cond-mat0302347/rpp.tex
1: \documentclass[12pt]{iopart}
2: \usepackage{epsf}
3: \begin{document}
4: \title{The Electronic Nature of High Temperature Cuprate Superconductors}
5: \author{M. R. Norman$^{1,2}$ and C. P\'{e}pin$^2$}
6: \address{$^1$Materials Science Division, Argonne National Laboratory, 
7: Argonne, IL 60439, USA}
8: \address{$^2$SPhT, L'Orme des Merisiers, CEA-Saclay, 91191 Gif-sur-Yvette,
9: France}
10: \ead{norman@anl.gov}
11: \begin{abstract}
12: We review the field of high temperature cuprate superconductors, with an
13: emphasis on the nature of their electronic properties.  After a general
14: overview of experiment and theory, we concentrate on recent results
15: obtained by angle resolved photoemission, inelastic neutron scattering,
16: and optical conductivity, along with various proposed explanations for
17: these results.  We conclude by reviewing efforts which attempt to
18: identify the energy savings
19: involved in the formation of the superconducting ground state.
20: \end{abstract}
21: \pacs{74.25.-q, 74.25.Jb, 74.72.-h}
22: \submitto{\RPP}
23: 
24: \maketitle
25: 
26: \tableofcontents
27: 
28: \section{General Overview}
29: 
30: \subsection{History}
31: 
32: As with all endeavors in science, there is a prehistory involved, and 
33: cuprates are no exception.  Research in superconductivity was not the 
34: languishing field that it is often portrayed as being prior to the 
35: Bednorz-Muller ``revolution'' of 1986.  What had begun to be 
36: diminished, though, was the hope that a truly high temperature 
37: superconductor would ever be discovered.  At the time of the 
38: Bednorz-Muller discovery, the highest 
39: temperature superconductor known was Nb$_3$Ge (23K).  That
40: material had been known since 1973, and was not much of an 
41: improvement over NbN (15K) which had been discovered all the way back in 
42: 1941 \cite{DAHL}.  This pessimistic outlook was best articulated by 
43: Bernd Matthias in a number of papers which 
44: still make interesting reading today \cite{MATTHIAS}.  Such pessimism was 
45: not confined to experiment, as witnessed by the famous paper of Cohen 
46: and Anderson \cite{COHEN}.  As was well appreciated by that time, the 
47: A15 materials with highest $T_c$ were on the verge of a structural 
48: transition, and thus it was anticipated that one could not push $T_c$ 
49: much higher before the lattice became unstable \cite{TESTARDI}.
50: 
51: Despite this, a number of new classes of superconductors had 
52: been discovered in the period before 1986, including the ternary magnetic 
53: superconductors such as ErRh$_4$B$_4$ and HoMo$_6$S$_8$, and 
54: various uranium based superconductors such as $\alpha$-U and U$_6$Fe,
55: many of these discovered by Matthias and his various associates.  
56: Matthias' speculation that something really different was going on in 
57: f electron superconductors was spectacularly confirmed with the 
58: discovery by Frank Steglich in 1979 of ``heavy fermion'' superconductivity in 
59: CeCu$_2$Si$_2$ \cite{FRANK}, followed by the discovery of superconductivity
60: in UPt$_3$ and UBe$_{13}$ \cite{GREG}.
61: 
62: Heavy fermion superconductivity was one of the main 
63: research topics in fundamental physics prior to 1986, and its history 
64: has had some impact on the cuprate field.  Unlike the magnetic 
65: superconductors such as ErRh$_4$B$_4$ where the magnetic moments are 
66: confined to the rare earth site and the superconductivity to the 
67: ligand sites, in heavy fermion superconductors, the f electrons 
68: themselves become superconducting.  This is known from the extremely 
69: high effective mass of the superconducting carriers.  More properly, 
70: the carriers should be thought of as composite objects of conduction 
71: electron charge and f electron spin \cite{MILLIS}.  The fascinating 
72: thing about these materials, though, is that their superconducting 
73: ground states do not appear to have the L=0, S=0 symmetry that Cooper 
74: pairs exhibit in normal superconductors \cite{RMP91,RMP02}.
75: 
76: As with cuprates, heavy fermion superconductivity had its 
77: own prehistory as well, that being the field of superfluid $^3$He.  
78: $^3$He had been speculated in the 1960s to possibly be a paired 
79: superfluid with non-zero orbital angular momentum, in particular 
80: L=2 pairing \cite{EMERY}.  The idea was that the hard core repulsion of 
81: the atoms would prevent L=0 pairing, but that longer range pairs could 
82: be stabilized by the attractive van der Waals interaction between the 
83: He atoms.  Subsequently, Layzer and Fay \cite{FAY} showed that for 
84: nearly ferromagnetic metals, and $^3$He as well, spin dependent
85: interactions instead could stabilize L=1, S=1 pairs, leading to the 
86: concept of paramagnon mediated pairing.
87: 
88: At that time, experimentalists were beginning to push $^3$He to low 
89: temperatures, with the idea of searching for magnetic order under 
90: pressure.  This was subsequently discovered by Bill Halperin.  But 
91: along the way, Doug Osheroff found superfluidity, and various 
92: experiments did indeed confirm the L=1, S=1 nature of the pairs
93: \cite{3He}.  More interestingly, two paired states were found, the 
94: so-called A and B phases.  Anderson and Brinkman later proposed 
95: that the stabilization of the anisotropic A phase relative to the 
96: isotropic B phase 
97: could be understood by feedback of the pair formation on the spin 
98: fluctuation interactions which supposedly gave rise to the pairs to 
99: begin with \cite{BA}.  Such feedback effects are much in vogue lately 
100: in regards to spin fluctuation mediated theories of cuprates
101: \cite{AC}.
102: 
103: What does this imply for the lattice case?  Early theories for heavy 
104: fermions were indeed based on the $^3$He paradigm, but with the 
105: discovery of antiferromagnetic correlations by inelastic neutron 
106: scattering \cite{GABE}, people turned away from these nearly ferromagnetic 
107: models 
108: (though they have seen a resurgence of late, with the discovery of 
109: superconductivity in UGe$_{2}$ \cite{UGe2} and ZrZn$_2$ 
110: \cite{ZrZn2}).  Rather, theoretical work published in 1986 led 
111: to the concept of L=2, S=0 pairs in the nearly antiferromagnetic 
112: case \cite{1986}.  This 
113: d-wave model is still one of the leading candidates to describe 
114: superconductivity in UPt$_3$, though a competing model based on 
115: f-wave pairs has been proposed by Norman\cite{MIKE92} and 
116: Sauls \cite{JIM94}. 
117: The problems with determining the pair symmetry in heavy fermions are 
118: the multiple band nature of the problem (orbital degeneracy), along 
119: with the effects of non-trivial crystal structures (UPt$_3$ has a non 
120: symmorphic lattice, for instance) and spin-orbit (which destroys L and 
121: S as good quantum numbers), not to mention the complications 
122: of dealing with three dimensions.  (Fortunately none of these problems 
123: exist in the cuprates.)  One of the interesting observations 
124: of these early calculations was the prediction that for a simple cubic 
125: lattice, the d-wave pairs should be of the form $(x^2-y^2) \pm i (3z^2-r^2)$ 
126: \cite{1986}.  If 
127: one simply eliminates the third dimension, one obtains the order 
128: paramater now known to be the pair state of the cuprates.  In some 
129: sense, the prediction of $d_{x^2-y^2}$ pairing in the cuprates was the 
130: ultimate one liner.
131: 
132: The above path, though, is not what led to the discovery of cuprate 
133: superconductors.  The history of this is rather lucidly described in 
134: Bednorz and Muller's Nobel lecture \cite{RMP88}.  Of particular 
135: interest to them was the case of doped SrTiO$_3$.  This material had a 
136: $T_c$ of less than 1K, but as it had such an incredibly low carrier density, 
137: it shouldn't have been superconducting at all, at least according to 
138: what people thought at the time.  In fact, the properties of this 
139: material led to a speculation by Eagles about the possibility of Bose 
140: condensation \cite{EAGLES}, with pairs existing 
141: above the superconducting transition temperature, a forerunner 
142: of the pseudogap 
143: physics currently being discussed for the cuprates.  Although Binning 
144: and Bednorz did some work on this material, it never led to much,
145: so Binning got bored and moved on to the discovery of the scanning 
146: tunneling microscope, which he later got the Nobel prize for.
147: 
148: After this, Bednorz began to work under Alex Muller, who was also 
149: interested in the possibility of oxide superconductors.  Alex was 
150: particularly intrigued by the role that Jahn-Teller effects
151: played in the perovskite structure; that is, in the distortions of the 
152: oxygen octahedra surrounding the transition metal ions
153: which lead to the lifting of the degeneracy of the 3d crystal field levels.
154: While searching around, they 
155: became aware of work that Raveau's group had done on LaBaCuO, and in 
156: the course of reproducing this work, they discovered high temperature 
157: superconductivity.  The story subsequently circulated was that Raveau 
158: took his samples off the shelf, and found that they too were 
159: superconducting.  It is on such twists of fate that careers in 
160: science are often decided.
161: 
162: After the original discovery, several groups got in the act, and by 
163: use of pressure, Paul Chu's group was able to drive the initial 
164: transition temperature of 35K up to 50K.  The real quest, though, was 
165: to find a related structure with a higher transition temperature, and 
166: this was rapidly discovered in early 1987, when Chu and collaborators 
167: found 90K superconductivity in YBCO \cite{CHU}.  The liquid 
168: air barrier (77K) had finally been breached, and true high temperature 
169: superconductivity had at last been discovered.  By varying the crystal
170: structure and again exploiting pressure, transition temperatures 
171: up to 160K have been achieved, again by Chu's group.  Matthias must have 
172: been smiling from on high.
173: 
174: \subsection{Crystal Symmetry and Electronic Structure}
175: 
176: The crystal structures of the cuprates were one of the first things 
177: elucidated, which is obvious because of the patent rights involved 
178: (after a very long struggle, Bell Labs eventually won that 
179: one \cite{BELL}; for an 
180: illuminating account of those heady days, the reader is referred to 
181: the book by Hazen \cite{HAZEN}).  Though they come in many variants, 
182: the basic structure is quite simple (Fig.~\ref{fig1}).  The material consists of 
183: CuO$_2$ planes, where each Cu ion is four fold coordinated with O ions,
184: separated by insulating spacer layers.
185: \begin{figure}
186: \centerline{\epsfxsize=0.6\textwidth{\epsfbox{fig1.eps}}}
187: \caption{Crystal structure of $La_2CuO_4$.  Left panel shows the 
188: layer structure along the c-axis, the right panel the structure of 
189: the $CuO_{2}$ plane.}
190: \label{fig1}
191: \end{figure}
192: The exception 
193: to this is YBCO, which has metallic CuO chain layers as one of the 
194: spacer layers.  The orginal 
195: LBCO material had two apical oxygens (the standard
196: perovskite structure, where the transition metal ion is in the center 
197: of an octahedron formed from six surrounding oxygen ions), but in 
198: YBCO, one of these oxygens is absent, and in other structures, 
199: such as that formed by electron doped cuprates, the apical oxygens are 
200: totally missing.  What this means is that despite all of these 
201: complications, the essential structure to worry about are the CuO$_2$ 
202: planes, which was understood early on, particularly by 
203: Anderson \cite{RVB}.  Of course, 
204: the superconducting transition temperature varies a lot from structure to 
205: structure, and is generally higher the more CuO$_2$ planes per unit 
206: cell that there are, but after years of study, most researchers have come to the 
207: conclusion that the main c-axis effect is simply to tune the 
208: electronic structure of the CuO$_2$ planes.
209: 
210: When considering these planes, one immediately comes across a basic 
211: fact.  Most transition metal oxides are insulators with
212: a particular electronic 
213: structure.  This is due to the fact that the transition metal 3d level 
214: and oxygen 
215: 2p level are separated by a greater energy than the energy spread of these 
216: levels from band formation.  The net result is one gets separate 3d 
217: and 2p energy bands.  The Coulomb repulsion on the transition metal 
218: site is so large that the 3d band 
219: ``Mott-Hubbardizes'', spliting into upper and lower Hubbard bands 
220: separated by this energy scale, U (typically 8-10 eV in the solid).  The 
221: true energy gap then becomes of the charge transfer type,
222: separating the filled oxygen 2p valence band from the empty 3d 
223: conduction (i.e., upper Hubbard) band \cite{SAWATZKY}.
224: 
225: The cuprate case is different, though (Fig.~\ref{fig2}) \cite{PICKETT}.  In the solid,
226: the Cu ion is in a $d^9$ configuration ($Cu^{++}$) and the O ion in a $p^6$
227: configuration ($O^{--}$), with the  Cu 3d energy level above but relatively
228: close to the O 2p energy level.  In the layered perovskites, the tetragonal
229: environment of the Cu ion leads to the single 3d hole having $d_{x^2-y^2}$
230: symmetry.  In this case, the dominant energy  is the bonding-antibonding splitting 
231: involving the
232: quantum mechanical mixture of the
233: Cu 3d $x^2-y^2$ orbital and the planar O $2p_x$ and $2p_y$ orbitals (with an
234: energy of 6 eV).
235: \begin{figure}
236: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig2.eps}}}
237: \caption{Electronic structure of the undoped cuprates.  Left panel shows the
238: atomic Cu d and O p levels, middle panel the band structure of the 
239: solid (where B is the bonding combination of the atomic levels, AB the 
240: antibonding one), and right panel the effect of correlations
241: (Mott-Hubbard gap) on the AB band (LHB and UHB are the lower and upper
242: Hubbard bands).}
243: \label{fig2}
244: \end{figure}
245: The net 
246: result is that in the parent (undoped) structure, one is left with a 
247: half filled band which is the antibonding combination of these three orbitals, 
248: with the bonding, non-bonding, and the rest of the Cu and O
249: orbitals filled.  As Anderson 
250: speculated early on \cite{RVB},
251: it is this copper-oxygen antibonding band which ``Mott-Hubbardizes'', forming 
252: an insulating gap of order 2 eV in the parent compound (the effective 
253: U being reduced because of the Cu-O orbital admixture).
254: 
255: Therefore, in the end, the complicated electronic structure leads to 
256: a single 2D energy band near the Fermi energy, which is what makes the 
257: cuprates so 
258: attractive from a theoretical perspective \cite{ANDERSON}.
259: But one can even reduce 
260: this ``one band Hubbard model'' further by taking the limit of large 
261: U.  In this case, the upper Hubbard band is projected out (assuming we 
262: are considering hole doping the insulator), and the effect of U 
263: becomes virtual, leading to a superexchange interaction between the 
264: Cu spins, J 
265: (t$^2$/U, where t is the effective Cu-Cu hopping mediated by intervening
266: O sites).  This is easily understood by 
267: noting that two parallel spins are not allowed to occupy the same Cu site 
268: because of the Pauli exclusion principle, but antiparallel spins can, 
269: leading to an energy savings of t$^2$/U from second order perturbation 
270: theory.  This so-called ``t-J'' model is the minimal model for the 
271: cuprates.  Despite the success of motivating this model from first 
272: principles calculations \cite{HYBERTSEN}, it is not generally agreed 
273: upon.  For instance, Varma has advocated that one must consider the full 
274: three band Hubbard model (one band from each of the three states, Cu 
275: 3d $x^2-y^2$, O 2p$_x$, and O 2p$_y$).  His claim is that upon projection to 
276: the low energy sector, non-trivial phase factors between the 
277: three bands become
278: possible, which can lead to an orbital current state which 
279: he associates with the pseudogap phase \cite{VARMA}.
280: 
281: \subsection{Phase Diagram}
282: 
283: The original hope of Anderson was that the insulating phase of the 
284: cuprates would turn out to be a spin liquid \cite{RVB}.
285: \begin{figure}
286: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig3.eps}}}
287: \caption{Phase diagram of the cuprates (x is the hole doping).  AF is 
288: the antiferromagnetic insulator.  The dotted line is a crossover line 
289: between the normal metal phase and the pseudogap phase.}
290: \label{fig3}
291: \end{figure}
292: The issue 
293: here is that most Mott insulators exhibit broken symmetry, such as 
294: antiferromagnetism.  This means that such insulators can be 
295: adiabatically continued to an ordinary band insulator with magnetic 
296: order, as originally proposed by Slater \cite{SLATER}.  The cuprates 
297: would be another example of this, since the half filled band discussed
298: above would become one full and one empty band upon magnetic 
299: ordering due to the unit cell doubling
300: (in this picture, the lower Hubbard band would correspond to 
301: the up spin band of the band insulator, the upper Hubbard band to the 
302: down spin band).  Anderson believed, though, that the Mott phenomenon
303: should be unrelated to this argument, and that the cuprates 
304: would be the ideal place to demonstrate this.  Since there is only 
305: one d hole per Cu site (and thus, S=1/2), and given the 2D nature of 
306: the material, he felt that quantum fluctuations would be sufficient 
307: to destroy the order, leading to a spin liquid ground state, which he 
308: called a resonating valence bond (RVB) state (harking back to the 
309: theory of benzene rings, where each C-C link resonates between a single bond 
310: and a double bond state).
311: 
312: As was discovered soon after, though, the undoped phase is indeed 
313: magnetic, though the moment is reduced by 1/3 from the free ion value 
314: due to fluctuations \cite{INS}.  On the other hand, magnetic order is rapidly 
315: destroyed upon hole doping, so in fact the magnetic phase only takes 
316: up a small sliver (Fig.~\ref{fig3}) of the phase diagram (in the electron doped case, 
317: though, the magnetism exists over a much larger doping range).  So, 
318: in that sense, Anderson's intuition was quite good.
319: 
320: For dopings beyond a few percent, the system either enters a messy 
321: disordered phase exhibiting spin glass behavior (as in LSCO) before
322: superconducting order sets in, or 
323: immediately goes to the superconducting phase (as in YBCO).  The 
324: superconducting transition monotonically rises with doping, reaching a 
325: maximum at about 16\% doping, after which $T_c$ declines to zero.  The 
326: net effect is to form a superconducting ``dome'' which extends from 
327: about 5\% to 25\% doping.
328: 
329: At first sight, the superconducting phase is not so different from 
330: that of classical superconductors.  We know that it exhibits a zero 
331: resistance state with a Meissner effect.  Experiments show 
332: that the superconducting objects are charge 2e, and thus pairs are 
333: formed.  What is unusual, though, are the short coherence lengths.  
334: For typical superconductors, the coherence length is quite long, usually 
335: several hundered $\AA$ or more.  This is in contrast to magnets, which have 
336: quite short coherence lengths.  Therefore, for most superconductors we 
337: know, mean field theory works extremely well, as opposed to magnets 
338: where it almost always fails.  But cuprates exhibit short coherence 
339: lengths, of order 20$\AA$ in the plane, and a paltry 2$\AA$ between 
340: planes.  The latter is so short, the cuprates are essentially 
341: composed of Josephson coupled planes, as has been experimentally 
342: verified by a number of groups \cite{KLMU}.  Such coupling is necessary, 
343: of course, since long range superconducting order cannot occur in two 
344: dimensions (except the Kosterlitz-Thouless phase, whose existence in the 
345: cuprates is still debated \cite{JOE}).
346: 
347: Another unusual finding is the symmetry of the order parameter (Fig.~\ref{fig4}).  
348: For many 
349: years, it was felt that the order parameter probably had s-wave 
350: symmetry.
351: \begin{figure}
352: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig4.eps}}}
353: \caption{Angular variation of the d-wave gap around the Fermi surface 
354: is shown in the left panel (with nodes at $k_x=\pm k_y$).  
355: The resulting density of states is plotted 
356: in the right panel.  $\Delta$ is the maximum superconducting energy 
357: gap.}
358: \label{fig4}
359: \end{figure}
360: There was no evidence from thermodynamic measurements for nodes in 
361: the gap as in heavy fermion superconductors, except for an early
362: report of a non-exponential temperature dependence of the Knight 
363: shift \cite{TAKIGAWA}.  And the cuprates were 
364: viewed as quite disordered (doping being achieved by chemical 
365: substitution), which is known to be pair breaking for unconventional 
366: superconductors.  This was despite the prediction of d-wave pairing 
367: from both spin fluctuation models \cite{BICKERS} and RVB theory 
368: \cite{GROS,KOTLIAR}.  But this begin to change when NMR measurements found 
369: a $T^3$ variation of the spin lattice relaxation rate, as found in 
370: heavy fermion superconductors, as opposed to the exponential behavior 
371: found in s-wave superconductors \cite{MARTINDALE}.
372: This was soon followed by penetration depth
373: measurements, where a corresponding linear in T behavior was found
374: \cite{HARDY}.  At the same time, angle resolved 
375: photoemission measurements gave direct spectroscopic evidence for 
376: nodes in the gap \cite{SHEN93}.
377: 
378: Of course, the possibility still remained that the order parameter 
379: was s-wave, but with a highly anisotropic gap.  These worries were 
380: put to rest once and for all by phase sensitive measurements.  The 
381: first of these was by van Harlingen's group.  Following suggestions 
382: by Leggett, they formed SIS tunneling junctions on the ac and bc faces 
383: of YBCO, using an ordinary s-wave superconductor as the 
384: counterelectrode.  These two junctions were then connected, and the 
385: superconducting phase difference measured from the dependence
386: of the Josephson critical current on applied magnetic field.
387: They found exactly the $\pi$ 
388: phase shift expected for a d-wave state (which differs by a minus 
389: sign between the two orthogonal a and b directions) \cite{DALE}.
390: This was soon followed by the tricrystal grain boundary experiments 
391: of Tsuei and Kirtley \cite{TSUEI}, where three grain boundaries 
392: at different orientations were brought together at a point.  Thus, about
393: the tricrystal point, there are three junctions.  Each junction will act as a
394: ``zero" junction or a $\pi$ junction depending on the superconducting
395: phase difference across the junction.  If the number of such $\pi$ junctions
396: is odd, then a half integral flux quantum will appear at the tricrystal point.
397: The advantage of this method is that by varying the crystallographic
398: orientation of the three grains, the symmetry of the order parameter can
399: be mapped out in detail.  The net result was that for only those orientations
400: where d-wave symmetry predicted a half integral flux quantum was one 
401: observed.  As YBCO is orthorhombic, though, 
402: there still remained an out (since s-wave and d-wave are the same 
403: group representation in that crystal structure), but the tricrystal 
404: experiments were repeated for Tl2201, which has 
405: tetragonal symmetry, with the same results \cite{TLPH}.  After that, 
406: there was no question anymore about the order parameter symmetry, and 
407: for these pioneering efforts, four of the researchers were awarded the 
408: Buckley prize in 1998.
409: 
410: Perhaps the most unusual finding, though, is the difference of the 
411: dynamics between the normal and superconducting states.  As will be
412: discussed below, the normal state of the cuprates (away from 
413: the overdoped side of the phase diagram) does not appear to be a 
414: Landau Fermi liquid.  On the other hand, a variety of experiments, 
415: first microwave conductivity \cite{BONN}, then thermal 
416: conductivity \cite{ONG}, infrared 
417: conductivity \cite{PUCH}, and photoemission \cite{ADAM}, revealed that 
418: the scattering rate of the electrons at low energies
419: drops precipitously in the superconducting state (right panel, Fig.~\ref{fig5}).  At 
420: low temperatures in YBCO, mean free paths of the order of microns 
421: have been inferred for the electrons, as opposed to the very short mean 
422: free paths found in 
423: the normal state.  This strong loss in inelastic scattering would be unusual
424: for an electron-phonon mediated superconductor, since the phonons are not 
425: gapped in the superconducting state.  The implication, then, is that 
426: the primary scattering is electron-electron like in character, 
427: and thus is strongly reduced in the superconducting state since the 
428: electrons become gapped.  This can be easily seen from the lowest 
429: order Feynman diagram (left panel, Fig.~\ref{fig5}, showing an electron 
430: scattering off a particle-hole excitation), since every internal line in the diagram 
431: is gapped by $\Delta$ (a 
432: point first noted by Nozieres in his famous book\cite{NOZ}).
433: \begin{figure}
434: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig5.eps}}}
435: \caption{Left panel: lowest order Feynman diagram for 
436: electron-electron scattering.  Right panel: resulting temperature 
437: dependence of the zero energy scattering rate.  $T_{c}$ is the superconducting 
438: transition temperature (dotted line).}
439: \label{fig5}
440: \end{figure}
441: This obviously points to an electron-electron 
442: origin to the pairing as well.
443: 
444: This brings us to a consideration of the rest of the phase diagram.
445: As discussed above, cuprates should be thought of as doped 
446: Mott insulators.  What this means is that for low doping, the number 
447: of carriers is small.  As the superconducting phase is conjugate to 
448: the number operator, this implies that phase fluctuations could play 
449: an important role on the underdoped side of the phase diagram.  Again, this was 
450: realized early on by Anderson, who proposed that the doped holes would 
451: only be phase coherent 
452: below a temperature which scaled linearly with 
453: doping \cite{RVB2}.  This should be contrasted with the ``pairing'' 
454: scale, which within the RVB model would be maximal for the insulator,
455: and then drops to zero on the overdoped side when the bandwidth $xt$
456: of the doped holes
457: becomes comparable to the superexchange energy $J$ \cite{BZA,GROS}.
458: \begin{figure}
459: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig6.eps}}}
460: \caption{Two proposed theoretical phase diagrams for the cuprates:  
461: RVB picture (left panel) \cite{LEE} and the quantum critical scenario (right 
462: panel).}
463: \label{fig6}
464: \end{figure}
465: These two crossing 
466: lines led to the 
467: proposal of a generic ``RVB'' phase diagram (left panel, Fig.~\ref{fig6}) \cite{LEE} 
468: composed of 
469: four phases, a superconductor (bottom quadrant), a Fermi liquid 
470: (right quadrant), a strange metal phase (upper quadrant), 
471: and a spin gap phase (left quadrant, now known as the 
472: pseudogap phase).  In this picture, only the superconducting phase 
473: (which lies below both crossing lines) should be considered as having 
474: true long range order, otherwise, these ``phase'' lines should be 
475: considered as crossover lines.
476: 
477: The first of these ``phases'' which was studied in detail was the 
478: strange metal phase.  Transport measurements revealed that the 
479: resistivity was dead linear in temperature over a large range, 
480: the most amazing example of this being in single 
481: layer Bi2201, where linearity persisted down to 10K (when 
482: superconductivity finally occurred) \cite{Bi2201}.  No saturation at high 
483: temperatures was observed as occurs in A15 superconductors.  This 
484: behavior was further confirmed by a generalized Drude analysis of 
485: infrared data, which shows a scattering rate linear in energy up to half an 
486: eV \cite{PUCH}.  These striking observations led Varma and colleagues to 
487: propose the so-called marginal Fermi liquid phenomenology for the 
488: strange metal phase \cite{MFL}.  In this model, the electrons are 
489: assumed to be scattering off a bosonic spectrum which is linear in 
490: energy up to an energy scale T, then constant afterwards.  Because of 
491: this, no energy enters the problem except the temperature (modulo an 
492: ultraviolet cut-off), a phenomenon referred to as quantum critical 
493: scaling.  This in turn has led to the proposal of an alternate (to the 
494: ``RVB'') phase diagram based on a quantum critical point (right panel,
495: Fig.~\ref{fig6}).  In such a 
496: picture, the ordered phase (to the left of the critical point) would 
497: correspond to the pseudogap phase, its disordered analogue (to the 
498: right of the critical point) to the Fermi liquid phase, and the 
499: quantum critical regime (above the critical point) to the strange 
500: metal phase.  The superconducting ``dome'' surrounds the critical 
501: point, screening it like an event horizon of a black hole.  We note 
502: that the Fermi liquid/strange metal boundary is a crossover line in 
503: both phase diagram scenarios, but the strange metal/pseudogap boundary is a 
504: crossover line in the RVB model and a true phase line in the quantum 
505: critical model.  An exception is in certain antiferromagnetic quantum critical
506: scenarios where the ``phase line" corresponds to short range 2D order \cite{AC}.
507: 
508: This brings us to the most controversial aspect of the cuprate field, 
509: the nature of the pseudogap phase \cite{TIMRPP,VARENNA}.
510: The first experimental indication 
511: of such a phase was from NMR measurements by the Bell group, which 
512: showed that the spin lattice relaxation rate of underdoped cuprates begins to 
513: decrease well above $T_c$ (left panel, Fig.~\ref{fig7}) \cite{WARREN}.  
514: A similar decrease is seen in
515: the Knight shift \cite{ALLOUL}, which measures 
516: the bulk susceptibility (NMR measuring the zero energy limit of the imaginary 
517: part of the dynamic susceptibility divided by the energy).  A 
518: signature of a gap was also evident in infrared measurements, which 
519: showed a dip in the conductance
520: separating the low energy Drude peak from the so-called 
521: mid-infrared bump \cite{COOPER,ROTTER}, with the temperature dependence of 
522: the conductance near the dip energy  \cite{ROTTER} scaling 
523: with the spin lattice relaxation rate \cite{TAKI2}.
524: The resulting sharpening of the Drude
525: peak leads to a decrease of the planar resistivity in the pseudogap 
526: phase \cite{RESA}.
527: \begin{figure}
528: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig7a.eps}}
529: \epsfxsize=0.5\textwidth{\epsfbox{fig7b.eps}}}
530: \caption{Experimental evidence for a pseudogap.  Left panel is the NMR 
531: relaxation rate for various samples of Bi2212 \cite{ISHIDA}, with a 
532: suppression of $1/T_{1}T$ (spin gap) starting at $T^*$ well above $T_{c}$ 
533: for underdoped 
534: samples.  Right panel is the c-axis
535: conductivity for underdoped YBCO, with a pseudogap which
536: fills in with temperature \cite{PUCH}.  The inset shows that the 
537: subgap conductance scales with the Knight shift.}
538: \label{fig7}
539: \end{figure}
540: But the most dramatic effect was in 
541: c-axis polarized infrared measurements, which showed a significant gap 
542: at low energies (no Drude peak), with the temperature dependence of the subgap 
543: conductance \cite{HOMES} tracking the Knight shift (right panel, Fig.~\ref{fig7})
544: \cite{TAKI2}.  This 
545: leads to an insulating up turn below the pseudogap temperature, 
546: $T^*$, in the c-axis resistivity \cite{RESC}.
547: 
548: What brought the pseudogap effect to the forefront, though, was its 
549: observation by angle resolved photoemission \cite{MARSHALL,LOESER,DING}.
550: These experiments found that although the quasiparticle peak in the 
551: spectral function was destroyed above $T_c$, the spectral gap 
552: persisted to the higher temperature, $T^*$.  This so-called leading 
553: edge gap had a similar magnitude and momentum anisotropy as the 
554: superconducting energy gap, leading to the speculation that this gap 
555: was a precursor to the superconducting gap, that is, that the 
556: pseudogap phase represented pairs without long range phase
557: order \cite{DING}.  This picture was consistent with the NMR and Knight 
558: shift data, in that pair formation is equivalent to singlet formation, 
559: and thus the Knight shift and the spin lattice relaxation rate should 
560: decrease accordingly \cite{MOHIT92}.  The strong coupling limit of this 
561: picture is simply the RVB physics mentioned above, where the pseudogap 
562: state corresponds to d-wave pairing of spins.  This picture received 
563: further support by later ARPES experiments which showed that the 
564: pseudogap's minimum gap locus in momentum space
565: coincided with the normal state Fermi 
566: surface.  That is, the pseudogap is locked to the Fermi surface,
567: as would be expected for a Q=0 instability \cite{DING97} (in 
568: superconductors, pairs have zero center of mass momentum).  
569: Later tunneling experiments were found to be in support of this 
570: picture as well \cite{FISCHER}.
571: 
572: Despite this, the situation, even from ARPES, remains controversial.  
573: ARPES experiments reveal that the pseudogap turns off at different 
574: momentum points at different temperatures, leading to the presence of 
575: temperature dependent Fermi arcs (Fig.~\ref{fig8}) \cite{NAT98}.
576: \begin{figure}
577: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig8.eps}}}
578: \caption{Momentum anisotropy of the pseudogap from ARPES.  Spectral 
579: gap around the Fermi surface is plotted in the left panel (SC is the 
580: superconducting state, PG the pseudogap phase), the locus of gapless 
581: excitations in the right panel (NS is the normal state Fermi surface, 
582: PG the Fermi arc of the pseudogap phase, and SC the node of the 
583: d-wave superconducting state).}
584: \label{fig8}
585: \end{figure}
586: These arcs persist to 
587: low doping; they have even been observed for LSCO at a 3\% doping level,
588: where the system is on the verge of magnetism \cite{YOSHIDA}.  Some 
589: authors have taken this as evidence that these arcs represent one side 
590: of a small hole pocket, the latter picture expected when doping the 
591: magnetic insulating phase (the idea being that SDW coherence factors 
592: suppress the intensity on the back side of the pocket).  This magnetic 
593: precursor scenario is one of the leading alternates to the preformed 
594: pairs picture.  Despite much searching, though, no ARPES experiment 
595: to date as ever seen a true hole pocket centered about $(\pi/2,\pi/2)$.
596: 
597: Another explanation has been put forward that the pseudogap represents 
598: an orbital current phase.  This was implicit in certain treatments of 
599: the RVB model, which predicted at low dopings the presence of a 
600: so-called staggered flux phase, which is quantum mechanically 
601: equivalent to the d-wave pair state in the zero doping 
602: limit \cite{LEESF}.
603: This has been generalized to the d density wave state, first 
604: discussed by Heinz Schulz \cite{SCHULZ}, but popularized by Laughlin 
605: and colleagues \cite{NAYAK}.  A related picture has been put forth by 
606: Varma, where his orbital current phase is the result of a non-trivial 
607: projection of the three band Hubbard model onto the low energy 
608: sector \cite{VARMA}.  Some experimental evidence for such a state was 
609: obtained from inelastic neutron scattering which indicated a 
610: momentum form factor inconsistent with simple Cu spins \cite{MOOK}, but after 
611: studies by several groups, the feeling is that the observed effect may
612: represent an impurity phase (always a problem for neutrons given the 
613: large crystals needed for such measurements).  The latest evidence, 
614: though, has been given again by ARPES, where Campuzano's group has 
615: done measurements with circularly polarized light \cite{NAT02}.  What 
616: they have found is the presence of chiral symmetry breaking below 
617: $T^*$.  The momentum dependence of the effect, though, is not what is 
618: expected from the d density wave scenario, though one of the two 
619: orbital current states proposed by Varma appears to be consistent 
620: with these observations \cite{VARMA02}.
621: 
622: The main debate, though, is whether the pseudogap phase represents a 
623: state with true long range order (which the neutron and circularly 
624: polarized ARPES give some evidence for), or simply some precursor 
625: phase.  If it is the former, then the preformed pairs scenario is 
626: probably wrong, unless the ordering is some parasitical effect.  Long 
627: range order is certainly in line with a quantum critical point
628: scenario.  Additional support for such a scenario comes from
629: various measurements by Loram and Tallon (Fig.~\ref{fig9}), who claim that the 
630: pseudogap phase line passes through the superconducting dome and goes 
631: to zero at some critical point within the dome
632: (19\% doping) \cite{LORAM}.
633: \begin{figure}
634: \centerline{\epsfxsize=1.0\textwidth{\epsfbox{fig9.eps}}}
635: \caption{Left:  Tallon/Loram picture of the phase diagram, with a
636: quantum critical point at $p$=0.19 where $E_g$, the pseudogap energy scale, 
637: vanishes ($p$ is the hole doping) .  $SG$ is a spin glass phase.
638: Middle: Variation with doping of $E_g$ and the superconducting condensation energy
639: $U_0$ as extracted from specific heat data on YBCO.
640: Right: Collapse of the superconducting dome about the $E_g$ crossover line
641: with increasing cobalt doing for Bi2212.  From Ref.~\cite{LORAM}.}
642: \label{fig9}
643: \end{figure}
644: One of the strongest points given as evidence 
645: for their conjecture is that upon impurity doping, 
646: the superconducting dome appears to collapse about the pseudogap 
647: phase line (right panel, Fig.~\ref{fig9}).  In their picture, the specific heat data indicate a 
648: loss of states in the pseudogap phase.  This implies that the 
649: pseudogap eats up part of the Fermi surface, leaving a smaller part 
650: available for pairing, thus explaining the collapse of $T_c$ on the 
651: underdoped side.  This ``mean field'' picture is in total contrast to 
652: the phase fluctuation picture discussed in the context of the 
653: precursor pairing scenario \cite{EMERY95}.
654: 
655: Recently, there has been a new measurement which comes out in support 
656: of the precursor pair scenario.  Ong's group has measured the Nernst 
657: effect, the coefficient of a higher order transport tensor which is 
658: very small in normal metals, but is appreciable in 
659: superconductors because of the presence of vortices.  What they find 
660: is a sizable Nernst signal on the underdoped side of the phase 
661: diagram which persists well above $T_c$, though not as high in 
662: temperature as the ARPES pseudogap \cite{NERNST}.  The only 
663: explanation that has been put forth for this amazing observation is 
664: that the pseudogap phase does indeed contain vortices.  This may be 
665: connected to the results of STM measurements, which reveal that the 
666: pseudogap forms in the vortex cores in the superconducting state for 
667: underdoped samples \cite{CORE}.
668: 
669: \subsection{Inhomogeneities}
670: 
671: How one crosses over from the Fermi arc state to the insulator is 
672: still an unresolved issue.  In one scenario, the chemical potential jumps from the middle 
673: of the Mott-Hubbard gap to either the lower Hubbard band upon hole 
674: doping, or the upper Hubbard band upon electron doping.  There is some 
675: evidence of this from photoemission. In particular, 
676: the Fermi arc is in a momentum region near the $(\pi/2,\pi/2)$ 
677: point, and the latter is the top of the valence band in the 
678: insulator, as seen by photoemission in Sr$_2$CuO$_2$Cl$_2$ \cite{WELLS}.
679: This picture has been bolstered recently by 
680: photoemission experiments on the sodium doped version of this 
681: insulator, which also find a Fermi arc \cite{NADOP}.
682: 
683: The other 
684: scenario is that upon doping, one creates new states inside the gap.  
685: Shen's group has given evidence that the latter scenario occurs in 
686: LSCO, and argues that this is associated with the strong inhomogeneities present in 
687: that material \cite{RMP03}, though it should be remarked that 3\% doped 
688: LSCO has the same Fermi arc that is seen in other hole doped cuprates
689: such as Na doped Sr$_2$CuO$_2$Cl$_2$ \cite{YOSHIDA} and underdoped 
690: Bi2212 \cite{MARSHALL,NAT98}.
691: 
692: This brings us to the question of stripes.  At low doping, materials can 
693: be subject to electronic phase separation.  This tendency occurs 
694: since each doped hole breaks four magnetic bonds, and thus this 
695: magnetic 
696: energy loss can be minimized by the holes clumping together.  This
697: clumping is opposed by the Coulomb repulsion of the holes.  This led 
698: to the picture that the doped holes, as a compromise,
699: might form rivers of charge, known 
700: as stripes \cite{ZAANEN,EMKIV}.
701: 
702: The first evidence for such stripes was given by Tranquada and 
703: co-workers 
704: using neutron scattering (Fig.~\ref{fig10}) \cite{JOHN95}.  To understand their result, 
705: it should be noted that LSCO has a peculiarity in its phase diagram.  
706: LSCO is normally orthorhombic, the tetragonal phase existing 
707: either at high temperatures or under pressure.  
708: Near 1/8 hole doping, though, LSCO has a tendency to distort from its normal 
709: orthorhombic phase to another distorted phase known as low 
710: temperature tetragonal, which differs from the high temperature 
711: tetragonal phase mentioned above \cite{AXE}.
712: \begin{figure}
713: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig10.eps}}}
714: \caption{Stripe picture.  Left panel illustrates stripes for $\delta$=1/8 
715: doping, arrows represent spins, dark circles doped holes.  Right 
716: panel plots the resulting neutron scattering peaks (averaged between 
717: the x and y directions), with charge peaks at $\pm 2\delta$
718: about (0,0) and spin peaks at $\pm \delta$ about (0.5,0.5).
719: Adapted from Ref.~\cite{JOHN95}.}
720: \label{fig10}
721: \end{figure}
722: This LTT phase is 
723: stabilized by neodynium doping.  What Tranquada and co-workers found was that the 
724: LTT phase exhibited long range ordering consistent with the 
725: formation of a 1D density wave state.  Both charge and spin ordering 
726: occurs, but the former sets in at a higher temperature.  This 
727: stripe formation is consistent with later photoemission \cite{SHENND} and 
728: transport \cite{ANDO} measurements, which again indicate 1D behavior.
729: 
730: What remains controversial, though, is whether stripes exist only
731: for this anomalous Nd-doped LSCO compound.  Strong incommensurate 
732: magnetic spots are seen by inelastic neutron scattering for various dopings 
733: in LSCO \cite{NS-LSCO} and YBCO \cite{NS-YBCO}, and have been taken as 
734: evidence for the existence of dynamic stripes \cite{MOOK02}, given their 
735: resemblance to the static spot pattern of Nd-doped LSCO \cite{JOHN95}.  On the 
736: other hand, these spot patterns can also be reproduced from standard 
737: linear response calculations based on the known Fermi 
738: surface geometry \cite{NORM00}.
739: 
740: What is clear, though, is that there is a definite tendency for 
741: underdoped materials to exhibit electronic inhomogeneity.  The most 
742: dramatic example of this has been recently provided by STM studies on 
743: underdoped Bi2212 by Davis' group \cite{PAN,LANG}.  What they find is the 
744: existence 
745: of large gap regions (which have spectra reminiscent of the pseuodgap 
746: phase) imbedded in smaller gap regions, with the relative fraction of 
747: the larger gap regions increasing with underdoping, similar to earlier 
748: results by Roditchev's group \cite{CREN}.  The large gap 
749: domains have a size of order 30$\AA$.  This granular picture would 
750: certainly suggest that the pseudogap phase is not a simple precursor 
751: to the superconducting phase as has been asserted by previous ARPES 
752: and STM studies.
753: 
754: More recently, the same group has seen a charge density wave state 
755: associated with the vortex cores which they inferred from Fourier 
756: transformation of the real space STM spectra \cite{HOFFMAN}.  An 
757: anisotropy of the pattern gave some indication that this might involve 
758: stripe formation.  But even more recently, the same group has done a 
759: Fourier transformation of zero field data \cite{HOFF2}.  They find a weaker 
760: inhomogeneity in this case, but more interestingly, the Fourier peaks
761: dispersed with energy.  Although the interpretation of such Fourier 
762: transforms remain controversial (Kapitulnik's
763: group \cite{HOWALD} sees similar patterns which they attribute to 
764: stripe formation), the latest 
765: results \cite{MCELROY} are consistent with Friedel oscillations from 
766: impurities whose momentum wavevectors can be used to map out 
767: the Fermi surface and gap anisotropy.  The results are consistent with 
768: previous ARPES studies, and have been taken as support of the
769: interpretation of incommensurability in neutron scattering as due to 
770: the Fermi surface geometry, as opposed to stripes.
771: 
772: \subsection{Electron Doped Materials}
773: 
774: Electron doped materials have been less studied, mainly due to 
775: metallurgical problems.  What has been learned, though, is 
776: that the magnetic phase extends much further in doping than on the hole 
777: doped side \cite{TOKURA,LUKE90}.  The superconducting phase has a 
778: lower $T_c$ 
779: than on the hole doped side, probably for the same reason.  A pseudogap 
780: phase is observed which appears to be a precursor to the magnetic 
781: phase in that they exist over the same doping range \cite{ONOSE}, though 
782: it should be remarked that the pseudogap 
783: seen is not the leading edge gap (discussed above in the context of 
784: ARPES), but rather the ``high energy pseudogap'' to be discussed later 
785: on in the ARPES section.
786: 
787: The symmetry of the superconducting order parameter is still somewhat 
788: controversial in these materials.  Earlier penetration depth 
789: measurements \cite{WU93} and point contact tunneling \cite{JOHN90}
790: showed behavior expected for s-wave pairing, but more recent 
791: penetration depth measurements have found a power law temperature 
792: dependence consistent with a disordered d-wave 
793: state \cite{NCCO}.  Recent ARPES measurements are also consistent 
794: with d-wave symmetry \cite{SATO,PETER1}, but these experiments are near 
795: the resolution 
796: limit because of the small energy gap.  It should be mentioned that 
797: the tri-crystal experiments mentioned above in the context of hole 
798: doped superconductors have been performed for the electron-doped case as 
799: well, and again find a half integral flux quantum in geometries predicted 
800: by d-wave symmetry \cite{TSUEI2}.  Based on these developments, it is 
801: fairly certain that the pairing symmetry is d-wave in these systems.
802: 
803: More recently, Raman studies by 
804: Blumberg and co-workers \cite{GIRSH} find evidence that the 
805: superconducting gap maximum 
806: is displaced away from the Fermi surface crossing
807: along $(\pi,0)-(\pi,\pi)$ (as expected for a d-wave gap based on near 
808: neighbor pairs)
809: to the ``hot spots'' (where the Fermi surface crosses the 
810: magnetic Brillouin zone boundary).  This result is consistent with spin 
811: fluctuation mediated pairing if the magnetic correlation length is
812: long (not surprising, given the persistence of long range magnetic order 
813: over a larger part of the electron doped phase diagram).  
814: It should be noted that ARPES sees
815: an intensity suppression at these ``hot spots'' \cite{PETER1} associated 
816: with the formation of the ``high energy pseudogap'' \cite{PETER2} 
817: mentioned above.  In 
818: addition, at low dopings, low energy spectral weight is found around 
819: the $(\pi,0)$ point \cite{PETER3}, as opposed to the $(\pi/2,\pi/2)$ 
820: point characteristic of the hole-doped material.  This electron-hole 
821: asymmetry is what is expected if the chemical potential jumps to the 
822: upper Hubbard band upon electron doping.
823: 
824: \section{Theory}
825: 
826: \subsection{BCS}
827: 
828: The first microscopic theory of superconductivity, the much 
829: celebrated BCS theory \cite{BCS}, took many years to come about \cite{SUPER}.
830: The reason was 
831: that the machinery needed to construct a proper many-body theory of 
832: electrons did not emerge until the 1950s.  The motivation behind the 
833: theory was the isotope experiments of 1950 and their simultaneous 
834: prediction by Frohlich based on the electron-phonon interaction.  
835: Bardeen also understood the critical role that the concept of an energy gap 
836: would play in the ultimate theory.  Once Leon Cooper, an 
837: expert in many-body theory, joined Bardeen's group as a postdoc, 
838: progress was rapidly made.
839: 
840: What was known by that time was that the electron-phonon interaction 
841: could provide attraction among the electrons.  The way this 
842: works is as follows (left panel, Fig.~\ref{fig11}).  Positive ions are attracted to an electron 
843: because of the Coulomb interaction.  But, the ion dynamics are slow 
844: because of their heavy mass.  Thus, once the electron moves away, 
845: another electron can move into this ``ionic hole'' before the ions 
846: have a chance to relax back.  This provides attraction at the same 
847: point in space which can lead to pair 
848: formation.  The interaction is retarded in time, though, which is what 
849: ultimately puts a limit on the transition temperature (energy, and 
850: thus temperature, being conjugate to time).
851: 
852: Such a ``pair'' theory had been proposed in the past, but the physics 
853: of conventional superconductors does not resemble simple Bose 
854: condensation, despite the fact that a pair of fermions behaves quantum 
855: mechanically like a 
856: boson.  The key discovery by Cooper was the concept of Cooper 
857: pairs.  The idea is that one is dealing with a degenerate system, 
858: that is, a filled Fermi sea.  The problem Cooper considered was 
859: two electrons sitting in unoccupied states above the Fermi sea.  As 
860: the temperature is lowered, the particle-particle response function 
861: diverges logarithmically because the 
862: Fermi distribution function of the electrons approaches a step 
863: function.  This divergence is strongest when the two electrons are in
864: time reversed 
865: states (that is, a state k and a state -k, thus the center of mass 
866: momentum of the pair is Q=0).
867: Note the difference from the particle-hole response at Q=0, which 
868: simply measures the density of states at the Fermi energy.
869: 
870: This logarithmic divergence is cut-off at some ultraviolet energy 
871: scale, which in the electron-phonon problem is the Debye energy, 
872: $\omega_D$.  This 
873: leads to a bare response function that goes as $\chi_0=N\ln(\omega_D/T)$ 
874: where $N$ is the density of states.
875: \begin{figure}
876: \centerline{\epsfxsize=0.6\textwidth{\epsfbox{fig11.eps}}}
877: \caption{Electron-phonon interaction leads to attraction (left 
878: panel).  Arrows joining circles show displaced ions; the time scale 
879: of these ions for relaxation back is slow compared to the electron 
880: dynamics.  Right panel is the ladder sum for repeated electron-phonon 
881: scattering which leads to an electron pairing instability.}
882: \label{fig11}
883: \end{figure}
884: Now 
885: one sums a ladder series (repeated scattering of the two electrons, right
886: panel, Fig.~\ref{fig11}),
887: which leads to an expression for the 
888: full response function of $\chi=\chi_{0}/(1-V\chi_0)$ where $V$ is 
889: the interaction.  For positive 
890: $V$ (attractive in our sign notation), the denominator will have a 
891: pole when $T_{div}=\omega_D e^{-1/\lambda}$ with $\lambda=NV$.  This is 
892: the famous Cooper pair divergence.
893: 
894: This doesn't answer the question of what the ground state is.  This 
895: problem was solved by Schrieffer, Bardeen's graduate student (thus 
896: BCS).  Based on Cooper's solution, he guessed the many-body ground 
897: state at T=0 (a rare accomplishment, the other well known example of 
898: this was Laughlin's guess for the fractional quantum Hall state).  It 
899: is of the form $\prod_k (u_k + v_k c^{\dag}_k c^{\dag}_{-k} |0>$.  
900: Here $|0>$ is the vacuum and $u,v$ are ``coherence'' 
901: factors (the sum of whose squares equals one).  What can be seen here 
902: is that the BCS ground state is a superposition of states where the 
903: pair $k,-k$ is either occupied or filled.  Solving the variational 
904: problem (equivalent to replacing the product of the two creation 
905: operators by a c number), BCS found that
906: $u_k^2,v_k^2 = 1/2(1 \pm \epsilon_k/E_k)$ where 
907: $E_k=\sqrt{\epsilon_k^2+\Delta_k^2}$ and $\Delta_k$, gotten from 
908: solving an integral equation (the so-called gap equation), has the same
909: form as $T_{div}$ above.
910: 
911: Note that the Fermi distribution function has been replaced by 
912: $v_k^2$, thus leading to particle-hole mixing (Fig.~\ref{fig12}).
913: \begin{figure}
914: \centerline{\epsfxsize=0.4\textwidth{\epsfbox{fig12.eps}}}
915: \caption{Momentum distribution function for the normal state (NS) and 
916: superconducting state (SC).  $\xi_0$ is the BCS correlation length.}
917: \label{fig12}
918: \end{figure}
919: In essence, the BCS 
920: instability is a consequence of Nature's abhorence of singularities 
921: (in this case, the step function behavior of the Fermi function at 
922: T=0).  This distribution is smeared over a momentum range of $\sim 
923: \Delta/v_F$ where $v_F$ is the Fermi velocity, thus defining the 
924: inverse correlation length.  Excitations from the 
925: ground state can be formed by breaking up a pair.  These excitations 
926: are of the form $\gamma^{\dag}_k = u_k c^{\dag}_k - v_k c_{-k}$ and have an 
927: energy $E_k$.  That is, on the Fermi surface, the quantity $\Delta$, 
928: which is related to the order parameter, is 
929: nothing more than the spectral gap, and thus the energy gap emerges 
930: quite naturally from the theory.
931: 
932: What allows this conceptually simple picture to work is Migdal's 
933: theorem \cite{BOB}.  It states that the single particle self-energy 
934: can be treated to lowest order, since $\omega_D/E_F$ is a small 
935: expansion parameter (where $E_F$ is the Fermi energy).  Thus, the only 
936: diagram series which has to be summed is the particle-particle ladder 
937: mentioned above.  Such a theorem obviously does not apply if the 
938: pairing is due to electron-electron interactions.
939: 
940: \subsection{Spin Fluctuation Models}
941: 
942: As mentioned above, the electron-phonon attraction is local in space 
943: and retarded in time.  This leads to L=0 pairs.  By fermion 
944: antisymmetry, this requires that the pair state be a spin singlet.  
945: For the case of electron-electron interactions, L=0 pairs are usually 
946: not favored (because of the direct Coulomb repulsion between the 
947: electrons).  In fact, one might wonder how one can ever get an
948: ``attractive'' interaction in this case.
949: 
950: Let us start with the nearly ferromagnetic case \cite{FAY}.
951: \begin{figure}
952: \centerline{\epsfxsize=0.3\textwidth{\epsfbox{fig13.eps}}}
953: \caption{Particle-particle diagram for the spin fluctuation case.  
954: Note the particle-hole ladder sum buried inside this diagram.}
955: \label{fig13}
956: \end{figure}
957: The 
958: particle-particle ladder sum in this case involves exchanging the ends 
959: of one of the particle lines, thus representing a particle-hole ladder sum 
960: buried inside of a particle-particle one (Fig.~\ref{fig13}).  Thus, the diverging 
961: particle-hole response (representing a ferromagnetic instability) drives a 
962: diverging particle-particle response.  For S=1 pairs, this is 
963: attractive.  In essence, the bare triplet interaction is zero (due to 
964: the Pauli exclusion principle) and the induced interaction is 
965: attractive (representing the tendency for an up spin electron to have 
966: another up spin electron nearby).  By fermion antisymmetry, the L 
967: state must be odd, thus allowing the two electrons in the pair to avoid 
968: coming too close to one another (thus minimizing the direct Coulomb 
969: repulsion).
970: 
971: For heavy fermions and cuprates, though, the nearly antiferromagnetic 
972: case is of more interest.  In that case, one again wants to avoid the 
973: direct Coulomb repulsion, but now a spin up electron wants to have a 
974: spin down electron nearby.  In the absence of spin-orbit, this implies 
975: S=0, L=2 pairs.  For strong spin-orbit, $S=1,S_z=0$ pairs can be 
976: stable as well, which is the basis of the ``f-wave'' scenario postulated by 
977: Norman in the case of heavy fermions \cite{MIKE92}, but this is not 
978: relevant for the cuprate case.
979: 
980: The above considerations on a 2D square lattice leads to a pair state 
981: with $d_{x^2-y^2}$ symmetry \cite{1986,BICKERS}.  In fact, the theory 
982: of this is a bit counterintuitive.  Unlike the nearly ferromagnetic 
983: case, in the nearly antiferromagnetic case, the interaction is always 
984: repulsive (in a momentum space representation, left panel, Fig.~\ref{fig14}), 
985: and in fact is most 
986: repulsive at $Q=(\pi,\pi)$ where the antiferromagnetic instability would 
987: occur.  But the d-wave version of $\Delta_k$ changes sign under translation 
988: by $Q$ 
989: (it is of the form $\cos(k_x)-\cos(k_y)$), and this sign change 
990: compensates for the repulsive sign of the interaction when solving the 
991: integral (gap) equation for $\Delta_k$.  In real space, the picture is more 
992: clear (right panel, Fig.~\ref{fig14}) \cite{DOUG}.
993: \begin{figure}
994: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig14.eps}}}
995: \caption{Effective interaction for spin fluctuation mediated 
996: pairing (AF case).  Left panel is for momentum space (overall repulsive, peaked 
997: at $(\pi,\pi)$), right panel for real space (repulsive on-site,
998: with first attractive minimum at a near neighbor separation).  Adapted from 
999: Ref.~\cite{DOUG}.}
1000: \label{fig14}
1001: \end{figure}
1002: The on-site interaction is repulsive (which is 
1003: why the overall interaction is repulsive in momentum space), but this 
1004: interaction contains Friedel oscillations, with the first (attractive) 
1005: minimum at a near neighbor separation, representing the tendency of
1006: opposite spin electrons to be on neighboring sites.
1007: 
1008: Note that retardation does not play the central role in the above arguments
1009: as in the phonon case.  In 
1010: essence, there is no Migdal's theorm in this case \cite{HL}, and so 
1011: one may question such a theory which does not take into account vertex 
1012: corrections.  The justification that is often given is that if one 
1013: considers the electrons and spin fluctuations as separate objects (like 
1014: electrons and phonons), then as the spin fluctuations are ``slow'' 
1015: relative to electrons, one gets something like a Migdal's theorem.  
1016: But this simple argument usually breaks down \cite{HL}, though 
1017: Chubukov has recently made arguments about why an effective Migdal 
1018: theory would apply in the cuprate case \cite{AC}.  Regardless, 
1019: feedback effects definitely have to be considered whenvever 
1020: electron-electron interactions are involved, since the spin 
1021: fluctuation propagator is drastically changed by the introduction of 
1022: the superconducting gap for the electrons \cite{AC}.  The 
1023: classic example of this is the stabilization of the $A$ phase relative to 
1024: the $B$ phase in superfluid $^3He$ \cite{BA}.
1025: 
1026: \subsection{RVB}
1027: 
1028: The spin fluctuation theory is essentially a weak coupling approach.  The RVB 
1029: picture mentioned in the introduction is the strong coupling 
1030: version of the spin fluctuation approach \cite{RVB} (though 
1031: Anderson differs on this \cite{ADV}).  The amazing thing was how 
1032: quickly the RVB concept emerged (Anderson first spoke on this 
1033: before the discovery of YBCO).  It has certainly been 
1034: controversial (one well known scientist, whose name will not be 
1035: mentioned here, quipped that RVB actually stood for ``rather vague 
1036: bullshit'').
1037: 
1038: To understand this approach, consider the undoped insulator, where 
1039: there is one Cu spin per site (the Cu being in a $d^9$ configuration).
1040: The ground state of this system is an antiferromagnet.  The reason is
1041: that if the spins on each site are parallel, then they cannot virtually 
1042: hop because of the Pauli exclusion principle, but they can if the 
1043: spins are antiparallel.  In this case, the virtual hopping leads to 
1044: an energy lowering of $J = t^2/U$ per bond, where $t$ is the effective 
1045: Cu-Cu hopping integral, and $U$ the Coulomb repulsion for double 
1046: occupation.  Mean field theory would then predict that the arrangement 
1047: of spins forms a Neel lattice of alternating up and down spins (left panel,
1048: Fig.~\ref{fig15}).
1049: 
1050: \begin{figure}
1051: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig15.eps}}}
1052: \caption{Neel lattice (left panel) versus RVB (right panel).  
1053: The RVB state is a liquid of spin singlets.}
1054: \label{fig15}
1055: \end{figure}
1056: 
1057: Anderson, though, suggested that the Cu case was special, since the 
1058: spin of the single d hole was only 1/2.  Because of this, he 
1059: anticipated that quantum fluctuations would melt the Neel lattice, 
1060: leading to a spin liquid ground state, a fluid of singlet pairs of 
1061: spins (right panel, Fig.~\ref{fig15}) \cite{RVB}.  
1062: The ``RVB'' notation comes from the fact that a given spin 
1063: could be taken as paired with any one of its four neighbors, thus 
1064: each bond fluctuates from being paired to not paired.  This is 
1065: analogous to benzene rings, where each C-C link fluctuates between a 
1066: single bond and a double bond.
1067: Although it was discovered soon after \cite{INS} that the undoped 
1068: insulator is indeed a Neel lattice, this lattice does indeed ``melt'' 
1069: with only a few percent of doped holes.  This is rather easy to understand,
1070: since the kinetic energy of the doped holes is frustrated in the Neel state.
1071: 
1072: After Anderson's original conjecture, it was realized that at the mean 
1073: field level, there were a number of possible ground states for such a 
1074: spin fluid.  For the undoped case, these states are quantum
1075: mechanically equivalent, since
1076: the presence of a spin up state is equivalent to the absence of 
1077: a spin down state, an effective SU(2) symmetry \cite{SU2}.
1078: That is, various mean field decouplings of the 
1079: Hamiltonian are equivalent since
1080: $<c^{\dag}_{\uparrow}c^{\dag}_{\downarrow}> \equiv 
1081: <c^{\dag}_{\uparrow}c_{\uparrow}>$.
1082: The first decomposition is equivalent to a BCS pairing of spins, 
1083: the second to a bond current state (left panels, Fig.~\ref{fig16}).  
1084: For the spin pairing case, the 
1085: d-wave state is favored since it
1086: does the best job of localizing the two spins on neighboring 
1087: sites.
1088: \begin{figure}
1089: \centerline{\epsfxsize=1.0\textwidth{\epsfbox{fig16.eps}}}
1090: \caption{Two RVB states which are equivalent at half filling.  The 
1091: left panel is a d-wave pairing of spins, the middle panel a $\pi$ flux 
1092: state.  Dots are Cu ions, and arrows are bond currents.  Right:
1093: Variation of the RVB gap parameter, $\Delta$, and the superconducting
1094: order paramter, $\Delta_{SC}$, with doping \cite{GROS}.}
1095: \label{fig16}
1096: \end{figure}
1097: Its bond current equivalent is the $\pi$ flux phase state, 
1098: where the bond currents flow around an elementary plaquette (square
1099: formed from four Cu-Cu bonds), yielding 
1100: a net phase of $\pi$ per plaquette.
1101: 
1102: Upon doping, this SU(2) symmetry is broken to U(1).  At the mean 
1103: field level, the d-wave state has the lowest energy \cite{KOTLIAR,GROS}, as
1104: it minimizes the kinetic energy of the doped holes.  The 
1105: variational $\Delta$ associated with the d-wave state is maximal for 
1106: the undoped case and decreases linearly with doping
1107: (right panel, Fig.~\ref{fig16}).
1108: To understand the implications of this for superconductivity, one 
1109: should note that the above variational parameter applies to the 
1110: spins.  But only the doped holes carry the current.  As their density 
1111: increases linearly with doping, then the superfluid density of the real
1112: electrons (a product of spin and charge) varies 
1113: linearly with doping, despite the fact that the variational parameter 
1114: does not.  That is, there is a complete decoupling of $\Delta$ (the
1115: excitation gap) from the order parameter (the superfluid density),
1116: unlike in BCS theory.  These simple 
1117: mean field considerations have been confirmed by recent variational Monte 
1118: Carlo calculations of a ``projected'' d-wave BCS pair state (where double 
1119: occupied states are projected out) \cite{PARM}.  Such projected 
1120: states can also be shown to contain orbital current correlations \cite{ORB}.
1121: 
1122: Although a finite temperature generalization of RVB theory is 
1123: non-trivial, the overall phase diagram can be easily appreciated by 
1124: noting that the ``pairing'' temperature scale, $T_{RVB}$, will be 
1125: proportional to $\Delta$, and that the phase coherence temperature of 
1126: the doped holes will be proportional to the doping.  The net result 
1127: are two crossing lines with doping, with the spin gap phase in the left
1128: quadrant, 
1129: the Fermi liquid phase in the right quandrant, the superconducting 
1130: phase in the bottom quadrant (below both lines), and the strange metal 
1131: phase in the upper quadrant (see left panel, Fig.~\ref{fig6}) \cite{LEE}.
1132: 
1133: One of the most important concepts to be introduced by RVB theory is 
1134: the concept of spin-charge separation \cite{ANDERSON}.  This idea can 
1135: be most easily appreciated by the RVB explanation of the spin gap 
1136: phase \cite{PLEE}.  Since only the spins are paired, then strong effects are 
1137: expected for spin probes but only weak effects for charge 
1138: probes.  This is consistent with planar properties 
1139: of the pseudogap phase, which show a strong spin gap in NMR (left 
1140: panel, Fig.~\ref{fig7}), but only a weak gap-like depression
1141: in the in-plane infrared conductivity.  In fact, the drop in 
1142: in-plane resistance in the pseudogap phase
1143: is easily understood since when the spins pair up and become gapped,
1144: there are less states for the doped holes to scatter off of.  On the other hand,
1145: spin-charge separation, being a 2D effect, only occurs within a plane.  As
1146: spins and charges must thus recombine into real physical electrons to 
1147: tunnel from plane to plane, then large effects are expected in any 
1148: experiment measuring a c-axis current.  This is consistent with 
1149: experiment, since a hard gap is seen in c-axis infrared conductivity 
1150: (right panel, Fig.~\ref{fig7}), 
1151: as well as tunneling and photoemission.
1152: 
1153: Going beyond mean field considerations, though, has proven to be 
1154: difficult in the RVB scheme.  The most promising approach is to use 
1155: gauge theory to treat the various SU(2) and U(1) symmetries of the 
1156: model \cite{PLEE}.  The resulting gauge field fluctuations coupling the 
1157: ``spinons'' and 
1158: ``holons'', though, are extremely strong, leading to an uncontrolled 
1159: theory, as expected given the strong coupling nature of the problem.
1160: Still, these calculations have given a number of important insights 
1161: into understanding various excited state properties of the cuprates,
1162: particularly in the spin gap phase.
1163: 
1164: \subsection{Marginal Fermi Liquid}
1165: 
1166: The striking linear temperature dependence of the in-plane 
1167: resistivity led Varma and co-workers to propose a marginal Fermi 
1168: liquid phenomenology \cite{MFL} to explain many of the anomalous behaviors in 
1169: cuprates.  Their idea was that the electrons are interacting with a 
1170: spectrum of bosonic excitations which has the following form
1171: (left panel, Fig.~\ref{fig17})
1172: \begin{equation}
1173: B(\omega) \propto min(\omega/T,1)
1174: \end{equation}
1175: \begin{figure}
1176: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig17.eps}}}
1177: \caption{Bosonic spectrum which yields a marginal Fermi liquid (left 
1178: panel).  Resulting real and imaginary parts of the electron 
1179: self-energy (right panel).  $\omega_{c}$ is the ultraviolet cut-off.  T=0.
1180: Adapted from Ref.\cite{MFL}.}
1181: \label{fig17}
1182: \end{figure}
1183: That is, the bosonic spectrum has no other energy scale present 
1184: besides the temperature, that is, it exhibits quantum critical scaling.
1185: Such a spectrum, though, does not yield a 
1186: convergent fermion self-energy, necessitating the presence of an ultraviolet 
1187: cut-off, $\omega_c$, in the theory.  At zero temperature, the 
1188: resulting self-energy is (right panel, Fig.~\ref{fig17})
1189: \begin{eqnarray}
1190: Im\Sigma \propto \omega \nonumber \\
1191: Re\Sigma \propto \omega \ln \frac{\omega}{\omega_c}
1192: \end{eqnarray}
1193: This result, obtained by a convolution of $B$ and $ImG$ (where $G$ is 
1194: the bare fermion Greens function) can be most easily appreciated by 
1195: noting that for electrons interacting with an Einstein mode, 
1196: $B(\omega)=\delta(\omega-\omega_0)$, $Im\Sigma$ is a step function, 
1197: ($Im\Sigma=0, \omega < \omega_0$; $Im\Sigma \propto 1, \omega > 
1198: \omega_0$).  For an array of $\delta$ functions for $B$, $Im\Sigma$ 
1199: becomes a ramp of steps, which in the limit as the energy spacing of the 
1200: $\delta$ functions goes to zero becomes a linear $\omega$ 
1201: behavior.
1202: 
1203: The ``marginal'' notation comes from the fact that the quantity $1 - 
1204: dRe\Sigma/d\omega$ is logarithmically divergent as $\omega$ 
1205: approaches 0.  As a result, the momentum distribution function, $n(k)$, no 
1206: longer has a step discontinuity at $k_F$ as in a Fermi liquid, but 
1207: rather an inflection behavior.  This logarithm is cut-off by the 
1208: temperature, and a standard calculation of the longitudinal 
1209: conductivity leads to a linear temperature dependence of the 
1210: resistivity \cite{MFL}.
1211: 
1212: Although an attractive phenomenology for thinking about various 
1213: properties of the cuprates, the deficiency of the model is that there 
1214: is no explicit momentum dependence, leading to the question of
1215: where d-wave pairing would come from.  In later work, Varma has 
1216: claimed that d-wave pairing could arise from vertex corrections 
1217: \cite{VARMA}, but certainly the underlying microscopics behind this 
1218: very successful idea remain somewhat unclear at the present time.
1219: 
1220: These problems have led to a proposal of another phenomenology to 
1221: explain transport data, the ``cold spots'' model of Ioffe and Millis 
1222: \cite{COLD}.  In this picture, there is a Fermi liquid like 
1223: scattering rate, but it is confined to the vicinity of the d-wave 
1224: nodes.  Such a model can reproduce the linear $T$ resistivity, but this 
1225: nice idea now seems to be ruled out by recent photoemission data
1226: which find that even at the d-wave node, $Im\Sigma$ has 
1227: the linear $\omega$ behavior \cite{VALLA} predicted by the marginal 
1228: Fermi liquid phenomenology \cite{MFL}.
1229: 
1230: The MFL phenomenology can be easily extended to the superconducting 
1231: state.  Since the $B(\omega)$ spectrum is considered to be electronic 
1232: in origin, then it will acquire a $2\Delta$ gap in the 
1233: superconducting state (see bubble in the left panel, Fig.~\ref{fig5}), 
1234: thus being able to account for the scattering 
1235: rate gap seen in various measurements \cite{KURODA,LITTLEW}.  Such a
1236: ``gapped marginal Fermi liquid'', though, cannot account for all 
1237: observations, and has to be supplemented by collective effects 
1238: \cite{NDING}, which we now discuss.
1239: 
1240: \subsection{SO(5)}
1241: 
1242: Inelastic neutron scattering measurements by the group of
1243: Rossat-Mignod \cite{ROSSAT} revealed the presence of a
1244: narrow (in energy) resonance in
1245: the superconducting state of YBCO in a small region of momentum centered at 
1246: the $(\pi,\pi)$ wavevector \cite{ROSSAT}.  Subsequent polarized measurements by
1247: the Oak Ridge group \cite{MOOK93} verified that the resonance was 
1248: magnetic in character.  Since the BCS ground state involves S=0 pairs 
1249: with zero center-of-mass momentum, this implies that this 
1250: excited state must involve S=1 pairs with center-of-mass momentum 
1251: $Q=(\pi,\pi)$.  To see this, note that because of the energy gap, only pair 
1252: creation 
1253: processes are present in the particle-hole response at T=0, these processes 
1254: being possible because of particle-hole mixing \cite{SCALAPINO}.  
1255: Since the magnetic signal detected by neutrons involves a spin flip process, 
1256: then the excited pair must have spin one as the ground state is spin
1257: zero \cite{DEMLER}.  Fermion antisymmetry
1258: then implies that the excited pair has odd L.  This is evident as well, since
1259: the d-wave gap function $\cos(k_x)-\cos(k_y)$  translated by $Q/2$
1260: becomes $\sin(k_x)-\sin(k_y)$ \cite{DEMLER}.
1261: 
1262: At first sight, such a triplet collective mode is a surprise, since they have 
1263: never been found in classic superconductors.  But in the d-wave case, 
1264: since the order parameter changes sign under translation by $Q$, then 
1265: the BCS coherence factor for the pair creation process takes its maximal 
1266: value on the 
1267: Fermi surface, as opposed to the s-wave case where it is zero 
1268: \cite{FONG95}.
1269: \begin{figure}
1270: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig18.eps}}}
1271: \caption{Real and imaginary parts of the bare bubble for a d-wave 
1272: superconductor.  Intersection of real part with 1/J, where J is the 
1273: superexchange energy, marks the location of the pole in the RPA 
1274: response function (arrow).}
1275: \label{fig18}
1276: \end{figure}
1277: The net result (Fig.~\ref{fig18}) is that the imaginary part of the bare 
1278: particle-hole response, $\chi_0$, has a step function jump from zero to a finite 
1279: value at a threshold of $2\Delta_{hs}$ \cite{NORM00,AC}, where 
1280: $\Delta_{hs}$ is the value of the superconducting gap at the ``hot 
1281: spots'' (points on the Fermi surface connected by $Q$).  By 
1282: Kramers-Kronig, $Re\chi_0$ will then have a 
1283: logarithmic divergence at $2\Delta_{hs}$ because of the step in the 
1284: imaginary part.  Thus, the full response function 
1285: ($\chi=\chi_{0}/(1-J\chi_0)$, where $J$ is the superexchange energy) 
1286: will always have an undamped pole at some energy less than the 
1287: threshold energy.  That is, linear response theory (RPA) for a d-wave 
1288: superconductor predicts the presence of a spin triplet collective mode 
1289: below the $2\Delta$ continuum edge.
1290: 
1291: Experiments, though, reveal that the resonance 
1292: energy does not depend on temperature, and its amplitude scales 
1293: with the temperature dependence of the d-wave order parameter 
1294: \cite{FONG96}.  This is not easily understood in the ``RPA'' framework.
1295: This led Demler and Zhang to 
1296: propose that the spin resonance was in fact a particle-particle 
1297: antibound resonance (antibound, because the triplet interaction is
1298: repulsive).  In such a picture, the resonance is always present, but
1299: can only be detected below $T_c$ by neutron scattering because of 
1300: particle-hole mixing, which allows the particle-particle resonance to 
1301: appear in the particle-hole response \cite{DEMLER}.  This idea 
1302: naturally resolves the two puzzles mentioned above.
1303: Although this original idea has been 
1304: put into question on formal grounds by Greiter \cite{GREITER} (in the 
1305: t-J model, the triplet interaction is formally zero), and on 
1306: kinematics grounds by Tchernyshyov \etal \cite{OLEG} (an 
1307: antibound state is inconsistent with photoemission and tunneling, 
1308: which indicate that the resonance has an energy lower than the 
1309: two-particle continuum), the idea led 
1310: Zhang to propose a very interesting SO(5) phenomenology to explain the 
1311: cuprate phase diagram \cite{ZHANG}.
1312: 
1313: In the SO(5) picture, the ``5'' stands for the three degrees 
1314: of freedom of the Neel order ($N_x,N_y,N_z$), and the two degrees of 
1315: freedom of the superconducting order (real and imaginary parts of 
1316: $\Delta$).  In an imaginary world where these two order parameters 
1317: were degenerate, then the underlying Hamiltonian would have SO(5) 
1318: symmetry.  This group has ten generators, the three components of the 
1319: spin operator, the charge operator, and six new generators which 
1320: rotate the ``superspin'' between the Neel and superconducting 
1321: sectors.  These new generators are nothing more than the spin 
1322: resonance discussed above (a spin triplet pair with complex 
1323: $\Delta$).  This idea provides a new framework for thinking about 
1324: the phase diagram and the various collective excitations of cuprates
1325: \cite{ZHANG}.  Of course, cuprates are doped Mott insulators, and 
1326: this effect is not present in the theory as stands (that is, charge 
1327: fluctuations are suppressed strongly at low doping).  This has led to 
1328: the development of a version of the theory known as 
1329: ``projected SO(5)''  where double occupation has been projected
1330: out \cite{PROJ}.  One result of projected SO(5) 
1331: theory is the claim that it explains the ``d-wave-like'' dispersion of 
1332: the valence band seen in the undoped insulator by ARPES \cite{HANKE}.
1333: 
1334: \subsection{Stripes}
1335: 
1336: A number of models of correlated electron systems predict the 
1337: presence of phase separation at small doping, with the system 
1338: bifurcating into hole rich (metallic) and hole poor (magnetic 
1339: insulating) regions.  In some models, these regions form a lamellar 
1340: pattern, i.e., one dimensional ``stripes''.
1341: 
1342: To connect with experiment, it had been known for some time that 
1343: inelastic neutron scattering experiments for LSCO indicated the 
1344: presence of four incommensurate peaks displaced a distance $\delta$ 
1345: from the commensurate wavevector $Q=(\pi,\pi)$ which characterizes 
1346: the magnetic insulator (right panel, Fig.~\ref{fig10}) \cite{NS-LSCO}.  The standard way of thinking 
1347: about these peaks was that they were due to the Fermi surface geometry 
1348: \cite{LEVIN}.
1349: 
1350: An alternate way, though, was to consider having holes residing in 1D 
1351: stripes, with magnetic domains between these stripes (left panel, 
1352: Fig.~\ref{fig10}).  Even if the 
1353: local ordering within the magnetic domains is commensurate, if the 
1354: stripes represent an antiphase domain wall, then neutron scattering 
1355: will see incommensurate magnetic peaks, with $\delta$ a measure of the 
1356: spacing between the stripes.  Since the hole density is the doping, 
1357: then this predicts that $\delta$ will have a linear variation with 
1358: doping.  This is indeed what is seen in LSCO, and is known as the Yamada plot 
1359: \cite{YAMADA}.  Associated with these magnetic peaks should be charge 
1360: peaks at positions of $2\delta$ relative to the Bragg peaks.  These have 
1361: been observed as well (right panel, Fig.~\ref{fig10}) \cite{JOHN95}.
1362: 
1363: An attractive feature of the stripes model is its 1D physics 
1364: \cite{EMKIV}.  The jury is still out whether Fermi liquids are 
1365: inherently unstable in 2D \cite{ANDERSON}, but they definitely are in 
1366: 1D.  So, such 1D models naturally contain non Fermi liquid normal 
1367: states exhibiting spin-charge separation.  Moreover, in this picture, 
1368: the pseudogap is nothing more than the spin gap associated with the 
1369: magnetic domains.  Pairs of holes from the stripes 
1370: can obtain pairing correlations by virtually hopping into the magnetic 
1371: domains (the fluctuating Neel order in the magnetic domains 
1372: favors antiparallel spins on neighboring Cu sites).
1373: \begin{figure}
1374: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig19.eps}}}
1375: \caption{Stripes model for cuprates.  Pairs of doped holes (dark 
1376: circles) virtually 
1377: hop into AF (spin gap) domains, acquiring spin pairing correlations.  
1378: Josephson coupling of the stripes leads to long range superconducting 
1379: order.  Adapted from Ref.~\cite{EMKIV}.}
1380: \label{fig19}
1381: \end{figure}
1382: Below some 
1383: temperature, the stripes phase coherently lock via Josephson coupling, 
1384: leading to long range (3D) superconducting order (Fig.~\ref{fig19}).  That is, the system 
1385: crosses over from a 1D non-Fermi liquid normal state to a 3D coherent
1386: superconducting state \cite{EMKIV}.
1387: 
1388: \subsection{Pseudogap}
1389: 
1390: Most authors agree that the superconducting state is isomorphic to a 
1391: BCS ground state of d-wave pairs.  There is no agreement, however, on 
1392: the nature of the pseudogap phase.  The general hope is that once 
1393: the pseudogap phase is sorted out experimentally, then the number of 
1394: possible theories for cuprates will be drastically reduced.
1395: 
1396: One general class of theories is that the pseudogap phase represents 
1397: preformed pairs \cite{VARENNA}.  Cuprates are characterized by 
1398: short superconducting coherence lengths, low carrier densities, and 
1399: quasi-two dimensionality.  All of these conditions favor a 
1400: suppression of the transition temperature relative to its mean field 
1401: value due to phase fluctuations.  In the intermediate region between 
1402: these two temperatures, 
1403: preformed pairs are possible.  The cuprates, though, are not in the 
1404: Bose condensation (local pair) limit, in that photoemission still reveals the 
1405: presence of a large Fermi surface (in the Bose limit, the chemical 
1406: potential would actually lie beneath the bottom of the energy band).  
1407: Still, specific heat data clearly reveal the non-mean-field like 
1408: character of the superconducting phase transition, particulary for 
1409: underdoped samples \cite{JUNOD}.
1410: 
1411: The ``RVB'' picture has subtle differences from that of pre-formed 
1412: pairs.  In this case, the pseudogap phase is a spin gap phase (that 
1413: is, the spins bind into singlets).  As an electron is a product of 
1414: spin and charge, then real electrons acquire an energy gap because of 
1415: the spin gap, though charge excitations confined to the plane do not.  
1416: At low enough temperatures, the doped holes become phase coherent, 
1417: leading to rebinding of spin and charge, and the formation of a true 
1418: superconducting ground state.
1419: The stripes picture is not unrelated to the RVB picture, in that the 
1420: pseudogap is due to the spin gap present in the magnetic insulating 
1421: domains between the stripes.
1422: 
1423: In the other class of scenarios, the pseudogap is not related to 
1424: superconductivity per se, but rather is competitive with it.  Most of 
1425: these scenarios involve either a charge density wave or spin density 
1426: wave, usually without long range order.  As the spectral gap 
1427: associated with the ordering grows with reduced doping, then more and 
1428: more of the Fermi surface becomes unavailable for pairing, leading to 
1429: an increasing suppression of the superconducting transition 
1430: temperature on the underdoped side.  This is an old idea, going back to 
1431: the A15 superconductors where the martensitic phase transition 
1432: competes with superconductivity \cite{BILBRO}.  These models 
1433: have a certain attractiveness, since the basic physics can 
1434: be appreciated at the mean field level.
1435: 
1436: The most interesting example of the competitive scenario is that 
1437: of orbital currents.
1438: These involve bond currents either circulating around an elementary 
1439: plaquette of four coppers (middle panel, Fig.~\ref{fig16})
1440: \cite{LEESF,SCHULZ,PLEE,ORB,NAYAK} or within a 
1441: subplaquette involving just the Cu-O bonds (Fig.~\ref{fig20}) \cite{VARMA,VARMA02}.
1442: \begin{figure}
1443: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig20.eps}}}
1444: \caption{Orbital currents states proposed by Varma 
1445: (Ref.~\cite{VARMA02}).  Solid dots are Cu ions, open dots O 
1446: ions, and arrows are bond currents.  The right panel has a form factor 
1447: consistent with recent ARPES results \cite{NAT02}.}
1448: \label{fig20}
1449: \end{figure}
1450: So far, only two pieces of experimental evidence point to 
1451: such a state.  Inelastic neutron scattering experiments find small 
1452: moment magnetism in underdoped YBCO whose form factor drops off 
1453: more rapidly in momentum space than that associated with Cu spins 
1454: \cite{MOOK}, indicating the presence of a moment extended in real
1455: space.  And recent circularly polarized ARPES experiments reveal 
1456: the presence of time reversal symmetry breaking below $T^*$ 
1457: \cite{NAT02}, whose form factor in momentum space (if interpreted in 
1458: terms of orbital currents) favors the Varma picture \cite{VARMA02}.
1459: 
1460: One of the most interesting aspects of the competitive scenarios is 
1461: the prediction of a quantum critical point where the pseudogap effect 
1462: disappears.  From experiment, it has been claimed that the $T^*$ line 
1463: passes through the $T_c$ line and vanishes within the superconducting 
1464: dome at a concentration of 19\%, just beyond optimal doping (Fig.~\ref{fig9})
1465: \cite{LORAM}.
1466: 
1467: In fact, there are several possible quantum critical points in the 
1468: cuprate phase diagram (Fig.~\ref{fig21}).
1469: \begin{figure}
1470: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig21.eps}}}
1471: \caption{Four possible quantum critical points (dark circles)
1472: in the cuprate phase 
1473: diagram.  Dotted line is the pseudogap phase line.}
1474: \label{fig21}
1475: \end{figure}
1476: Starting from the undoped material, as the 
1477: doping progresses, one first finds the point where the Neel 
1478: temperature vanishes, then the point where superconductivity first 
1479: occurs, then the critical point mentioned above, and finally at higher 
1480: doping the point where superconductivity disappears, for a total of 
1481: four possible quantum critical points.  The last point may correspond 
1482: to where the Fermi surface topology changes from hole-like to 
1483: electron-like (that is, the saddle point in the dispersion at $(\pi,0)$ passes 
1484: through the Fermi energy), as there is some evidence of this from ARPES.
1485: 
1486: It remains to be seen whether the ``quantum critical paradigm'' with
1487: its emphasis on competing phases is the proper way of thinking about 
1488: cuprates \cite{LAUGH}.  It has certainly led to an enrichment in our 
1489: understanding of these novel materials \cite{SUBIR}.
1490: 
1491: \section{Photoemission}
1492: 
1493: \subsection{General Principles}
1494: 
1495: As emphasized by Anderson, angle resolved 
1496: photoemission has emerged as one of the most important spectroscopic 
1497: probes of cuprate superconductors, in some sense playing the role 
1498: that tunneling spectroscopy played in conventional superconductors 
1499: \cite{ANDERSON}.  Of course, much has 
1500: been discovered about cuprates using tunneling, but Anderson's statement 
1501: was meant to emphasize the fact that photoemission played no role in 
1502: the past, and then all of the sudden stepped up to play a major role 
1503: in the cuprate problem.  For those of us working in the ``old days'', 
1504: these developments have been nothing short of amazing.
1505: 
1506: To understand how this came about, a few comments are in order about 
1507: photoemission.  This technique has a venerable history; in fact, it 
1508: was for explaining the photoelectric effect, discovered by 
1509: Hertz in 1887, that Einstein got his Nobel 
1510: prize.  Although the concept for doing angle resolved experiments had 
1511: been recognized, the general perception was that not much 
1512: useful would be learned.  This changed in the early 1960s when Spicer 
1513: developed the three-step model for photoemission, showing that in 
1514: principle, important information about the electronic structure could 
1515: be elucidated.  Subsequent experiments by a number of groups, including
1516: Dean Eastman's, were able 
1517: to determine the electronic dispersion of transition metals using 
1518: this technique \cite{HUFNER}.  For these developments, Spicer and 
1519: Eastman received the Buckley Prize in 1980.
1520: 
1521: The technique involves shining photons on a sample with a specific 
1522: energy.  If the photons have an energy larger than the work function 
1523: of the metal, then electrons will be emitted.  In angle resolved 
1524: mode, an electrostatic detector measures the azimuthal and polar 
1525: angles of the electrons (relative to the surface normal), as well as 
1526: their energy.  Knowing the energy of the photon, then the initial 
1527: energy of the electrons in the crystal can be determined, as well as 
1528: the components of the momentum parallel to the sample surface.  In 
1529: principle, the perpendicular component of the momentum is also 
1530: determined from the energy-momentum relation (the electrons in vacuum 
1531: having an energy which is quadratic in momentum), but there are
1532: subtleties connected with the breaking of the crystal symmetry by 
1533: the surface in this direction.
1534: 
1535: The actual photocurrent is very complicated, since it is formally a 
1536: three current correlation function (this is the so-called one-step 
1537: model for photoemission).  But in the three-step approximation of 
1538: Spicer, where the initial electron is photoexcited, this 
1539: photoelectron transports through the crystal, then out in the 
1540: vacuum to the detector, the photocurrent (for one band) can be written as
1541: \begin{equation}
1542: I({\bf k},\omega) = c_{\bf k} \int_{\delta {\bf k}} d{\bf k'} \int d\omega'
1543: A({\bf k'},\omega') f(\omega') R(\omega,\omega')
1544: \end{equation}
1545: where $c_{\bf k}$ is the modulus squared of the matrix element of the 
1546: operator ${\bf A}\cdot{\bf p}$ between initial and final states 
1547: (${\bf A}$ is the vector potential, ${\bf p}$ the momentum operator), $A$
1548: the single particle spectral function ($-ImG/\pi$, where $G$ is the 
1549: electron Greens function), $f$ the Fermi-Dirac function, and $R$ the 
1550: energy resolution function (a guassian).  The momentum integration is 
1551: a window ($\delta {\bf k}$)
1552: centered about ${\bf k}$ which represents the finite 
1553: momentum resolution of the spectrometer.  This expression assumes the 
1554: impulse (or sudden) approximation, where the interaction of the 
1555: photoelectron with the photohole is ignored.  Moreover, the 
1556: expression implicitly assumes the 2D limit, which fortunately is 
1557: relevant for the cuprate case, where $k_z$ dispersion effects are 
1558: weak (``bilayer splitting'' is a different matter).
1559: 
1560: The significance of this expression is obvious.  The single particle 
1561: spectral function is the simplest quantity which emerges from a 
1562: many-body theory of electrons.  We note that $G^{-1}({\bf 
1563: k},\omega) = \omega -\epsilon_{\bf k} -\Sigma({\bf k},\omega)$
1564: where $\epsilon$ is the bare energy and $\Sigma$ the Dyson 
1565: self-energy.  A simple example of this is 
1566: BCS theory, where $\Sigma_{BCS}({\bf k},\omega) = \Delta_{\bf 
1567: k}^2/(\omega+\epsilon_{\bf k})$.  Since the momentum distribution 
1568: function (many-body occupation factor) is given by $n_{\bf k} = \int 
1569: d\omega A({\bf k},\omega) f(\omega)$, then modulo resolution and dipole 
1570: matrix elements, the frequency integral of the ARPES spectrum is 
1571: $n_{\bf k}$ \cite{MOHIT95}.
1572: 
1573: One problem with photoemission is that it is a surface sensitive 
1574: probe, particularly for the low energy ($\sim$ 20 eV) photons typically used 
1575: to achieve high energy and momentum resolution.  Even in the quasi-2D 
1576: cuprates, this can lead to problems unless a natural cleavage plane 
1577: exists.  This is why most measurements have been done on BSCCO, which 
1578: contains a double BiO spacer layer, with the two BiO layers having 
1579: the biggest interplanar separation in the cuprates (these layers are 
1580: at a separation typical of van der Waals interactions).  This 
1581: provides the best possible cleavage in the cuprates.  The penalty one 
1582: pays is that the BiO layers have planar bonds which are longer than 
1583: the CuO bonds.  The material tries to compensate for this by 
1584: developing a superstructure deformation of the BiO planes.  This can 
1585: lead to ``ghost'' images of the main CuO signal due to diffraction of 
1586: the photoelectrons off the BiO surface layer, which have to be taken 
1587: into account when interpreting ARPES spectra \cite{REVIEW}.
1588: 
1589: The advantage of ARPES, though, is now obvious.  It is both a 
1590: momentum and frequency resolved probe.  The only other probe 
1591: comparable to this is inelastic neutron scattering, which measures a 
1592: more complicated function (the spin part of the particle-hole response).
1593: Such a 
1594: momentum resolved probe was not essential in classic s-wave
1595: superconductors where momentum dependent effects are not important,
1596: but we know they are essential to consider in the d-wave case.  This 
1597: is an obvious advantange of ARPES over tunneling, though the latter 
1598: still has much better energy resolution, and has the advantage of 
1599: being able to see unoccupied states as well.  On the other hand, 
1600: important spatial information can be obtained by STM, which measures
1601: the local density of states, which is beyond 
1602: the scope of this review.
1603: 
1604: The true impact of ARPES was recently realized by the development of 
1605: the Scienta detector, which allows collection of data simulataneously 
1606: as a function of momentum and energy.  In angle integrated mode, 
1607: energy resolutions of order 2 meV become possible, allowing even 
1608: conventional superconductors to be studied by photoemission 
1609: \cite{YOKOYA}.  But even though the energy resolution is not as good 
1610: in angle resolved mode (typically 10-20 meV), it is adequate for the
1611: cuprates, and the momentum resolution is 
1612: quite good, of order 0.005 of a reciprocal lattice vector.  This 
1613: precision in momentum allows fine details of the spectral function to 
1614: now be resolved.  It is certainly a far cry from the pre-cuprate era, 
1615: when momentum resolutions were typically 0.1 of a reciprocal lattice 
1616: vector, and energy resolutions were typically 100 meV.
1617: 
1618: The power of the Scienta detectors can be appreciated by looking at 
1619: the expression for the spectral function
1620: \begin{equation}
1621: A({\bf k},\omega) = \frac{1}{\pi}\frac{Im\Sigma({\bf k},\omega)}
1622: {(\omega-\epsilon_{\bf k} -Re\Sigma({\bf k},\omega))^2+
1623: (Im\Sigma({\bf k},\omega))^2}
1624: \end{equation}
1625: In the past, spectra were typically analyzed at fixed ${\bf k}$ as a 
1626: function of energy (EDC).  This energy lineshape is obviously complicated 
1627: given the 
1628: non-trivial $\omega$ dependence of $\Sigma$.  But at fixed $\omega$
1629: as a function of ${\bf k}$
1630: instead, the momentum lineshape (so-called MDC \cite{VALLA}) is
1631: considerably simpler.  In the normal state near the Fermi energy, we 
1632: can typically linearize $\epsilon_{\bf k}$ in momenta normal to the
1633: Fermi surface.  As long as 
1634: $Re\Sigma$ can also be linearized, then the MDC 
1635: reduces to a Lorentzian, with half width $\Sigma(\omega)/v_{F0}$, 
1636: where $v_{F0}$ is (modulo $dRe\Sigma(k)/dk$) the bare Fermi velocity 
1637: (obtained from 
1638: $\epsilon_{\bf k}$) \cite{ADAM2}.  Given an estimate for the bare 
1639: velocity, then $Re\Sigma$ can be read off from the MDC dispersion, 
1640: and $Im\Sigma$ from the MDC width, though it should be noted that the 
1641: latter is only true if the resolution is accounted for in the analysis.
1642: 
1643: \subsection{Normal State}
1644: 
1645: The undoped version of the cuprates is a magnetic insulator.  Simple 
1646: considerations lead one to expect that the valence band maximum is at 
1647: the $(\pi/2,\pi/2)$ points of the zone (left panel, Fig.~\ref{fig22}).
1648: \begin{figure}
1649: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig22.eps}}}
1650: \caption{Left panel is the dispersion in the non magnetic phase (thick
1651: line) and magnetic phase (thin lines), the latter assuming an SDW gap 
1652: of 200 meV.  Resulting Fermi surfaces 
1653: are shown in the right panel, with the dashed line the magnetic zone
1654: boundary.}
1655: \label{fig22}
1656: \end{figure}
1657: This has been beautifully 
1658: confirmed by photoemission measurements \cite{WELLS}.  When hole doping 
1659: such a state, then one might expect to form small hole pockets with 
1660: volume $x$ centered at these points (right panel, Fig.~\ref{fig22}).  This would certainly be
1661: consistent with transport and optics data, which indicate a carrier 
1662: density of $x$.  Somewhat surprisingly, though, this is not what is 
1663: seen by ARPES \cite{JC90}.  Rather, what is seen is a large hole 
1664: surface (right panel, Fig.~\ref{fig22}) with volume $1+x$ centered about the $(\pi,\pi)$ points
1665: (remembering the full Brillouin zone corresponds to 2 filled states
1666: because of spin degeneracy).  This 
1667: is more or less what is predicted by paramagnetic band theory.  To be 
1668: consistent with transport, this would mean that the spectral weight 
1669: would have to scale with $x$.  This was subsequently found to be the 
1670: case from ARPES, first at 
1671: the $(\pi,0)$ points in Bi2212
1672: \cite{JC99,FENG00,DING01}, then most recently along the nodal direction
1673: in LSCO \cite{YOSHIDA}.  In some sense, the Fermi surface disappears 
1674: by losing its spectral weight, much like the Cheshire Cat in Alice in 
1675: Wonderland.
1676: 
1677: The energy dispersion seen in the doped case involves the presence of 
1678: a saddle point at $(\pi,0)$ on the occupied side which is relatively 
1679: close to the Fermi energy (left panel, Fig.~\ref{fig22}).  
1680: This has led to many theories based on van 
1681: Hove singularities in the density of states.  The dispersion near the 
1682: saddle point is quite flat, especially in the superconducting state, 
1683: which has led to the ``extended'' van Hove singularity concept 
1684: proposed by Abrikosov \cite{ALEX}.  As the hole doping 
1685: increases, the saddle point approaches the Fermi energy.  In the case 
1686: of Bi2201, which can be heavily overdoped, the saddle point appears 
1687: to be almost degenerate with the Fermi energy at a concentration 
1688: where $T_{c}$ is essentially zero \cite{SATOB}.  Beyond this 
1689: concentration, the saddle point is expected to pass through the Fermi 
1690: energy, leading to an electron surface centered at $(0,0)$.  There is
1691: evidence from ARPES that this occurs in LSCO \cite{INO}.  This 
1692: would be consistent with Hall measurements, which see a sign change 
1693: in the Hall number near this concentration \cite{HALL}.
1694: 
1695: This dispersion, though, should be taken with a very large grain of 
1696: salt.  In particular, no well defined spectral peaks appear in the 
1697: normal state, at least for optimal and underdoped samples.  That is, 
1698: the widths of the peaks are of order their energy separation from the 
1699: Fermi energy.  This is why no van Hove singularity in the 
1700: density of states has been inferred from any experimental measurement.
1701: 
1702: The momentum and energy dependence of the spectral peaks in the 
1703: normal state was of great interest from the beginning.  The original 
1704: ARPES analysis of Olson \etal \cite{OLSON} found the peak width to scale 
1705: linearly with peak energy.  This is the behavior predicted by the 
1706: marginal Fermi liquid phenomenology \cite{MFL}.
1707: \begin{figure}
1708: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig23.eps}}}
1709: \caption{Imaginary part of the self-energy determined by ARPES along 
1710: the nodal direction in Bi2212, demonstrating marginal behavior (quantum
1711: critical scaling) as a function of 
1712: both $\omega$ and $T$.  From Ref.~\cite{VALLA}.}
1713: \label{fig23}
1714: \end{figure}
1715: This behavior has 
1716: been confirmed to much better precision with high momentum resolution 
1717: data along the nodal direction (Fig.~\ref{fig23}) \cite{VALLA}, which also indicate a 
1718: linear $T$ dependence of the linewidth as well.
1719: 
1720: The momentum anisotropy of the linewidth, though, is still a matter of 
1721: controversy.  Data for optimal doping indicate that the linewidths get 
1722: broader as the $(\pi,0)$ point is approached.  This is as opposed to 
1723: heavily overdoped Bi2201, where the lineshape is relatively isotropic 
1724: around the Fermi surface.  Such momentum anisotropies are not 
1725: unexpected, given that the d-wave nature of the superconducting order 
1726: parameter implies momentum dependent interactions.
1727: On the other hand, some authors have suggested that the 
1728: intrinsic lineshape might be fairly isotropic, with the observed 
1729: anisotropy in Bi2212 a combination of overlap of features due to the ``ghost'' 
1730: images associated with the superstructure, along with bilayer 
1731: splitting of the energy bands \cite{BOGDANOV}.
1732: 
1733: The issue of bilayer splitting has been somewhat controversial, and 
1734: thus deserves some attention.  Most experiments have been done in 
1735: Bi2212, which has two CuO layers in each formula unit, the two layers
1736: being separated by Ca 
1737: ions.  Mixing of the levels on the two planes will lead to an
1738: antibonding and bonding combination, thus bilayer splitting (Fig.~\ref{fig24}).  Band 
1739: theory predicts that the splitting should be sizable (of order 1/4 
1740: eV), with the splitting varying with planar momentum like 
1741: $(\cos k_{x}-\cos k_{y})^{2}$.  Therefore, the effect is largest at 
1742: $(\pi,0)$ \cite{INTLAY}.  As spectral peaks are broad in the normal state, 
1743: then it is quite possible that the two overlapping features would smear 
1744: together into a single feature, making the $(\pi,0)$ 
1745: lineshape look anomalously broad.
1746: 
1747: \begin{figure}
1748: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig24.eps}}}
1749: \caption{ARPES spectrum at $(\pi,0)$ for optimal doped and 
1750: overdoped Bi2212 samples in the normal state.  Separate peaks in the 
1751: overdoped case are due to bilayer splitting.  Data from Ref.~\cite{JCNEW}.}
1752: \label{fig24}
1753: \end{figure}
1754: 
1755: Most early reports of bilayer 
1756: splitting were later attributed to the ``ghost'' images associated 
1757: with the superstructure.  The claim was that once this was factored 
1758: out, no bilayer splitting was apparent in the data, at least for an
1759: optimal doped sample \cite{DING96}.  
1760: This was of interest, since several theories of cupratres predicted 
1761: incoherent behavior along the c-axis in the normal state 
1762: \cite{ANDERSON}.
1763: 
1764: Subsequent measurements at high momentum resolution, though, revealed 
1765: the presence of bilayer splitting in heavily overdoped samples 
1766: \cite{BILAYER}.  The $(\pi,0)$ spectrum is characterized by a 
1767: relatively sharp antibonding peak near the Fermi energy (consistent 
1768: with the more Fermi liquid like character of heavily overdoped 
1769: samples), plus a broader (bonding) peak at higher binding energy  
1770: (its width being larger due to its greater binding energy).
1771: This leads to a peak-shoulder type spectrum, as opposed to the single 
1772: broad spectrum seen for optimal and underdoped samples (Fig.~\ref{fig24}).  The doping
1773: dependence of the bilayer splitting, though, is still a controversial issue.
1774: Recently,
1775: Campuzano's group has presented evidence that bilayer effects
1776: disappear in the ``strange metal" (quantum critical) phase of 
1777: Fig.~\ref{fig6} \cite{JCNEW}.
1778: 
1779: \subsection{Superconducting State}
1780: 
1781: There are two remarkable features revealed by ARPES in the superconducting 
1782: state.  First, the opening of an anisotropic superconducting gap in the 
1783: spectrum.  Second, the appearance of a sharp coherent peak below $T_{c}$.
1784: 
1785: The first observation of a gap was made in 1989 by Yves 
1786: Baer's group \cite{BAER}, and was considered a tour-de-force at the 
1787: time.  Later, angle resolved, measurements did not detect any gap 
1788: anisotropy, probably due to the sample qualities at the time.  This 
1789: all changed in 1993 with the observation by Z-X Shen's group of an 
1790: anisotropic gap consistent with d-wave symmetry \cite{SHEN93}.
1791: 
1792: \begin{figure}
1793: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig25.eps}}}
1794: \caption{Doping dependence of the spectral gap in Bi2212 from ARPES.  
1795: Left panel for an overdoped sample, right panel for an underdoped 
1796: one.  B=1 is a $\cos(2\phi)$ dependence of the gap on Fermi surface 
1797: angle, solid line in the right panel includes a contribution from the 
1798: next d-wave 
1799: harmonic, $\cos(6\phi)$.  From Ref.~\cite{MESOT99}.}
1800: \label{fig25}
1801: \end{figure}
1802: 
1803: Not all groups found this behavior, though.  It was realized 
1804: later that the discrepancies were due to complications caused 
1805: by the ``ghost'' images associated with the Bi2212 superstructure.  Once 
1806: this was appreciated, then it was evident that the best results could be 
1807: obtained in the $Y$ quadrant of the Brillouin zone where these images 
1808: were well separated from the main image.  By fitting the leading edge 
1809: of the spectrum (taking into account the known resolution), precise 
1810: values of the energy gap can be obtained.  The resultant plot of the
1811: gap along the Fermi surface shows a 
1812: clear V-shaped behavior around the node, as expected for d-wave 
1813: symmetry (left panel, Fig.~\ref{fig25}) \cite{GAP96}.  Moreover, the functional dependence on 
1814: momentum is precisely of the form $|\cos k_{x} - \cos k_{y}|$ 
1815: \cite{SHEN93,GAP96} as would be expected for pairs of electrons 
1816: sitting on near neighbor Cu sites.  In fact, along the observed Fermi
1817: surface, the functional dependence is essentially $\cos(2\phi)$, where
1818: $\phi$ is the angle of the line connecting $(\pi,\pi)$ to the 
1819: Fermi surface with the line $(\pi,0)-(\pi,\pi)$.  Subsequent measurements have 
1820: revealed a deviation from this form with increasing underdoping (right panel,
1821: Fig.~\ref{fig25}) \cite{MESOT99}, which can be fit by inclusion of the next 
1822: harmonic in the gap expansion, $\cos(6\phi)$, which is related to
1823: $\cos 2k_x - \cos 2k_y$.  This indicates that the 
1824: pair interaction is becoming longer range in real space as the doping is 
1825: reduced.  The trend is expected, given that the correlation length 
1826: associated with magnetic fluctuations increases with underdoping.  
1827: Recently, the same deviation from the $\cos(2\phi)$ form has been 
1828: inferred from Fourier transformation of STM data \cite{MCELROY}.
1829: 
1830: One of the most interesting aspects of the superconducting state is 
1831: that the low energy states have a Dirac-like dispersion, i.e., Dirac cones,
1832: whose constant energy contours are centered 
1833: about the nodes (actually, these contours are banana shaped due to the
1834: Fermi surface curvature, Fig.~\ref{fig26}).
1835: \begin{figure}
1836: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig26.eps}}}
1837: \caption{Left: Fermi surface (dashed line) and a constant energy contour for
1838: quasiparticle excitations in the d-wave superconducting state (solid 
1839: line) based on ARPES data for Bi2212.  Right: Modified version proposed to
1840: explain the incommensurability seen in neutron scattering data in YBCO.
1841: From Ref.~\cite{NORM00}.}
1842: \label{fig26}
1843: \end{figure}
1844: According to ARPES, these 
1845: cones are quite anisotropic, with the ratio of the velocity normal to 
1846: the Fermi surface to that along the Fermi surface (the latter 
1847: the slope of the gap around the node) of 20 
1848: \cite{MESOT99}.  This value has also been inferred from thermal 
1849: conductivity measurements \cite{CHIAO}.  In principle, by comparison 
1850: of this ratio to the value of the linear T coefficient of the
1851: penetration depth, important information can be obtained about the 
1852: electromagnetic coupling of the quasiparticles.  Present results are 
1853: consistent with a linear doping variation of the particular Landau 
1854: interaction parameter involved \cite{MESOT99}, but the associated 
1855: error bars are quite large.  More precise ARPES and 
1856: penetration depth measurements on Bi2212 would be gratifying in this 
1857: regard.  In particular, the doping variation of the gap slope around 
1858: the node has not been studied yet with high resolution detectors.  It 
1859: is of some interest to see whether this quantity scales with $T_c$ on 
1860: the underdoped side of the phase diagram.  Recent thermal conductivity
1861: data indicate that this is not the case \cite{SUTHER}.
1862: 
1863: What is known, though, is that the maximum superconducting energy gap 
1864: does not scale with $T_c$ on the underdoped side (Fig.~\ref{fig27}) \cite{HARRIS,JC99}.
1865: \begin{figure}
1866: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig27.eps}}}
1867: \caption{Spectral peak energy (maximum superconducting energy gap), 
1868: energy of the hump, and pseudogap temperature, $T^{*}$, versus doping, 
1869: x, from ARPES data on Bi2212.  From Ref.~\cite{JC99}.}
1870: \label{fig27}
1871: \end{figure}
1872: Instead, this quantity monotonically increases with underdoping, 
1873: scaling with the pseudogap temperature, $T^{*}$.  The same trend has 
1874: been seen by tunneling \cite{JOHNZ}.  The observed behavior would 
1875: be consistent with $T^{*}$ representing some mean field transition 
1876: temperature for pairing.  This doping trend was actually predicted 
1877: many years ago by RVB theory \cite{GROS}, where the spin pairing 
1878: energy scale is decoupled from the phase stiffness energy
1879: associated with the doped holes, since the latter is proportional to $x$
1880: (see Fig.~\ref{fig16}).
1881: 
1882: The lineshape changes between normal and superconducting states, though, 
1883: are perhaps the most fascinating aspect of the ARPES data (Fig.~\ref{fig28}).  The 
1884: changes are most spectacular near the $(\pi,0)$ point for optimal and 
1885: underdoped samples.
1886: \begin{figure}
1887: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig28.eps}}}
1888: \caption{ARPES spectra at $(\pi,0)$ for an overdoped (87K) Bi2212 sample in 
1889: the normal state (NS) and superconducting state (SC).  Data from 
1890: Ref.~\cite{NORM97}.}
1891: \label{fig28}
1892: \end{figure}
1893: As the temperature is lowered, the broad peak in 
1894: the normal state develops into a sharp coherent peak separated by a 
1895: spectral dip (near $(\pi,0)$) or a spectral break (near the 
1896: $(\pi,0)-(\pi\pi)$ Fermi surface crossing) from the higher energy 
1897: incoherent part \cite{MOHIT95}.  This behavior is consistent with a 
1898: strong increase in the lifetime of the electrons
1899: as the temperature is lowered below $T_{c}$, as has 
1900: been earlier inferred from microwave conductivity \cite{BONN} and 
1901: thermal conductivity \cite{ONG} experiments.  That is, a gap is being 
1902: opened in the scattering rate, as also derived from infrared 
1903: conductivity measurements \cite{PUCH}.  In ARPES, this can be seen 
1904: very clearly by ``inverting'' the data to directly extract the 
1905: temperature dependence of the electron self-energy \cite{TEMP01}.
1906: 
1907: An alternate interpretation has been given to the data, however 
1908: \cite{FEDEROV,FENG00,DING01}.  In this picture, a gap 
1909: develops in the incoherent part of the spectrum, with a quasiparticle 
1910: pole appearing inside the gap.  The pole weight monotonically 
1911: increases with decreasing temperature, and it has been suggested that 
1912: this behavior tracks the superfluid density \cite{FENG00}.  In some 
1913: sense, this would imply that the quasiparticle weight was equal to 
1914: the superconducting order parameter.  One particular model which 
1915: is suggestive of this is the Josephson coupling of stripes below 
1916: $T_{c}$ \cite{EMKIV}.
1917: 
1918: The remarkable spectral changes near $(\pi,0)$ leading to the unusual 
1919: peak-dip-hump lineshape below $T_{c}$ were actually first 
1920: observed 
1921: by tunneling \cite{HUANG}.  When they were subsequently observed by
1922: ARPES, the obvious explanation was 
1923: that they were due to bilayer splitting (the ``hump'' representing the 
1924: bonding band, the ``peak'' the antibonding band).  There are a number 
1925: of arguments against this (including the fact that tunneling 
1926: spectroscopy sees this lineshape for single layer materials like 
1927: Tl2201).  What is clear, though, is that bilayer splitting alone
1928: is not sufficient to explain the lineshape.  In particular, 
1929: the spectral dip represents a depletion of states which fills in as 
1930: the temperature is raised \cite{TEMP01}.  Moreover, the dip energy 
1931: scale appears to exist at the same energy throughout the Brillouin 
1932: zone \cite{ADAM2}.
1933: 
1934: These considerations have led to many speculations that the spectral dip 
1935: represents some sort of many body effect.  One of the first 
1936: treatments of this problem was by Arnold \etal \cite{ARNOLD}, where 
1937: they applied the McMillan-Rowell ``inversion" procedure \cite{MCMR} to
1938: the data to determine the boson spectral function from the frequency 
1939: dependence of the gap function, $\Delta(\omega)$.  From this analysis, a 
1940: sharp bosonic mode was inferred at about 10 meV.
1941: The problem with this pioneering analysis was that it assumed the data 
1942: represent an isotropic 
1943: density of states proportional to $\frac{\omega}{\sqrt{\omega^2-\Delta^2(\omega)}}$,
1944: with the spectral dip corresponding to a strong frequency variation
1945: of $\Delta(\omega)$.
1946: In this case, the ``normal'' part of 
1947: the self-energy (diagonal in particle-hole space)
1948: drops out, and so all structure in the data can 
1949: be associated with the pairing self-energy (off-diagonal part).
1950: This is not the case if 
1951: the data represent a spectral function.
1952: 
1953: The data were later analyzed assuming the primary effects 
1954: were due to the normal self-energy (the pairing part being 
1955: treated in a BCS approximation) \cite{NDING}.  In this 
1956: analysis, the spectral dip can be understood as a sharp threshold for 
1957: inelastic scattering.  To understand this, consider the Feynman diagram 
1958: for an electron scattering off particle-hole excitations (left panel, 
1959: Fig.~\ref{fig5}).
1960: \begin{figure}
1961: \centerline{\epsfxsize=1.0\textwidth{\epsfbox{fig29.eps}}}
1962: \caption{Left: boson spectrum (``$\alpha^2F$") in the normal state (dashed line, 
1963: corresponding to a marginal Fermi liquid) and superconducting state (solid line, 
1964: corresponding to a gapped marginal Fermi liquid).  The arrow marks the 
1965: possibility of a collective mode with energy $\Omega_0$ inside the 
1966: continuum gap of $2\Delta$ (corresponding to 
1967: a gapped marginal Fermi liquid plus a mode).  Middle: 
1968: Resulting imaginary self-energy for the electrons.
1969: Right:  Superconducting state spectral function from this model
1970: as compared to ARPES data at $(\pi,0)$ for slightly overdoped Bi2212.
1971: The mode energy, $\Omega_0$, equals the spin resonance energy
1972: determined independently from neutron scattering.
1973: Adapted from Ref.~\cite{NDING}.}
1974: \label{fig29}
1975: \end{figure}
1976: In the superconducting state, the particle-hole continuum will have a 
1977: gap of order $2\Delta$ (left panel, Fig.~\ref{fig29}).  If interactions are strong enough that a
1978: bound state (energy $\Omega_{0}$) emerges inside of this continuum gap, 
1979: then $Im\Sigma$ will develop a sharp threshold, as implied by the 
1980: data, at an energy $\Delta+\Omega_{0}$ \cite{NORM97} (middle panel, Fig.~\ref{fig29}).  
1981: In this picture, the energy of the bosonic mode 
1982: will be equal to the energy difference of the dip ($\Delta + 
1983: \Omega_{0}$) and the peak ($\Delta$) \cite{NDING,AC2} (right panel, Fig.~\ref{fig29}).  
1984: Moreover, the resulting spectral function 
1985: will consist of two features:  a broad feature at higher binding
1986: energy whose dispersion roughly tracks the dispersion of the 
1987: single feature in the normal state (hump), and a sharp, weakly dispersive feature 
1988: for smaller
1989: binding energies representing the renormalized quasiparticle branch (peak).  
1990: These two features are separated by the dip energy, which is 
1991: roughly constant in momentum.  This model gives a natural explanation of the 
1992: unusual dispersions associated with the peak and hump \cite{NORM97}.
1993: 
1994: Moreover, the peak-dip-hump is strongest at the $(\pi,0)$ points.  As 
1995: these points are connected by $(\pi,\pi)$ wavevectors, this would 
1996: imply that the bosonic excitations involved are associated with this 
1997: wavevector \cite{SS}.  This, coupled with the inferred mode energy (40 
1998: meV), points to the spin resonance as the boson \cite{NORM97}.
1999: At the time of this conjecture, the resonance had only been seen in 
2000: YBCO, but later experiments found it in Bi2212 at the energy inferred 
2001: from ARPES \cite{KEIMER}.  Despite this, a criticism offered 
2002: against such an interpretation was that all energy scales from ARPES 
2003: and tunneling appear to increase with underdoping 
2004: (Fig.~\ref{fig27}), but the resonance energy decreases 
2005: \cite{TIMRPP}.  This was answered later by a doping dependent ARPES 
2006: study, which found that
2007: the mode energy inferred from the data had a doping 
2008: dependence which indeed tracks
2009: the resonance energy (left panel, Fig.~\ref{fig30}) \cite{JC99}.
2010: \begin{figure}
2011: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig30a.eps}}
2012: \epsfxsize=0.5\textwidth{\epsfbox{fig30b.eps}}}
2013: \caption{Left panel:  mode energy inferred from ARPES as compared to 
2014: that determined directly from neutron data, showing the two data sets agree.
2015: From Ref.~\cite{JC99}.  Right panel: mode energy inferred from tunneling data, 
2016: showing scaling with $T_{c}$.  Inset demonstrates that the mode energy 
2017: saturates to $2\Delta$ in the overdoped limit.  From Ref.~\cite{JOHNZ}.}
2018: \label{fig30}
2019: \end{figure}
2020: This was confirmed in greater 
2021: detail by tunneling \cite{JOHNZ}, where the energies were tracked over 
2022: a larger doping range with higher energy resolution (right panel, Fig.~\ref{fig30}).  
2023: Not only were 
2024: the inferred mode energy the same as the resonance energy, but the 
2025: mode energy saturated to $2\Delta$ in the overdoped limit as would be 
2026: expected for a collective mode inside a continuum gap (inset, right panel,
2027: Fig.~\ref{fig30}) \cite{JOHNZ}.
2028: 
2029: There are other things revealed by the doping dependence of the ARPES 
2030: data as well.  
2031: Both the peak energy and the hump energy increase strongly with 
2032: underdoping (Fig.~\ref{fig27}), yet the ratio of their energies is 
2033: roughly constant (3.5-4)
2034: as would be expected for a strong-coupling superconductor 
2035: \cite{JC99} (a result difficult to explain if their energy 
2036: separation were simply due to bilayer splitting).  Moreover, the hump 
2037: dispersion (Fig.~\ref{fig36})
2038: increasingly begins to resemble that (Fig.~\ref{fig22})
2039: expected of a spin 
2040: density wave insulator as the doping is reduced \cite{JC99}.  That is, 
2041: as far as the hump dispersion is concerned, the wavevector $(\pi,\pi)$ 
2042: begins to look more and more like a reciprocal lattice vector.  This 
2043: is not unexpected, since as the resonance mode energy goes soft with
2044: underdoping, the 
2045: material will be unstable to long range order at this wavevector.
2046: 
2047: Similar effects to the ARPES ones have been inferred from a generalized 
2048: Drude analysis of optics data, where the gap in the optical 
2049: scattering rate has been interpreted in a similar fashion 
2050: \cite{JULES}.  On the other hand, the optical scattering rate 
2051: resembles most closely the behavior of the ARPES self-energy at the 
2052: node, rather than at the $(\pi,0)$ point \cite{ADAM} (not surprising,
2053: since as $(\pi,0)$ is 
2054: a saddle point, the velocity there is zero, and so it does not 
2055: contribute to the in-plane optical response).
2056: 
2057: As mentioned previously, the normal state scattering rate at the node 
2058: from ARPES resembles that expected for a marginal Fermi liquid 
2059: \cite{OLSON,VALLA}.  What has been controversial, though, is how this 
2060: scattering rate changes below $T_{c}$.
2061: \begin{figure}
2062: \centerline{\epsfxsize=1.0\textwidth{\epsfbox{fig31.eps}}}
2063: \caption{Left:  Dispersion obtained from ARPES MDCs along the nodal 
2064: direction for optimal doped (90K) Bi2212.  
2065: The difference of the superconducting state (SC) as compared to the
2066: normal state (NS) gives rise to a kink in the dispersion.  Middle:
2067: The change in the EDC linewidth, with the drop in the 
2068: scattering rate in the superconducting state connected to the 
2069: dispersion kink by Kramers-Kronig relations.  Right:
2070: EDC at the nodal point.  Note the break in the superconducting
2071: case marking the
2072: separation of the coherent peak from the incoherent part.
2073: Adapted from Ref.~\cite{ADAM2}.}
2074: \label{fig31}
2075: \end{figure}
2076: The latest results are consistent 
2077: with a strong drop in the scattering rate below some threshold energy (middle
2078: panel, Fig.~\ref{fig31})
2079: \cite{ADAM}, though the expected superlinear behavior this implies 
2080: with temperature has yet to be positively identified \cite{VALLA}.  By 
2081: Kramers-Kronig, this drop implies a ``kink'' in the dispersion (left panel,
2082: Fig.~\ref{fig31}).  For 
2083: binding energies smaller than the ``kink'', the spectral peak is 
2084: sharper and less dispersive, for larger energies, broader and more 
2085: dispersive.  Surprisingly, the kink was not recognized at first
2086: (it was later identified by Shen's group 
2087: \cite{KINK}).  This kink is present throughout the zone 
2088: \cite{KINK}, and occurs at the same energy as the  ``break'' in the ARPES 
2089: lineshape at the Fermi surface separating the quasiparticle peak from 
2090: the incoherent part (right panel, Fig.~\ref{fig31}) \cite{ADAM}.  
2091: Moreover, it was later shown that this spectral 
2092: break evolves into the spectral dip as the momentum is swept in the 
2093: zone from the node to the $(\pi,0)$ point \cite{ADAM2}.  This led to 
2094: the speculation that the ``kink'' effect was due to the resonance as
2095: suggested earlier for  the spectral dip \cite{NORM97}.  This
2096: was later confirmed by theoretical simulations \cite{ESCHRIG}.  
2097: Strong support for this conjecture was offered by data from 
2098: Johnson's group \cite{PETER01}, where the energy scale associated with 
2099: the kink was found to track in doping with the neutron resonance energy.
2100: \begin{figure}
2101: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig32.eps}}}
2102: \caption{Variation of the real part of the self-energy at the kink 
2103: energy with temperature as determined from ARPES for Bi2212, compared to the 
2104: intensity of the magnetic resonance for YBCO.  From Ref.~\cite{PETER01}.}
2105: \label{fig32}
2106: \end{figure}
2107: More 
2108: importantly, the change in the self-energy with temperature
2109: associated with the kink has the same temperature dependence as
2110: the resonance intensity (Fig.~\ref{fig32}) \cite{PETER01}.
2111: 
2112: This picture has been challenged by Shen's group \cite{LANZARA}.  
2113: They observed that not only did the kink effect persist above $T_c$, 
2114: it was universally present in all cuprates (Bi2212, Bi2201, LSCO) at 
2115: roughly the same energy.  They argued that this implied the effect was 
2116: due to a phonon, since the dynamic spin susceptibility of Bi2212 and 
2117: LSCO look very different.  There is some attractiveness to this 
2118: phonon picture, but one should recognize that (1) most of the 
2119: ``normal'' state data were actually taken in the pseudogap phase and 
2120: (2) the constancy of the energy scale is somewhat surprising in a 
2121: phonon model as well, since the kink energy, even at the node, should be the sum 
2122: of the maximum superconducting gap energy plus the mode energy (phonon 
2123: or otherwise).
2124: Also, it is somewhat surprising that only a single 
2125: phonon energy would appear in the data.  Still, the arguments 
2126: being invoked in the Lanzara \etal paper \cite{LANZARA} are quite important, 
2127: in that they address the fundamental issue of whether the 
2128: many-body effects in the cuprates should be associated with phonons 
2129: (as in classic superconductors) or with electron-electron 
2130: interactions (as has been commonly assumed in the literature).
2131: 
2132: \subsection{Pseudogap Phase}
2133:   
2134: ARPES has revealed many unique features connected with the pseudogap phase, and 
2135: has had a profound influence on our understanding of this unusual 
2136: state of matter.
2137: 
2138: We start our discussion by considering states near the $(\pi,0)$ point 
2139: of the zone for underdoped samples.  Upon heating above $T_{c}$, the 
2140: sharp spectral peak disappears, but the leading edge of the spectrum
2141: is still pulled
2142: back from the chemical potential (leading edge gap, see Fig.~\ref{fig33})
2143: \cite{LOESER,DING}.
2144: \begin{figure}
2145: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig33.eps}}}
2146: \caption{ARPES spectrum at $(\pi,0)$ for an underdoped Bi2212 sample 
2147: in the superconducting state (30K) and the pseudogap phase (90K).  The 
2148: sharp peak in the superconducting state is replaced by a leading edge gap 
2149: in the pseudogap phase.  
2150: Data courtesy of A. Kaminski and J. C. Campuzano.}
2151: \label{fig33}
2152: \end{figure}
2153: This is quite unusual, in that the spectral function is completely
2154: incoherent in nature, but the leading edge is still quite sharp.  As 
2155: the temperature is raised, the leading edge gap appears to go away, 
2156: with the leading edge becoming degenerate with the Fermi function at 
2157: a temperature $T^{*}$ \cite{DING}, similar to $T^{*}$ inferred from NMR 
2158: measurements of the spin gap.
2159: 
2160: One of the more surprising findings, though, was that this leading 
2161: edge gap has an anisotropy in momentum space quite similar to the d-wave 
2162: gap in the 
2163: superconductor (Fig.~\ref{fig8}) \cite{LOESER,DING}.  This has been taken as strong support 
2164: for those theories proposing that the pseuodgap involves pairs of 
2165: some kind.
2166: \begin{figure}
2167: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig34.eps}}}
2168: \caption{Left panel shows the determination of the minimum gap locus 
2169: along a cut (open circles of right panel)
2170: in momentum space from the leading edge shift of ARPES 
2171: EDCs.  The right panel shows that the minimum gap locus (MGL) of the 
2172: pseudogap phase matches the normal state Fermi surface (FS).  From 
2173: Ref.~\cite{DING97}.}
2174: \label{fig34}
2175: \end{figure}
2176: In further support of this picture, it was observed that 
2177: the minimum gap locus coincided with the normal state
2178: Fermi surface (Fig.~\ref{fig34}) \cite{DING97}, 
2179: as occurs for superconductors.  This set of points is obtained by 
2180: taking various cuts in momentum space, and looking for that point along 
2181: the cut where the leading edge gap is smallest.  This behavior can be 
2182: contrasted with a spin density wave precursor, for instance, where 
2183: the minimum gap locus would have a new symmetry defined by the 
2184: magnetic Brillouin zone boundary running from $(\pi,0)$ to $(0,\pi)$ 
2185: (see Fig.~\ref{fig22}).
2186: 
2187: There are, though, a number of unusual features of the data
2188: which are not as easily
2189: understood in terms of a precursor pairing scenario.  In 
2190: particular, the pseudogap phase does not have a node like for a d-wave 
2191: superconductor, instead, it possesses a ``Fermi arc'' 
2192: (Fig.~\ref{fig8}) centered at the 
2193: d-wave node \cite{MARSHALL}.  This arc expands in temperature, 
2194: eventually recovering the full Fermi surface at $T^{*}$ \cite{NAT98}.
2195: 
2196: For states near the $(\pi,0)$ point, the pseuodgap appears to fill in 
2197: with temperature rather than close \cite{NAT98}, much like what is observed 
2198: in c-axis 
2199: conductivity (Fig.~\ref{fig7}) \cite{HOMES}.
2200: \begin{figure}
2201: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig35.eps}}}
2202: \caption{Filling in of the spectral gap at the antinode 
2203: (I, $(\pi,0)-(\pi,\pi)$ Fermi crossing) as compared to the closing of the 
2204: spectral gap about halfway between the antinode and the node (II).  
2205: Symmetrized ARPES data for underdoped (75K) Bi2212 from Ref.~\cite{NAT98}.}
2206: \label{fig35}
2207: \end{figure}
2208: On the other hand, away from this region, 
2209: the spectral gap clearly closes \cite{NAT98}.  This is most easily 
2210: visualized (Fig.~\ref{fig35}) by ``symmetrizing'' the data (a way of removing the Fermi 
2211: function from the data by assuming particle-hole symmetry in the spectral 
2212: function).  This behavior has been further confirmed by data fitting 
2213: \cite{PHEN98}, where the spectral gap parameter, $\Delta$, at the antinode 
2214: ($(\pi,0)-(\pi,\pi)$ Fermi crossing) is found to be relatively temperature 
2215: independent.  The self-energy has the form 
2216: $\Sigma=-i\Gamma_{1}+\Delta^{2}/(\omega+i\Gamma_{0})$.  $\Gamma_{1}$ is a 
2217: crude approximation for the normal self-energy (this term becomes
2218: strongly reduced in the superconducting state with the onset of 
2219: coherence), whereas $\Gamma_{0}$ is the lifetime of the pair propagator, which 
2220: is proportional to $T-T_{c}$.
2221: Note the contrast with
2222: Eliashberg theory, where $\Gamma_0$ would be equal to $\Gamma_1$.
2223: An incoherent spectrum with a pseudogap is
2224: formed when $\Gamma_0 \ll \Delta \ll \Gamma_1$.  The strong
2225: temperature variation of $\Gamma_0$ leads to a filling in of the pseuodgap
2226: ($\Delta$ being roughly constant in $T$).
2227: $T^{*}$ is then simply 
2228: the temperature where $\Delta = \Gamma_{0}(T)$.
2229: This behavior can be contrasted with that away from the $(\pi,0)$ 
2230: region, where $\Delta$ closes with temperature in a BCS 
2231: like fashion \cite{PHEN98}.
2232: 
2233: These findings seem to imply the possibility of two regions in the 
2234: Brillouin
2235: zone, a ``pseudogap'' region centered at the $(\pi,0)$ point and an 
2236: ``arc'' region centered at the d-wave node.  This picture would be in 
2237: support of a competitive scenario, where the pseudogap and 
2238: superconducting gap were different phenomena.  On the other hand, 
2239: newer high resoultion data do not necessarily support the picture of two 
2240: regions of the zone, rather it appears that the gap ``closing'' and 
2241: gap ``filling in'' behaviors smoothly evolve into one another as a 
2242: function of momentum.  In fact, there are several pair precursor 
2243: calculations \cite{JAN,GIAN} which predict the presence of Fermi 
2244: arcs.  In the strong-coupling RVB approaches, these Fermi arcs are 
2245: also found, and are
2246: due to fluctuations in the pseudogap regime between d-wave pairs and 
2247: the staggered flux phase state, which are nearly degenerate in energy
2248: \cite{PLEE}.  Arcs are also found in a one loop renormalization group 
2249: treatment of interacting fermions in 2D \cite{FRS}, and in a high 
2250: temperature expansion study of the 2D t-J model \cite{PUTIKKA}.
2251: 
2252: The resemblance of the arcs to one side of a hole pocket 
2253: (Fig.~\ref{fig22}) has been 
2254: noted by a number of authors \cite{MARSHALL}, and as such hole pockets 
2255: are expected when doping a magnetic insulator, then a magnetic 
2256: precursor scenario is a possibility.  On the other hand, there 
2257: is no clear evidence from ARPES that the arc deviates from the large
2258: Fermi surface and ``turns in" so that its normal would be parallel to the magnetic
2259: zone boundary at the magnetic zone crossing as would be expected in such a scenario
2260: (see Fig.~\ref{fig22}).  And
2261: although ``shadow'' bands have been seen in ARPES \cite{AEBI} (the 
2262: image of the main band translated by $Q=(\pi,\pi)$, which would thus form the 
2263: back side of this pocket), their intensity seems to scale with 
2264: $T_{c}$ \cite{DRESDEN}, and thus drops off as the 
2265: doping is reduced, in complete contrast with the expected behavior if 
2266: the shadows were due to magnetic correlations.
2267: 
2268: All of the above discussion concerns the leading edge gap, also known 
2269: as the strong pseudogap.  ARPES studies also find a higher energy 
2270: pseudogap, known as the weak pseudogap.  The presence of the latter was 
2271: evident from the earliest studies \cite{MARSHALL}, but it was not 
2272: until later that the two effects were clearly differentiated 
2273: \cite{JC99}.  In constrast to the leading edge gap, which appears to 
2274: be a precursor to the d-wave superconducting gap, the high energy 
2275: pseudogap behaves differently.  It is simply the continuation of the 
2276: ``hump'' from the superconducting state, and has a dispersion 
2277: which increasingly resembles that of a spin density wave insulator as 
2278: the doping is reduced (Fig.~\ref{fig36}) \cite{JC99}.
2279: \begin{figure}
2280: \centerline{\epsfxsize=0.4\textwidth{\epsfbox{fig36.eps}}}
2281: \caption{Peak (top) and hump (bottom) dispersion from ARPES data of
2282: Bi2212 as a function 
2283: of doping 
2284: (superconducting state).  These quantities become the leading edge 
2285: (strong) pseudogap and the high energy (weak) pseudogap in the 
2286: pseudogap phase.  The two dispersion directions shown become increasingly 
2287: similar as the doping is reduced.  If magnetic long range order was
2288: present, the directions would be equivalent.  From Ref.~\cite{JC99}.}
2289: \label{fig36}
2290: \end{figure}
2291: In essence, the sharp spectral 
2292: peak in the superconducting state is
2293: replaced by the leading edge gap, the spectral dip is filled in, and 
2294: the high energy hump becomes the weak pseudogap.
2295: This high energy gap is 
2296: what is commonly observed by ARPES in LSCO and NCCO, as the actual 
2297: superconducting gaps are difficult to see in these materials because 
2298: of their small size.  This gap strongly increases with underdoping, 
2299: and adiabatically connects to the Mott insulating gap of the undoped 
2300: phase \cite{BIGBOB}.  The presence of these two gaps may resolve the 
2301: precursor pair versus competitive scenario debate, in that the leading 
2302: edge gap 
2303: seems to be the precursor to the superconducting gap, whereas the 
2304: high energy gap is the precursor to the magnetic insulating gap.  Of 
2305: course, these two gaps are connected, in that their energies scale 
2306: together with doping, with a ratio of 3.5-4 (Fig.~\ref{fig27}) \cite{JC99}.
2307: This again demonstrates the intimate 
2308: relation of magnetic and pairing correlations in the cuprates.
2309: 
2310: None of the spectroscopic data, though, support a picture where the 
2311: pseudogap phase represents a phase with true long range order, as 
2312: advocated by a number of theories, in particular those involving a 
2313: quantum critical point near optimal doping.  On the other hand, a 
2314: recent ARPES experiment does find evidence for broken symmetry in the 
2315: pseudogap phase (Fig.~\ref{fig37}) 
2316: \cite{NAT02}.  In these experiments, circularly polarized light is 
2317: employed.  In general, the signal for left and right polarized light 
2318: is different, but in a mirror plane, they should be equivalent.  This 
2319: mirror plane effect is seen in overdoped samples along the 
2320: $(0,0)-(\pi,0)$ ($\Gamma-M$) line.  But in underdoped 
2321: samples, the signals are no longer equivalent in this direction in 
2322: the pseudogap phase, rather, they become degenerate at some other k 
2323: point shifted off this line.
2324: \begin{figure}
2325: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig37.eps}}}
2326: \caption{Circularly polarized ARPES of underdoped Bi2212.  The intensity 
2327: difference between left and right polarized light normalized to the 
2328: average is plotted along $(\pi,0)-(\pi,\pi)$ in the 
2329: left panels, with M1 (panel a) and M2 (panel b) indicating orthogonal 
2330: ($(\pi,0)$ and $(0,\pi)$) directions (circles for T=250K, squares for T=100K).  
2331: $k_{y}=0$ is the mirror plane.
2332: The shift with temperature at $k_y=0$
2333: represents chiral symmetry breaking, and is plotted in panel c versus 
2334: temperature for an 
2335: overdoped (squares) and underdoped (circles) sample.  
2336: Panel d shows that this shift (diamonds) only exists for temperatures 
2337: below the pseudogap temperature, $T^{*}$ (circles indicate no 
2338: shift).  From Ref.~\cite{NAT02}.}
2339: \label{fig37}
2340: \end{figure}
2341: Moreover, the size of this shift 
2342: increases below $T^{*}$ as would be expected if an order parameter
2343: developed (Fig.~\ref{fig37}c).  The 
2344: implication is that time reversal symmetry (or chiral symmetry, 
2345: depending upon interpretation), is being broken in the pseudogap phase.
2346: 
2347: The first worry about such an experiment is that the $(0,0)-(\pi,0)$ 
2348: direction is technically not a mirror plane in the Bi2212 crystal 
2349: structure (due to the orthorhombicity and superstructure).  On the 
2350: other hand, it has been known for some time that the main band signal 
2351: from ARPES in Bi2212 appears to obey dipole selection rules consistent 
2352: with tetragonal symmetry \cite{DING96}, and such is the case in these 
2353: measurements as well.  Of course, there could be some structural 
2354: effect associated with $T^{*}$, but these authors did x-ray 
2355: scattering on the samples, and found no evidence for a structural 
2356: change below $T^{*}$ \cite{NAT02}.  Moreover, the effect of the pseudogap is to 
2357: shift the overall intensity of the left and right signals relative to 
2358: one another, as if chiral symmetry was being broken.
2359: 
2360: A similar effect has not been seen in the $(0,0)-(\pi,\pi)$ 
2361: ($\Gamma-Y$)
2362: mirror plane, though this has not been studied as extensively yet.  
2363: If this continues to hold, then it has definite implications.  
2364: Simple ferromagnetism would cause an effect in both mirror planes.  In 
2365: addition, most orbital current models (the d density wave state, for 
2366: example) would predict an effect along $\Gamma-Y$ and not along 
2367: $\Gamma-M$ (opposite to experiment).  One of the two orbital current 
2368: patterns discussed by Varma (left panel, Fig.~\ref{fig20})
2369: behaves the same way, but the other (right panel, Fig.~\ref{fig20}) has 
2370: a signature similar to experiment \cite{VARMA02}.  So, it is indeed 
2371: possible that the data represent an effect which can be attributed to 
2372: orbital currents, but more experiments would certainly be 
2373: desirable.  What this particular experiment illustrates, though, is the 
2374: power
2375: of photoemission in addressing fundamental issues connected with 
2376: the cuprates.
2377: 
2378: \section{Inelastic Neutron Scattering}
2379: 
2380: The other momentum resolved probe in the cuprates is inelastic 
2381: neutron scattering.  The part of the signal of interest here is the 
2382: magnetic part, which is proportional to the imaginary part of the
2383: spin-spin response function, $\chi({\bf q},\omega)$, times a Bose 
2384: population factor.  For elastic scattering, one sees Bragg peaks 
2385: associated with the magnetism if the material is magnetically ordered.
2386: (Phonons and structure are measured by neutrons as well, but this takes us 
2387: beyond the scope of this review).
2388: 
2389: The first result with neutrons was finding the antiferromagnetic order in the 
2390: undoped phase \cite{INS}.  Magnetic moments of 2/3 $\mu_{B}$ per Cu 
2391: site are 
2392: found \cite{JOHN88}, the reduction from 1 being due to quantum 
2393: fluctuations associated with the small spin (S=1/2) of the Cu ion.  
2394: The ordering wavevector is  $Q=(\pi,\pi,\pi)$, which means that 
2395: successive planes are antiferromagnetically coupled as well.
2396: 
2397: For bilayer systems like YBCO, there are two branches of the spectrum
2398: (Fig.~\ref{fig38}), an 
2399: acoustic branch with form factor $\sin^{2}(Q_{z}d/2)$ and an optic branch 
2400: with form factor $\cos^{2}(Q_{z}d/2)$,
2401: where $d$ is the separation of the two CuO layers of the bilayer \cite{REZNIK}.
2402: \begin{figure}
2403: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig38.eps}}}
2404: \caption{Acoustic and optic branches of the spin wave dispersion in the 
2405: magnetic phase of YBCO as revealed by inelastic neutron scattering.
2406: Adapted from Ref.~\cite{REZNIK}.}
2407: \label{fig38}
2408: \end{figure}
2409: The acoustic branch has the classic spin-wave dispersion with respect 
2410: to $(\pi,\pi)$, with the planar exchage energy $J_{\parallel} \sim 100$ meV
2411: determined from the slope.  The optic branch has a gap of order 60 meV.  As the
2412: optic gap is equal to $2\sqrt{J_{\parallel}J_{\perp}}$, where 
2413: $J_{\perp}$ is the intrabilayer exchange, then $J_{\perp} \sim 10$ meV \cite{FONG00}.
2414: 
2415: \subsection{LSCO}
2416: 
2417: The first doping dependent studies were done on LSCO.  They revealed the 
2418: presence of four incommensurate peaks (see Fig.~\ref{fig10}) at locations 
2419: $2\pi(0.5\pm\delta,0.5)$ and $2\pi(0.5,0.5\pm\delta)$, with $\delta$ scaling 
2420: with the doping, $x$ \cite{NS-LSCO}.
2421: The original explanation for this incommensurability was related to 
2422: the Fermi surface geometry.  As the doping increases, the 
2423: Fermi surface hole volume expands, and the predicted 
2424: incommensurability with doping from RPA calculations more or less agrees 
2425: with experiment \cite{LEVIN}.
2426: 
2427: This view changed, though, with the observation of elastic scattering 
2428: peaks in the LTT (low temperature tetragonal)
2429: phase of Nd doped LSCO \cite{JOHN95}.  The 
2430: incommensurate elastic peaks were accompained by charge ordering 
2431: peaks (see Fig.~\ref{fig10}) 
2432: at $2\pi(\pm2\delta,0)$ and $2\pi(0,\pm2\delta)$.  Tranquada and co-workers 
2433: interpreted this behavior as due to the formation of stripes of doped 
2434: holes with commensurate antiferromagnetic domains between the 
2435: stripes.  If the stripes act as antiphase domain walls, then the 
2436: prediction is that the magnetic signal will be incommensurate, with 
2437: $\delta$ proportional to $x$ (upper panel, Fig.~\ref{fig39}).
2438: \begin{figure}
2439: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig39.eps}}}
2440: \caption{Neutron scattering peaks versus doping for LSCO.  The spot 
2441: pattern rotates by 45$^{\circ}$ at the spin glass/superconducting 
2442: boundary.  $\delta$ is the incommenusability, and $\alpha$ the angle 
2443: of the spots in momentum space relative to the (1/2,1/2) wavevector.  
2444: From Ref.~\cite{YAMADA}.}
2445: \label{fig39}
2446: \end{figure}
2447: This picture received further support 
2448: when it was observed in LSCO that 
2449: for dopings smaller than those where superconductivity occured, only 
2450: two spots were present, and they were rotated $45^{\circ}$ relative 
2451: to the previous spots (bottom panel, Fig.~\ref{fig39}) \cite{YAMADA}.  This implies one 
2452: dimensional behavior, consistent with stripes.
2453: 
2454: A similar incommensurate pattern was later seen in YBCO 
2455: \cite{NS-YBCO}, and this pattern is also found to be 1D like in detwinned
2456: samples (where the CuO chains are all aligned) \cite{YBCO1D}.
2457: This again gives evidence for 
2458: stripes, though the effect may have a more benign origin due to the
2459: influence of the CuO chains.  At low dopings, charge ordering 
2460: peaks are seen in YBCO as well \cite{MOOK02}.
2461: 
2462: The effect of superconductivity on LSCO is to lead to a sharpening of 
2463: the incommensurate peaks, and the formation of a ``spin gap'' at low 
2464: energies of $\sim$ 6 meV (Fig.~\ref{fig40}).  This spin gap is fairly isotropic in momentum space 
2465: \cite{LAKE1}, which was taken as evidence against an
2466: interpretation of it being the 2$\Delta$ continuum gap since 
2467: $\Delta$ is anisotropic for a d-wave superconductor (in particular, there should be
2468: a significant continuum gap at $(\pi,\pi)$).
2469: This statement should be treated with care, though, since the only low 
2470: frequency structure is at these incommensurate wavevectors, with the 
2471: intensity at other wavevectors, like $(\pi,\pi)$,
2472: due to overlap from these peaks given their finite width in momentum.
2473: 
2474: Subsequent experiments found that this spin gap filled in at modest 
2475: values ($H \ll H_{c2}$) of the magnetic field (Fig.~\ref{fig40}) \cite{LAKE2}.
2476: \begin{figure}
2477: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig40.eps}}}
2478: \caption{Imaginary part of the dynamic susceptibility versus energy 
2479: (panel A)
2480: for the superconducting state of optimal doped (x=0.163) LSCO
2481: at zero field (lower set of circles), at a field 
2482: of 7.5 T (upper set of circles), and in the normal state (triangles).  The 
2483: difference of zero and 7.5 T data are plotted in the panel B.  
2484: From Ref.~\cite{LAKE2}.}
2485: \label{fig40}
2486: \end{figure}
2487: For more underdoped 
2488: samples, magnetic ordering could even be induced by applying a field
2489: \cite{LAKE3}.  The 
2490: obvious idea is that some kind of SDW ordering is being stabilized by 
2491: the vortices \cite{AROVAS,DSZ}.  But, the vortex density 
2492: is quite 
2493: low at the field values studied, which would imply that there is a
2494: very large magnetic polarization cloud around the vortices.  This
2495: is certainly consistent with the neutron results, in that the magnetic 
2496: correlation length is quite long in underdoped LSCO samples.
2497: 
2498: On the other hand, one might interpret these results as a 
2499: stabilization of stripe formation.  In this context, STM experiments 
2500: on Bi2212 find a charge density wave pattern associated with the 
2501: vortex cores \cite{HOFFMAN}.  The Fourier pattern was anisotropic (factor
2502: of 3 intensity difference between orthogonal planar directions), which 
2503: would argue for 1D behavior, although it is possible that this 
2504: could be an extrinsic effect due to the STM tip.
2505: 
2506: \subsection{YBCO}
2507: 
2508: Perhaps the most dramatic effect associated with neutron scattering in 
2509: the superconducting state is the formation of a sharp commensurate
2510: resonance at 
2511: about 40 meV in YBCO (Fig.~\ref{fig41}) \cite{ROSSAT}.
2512: \begin{figure}
2513: \centerline{\epsfxsize=1.0\textwidth{\epsfbox{fig41.eps}}}
2514: \caption{Wavevector integrated dynamic susceptibility for underdoped 
2515: YBCO at 35K (superconducting state), 80K (pseudogap phase), and 
2516: 290K (normal state) in the acoustic and optic channels.  The gray 
2517: shaded area represents the resonance.  From Ref.\cite{DAI99}.}
2518: \label{fig41}
2519: \end{figure}
2520: The magnetic nature of the 
2521: resonance was confirmed by later polarized measurements 
2522: \cite{MOOK93}.  Subsequently, the resonance was seen in Bi2212 
2523: \cite{KEIMER}.  
2524: The resonance has a form factor equivalent to the 
2525: acoustic branch of the undoped material, and so is centered about 
2526: the $(\pi,\pi,\pi)$ wavevector.  In particular, the ``optic branch'' 
2527: has no resonance, and remains gapped as in the insulator (Fig.~\ref{fig41})
2528: \cite{FONG00}.  These observations led to speculations that the 
2529: resonance might be a bilayer effect, since is it not seen in LSCO, 
2530: but a new experiment has identified the resonance in single layer 
2531: Tl2201 \cite{TL2201}.  Several 
2532: theories concerning the resonance were discussed in Section 2, in 
2533: particular the controversy concerning whether it represents a 
2534: particle-hole or particle-particle collective mode.
2535: 
2536: The resonance energy scales with doping like 
2537: $5T_{c}$ \cite{FONG00}, 
2538: and its intensity has a variation with temperature much like that of
2539: the superconducting order parameter \cite{FONG96}.  Based on previous theoretical work 
2540: \cite{WHITE}, Demler and Zhang made the provocative suggestion
2541: that these results implied an 
2542: equivalence between the exchange energy difference between the normal 
2543: and superconducting state and the resonance weight \cite{DZNAT}; that 
2544: is, the superconducting condensation energy was related to the 
2545: formation of the resonance.
2546: \begin{figure}
2547: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig42.eps}}}
2548: \caption{Neutron data for YBCO.  Panels A-C are the temperature 
2549: variation of the resonance peak intensity for samples at three different 
2550: dopings.  Panel D is the temperature derivative of the intensity data, in 
2551: comparison to specific heat data in panel E.  From Ref.~\cite{DAI99}.}
2552: \label{fig42}
2553: \end{figure}
2554: Their suggestion was tested by
2555: measurements on YBCO, which demonstrated a 
2556: similarity of the specific heat anomaly at $T_{c}$ and the temperature 
2557: derivative of the resonance weight (Fig.~\ref{fig42}) \cite{DAI99}.  As magnetic fields 
2558: along the c-axis are known to suppress the specific heat anomaly, this 
2559: motivated experiments which looked at the field dependence of the 
2560: resonance.  It was found that a field applied 
2561: perpendicular to the planes did lead to a strong suppression of the 
2562: resonance in an underdoped YBCO sample \cite{DAI00}, as opposed to a 
2563: parallel field which did not \cite{BOURGES}.  To understand this result,
2564: we note that STM measurements find in underdoped Bi2212 samples that the 
2565: pseudogap phase is present in the vortex cores, with no coherence
2566: peaks \cite{CORE}.  This
2567: would imply that the resonance, which is a property of the coherent
2568: superconducting state, is suppressed in the cores.  On the other 
2569: hand, an analysis of the neutron data indicates that the effective core radius 
2570: needed to explain the observed suppression of the resonance is significantly larger 
2571: than the known superconducting correlation length \cite{RESVOR}.  This 
2572: again demonstrates that for YBCO, like for LSCO, the 
2573: cores polarize the surrounding medium.  This leads to a suppression of the
2574: resonance out to a magnetic correlation length about the cores.  In that sense, 
2575: it should be remarked that the YBCO sample studied in the field 
2576: dependent experiment exhibits an anomalously large magnetic 
2577: correlation length (28 $\AA$) as determined from the resonance width
2578: in momentum space \cite{DAI01}.
2579: 
2580: Perhaps the biggest controversy surrounding the resonance is whether it is 
2581: responsible for structure seen in other spectroscopic probes, like 
2582: ARPES and tunneling \cite{NORM97}.  The smallness of the resonance 
2583: weight (a few percent of the local moment sum rule) argues against 
2584: this \cite{KEE}.  On the other hand, phase space considerations can be used 
2585: as a counterargument \cite{GANG5}.  The resonance weight is small 
2586: because it is localized in momentum space around $(\pi,\pi)$.  But as 
2587: electronic states near $(\pi,0)$ are connected by these wavevectors,
2588: there is no problem for these states to 
2589: be strongly affected by the resonance despite its overall small 
2590: weight.  These arguments can be extended to states in other regions 
2591: of the zone as well \cite{ESCHRIG}.
2592: 
2593: One reason this controversy has arisen is the suggestion by certain 
2594: researchers that the resonance interpretation of ARPES and tunneling
2595: implies that the 
2596: resonance alone is responsible for pairing.  Actually, the 
2597: formation of the resonance, as well as the profound changes in the 
2598: ARPES lineshape, is a consequence of the onset of superconductivity, 
2599: rather than the cause of it.  Such ``feedback'' effects are 
2600: unavoidable in any theory where the pairing is not due to 
2601: phonons \cite{AC}.  The classic example is the stabilization of the A phase in 
2602: $^{3}He$ relative to the B phase due to the feedback of pairing on 
2603: the spin fluctuations \cite{BA}.
2604: 
2605: Below the resonance energy in YBCO, the magnetic response becomes 
2606: incommensurate (top left panel, Fig.~\ref{fig43}) \cite{NS-YBCO}.
2607: \begin{figure}
2608: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig43.eps}}}
2609: \caption{RPA calculation of the dynamic susceptibility for YBCO in the 
2610: superconducting state based 
2611: on ARPES dispersions.  Momentum dependence of the intensity is 
2612: plotted at resonance in top right panel, and below resonance in top
2613: left panel.  Note the commensurate pattern at resonance compared to 
2614: the incommensurate (diamond shaped) pattern below resonance.  
2615: Resulting $\omega(q)$ dispersion relation (hourglass shaped)
2616: is plotted in the bottom left panel.  (The maximum superconducting gap 
2617: was 29 meV and the superexchange energy 110 meV.)  This is
2618: further illustrated in the bottom right panel, where the pole of the RPA response
2619: function is plotted (circles), with the solid line representing the edge of the
2620: continuum.
2621: Adapted from Refs.~\cite{NORM00} and \cite{NORMAN2}.}
2622: \label{fig43}
2623: \end{figure}
2624: Detailed studies of the $\omega({\bf 
2625: q})$ dispersion relation find an ``hourglass'' shape (bottom left panel,
2626: Fig.~\ref{fig43}) \cite{ARAI}, with 
2627: incommensurate ``sidebranches'' appearing both above and below the 
2628: resonance energy, though the upper branch is damped (it lies 
2629: above the two particle continuum).  This sidebranch behavior can be 
2630: understood from 
2631: linear response calculations for a d-wave superconductor 
2632: (Fig.~\ref{fig18}) 
2633: \cite{FLORA,KAO,NORM00}.  Under certain conditions, the 
2634: lower incommensurate branch is a collective mode
2635: pulled below the two particle continuum, with the commensurate
2636: resonance at the top of 
2637: this lower branch (bottom right panel, Fig.~\ref{fig43})  \cite{FLORA,NORMAN2}.
2638: This is consistent with an interpretation of recent 
2639: data on slightly underdoped YBCO \cite{BOURGES00}, but differs from an 
2640: interpretation of more heavily underdoped samples, where the resonance 
2641: and sidebranches appear to represent separate effects \cite{MOOK02}.  
2642: In the latter case, the incommensurate sidebranches have been interpreted as 
2643: due to antiphase domain stripes, with the commensurate resonance 
2644: presumably due in phase domain stripes.
2645: It should be remarked that the 
2646: condition to get pole-like behavior for the lower incommensurate 
2647: branch is difficult in linear response calculations, and requires 
2648: reduced curvature of the
2649: Fermi surface around the nodes, so that the constant energy contours of the Dirac cones 
2650: discussed in Section 3 (Fig.~\ref{fig26}) are flat instead of banana 
2651: shaped \cite{BLEE,NORM00,NORMAN2}.  For instance, the ARPES Fermi 
2652: surface of Bi2212 does not indicate such flat contours, 
2653: and the prediction would be that the incommensurate effects in
2654: Bi2212 are weak and 
2655: unconnected to the resonance \cite{NORM00,NORMAN2}, as inferred 
2656: experimentally in heavily underdoped YBCO \cite{MOOK02}.  Unfortunately, 
2657: neutron scattering experiments on Bi2212 are difficult due to the 
2658: small crystals available, and ARPES results for YBCO are controversial 
2659: because of surface related problems.  Still, such studies would 
2660: be useful to test these ideas.
2661: 
2662: As discussed by Batista and 
2663: co-workers \cite{BALATSKY}, differentiation between these various 
2664: interpretations is not as straightforward as it might seem.
2665: \begin{figure}
2666: \centerline{\epsfxsize=0.4\textwidth{\epsfbox{fig44.eps}}}
2667: \caption{Proposed $\omega(q)$ dispersion relation for the dynamic 
2668: susceptibility of YBCO based on overlapping spin waves from the AF 
2669: domains between stripes.  Adapted from Ref.~\cite{BALATSKY}.}
2670: \label{fig44}
2671: \end{figure}
2672: The 
2673: picture they offer is that the lower incommensurate branch is 
2674: analogous to the spin wave dispersion in an incommensurate antiferromagnet, 
2675: with the resonance being where the two spin wave branches from +Q and -Q 
2676: intersect (Fig.~\ref{fig44}).  Actually, the dispersion they plot follows the two 
2677: particle continuum gap of the linear response calculations.  In these 
2678: RPA calculations, the incommensurate branch either follows the 
2679: continuum gap (and thus is not a pole), or is pulled beneath (becoming 
2680: a pole), depending on how flat the constant energy contours of the Dirac cones are
2681: \cite{FLORA,NORMAN2}.  Presumably, this physics is not unrelated to a 
2682: stripes interpretation, where quasi 1D behavior would cause the same 
2683: effects as the flat contours.
2684: 
2685: Finally, although a lot has been made about the differences between 
2686: the dynamic susceptibilities of YBCO and LSCO, it should be noted that 
2687: the most recent data on LSCO find the presence of energy dispersion in 
2688: the incommensurate response (like in YBCO), with commensurate excitations
2689: present beyond 20 meV \cite{HIRAKA}.  Thus, it may be that 
2690: the central difference between these two materials is that the bulk 
2691: of the spectral weight sits in the lower incommensurate branch for 
2692: LSCO, but in the commensurate resonance for YBCO.  This difference
2693: is probably 
2694: due to a variety of factors:  the smaller energy gap in LSCO, 
2695: differences in the Fermi surface topology, and the stronger 
2696: tendency for disorder and inhomogeneous behavior (stripes) in LSCO.  In this 
2697: respect, Bi2212 is probably intermediate between YBCO and LSCO, so 
2698: more neutron studies of Bi2212 would be desirable.
2699: 
2700: \section{Optical Conductivity}
2701: 
2702: Optics measures a two-particle response function as well, the 
2703: current-current correlation function.  Because of the tiny momentum 
2704: associated with the light used, optics measures the zero $q$ limit of 
2705: this function.  
2706: This can be represented as a particle-hole bubble, like in the 
2707: previous section, but with the spin operators replaced by current 
2708: operators at the vertices (the Kubo bubble).
2709: 
2710: We start with the planar response.  The normal state is characterized 
2711: by a broad, Drude-like response centered at $\omega=0$ (Fig.~\ref{fig45}).  
2712: The Drude 
2713: tail, though, has an anomalous behavior.  The data are best 
2714: appreciated by representing the optical response in a generalized 
2715: Drude form \cite{ZACK,PUCH}
2716: \begin{equation}
2717: \sigma(\omega) = \frac{1}{4\pi}
2718: \frac{\omega_{pl}^{2}}{1/\tau(\omega)-i\omega[1+\lambda(\omega)]}
2719: \end{equation}
2720: where $\omega_{pl}$ (the plasma frequency) is given by the sum rule
2721: \begin{equation}
2722: \int_{0}^{\infty} d\omega Re\sigma(\omega) = \omega_{pl}^{2}/8
2723: \end{equation}
2724: (the last integral is usually cut-off at 1 eV or so to avoid interband 
2725: contributions).  In this form, $1/\tau$ is the scattering rate, and 
2726: $1+\lambda$ the mass renormalization.  The two terms are related by 
2727: a Kramers-Kronig tranform.  Such an analysis reveals that the 
2728: scattering rate has the form $a+b\omega$ (Fig.~\ref{fig46}).  The $b$ term is what is
2729: expected for a marginal Fermi liquid, but the $a$ term is not,
2730: since it does not appear to be proportional to T.
2731: Abrahams and Varma attribute it to scattering from 
2732: off planar impurities \cite{AV}, but more likely, it is a signature 
2733: of the non-Fermi liquid nature of the normal state.  In particular, there is some
2734: evidence from ARPES that the $a$ term is associated with the pseudogap,
2735: since it has a similar momentum anisotropy.
2736: 
2737: \begin{figure}
2738: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig45.eps}}}
2739: \caption{Real part of planar infrared conductivity of underdoped YBCO 
2740: at three temperatures corresponding to the superconducting state (10K), 
2741: pseudogap phase (65K), and normal state (300K).  Note gap-like 
2742: depression which develops at around 50 meV.  From Ref.~\cite{PUCH}.}
2743: \label{fig45}
2744: \end{figure}
2745: 
2746: When superconductivity sets in, $Re\sigma$ develops a depression at 
2747: an energy close to the value $2\Delta$ expected for a superconductor 
2748: \cite{COOPER,ZACK,JOE2,ROTTER,PUCH}, but a true gap never fully 
2749: develops.  Instead, a narrower Drude peak is present at energies
2750: below this depression, 
2751: representing uncondensed carriers.  For a superconductor, one 
2752: expects the presence of a $\delta$ function at $\omega=0$ due to the 
2753: dissipationless response of the superfluid, and 
2754: this is indeed seen as a $1/\omega$ term in $Im\sigma$ at low 
2755: frequencies.
2756: 
2757: A generalized Drude analysis in the superconducting state \cite{PUCH} 
2758: reveals that $1/\tau$ becomes strongly gapped below some threshold 
2759: energy (Figs.~\ref{fig46} and \ref{fig47}), and the behavior seen is 
2760: very similar to that inferred from ARPES along the nodal direction (Fig.~\ref{fig46})
2761: \cite{ADAM}.
2762: \begin{figure}
2763: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig46.eps}}}
2764: \caption{Comparison of optical scattering rate, $1/\tau$, from 
2765: Ref.~\cite{PUCH} to ARPES nodal linewidth in the superconducting 
2766: state and pseudogap phase of optimal doped Bi2212.  From 
2767: Ref.~\cite{ADAM}.}
2768: \label{fig46}
2769: \end{figure}
2770: The low frequency limit is most easily seen from 
2771: microwave \cite{BONN} and thermal conductivity \cite{ONG} measurements,
2772: which find a strong collapse with temperature (see Fig.~\ref{fig5})
2773: of the scattering rate 
2774: below $T_{c}$, with a residual low temperature
2775: mean free path of order microns for 
2776: clean samples of YBCO.  It should be remembered that for 
2777: electron-electron scattering, only Umklapp processes 
2778: contribute to the electromagnetic response, whereas normal processes 
2779: contribute to the ARPES and thermal response as well.  This has been 
2780: used to quantitatively account for differences between the microwave 
2781: and thermal conductivity scattering rates \cite{DUFFY}.  We should 
2782: note that although the strong collapse of the scattering rate
2783: with temperature below $T_c$ is consistent 
2784: with ARPES results near $(\pi,0)$ \cite{PHEN98}, the ARPES results along 
2785: the nodal direction are still controversial, where a linear T behavior below $T_c$
2786: has been claimed by one group (Fig.~\ref{fig23}) \cite{VALLA} but not by others.  All ARPES 
2787: results are on Bi2212, which is more disordered than YBCO, and in fact 
2788: terahertz ($\omega \sim 1$ meV) measurements on Bi2212 claim a linear T scattering rate of 
2789: the residual carriers as well \cite{JOE3}.  The size of this rate
2790: is difficult to compare to 
2791: ARPES, given the fact that the ARPES resolution width greatly exceeds 
2792: the frequency of this measurement.
2793: 
2794: \begin{figure}
2795: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig47.eps}}}
2796: \caption{Variation of optical scattering rate with energy for 
2797: underdoped YBCO (10K - superconducting state, 65K - pseudogap phase, 
2798: 300K - normal state).  From Ref.~\cite{PUCH}.}
2799: \label{fig47}
2800: \end{figure}
2801: 
2802: The other interesting point is that for underdoped samples, a partial 
2803: scattering rate gap persists in the pseudogap phase (Fig.~\ref{fig47}), likely due to the
2804: gapping of electronic states near $(\pi,0)$ \cite{PUCH}.  In 
2805: fact, the optics data have a smooth evolution through $T_{c}$ for 
2806: underdoped samples.  Even a finite frequency signature of the 
2807: superfluid response persists above $T_c$, as revealed by terahertz
2808: measurements of the electrodynamic response \cite{JOE}.  These data
2809: have been successfully modeled assuming the superconducting transition is of
2810: the Kosterlitz-Thouless type.
2811: 
2812: All of these behaviors would be difficult to attribute to phonons, in 
2813: that the latter are not gapped in the superconducting state.  There is 
2814: a drop in the electron-phonon scattering rate, since the electrons 
2815: themselves are gapped, but the effect is not as dramatic as in the 
2816: electron-electron case.
2817: \begin{figure}
2818: \centerline{\epsfxsize=0.4\textwidth{\epsfbox{fig48.eps}}}
2819: \caption{Kubo bubble for optical conductivity.  Dots represents 
2820: current vertices.  The attached part at the top is the lowest order 
2821: self-energy insertion to the bubble due to electron-electron scattering.}
2822: \label{fig48}
2823: \end{figure}
2824: A Feynman diagram analysis of the 
2825: Kubo bubble (Fig.~\ref{fig48}) would give a scattering rate drop setting in below 4$\Delta$
2826: (analogous to the 3$\Delta$ gap in the single particle scattering 
2827: rate (Fig.~\ref{fig5}) \cite{NOZ}) for the electron-electron scattering case (in the electron-phonon case,
2828: the self-energy insertion in Fig.~\ref{fig48} would be replaced by a phonon).
2829: The actual optics data, though, show an abrupt 
2830: onset of the drop at a frequency smaller than this.  This implies 
2831: collective effects, and in fact several researchers 
2832: \cite{MUNZAR,JULES,AC3} 
2833: have used the same explanation for the optics data as invoked previously 
2834: to explain the ARPES data \cite{NORM97,NDING}.  In particular, the 
2835: scattering rate onset corresponds to $2\Delta + \Omega_{res}$, 
2836: where $\Omega_{res}$ is the spin resonance energy
2837: seen by inelastic neutron scattering.
2838: 
2839: The c-axis response, though, is quite different (see Fig.~\ref{fig7}).
2840: Most studies have 
2841: been done on YBCO, since the electronic part of the c-axis 
2842: conductivity is small, and non-trivial to separate from the phonons 
2843: present in the data, particularly for Bi2212 which is more 2D like.  The normal state 
2844: response is non-Drude like in nature, $Re\sigma$ being more or less flat in 
2845: frequency \cite{HOMES}.  A hard gap begins to open up in this spectrum 
2846: below $T_{c}$ at a
2847: frequency near $2\Delta$ \cite{HOMES}.  Beyond this energy, a peak is present 
2848: in the data whose origin is controversial.  It has been identified as 
2849: the optic plasmon of the bilayer \cite{DIRK}, but also appears to be related to 
2850: the neutron resonance as well \cite{TOM}.
2851: 
2852: Recent developments in optics have focused on the issue of the 
2853: condensation energy, which brings us to the topic of the next section.
2854: 
2855: \section{Condensation Energy}
2856: 
2857: The condensation energy is defined as the energy difference between 
2858: the normal and superconducting states at T=0 (at finite temperature, 
2859: the entropy must be taken into account).  Since normal state properties 
2860: are temperature dependent, then this requires an extrapolation 
2861: to infer a hypothetical zero temperature normal state.  Most 
2862: estimates of the condensation energy have been made from specific heat 
2863: data, which have been recently used to suggest a quantum critical point 
2864: just beyond optimal doping based in part on the doping dependence of 
2865: the inferred condensation energy \cite{LORAM}, which has a maximum near
2866: the suggested critical point (middle panel, Fig.~\ref{fig9}).
2867: 
2868: We begin our story with a piece of pre-BCS history.  In 1956, Chester 
2869: published a paper demonstrating where the condensation energy was 
2870: coming from in conventional superconductors \cite{CHESTER}.  What he 
2871: chose to study was the full Hamiltonian of the solid, which is 
2872: composed of the kinetic energies of the electrons and ions,
2873: and the electron-electron, electron-ion, and ion-ion interactions.  
2874: Exploiting the dependence of $T_{c}$ on ion mass (isotope effect)
2875: and the virial 
2876: theorem, he was able to demonstrate that: (1) the potential energy of 
2877: the electrons is reduced in the superconducting state, (2) the kinetic 
2878: energy of the electrons is increased, and (3) the ion kinetic energy is 
2879: reduced.  For the classic value of the isotope coefficient (1/2), the 
2880: potential and kinetic energy changes of the electrons actually 
2881: cancel, which means that the entire condensation energy is 
2882: equivalent to the lowering of the ion kinetic energy.
2883: 
2884: This is not what occurs in BCS theory, of course.  The reason is that 
2885: it is an effective theory gotten by projecting the full 
2886: Hamiltonian onto the low energy subspace.  In such a theory, the ion 
2887: kinetic energy term is absorbed into the definition of the potential 
2888: energy, and so superconductivity is due to a lowering of the 
2889: effective potential energy of the low energy electrons.  Note that 
2890: the virial theorem does not apply to the projected Hamiltonian, and so 
2891: cannot be used as a guide.
2892: 
2893: As for the kinetic energy, in BCS theory, the normal state is taken to 
2894: be that of non-interacting electrons.  Because of this, the effective 
2895: kinetic energy of the electrons is exactly diagonalized.  That is, 
2896: the momentum distribution function is equal to 1 for $k < k_{F}$ and 
2897: 0 for $k > k_{F}$.  In the superconducting state, particle-hole 
2898: mixing occurs, leading to $n_{k} = v_{k}^{2} = \frac{1}{2}
2899: (1-\epsilon_{k}/\sqrt{\epsilon_{k}^{2}+\Delta_{k}^{2}})$ (see 
2900: Fig.~\ref{fig12}).  So, the 
2901: ``smearing'' of $n_{k}$ (Fig.~\ref{fig49}) leads to an increase of the kinetic energy.
2902: 
2903: \begin{figure}
2904: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig49.eps}}}
2905: \caption{Model calculations of the momentum distribution function 
2906: in the normal state (NS) and 
2907: the superconducting state (SC) compared to a hypothetical Fermi 
2908: liquid normal state (FL).  From Ref.~\cite{GANG4}.}
2909: \label{fig49}
2910: \end{figure}
2911: 
2912: The potential and kinetic energy differences are straightforward to 
2913: calculate in BCS theory, and this exercise can be found in Tinkham's book 
2914: \cite{TINKHAM}.  Although each term is ultraviolet divergent (and 
2915: thus cut-off at the Debye energy), the sum is convergent, equal to 
2916: $\frac{1}{2} N \Delta^{2}$ where $N$ is the density of states.
2917: Of course, the normal state of the cuprates is not a Fermi liquid (at 
2918: least for optimal and underdoped samples), so BCS considerations 
2919: could be misleading.
2920: 
2921: In Section 4, we discussed one attempt to get at a piece of the 
2922: condensation energy.  It was noted by Scalapino and White \cite{WHITE} 
2923: that the difference of the exchange energy between the normal and 
2924: superconducting states could be obtained from an integral involving 
2925: the difference of $Im\chi$ between the normal and 
2926: superconducting states.  Demler and Zhang \cite{DZNAT} noted that for 
2927: optimal doped YBCO, the normal state $Im\chi$ is small, and so 
2928: an estimate could be made in that case by simply integrating the 
2929: neutron resonance (see Figs.~\ref{fig41} and \ref{fig42}).
2930: The resulting energy change (the exchange energy 
2931: is lowered in the superconducting state) is in excess of the known 
2932: condensation energy, and they suggested this was compensated by an 
2933: expected increase in the kinetic energy in the superconducting state.
2934: Although the latter is true in BCS theory, is it necessarily true in 
2935: general?  Recent optics data give evidence to the contrary, as we now
2936: discuss.
2937: 
2938: \subsection{Optics}
2939: 
2940: Anderson has noted that several unusual properties of the cuprates 
2941: could be understood if spectral weight from high energies were 
2942: transfered to low energies when going from the normal to the 
2943: superconducting state \cite{PWA90,ANDERSON}.  Under such conditions, one 
2944: might anticipate the kinetic energy of the electrons to actually be 
2945: lowered in the superconducting state rather than raised as in BCS 
2946: theory.  This was also suggested by Hirsch, who predicted a 
2947: ``violation'' of the optical sum rule due to this effect \cite{HIRSCH}.
2948: That is, in a conventional superconductor, the weight appearing in the 
2949: zero frequency condensate peak comes entirely from the weight removed 
2950: by the 
2951: $2\Delta$ gap in the optical response (a result known as the 
2952: Tinkham-Ferrell-Glover sum rule \cite{TINKHAM}).  But a ``violation'' 
2953: would indicate that more weight was present in the condensate peak 
2954: than expected (Fig.~\ref{fig50}) \cite{KLEIN}.
2955: 
2956: \begin{figure}
2957: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig50.eps}}}
2958: \caption{Top panel:  schematic of optical conductivity (real part) in the 
2959: normal state (A is low energy, B high energy).  Middle panel:  
2960: Comparison of normal state to superconducting state if the single band 
2961: sum rule is preserved.  Gapped weight appears as a $\delta$ function 
2962: at zero energy.  Bottom panel:  When a sum rule ``violation'' is present, 
2963: more weight appears in the $\delta$ function than is gapped out.  
2964: This implies that the remaining weight must come from high energies to 
2965: satisfy the full sum rule.  From Ref.~\cite{KLEIN}.}
2966: \label{fig50}
2967: \end{figure}
2968: 
2969: To understand this, it is necessary to find a relation between the 
2970: optics and kinetic energy.  The optical sum rule of interest is the 
2971: same one discussed in Section 5
2972: \begin{equation}
2973: \int_{0}^{\infty} d\omega Re\sigma_{jj}(\omega) = \omega_{pl}^{2}/8
2974: \end{equation}
2975: This integral is well known when integrated over all energy bands, and 
2976: is simply proportional to the bare 
2977: carrier density, $n$, over the bare electron mass, $m$.
2978: By charge conservation ($n$ fixed), 
2979: this integral is always conserved.
2980: 
2981: On the other hand, what is of 
2982: interest here is the ``single band'' sum rule
2983: \begin{equation}
2984: {\hat P} \int_{0}^{\infty} d\omega Re\sigma_{jj}(\omega) =
2985: \frac{\pi e^{2} a^{2}}{2\hbar^{2}V} E_{K}
2986: \end{equation}
2987: where
2988: \begin{equation}
2989: E_{K} = \frac{2}{a^{2}N} \sum_{k}
2990: \frac{\partial^{2} \epsilon_{k}}{\partial k_{j}^{2}} n_{k}
2991: \end{equation}
2992: and ${\hat P}$ projects onto the single band subspace.  In these 
2993: expressions,
2994: $V$ is the unit cell volume, $a$ the in-plane lattice constant,
2995: $N$ the number of $k$ vectors, $\epsilon_{k}$ the dispersion 
2996: defined from the kinetic energy part of the effective single band 
2997: Hamiltonian, and $n_{k}$ the momentum distribution function.  This was 
2998: first derived by Kubo \cite{KUBO}.   On the other hand, the kinetic 
2999: energy for this band is
3000: \begin{equation}
3001: E_{kin} = \frac{2}{N} \sum_{k} \epsilon_{k} n_{k}
3002: \end{equation}
3003: Thus, the optical integral is of similar form, but not identical, to 
3004: the kinetic energy \cite{MNCP}, except for a near neighbor tight binding 
3005: form for $\epsilon_{k}$, where $E_{K} = -E_{kin}$.
3006: In practice, the optical integral must 
3007: be cut off at some energy so that interband terms are not included 
3008: (typically 1 eV in the cuprates).
3009: 
3010: There have been a number of studies which show anomalous changes in 
3011: the optical response between normal and superconducting states at
3012: energies beyond 1 eV in the cuprates \cite{FUGOL,LITTLE,RUBHAUSEN}.
3013: But matters came to the forefront 
3014: when Basov and co-workers demonstrated an explicit violation of the 
3015: single band sum rule for the c-axis optical response of underdoped 
3016: cuprates \cite{BASOV}.  Since a near neighbor tight binding model 
3017: should be adequate to describe the hopping along the c-axis, this 
3018: finding gives direct evidence that the c-axis kinetic energy is 
3019: lowered in the superconducting state, as suggested by Anderson and 
3020: co-workers \cite{INTLAY,ANDERSON} and Hirsch as well
3021: \cite{HIRSCH}.  There has been an alternate interpretation of this 
3022: observation, though, from 
3023: Ioffe and Millis \cite{LEV}.  They claim that a ``violation'' is 
3024: possible for the c-axis response if the normal state reference is 
3025: taken to be a pseudogap phase involving pairs without long range 
3026: phase order.
3027: 
3028: Regardless, the c-axis kinetic energy is 
3029: so small, the energy savings inferred is far below what is needed to 
3030: account for the actual condensation energy of the cuprates, which is
3031: about 3K per CuO plane in optimal doped YBCO \cite{LORSH}.
3032: On the other hand, the in-plane kinetic energy is quite large, of 
3033: order 1 eV.  Therefore, if a similar relative violation of the size seen 
3034: for the c-axis occurs for the in-plane response, then the energy 
3035: savings could be enough to account for the condensation energy.
3036: 
3037: Motivated by this, two recent experiments on optimal and underdoped 
3038: Bi2212 found evidence for a change in the planar optical integral between 
3039: normal and superconducting states large enough to account for the 
3040: condensation energy (Fig.~\ref{fig51}) \cite{DVMSC,NICOLE} (though an earlier study of
3041: underdoped YBCO did not \cite{JOE2}).
3042: \begin{figure}
3043: \centerline{\epsfxsize=0.5\textwidth{\epsfbox{fig51a.eps}}
3044: \epsfxsize=0.5\textwidth{\epsfbox{fig51b.eps}}}
3045: \caption{Left:  Variation of the optical spectral weight (integrated to 1.25 eV)
3046: with temperature
3047: for an optimal (top left) and an underdoped (bottom left)
3048: sample of Bi2212.  Anomalous rise below $T_c$
3049: implies kinetic energy lowering in the 
3050: superconducting state.  From 
3051: Ref.\cite{DVMSC}.  Right:  Optical spectral weight difference between a 
3052: temperature above $T_{c}$ and 10K integrated to the energy 
3053: plotted on the x axis normalized to the weight of the superconducting 
3054: condensate for various samples of Bi2212 (diamonds for overdoped, 
3055: triangles for optimal doped, and circles for underdoped). Spectral 
3056: weight balance (sum rule) would correspond to a value of 1.
3057: From Ref.~\cite{NICOLE}.}
3058: \label{fig51}
3059: \end{figure}
3060: No such violation was found 
3061: for an overdoped Bi2212 sample \cite{NICOLE}.  In both experiments, more 
3062: spectral weight showed up in the zero frequency condensate peak than can 
3063: be accounted for by the loss of finite frequency weight up to about 1 eV.
3064: Integrating up 
3065: to about 2 eV, though, spectral weight balance is found.  This confirms 
3066: the earlier speculations by Anderson \cite{PWA90,ANDERSON} of transfer of 
3067: spectral weight from high to low energies.
3068: 
3069: Using the observed form of
3070: the scattering rate in the normal state,
3071: $a_{k} + b\omega$ \cite{AV}, and assuming both of these terms are 
3072: gapped below some threshold energy in the superconducting state
3073: (left panel, Fig.~\ref{fig52}),
3074: the optical integral 
3075: difference was calculated by Norman and P\'{e}pin \cite{MNCP} (right
3076: panel, Fig.~\ref{fig52}) from 
3077: an $\epsilon_{k}$ extracted from ARPES data (more properly, band structure
3078: values should be used for $\epsilon_k$).  They 
3079: find that such a calculation gives a good estimate of the optical 
3080: integral change, including its doping dependence.
3081: \begin{figure}
3082: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig52.eps}}}
3083: \caption{Left: $1/\tau$ vs $\omega$ in the superconducting state
3084: as extracted from optics for an overdoped (OD70)
3085: and an underdoped (UD67) sample of Bi2212 \cite{PUCH}.  Dotted lines are 
3086: $a+b\omega$ fits to normal state data, arrows the locations of the scattering rate
3087: gaps.  Note the near zero value of $a$ in the OD case,
3088: as contrasted to the large value in the UD case.
3089: Right: Calculated change in the kinetic energy (open circles) and 
3090: optical integral (full circles) versus hole doping, x, with parameters determined
3091: from optics (left panel)  and ARPES.
3092: The open squares are the data of 
3093: Ref.~\cite{DVMSC} and the full squares that of Ref.~\cite{NICOLE}.  
3094: Adapted from Ref.~\cite{MNCP}.}
3095: \label{fig52}
3096: \end{figure}
3097: In these 
3098: calculations, the sum rule violation is coming from the $a_{k}$ term (that 
3099: is, if the normal state were a pure marginal Fermi liquid, there would 
3100: be no sum rule violation).  This $a_k$ term, which as stated earlier seems to
3101: be associated with the pseudogap, has a strong dependence on
3102: doping (left panel, Fig.~\ref{fig52}), which explains the large doping
3103: dependence of the sum rule violation.
3104: The same calculations find that the 
3105: actual kinetic energy change is about twice that indicated by the 
3106: optical integral, the difference due to the fact that the 
3107: inverse mass tensor is not simply the negative of $\epsilon_{k}$ as in a 
3108: near neighbor tight binding model of the energy dispersion.  In summary, the 
3109: origin of the sum rule ``violation'' and kinetic energy lowering
3110: can be traced to the formation of 
3111: quasiparticle peaks in the superconducting state due to the opening of 
3112: a scattering rate gap.
3113: 
3114: \subsection{ARPES}
3115: 
3116: These results can be generalized by considering the entire free 
3117: energy.  If only two particle interactions are involved, it is easily 
3118: shown that the full condensation energy can be written as \cite{GANG4}
3119: \begin{equation}
3120: E_{cond} = \sum_{k} \int_{-\infty}^{\infty} d\omega (\omega + 
3121: \epsilon_{k}) f(\omega) [A_{N}(k,\omega)-A_{S}(k,\omega)]
3122: \end{equation}
3123: where $A_{N}$ is the normal state single particle spectral function, 
3124: and $A_{S}$ that of the superconducting state.  This expression is 
3125: the sum of two terms, a kinetic energy term (where $\omega+\epsilon_{k}$ 
3126: is replaced by $2\epsilon_{k}$) and a potential energy term (where 
3127: $\omega+\epsilon_{k}$ is replaced by $\omega-\epsilon_{k}$).
3128: 
3129: This expression can be reduced further by performing some of the 
3130: integrals and sums
3131: \begin{equation}
3132: E_{cond} = \sum_{k} \epsilon_{k} [n_{N}(k) - n_{S}(k)]
3133: +\int_{-\infty}^{\infty} d\omega \omega f(\omega) 
3134: [N_{N}(\omega)-N_{S}(\omega)]
3135: \end{equation}
3136: where $n(k)$ is the momentum distribution function and $N(\omega)$ the density 
3137: of states.  The first term is related to (but not the same as)
3138: the optical integral we just discussed, the second term 
3139: could be obtained from tunneling spectroscopy.  Both terms, though, 
3140: can in principle be obtained from angle resolved photoemission, since 
3141: only the occupied states enter.  In practice, resolution, 
3142: normalization, and matrix element effects will be a limiting factor 
3143: \cite{GANG4}.
3144: 
3145: The above expressions, though, represent a simple conceptual formalism 
3146: for tackling the condensation energy issue which avoids the problem 
3147: of considering complicated two particle correlation functions.  This 
3148: was illustrated by Norman \etal \cite{GANG4}, who evaluated these 
3149: expressions using the ``mode'' model of Norman and Ding 
3150: (Fig.~\ref{fig29}) \cite{NDING} 
3151: for fitting ARPES spectra (the same model was employed by Ioffe and 
3152: Millis \cite{LEV} to analyze the c-axis optical sum rule violation, 
3153: and Norman and P\'{e}pin \cite{MNCP} to analyze the in-plane one).  What 
3154: these authors found was that for small normal state scattering rates, 
3155: a result similar to BCS theory occurs, which presumably applies on 
3156: the overdoped side of the phase diagram.  For scattering rates much 
3157: larger than the superconducting energy gap, though, a result opposite 
3158: to BCS theory was found, in that the kinetic energy was lowered and 
3159: the potential energy raised in the superconducting state.
3160: 
3161: The kinetic energy result is rather straightforward to understand 
3162: (Fig.~\ref{fig49}).  
3163: There is the BCS effect of particle-hole mixing which raises the 
3164: kinetic energy.  Opposed to this is the formation of coherent 
3165: quasiparticle states from the incoherent normal state, which acts to 
3166: lower the kinetic energy.  For a small scattering rate, the 
3167: particle-hole mixing effect wins out, and the kinetic energy is 
3168: raised, but for larger scattering rate, the quasiparticle formation 
3169: effect wins out, and the kinetic energy is lowered.
3170: 
3171: The potential energy part is also straightforward to understand (Fig.~\ref{fig53}).
3172: \begin{figure}
3173: \centerline{\epsfxsize=0.8\textwidth{\epsfbox{fig53.eps}}}
3174: \caption{Left panel:  calculated spectral function in the normal state (NS) 
3175: and superconducting state (SC).  Right panel:  resulting difference in 
3176: first moments of the spectrum
3177: versus cut-off energy.  Note positive contribution (potential energy 
3178: decrease in the SC state) from spectral peak, negative contribution 
3179: (potential energy increase in the SC state) from the difference in the 
3180: spectral tails.
3181: From Ref.~\cite{GANG4}.}
3182: \label{fig53}
3183: \end{figure}
3184: There is a competition between the formation of a spectral gap (which 
3185: lowers the potential energy as in BCS theory) and spectral weight 
3186: transfer from high to low energy to form the coherent peak 
3187: (which raises the potential energy).  For small scattering 
3188: rate, the spectral gap effect wins out and the potential energy is 
3189: lowered, whereas for large scattering rate, the weight transfer 
3190: effect wins out and the potential energy is raised.
3191: 
3192: What this implies for the phase diagram is that on the overdoped side, 
3193: one expects more or less BCS like physics.  But on the underdoped 
3194: side, one expects dramatic departures.  In particular, the potential 
3195: energy is lowered across the pseudogap $T^{*}$ line due to the formation 
3196: of a spectral gap, then the kinetic energy lowered across the 
3197: superconducting $T_{c}$ line due to the onset of coherence.
3198: 
3199: We would like to end with a ``cautionary'' remark.  What one calls 
3200: kinetic or potential energy depends on what effective Hamiltonian is 
3201: being employed.  The definition of this changes, for instance, if one 
3202: goes from the three band Hubbard model, to the single band Hubbard 
3203: model, to the t-J model.  The purpose of going through the above 
3204: exercise is to demonstrate that one's preconceptions based on BCS 
3205: theory could well 
3206: be wrong in pairing models driven by electron-electron interactions, 
3207: particularly if the normal state reference is non Fermi liquid like.
3208: 
3209: On that note, we would like to bring this review to a close.
3210: 
3211: \ack
3212: 
3213: This work was supported by the U.S. Dept.~of Energy, Office of Science, under
3214: contract W-31-109-ENG-38.  This review is based on a series of lectures given
3215: at the SPhT in the fall of 2001 by MRN.  We would like to thank the
3216: staff of the SPhT, in particular Dr. J. P. Blaizot and Dr. G. Misguich, for
3217: making these lectures possible.
3218: 
3219: \section*{References}
3220: 
3221: \begin{thebibliography}{999}
3222: 
3223: \bibitem{DAHL}
3224: Dahl P F 1992 {\it Superconductivity} (New York: American Institute of 
3225: Physics)
3226: \bibitem{MATTHIAS}
3227: Matthias B T 1970 {\it Comments on Solid State Physics} {\bf 3} 93;
3228: 1969
3229: {\it Superconductivity} ed F Chilton (Amsterdam: North-Holland) p 69; 
3230: 1973
3231: {\it The Science and Technology of Superconductivity} ed W D Gregory, 
3232: W N Mathews Jr and E A Edelsack (New York: Plenum Press) p 263
3233: \bibitem{COHEN}
3234: Cohen M and Anderson P W 1972 {\it Superconductivity in d- and f-Band 
3235: Metals} ed D H Douglass (New York: American Institute of Physics) p 17
3236: \bibitem{TESTARDI}
3237: Testardi L R 1975 \RMP {\bf 47} 637
3238: \bibitem{FRANK}
3239: Steglich F, Aarts J, Bredl C D, Lieke W, Meschede D, Franz W and Schafer 
3240: H 1979 \PRL {\bf 43} 1892
3241: \bibitem{GREG}
3242: Stewart G R 1984 \RMP {\bf 56} 755
3243: \bibitem{MILLIS}
3244: Millis A J and Lee P A 1987 \PR B {\bf 35} 3394
3245: \bibitem{RMP91}
3246: Sigrist M and Ueda K 1991 \RMP {\bf 63} 239
3247: \bibitem{RMP02}
3248: Joynt R and Taillefer L 2002 \RMP {\bf 74} 235
3249: \bibitem{EMERY}
3250: Emery V J and Sessler A M 1960 \PR {\bf 119} 43
3251: \bibitem{FAY}
3252: Layzer A and Fay D 1971 {\it Intl. J. Magnetism} {\bf 1} 135
3253: \bibitem{3He}
3254: Vollhardt D and Wolfle P 1990 {\it The Superfluid Phases of Helium 
3255: 3} (London: Taylor and Francis)
3256: \bibitem{BA}
3257: Anderson P W and Brinkman W F 1973 \PRL {\bf 30} 1108
3258: \bibitem{AC}
3259: Abanov Ar, Chubukov A V and Schmalian J 2001 {\it J. Elec. Spec.} {\bf 117} 129
3260: \nonum Chubukov A V, Pines D and Schmalian J 2002 {\it Preprint} 
3261: cond-mat/0201140
3262: \nonum Abanov Ar, Chubukov A V and Schmalian J 2003 {\it Adv. Phys.} {\bf 52} 119
3263: \bibitem{GABE}
3264: Aeppli G, Goldman A, Shirane G, Bucher E and Lux-Steiner M-Ch 1987 
3265: \PRL {\bf 58} 808
3266: \bibitem{UGe2}
3267: Saxena S S, Agarwal P, Ahilan K, Grosche F M, Haselwimmer R K W, 
3268: Steiner M J, Pugh E, Walker I R, Julian S R, Mouthoux P, Lonzarich G 
3269: G, Huxley A, Sheikin I, Braithwaite D and Flouquet J 2000 {\it Nature} 
3270: {\bf 406} 587
3271: \bibitem{ZrZn2}
3272: Pfleiderer C, Uhlarz M, Hayden S M, Vollmer R, von Lohneysen H, 
3273: Bernhoeft N R and Lonzarich G G 2001 {\it Nature} {\bf 412} 58
3274: \bibitem{1986}
3275: Miyake K, Schmitt-Rink S and Varma C M 1986 \PR B {\bf 34} 6554
3276: \nonum Scalapino D J, Loh E Jr and Hirsch J E 1986 \PR B {\bf 34} 8190
3277: \bibitem{MIKE92}
3278: Norman M R 1992 {\it Physica} C {\bf 194} 203
3279: \bibitem{JIM94}
3280: Sauls J A 1994 {\it Adv. Phys.} {\bf 43} 113
3281: \bibitem{RMP88}
3282: Bednorz J G and Muller K A 1988 \RMP {\bf 60} 585
3283: \bibitem{EAGLES}
3284: Eagles D M 1969 \PR {\bf 186} 456
3285: \bibitem{CHU}
3286: Wu M K, Ashburn J R, Torng C J, Hor P H, Meng R L, Goa L, Huang Z J, 
3287: Wang Y Q and Chu C W 1987 \PRL {\bf 58} 908
3288: \bibitem{BELL}
3289: Feder T 2000 {\it Physics Today} {\bf 53} No 4 p 56
3290: \bibitem{HAZEN}
3291: Hazen R M 1988 {\it The Breakthrough} (New York: Summit Books)
3292: \bibitem{RVB}
3293: Anderson P W 1987 {\it Science} {\bf 235} 1196
3294: \bibitem{SAWATZKY}
3295: Sawatzky G A and Allen J W 1984 \PRL {\bf 53} 2339
3296: \bibitem{PICKETT}
3297: Pickett W E 1989 \RMP {\bf 61} 433
3298: \bibitem{ANDERSON}
3299: Anderson P W 1997 {\it The Theory of Superconductivity in the High-$T_{c}$ 
3300: Cuprates} (Princeton: Princeton Univ. Pr.)
3301: \bibitem{HYBERTSEN}
3302: Hybertsen M S, Stechel E B, Schluter M and Jennison D R 1990 \PR B {\bf 
3303: 41} 11068
3304: \bibitem{VARMA}
3305: Varma C M 1997 \PR B {\bf 55} 14554
3306: \bibitem{SLATER}
3307: Slater J C 1951 \PR {\bf 82} 538
3308: \bibitem{INS}
3309: Vaknin D, Sinha S K, Moncton D E, Johnston D C, Newsam J M, Safinya C 
3310: R and King H E Jr 1987 \PRL {\bf 58} 2802
3311: \bibitem{KLMU}
3312: Kleiner R and Muller P 1994 \PR B {\bf 49} 1327
3313: \bibitem{JOE}
3314: Corson J, Mallozzi R, Orenstein J, Eckstein J N and Bozovic I 1999 
3315: {\it Nature} {\bf 398} 221
3316: \bibitem{TAKIGAWA}
3317: Takigawa M, Hammel P C, Heffner R H and Fisk Z 1989 \PR B {\bf 39} 
3318: 7371
3319: \bibitem{BICKERS}
3320: Bickers N E, Scalapino D J and Scalettar R T 1987 {\it Intl. J. Modern 
3321: Physics} B {\bf 1} 687
3322: \bibitem{KOTLIAR}
3323: Kotliar G 1988 \PR B {\bf 37} 3664
3324: \nonum Kotliar G and Liu J 1988 \PR B {\bf 38} 5142
3325: \bibitem{GROS}
3326: Zhang F C, Gros C, Rice T M and Shiba H 1988 {\it Supercond. Sci. 
3327: Technol.} {\bf 1} 36
3328: \bibitem{MARTINDALE}
3329: Martindale J A, Barrett S E, O'Hara K E, Slichter C P, Lee W C and 
3330: Ginsberg D M 1993 \PR B {\bf 47} 9155
3331: \bibitem{HARDY}
3332: Hardy W N, Bonn D A, Morgan D C, Liang R and Zhang K
3333: 1993 \PRL {\bf 70} 3999
3334: \bibitem{SHEN93}
3335: Shen Z-X, Dessau D S, Wells B O, King D M, Spicer W E, Arko A J,
3336: Marshall D, Lombardo L W, Kapitulnik A, Dickinson P, Doniach S,
3337: DiCarlo J, Loeser A G and Park C H 1993 \PRL {\bf 70} 1553
3338: \bibitem{DALE}
3339: Wollman D A, van Harlingen D J, Lee W C, Ginsberg D M and Leggett A J 
3340: 1993 \PRL {\bf 71} 2134
3341: \bibitem{TSUEI}
3342: Tsuei C C, Kirtley J R, Chi C C, Yu-Jahnes L S, Gupta A, Shaw T, Sun 
3343: J Z and Ketchen M B 1994 \PRL {\bf 73} 593
3344: \bibitem{TLPH}
3345: Tsuei C C, Kirtley J R, Rupp M, Sun J Z, Gupta A, Ketchen M B, Wang C 
3346: A, Ren Z F, Wang J H and Bhushan M 1996 {\it Science} {\bf 271} 329
3347: \bibitem{BONN}
3348: Bonn D A, Dosanjh P, Liang R and Hardy W N 1992 \PRL {\bf 68} 2390
3349: \bibitem{ONG}
3350: Krishana K, Harris J M and Ong N P 1995 \PRL {\bf 75} 3529
3351: \bibitem{PUCH}
3352: Puchkov A V, Basov D N and Timusk T 1996 \JPCM {\bf 8} 10049
3353: \bibitem{ADAM}
3354: Kaminski A, Mesot J, Fretwell H, Campuzano J C, Norman M R,
3355: Randeria M,
3356: Ding H, Sato T, Takahashi T, Mochiku T, Kadowaki K and Hoechst H 2000
3357: \PRL {\bf 84} 1788
3358: \bibitem{NOZ}
3359: Nozieres P 1964 {\it Theory of Interacting Fermi Systems} (Reading: 
3360: Addison-Wesley)
3361: \bibitem{RVB2}
3362: Anderson P W, Baskaran G, Zou Z and Hsu T 1987 \PRL {\bf 58} 2790
3363: \bibitem{BZA}
3364: Baskaran G, Zou Z and Anderson P W 1987 \SSC {\bf 63} 973
3365: \bibitem{LEE}
3366: Nagaosa N and Lee P A 1992 \PR B {\bf 45} 966
3367: \bibitem{Bi2201}
3368: Martin S, Fiory A T, Fleming R M, Schneemeyer L F and Waszczak J V 
3369: 1990 \PR B {\bf 41} 846
3370: \bibitem{MFL}
3371: Varma C M, Littlewood P B, Schmitt-Rink S, Abrahams E and Ruckentstein A E
3372: 1989 \PRL {\bf 63} 1996
3373: \bibitem{TIMRPP}
3374: Timusk T and Statt B 1999 \RPP {\bf 62} 61
3375: \bibitem{VARENNA}
3376: Randeria M 1999 {\it Proceedings of the International School of Physics ``Enrico
3377: Fermi'' Course CXXXVI on High Temperature Superconductors}
3378: ed G Iadonisi, J R Schrieffer and M L Chiafalo
3379: (Amsterdam: IOS Press) p 53 (cond-mat/9710223)
3380: \bibitem{WARREN}
3381: Warren W W Jr, Walstedt R E, Brennert G F, Cava R J, Tycko R, Bell R 
3382: and Dabbagh G 1989 \PRL {\bf 62} 1193
3383: \bibitem{ALLOUL}
3384: Alloul H, Ohno T and Mendels P 1989 \PRL {\bf 63} 1700
3385: \bibitem{COOPER}
3386: Cooper S L, Thomas G A, Orenstein J, Rapkine D H, Capizzi M, Timusk T, 
3387: Millis A J, Schneemeyer L F and Waszczak J V 1989 \PR B {\bf 40} 11358
3388: \bibitem{ROTTER}
3389: Rotter L D, Schlesinger Z, Collins R T, Holtzberg F, Field C, Welp U 
3390: W, Crabtree G W, Liu J Z, Fang Y, Vandervoort K G and Fleshler S 1991 
3391: \PRL {\bf 67} 2741
3392: \bibitem{TAKI2}
3393: Takigawa M, Reyes A P, Hammel P C, Thompson J D, Heffner R H, Fisk Z 
3394: and Ott K C 1989 \PR B {\bf 43} 247
3395: \bibitem{RESA}
3396: Bucher B, Steiner P, Karpinksi J, Kaldis E and Wachter P 1993 \PRL 
3397: {\bf 70} 2012
3398: \nonum Ito T, Takenaka K and Uchida S 1993 \PRL {\bf 70} 3995
3399: \bibitem{ISHIDA}
3400: Ishida K, Yoshida K, Mito T, Tokunaga Y, Kitaoka Y, Asayama K, 
3401: Nakayama Y, Shimoyama J and Kishio K 1998 \PR B {\bf 58} 5960
3402: \bibitem{HOMES}
3403: Homes C C, Timusk T, Liang R, Bonn D A and Hardy W N 1993 \PRL {\bf 71} 
3404: 1645; 1995 {\it Physica} C {\bf 254} 265
3405: \bibitem{RESC}
3406: Takenaka K, Mizuhashi K, Takagi H and Uchida S 1994 \PR B {\bf 50} 6534
3407: \bibitem{MARSHALL}
3408: Marshall D S, Dessau D S, Loeser A G, Park C H, Matsuura A V, 
3409: Eckstein J N, Bozovic I, Fournier P, Kapitulnik A, Spicer W E
3410: and Shen Z-X 1996 \PRL {\bf 76} 4841
3411: \bibitem{LOESER}
3412: Loeser A G, Shen Z-X, Dessau D S, Marshall D S, Park C H,
3413: Fournier P and Kapitulnik A 1996 {\it Science} {\bf 273} 325
3414: \bibitem{DING}
3415: Ding H, Yokoya T, Campuzano J C, Takahashi T, Randeria M, 
3416: Norman M R, Mochiku T, Kadowaki K and Giapintzakis J 1996
3417: {\it Nature} {\bf 382} 51
3418: \bibitem{MOHIT92}
3419: Randeria M, Trivedi N, Moreo A and Scalettar R T 1992 \PRL {\bf 69} 
3420: 2001
3421: \bibitem{DING97}
3422: Ding H, Norman M R, Yokoya T, Takuechi T, Randeria M, 
3423: Campuzano J C, Takahashi T, Mochiku T and Kadowaki K 1997
3424: \PRL {\bf 78} 2628
3425: \bibitem{FISCHER}
3426: Renner Ch, Revaz B, Genoud J-Y, Kadowaki K and Fischer O 1998 \PRL 
3427: {\bf 80} 149
3428: \bibitem{NAT98}
3429: Norman M R, Ding H, Randeria M, Campuzano J C,
3430: Yokoya T, Takeuchi T, Takahashi T, Mochiku T,
3431: Kadowaki K, Guptasarma P and Hinks D G 1998 {\it Nature} {\bf 392} 157
3432: \bibitem{YOSHIDA}
3433: Yoshida T, Zhou X J, Sasagawa T, Yang W L, Bogdanov P V, Lanzara A, 
3434: Hussain Z, Mizokawa T, Fujimori A, Eisaki H, Shen Z-X, Kakeshita T 
3435: and Uchida S 2002 {\it Preprint} cond-mat/0206469
3436: \bibitem{LEESF}
3437: Affleck I and Marston J B 1988 \PR B {\bf 37} 3774
3438: \bibitem{SCHULZ}
3439: Schulz H J 1989 \PR B {\bf 39} 2940
3440: \bibitem{NAYAK}
3441: Chakravarty S, Laughlin R B, Morr D K and Nayak C 2001 \PR B {\bf 63} 
3442: 094503
3443: \bibitem{MOOK}
3444: Mook H A, Dai P and Dogan F 2001 \PR B {\bf 64} 012502
3445: \bibitem{NAT02}
3446: Kaminski A, Rosenkranz S, Fretwell H, Campuzano J C, Li Z, Raffy H,
3447: Cullen W G, You H, Olson C G, Varma C M and Hoechst H 2002 {\it Nature} 
3448: {\bf 416} 610
3449: \bibitem{VARMA02}
3450: Simon M E and Varma C M 2002 \PRL {\bf 89} 247003
3451: \bibitem{LORAM}
3452: Tallon J L, Loram J W, Williams G V M, Cooper J R, Fisher I R, Johnson 
3453: J D, Staines M P and Bernhard C 1999 {\it Phys. Stat. Sol.} (b) {\bf 215} 531
3454: \bibitem{EMERY95}
3455: Emery V and Kivelson S 1995 {\it Nature} {\bf 374} 434
3456: \bibitem{NERNST}
3457: Xu Z A, Ong N P, Wang Y, Kakeshita T and Uchida S 2000 {\it Nature} 
3458: {\bf 406} 486
3459: \bibitem{CORE}
3460: Renner Ch, Revaz B, Kadowaki K, Maggio-Aprile I and Fischer O 1998 
3461: \PRL {\bf 80} 3606
3462: \bibitem{WELLS}
3463: Wells B O, Shen Z-X, Matsuura A, King D M,
3464: Kastner M A, Greven M and Birgeneau R J 1995 \PRL {\bf 74} 964
3465: \bibitem{NADOP}
3466: Kohsaka Y, Sasagawa T, Ronning F, Yoshida T, Kim C, Hanaguri T, Azuma 
3467: M, Takano M, Shen Z-X, and Takagi H 2002 {\it Preprint} cond-mat/0209339
3468: \bibitem{RMP03}
3469: Damascelli A, Shen Z-X and Hussain Z 2002 {\it Preprint} 
3470: cond-mat/0208504 (\RMP Apr. 2003)
3471: \bibitem{ZAANEN}
3472: Zaanen J and Gunnarsson O 1989 \PR B {\bf 40} 7391
3473: \bibitem{EMKIV}
3474: Carlson E W, Emery V J, Kivelson S A and Orgad D 2002 {\it Preprint} 
3475: cond-mat/0206217
3476: \bibitem{JOHN95}
3477: Tranquada J M, Sternlieb B J, Axe J D, Nakamura Y and Uchida S 1995 
3478: {\it Nature} {\bf 375} 561
3479: \bibitem{AXE}
3480: Axe J D, Moudden A H, Hohlwein D, Cox D E, Mohanty K M, Moodenbaugh A 
3481: R and Xu Y 1989 \PRL {\bf 62} 2751
3482: \bibitem{SHENND}
3483: Zhou X J, Bogdanov P, Kellar S A, Noda T, Eisaki H, Uchida S, Hussain 
3484: Z and Shen Z-X 1999 {\it Science} {\bf 286} 268
3485: \bibitem{ANDO}
3486: Noda T, Eisaki H and Uchida S 1999 {\it Science} {\bf 286} 265
3487: \bibitem{NS-LSCO}
3488: Cheong S-W, Aeppli G, Mason T E, Mook H, Hayden S M, Canfield P C, 
3489: Fisk Z, Clausen K N and Martinez J L 1991 \PRL {\bf 67} 1791
3490: \bibitem{NS-YBCO}
3491: Mook H A, Dai P, Hayden S M, Aeppli G, Perring T G and Dogan F 1998 
3492: {\it Nature} {\bf 395} 580
3493: \bibitem{MOOK02}
3494: Mook H A, Dai P and Dogan F 2002 \PRL {\bf 88} 097004
3495: \bibitem{NORM00}
3496: Norman M R 2000 \PR B {\bf 61} 14751
3497: \bibitem{PAN}
3498: Pan S H, O'Neal J P, Badzey R L, Chamon C, Ding H, Engelbrecht J R,
3499: Wang Z, Eisaki H, Uchida S, Gupta A K, Ng K-W, Hudson E W, Lang K M 
3500: and Davis J C 2001 {\it Nature} {\bf 413} 282
3501: \bibitem{LANG}
3502: Lang K M, Madhavan V, Hoffman J E, Hudson E W, Eisaki H, Uchida S and 
3503: Davis J C 2002 {\it Nature} {\bf 415} 412
3504: \bibitem{CREN}
3505: Cren T, Roditchev D, Sacks W, Klein J, Moussy J-B, Deville-Cavellin C 
3506: and Lagues M 2000 \PRL {\bf 84} 147
3507: \bibitem{HOFFMAN}
3508: Hoffman J E, Hudson E W, Lang K M, Madhavan V, Eisaki H, Uchida S and
3509: Davis J C 2002 {\t Science} {\bf 295} 466
3510: \bibitem{HOFF2}
3511: Hoffman J E, McElroy K, Lee D-H, Lang K M, Eisaki H, Uchida S and 
3512: Davis J C 2002 {\it Science} {\bf 297} 1148
3513: \bibitem{HOWALD}
3514: Howald C, Eisaki H, Kaneko N, Greven M and Kapitulnik A 2003 \PR B {\bf 67} 
3515: 014533 (2003)
3516: \bibitem{MCELROY}
3517: McElroy K, Simmonds, R W, Hoffman J E, Lee D-H, Orenstein J, Eisaki 
3518: H, Uchida S and Davis J C 2002 {\it Preprint} ({\it Nature} Feb. 2003)
3519: \bibitem{TOKURA}
3520: Tokura Y, Takagi H and Uchida S 1989 {\it Nature} {\bf 337} 345
3521: \bibitem{LUKE90}
3522: Luke G M \etal 1990 \PR B {\bf 42} 7981
3523: \bibitem{ONOSE}
3524: Onose Y, Taguchi Y, Ishizaka K and Tokura Y 2001 \PRL {\bf 87} 217001
3525: \bibitem{WU93}
3526: Wu D H, Mao J, Mao S N, Peng J L, Xi X X, Venkatesan T, Greene R L 
3527: and Anlage S M 1993 \PRL {\bf 70} 85
3528: \bibitem{JOHN90}
3529: Huang Q, Zasadzinski J F, Tralshawala N, Gray K E, Hinks D G, Peng J L 
3530: and Greene R L 1990 {\it Nature} {\bf 347} 369
3531: \bibitem{NCCO}
3532: Kokales D K, Fournier P, Mercaldo L V, Talanov V V, Greene R L and 
3533: Anlage S M 2000 \PRL {\bf 85} 3696
3534: \nonum Prozorov R, Giannetta R W, Fournier P and Greene R L 2000 \PRL {\bf 
3535: 85} 3700
3536: \bibitem{SATO}
3537: Sato T, Kamiyama T, Takahashi T, Kurahashi K and Yamada K 2001 {\it 
3538: Science} {\bf 291} 1517
3539: \bibitem{PETER1}
3540: Armitage N P, Lu D H, Feng D L, Kim C, Damascelli A, Shen K M, 
3541: Ronning F, Shen Z-X, Onose Y, Taguchi Y and Tokura Y 2001 \PRL {\bf 86} 1126
3542: \bibitem{TSUEI2}
3543: Tsuei C C and Kirtley J R 2000 \PRL {\bf 85} 182
3544: \bibitem{GIRSH}
3545: Blumberg G, Koitzsch A, Gozar A, Dennis B S, Kendziora C A, Fournier P 
3546: and Greene R L 2002 \PRL {\bf 88} 107002
3547: \bibitem{PETER2}
3548: Armitage N P, Lu D H, Kim C, Damascelli A, Shen K M, Ronning F,
3549: Feng D L, Bogdanov P, Shen Z-X, Onose Y, Taguchi Y, Tokura Y,
3550: Mang P K, Kaneko N and Greven M 2001 \PRL {\bf 87} 147003
3551: \bibitem{PETER3}
3552: Armitage N P, Ronning F, Lu D H, Kim C, Damascelli A, Shen K M,
3553: Feng D L, Eisaki H, Shen Z-X, Mang P K, Kaneko N, Greven M,
3554: Onose Y, Taguchi Y and Tokura Y 2002 \PRL {\bf 88} 257001
3555: \bibitem{BCS}
3556: Bardeen J, Cooper L N and Schrieffer J R 1957 \PR {\bf 108} 1175
3557: \bibitem{SUPER}
3558: Bardeen J 1963 {\it 8th Intl. Conf. on Low Temp.} ed R O Davies 
3559: (London: Butterworths) p 3
3560: \nonum Cooper L N 1987 {\it IEEE Trans. on Magnetics} {\bf MAG-23} 376
3561: \bibitem{BOB}
3562: Schrieffer J R 1964 {\it Theory of Superconductivity} (Reading: 
3563: Bejamin/Cummings)
3564: \bibitem{DOUG}
3565: Scalapino D J 1994 {\it Random Magnetism, High Temperature 
3566: Superconductivity} ed W P Beyermann, N L Huang-Liu and D E 
3567: MacLaughlin (Singapore: World Scientific) p 155
3568: \bibitem{HL}
3569: Hertz J A, Levin K and Beal-Monod M T 1976 {\it Solid State Comm.} 
3570: {\bf 18} 803
3571: \bibitem{ADV}
3572: Anderson PW 1997 {\it Adv. Phys.} {\bf 46} 3
3573: \bibitem{SU2}
3574: Affleck I, Zou Z, Hsu T and Anderson P W 1988 \PR B {\bf 38} 745
3575: \bibitem{PARM}
3576: Paramekanti A, Randeria M and Trivedi N 2001 \PRL {\bf 87} 217002
3577: \bibitem{ORB}
3578: Ivanov D A, Lee P A and Wen X-G 2000 \PRL {\bf 84} 3958
3579: \bibitem{PLEE}
3580: Wen X-G and Lee P A 1996 \PRL {\bf 76} 503
3581: \nonum Lee P A, Nagaosa N, Ng T-K and Wen X-G 1998 \PR B {\bf 57} 6003
3582: \bibitem{COLD}
3583: Ioffe L B and Millis A J 1998 \PR B {\bf 58} 11631
3584: \bibitem{VALLA}
3585: Valla T, Fedorov A V, Johnson P D, Wells B O,
3586: Hulbert S L, Li Q, Gu G D and Koshizuka N 1999
3587: {\it Science} {\bf 285} 2110
3588: \bibitem{KURODA}
3589: Kuroda Y and Varma C M 1990 \PR B {\bf 42} 8619
3590: \bibitem{LITTLEW}
3591: Littlewood P B and Varma C M 1992 \PR B {\bf 46} 405
3592: \bibitem{NDING}
3593: Norman M R and Ding H 1998 \PR B {\bf 57} 11089
3594: \bibitem{ROSSAT}
3595: Rossat-Mignod J, Regnault L P, Vettier C, Bourges P, Burlet P, Bossy 
3596: J, Henry J Y and Lapertot G 1991 {\it Physica} C {\bf 185-189} 86
3597: \bibitem{MOOK93}
3598: Mook H A, Yethraj M, Aeppli G, Mason T E and Armstrong T 1993 \PRL 
3599: {\bf 70} 3490
3600: \bibitem{SCALAPINO}
3601: Bulut N and Scalapino D J 1996 \PR B {\bf 53} 5149
3602: \bibitem{DEMLER}
3603: Demler E and Zhang S-C 1995 \PRL {\bf 75} 4126
3604: \bibitem{FONG95}
3605: Fong H F, Keimer B, Anderson P W, Reznik D, Dogan F and Aksay I A 
3606: 1995 \PRL {\bf 75} 316
3607: \bibitem{FONG96}
3608: Fong H F, Keimer B, Reznik D, Milius D L and Aksay I A 
3609: 1996 \PR B {\bf 54} 6708
3610: \bibitem{GREITER}
3611: Greiter M 1997 \PRL {\bf 79} 4898
3612: \bibitem{OLEG}
3613: Tchernyshyov O, Norman M R and Chubukov A V 2001 \PR B {\bf 63} 144507
3614: \bibitem{ZHANG}
3615: Zhang S-C 1997 {\it Science} {\bf 275} 1089
3616: \bibitem{PROJ}
3617: Zhang S-C, Hu J-P, Arrigoni E, Hanke W and Auerbach A 1999 \PR B {\bf 60} 
3618: 13070
3619: \bibitem{HANKE}
3620: Zachar M G, Hanke W, Arrigoni E and Zhang S-C 2000 \PRL {\bf 85} 824
3621: \bibitem{LEVIN}
3622: Si Q, Zha Y, Levin K and Lu J P 1993 \PR B {\bf 47} 9055
3623: \bibitem{YAMADA}
3624: Fujita M, Yamada K, Hiraka H, Gehring P M, Lee S H, Wakimoto S and 
3625: Shirane G 2002 \PR B {\bf 65} 064505
3626: \bibitem{JUNOD}
3627: Junod A, Erb A and Renner C 1999 {\it Physica} C {\bf 317} 333
3628: \bibitem{BILBRO}
3629: Bilbro G and McMillan W L 1976 \PR B {\bf 14} 1887
3630: \bibitem{LAUGH}
3631: Laughlin R B 1998 {\it Adv. Phys.} {\bf 47} 943
3632: \bibitem{SUBIR}
3633: Sachdev S 2000 {\it Science} {\bf 288} 475
3634: \bibitem{HUFNER}
3635: H\"{u}fner S 1996 {\it Photoelectron Spectroscopy} 
3636: (Berlin: Springer-Verlag)
3637: \bibitem{MOHIT95}
3638: Randeria M, Ding H, Campuzano J C, Bellman A, Jennings G, Yokoya T,
3639: Takahashi T, Katayama-Yoshida H, Mochiku T and Kadowaki K 1995 \PRL {\bf 74}
3640: 4951
3641: \bibitem{REVIEW}
3642: Campuzano J C, Norman M R and Randeria M 2002 {\it Preprint} 
3643: cond-mat/0209476
3644: \bibitem{YOKOYA}
3645: Chainani A, Yokoya T, Kiss T and Shin S 2000 \PRL {\bf 85} 1966
3646: \bibitem{ADAM2}
3647: Kaminski A, Randeria M, Campuzano J C, Norman M R, Fretwell H,
3648: Mesot J, Sato T, Takahashi T and and Kadowaki K 2001
3649: \PRL {\bf 86} 1070
3650: \bibitem{JC90}
3651: Campuzano J C, Jennings G, Faiz M, Beaulaigue L, Veal B W, Liu J Z,
3652: Paulikas A P, Vandervoort K and Claus H 1990 \PRL {\bf 64} 2308
3653: \bibitem{JC99}
3654: Campuzano J C, Ding H, Norman M R, Fretwell H M, Randeria M,
3655: Kaminski A, Mesot J, Takeuchi T, Sato T, Yokoya T, Takahashi T,
3656: Kadowaki K, Guptasarma P, Hinks D G, Konstantinovic Z, Li Z Z and
3657: Raffy H 1999 \PRL {\bf 83} 3709
3658: \bibitem{FENG00}
3659: Feng D L, Lu D H, Shen K M, Kim C, Eisaki H, Damascelli A,
3660: Yoshizaki R, Shimoyama J-I, Kishio K, Gu G D, Oh S, Andrus A,
3661: O'Donell J, Eckstein J N and Shen Z-X 2000
3662: {\it Science} {\bf 289} 277
3663: \bibitem{DING01}
3664: Ding H, Engelbrecht J R, Wang Z, Campuzano J C, Wang S-C, Yang H-B, 
3665: Rogan R, Takahashi T, Kadowaki K and Hinks D G 2001
3666: \PRL {\bf 87} 227001
3667: \bibitem{ALEX}
3668: Abrikosov A A, Campuzano J C and Gofron K 1993 {\it Physica} C {\bf 214} 
3669: 73
3670: \bibitem{SATOB}
3671: Sato T, Kamiyama T, Takahashi T, Mesot J, Kaminski A, Campuzano J C,
3672: Fretwell H M, Takeuchi T, Ding H, Chong I, Terashima T and Takano M 
3673: 2001 \PR B {\bf 64} 054502
3674: \bibitem{INO}
3675: Ino A, Kim C, Nakamura M, Yoshida T, Mizokawa T, Shen Z-X,
3676: Fujimori A, Kakeshita T, Eisaki H and Uchida S 2002
3677: \PR B {\bf 65} 094504 (2002)
3678: \bibitem{HALL}
3679: Takagi H, Ido T, Ishibashi S, Uota M, Uchida S and Tokura Y 1989 \PR B 
3680: {\bf 40} 2254
3681: \bibitem{OLSON}
3682: Olson C G, Liu R, Lynch D W, List R S, Arko A J, Veal B W,
3683: Chang Y C, Jiang P Z and Paulikas A P 1990 \PR B {\bf 42} 381
3684: \bibitem{BOGDANOV}
3685: Bogdanov P V, Lanzara A, Zhou X J, Yang W L, Eisaki H, Hussain Z and 
3686: Shen Z-X 2002 \PRL {\bf 89} 167002
3687: \bibitem{INTLAY}
3688: Chakravarty S, Sudbo A, Anderson P W and Strong S 1994
3689: {\it Science} {\bf 261} 337
3690: \bibitem{DING96}
3691: Ding H, Bellman A F, Campuzano J C, Randeria M, Norman M R,
3692: Yokoya T, Takahashi T, Katayama-Yoshida H, Mochiku T, Kadowaki K, Jennings 
3693: G and Brivio G P 1996 \PRL {\bf 76} 1533
3694: \bibitem{BILAYER}
3695: Feng D L, Armitage N P, Lu D H, Damascelli A, Hu J P, Bogdanov P,
3696: Lanzara A, Ronning F, Shen K M, Eisaki H, Kim C and Shen Z-X
3697: 2001 \PRL {\bf 86} 5550
3698: \bibitem{JCNEW}
3699: Kaminski A, Rosenkranz S, Fretwell H M, Li Z, Raffy H, Randeria M,
3700: Norman M R and Campuzano J C 2002 {\it Preprint} cond-mat/0210531
3701: \bibitem{BAER}
3702: Imer J-M, Patthey F, Dardel B, Schneider W-D, Baer Y, Petroff Y and
3703: Zettl A 1989 \PRL {\bf 62} 336
3704: \bibitem{GAP96}
3705: Ding H, Norman M R, Campuzano J C, Randeria M, Bellman A, 
3706: Yokoya T, Takahashi T, Mochiku T and Kadowaki K 1996
3707: \PR B {\bf 54} R9678
3708: \bibitem{MESOT99}
3709: Mesot J, Norman M R, Ding H, Randeria M,
3710: Campuzano J C, Paramekanti A, Fretwell H M, Kaminski A,
3711: Takeuchi T, Yokoya T, Sato T, Takahashi T, Mochiku T and Kadowaki K
3712: 1999 \PRL {\bf 83} 840
3713: \bibitem{CHIAO}
3714: Chiao M, Lambert P, Hill R W, Lupien C, Gagnon R, Taillefer L and
3715: Fournier P 2000 \PR B {\bf 62} 3554
3716: \bibitem{SUTHER}
3717: Sutherland M, Hawthorn D G, Hill R W, Ronning F, Wakimoto S,
3718: Zhang H, Proust C, Boaknin E, Lupien C, Taillefer L, Liang R,
3719: Bonn D A, Hardy W N, Gagnon R, Hussey N E, Kimura T, Nohara M and
3720: Takagi H 2003 {\it Preprint} cond-mat/0301105
3721: \bibitem{HARRIS}
3722: Harris J M, Shen Z-X, White P J, Marshall D S, Schabel M C,
3723: Eckstein J N and Bozovic I 1996 \PR B {\bf 54} R15665
3724: \bibitem{JOHNZ}
3725: Zasadzinski J F, Ozyuzer L, Miyakawa N, Gray K E,
3726: Hinks D G and Kendziora C 2001 \PRL {\bf 87} 067005
3727: \bibitem{TEMP01}
3728: Norman M R, Kaminski A, Mesot J and Campuzano J C 2001 \PR B {\bf 63}
3729: 140508 (R)
3730: \bibitem{FEDEROV}
3731: Federov A V, Valla T, Johnson P D, Li Q, Gu G D and Koshizuka N 1999
3732: \PRL {\bf 82} 2179
3733: \bibitem{HUANG}
3734: Huang Q, Zasadzinski J F, Gray K E, Liu J Z and Claus H 1989
3735: \PR B {\bf 40} 9366
3736: \bibitem{ARNOLD}
3737: Arnold G B, Mueller F M and Swihart J C 1991 \PRL {\bf 67} 2569
3738: \bibitem{MCMR}
3739: McMillan W L and Rowell J M 1965 \PRL {\bf 14} 108
3740: \bibitem{NORM97}
3741: Norman M R, Ding H, Campuzano J C, Takeuchi T, Randeria M, Yokoya T,
3742: Takahashi T, Mochiku T and Kadowaki K 1997 \PRL {\bf 79} 3506
3743: \bibitem{AC2}
3744: Abanov Ar and Chubukov A V 1999 \PRL {\bf 83} 1652
3745: \bibitem{SS}
3746: Shen Z-X and Schrieffer J R 1997 \PRL {\bf 78} 1771
3747: \bibitem{KEIMER}
3748: Fong H F, Bourges P, Sidis Y, Regnault L P, Ivanov A, Gu G D,
3749: Koshizuka N and Keimer B 1999 {\it Nature } {\bf 398} 588
3750: \bibitem{JULES}
3751: Carbotte J P, Schachinger E and Basov D N 1999 {\it Nature} {\bf 401} 
3752: 354
3753: \bibitem{KINK}
3754: Bogdanov P V, Lanzara A, Kellar S A, Zhou X J, Lu E D, Zheng W J,
3755: Gu G, Shimoyama J-I, Kishio K, Ikeda H, Yoshizaki R,
3756: Hussain Z and Shen Z-X 2000 \PRL {\bf 85} 2581
3757: \bibitem{ESCHRIG}
3758: Eschrig M and Norman M R 2000 \PRL {\bf 85} 3261; 2002 {\it Preprint} 
3759: cond-mat/0202083; 2002 \PRL {\bf 89} 277005
3760: \bibitem{PETER01}
3761: Johnson P D, Valla T, Fedorov A V, Yusof Z, Wells B O, Li Q,
3762: Moodenbaugh A R, Gu G D, Koshizuka N, Kendziora C, Jian S
3763: and Hinks D G 2001 \PRL {\bf 87} 177007
3764: \bibitem{LANZARA}
3765: Lanzara A, Bogdanov P V, Zhou X J, Kellar S A, Feng D L,
3766: Lu E D, Yoshida T, Eisaki H, Fujimori A, Kishio K, Shimoyama J-I,
3767: Noda T, Uchida S, Hussain Z and Shen Z-X 2001
3768: {\it Nature} {\bf 412} 510
3769: \bibitem{PHEN98}
3770: Norman M R, Randeria M, Ding H and Campuzano J C 1998 \PR B
3771: {\bf 57} R11093
3772: \bibitem{JAN}
3773: Engelbrect J R, Nazarenko A, Randeria M and Dagotto E 1998 \PR B {\bf 
3774: 57} 13406
3775: \bibitem{GIAN}
3776: Preosti G, Vilk Y M and Norman M R 1999 \PR B {\bf 59} 1474
3777: \bibitem{FRS}
3778: Furukawa N, Rice T M and Salmhofer M 1998 \PRL {\bf 81} 3195
3779: \bibitem{PUTIKKA}
3780: Putikka W O, Luchini M U and Singh R R P 1998 \PRL {\bf 81} 2966
3781: \bibitem{AEBI}
3782: Aebi P, Osterwalder J, Schwaller P, Schlapbach L, Shimoda M,
3783: Mochiku T and Kadowaki K 1994 \PRL {\bf 72}  2757
3784: \bibitem{DRESDEN}
3785: Kordyuk A A, Borisenko S V, Golden M S, Legner S, Nenkov K A,
3786: Knupfer M, Fink J, Berger H, Forro L and Follath R 2002 \PR B 
3787: {\bf 66} 014502
3788: \bibitem{BIGBOB}
3789: Laughlin R B 1997 \PRL {\bf 79} 1726
3790: \bibitem{JOHN88}
3791: Tranquada J M, Moudden A H, Goldman A I, Zolliker P, Cox D E, Shirane 
3792: G, Sinha S K, Vaknin D, Johnston D C, Alvarez M S, Jacobson A J, 
3793: Lewandowski J T and Newsam J M 1988 \PR B {\bf 38} 2477
3794: \bibitem{REZNIK}
3795: Reznik D, Bourges P, Fong H F, Regnault L P, Bossy J, Vettier C, Milius D L,
3796: Aksay I A and Keimer B 1996 \PR B {\bf 53} 14741
3797: \bibitem{FONG00}
3798: Fong H F, Bourges P, Sidis Y, Regnault L P, Bossy J, Ivanov A, Milius 
3799: D L, Aksay I A and Keimer B 2000 \PR B {\bf 61} 14773
3800: \bibitem{YBCO1D}
3801: Mook H A, Dai P, Dogan F and Hunt R D 2000 {\it Nature} {\bf 404} 729
3802: \bibitem{LAKE1}
3803: Lake B, Aeppli G, Mason T E, Schroder A, McMorrow D F, Lefmann K, 
3804: Isshiki M, Nohara M, Takagi H and Hayden S M 1999 {\it Nature} {\bf 
3805: 400} 43
3806: \bibitem{LAKE2}
3807: Lake B, Aeppli G, Clausen K N, McMorrow D F, Lefmann K, Hussey N E, 
3808: Mangkorntong N, Nohara M, Takagi H, Mason T E and Schroder A 2001 {\it 
3809: Science} {\bf 291} 1759
3810: \bibitem{LAKE3}
3811: Lake B, Ronnow H M, Christensen N B, Aeppli G, Lefmann K, McMorrow D 
3812: F, Vorderwisch P, Smeibidl P, Mangkorntong N, Sasagawa T, Nohara M, 
3813: Takagi H and Mason T E 2002 {\it Nature} {\bf 415} 299
3814: \bibitem{AROVAS}
3815: Arovas D P, Berlinsky A J, Kallin C and Zhang S-C 1997 \PRL {\bf 79} 2871
3816: \bibitem{DSZ}
3817: Demler E, Sachdev S and Zhang Y 2001 \PRL {\bf 87} 067202
3818: \bibitem{TL2201}
3819: He H, Bourges P, Sidis Y, Ulrich C, Regnault L P, Pailhes S, 
3820: Berzigiarova N S, Kolesnikov N N and Keimer B 2002 {\it Science} {\bf 
3821: 295} 1045
3822: \bibitem{WHITE}
3823: Scalapino D J and White S R 1998 \PR B {\bf 58} 8222
3824: \bibitem{DZNAT}
3825: Demler E and Zhang S-C 1998 {\it Nature} {\bf 396} 733
3826: \bibitem{DAI99}
3827: Dai P, Mook H A, Hayden S M, Aeppli G, Perring T G, Hunt R D and 
3828: Dogan F 1999 {\it Nature} {\bf 284} 1344
3829: \bibitem{DAI00}
3830: Dai P, Mook H A, Aeppli G, Hayden S M and Dogan F 2000 {\it Nature} 
3831: {\bf 406} 965
3832: \bibitem{BOURGES}
3833: Bourges P, Casalta H, Regnault L P, Bossy J, Burlet P, Vettier C, 
3834: Beaugnon E, Gautier-Picard P and Tournier R 1997 {\it Physica} B {\bf 
3835: 234-236} 830
3836: \bibitem{RESVOR}
3837: Eschrig M, Norman M R and Janko B 2001 \PR B {\bf 64} 134509
3838: \bibitem{DAI01}
3839: Dai P, Mook H A, Hunt R D and Dogan F 2001 \PR B {\bf 63} 054525
3840: \bibitem{KEE}
3841: Kee H-Y, Kivelson S A and Aeppli G 2002 \PRL {\bf 88} 257002
3842: \bibitem{GANG5}
3843: Abanov Ar, Chubukov A V, Eschrig M, Norman M R and Schmalian J 2002 
3844: \PRL {\bf 89} 177002
3845: \bibitem{ARAI}
3846: Arai M, Nishijima T, Endoh Y, Egami T, Tajima S, Tomimoto K, Shiohara 
3847: Y, Takahashi M, Garrett A and Bennington S M 1999 \PRL {\bf 83} 608
3848: \bibitem{FLORA}
3849: Onufrieva F and Pfeuty P 1999 {\it Preprint} cond-mat/9903097; 2002
3850: \PR B {\bf 65} 054515
3851: \bibitem{KAO}
3852: Kao Y-J, Si Q and Levin K 2000 \PR B {\bf 61} R11898
3853: \bibitem{NORMAN2}
3854: Norman M R 2001 \PR B {\bf 63} 092509
3855: \bibitem{BOURGES00}
3856: Bourges P, Sidis Y, Fong H F, Regnault L P, Bossy J, Ivanov A and 
3857: Keimer B 2000 {\it Science} {\bf 288} 1234
3858: \bibitem{BLEE}
3859: Brinkmann J and Lee P A 1999 \PRL {\bf 82} 2915; 2002 \PR B {\bf 65} 
3860: 014502
3861: \bibitem{BALATSKY}
3862: Batista C D, Ortiz G and Balatsky A V 2001 \PR B {\bf 64} 172508
3863: \bibitem{HIRAKA}
3864: Hiraka H, Endoh Y, Fujita M, Lee Y S, Kulda J, Ivanov A and Birgenau R 
3865: J 2001 {\it J. Phys. Soc. Japan} {\bf 70} 853
3866: \bibitem{ZACK}
3867: Schlesinger Z, Collins R T, Holtzberg F, Feild C, Koren G and Gupta A 
3868: 1990 \PR B {\bf 41} 11237
3869: \bibitem{AV}
3870: Abrahams E and Varma C M 2000 {\it Proc. Natl. Acad. Sci.} {\bf 97} 
3871: 5714
3872: \bibitem{JOE2}
3873: Orenstein J, Thomas G A, Millis A J, Cooper S L, Rapkine D H, Timusk 
3874: T, Schneemeyer L F and Waszczak J V 1990 \PR B {\bf 42} 6342
3875: \bibitem{DUFFY}
3876: Duffy D, Hirschfeld P J and Scalapino D J 2001 \PR B {\bf 64} 224522
3877: \bibitem{JOE3}
3878: Corson J, Orenstein J and Eckstein J N 2000 \PRL {\bf 85} 2569
3879: \bibitem{MUNZAR}
3880: Munzar D, Bernhard C and Cardona M 1999 {\it Physica} C {\bf 312} 121
3881: \bibitem{AC3}
3882: Abanov Ar, Chubukov A V and Schmalian J 2001 \PR B {\bf 63} 180510
3883: \bibitem{DIRK}
3884: Gruninger M, van der Marel D, Tsvetkov A A and
3885: Erb A 2000 \PRL {\bf 84} 1575
3886: \bibitem{TOM}
3887: Timusk T and Homes C C 2002 {\it Preprint} cond-mat/0209371
3888: \bibitem{CHESTER}
3889: Chester G V 1956 \PR {\bf 103} 1693
3890: \bibitem{TINKHAM}
3891: Tinkham M 1975 {\it Introduction to Superconductivity} (New York: McGraw 
3892: Hill)
3893: \bibitem{PWA90}
3894: Anderson P W 1990 \PR B {\bf 42} 2624
3895: \bibitem{HIRSCH}
3896: Hirsch J E 1992 {\it Physica} C {\bf 201} 347
3897: \nonum Hirsch J E and Marsiglio F 2000 {\it Physica} C {\bf 331} 150;
3898: 2000 \PR B {\bf 62} 15131
3899: \nonum Hirsch J E 2002 {\it Science} {\bf 295} 2226
3900: \bibitem{KLEIN}
3901: Klein M V and Blumberg G 1999 {\it Science} {\bf 283} 42
3902: \bibitem{KUBO}
3903: Kubo R 1957 \JPSJ {\bf 12} 570
3904: \bibitem{MNCP}
3905: Norman M R and P\'{e}pin C 2002 \PR B {\bf 66} 100506 (R)
3906: \bibitem{FUGOL}
3907: Fugol I, Samovarov V, Ratner A, Zhuravlev V, Saemann-Ischenko G, 
3908: Holzapfel B and Meyer O 1993 \SSC {\bf 86} 385
3909: \bibitem{LITTLE}
3910: Holcomb M J, Collman J P and Little W A 1994 \PRL {\bf 73} 2360
3911: \bibitem{RUBHAUSEN}
3912: Rubhausen M, Gozar A, Klein M V, Guptasarma P and Hinks D G 2001 \PR B 
3913: {\bf 63} 224514
3914: \bibitem{BASOV}
3915: Basov D N, Woods S I, Katz A S, Singley E J, Dynes R C, Xu M, Hinks D 
3916: G, Homes C C and Strongin M 1999 {\it Science} {\bf 283} 49
3917: \bibitem{LEV}
3918: Ioffe L B and Millis A J 1999 {\it Science} {\bf 285} 1241; 2000 \PR 
3919: B {\bf 61} 9077
3920: \bibitem{LORSH}
3921: Loram J W, Mirza K A and Freeman P F 1990 {\it Physica} C {\bf 171} 243
3922: \bibitem{DVMSC}
3923: Molegraaf H J A, Presura C, van der Marel D, Kes P H and Li M 2002 
3924: {\it Science} {\bf 295} 2239
3925: \nonum van der Marel D 2003 {\it Preprint} cond-mat/0301506
3926: \bibitem{NICOLE}
3927: Santander-Syro A F, Lobo R P S M, Bontemps N, Konstantinovic Z, Li Z 
3928: Z and Raffy H 2001 {\it Preprint} cond-mat/0111539
3929: \bibitem{GANG4}
3930: Norman M R, Randeria M, Janko B and Campuzano J C 2000 \PR B {\bf 61} 
3931: 14742
3932: 
3933: \end{thebibliography}
3934: 
3935: \end{document}
3936: