \bibitem[{\citenamefont{Ivanov}(2000)}]{I00}
8802: \bibinfo{author}{\bibnamefont{Ivanov}, \bibfnamefont{D.~A.}},
8803:   \bibinfo{year}{2000}, \bibinfo{journal}{MIT PhD Thesis} .
8804: 
8805: \bibitem[{\citenamefont{Ivanov}(2003)}]{I0365}
8806: \bibinfo{author}{\bibnamefont{Ivanov}, \bibfnamefont{D.~A.}},
8807:   \bibinfo{year}{2003}, \bibinfo{journal}{cond-mat/0309265}.
8808: 
8809: \bibitem[{\citenamefont{Ivanov and Lee}(2003)}]{IL0301}
8810: \bibinfo{author}{\bibnamefont{Ivanov}, \bibfnamefont{D.~A.}}, and
8811:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
8812:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. B}
8813:   \textbf{\bibinfo{volume}{68}}, \bibinfo{pages}{132501}.
8814: 
8815: \bibitem[{\citenamefont{Ivanov} \emph{et~al.}(2000)\citenamefont{Ivanov, Lee,
8816:   and Wen}}]{ILW0053}
8817: \bibinfo{author}{\bibnamefont{Ivanov}, \bibfnamefont{D.~A.}},
8818:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}}, and
8819:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
8820:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. Lett.}
8821:   \textbf{\bibinfo{volume}{84}}, \bibinfo{pages}{3953}.
8822: 
8823: \bibitem[{\citenamefont{Kakuyanagi}
8824:   \emph{et~al.}(2002)\citenamefont{Kakuyanagi, Kumagai, and Matsuda}}]{KKM0203}
8825: \bibinfo{author}{\bibnamefont{Kakuyanagi}, \bibfnamefont{K.}},
8826:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Kumagai}}, and
8827:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Matsuda}},
8828:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. B}
8829:   \textbf{\bibinfo{volume}{65}}, \bibinfo{pages}{060503}.
8830: 
8831: \bibitem[{\citenamefont{Kalmeyer and Laughlin}(1987)}]{KL8795}
8832: \bibinfo{author}{\bibnamefont{Kalmeyer}, \bibfnamefont{V.}}, and
8833:   \bibinfo{author}{\bibfnamefont{R.~B.} \bibnamefont{Laughlin}},
8834:   \bibinfo{year}{1987}, \bibinfo{journal}{Phys. Rev. Lett.}
8835:   \textbf{\bibinfo{volume}{59}}, \bibinfo{pages}{2095}.
8836: 
8837: \bibitem[{\citenamefont{Kane} \emph{et~al.}(1989)\citenamefont{Kane, Lee, and
8838:   Read}}]{KLR8980}
8839: \bibinfo{author}{\bibnamefont{Kane}, \bibfnamefont{C.}},
8840:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Lee}}, and
8841:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Read}}, \bibinfo{year}{1989},
8842:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{39}},
8843:   \bibinfo{pages}{6880}.
8844: 
8845: \bibitem[{\citenamefont{Kao} \emph{et~al.}(2000)\citenamefont{Kao, Si, and
8846:   Levin}}]{KSL0098}
8847: \bibinfo{author}{\bibnamefont{Kao}, \bibfnamefont{Y.-J.}},
8848:   \bibinfo{author}{\bibfnamefont{Q.}~\bibnamefont{Si}}, and
8849:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Levin}},
8850:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. B}
8851:   \textbf{\bibinfo{volume}{61}}, \bibinfo{pages}{11898}.
8852: 
8853: \bibitem[{\citenamefont{Kastner} \emph{et~al.}(1998)\citenamefont{Kastner,
8854:   Birgeneau, Shirane, and Endoh}}]{KBS9897}
8855: \bibinfo{author}{\bibnamefont{Kastner}, \bibfnamefont{M.~A.}},
8856:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Birgeneau}},
8857:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Shirane}}, and
8858:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Endoh}},
8859:   \bibinfo{year}{1998}, \bibinfo{journal}{Rev. Mod. Phys.}
8860:   \textbf{\bibinfo{volume}{70}}, \bibinfo{pages}{897}.
8861: 
8862: \bibitem[{\citenamefont{Khasanov} \emph{et~al.}(2003)\citenamefont{Khasanov,
8863:   Shengelaya, Conder, Morenzoni, Savic, and Keller}}]{KSC0317}
8864: \bibinfo{author}{\bibnamefont{Khasanov}, \bibfnamefont{R.}},
8865:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Shengelaya}},
8866:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Conder}},
8867:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Morenzoni}},
8868:   \bibinfo{author}{\bibfnamefont{I.~M.} \bibnamefont{Savic}}, and
8869:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Keller}},
8870:   \bibinfo{year}{2003}, \bibinfo{journal}{J. Phys. -- cond. matt.}
8871:   \textbf{\bibinfo{volume}{15}}, \bibinfo{pages}{L17}.
8872: 
8873: \bibitem[{Khasanov \emph{et~al.}(2004)\citenamefont{Khasanov}
8874:   \emph{et~al.}}]{Ko0402}
8875: \bibinfo{author}{\bibnamefont{Khasanov}, \bibfnamefont{R.}}, \emph{et~al.},
8876:   \bibinfo{year}{2004}, \bibinfo{journal}{Phys. Rev. Lett.}
8877:   \textbf{\bibinfo{volume}{92}}, \bibinfo{pages}{057602}.
8878: 
8879: \bibitem[{\citenamefont{Khaykovich}
8880:   \emph{et~al.}(2002)\citenamefont{Khaykovich, Lee, Erwin, Lee, Wakimoto,
8881:   Thomas, Kastner, and Birgeneau}}]{KLE0228}
8882: \bibinfo{author}{\bibnamefont{Khaykovich}, \bibfnamefont{B.}},
8883:   \bibinfo{author}{\bibfnamefont{Y.~S.} \bibnamefont{Lee}},
8884:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Erwin}},
8885:   \bibinfo{author}{\bibfnamefont{S.-H.} \bibnamefont{Lee}},
8886:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Wakimoto}},
8887:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Thomas}},
8888:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Kastner}}, and
8889:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Birgeneau}},
8890:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. B}
8891:   \textbf{\bibinfo{volume}{66}}, \bibinfo{pages}{014528}.
8892: 
8893: \bibitem[{\citenamefont{Khveshchenko and Wiegmann}(1989)}]{KW8983}
8894: \bibinfo{author}{\bibnamefont{Khveshchenko}, \bibfnamefont{D.}}, and
8895:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Wiegmann}},
8896:   \bibinfo{year}{1989}, \bibinfo{journal}{Mod. Phys. Lett.}
8897:   \textbf{\bibinfo{volume}{3}}, \bibinfo{pages}{1383}.
8898: 
8899: \bibitem[{\citenamefont{Khveshchenko}(2002)}]{K0206}
8900: \bibinfo{author}{\bibnamefont{Khveshchenko}, \bibfnamefont{D.~V.}},
8901:   \bibinfo{year}{2002}, \bibinfo{journal}{cond-mat/0205106}.
8902: 
8903: \bibitem[{\citenamefont{Kim and Lee}(1999)}]{KL9930}
8904: \bibinfo{author}{\bibnamefont{Kim}, \bibfnamefont{D.~H.}}, and
8905:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
8906:   \bibinfo{year}{1999}, \bibinfo{journal}{Annals of Physics}
8907:   \textbf{\bibinfo{volume}{272}}, \bibinfo{pages}{130}.
8908: 
8909: \bibitem[{\citenamefont{Kim} \emph{et~al.}(1997)\citenamefont{Kim, Lee, and
8910:   Wen}}]{KLW9709}
8911: \bibinfo{author}{\bibnamefont{Kim}, \bibfnamefont{D.~H.}},
8912:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}}, and
8913:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
8914:   \bibinfo{year}{1997}, \bibinfo{journal}{Phys. Rev. Lett.}
8915:   \textbf{\bibinfo{volume}{79}}, \bibinfo{pages}{2109}.
8916: 
8917: \bibitem[{\citenamefont{Kimura} \emph{et~al.}(1999)\citenamefont{Kimura,
8918:   Hiroki, Yamada, Endoh, Lee, Majkrzak, Erwin, Shirane, Greven, Lee, Kastner,
8919:   and Birgeneau}}]{KHY9917}
8920: \bibinfo{author}{\bibnamefont{Kimura}, \bibfnamefont{H.}},
8921:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Hiroki}},
8922:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Yamada}},
8923:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Endoh}},
8924:   \bibinfo{author}{\bibfnamefont{S.-H.} \bibnamefont{Lee}},
8925:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Majkrzak}},
8926:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Erwin}},
8927:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Shirane}},
8928:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Greven}},
8929:   \bibinfo{author}{\bibfnamefont{Y.~S.} \bibnamefont{Lee}},
8930:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Kastner}}, and
8931:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Birgeneau}},
8932:   \bibinfo{year}{1999}, \bibinfo{journal}{Phys. Rev. B}
8933:   \textbf{\bibinfo{volume}{59}}, \bibinfo{pages}{6517}.
8934: 
8935: \bibitem[{\citenamefont{Kishine} \emph{et~al.}(2002)\citenamefont{Kishine, Lee,
8936:   and Wen}}]{KLW0226}
8937: \bibinfo{author}{\bibnamefont{Kishine}, \bibfnamefont{J.}},
8938:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}}, and
8939:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
8940:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. B}
8941:   \textbf{\bibinfo{volume}{65}}, \bibinfo{pages}{064526}.
8942: 
8943: \bibitem[{\citenamefont{Kitaev}(2003)}]{K032}
8944: \bibinfo{author}{\bibnamefont{Kitaev}, \bibfnamefont{A.~Y.}},
8945:   \bibinfo{year}{2003}, \bibinfo{journal}{Ann. Phys. (N.Y.)}
8946:   \textbf{\bibinfo{volume}{303}}, \bibinfo{pages}{2}.
8947: 
8948: \bibitem[{\citenamefont{Kitano} \emph{et~al.}(2000)\citenamefont{Kitano,
8949:   Yamada, Suzuki, and Fukase}}]{KYS0077}
8950: \bibinfo{author}{\bibnamefont{Kitano}, \bibfnamefont{S.~M.}},
8951:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Yamada}},
8952:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Suzuki}}, and
8953:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Fukase}},
8954:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. B}
8955:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{14677}.
8956: 
8957: \bibitem[{\citenamefont{Kivelson} \emph{et~al.}(2003)\citenamefont{Kivelson,
8958:   Bindloss, Fradkin, Oganesyan, Tranquada, Kapitulnik, and Howard}}]{KBF0301}
8959: \bibinfo{author}{\bibnamefont{Kivelson}, \bibfnamefont{S.~A.}},
8960:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Bindloss}},
8961:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Fradkin}},
8962:   \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Oganesyan}},
8963:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Tranquada}},
8964:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kapitulnik}}, and
8965:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Howard}},
8966:   \bibinfo{year}{2003}, \bibinfo{journal}{Rev. Mod. Phys.}
8967:   \textbf{\bibinfo{volume}{75}}, \bibinfo{pages}{1201}.
8968: 
8969: \bibitem[{\citenamefont{Kivelson} \emph{et~al.}(1987)\citenamefont{Kivelson,
8970:   Rokhsar, and Sethna}}]{KRS8765}
8971: \bibinfo{author}{\bibnamefont{Kivelson}, \bibfnamefont{S.~A.}},
8972:   \bibinfo{author}{\bibfnamefont{D.~S.} \bibnamefont{Rokhsar}}, and
8973:   \bibinfo{author}{\bibfnamefont{J.~P.} \bibnamefont{Sethna}},
8974:   \bibinfo{year}{1987}, \bibinfo{journal}{Phys. Rev. B}
8975:   \textbf{\bibinfo{volume}{35}}, \bibinfo{pages}{8865}.
8976: 
8977: \bibitem[{\citenamefont{Kogut and Susskind}(1975)}]{KS7595}
8978: \bibinfo{author}{\bibnamefont{Kogut}, \bibfnamefont{J.}}, and
8979:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Susskind}},
8980:   \bibinfo{year}{1975}, \bibinfo{journal}{Phys. Rev. D}
8981:   \textbf{\bibinfo{volume}{11}}, \bibinfo{pages}{395}.
8982: 
8983: \bibitem[{\citenamefont{Kosterlitz and Thouless}(1973)}]{KT7381}
8984: \bibinfo{author}{\bibnamefont{Kosterlitz}, \bibfnamefont{J.~M.}}, and
8985:   \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Thouless}},
8986:   \bibinfo{year}{1973}, \bibinfo{journal}{J. Phys. C}
8987:   \textbf{\bibinfo{volume}{6}}, \bibinfo{pages}{1181}.
8988: 
8989: \bibitem[{\citenamefont{Kotliar and Liu}(1988)}]{KL8842}
8990: \bibinfo{author}{\bibnamefont{Kotliar}, \bibfnamefont{G.}}, and
8991:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Liu}}, \bibinfo{year}{1988},
8992:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{38}},
8993:   \bibinfo{pages}{5142}.
8994: 
8995: \bibitem[{\citenamefont{Kugler} \emph{et~al.}(2001)\citenamefont{Kugler,
8996:   Fischer, Renner, Ono, and Ando}}]{KFR0111}
8997: \bibinfo{author}{\bibnamefont{Kugler}, \bibfnamefont{M.}},
8998:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Fischer}},
8999:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Renner}},
9000:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Ono}}, and
9001:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Ando}}, \bibinfo{year}{2001},
9002:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{86}},
9003:   \bibinfo{pages}{4911}.
9004: 
9005: \bibitem[{\citenamefont{Kuzmenko} \emph{et~al.}(2003)\citenamefont{Kuzmenko,
9006:   Tombros, Molegraaf, Grueninger, van~der Marel, and Uchida}}]{KTM0304}
9007: \bibinfo{author}{\bibnamefont{Kuzmenko}, \bibfnamefont{A.~B.}},
9008:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Tombros}},
9009:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Molegraaf}},
9010:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Grueninger}},
9011:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{van~der Marel}}, and
9012:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Uchida}},
9013:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. Lett.}
9014:   \textbf{\bibinfo{volume}{91}}, \bibinfo{pages}{037004}.
9015: 
9016: \bibitem[{\citenamefont{Lake} \emph{et~al.}(2001)\citenamefont{Lake, Aeppli,
9017:   Clausen, McMorrow, Lefmann, Hussey, Mangkorntong, Nohara, Takagi, Mason, and
9018:   Schroder}}]{LAC0159}
9019: \bibinfo{author}{\bibnamefont{Lake}, \bibfnamefont{B.}},
9020:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Aeppli}},
9021:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Clausen}},
9022:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{McMorrow}},
9023:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Lefmann}},
9024:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Hussey}},
9025:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Mangkorntong}},
9026:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Nohara}},
9027:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Takagi}},
9028:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Mason}}, and
9029:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Schroder}},
9030:   \bibinfo{year}{2001}, \bibinfo{journal}{Science}
9031:   \textbf{\bibinfo{volume}{291}}, \bibinfo{pages}{1759}.
9032: 
9033: \bibitem[{\citenamefont{Lake} \emph{et~al.}(2002)\citenamefont{Lake, Ronnow,
9034:   Christensen, Aeppli, Lefmann, McMorrow, Vorderwisch, Smeibidl, Mangkorntong,
9035:   Sasagawa, Nohara, Takagi} \emph{et~al.}}]{LRC0299}
9036: \bibinfo{author}{\bibnamefont{Lake}, \bibfnamefont{B.}},
9037:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ronnow}},
9038:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Christensen}},
9039:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Aeppli}},
9040:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Lefmann}},
9041:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{McMorrow}},
9042:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Vorderwisch}},
9043:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Smeibidl}},
9044:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Mangkorntong}},
9045:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Sasagawa}},
9046:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Nohara}},
9047:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Takagi}}, \emph{et~al.},
9048:   \bibinfo{year}{2002}, \bibinfo{journal}{Nature}
9049:   \textbf{\bibinfo{volume}{415}}, \bibinfo{pages}{299}.
9050: 
9051: \bibitem[{\citenamefont{Laughlin}(1995)}]{L9527}
9052: \bibinfo{author}{\bibnamefont{Laughlin}, \bibfnamefont{R.~B.}},
9053:   \bibinfo{year}{1995}, \bibinfo{journal}{J. Phys. Chem. Solid}
9054:   \textbf{\bibinfo{volume}{56}}, \bibinfo{pages}{1627}.
9055: 
9056: \bibitem[{\citenamefont{Laughlin}(1997)}]{L9726}
9057: \bibinfo{author}{\bibnamefont{Laughlin}, \bibfnamefont{R.~B.}},
9058:   \bibinfo{year}{1997}, \bibinfo{journal}{Phys. Rev. Lett.}
9059:   \textbf{\bibinfo{volume}{79}}, \bibinfo{pages}{1726}.
9060: 
9061: \bibitem[{\citenamefont{Lee}(2000)}]{L0094}
9062: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{D.-H.}},
9063:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. Lett.}
9064:   \textbf{\bibinfo{volume}{84}}, \bibinfo{pages}{2694}.
9065: 
9066: \bibitem[{\citenamefont{Lee} \emph{et~al.}(1996)\citenamefont{Lee, Kim, and
9067:   Lee}}]{LKL9601}
9068: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{D.~K.~K.}},
9069:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Kim}}, and
9070:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
9071:   \bibinfo{year}{1996}, \bibinfo{journal}{Phys. Rev. Lett.}
9072:   \textbf{\bibinfo{volume}{76}}, \bibinfo{pages}{4801}.
9073: 
9074: \bibitem[{\citenamefont{Lee}(1993)}]{L9387}
9075: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{P.~A.}},
9076:   \bibinfo{year}{1993}, \bibinfo{journal}{Phys. Rev. Lett.}
9077:   \textbf{\bibinfo{volume}{71}}, \bibinfo{pages}{1887}.
9078: 
9079: \bibitem[{\citenamefont{Lee}(2002)}]{L0249}
9080: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{P.~A.}},
9081:   \bibinfo{year}{2002}, \bibinfo{journal}{J. Phys. and Chem. Solids}
9082:   \textbf{\bibinfo{volume}{63}}, \bibinfo{pages}{2149}.
9083: 
9084: \bibitem[{\citenamefont{Lee and Nagaosa}(1992)}]{LN9221}
9085: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{P.~A.}}, and
9086:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Nagaosa}},
9087:   \bibinfo{year}{1992}, \bibinfo{journal}{Phys. Rev. B}
9088:   \textbf{\bibinfo{volume}{45}}, \bibinfo{pages}{5621}.
9089: 
9090: \bibitem[{\citenamefont{Lee and Nagaosa}(2003)}]{LN0316}
9091: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{P.~A.}}, and
9092:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Nagaosa}},
9093:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. B}
9094:   \textbf{\bibinfo{volume}{68}}, \bibinfo{pages}{024516}.
9095: 
9096: \bibitem[{\citenamefont{Lee} \emph{et~al.}(1998)\citenamefont{Lee, Nagaosa, Ng,
9097:   and Wen}}]{LNNWsu2}
9098: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{P.~A.}},
9099:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Nagaosa}},
9100:   \bibinfo{author}{\bibfnamefont{T.-K.} \bibnamefont{Ng}}, and
9101:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
9102:   \bibinfo{year}{1998}, \bibinfo{journal}{Phys. Rev. B}
9103:   \textbf{\bibinfo{volume}{57}}, \bibinfo{pages}{6003}.
9104: 
9105: \bibitem[{\citenamefont{Lee and Sha}(2003)}]{LS0371}
9106: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{P.~A.}}, and
9107:   \bibinfo{author}{\bibfnamefont{G.~B.} \bibnamefont{Sha}},
9108:   \bibinfo{year}{2003}, \bibinfo{journal}{Sol. States Comm.}
9109:   \textbf{\bibinfo{volume}{126}}, \bibinfo{pages}{71}.
9110: 
9111: \bibitem[{\citenamefont{Lee and Wen}(1997)}]{LW9711}
9112: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{P.~A.}}, and
9113:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
9114:   \bibinfo{year}{1997}, \bibinfo{journal}{Phys. Rev. Lett.}
9115:   \textbf{\bibinfo{volume}{78}}, \bibinfo{pages}{4111}.
9116: 
9117: \bibitem[{\citenamefont{Lee and Wen}(2001)}]{LW0117}
9118: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{P.~A.}}, and
9119:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
9120:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. B}
9121:   \textbf{\bibinfo{volume}{63}}, \bibinfo{pages}{224517}.
9122: 
9123: \bibitem[{\citenamefont{Lee and Salk}(2001)}]{LS0101}
9124: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{S.-S.}}, and
9125:   \bibinfo{author}{\bibfnamefont{S.-H.} \bibnamefont{Salk}},
9126:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. B}
9127:   \textbf{\bibinfo{volume}{64}}, \bibinfo{pages}{052501}.
9128: 
9129: \bibitem[{\citenamefont{Lee and Feng}(1988)}]{LF8809}
9130: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{T.-K.}}, and
9131:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Feng}}, \bibinfo{year}{1988},
9132:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{38}},
9133:   \bibinfo{pages}{11809}.
9134: 
9135: \bibitem[{\citenamefont{Lee}
9136:   \emph{et~al.}(2003{\natexlab{a}})\citenamefont{Lee, Ho, and
9137:   Nagaosa}}]{LHN0301}
9138: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{T.~K.}},
9139:   \bibinfo{author}{\bibfnamefont{C.-M.} \bibnamefont{Ho}}, and
9140:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Nagaosa}},
9141:   \bibinfo{year}{2003}{\natexlab{a}}, \bibinfo{journal}{Phys. Rev. Lett.}
9142:   \textbf{\bibinfo{volume}{90}}, \bibinfo{pages}{067001}.
9143: 
9144: \bibitem[{\citenamefont{Lee}
9145:   \emph{et~al.}(2003{\natexlab{b}})\citenamefont{Lee, Lee, Ho, and
9146:   Leung}}]{LLH0301}
9147: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{W.-C.}},
9148:   \bibinfo{author}{\bibfnamefont{T.~K.} \bibnamefont{Lee}},
9149:   \bibinfo{author}{\bibfnamefont{C.-M.} \bibnamefont{Ho}}, and
9150:   \bibinfo{author}{\bibfnamefont{P.~W.} \bibnamefont{Leung}},
9151:   \bibinfo{year}{2003}{\natexlab{b}}, \bibinfo{journal}{Phys. Rev. Lett.}
9152:   \textbf{\bibinfo{volume}{91}}, \bibinfo{pages}{057001}.
9153: 
9154: \bibitem[{\citenamefont{Lee} \emph{et~al.}(1999)\citenamefont{Lee, Birgeneau,
9155:   Kastner, Endoh, Wakimoto, Yamada, Erwin, Lee, and Shirane}}]{LBK9943}
9156: \bibinfo{author}{\bibnamefont{Lee}, \bibfnamefont{Y.~S.}},
9157:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Birgeneau}},
9158:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Kastner}},
9159:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Endoh}},
9160:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Wakimoto}},
9161:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Yamada}},
9162:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Erwin}},
9163:   \bibinfo{author}{\bibfnamefont{S.-H.} \bibnamefont{Lee}}, and
9164:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Shirane}},
9165:   \bibinfo{year}{1999}, \bibinfo{journal}{Phys. Rev. B}
9166:   \textbf{\bibinfo{volume}{60}}, \bibinfo{pages}{3643}.
9167: 
9168: \bibitem[{\citenamefont{Leung}(2000)}]{L0012}
9169: \bibinfo{author}{\bibnamefont{Leung}, \bibfnamefont{P.~W.}},
9170:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. B}
9171:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{6112}.
9172: 
9173: \bibitem[{\citenamefont{Levin and Wen}(2003)}]{LWsta}
9174: \bibinfo{author}{\bibnamefont{Levin}, \bibfnamefont{M.}}, and
9175:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
9176:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. B}
9177:   \textbf{\bibinfo{volume}{67}}, \bibinfo{pages}{245316}.
9178: 
9179: \bibitem[{\citenamefont{Littlewood}
9180:   \emph{et~al.}(1993)\citenamefont{Littlewood, Zaanen, Aeppli, H, and
9181:   Monien}}]{LZA9387}
9182: \bibinfo{author}{\bibnamefont{Littlewood}, \bibfnamefont{P.~B.}},
9183:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Zaanen}},
9184:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Aeppli}},
9185:   \bibinfo{author}{\bibnamefont{H}}, and
9186:   \bibinfo{author}{\bibnamefont{Monien}}, \bibinfo{year}{1993},
9187:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{48}},
9188:   \bibinfo{pages}{487}.
9189: 
9190: \bibitem[{\citenamefont{Liu} \emph{et~al.}(1995)\citenamefont{Liu, Zha, and
9191:   Levin}}]{LZL9530}
9192: \bibinfo{author}{\bibnamefont{Liu}, \bibfnamefont{D.~Z.}},
9193:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Zha}}, and
9194:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Levin}},
9195:   \bibinfo{year}{1995}, \bibinfo{journal}{Phys. Rev. Lett.}
9196:   \textbf{\bibinfo{volume}{75}}, \bibinfo{pages}{4130}.
9197: 
9198: \bibitem[{\citenamefont{Liu and Manousakis}(1992)}]{LM9225}
9199: \bibinfo{author}{\bibnamefont{Liu}, \bibfnamefont{Z.}}, and
9200:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Manousakis}},
9201:   \bibinfo{year}{1992}, \bibinfo{journal}{Phys. Rev. B}
9202:   \textbf{\bibinfo{volume}{45}}, \bibinfo{pages}{2425}.
9203: 
9204: \bibitem[{\citenamefont{Loeser} \emph{et~al.}(1996)\citenamefont{Loeser, Shen,
9205:   Desau, Marshall, Park, Fournier, and Kapitulnik}}]{LSD9625}
9206: \bibinfo{author}{\bibnamefont{Loeser}, \bibfnamefont{A.~C.}},
9207:   \bibinfo{author}{\bibfnamefont{Z.-X.} \bibnamefont{Shen}},
9208:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Desau}},
9209:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Marshall}},
9210:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Park}},
9211:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Fournier}}, and
9212:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kapitulnik}},
9213:   \bibinfo{year}{1996}, \bibinfo{journal}{Science}
9214:   \textbf{\bibinfo{volume}{273}}, \bibinfo{pages}{325}.
9215: 
9216: \bibitem[{\citenamefont{Loram} \emph{et~al.}(2001)\citenamefont{Loram, Luo,
9217:   Cooper, Liang, and Tallon}}]{LLC0159}
9218: \bibinfo{author}{\bibnamefont{Loram}, \bibfnamefont{J.~W.}},
9219:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Luo}},
9220:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Cooper}},
9221:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Liang}}, and
9222:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Tallon}},
9223:   \bibinfo{year}{2001}, \bibinfo{journal}{J. Phys. Chem. Solids}
9224:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{59}.
9225: 
9226: \bibitem[{\citenamefont{Loram} \emph{et~al.}(1993)\citenamefont{Loram, Mirza,
9227:   Cooper, and Liang}}]{LMC9340}
9228: \bibinfo{author}{\bibnamefont{Loram}, \bibfnamefont{J.~W.}},
9229:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Mirza}},
9230:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Cooper}}, and
9231:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Liang}},
9232:   \bibinfo{year}{1993}, \bibinfo{journal}{Phys. Rev. Lett.}
9233:   \textbf{\bibinfo{volume}{71}}, \bibinfo{pages}{1740}.
9234: 
9235: \bibitem[{\citenamefont{Maggio-Aprile}
9236:   \emph{et~al.}(1995)\citenamefont{Maggio-Aprile, Renner, Erb, Walker, and
9237:   Fischer}}]{MRE9554}
9238: \bibinfo{author}{\bibnamefont{Maggio-Aprile}, \bibfnamefont{I.}},
9239:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Renner}},
9240:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Erb}},
9241:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Walker}}, and
9242:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Fischer}},
9243:   \bibinfo{year}{1995}, \bibinfo{journal}{Phys. Rev. Lett.}
9244:   \textbf{\bibinfo{volume}{75}}, \bibinfo{pages}{2754}.
9245: 
9246: \bibitem[{\citenamefont{Mandelstam}(1979)}]{M7991}
9247: \bibinfo{author}{\bibnamefont{Mandelstam}, \bibfnamefont{S.}},
9248:   \bibinfo{year}{1979}, \bibinfo{journal}{Phys. Rev. D}
9249:   \textbf{\bibinfo{volume}{19}}, \bibinfo{pages}{2391}.
9250: 
9251: \bibitem[{\citenamefont{Marshall} \emph{et~al.}(1996)\citenamefont{Marshall,
9252:   Dessau, Loeser, Park, Matsuura, Eckstein, Bozovic, Fournier, Kapitulnik,
9253:   Spicer, and Shen}}]{MDL9641}
9254: \bibinfo{author}{\bibnamefont{Marshall}, \bibfnamefont{D.~S.}},
9255:   \bibinfo{author}{\bibfnamefont{D.~S.} \bibnamefont{Dessau}},
9256:   \bibinfo{author}{\bibfnamefont{A.~G.} \bibnamefont{Loeser}},
9257:   \bibinfo{author}{\bibfnamefont{C.-H.} \bibnamefont{Park}},
9258:   \bibinfo{author}{\bibfnamefont{A.~Y.} \bibnamefont{Matsuura}},
9259:   \bibinfo{author}{\bibfnamefont{J.~N.} \bibnamefont{Eckstein}},
9260:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Bozovic}},
9261:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Fournier}},
9262:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kapitulnik}},
9263:   \bibinfo{author}{\bibfnamefont{W.~E.} \bibnamefont{Spicer}}, and
9264:   \bibinfo{author}{\bibfnamefont{Z.-X.} \bibnamefont{Shen}},
9265:   \bibinfo{year}{1996}, \bibinfo{journal}{Phys. Rev. Lett.}
9266:   \textbf{\bibinfo{volume}{76}}, \bibinfo{pages}{4841}.
9267: 
9268: \bibitem[{\citenamefont{Matsuda} \emph{et~al.}(2000)\citenamefont{Matsuda,
9269:   Fujita, Yamada, Birgeneau, Kastner, Hiraka, Endoh, Wakimoto, and
9270:   Shirane}}]{MFY0048}
9271: \bibinfo{author}{\bibnamefont{Matsuda}, \bibfnamefont{M.}},
9272:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Fujita}},
9273:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Yamada}},
9274:   \bibinfo{author}{\bibfnamefont{R.~J.} \bibnamefont{Birgeneau}},
9275:   \bibinfo{author}{\bibfnamefont{M.~A.} \bibnamefont{Kastner}},
9276:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Hiraka}},
9277:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Endoh}},
9278:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Wakimoto}}, and
9279:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Shirane}},
9280:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. B}
9281:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{9148}.
9282: 
9283: \bibitem[{\citenamefont{Mattheiss}(1987)}]{M8728}
9284: \bibinfo{author}{\bibnamefont{Mattheiss}, \bibfnamefont{L.~F.}},
9285:   \bibinfo{year}{1987}, \bibinfo{journal}{Phys. Rev. Lett.}
9286:   \textbf{\bibinfo{volume}{58}}, \bibinfo{pages}{1028}.
9287: 
9288: \bibitem[{\citenamefont{McElroy} \emph{et~al.}(2004)\citenamefont{McElroy, Lee,
9289:   Hoffman, Lay, Hudson, Eisaki, Uchida, Lee, and Davis}}]{MLH0405}
9290: \bibinfo{author}{\bibnamefont{McElroy}, \bibfnamefont{K.}},
9291:   \bibinfo{author}{\bibfnamefont{D.-H.} \bibnamefont{Lee}},
9292:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Hoffman}},
9293:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Lay}},
9294:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Hudson}},
9295:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Eisaki}},
9296:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Uchida}},
9297:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Lee}}, and
9298:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Davis}},
9299:   \bibinfo{year}{2004}, \bibinfo{journal}{cond-mat/0404005}.
9300: 
9301: \bibitem[{\citenamefont{Miller} \emph{et~al.}(2002)\citenamefont{Miller, Kiefl,
9302:   Brewer, Sonier, Chakhalian, Dunsiger, Morris, Price, Bonn, Hardy, and
9303:   Liang}}]{MKB0202}
9304: \bibinfo{author}{\bibnamefont{Miller}, \bibfnamefont{R.~I.}},
9305:   \bibinfo{author}{\bibfnamefont{R.~F.} \bibnamefont{Kiefl}},
9306:   \bibinfo{author}{\bibfnamefont{J.~H.} \bibnamefont{Brewer}},
9307:   \bibinfo{author}{\bibfnamefont{J.~E.} \bibnamefont{Sonier}},
9308:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Chakhalian}},
9309:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Dunsiger}},
9310:   \bibinfo{author}{\bibfnamefont{G.~D.} \bibnamefont{Morris}},
9311:   \bibinfo{author}{\bibfnamefont{A.~N.} \bibnamefont{Price}},
9312:   \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Bonn}},
9313:   \bibinfo{author}{\bibfnamefont{W.~H.} \bibnamefont{Hardy}}, and
9314:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Liang}},
9315:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. Lett.}
9316:   \textbf{\bibinfo{volume}{88}}, \bibinfo{pages}{137002}.
9317: 
9318: \bibitem[{\citenamefont{Millis} \emph{et~al.}(1998)\citenamefont{Millis,
9319:   Girvin, Ioffe, and Larkin}}]{MGI9842}
9320: \bibinfo{author}{\bibnamefont{Millis}, \bibfnamefont{A.~J.}},
9321:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Girvin}},
9322:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Ioffe}}, and
9323:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Larkin}},
9324:   \bibinfo{year}{1998}, \bibinfo{journal}{J. Phys. and Chem. of Solids}
9325:   \textbf{\bibinfo{volume}{59}}, \bibinfo{pages}{1742}.
9326: 
9327: \bibitem[{\citenamefont{Millis and Monien}(1993)}]{MM9310}
9328: \bibinfo{author}{\bibnamefont{Millis}, \bibfnamefont{A.~J.}}, and
9329:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Monien}},
9330:   \bibinfo{year}{1993}, \bibinfo{journal}{Phys. Rev. Lett.}
9331:   \textbf{\bibinfo{volume}{70}}, \bibinfo{pages}{2810}.
9332: 
9333: \bibitem[{\citenamefont{Mishchenko and Nagaosa}(2004)}]{MN0402}
9334: \bibinfo{author}{\bibnamefont{Mishchenko}, \bibfnamefont{A.~S.}}, and
9335:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Nagaosa}},
9336:   \bibinfo{year}{2004}, \bibinfo{journal}{Phys. Rev. Lett.}
9337:   \textbf{\bibinfo{volume}{93}}, \bibinfo{pages}{036402}.
9338: 
9339: \bibitem[{\citenamefont{Mitrovic} \emph{et~al.}(2001)\citenamefont{Mitrovic,
9340:   Sigmund, Bachman, Eschrig, Halperin, Reyes, Kuhns, and Moulton}}]{MSB0105}
9341: \bibinfo{author}{\bibnamefont{Mitrovic}, \bibfnamefont{V.~F.}},
9342:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Sigmund}},
9343:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Bachman}},
9344:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Eschrig}},
9345:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Halperin}},
9346:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Reyes}},
9347:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Kuhns}}, and
9348:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Moulton}},
9349:   \bibinfo{year}{2001}, \bibinfo{journal}{Nature}
9350:   \textbf{\bibinfo{volume}{413}}, \bibinfo{pages}{505}.
9351: 
9352: \bibitem[{\citenamefont{Mitrovic} \emph{et~al.}(2003)\citenamefont{Mitrovic,
9353:   Sigmund, Halperin, Reyes, Kuhns, and Moulton}}]{MSH0303}
9354: \bibinfo{author}{\bibnamefont{Mitrovic}, \bibfnamefont{V.~F.}},
9355:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Sigmund}},
9356:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Halperin}},
9357:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Reyes}},
9358:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Kuhns}}, and
9359:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Moulton}},
9360:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. B}
9361:   \textbf{\bibinfo{volume}{67}}, \bibinfo{pages}{220503}.
9362: 
9363: \bibitem[{\citenamefont{Miyake} \emph{et~al.}(1986)\citenamefont{Miyake,
9364:   Schmitt-Rink, and Varma}}]{MSV8654}
9365: \bibinfo{author}{\bibnamefont{Miyake}, \bibfnamefont{K.}},
9366:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Schmitt-Rink}}, and
9367:   \bibinfo{author}{\bibfnamefont{C.~M.} \bibnamefont{Varma}},
9368:   \bibinfo{year}{1986}, \bibinfo{journal}{Phys. Rev. B}
9369:   \textbf{\bibinfo{volume}{34}}, \bibinfo{pages}{6554}.
9370: 
9371: \bibitem[{\citenamefont{Moessner and Sondhi}(2001)}]{MS0181}
9372: \bibinfo{author}{\bibnamefont{Moessner}, \bibfnamefont{R.}}, and
9373:   \bibinfo{author}{\bibfnamefont{S.~L.} \bibnamefont{Sondhi}},
9374:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. Lett.}
9375:   \textbf{\bibinfo{volume}{86}}, \bibinfo{pages}{1881}.
9376: 
9377: \bibitem[{\citenamefont{Moler} \emph{et~al.}(1994)\citenamefont{Moler, Baar,
9378:   Urbach, Liang, Hardy, and Kapitulnik}}]{MBU9444}
9379: \bibinfo{author}{\bibnamefont{Moler}, \bibfnamefont{K.~A.}},
9380:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Baar}},
9381:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Urbach}},
9382:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Liang}},
9383:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Hardy}}, and
9384:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kapitulnik}},
9385:   \bibinfo{year}{1994}, \bibinfo{journal}{Phys. Rev. Lett.}
9386:   \textbf{\bibinfo{volume}{73}}, \bibinfo{pages}{2744}.
9387: 
9388: \bibitem[{\citenamefont{Monthonx and Pines}(1993)}]{MP9369}
9389: \bibinfo{author}{\bibnamefont{Monthonx}, \bibfnamefont{P.}}, and
9390:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Pines}},
9391:   \bibinfo{year}{1993}, \bibinfo{journal}{Phys. Rev. B}
9392:   \textbf{\bibinfo{volume}{47}}, \bibinfo{pages}{6069}.
9393: 
9394: \bibitem[{\citenamefont{Mook} \emph{et~al.}(2000)\citenamefont{Mook, Dai,
9395:   Dogan, and Hunt}}]{MDD0004}
9396: \bibinfo{author}{\bibnamefont{Mook}, \bibfnamefont{H.~A.}},
9397:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Dai}},
9398:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Dogan}}, and
9399:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Hunt}}, \bibinfo{year}{2000},
9400:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{88}},
9401:   \bibinfo{pages}{097004}.
9402: 
9403: \bibitem[{\citenamefont{Mook} \emph{et~al.}(1993)\citenamefont{Mook, Yethraj,
9404:   Aeppli, Mason, and Armstrong}}]{MYA9390}
9405: \bibinfo{author}{\bibnamefont{Mook}, \bibfnamefont{H.~A.}},
9406:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Yethraj}},
9407:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Aeppli}},
9408:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Mason}}, and
9409:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Armstrong}},
9410:   \bibinfo{year}{1993}, \bibinfo{journal}{Phys. Rev. Lett.}
9411:   \textbf{\bibinfo{volume}{70}}, \bibinfo{pages}{3490}.
9412: 
9413: \bibitem[{\citenamefont{Motrunich}(2003)}]{M0308}
9414: \bibinfo{author}{\bibnamefont{Motrunich}, \bibfnamefont{O.~I.}},
9415:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. B}
9416:   \textbf{\bibinfo{volume}{67}}, \bibinfo{pages}{115108}.
9417: 
9418: \bibitem[{\citenamefont{Motrunich and Senthil}(2002)}]{MS0204}
9419: \bibinfo{author}{\bibnamefont{Motrunich}, \bibfnamefont{O.~I.}}, and
9420:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Senthil}},
9421:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. Lett.}
9422:   \textbf{\bibinfo{volume}{89}}, \bibinfo{pages}{277004}.
9423: 
9424: \bibitem[{\citenamefont{Mott}(1949)}]{M4916}
9425: \bibinfo{author}{\bibnamefont{Mott}, \bibfnamefont{N.~F.}},
9426:   \bibinfo{year}{1949}, \bibinfo{journal}{Proc. Phys. Soc.}
9427:   \textbf{\bibinfo{volume}{A 62}}, \bibinfo{pages}{416}.
9428: 
9429: \bibitem[{\citenamefont{Nagaosa}(1993)}]{N9310}
9430: \bibinfo{author}{\bibnamefont{Nagaosa}, \bibfnamefont{N.}},
9431:   \bibinfo{year}{1993}, \bibinfo{journal}{Phys. Rev. Lett.}
9432:   \textbf{\bibinfo{volume}{71}}, \bibinfo{pages}{4210}.
9433: 
9434: \bibitem[{\citenamefont{Nagaosa and Lee}(1990)}]{NL9050}
9435: \bibinfo{author}{\bibnamefont{Nagaosa}, \bibfnamefont{N.}}, and
9436:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
9437:   \bibinfo{year}{1990}, \bibinfo{journal}{Phys. Rev. Lett.}
9438:   \textbf{\bibinfo{volume}{64}}, \bibinfo{pages}{2450}.
9439: 
9440: \bibitem[{\citenamefont{Nagaosa and Lee}(1991)}]{NL9133}
9441: \bibinfo{author}{\bibnamefont{Nagaosa}, \bibfnamefont{N.}}, and
9442:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
9443:   \bibinfo{year}{1991}, \bibinfo{journal}{Phys. Rev. B}
9444:   \textbf{\bibinfo{volume}{43}}, \bibinfo{pages}{1233}.
9445: 
9446: \bibitem[{\citenamefont{Nagaosa and Lee}(1992)}]{NL9266}
9447: \bibinfo{author}{\bibnamefont{Nagaosa}, \bibfnamefont{N.}}, and
9448:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
9449:   \bibinfo{year}{1992}, \bibinfo{journal}{Phys. Rev. B}
9450:   \textbf{\bibinfo{volume}{45}}, \bibinfo{pages}{966}.
9451: 
9452: \bibitem[{\citenamefont{Nagaosa and Lee}(2000)}]{NL0066}
9453: \bibinfo{author}{\bibnamefont{Nagaosa}, \bibfnamefont{N.}}, and
9454:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
9455:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. B}
9456:   \textbf{\bibinfo{volume}{61}}, \bibinfo{pages}{9166}.
9457: 
9458: \bibitem[{\citenamefont{Nakano} \emph{et~al.}(1994)\citenamefont{Nakano, Oda,
9459:   Manabe, Momono, Miura, and Ido}}]{NOM9400}
9460: \bibinfo{author}{\bibnamefont{Nakano}, \bibfnamefont{T.}},
9461:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Oda}},
9462:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Manabe}},
9463:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Momono}},
9464:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Miura}}, and
9465:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Ido}}, \bibinfo{year}{1994},
9466:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{49}},
9467:   \bibinfo{pages}{16000}.
9468: 
9469: \bibitem[{\citenamefont{Nayak}(2000)}]{N0078}
9470: \bibinfo{author}{\bibnamefont{Nayak}, \bibfnamefont{C.}}, \bibinfo{year}{2000},
9471:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{85}},
9472:   \bibinfo{pages}{178}.
9473: 
9474: \bibitem[{\citenamefont{Nayak}(2001)}]{N0193}
9475: \bibinfo{author}{\bibnamefont{Nayak}, \bibfnamefont{C.}}, \bibinfo{year}{2001},
9476:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{86}},
9477:   \bibinfo{pages}{943}.
9478: 
9479: \bibitem[{\citenamefont{Negele and Orland}(1987)}]{NegO87}
9480: \bibinfo{author}{\bibnamefont{Negele}, \bibfnamefont{J.~W.}}, and
9481:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Orland}},
9482:   \bibinfo{year}{1987}, \emph{\bibinfo{title}{Quantum Many-Particle Systems}}
9483:   (\bibinfo{publisher}{Perseus Publishing}).
9484: 
9485: \bibitem[{\citenamefont{Nelson and Kosterlitz}(1977)}]{NK7701}
9486: \bibinfo{author}{\bibnamefont{Nelson}, \bibfnamefont{D.~R.}}, and
9487:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kosterlitz}},
9488:   \bibinfo{year}{1977}, \bibinfo{journal}{Phys. Rev. Lett.}
9489:   \textbf{\bibinfo{volume}{39}}, \bibinfo{pages}{1201}.
9490: 
9491: \bibitem[{\citenamefont{Norman}(2000)}]{N0051}
9492: \bibinfo{author}{\bibnamefont{Norman}, \bibfnamefont{M.~R.}},
9493:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. B}
9494:   \textbf{\bibinfo{volume}{61}}, \bibinfo{pages}{14751}.
9495: 
9496: \bibitem[{\citenamefont{Norman}(2001)}]{N0109}
9497: \bibinfo{author}{\bibnamefont{Norman}, \bibfnamefont{M.~R.}},
9498:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev.}
9499:   \textbf{\bibinfo{volume}{63}}, \bibinfo{pages}{092509}.
9500: 
9501: \bibitem[{\citenamefont{Norman} \emph{et~al.}(1998)\citenamefont{Norman, Ding,
9502:   Randeria, Campuzano, Yokoya, Takeuchi, Takahashi, Michiku, Kadowaki,
9503:   Guptasarma, and Hinks}}]{NDR9857}
9504: \bibinfo{author}{\bibnamefont{Norman}, \bibfnamefont{M.~R.}},
9505:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ding}},
9506:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Randeria}},
9507:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Campuzano}},
9508:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Yokoya}},
9509:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Takeuchi}},
9510:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Takahashi}},
9511:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Michiku}},
9512:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Kadowaki}},
9513:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Guptasarma}}, and
9514:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Hinks}},
9515:   \bibinfo{year}{1998}, \bibinfo{journal}{Nature}
9516:   \textbf{\bibinfo{volume}{392}}, \bibinfo{pages}{157}.
9517: 
9518: \bibitem[{\citenamefont{Norman and Pepin}(2003)}]{NP0347}
9519: \bibinfo{author}{\bibnamefont{Norman}, \bibfnamefont{M.~R.}}, and
9520:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Pepin}},
9521:   \bibinfo{year}{2003}, \bibinfo{journal}{Rep. Prog. Phys.}
9522:   \textbf{\bibinfo{volume}{66}}, \bibinfo{pages}{1547}.
9523: 
9524: \bibitem[{\citenamefont{Onufrieva and Pfeuty}(2002)}]{OP0215}
9525: \bibinfo{author}{\bibnamefont{Onufrieva}, \bibfnamefont{F.}}, and
9526:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Pfeuty}},
9527:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. B}
9528:   \textbf{\bibinfo{volume}{65}}, \bibinfo{pages}{054515}.
9529: 
9530: \bibitem[{\citenamefont{Orenstein and Millis}(2000)}]{OM0068}
9531: \bibinfo{author}{\bibnamefont{Orenstein}, \bibfnamefont{J.}}, and
9532:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Millis}},
9533:   \bibinfo{year}{2000}, \bibinfo{journal}{Science}
9534:   \textbf{\bibinfo{volume}{288}}, \bibinfo{pages}{468}.
9535: 
9536: \bibitem[{\citenamefont{Orenstein} \emph{et~al.}(1990)\citenamefont{Orenstein,
9537:   Thomas, Millis, Cooper, Rapkine, Timusk, Schneemeyer, and
9538:   Waszczak}}]{OTM9042}
9539: \bibinfo{author}{\bibnamefont{Orenstein}, \bibfnamefont{J.}},
9540:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Thomas}},
9541:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Millis}},
9542:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Cooper}},
9543:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Rapkine}},
9544:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Timusk}},
9545:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Schneemeyer}}, and
9546:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Waszczak}},
9547:   \bibinfo{year}{1990}, \bibinfo{journal}{Phys. Rev. B}
9548:   \textbf{\bibinfo{volume}{42}}, \bibinfo{pages}{6342}.
9549: 
9550: \bibitem[{\citenamefont{Oshikawa}(2003)}]{O0301}
9551: \bibinfo{author}{\bibnamefont{Oshikawa}, \bibfnamefont{M.}},
9552:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. Lett.}
9553:   \textbf{\bibinfo{volume}{91}}, \bibinfo{pages}{109901}.
9554: 
9555: \bibitem[{\citenamefont{Padilla} \emph{et~al.}(2004)\citenamefont{Padilla, Lee,
9556:   Deemm, Blumberg, Ono, Segawa, Komiya, Ando, and Basov}}]{PLD04}
9557: \bibinfo{author}{\bibnamefont{Padilla}, \bibfnamefont{W.~J.}},
9558:   \bibinfo{author}{\bibfnamefont{Y.~S.} \bibnamefont{Lee}},
9559:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Deemm}},
9560:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Blumberg}},
9561:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Ono}},
9562:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Segawa}},
9563:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Komiya}},
9564:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Ando}}, and
9565:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Basov}},
9566:   \bibinfo{year}{2004}, \bibinfo{journal}{preprint} .
9567: 
9568: \bibitem[{\citenamefont{Pailhes} \emph{et~al.}(2004)\citenamefont{Pailhes,
9569:   Sidis, Bourges, Hinkov, Ivanov, Ulrich, Regnault, and Keimer}}]{PSB0409}
9570: \bibinfo{author}{\bibnamefont{Pailhes}, \bibfnamefont{A.}},
9571:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Sidis}},
9572:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Bourges}},
9573:   \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Hinkov}},
9574:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Ivanov}},
9575:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Ulrich}},
9576:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Regnault}}, and
9577:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Keimer}},
9578:   \bibinfo{year}{2004}, \bibinfo{journal}{cond-mat/0403609}.
9579: 
9580: \bibitem[{\citenamefont{Pan} \emph{et~al.}(2000)\citenamefont{Pan, Hudson,
9581:   Gupta, Ng, Eisaki, Uchida, and David}}]{PHG0036}
9582: \bibinfo{author}{\bibnamefont{Pan}, \bibfnamefont{S.-H.}},
9583:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Hudson}},
9584:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Gupta}},
9585:   \bibinfo{author}{\bibfnamefont{K.-W.} \bibnamefont{Ng}},
9586:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Eisaki}},
9587:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Uchida}}, and
9588:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{David}},
9589:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. Lett.}
9590:   \textbf{\bibinfo{volume}{85}}, \bibinfo{pages}{1536}.
9591: 
9592: \bibitem[{\citenamefont{Paramekanti}(2002)}]{P0221}
9593: \bibinfo{author}{\bibnamefont{Paramekanti}, \bibfnamefont{A.}},
9594:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. B}
9595:   \textbf{\bibinfo{volume}{65}}, \bibinfo{pages}{104521}.
9596: 
9597: \bibitem[{\citenamefont{Paramekanti}
9598:   \emph{et~al.}(2000)\citenamefont{Paramekanti, Randeria, Ramakrishnan, and
9599:   Mandal}}]{PRR0086}
9600: \bibinfo{author}{\bibnamefont{Paramekanti}, \bibfnamefont{A.}},
9601:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Randeria}},
9602:   \bibinfo{author}{\bibfnamefont{T.~V.} \bibnamefont{Ramakrishnan}}, and
9603:   \bibinfo{author}{\bibfnamefont{S.~S.} \bibnamefont{Mandal}},
9604:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. B}
9605:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{6786}.
9606: 
9607: \bibitem[{\citenamefont{Paramekanti}
9608:   \emph{et~al.}(2001)\citenamefont{Paramekanti, Randeria, and
9609:   Trivedi}}]{PRT0102}
9610: \bibinfo{author}{\bibnamefont{Paramekanti}, \bibfnamefont{A.}},
9611:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Randeria}}, and
9612:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Trivedi}},
9613:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. Lett.}
9614:   \textbf{\bibinfo{volume}{87}}, \bibinfo{pages}{217002}.
9615: 
9616: \bibitem[{\citenamefont{Paramekanti}
9617:   \emph{et~al.}(2004{\natexlab{a}})\citenamefont{Paramekanti, Randeria, and
9618:   Trivedi}}]{PRT0453}
9619: \bibinfo{author}{\bibnamefont{Paramekanti}, \bibfnamefont{A.}},
9620:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Randeria}}, and
9621:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Trivedi}},
9622:   \bibinfo{year}{2004}{\natexlab{a}}, \bibinfo{journal}{cond-mat/0405353}.
9623: 
9624: \bibitem[{\citenamefont{Paramekanti}
9625:   \emph{et~al.}(2004{\natexlab{b}})\citenamefont{Paramekanti, Randeria, and
9626:   Trivedi}}]{PRT0404}
9627: \bibinfo{author}{\bibnamefont{Paramekanti}, \bibfnamefont{A.}},
9628:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Randeria}}, and
9629:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Trivedi}},
9630:   \bibinfo{year}{2004}{\natexlab{b}}, \bibinfo{journal}{Phys. Rev. B}
9631:   \textbf{\bibinfo{volume}{70}}, \bibinfo{pages}{054504}.
9632: 
9633: \bibitem[{\citenamefont{Pavarini} \emph{et~al.}(2001)\citenamefont{Pavarini,
9634:   Dasgupta, Saha-Dasgupta, Jasperson, and Andersen}}]{PDS0103}
9635: \bibinfo{author}{\bibnamefont{Pavarini}, \bibfnamefont{E.}},
9636:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Dasgupta}},
9637:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Saha-Dasgupta}},
9638:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Jasperson}}, and
9639:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Andersen}},
9640:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. Lett.}
9641:   \textbf{\bibinfo{volume}{87}}, \bibinfo{pages}{047003}.
9642: 
9643: \bibitem[{\citenamefont{Polyakov}(1978)}]{P7877}
9644: \bibinfo{author}{\bibnamefont{Polyakov}, \bibfnamefont{A.}},
9645:   \bibinfo{year}{1978}, \bibinfo{journal}{Phys. Lett. B}
9646:   \textbf{\bibinfo{volume}{72}}, \bibinfo{pages}{477}.
9647: 
9648: \bibitem[{\citenamefont{Polyakov}(1975)}]{P7582}
9649: \bibinfo{author}{\bibnamefont{Polyakov}, \bibfnamefont{A.~M.}},
9650:   \bibinfo{year}{1975}, \bibinfo{journal}{Phys. Lett. B}
9651:   \textbf{\bibinfo{volume}{59}}, \bibinfo{pages}{82}.
9652: 
9653: \bibitem[{\citenamefont{Polyakov}(1977)}]{P7729}
9654: \bibinfo{author}{\bibnamefont{Polyakov}, \bibfnamefont{A.~M.}},
9655:   \bibinfo{year}{1977}, \bibinfo{journal}{Nucl. Phys. B}
9656:   \textbf{\bibinfo{volume}{120}}, \bibinfo{pages}{429}.
9657: 
9658: \bibitem[{\citenamefont{Polyakov}(1979)}]{P7947}
9659: \bibinfo{author}{\bibnamefont{Polyakov}, \bibfnamefont{A.~M.}},
9660:   \bibinfo{year}{1979}, \bibinfo{journal}{Phys. Lett. B}
9661:   \textbf{\bibinfo{volume}{82}}, \bibinfo{pages}{247}.
9662: 
9663: \bibitem[{\citenamefont{Polyakov}(1987)}]{Pol87}
9664: \bibinfo{author}{\bibnamefont{Polyakov}, \bibfnamefont{A.~M.}},
9665:   \bibinfo{year}{1987}, \emph{\bibinfo{title}{Gauge Fields and Strings}}
9666:   (\bibinfo{publisher}{Harwood Academic Publishers},
9667:   \bibinfo{address}{London}).
9668: 
9669: \bibitem[{\citenamefont{Proust} \emph{et~al.}(2002)\citenamefont{Proust,
9670:   Boaknin, Hill, Taillefer, and MacKenzie}}]{PBH0203}
9671: \bibinfo{author}{\bibnamefont{Proust}, \bibfnamefont{C.}},
9672:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Boaknin}},
9673:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Hill}},
9674:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Taillefer}}, and
9675:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{MacKenzie}},
9676:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. Lett.}
9677:   \textbf{\bibinfo{volume}{89}}, \bibinfo{pages}{147003}.
9678: 
9679: \bibitem[{\citenamefont{Rantner and Wen}(2001{\natexlab{a}})}]{RW0171}
9680: \bibinfo{author}{\bibnamefont{Rantner}, \bibfnamefont{W.}}, and
9681:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
9682:   \bibinfo{year}{2001}{\natexlab{a}}, \bibinfo{journal}{Phys. Rev. Lett.}
9683:   \textbf{\bibinfo{volume}{86}}, \bibinfo{pages}{3871}.
9684: 
9685: \bibitem[{\citenamefont{Rantner and Wen}(2001{\natexlab{b}})}]{RW0140}
9686: \bibinfo{author}{\bibnamefont{Rantner}, \bibfnamefont{W.}}, and
9687:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
9688:   \bibinfo{year}{2001}{\natexlab{b}}, \bibinfo{journal}{cond-mat/0105540}.
9689: 
9690: \bibitem[{\citenamefont{Rantner and Wen}(2002)}]{RWspin}
9691: \bibinfo{author}{\bibnamefont{Rantner}, \bibfnamefont{W.}}, and
9692:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
9693:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. B}
9694:   \textbf{\bibinfo{volume}{66}}, \bibinfo{pages}{144501}.
9695: 
9696: \bibitem[{\citenamefont{Read and Chakraborty}(1989)}]{RC8933}
9697: \bibinfo{author}{\bibnamefont{Read}, \bibfnamefont{N.}}, and
9698:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Chakraborty}},
9699:   \bibinfo{year}{1989}, \bibinfo{journal}{Phys. Rev. B}
9700:   \textbf{\bibinfo{volume}{40}}, \bibinfo{pages}{7133}.
9701: 
9702: \bibitem[{\citenamefont{Read and Sachdev}(1990)}]{RS9068}
9703: \bibinfo{author}{\bibnamefont{Read}, \bibfnamefont{N.}}, and
9704:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Sachdev}},
9705:   \bibinfo{year}{1990}, \bibinfo{journal}{Phys. Rev. B}
9706:   \textbf{\bibinfo{volume}{42}}, \bibinfo{pages}{4568}.
9707: 
9708: \bibitem[{\citenamefont{Read and Sachdev}(1991)}]{RS9173}
9709: \bibinfo{author}{\bibnamefont{Read}, \bibfnamefont{N.}}, and
9710:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Sachdev}},
9711:   \bibinfo{year}{1991}, \bibinfo{journal}{Phys. Rev. Lett.}
9712:   \textbf{\bibinfo{volume}{66}}, \bibinfo{pages}{1773}.
9713: 
9714: \bibitem[{\citenamefont{Reizer}(1989)}]{R8902}
9715: \bibinfo{author}{\bibnamefont{Reizer}, \bibfnamefont{M.}},
9716:   \bibinfo{year}{1989}, \bibinfo{journal}{Phys. Rev. B}
9717:   \textbf{\bibinfo{volume}{39}}, \bibinfo{pages}{1602}.
9718: 
9719: \bibitem[{\citenamefont{Rokhsar and Kivelson}(1988)}]{RK8876}
9720: \bibinfo{author}{\bibnamefont{Rokhsar}, \bibfnamefont{D.~S.}}, and
9721:   \bibinfo{author}{\bibfnamefont{S.~A.} \bibnamefont{Kivelson}},
9722:   \bibinfo{year}{1988}, \bibinfo{journal}{Phys. Rev. Lett.}
9723:   \textbf{\bibinfo{volume}{61}}, \bibinfo{pages}{2376}.
9724: 
9725: \bibitem[{\citenamefont{Ronning} \emph{et~al.}(2003)\citenamefont{Ronning,
9726:   Sasagawa, Kohsaka, Shen, Damascelli, Kim, Yoshida, Armitage, Lu, Feng,
9727:   Miller, Takagi} \emph{et~al.}}]{RSK0301}
9728: \bibinfo{author}{\bibnamefont{Ronning}, \bibfnamefont{F.}},
9729:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Sasagawa}},
9730:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Kohsaka}},
9731:   \bibinfo{author}{\bibfnamefont{K.~M.} \bibnamefont{Shen}},
9732:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Damascelli}},
9733:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Kim}},
9734:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Yoshida}},
9735:   \bibinfo{author}{\bibfnamefont{N.~P.} \bibnamefont{Armitage}},
9736:   \bibinfo{author}{\bibfnamefont{D.~H.} \bibnamefont{Lu}},
9737:   \bibinfo{author}{\bibfnamefont{D.~L.} \bibnamefont{Feng}},
9738:   \bibinfo{author}{\bibfnamefont{L.~L.} \bibnamefont{Miller}},
9739:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Takagi}}, \emph{et~al.},
9740:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. B}
9741:   \textbf{\bibinfo{volume}{67}}, \bibinfo{pages}{165101}.
9742: 
9743: \bibitem[{Ronning \emph{et~al.}(1998)\citenamefont{Ronning}
9744:   \emph{et~al.}}]{Ro9867}
9745: \bibinfo{author}{\bibnamefont{Ronning}, \bibfnamefont{F.}}, \emph{et~al.},
9746:   \bibinfo{year}{1998}, \bibinfo{journal}{Science}
9747:   \textbf{\bibinfo{volume}{282}}, \bibinfo{pages}{2067}.
9748: 
9749: \bibitem[{\citenamefont{Rossat-Mignod}
9750:   \emph{et~al.}(1991)\citenamefont{Rossat-Mignod, Regnault, Vettier, Bourges,
9751:   Burlet, Bossy, Henry, and Lapertot}}]{RRV9186}
9752: \bibinfo{author}{\bibnamefont{Rossat-Mignod}, \bibfnamefont{J.}},
9753:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Regnault}},
9754:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Vettier}},
9755:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Bourges}},
9756:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Burlet}},
9757:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Bossy}},
9758:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Henry}}, and
9759:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Lapertot}},
9760:   \bibinfo{year}{1991}, \bibinfo{journal}{Physica C}
9761:   \textbf{\bibinfo{volume}{185}}, \bibinfo{pages}{86}.
9762: 
9763: \bibitem[{\citenamefont{Runge}(1992)}]{R9292}
9764: \bibinfo{author}{\bibnamefont{Runge}, \bibfnamefont{K.~J.}},
9765:   \bibinfo{year}{1992}, \bibinfo{journal}{Phys. Rev. B}
9766:   \textbf{\bibinfo{volume}{45}}, \bibinfo{pages}{12292}.
9767: 
9768: \bibitem[{\citenamefont{Sachdev}(1992)}]{S9289}
9769: \bibinfo{author}{\bibnamefont{Sachdev}, \bibfnamefont{S.}},
9770:   \bibinfo{year}{1992}, \bibinfo{journal}{Phys. Rev. B}
9771:   \textbf{\bibinfo{volume}{45}}, \bibinfo{pages}{389}.
9772: 
9773: \bibitem[{\citenamefont{Sachdev}(2003)}]{S0313}
9774: \bibinfo{author}{\bibnamefont{Sachdev}, \bibfnamefont{S.}},
9775:   \bibinfo{year}{2003}, \bibinfo{journal}{Rev. Mod. Phys.}
9776:   \textbf{\bibinfo{volume}{75}}, \bibinfo{pages}{913}.
9777: 
9778: \bibitem[{\citenamefont{Sandvik} \emph{et~al.}(1997)\citenamefont{Sandvik,
9779:   Dagotto, and Scalapino}}]{SDS9701}
9780: \bibinfo{author}{\bibnamefont{Sandvik}, \bibfnamefont{A.~W.}},
9781:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Dagotto}}, and
9782:   \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Scalapino}},
9783:   \bibinfo{year}{1997}, \bibinfo{journal}{Phys. Rev. B}
9784:   \textbf{\bibinfo{volume}{56}}, \bibinfo{pages}{11701}.
9785: 
9786: \bibitem[{\citenamefont{Santander-Syro}
9787:   \emph{et~al.}(2002)\citenamefont{Santander-Syro, Lobo, Bontemps,
9788:   Konstantinovic, Li, and Raffy}}]{SLB0205}
9789: \bibinfo{author}{\bibnamefont{Santander-Syro}, \bibfnamefont{A.~F.}},
9790:   \bibinfo{author}{\bibfnamefont{R.~P. S.~M.} \bibnamefont{Lobo}},
9791:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Bontemps}},
9792:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Konstantinovic}},
9793:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Li}}, and
9794:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Raffy}},
9795:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. Lett.}
9796:   \textbf{\bibinfo{volume}{88}}, \bibinfo{pages}{097005}.
9797: 
9798: \bibitem[{\citenamefont{Savici} \emph{et~al.}(2002)\citenamefont{Savici,
9799:   Fudamoto, Gat, Ito, Larkin, Uemura, Luke, Kojima, Lee, Kastner, Birgeneau,
9800:   and Yamada}}]{SFG0224}
9801: \bibinfo{author}{\bibnamefont{Savici}, \bibfnamefont{A.~T.}},
9802:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Fudamoto}},
9803:   \bibinfo{author}{\bibfnamefont{I.~M.} \bibnamefont{Gat}},
9804:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Ito}},
9805:   \bibinfo{author}{\bibfnamefont{M.~I.} \bibnamefont{Larkin}},
9806:   \bibinfo{author}{\bibfnamefont{Y.~J.} \bibnamefont{Uemura}},
9807:   \bibinfo{author}{\bibfnamefont{G.~M.} \bibnamefont{Luke}},
9808:   \bibinfo{author}{\bibfnamefont{K.~M.} \bibnamefont{Kojima}},
9809:   \bibinfo{author}{\bibfnamefont{Y.~S.} \bibnamefont{Lee}},
9810:   \bibinfo{author}{\bibfnamefont{M.~A.} \bibnamefont{Kastner}},
9811:   \bibinfo{author}{\bibfnamefont{R.~J.} \bibnamefont{Birgeneau}}, and
9812:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Yamada}},
9813:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. B}
9814:   \textbf{\bibinfo{volume}{66}}, \bibinfo{pages}{014524}.
9815: 
9816: \bibitem[{\citenamefont{Savit}(1980)}]{S8053}
9817: \bibinfo{author}{\bibnamefont{Savit}, \bibfnamefont{R.}}, \bibinfo{year}{1980},
9818:   \bibinfo{journal}{Rev. Mod. Phys.} \textbf{\bibinfo{volume}{52}},
9819:   \bibinfo{pages}{453}.
9820: 
9821: \bibitem[{\citenamefont{Scalapino} \emph{et~al.}(1986)\citenamefont{Scalapino,
9822:   Loh, and Hirsch}}]{SLH8690}
9823: \bibinfo{author}{\bibnamefont{Scalapino}, \bibfnamefont{D.~J.}},
9824:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Loh}}, and
9825:   \bibinfo{author}{\bibfnamefont{J.~E.} \bibnamefont{Hirsch}},
9826:   \bibinfo{year}{1986}, \bibinfo{journal}{Phys. Rev. B}
9827:   \textbf{\bibinfo{volume}{34}}, \bibinfo{pages}{8190}.
9828: 
9829: \bibitem[{\citenamefont{Scalapino} \emph{et~al.}(1987)\citenamefont{Scalapino,
9830:   Loh, and Hirsch}}]{SLH8794}
9831: \bibinfo{author}{\bibnamefont{Scalapino}, \bibfnamefont{D.~J.}},
9832:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Loh}}, and
9833:   \bibinfo{author}{\bibfnamefont{J.~E.} \bibnamefont{Hirsch}},
9834:   \bibinfo{year}{1987}, \bibinfo{journal}{Phys. Rev. B}
9835:   \textbf{\bibinfo{volume}{35}}, \bibinfo{pages}{6694}.
9836: 
9837: \bibitem[{\citenamefont{Senthil and Fisher}(1999)}]{SF9993}
9838: \bibinfo{author}{\bibnamefont{Senthil}, \bibfnamefont{T.}}, and
9839:   \bibinfo{author}{\bibfnamefont{M.~P.~A.} \bibnamefont{Fisher}},
9840:   \bibinfo{year}{1999}, \bibinfo{journal}{Phys. Rev. B}
9841:   \textbf{\bibinfo{volume}{60}}, \bibinfo{pages}{6893}.
9842: 
9843: \bibitem[{\citenamefont{Senthil and Fisher}(2000)}]{SF0050}
9844: \bibinfo{author}{\bibnamefont{Senthil}, \bibfnamefont{T.}}, and
9845:   \bibinfo{author}{\bibfnamefont{M.~P.~A.} \bibnamefont{Fisher}},
9846:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. B}
9847:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{7850}.
9848: 
9849: \bibitem[{\citenamefont{Senthil and Fisher}(2001{\natexlab{a}})}]{SF0119}
9850: \bibinfo{author}{\bibnamefont{Senthil}, \bibfnamefont{T.}}, and
9851:   \bibinfo{author}{\bibfnamefont{M.~P.~A.} \bibnamefont{Fisher}},
9852:   \bibinfo{year}{2001}{\natexlab{a}}, \bibinfo{journal}{J. Phys. A}
9853:   \textbf{\bibinfo{volume}{34}}, \bibinfo{pages}{L119}.
9854: 
9855: \bibitem[{\citenamefont{Senthil and Fisher}(2001{\natexlab{b}})}]{SF0192}
9856: \bibinfo{author}{\bibnamefont{Senthil}, \bibfnamefont{T.}}, and
9857:   \bibinfo{author}{\bibfnamefont{M.~P.~A.} \bibnamefont{Fisher}},
9858:   \bibinfo{year}{2001}{\natexlab{b}}, \bibinfo{journal}{Phys. Rev. Lett.}
9859:   \textbf{\bibinfo{volume}{86}}, \bibinfo{pages}{292}.
9860: 
9861: \bibitem[{\citenamefont{Senthil and Lee}(2004)}]{SL0466}
9862: \bibinfo{author}{\bibnamefont{Senthil}, \bibfnamefont{T.}}, and
9863:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
9864:   \bibinfo{year}{2004}, \bibinfo{journal}{cond-mat/0406066}.
9865: 
9866: \bibitem[{\citenamefont{Senthil and Motrunich}(2002)}]{SM0204}
9867: \bibinfo{author}{\bibnamefont{Senthil}, \bibfnamefont{T.}}, and
9868:   \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Motrunich}},
9869:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. B}
9870:   \textbf{\bibinfo{volume}{66}}, \bibinfo{pages}{205104}.
9871: 
9872: \bibitem[{\citenamefont{Senthil} \emph{et~al.}(2004)\citenamefont{Senthil,
9873:   Vishwanath, Balents, Sachdev, and Fisher}}]{SVB0490}
9874: \bibinfo{author}{\bibnamefont{Senthil}, \bibfnamefont{T.}},
9875:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Vishwanath}},
9876:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Balents}},
9877:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Sachdev}}, and
9878:   \bibinfo{author}{\bibfnamefont{M.~P.~A.} \bibnamefont{Fisher}},
9879:   \bibinfo{year}{2004}, \bibinfo{journal}{Science}
9880:   \textbf{\bibinfo{volume}{303}}, \bibinfo{pages}{1490}.
9881: 
9882: \bibitem[{\citenamefont{Shen} \emph{et~al.}(2004)\citenamefont{Shen, Ronning,
9883:   Lu, Lee, Ingle, Meevasana, Baumberger, Damascelli, Armitage, Miller, Kohsaka,
9884:   Azuma} \emph{et~al.}}]{SRL0402}
9885: \bibinfo{author}{\bibnamefont{Shen}, \bibfnamefont{K.~M.}},
9886:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Ronning}},
9887:   \bibinfo{author}{\bibfnamefont{D.~H.} \bibnamefont{Lu}},
9888:   \bibinfo{author}{\bibfnamefont{W.~S.} \bibnamefont{Lee}},
9889:   \bibinfo{author}{\bibfnamefont{N.~J.~C.} \bibnamefont{Ingle}},
9890:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Meevasana}},
9891:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Baumberger}},
9892:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Damascelli}},
9893:   \bibinfo{author}{\bibfnamefont{N.~P.} \bibnamefont{Armitage}},
9894:   \bibinfo{author}{\bibfnamefont{L.~L.} \bibnamefont{Miller}},
9895:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Kohsaka}},
9896:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Azuma}}, \emph{et~al.},
9897:   \bibinfo{year}{2004}, \bibinfo{journal}{cond-mat/0407002}.
9898: 
9899: \bibitem[{\citenamefont{Shih} \emph{et~al.}(1998)\citenamefont{Shih, Chen, Lin,
9900:   and Lee}}]{SCL9894}
9901: \bibinfo{author}{\bibnamefont{Shih}, \bibfnamefont{C.~T.}},
9902:   \bibinfo{author}{\bibfnamefont{Y.~C.} \bibnamefont{Chen}},
9903:   \bibinfo{author}{\bibfnamefont{H.~Q.} \bibnamefont{Lin}}, and
9904:   \bibinfo{author}{\bibfnamefont{T.-K.} \bibnamefont{Lee}},
9905:   \bibinfo{year}{1998}, \bibinfo{journal}{Phys. rev. Lett.}
9906:   \textbf{\bibinfo{volume}{81}}, \bibinfo{pages}{1294}.
9907: 
9908: \bibitem[{\citenamefont{Shih} \emph{et~al.}(2004)\citenamefont{Shih, Lee, Eder,
9909:   Mou, and Chen}}]{SLE0402}
9910: \bibinfo{author}{\bibnamefont{Shih}, \bibfnamefont{C.~T.}},
9911:   \bibinfo{author}{\bibfnamefont{T.-K.} \bibnamefont{Lee}},
9912:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Eder}},
9913:   \bibinfo{author}{\bibfnamefont{C.-Y.} \bibnamefont{Mou}}, and
9914:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Chen}}, \bibinfo{year}{2004},
9915:   \bibinfo{journal}{superconducting tendency decreases \j Phys. Rev. Lett.}
9916:   \textbf{\bibinfo{volume}{92}}, \bibinfo{pages}{227002}.
9917: 
9918: \bibitem[{\citenamefont{Shraiman and Siggia}(1988)}]{SS8867}
9919: \bibinfo{author}{\bibnamefont{Shraiman}, \bibfnamefont{B.~I.}}, and
9920:   \bibinfo{author}{\bibfnamefont{E.~D.} \bibnamefont{Siggia}},
9921:   \bibinfo{year}{1988}, \bibinfo{journal}{Phys. Rev. Lett.}
9922:   \textbf{\bibinfo{volume}{61}}, \bibinfo{pages}{467}.
9923: 
9924: \bibitem[{\citenamefont{Si} \emph{et~al.}(1993)\citenamefont{Si, Zha, Levin,
9925:   and Lu}}]{SZL9355}
9926: \bibinfo{author}{\bibnamefont{Si}, \bibfnamefont{Q.}},
9927:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Zha}},
9928:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Levin}}, and
9929:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Lu}}, \bibinfo{year}{1993},
9930:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{47}},
9931:   \bibinfo{pages}{9055}.
9932: 
9933: \bibitem[{\citenamefont{Singh and Ghosh}(2002)}]{SG0214}
9934: \bibinfo{author}{\bibnamefont{Singh}, \bibfnamefont{A.}}, and
9935:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ghosh}},
9936:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. B}
9937:   \textbf{\bibinfo{volume}{65}}, \bibinfo{pages}{134414}.
9938: 
9939: \bibitem[{\citenamefont{Sorella} \emph{et~al.}(2002)\citenamefont{Sorella,
9940:   Martins, Bocca, Gazza, Capriotti, Parola, , and Dagotto}}]{SMB0202}
9941: \bibinfo{author}{\bibnamefont{Sorella}, \bibfnamefont{S.}},
9942:   \bibinfo{author}{\bibfnamefont{G.~B.} \bibnamefont{Martins}},
9943:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Bocca}},
9944:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Gazza}},
9945:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Capriotti}},
9946:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Parola}}, , and
9947:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Dagotto}},
9948:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. Lett.}
9949:   \textbf{\bibinfo{volume}{88}}, \bibinfo{pages}{117002}.
9950: 
9951: \bibitem[{\citenamefont{Stajis} \emph{et~al.}(2003)\citenamefont{Stajis,
9952:   Iyengar, Levin, Boyce, and Lemberger}}]{SIL0320}
9953: \bibinfo{author}{\bibnamefont{Stajis}, \bibfnamefont{J.}},
9954:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Iyengar}},
9955:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Levin}},
9956:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Boyce}}, and
9957:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Lemberger}},
9958:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. B}
9959:   \textbf{\bibinfo{volume}{68}}, \bibinfo{pages}{024520}.
9960: 
9961: \bibitem[{\citenamefont{Stock}
9962:   \emph{et~al.}(2004{\natexlab{a}})\citenamefont{Stock, Buyers, Cowley, Clegg,
9963:   Coldea, Frost, Liang, Peets, Bonn, Hardy, and Birgeneau}}]{SBC0471}
9964: \bibinfo{author}{\bibnamefont{Stock}, \bibfnamefont{C.}},
9965:   \bibinfo{author}{\bibfnamefont{W.~J.~L.} \bibnamefont{Buyers}},
9966:   \bibinfo{author}{\bibfnamefont{R.~A.} \bibnamefont{Cowley}},
9967:   \bibinfo{author}{\bibfnamefont{P.~S.} \bibnamefont{Clegg}},
9968:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Coldea}},
9969:   \bibinfo{author}{\bibfnamefont{C.~D.} \bibnamefont{Frost}},
9970:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Liang}},
9971:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Peets}},
9972:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Bonn}},
9973:   \bibinfo{author}{\bibfnamefont{W.~N.} \bibnamefont{Hardy}}, and
9974:   \bibinfo{author}{\bibfnamefont{R.~J.} \bibnamefont{Birgeneau}},
9975:   \bibinfo{year}{2004}{\natexlab{a}}, \bibinfo{journal}{cond-mat/0408071}.
9976: 
9977: \bibitem[{\citenamefont{Stock}
9978:   \emph{et~al.}(2004{\natexlab{b}})\citenamefont{Stock, Buyers, Liang, Peets,
9979:   Tun, Bonn, Hardy, and Birgeneau}}]{SBL0402}
9980: \bibinfo{author}{\bibnamefont{Stock}, \bibfnamefont{C.}},
9981:   \bibinfo{author}{\bibfnamefont{W.~J.~L.} \bibnamefont{Buyers}},
9982:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Liang}},
9983:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Peets}},
9984:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Tun}},
9985:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Bonn}},
9986:   \bibinfo{author}{\bibfnamefont{W.~N.} \bibnamefont{Hardy}}, and
9987:   \bibinfo{author}{\bibfnamefont{R.~J.} \bibnamefont{Birgeneau}},
9988:   \bibinfo{year}{2004}{\natexlab{b}}, \bibinfo{journal}{Phys. Rev. B}
9989:   \textbf{\bibinfo{volume}{69}}, \bibinfo{pages}{014502}.
9990: 
9991: \bibitem[{\citenamefont{Sulewsky} \emph{et~al.}(1990)\citenamefont{Sulewsky,
9992:   Fleury, Lyons, Cheong, and Fisk}}]{SFL9025}
9993: \bibinfo{author}{\bibnamefont{Sulewsky}, \bibfnamefont{P.~E.}},
9994:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Fleury}},
9995:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Lyons}},
9996:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Cheong}}, and
9997:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Fisk}}, \bibinfo{year}{1990},
9998:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{41}},
9999:   \bibinfo{pages}{225}.
10000: 
10001: \bibitem[{\citenamefont{Susskind}(1979)}]{S7910}
10002: \bibinfo{author}{\bibnamefont{Susskind}, \bibfnamefont{L.}},
10003:   \bibinfo{year}{1979}, \bibinfo{journal}{Phys. Rev. D}
10004:   \textbf{\bibinfo{volume}{20}}, \bibinfo{pages}{2610}.
10005: 
10006: \bibitem[{\citenamefont{Sutherland}
10007:   \emph{et~al.}(2003)\citenamefont{Sutherland, Hawthorn, Hill, Ronning,
10008:   Wakimoto, Zhang, Proust, Boaknin, Lupien, Taillefer, Liang, Bonn}
10009:   \emph{et~al.}}]{SHH0320}
10010: \bibinfo{author}{\bibnamefont{Sutherland}},
10011:   \bibinfo{author}{\bibfnamefont{D.~G.} \bibnamefont{Hawthorn}},
10012:   \bibinfo{author}{\bibfnamefont{R.~W.} \bibnamefont{Hill}},
10013:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Ronning}},
10014:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Wakimoto}},
10015:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Zhang}},
10016:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Proust}},
10017:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Boaknin}},
10018:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Lupien}},
10019:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Taillefer}},
10020:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Liang}},
10021:   \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Bonn}}, \emph{et~al.},
10022:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. B}
10023:   \textbf{\bibinfo{volume}{67}}, \bibinfo{pages}{174520}.
10024: 
10025: \bibitem[{\citenamefont{Sutherland}
10026:   \emph{et~al.}(2004)\citenamefont{Sutherland, Hawthorn, Hill, Ronning,
10027:   Tanatar, Paglione, Boaknin, Zhang, Taillefer, DeBenedictis, Liang, Bonn}
10028:   \emph{et~al.}}]{SHH04}
10029: \bibinfo{author}{\bibnamefont{Sutherland}, \bibfnamefont{M.}},
10030:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Hawthorn}},
10031:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Hill}},
10032:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Ronning}},
10033:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Tanatar}},
10034:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Paglione}},
10035:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Boaknin}},
10036:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Zhang}},
10037:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Taillefer}},
10038:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{DeBenedictis}},
10039:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Liang}},
10040:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Bonn}}, \emph{et~al.},
10041:   \bibinfo{year}{2004}, \bibinfo{journal}{to be published} .
10042: 
10043: \bibitem[{\citenamefont{Suzumura} \emph{et~al.}(1988)\citenamefont{Suzumura,
10044:   Hasegawa, and Fukuyama}}]{SHF8868}
10045: \bibinfo{author}{\bibnamefont{Suzumura}, \bibfnamefont{Y.}},
10046:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Hasegawa}}, and
10047:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Fukuyama}},
10048:   \bibinfo{year}{1988}, \bibinfo{journal}{J. Phys. Soc. Jpn.}
10049:   \textbf{\bibinfo{volume}{57}}, \bibinfo{pages}{2768}.
10050: 
10051: \bibitem[{\citenamefont{Svetitsky}(1986)}]{S861}
10052: \bibinfo{author}{\bibnamefont{Svetitsky}, \bibfnamefont{B.}},
10053:   \bibinfo{year}{1986}, \bibinfo{journal}{Phys. Rep.}
10054:   \textbf{\bibinfo{volume}{132}}, \bibinfo{pages}{1}.
10055: 
10056: \bibitem[{\citenamefont{Taguchi} \emph{et~al.}(2001)\citenamefont{Taguchi,
10057:   Oohara, Yoshizawa, Nagaosa, and Tokura}}]{TOY0173}
10058: \bibinfo{author}{\bibnamefont{Taguchi}, \bibfnamefont{Y.}},
10059:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Oohara}},
10060:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Yoshizawa}},
10061:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Nagaosa}}, and
10062:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Tokura}},
10063:   \bibinfo{year}{2001}, \bibinfo{journal}{Science}
10064:   \textbf{\bibinfo{volume}{291}}, \bibinfo{pages}{2573}.
10065: 
10066: \bibitem[{\citenamefont{Taillefer} \emph{et~al.}(1997)\citenamefont{Taillefer,
10067:   Lussier, Gagnon, Behnia, and Aubin}}]{TLG9783}
10068: \bibinfo{author}{\bibnamefont{Taillefer}, \bibfnamefont{L.}},
10069:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Lussier}},
10070:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Gagnon}},
10071:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Behnia}}, and
10072:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Aubin}},
10073:   \bibinfo{year}{1997}, \bibinfo{journal}{Phys. Rev. Lett.}
10074:   \textbf{\bibinfo{volume}{79}}, \bibinfo{pages}{483}.
10075: 
10076: \bibitem[{\citenamefont{Takigawa} \emph{et~al.}(1993)\citenamefont{Takigawa,
10077:   Hults, and Smith}}]{THS9350}
10078: \bibinfo{author}{\bibnamefont{Takigawa}, \bibfnamefont{M.}},
10079:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Hults}}, and
10080:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Smith}},
10081:   \bibinfo{year}{1993}, \bibinfo{journal}{Phys. Rev. Lett.}
10082:   \textbf{\bibinfo{volume}{71}}, \bibinfo{pages}{2650}.
10083: 
10084: \bibitem[{\citenamefont{Takigawa} \emph{et~al.}(1991)\citenamefont{Takigawa,
10085:   Reyes, Hammel, Thompson, Heffner, Fisk, and Ott}}]{TRH9147}
10086: \bibinfo{author}{\bibnamefont{Takigawa}, \bibfnamefont{M.}},
10087:   \bibinfo{author}{\bibfnamefont{A.~P.} \bibnamefont{Reyes}},
10088:   \bibinfo{author}{\bibfnamefont{P.~C.} \bibnamefont{Hammel}},
10089:   \bibinfo{author}{\bibfnamefont{J.~D.} \bibnamefont{Thompson}},
10090:   \bibinfo{author}{\bibfnamefont{R.~H.} \bibnamefont{Heffner}},
10091:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Fisk}}, and
10092:   \bibinfo{author}{\bibfnamefont{K.~C.} \bibnamefont{Ott}},
10093:   \bibinfo{year}{1991}, \bibinfo{journal}{Phys. Rev. B}
10094:   \textbf{\bibinfo{volume}{43}}, \bibinfo{pages}{247}.
10095: 
10096: \bibitem[{\citenamefont{Tallon and Loram}(2000)}]{TL0053}
10097: \bibinfo{author}{\bibnamefont{Tallon}, \bibfnamefont{J.~L.}}, and
10098:   \bibinfo{author}{\bibfnamefont{J.~W.} \bibnamefont{Loram}},
10099:   \bibinfo{year}{2000}, \bibinfo{journal}{Physica C}
10100:   \textbf{\bibinfo{volume}{349}}, \bibinfo{pages}{53}.
10101: 
10102: \bibitem[{\citenamefont{Tanamoto} \emph{et~al.}(1993)\citenamefont{Tanamoto,
10103:   Konno, and Fukuyama}}]{TKF9355}
10104: \bibinfo{author}{\bibnamefont{Tanamoto}, \bibfnamefont{T.}},
10105:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Konno}}, and
10106:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Fukuyama}},
10107:   \bibinfo{year}{1993}, \bibinfo{journal}{J. Phys. Soc. Jpn.} ,
10108:   \bibinfo{pages}{1455}.
10109: 
10110: \bibitem[{\citenamefont{Tchernyshyov}
10111:   \emph{et~al.}(2001)\citenamefont{Tchernyshyov, Norman, and
10112:   Chubukov}}]{TNC0107}
10113: \bibinfo{author}{\bibnamefont{Tchernyshyov}, \bibfnamefont{O.}},
10114:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Norman}}, and
10115:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Chubukov}},
10116:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. B}
10117:   \textbf{\bibinfo{volume}{63}}, \bibinfo{pages}{144507}.
10118: 
10119: \bibitem[{\citenamefont{Timusk and Statt}(1999)}]{TS9961}
10120: \bibinfo{author}{\bibnamefont{Timusk}, \bibfnamefont{T.}}, and
10121:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Statt}},
10122:   \bibinfo{year}{1999}, \bibinfo{journal}{Rep. Prog. Phys.}
10123:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{61}.
10124: 
10125: \bibitem[{Tohyama \emph{et~al.}(2000)\citenamefont{Tohyama}
10126:   \emph{et~al.}}]{T009}
10127: \bibinfo{author}{\bibnamefont{Tohyama}, \bibfnamefont{T.}}, \emph{et~al.},
10128:   \bibinfo{year}{2000}, \bibinfo{journal}{J. Phys. Soc. Jpn.}
10129:   \textbf{\bibinfo{volume}{69}}, \bibinfo{pages}{9}.
10130: 
10131: \bibitem[{\citenamefont{Tranquada} \emph{et~al.}(1988)\citenamefont{Tranquada,
10132:   Moudden, Goldman, Zolliker, Cox, Shirane, Sinha, Vaknin, Johnston, Alvarez,
10133:   Jacobson, Lewandowski} \emph{et~al.}}]{TMG8877}
10134: \bibinfo{author}{\bibnamefont{Tranquada}, \bibfnamefont{J.~M.}},
10135:   \bibinfo{author}{\bibfnamefont{A.~H.} \bibnamefont{Moudden}},
10136:   \bibinfo{author}{\bibfnamefont{A.~I.} \bibnamefont{Goldman}},
10137:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Zolliker}},
10138:   \bibinfo{author}{\bibfnamefont{D.~E.} \bibnamefont{Cox}},
10139:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Shirane}},
10140:   \bibinfo{author}{\bibfnamefont{S.~K.} \bibnamefont{Sinha}},
10141:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Vaknin}},
10142:   \bibinfo{author}{\bibfnamefont{D.~C.} \bibnamefont{Johnston}},
10143:   \bibinfo{author}{\bibfnamefont{M.~S.} \bibnamefont{Alvarez}},
10144:   \bibinfo{author}{\bibfnamefont{A.~J.} \bibnamefont{Jacobson}},
10145:   \bibinfo{author}{\bibfnamefont{J.~T.} \bibnamefont{Lewandowski}},
10146:   \emph{et~al.}, \bibinfo{year}{1988}, \bibinfo{journal}{Phys. Rev. B}
10147:   \textbf{\bibinfo{volume}{38}}, \bibinfo{pages}{2477}.
10148: 
10149: \bibitem[{\citenamefont{Tranquada} \emph{et~al.}(1995)\citenamefont{Tranquada,
10150:   Sternkieb, Aye, Nakamura, and Uchida}}]{TSA9561}
10151: \bibinfo{author}{\bibnamefont{Tranquada}, \bibfnamefont{J.~M.}},
10152:   \bibinfo{author}{\bibfnamefont{B.~J.} \bibnamefont{Sternkieb}},
10153:   \bibinfo{author}{\bibfnamefont{J.~D.} \bibnamefont{Aye}},
10154:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Nakamura}}, and
10155:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Uchida}},
10156:   \bibinfo{year}{1995}, \bibinfo{journal}{Nature}
10157:   \textbf{\bibinfo{volume}{375}}, \bibinfo{pages}{561}.
10158: 
10159: \bibitem[{\citenamefont{Tranquada} \emph{et~al.}(2004)\citenamefont{Tranquada,
10160:   Wou, Perring, Goka, Gu, Xu, Fujita, and Yamada}}]{TWP0434}
10161: \bibinfo{author}{\bibnamefont{Tranquada}, \bibfnamefont{J.~M.}},
10162:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Wou}},
10163:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Perring}},
10164:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Goka}},
10165:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Gu}},
10166:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Xu}},
10167:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Fujita}}, and
10168:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Yamada}},
10169:   \bibinfo{year}{2004}, \bibinfo{journal}{Nature}
10170:   \textbf{\bibinfo{volume}{429}}, \bibinfo{pages}{534}.
10171: 
10172: \bibitem[{\citenamefont{Trivedi and Ceperley}(1989)}]{TC8937}
10173: \bibinfo{author}{\bibnamefont{Trivedi}, \bibfnamefont{N.}}, and
10174:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Ceperley}},
10175:   \bibinfo{year}{1989}, \bibinfo{journal}{Phys. Rev. B}
10176:   \textbf{\bibinfo{volume}{40}}, \bibinfo{pages}{2737}.
10177: 
10178: \bibitem[{\citenamefont{Tsuchiura} \emph{et~al.}(2003)\citenamefont{Tsuchiura,
10179:   Ogata, Tanaka, and Kashiwaya}}]{TOT0309}
10180: \bibinfo{author}{\bibnamefont{Tsuchiura}, \bibfnamefont{H.}},
10181:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Ogata}},
10182:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Tanaka}}, and
10183:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Kashiwaya}},
10184:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. Rev. B}
10185:   \textbf{\bibinfo{volume}{68}}, \bibinfo{pages}{012509}.
10186: 
10187: \bibitem[{\citenamefont{Tsuchiura} \emph{et~al.}(2000)\citenamefont{Tsuchiura,
10188:   Tanaka, Ogata, and Kashiwaya}}]{TTO0005}
10189: \bibinfo{author}{\bibnamefont{Tsuchiura}, \bibfnamefont{H.}},
10190:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Tanaka}},
10191:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Ogata}}, and
10192:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Kashiwaya}},
10193:   \bibinfo{year}{2000}, \bibinfo{journal}{Phys. Rev. Lett.}
10194:   \textbf{\bibinfo{volume}{84}}, \bibinfo{pages}{3105}.
10195: 
10196: \bibitem[{\citenamefont{Ubbens and Lee}(1992)}]{UL9234}
10197: \bibinfo{author}{\bibnamefont{Ubbens}, \bibfnamefont{M.}}, and
10198:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
10199:   \bibinfo{year}{1992}, \bibinfo{journal}{Phys. Rev. B}
10200:   \textbf{\bibinfo{volume}{46}}, \bibinfo{pages}{8434}.
10201: 
10202: \bibitem[{\citenamefont{Ubbens and Lee}(1994)}]{UL9453}
10203: \bibinfo{author}{\bibnamefont{Ubbens}, \bibfnamefont{M.~U.}}, and
10204:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
10205:   \bibinfo{year}{1994}, \bibinfo{journal}{Phys. Rev. B}
10206:   \textbf{\bibinfo{volume}{49}}, \bibinfo{pages}{6853}.
10207: 
10208: \bibitem[{\citenamefont{Uchida}(1997)}]{U9712}
10209: \bibinfo{author}{\bibnamefont{Uchida}, \bibfnamefont{S.}},
10210:   \bibinfo{year}{1997}, \bibinfo{journal}{Physica C}
10211:   \textbf{\bibinfo{volume}{282}}, \bibinfo{pages}{12}.
10212: 
10213: \bibitem[{\citenamefont{Uchida} \emph{et~al.}(1991)\citenamefont{Uchida, Ido,
10214:   Takagi, Arima, Tokura, and Tajima}}]{UIT9142}
10215: \bibinfo{author}{\bibnamefont{Uchida}, \bibfnamefont{S.}},
10216:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Ido}},
10217:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Takagi}},
10218:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Arima}},
10219:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Tokura}}, and
10220:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Tajima}},
10221:   \bibinfo{year}{1991}, \bibinfo{journal}{Phys. Rev. B}
10222:   \textbf{\bibinfo{volume}{43}}, \bibinfo{pages}{7942}.
10223: 
10224: \bibitem[{\citenamefont{Uemura} \emph{et~al.}(1989)\citenamefont{Uemura, Luke,
10225:   Sternlieb, Brewer, Carolan, Hardy, Kadono, Kempton, Kiefl, Kreitzman,
10226:   Mulhern, Riseman} \emph{et~al.}}]{ULS8917}
10227: \bibinfo{author}{\bibnamefont{Uemura}, \bibfnamefont{Y.~J.}},
10228:   \bibinfo{author}{\bibfnamefont{G.~M.} \bibnamefont{Luke}},
10229:   \bibinfo{author}{\bibfnamefont{B.~J.} \bibnamefont{Sternlieb}},
10230:   \bibinfo{author}{\bibfnamefont{J.~H.} \bibnamefont{Brewer}},
10231:   \bibinfo{author}{\bibfnamefont{J.~F.} \bibnamefont{Carolan}},
10232:   \bibinfo{author}{\bibfnamefont{W.~N.} \bibnamefont{Hardy}},
10233:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Kadono}},
10234:   \bibinfo{author}{\bibfnamefont{J.~R.} \bibnamefont{Kempton}},
10235:   \bibinfo{author}{\bibfnamefont{R.~F.} \bibnamefont{Kiefl}},
10236:   \bibinfo{author}{\bibfnamefont{S.~R.} \bibnamefont{Kreitzman}},
10237:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Mulhern}},
10238:   \bibinfo{author}{\bibfnamefont{T.~M.} \bibnamefont{Riseman}}, \emph{et~al.},
10239:   \bibinfo{year}{1989}, \bibinfo{journal}{Phys. Rev. Lett.}
10240:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{2317}.
10241: 
10242: \bibitem[{\citenamefont{Ussishkin} \emph{et~al.}(2002)\citenamefont{Ussishkin,
10243:   Sondhi, and Huse}}]{USH0201}
10244: \bibinfo{author}{\bibnamefont{Ussishkin}, \bibfnamefont{I.}},
10245:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Sondhi}}, and
10246:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Huse}}, \bibinfo{year}{2002},
10247:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{89}},
10248:   \bibinfo{pages}{287001}.
10249: 
10250: \bibitem[{\citenamefont{Valla} \emph{et~al.}(1999)\citenamefont{Valla, Fedorov,
10251:   Johnson, Wells, Hulbert, Li, Gu, and Koshizuka}}]{VFJ9910}
10252: \bibinfo{author}{\bibnamefont{Valla}, \bibfnamefont{T.}},
10253:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Fedorov}},
10254:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Johnson}},
10255:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Wells}},
10256:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Hulbert}},
10257:   \bibinfo{author}{\bibfnamefont{Q.}~\bibnamefont{Li}},
10258:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Gu}}, and
10259:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Koshizuka}},
10260:   \bibinfo{year}{1999}, \bibinfo{journal}{Science}
10261:   \textbf{\bibinfo{volume}{285}}, \bibinfo{pages}{2110}.
10262: 
10263: \bibitem[{\citenamefont{Varma}(1997)}]{V9754}
10264: \bibinfo{author}{\bibnamefont{Varma}, \bibfnamefont{C.~M.}},
10265:   \bibinfo{year}{1997}, \bibinfo{journal}{Phys. Rev. B}
10266:   \textbf{\bibinfo{volume}{55}}, \bibinfo{pages}{14554}.
10267: 
10268: \bibitem[{\citenamefont{Varma} \emph{et~al.}(1987)\citenamefont{Varma,
10269:   Schmitt-Rink, and Abrahams}}]{VSA8781}
10270: \bibinfo{author}{\bibnamefont{Varma}, \bibfnamefont{C.~M.}},
10271:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Schmitt-Rink}}, and
10272:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Abrahams}},
10273:   \bibinfo{year}{1987}, \bibinfo{journal}{Solid State Commun.}
10274:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{681}.
10275: 
10276: \bibitem[{\citenamefont{Vershinin} \emph{et~al.}(2004)\citenamefont{Vershinin,
10277:   Shashank, Ono, Abe, Ando, and Yazdani}}]{VSO0495}
10278: \bibinfo{author}{\bibnamefont{Vershinin}, \bibfnamefont{M.}},
10279:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Shashank}},
10280:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Ono}},
10281:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Abe}},
10282:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Ando}}, and
10283:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Yazdani}},
10284:   \bibinfo{year}{2004}, \bibinfo{journal}{Science}
10285:   \textbf{\bibinfo{volume}{303}}, \bibinfo{pages}{1995}.
10286: 
10287: \bibitem[{\citenamefont{Vojta and Sachdev}(1999)}]{VS9916}
10288: \bibinfo{author}{\bibnamefont{Vojta}, \bibfnamefont{M.}}, and
10289:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Sachdev}},
10290:   \bibinfo{year}{1999}, \bibinfo{journal}{Phys. Rev. Lett.}
10291:   \textbf{\bibinfo{volume}{83}}, \bibinfo{pages}{3916}.
10292: 
10293: \bibitem[{\citenamefont{Vollhardt}(1984)}]{V8499}
10294: \bibinfo{author}{\bibnamefont{Vollhardt}, \bibfnamefont{D.}},
10295:   \bibinfo{year}{1984}, \bibinfo{journal}{Rev. Mod. Phys.}
10296:   \textbf{\bibinfo{volume}{56}}, \bibinfo{pages}{99}.
10297: 
10298: \bibitem[{\citenamefont{Volovik}(1993)}]{V9369}
10299: \bibinfo{author}{\bibnamefont{Volovik}, \bibfnamefont{G.}},
10300:   \bibinfo{year}{1993}, \bibinfo{journal}{JETP Lett.}
10301:   \textbf{\bibinfo{volume}{58}}, \bibinfo{pages}{469}.
10302: 
10303: \bibitem[{\citenamefont{Wakimoto} \emph{et~al.}(1999)\citenamefont{Wakimoto,
10304:   Shirane, Endoh, Hirota, Ueki, Yamada, Birgeneau, Kastner, Lee, Gehring, and
10305:   Lee}}]{WSE9969}
10306: \bibinfo{author}{\bibnamefont{Wakimoto}, \bibfnamefont{S.}},
10307:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Shirane}},
10308:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Endoh}},
10309:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Hirota}},
10310:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Ueki}},
10311:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Yamada}},
10312:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Birgeneau}},
10313:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Kastner}},
10314:   \bibinfo{author}{\bibfnamefont{Y.~S.} \bibnamefont{Lee}},
10315:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Gehring}}, and
10316:   \bibinfo{author}{\bibfnamefont{S.-H.} \bibnamefont{Lee}},
10317:   \bibinfo{year}{1999}, \bibinfo{journal}{Phys. Rev. B}
10318:   \textbf{\bibinfo{volume}{60}}, \bibinfo{pages}{R769}.
10319: 
10320: \bibitem[{\citenamefont{Wang} \emph{et~al.}(2004)\citenamefont{Wang, Lu,
10321:   Naughton, Gu, Uchida, and Ong}}]{WLN04}
10322: \bibinfo{author}{\bibnamefont{Wang}, \bibfnamefont{Y.}},
10323:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Lu}},
10324:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Naughton}},
10325:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Gu}},
10326:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Uchida}}, and
10327:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Ong}}, \bibinfo{year}{2004},
10328:   \bibinfo{journal}{to appear} .
10329: 
10330: \bibitem[{\citenamefont{Wang and MacDonald}(1995)}]{WM9576}
10331: \bibinfo{author}{\bibnamefont{Wang}, \bibfnamefont{Y.}}, and
10332:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{MacDonald}},
10333:   \bibinfo{year}{1995}, \bibinfo{journal}{Phys. Rev. B}
10334:   \textbf{\bibinfo{volume}{52}}, \bibinfo{pages}{3876}.
10335: 
10336: \bibitem[{\citenamefont{Wang} \emph{et~al.}(2002)\citenamefont{Wang, Ong, Xu,
10337:   Kakeshita, Uchida, Bonn, Liang, and Hardy}}]{WOX0203}
10338: \bibinfo{author}{\bibnamefont{Wang}, \bibfnamefont{Y.}},
10339:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Ong}},
10340:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Xu}},
10341:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Kakeshita}},
10342:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Uchida}},
10343:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Bonn}},
10344:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Liang}}, and
10345:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Hardy}},
10346:   \bibinfo{year}{2002}, \bibinfo{journal}{Phys. Rev. Lett.}
10347:   \textbf{\bibinfo{volume}{88}}, \bibinfo{pages}{257003}.
10348: 
10349: \bibitem[{\citenamefont{Wang} \emph{et~al.}(2003)\citenamefont{Wang, Ono,
10350:   Onose, Gu, Ando, Tokura, Uchida, and Ong}}]{WOO0386}
10351: \bibinfo{author}{\bibnamefont{Wang}, \bibfnamefont{Y.}},
10352:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Ono}},
10353:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Onose}},
10354:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Gu}},
10355:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Ando}},
10356:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Tokura}},
10357:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Uchida}}, and
10358:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Ong}}, \bibinfo{year}{2003},
10359:   \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{299}},
10360:   \bibinfo{pages}{86}.
10361: 
10362: \bibitem[{\citenamefont{Wang} \emph{et~al.}(2001)\citenamefont{Wang, Yu,
10363:   Kakeshita, Uchida, Ono, Ando, and Ong}}]{WYK0119}
10364: \bibinfo{author}{\bibnamefont{Wang}, \bibfnamefont{Y.}},
10365:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Yu}},
10366:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Kakeshita}},
10367:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Uchida}},
10368:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Ono}},
10369:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Ando}}, and
10370:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Ong}}, \bibinfo{year}{2001},
10371:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{64}},
10372:   \bibinfo{pages}{224519}.
10373: 
10374: \bibitem[{\citenamefont{Warren} \emph{et~al.}(1989)\citenamefont{Warren,
10375:   Walstedt, Brennert, Cava, Tyeko, Bell, and Dabbaph}}]{WWB8993}
10376: \bibinfo{author}{\bibnamefont{Warren}, \bibfnamefont{W.~W.}},
10377:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Walstedt}},
10378:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Brennert}},
10379:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Cava}},
10380:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Tyeko}},
10381:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Bell}}, and
10382:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Dabbaph}},
10383:   \bibinfo{year}{1989}, \bibinfo{journal}{Phys. Rev. Lett.}
10384:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{1193}.
10385: 
10386: \bibitem[{\citenamefont{Wegner}(1971)}]{W7159}
10387: \bibinfo{author}{\bibnamefont{Wegner}, \bibfnamefont{F.}},
10388:   \bibinfo{year}{1971}, \bibinfo{journal}{J. Math. Phys.}
10389:   \textbf{\bibinfo{volume}{12}}, \bibinfo{pages}{2259}.
10390: 
10391: \bibitem[{\citenamefont{Wells} \emph{et~al.}(1995)\citenamefont{Wells, Shen,
10392:   Matsura, King, Kastner, Greven, and Birgeneau}}]{WSM9564}
10393: \bibinfo{author}{\bibnamefont{Wells}, \bibfnamefont{B.~O.}},
10394:   \bibinfo{author}{\bibfnamefont{Z.-X.} \bibnamefont{Shen}},
10395:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Matsura}},
10396:   \bibinfo{author}{\bibfnamefont{D.~M.} \bibnamefont{King}},
10397:   \bibinfo{author}{\bibfnamefont{M.~A.} \bibnamefont{Kastner}},
10398:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Greven}}, and
10399:   \bibinfo{author}{\bibfnamefont{S.~R.~J.} \bibnamefont{Birgeneau}},
10400:   \bibinfo{year}{1995}, \bibinfo{journal}{Phys. Rev. Lett.}
10401:   \textbf{\bibinfo{volume}{74}}, \bibinfo{pages}{964}.
10402: 
10403: \bibitem[{\citenamefont{Wen}(1991)}]{Wsrvb}
10404: \bibinfo{author}{\bibnamefont{Wen}, \bibfnamefont{X.-G.}},
10405:   \bibinfo{year}{1991}, \bibinfo{journal}{Phys. Rev. B}
10406:   \textbf{\bibinfo{volume}{44}}, \bibinfo{pages}{2664}.
10407: 
10408: \bibitem[{\citenamefont{Wen}(1995)}]{Wtoprev}
10409: \bibinfo{author}{\bibnamefont{Wen}, \bibfnamefont{X.-G.}},
10410:   \bibinfo{year}{1995}, \bibinfo{journal}{Advances in Physics}
10411:   \textbf{\bibinfo{volume}{44}}, \bibinfo{pages}{405}.
10412: 
10413: \bibitem[{\citenamefont{Wen}(2002{\natexlab{a}})}]{Wlight}
10414: \bibinfo{author}{\bibnamefont{Wen}, \bibfnamefont{X.-G.}},
10415:   \bibinfo{year}{2002}{\natexlab{a}}, \bibinfo{journal}{Phys. Rev. Lett.}
10416:   \textbf{\bibinfo{volume}{88}}, \bibinfo{pages}{11602}.
10417: 
10418: \bibitem[{\citenamefont{Wen}(2002{\natexlab{b}})}]{Wqoslpub}
10419: \bibinfo{author}{\bibnamefont{Wen}, \bibfnamefont{X.-G.}},
10420:   \bibinfo{year}{2002}{\natexlab{b}}, \bibinfo{journal}{Phys. Rev. B}
10421:   \textbf{\bibinfo{volume}{65}}, \bibinfo{pages}{165113}.
10422: 
10423: \bibitem[{\citenamefont{Wen}(2003{\natexlab{a}})}]{Walight}
10424: \bibinfo{author}{\bibnamefont{Wen}, \bibfnamefont{X.-G.}},
10425:   \bibinfo{year}{2003}{\natexlab{a}}, \bibinfo{journal}{Phys. Rev. B}
10426:   \textbf{\bibinfo{volume}{68}}, \bibinfo{pages}{115413}.
10427: 
10428: \bibitem[{\citenamefont{Wen}(2003{\natexlab{b}})}]{Wqoem}
10429: \bibinfo{author}{\bibnamefont{Wen}, \bibfnamefont{X.-G.}},
10430:   \bibinfo{year}{2003}{\natexlab{b}}, \bibinfo{journal}{Phys. Rev. D}
10431:   \textbf{\bibinfo{volume}{68}}, \bibinfo{pages}{065003}.
10432: 
10433: \bibitem[{\citenamefont{Wen}(2003{\natexlab{c}})}]{Wqoexct}
10434: \bibinfo{author}{\bibnamefont{Wen}, \bibfnamefont{X.-G.}},
10435:   \bibinfo{year}{2003}{\natexlab{c}}, \bibinfo{journal}{Phys. Rev. Lett.}
10436:   \textbf{\bibinfo{volume}{90}}, \bibinfo{pages}{016803}.
10437: 
10438: \bibitem[{\citenamefont{Wen}(2004)}]{Wen04}
10439: \bibinfo{author}{\bibnamefont{Wen}, \bibfnamefont{X.-G.}},
10440:   \bibinfo{year}{2004}, \emph{\bibinfo{title}{Quantum Field Theory of Many-Body
10441:   Systems -- From the Origin of Sound to an Origin of Light and Electrons}}
10442:   (\bibinfo{publisher}{Oxford Univ. Press}, \bibinfo{address}{Oxford}).
10443: 
10444: \bibitem[{\citenamefont{Wen and Lee}(1996)}]{WLsu2}
10445: \bibinfo{author}{\bibnamefont{Wen}, \bibfnamefont{X.-G.}}, and
10446:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
10447:   \bibinfo{year}{1996}, \bibinfo{journal}{Phys. Rev. Lett.}
10448:   \textbf{\bibinfo{volume}{76}}, \bibinfo{pages}{503}.
10449: 
10450: \bibitem[{\citenamefont{Wen and Lee}(1998)}]{WL9893}
10451: \bibinfo{author}{\bibnamefont{Wen}, \bibfnamefont{X.-G.}}, and
10452:   \bibinfo{author}{\bibfnamefont{P.~A.} \bibnamefont{Lee}},
10453:   \bibinfo{year}{1998}, \bibinfo{journal}{Phys. Rev. Lett.}
10454:   \textbf{\bibinfo{volume}{80}}, \bibinfo{pages}{2193}.
10455: 
10456: \bibitem[{\citenamefont{Wen} \emph{et~al.}(1989)\citenamefont{Wen, Wilczek, and
10457:   Zee}}]{WWZcsp}
10458: \bibinfo{author}{\bibnamefont{Wen}, \bibfnamefont{X.-G.}},
10459:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Wilczek}}, and
10460:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Zee}}, \bibinfo{year}{1989},
10461:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{39}},
10462:   \bibinfo{pages}{11413}.
10463: 
10464: \bibitem[{\citenamefont{Wen and Zee}(2002)}]{WZqoind}
10465: \bibinfo{author}{\bibnamefont{Wen}, \bibfnamefont{X.-G.}}, and
10466:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Zee}}, \bibinfo{year}{2002},
10467:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{66}},
10468:   \bibinfo{pages}{235110}.
10469: 
10470: \bibitem[{\citenamefont{Weng}(2003)}]{W0361}
10471: \bibinfo{author}{\bibnamefont{Weng}, \bibfnamefont{Z.-Y.}},
10472:   \bibinfo{year}{2003}, in \emph{\bibinfo{booktitle}{Proceedings of the
10473:   International Symposium on Frontiers of Science}},
10474:   \bibinfo{note}{cond-mat/0304261}.
10475: 
10476: \bibitem[{\citenamefont{Weng} \emph{et~al.}(2000)\citenamefont{Weng, Sheng, and
10477:   Ting}}]{WST0067}
10478: \bibinfo{author}{\bibnamefont{Weng}, \bibfnamefont{Z.~Y.}},
10479:   \bibinfo{author}{\bibfnamefont{D.~N.} \bibnamefont{Sheng}}, and
10480:   \bibinfo{author}{\bibfnamefont{C.~S.} \bibnamefont{Ting}},
10481:   \bibinfo{year}{2000}, \bibinfo{journal}{Physica}
10482:   \textbf{\bibinfo{volume}{C341-348}}, \bibinfo{pages}{67}.
10483: 
10484: \bibitem[{\citenamefont{White and Scalapino}(1999)}]{WS9953}
10485: \bibinfo{author}{\bibnamefont{White}, \bibfnamefont{S.~R.}}, and
10486:   \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Scalapino}},
10487:   \bibinfo{year}{1999}, \bibinfo{journal}{Phys. Rev. B}
10488:   \textbf{\bibinfo{volume}{60}}, \bibinfo{pages}{753}.
10489: 
10490: \bibitem[{\citenamefont{Wilson}(1974)}]{W7445}
10491: \bibinfo{author}{\bibnamefont{Wilson}, \bibfnamefont{K.~G.}},
10492:   \bibinfo{year}{1974}, \bibinfo{journal}{Phys. Rev. D}
10493:   \textbf{\bibinfo{volume}{10}}, \bibinfo{pages}{2445}.
10494: 
10495: \bibitem[{\citenamefont{Witten}(1979)}]{W7985}
10496: \bibinfo{author}{\bibnamefont{Witten}, \bibfnamefont{E.}},
10497:   \bibinfo{year}{1979}, \bibinfo{journal}{Nucl. Phys. B}
10498:   \textbf{\bibinfo{volume}{149}}, \bibinfo{pages}{285}.
10499: 
10500: \bibitem[{\citenamefont{Wu} \emph{et~al.}(1998)\citenamefont{Wu, Mou, Wen, and
10501:   Chang}}]{WMW9846}
10502: \bibinfo{author}{\bibnamefont{Wu}, \bibfnamefont{C.-L.}},
10503:   \bibinfo{author}{\bibfnamefont{C.-Y.} \bibnamefont{Mou}},
10504:   \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}}, and
10505:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Chang}},
10506:   \bibinfo{year}{1998}, \bibinfo{journal}{cond-mat/9811146}.
10507: 
10508: \bibitem[{\citenamefont{Wu} \emph{et~al.}(1987)\citenamefont{Wu, Ashburn,
10509:   Torng, Hor, Meng, Gao, Huang, Wang, and Chu}}]{WAT8708}
10510: \bibinfo{author}{\bibnamefont{Wu}, \bibfnamefont{M.~K.}},
10511:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Ashburn}},
10512:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Torng}},
10513:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Hor}},
10514:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Meng}},
10515:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Gao}},
10516:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Huang}},
10517:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Wang}}, and
10518:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Chu}}, \bibinfo{year}{1987},
10519:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{58}},
10520:   \bibinfo{pages}{908}.
10521: 
10522: \bibitem[{\citenamefont{Wynn} \emph{et~al.}(2001)\citenamefont{Wynn, Bonn,
10523:   Gardner, Lin, Liang, Hardy, Kirtley, and Moler}}]{WBG0102}
10524: \bibinfo{author}{\bibnamefont{Wynn}, \bibfnamefont{J.}},
10525:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Bonn}},
10526:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Gardner}},
10527:   \bibinfo{author}{\bibfnamefont{Y.-J.} \bibnamefont{Lin}},
10528:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Liang}},
10529:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Hardy}},
10530:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kirtley}}, and
10531:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Moler}},
10532:   \bibinfo{year}{2001}, \bibinfo{journal}{Phys. Rev. Lett.}
10533:   \textbf{\bibinfo{volume}{87}}, \bibinfo{pages}{197002}.
10534: 
10535: \bibitem[{Yamada \emph{et~al.}(1998)\citenamefont{Yamada}
10536:   \emph{et~al.}}]{Yo9865}
10537: \bibinfo{author}{\bibnamefont{Yamada}, \bibfnamefont{K.}}, \emph{et~al.},
10538:   \bibinfo{year}{1998}, \bibinfo{journal}{Phys. Rev. B}
10539:   \textbf{\bibinfo{volume}{57}}, \bibinfo{pages}{6165}.
10540: 
10541: \bibitem[{\citenamefont{Ye}(2002)}]{Y0217}
10542: \bibinfo{author}{\bibnamefont{Ye}, \bibfnamefont{J.}}, \bibinfo{year}{2002},
10543:   \bibinfo{journal}{cond-mat/0205417}.
10544: 
10545: \bibitem[{\citenamefont{Ye} \emph{et~al.}(1999)\citenamefont{Ye, Kim, Millis,
10546:   Shraiman, Majumda, and Tesanovic}}]{YKM9937}
10547: \bibinfo{author}{\bibnamefont{Ye}, \bibfnamefont{J.}},
10548:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Kim}},
10549:   \bibinfo{author}{\bibfnamefont{A.~J.} \bibnamefont{Millis}},
10550:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Shraiman}},
10551:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Majumda}}, and
10552:   \bibinfo{author}{\bibfnamefont{Z.}~\bibnamefont{Tesanovic}},
10553:   \bibinfo{year}{1999}, \bibinfo{journal}{Phys. Rev. Lett.}
10554:   \textbf{\bibinfo{volume}{83}}, \bibinfo{pages}{3737}.
10555: 
10556: \bibitem[{\citenamefont{Yokoyama and Ogata}(1996)}]{YO9615}
10557: \bibinfo{author}{\bibnamefont{Yokoyama}, \bibfnamefont{H.}}, and
10558:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Ogata}},
10559:   \bibinfo{year}{1996}, \bibinfo{journal}{J. Phys. Soc. Jpn}
10560:   \textbf{\bibinfo{volume}{65}}, \bibinfo{pages}{3615}.
10561: 
10562: \bibitem[{Yoshida \emph{et~al.}(2003)\citenamefont{Yoshida}
10563:   \emph{et~al.}}]{Yo0301}
10564: \bibinfo{author}{\bibnamefont{Yoshida}, \bibfnamefont{T.}}, \emph{et~al.},
10565:   \bibinfo{year}{2003}, \bibinfo{journal}{Phys. rev. Lett.}
10566:   \textbf{\bibinfo{volume}{91}}, \bibinfo{pages}{027001}.
10567: 
10568: \bibitem[{\citenamefont{Yoshioka}(1989)}]{Y8916}
10569: \bibinfo{author}{\bibnamefont{Yoshioka}, \bibfnamefont{D.}},
10570:   \bibinfo{year}{1989}, \bibinfo{journal}{J. Phys. Soc. Jpn.}
10571:   \textbf{\bibinfo{volume}{58}}, \bibinfo{pages}{1516}.
10572: 
10573: \bibitem[{\citenamefont{Yu} \emph{et~al.}(1987)\citenamefont{Yu, Freeman, and
10574:   Xu}}]{YFX8735}
10575: \bibinfo{author}{\bibnamefont{Yu}, \bibfnamefont{J.}},
10576:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Freeman}}, and
10577:   \bibinfo{author}{\bibfnamefont{J.-H.} \bibnamefont{Xu}},
10578:   \bibinfo{year}{1987}, \bibinfo{journal}{Phys. Rev. Lett.}
10579:   \textbf{\bibinfo{volume}{58}}, \bibinfo{pages}{1035}.
10580: 
10581: \bibitem[{\citenamefont{Zaanen} \emph{et~al.}(1985)\citenamefont{Zaanen,
10582:   Sauatzky, and Allen}}]{ZSA8518}
10583: \bibinfo{author}{\bibnamefont{Zaanen}, \bibfnamefont{J.}},
10584:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Sauatzky}}, and
10585:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Allen}},
10586:   \bibinfo{year}{1985}, \bibinfo{journal}{Phys. Rev. Lett.}
10587:   \textbf{\bibinfo{volume}{55}}, \bibinfo{pages}{418}.
10588: 
10589: \bibitem[{\citenamefont{Zhang} \emph{et~al.}(1988)\citenamefont{Zhang, Gros,
10590:   Rice, and Shiba}}]{ZGR8836}
10591: \bibinfo{author}{\bibnamefont{Zhang}, \bibfnamefont{F.~C.}},
10592:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Gros}},
10593:   \bibinfo{author}{\bibfnamefont{T.~M.} \bibnamefont{Rice}}, and
10594:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Shiba}},
10595:   \bibinfo{year}{1988}, \bibinfo{journal}{Supercond. Sci. Tech.}
10596:   \textbf{\bibinfo{volume}{1}}, \bibinfo{pages}{36}.
10597: 
10598: \bibitem[{\citenamefont{Zhang and Rice}(1988)}]{ZR8859}
10599: \bibinfo{author}{\bibnamefont{Zhang}, \bibfnamefont{F.~C.}}, and
10600:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Rice}}, \bibinfo{year}{1988},
10601:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{37}},
10602:   \bibinfo{pages}{3759}.
10603: 
10604: \bibitem[{\citenamefont{Zhang}(1997)}]{Z9789}
10605: \bibinfo{author}{\bibnamefont{Zhang}, \bibfnamefont{S.-C.}},
10606:   \bibinfo{year}{1997}, \bibinfo{journal}{Science}
10607:   \textbf{\bibinfo{volume}{275}}, \bibinfo{pages}{1089}.
10608: 
10609: \bibitem[{\citenamefont{Zheng} \emph{et~al.}(2003)\citenamefont{Zheng,
10610:   Odaguchi, Mito, Kitaoka, Asayama, and Kodama}}]{ZOM0391}
10611: \bibinfo{author}{\bibnamefont{Zheng}, \bibfnamefont{G.-Q.}},
10612:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Odaguchi}},
10613:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Mito}},
10614:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Kitaoka}},
10615:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Asayama}}, and
10616:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Kodama}},
10617:   \bibinfo{year}{2003}, \bibinfo{journal}{J. Phys. Soc. Jpn.}
10618:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{2591}.
10619: 
10620: \bibitem[{\citenamefont{Zhou} \emph{et~al.}(2004)\citenamefont{Zhou, Yoshida,
10621:   Lee, Yang, Brouet, Zhou, Ti, Xiong, Zhao, Sasagawa, Kakeshita, Eisaki}
10622:   \emph{et~al.}}]{ZYL0481}
10623: \bibinfo{author}{\bibnamefont{Zhou}, \bibfnamefont{Y.-J.}},
10624:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Yoshida}},
10625:   \bibinfo{author}{\bibfnamefont{D.-H.} \bibnamefont{Lee}},
10626:   \bibinfo{author}{\bibfnamefont{W.~L.} \bibnamefont{Yang}},
10627:   \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Brouet}},
10628:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Zhou}},
10629:   \bibinfo{author}{\bibfnamefont{W.~X.} \bibnamefont{Ti}},
10630:   \bibinfo{author}{\bibfnamefont{J.~W.} \bibnamefont{Xiong}},
10631:   \bibinfo{author}{\bibfnamefont{Z.~X.} \bibnamefont{Zhao}},
10632:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Sasagawa}},
10633:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Kakeshita}},
10634:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Eisaki}}, \emph{et~al.},
10635:   \bibinfo{year}{2004}, \bibinfo{journal}{cond-mat/0403181}.
10636: 
10637: \end{thebibliography}
10638: 
10639: \end{document}
10640: