cond-mat0307508/Rio.tex
1: \documentclass{elsart3}
2: \usepackage{graphicx}
3: 
4: \begin{document}
5: 
6: \begin{frontmatter}
7: 
8: \title{Pseudogap and Competing States in Underdoped Cuprates}
9: 
10: \author{Patrick A.  Lee\thanksref{thank1}},
11: 
12: \address{Center for Materials Science and Engineering and
13: Department of Physics, MIT, Cambridge, MA 02139 USA}
14: 
15: \thanks[thank1] {E-mail: palee@mit.edu}
16: 
17: 
18: %\date{\today}
19: 
20: %\maketitle
21: 
22: 
23: 
24: \begin{abstract}
25: I shall argue that the high $T_c$ problem is the problem of doping into a Mott insulator. Furthermore, the well
26: documented pseudo-gap phenomenon in underdoped cuprates holds the key to understanding this physics. Phase
27: fluctuation alone cannot explain this phenomenon, but  there is a clear need to identify  a competing state which
28: lives in the vortex core. The staggered flux state is a good candidate for the competing state and experimental
29: tests of these ideas will be discussed.
30: \end{abstract}
31: 
32: \begin{keyword}
33: Superconductivity; Pseudogap; Cuprate
34: \end{keyword}
35: \end{frontmatter}
36: 
37: 
38: 
39: In the past several years, a concensus has begun to emerge that the phenomenon of high temperature
40: superconductivity in cuprates is associated with doping into a Mott insulator.  The undoped
41: material is an antiferromagnet with a large exchange energy $J$ of order 1500~K.  The doped holes hop with a
42: matrix element $t$, which is estimated to be approximately $3J$.  However, the  N\'{e}el state is not favorable
43: for hole hopping, because after one hop the spin finds itself in a ferromagnetic environment.  Thus it
44: is clear that the physics is that of competition between the exchange energy $J$ and the hole
45: kinetic energy per hole $xt$.  Apparently the superconducting state emerges as the best
46: compromise, but how and why this occurs is the central question of the high T$_c$ puzzle.  In
47: the underdoped region this competition results in physical properties that are most anomalous. 
48: The metallic state above the superconducting T$_c$ behaves in a way unlike anything we have
49: encountered before.   Essentially, an energy gap appears in some properties and not others, and
50: this metallic state has been referred to as the pseudogap state.  We will focus our attention
51: on this region because the phenomenology is well established and we have the best chance of
52: sorting out the fundamental physics.  
53: 
54: The pseudogap phenomenon is most clearly seen in the uniform susceptibility.  For example,
55: Knight shift measurement in the YBCO 124 compound shows that while the spin susceptibility $\chi_s$ is
56: almost temperature independent between 700~K and 300~K, as in an ordinary metal, it decreases
57: below 300~K and by the time the T$_c$ of 80~K is reached, the system has lost 80\% of the spin
58: susceptibility.\cite{1}  
59: To emphasize the universality of this phenomenon, I reproduce in Fig. 1 some old data on YBCO and LSCO.  Figure 1(a)
60: shows the Knight shft data from Alloul {\it et al}. from 1989.\cite{2}  I have subtracted the orbital
61: contribution, which is generally agreed to be 150~ppm,\cite{3} and drawn in the zero line to highlight the spin
62: contribution to the Knight shift which is proportional to $\chi_s$.  The proportionality constant is known\cite{2}
63: which allows us to draw in the Knight shift which corresponds to the 2D square $S = {1\over 2}$ Heisenberg
64: antiferromagnet with
65: $J = 0.13$~eV.\cite{4,5}  The point of this exercist is to show that in the underdoped region, the spin 
66: susceptibility drops {\it below} that of the Heisenberg model at low temperatures before the onset of
67: superconductivity.  This trend continues even in the severely underdoped limit (O$_{0.53}$ to O$_{0.41}$), showing
68: that the $\chi_s$ reduction cannot simply be understood as fluctuations towards the antiferromagnet. 
69: Note that the discrepancy is worse if $J$ were replaced by a smaller $J_{eff}$ due to doping, since $\chi_s \sim
70: J_{eff}^{-1}$. 
71: The data seen
72: in this light strongly point to singlet formation as the origin of the pseudogap seen in the uniform spin
73: susceptibility.
74: 
75: \begin{figure}[h]
76: %h=here, t=top, b=bottom, p=separate figure page
77: %\begin{center}\leavevmode
78: \includegraphics[width=1.0125\linewidth]{Fig.1a.eps}
79: \includegraphics[width=1.0325\linewidth]{Fig.1b.eps}
80: \caption{(a) Knight shift data of YBCO for a variety of doping.\cite{2}  The zero level for the spin
81: contribution has been added and the solid line represents the prediction of the 2D $S={1\over 2}$
82: Heisenberg model for $J = 0.13$~eV.  (b) Uniform magnetic susceptibility for LSCO.\cite{7}  The orbital
83: contribution $\chi_0$  is shown (see text) and the solid line represents the Heisenberg model
84: prediction.}
85: %\end{center}
86: \end{figure}
87: 
88: It is worth noting that the trend shown in Fig. 1 is not so apparent if one looks at the measured spin
89: susceptibility directly.\cite{6}  This is because the van~Vleck part of the spin susceptibility is doping
90: dependent, due to the changing chain contribution.  This problem does not arise for LSCO, and in Fig. 1(b) we show
91: the uniform susceptibility data.\cite{7}  The zero of the spin part is determined by comparing susceptibility
92: measurements to $^{17}$O Knight shift data.\cite{8}  Nakano {\it et al}.\cite{7} find an excellent fit for the $x =
93: 0.15$ sample (see Fig. 9 of ref.  7) and determine the orbital contribution for this sample to be $\chi_0 \sim 0.4
94: \times 10^{-7}$~emu/g. This again allows us to plot the theoretical prediction for the Heisenberg model.\cite{9} 
95: Just as for YBCO, $\chi_s$ for the underdoped samples ($x = $0.1 and 0.08) drops below that of the Heisenberg model. 
96: In fact, the behavior of
97: $\chi_s$ for the two systems is remarkably similar, especially in the underdoped region.
98: 
99: 
100: 
101: A second indication of the pseudogap comes from the linear $T$ coefficient of the specific heat, which shows a marked
102: decrease below room temperature.\cite{10}  It is apparent that the spins are forming into singlets and
103: the spin entropy is gradually lost.  On the other hand, the frequency dependent conductivity
104: behaves very differently depending on whether the electric field is in the $ab$ plane
105: $(\sigma_{ab})$ or perpendicular to it $(\sigma_c)$.  At low frequencies (below 500~cm$^{-1}$)
106: $(\sigma_{ab})$ shows a typical Drude-like behavior for a metal with a width which decreases
107: with temperature, but an area (spectral weight) which is independent of temperature.\cite{11}  Thus
108: there is no sign of the pseudogap in the spectral weight.  This is surprising because in other
109: examples where an energy gap appears in a metal, such as the onset of charge or spin density
110: waves, there is a redistribution of the spectral weight from the Drude part to higher
111: frequencies.  On the other hand below 300~K $\sigma_c(\omega)$ is gradually reduced for frequencies below
112: 500~cm$^{-1}$ and a deep hole is carved out of $\sigma_c(\omega)$ by the time T$_c$ is
113: reached.\cite{12}   Finally, angle-resolved photoemission shows that an energy gap (in
114: the form of a pulling back of the leading edge of the electronic spectrum from the Fermi
115: energy) is observed near momentum $(0,\pi)$ and the onset of superconductivity is marked by the
116: appearance of a small coherent peak at this gap edge.
117: 
118: 
119: The pseudogap phenomenology is well explained by a cartoon picture which emerges from the RVB
120: (resonating valence band) theory of Anderson.\cite{13}  The spins are paired into singlet pairs. 
121: However, the pairs are  not static but are fluctuating due to quantum mechanical superposition,
122: hence the term quantum spin liquid.  The singlet formation explains the appearance of the spin gap and the
123: reduction of spin entropy.  The doped holes appear as vacancies in the background of singlet
124: pair liquid and can carry a current without any energy gap.  However in $c$-axis conductivity
125: and electron is removed from one plane and placed on the next.  The intermediate state is an
126: electron which carries spin 1/2 and therefore it is necessary to break a singlet pair and pay
127: the spin-gap energy.  The same consideration applies to the photoemission experiment.  Finally, according to RVB
128: theory, superconductivity emerges when the holes become phase coherent.  The spin singlet
129: familiar in the BCS theory has already been formed.
130: 
131: While the above picture is appealing, there has been another popular explanation.  The idea is that the pseudogap
132: phenomenology can be understood as a superconductor with robust amplitude but strong phase fluctuation.  The
133: superfluid density
134: $\rho_s$ which controls the phase stiffness is proportional to the doping concentration $x$ and
135: becomes small in the underdoped region.  As emphasized by Uemura\cite{4} and by Emery and Kivelson,\cite{5}
136: T$_c$ is controlled by $\rho_s$ and is much lower than the energy gap.  We shall now argue that
137: phase fluctuations cannot be the whole story.  Setting aside the question of where
138: the strong pairing amplitude comes from in the first place, that the phase fluctuation scenairo
139: is incomplete can be seen from the following argument.  In two dimensions the destruction of
140: superconducting order is via the Berezinskii-Kosterlitz-Thouless (BKT) theory of vortex
141: unbinding.  Above T$_c$ the number of vortices proliferate and the normal metallic state is
142: reached only when the vortex density is so high that the cores overlap.  At lower
143: vortex density, transport properties will resemble a superconductor in the flux flow regime. 
144: In ordinary superconductors, the BKT temperature is close to the mean field temperature, and
145: the core energy rapidly becomes small.  However, in the present case, it is postulated that the
146: mean field temperature is high, so that a large core energy is expected.  Indeed, in a
147: conventional core the order parameter and energy gap vanish, costing $\Delta^2_0/E_F$ per unit area of energy. 
148: Using a core radius of $\xi = V_F/\Delta_0$, the core energy of a conventional superconductor is $E_F$.  In our
149: case, we may replace $E_F$ by $J$.  If this were the case, the proliferation of vortices would not happen until a
150: high temperature $\sim J$ independent of $x$ is reached.  Thus for the phase fluctuation scenario to to work, it
151: is essential to have ``cheap'' vortices, with energy cost of order T$_c$.  Then the essential problem is to
152: understand what the vortex core is made of.  Put another way, there has to be a competing state with energy very
153: close to the $d$-wave superconductors which constitute the core.  The vortex core indeed offers a glimpse of the
154: normal state reached when $H$ exceeds
155: $H_{c2}$, and is an important constituent of the pseudogap state above T$_c$.
156: 
157: 
158: What are the candidates for the competing order?  A candidate which has attracted a lot of
159: attention is the stripe phase.\cite{16,17}    In the LSCO familty, dynamical stripes (spin
160: density waves) are clearly important, especially near $x = {1\over 8}$.  There are recent
161: report of incommensurate SDW nucleating around vortices.  However, until now there has been little
162: evidence for stripes outside of the LSCO family.  On the theoretical side, as a competing state
163: it is not clear how the stripes are connected to $d$-wave superconductivity and it is hard to
164: understand how the nodal quasiparticles turn out to be most sharply defined on the
165: Fermi surface, since these have to transverse the stripes at a $45^\circ$ angle.
166: 
167: A second candidate for the vortex core is the antiferromagnetic state.  This possibility was
168: first proposed several years ago in the context of the SO(5) theory.\cite{18}  This theory is
169: phenomenological in that it involves only bosonic degrees of freedom (the SDW and pairing order
170: parameters).  The quasiparticles are out of the picture.  Thus the fundamental question of
171: how the holes are accommodated has not really been addressed.   There are reports of enhanced antiferromagnetic spin
172: fluctuations, and perhaps even static order, using NMR.\cite{19,20}  I shall argue next that other considerations
173: also lead to antiferromagnetic fluctuations and possibly static orders inside the vortex core, so that the
174: observation of antiferromagnetic cores does not necessarily imply the existence of SO(5)
175: symmetry.
176: 
177: Finally, I come to the candidate which we favor --- the staggered flux state with orbital
178: currents.\cite{21}  Indeed, using the staggered flux state as the core, Lee and Wen have successfully
179: constructed a ``cheap'' vortex state.\cite{22}  The staggered flux phase has an
180: advantage over other possibilities in that its excitation spectrum is similar to the $d$-wave
181: superconductor and the SU(2) theory allows us to smoothly connect  it to the
182: superconductivity.   We also regard the staggered flux phase as the precursor to  N\'{e}el
183: order, so that antiferromagnetic  fluctuations or even SDW order are accommodated naturally. 
184: Of course, it is experiments which have the final say as to which candidate turns out to be
185: realized.  Our strategy is to work out as many experimental  consequences as we can and propose
186: experiments to confirm or falsify our theory.
187: 
188: The staggered flux state was first introduced as a mean field solution at half-filling\cite{23}
189: and later was extended to include finite doping.\cite{24}   At half-filling, due to the constraint of no double
190: occupation, the staggered flux state corresponds to an insulating state with power law decay in the spin
191: correlation function.  It is known that upon including gauge fluctuations which enforce the
192: constraint, the phenomenon of confinement and chiral symmetry breaking occurs, which directly
193: corresponds to  N\'{e}el ordering.\cite{25}  The idea is that with doping, confinement is
194: suppressed at some intermediate energy scale, due to screening by holes and to dissipation.\cite{26}
195: As the temperature is lowered,  the pseudogap state emerges which can be understood as fluctuating
196: between the staggered flux state and the $d$-wave superconducting state.  
197: As still lower tamperature, the staggered flux states become dilute and form the core of fluctuating $hc/2e$
198: vortices. Finally,  the vortices bind via the BKT transition and 
199: the $d$-wave superconducting state is the stable ground state. 
200: Thus the staggered flux state may be regarded as the ``mother state'' which is an unstable
201: fixed point due to gauge fluctuations.  It flows to  N\'{e}el ordering at half-filling and to
202: the $d$-wave superconductor for sufficiently large $x$.  Thus the staggered flux state plays a
203: central role in this kind of theory.  This picture is depicted schematically in Fig. 2.  
204: We should point out that the
205: staggered flux state (called the $D$-density wave state) has recently been proposed as the ordered state in the
206: pseudogap region.\cite{27}  As explained elsewhere,\cite{28} we think that this view is not supported
207: by experiment and we continue to favor the fluctuation picture.
208: 
209: 
210: \begin{figure}[h]
211: %h=here, t=top, b=bottom, p=separate figure page
212: \begin{center}\leavevmode
213: %\includegraphics[width=1.0\linewidth]{Fig.2.eps}
214: %\includegraphics[width=.95\linewidth]{Fig.2.eps}
215: \includegraphics{Fig.2.eps}
216: \caption{Schematic representation of our view of the phase diagram.  In the undoped limit the staggered flux
217: state develops into the N\'{e}el state once gauge fluctuations are included.  The
218: pseudogap phase
219: can be thought of as a fluctuating phase between the staggered flux state and the $d$-wave
220: superconductor.  The
221: SU(2) theory allows us to smoothly connect these fluctuations.  In this theory different states may be
222: represented
223: by a three-dimensional quantization axis with arrows pointing to the north and south poles representing
224: staggered
225: flux order and arrows in the equator representing the  $d$-wave superconductor.  These arrows are
226: schematically
227: shown in the figure.  As temperature is lowered, the staggered flux regions (arrows pointing to north
228: or south pole)
229: are localized to form cores of $hc/2e$~vortices.  Eventually these vortices disappear via the
230: Kosterlitz-Thouless
231: transition to the
232: $d$-wave superconducting state.  There is a broad range in  temperatures above this transition where
233: vortices may
234: give rise to Nernst effects.\cite{39}  This is sketched in the schematic phase diagram.
235: }
236: \end{center}
237: \end{figure}
238: 
239: 
240: The above picture finds support from studies of  projected wavefunctions, where the 
241: no-double-occupation constraint is enforced by hand on a computer.  With doping the best state
242: is a projected
243: $d$-wave state.   This state can explain many of the properties of the superconductor, as recently discussed by
244: Paramekanti {\it et al}.\cite{29}  It is natural to consider the projected staggered flux state (at finite doping) as
245: a trial wavefunction for the ``normal state'' which exists inside the vortex core.  The energy difference between
246: this and the projected $d$-wave superconductor may be considered the condensation energy.  The condensation energy per
247: site computed this way is shown in Fig. 3.\cite{30}  Note the dome shape which is reminiscent of the $T_c$ curve
248: and the rather small value for the condensation energy, consistent  with our expectation based on the SU(2) idea. 
249: Another interesting quantity we calculated\cite{31} is the current-current correlation function for the projected
250: $d$-wave BCS wavefunction:
251: $
252: c_j(k,\ell) = < j(k)j(\ell) >
253: $
254: where $j(k)$ is the physical electron current on the bond $k$.  The average current $<j(k)>$ is
255: obviously zero, but the correlator exhibits a staggered circulating pattern.\cite{31} 
256: Such a pattern is absent in the $d$-wave BCS state before
257: projection, and is a result of the Gutzwiller projection.  Our result for $c_j$ is consistent with exact
258: diagonalization of two holes in 32 sites.\cite{32}
259: 
260: 
261: The staggered current generates a staggered physical
262: magnetic field (estimated to be 10--40 gauss)\cite{24,28} which may be detected, in principle, by neutron
263: scattering.  In practice the small signal makes this a difficult, though not impossible experiment and we
264: are motivated to look for situations where the orbital current may become static or quasi-static.  Recently, we
265: analyzed the structure of the $hc/2e$ vortex in the superconducting state within the SU(2) theory and concluded
266: that in the vicinity of the vortex core, the orbital current becomes quasi-static, with a time scale determined by
267: the tunnelling between two degenerate staggered flux states.\cite{22}  It is very likely that this time is
268: long on the neutron time scale.   Thus we propose that a quasi-static peak centered around $(\pi,\pi)$ will appear
269: in neutron scattering in a magnetic field, with intensity proportional to the number of vortices.  The time
270: scale may actually be long enough for the small magnetic fields  generated by the orbital currents to be
271: detectable by
272: $\mu$-SR or Yttrium NMR.  Again, the signal should be proportional to the external fields.  (The NMR experiment
273: must be carried out in 2--4--7 or 3 layer samples to avoid the cancellation between bi-layers.)  We have also
274: computed the tunnelling density of states in the vicinity of the vortex core, and predicted a rather specific kind
275: of period doubling which should be detectable by atomic resolution STM.\cite{33}  The recent
276: report\cite{34} of a static field of $\pm 18$ gauss in underdoped YBCO which appears in the vortex state is
277: promising, even though muon cannot distinguish between orbital current or spin as the origin of the magnetic
278: field.  We remark that in the underdoped antiferromagnet, the local moment gives rise to a field of 340 gauss
279: at the muon site.  Thus if the 18 gauss signal is due to spin, it will correspond to roughly $1/20$th of the
280: full moment.
281: 
282: \begin{figure}[h]
283: %h=here, t=top, b=bottom, p=separate figure page
284: \begin{center}\leavevmode
285: \includegraphics[width=0.8\linewidth]{Fig.3.eps}
286: \caption{The condensation energy per site vs. doping
287: as estimated from the difference between the energy of the projected staggered flux state and the
288: projected $d$-wave superconductor.\cite{30}  Data are shown for a variety of
289: sample sizes.}
290: \end{center}
291: \end{figure}
292: 
293: 
294: 
295: We remark that our analytic model of the vortex core is in full agreement with the numerical solution of
296: unrestricted mean field $\Delta_{ij}$ and $\chi_{ij}$ by Wang, Han and Lee\cite{35} and Ogata and
297: collaborators.\cite{36} Recently we found that for small doping in the $t$-$J$ model a small moment SDW co-exists
298: with orbital currents in the vortex core.\cite{37}  More generally, we expect $(\pi,\pi)$ spin fluctuations to be
299: enhanced\cite{38} so that the tendency to antiferromagnetism is fully compatible with the staggered flux picture. 
300: This vortex solution is also interesting in that the tunnelling density of states show a gap, with no sign of the
301: large resonance associated with Caroli-deGennes-type core levels found in the standard BCS model of the vortex. 
302:  The low density of states inside the vortex core has an important implication.  In the standard
303: Bardeen-Stephen model of flux-flow resistivity, the friction coefficient of a moving vortex is due to dissipation
304: associated with the vortex core states.  Now that the core states are absent, we can expect anomalously small
305: friction coefficients for underdoped cuprates.  The vortex moves fast transverse to the current and gives rise to
306: large flux-flow resistivity.   Since the total conductivity is the sum of the flux-flow conductivity and the
307: quasiparticle conductivity, it is possible to get into a situation where the quasiparticle conductivity dominates
308: even for H
309: $\ll$ H$_{c2}$.   Thus the ``cheap'' and ``fast'' vortex opens the possibility of having vortex states above the
310: nominal T$_c$ and H$_{c2}$, when the resistivity looks like that of a metal, with little sign of flux-flow
311: contribution.  From this point of view, the large Nerst effect observed by Ong and
312: co-workers \cite{39} over a large region in the H-T plane above the nominal T$_c$ and H$_{c2}$ (as determined
313: by resistivity) may be qualitatively explained.  The schematic phase diagram is shown in Fig. 2.
314: 
315: I thank X.-G. Wen, N.  Nagaosa, D. Ivanov, J. Kishine and Y. Morita for their collaboration on the
316: work reviewed here.  I acknowledge the support of NSF through the MRSEC program with grant number DMR 02--13282.
317: 
318: 
319: \begin{thebibliography}{9}
320: \bibitem{1} N. Curro {\it et al}., Phys. Rev. B{\bf 56} (1997) 877.
321: 
322: \bibitem{2} H. Alloul, T. Ohno, and P. Mendels, Phys. Rev. Lett. {\bf 63} (1989) 1700.
323: 
324: \bibitem{3}  M. Takigawa, W.L. Hults, and J.L. Smith, Phys. Rev. Lett. {\bf 71} (1993) 2650.
325: 
326: \bibitem{4} H.O. Ding and M. Makivic, Phys. Rev. B{\bf 43} (1991) 3562.
327: 
328: \bibitem{5} A.W. Sandvik, E. Dagotto, and D.J. Scalapino, Phys. Rev. B{\bf56} (1997) 11701.
329: 
330: \bibitem{6} J.M. Tranquada {\it et al}., Phys. Rev. B{\bf 38} (1988) 2477.
331: 
332: \bibitem{7} T. Nakano {\it et al}., Phys. Rev. B{\bf 49}, 16000 (1994).
333: 
334: \bibitem{8} K. Isida, Y. Kitaoka, G.Q. Zhang, and K. Asayama, J. Phys. Soc. Jpn. {\bf 60}, 3516 (1991).
335: 
336: \bibitem{9} We note that a comparison of $\chi_s$ for YBCO and LSCO was made by A.J. Millis and H. Monien [Phys.
337: Rev. Lett. {\bf 70}, 2810 (1993)].  Their YBCO analysis is similar to ours.  However, for LSCO they find  a rather
338: different 
339: $\chi_0$ by matching the measurement above 600~K to that of the Heisenberg model.  Consequently, their $\chi_s$ looks
340: different for YBCO and LSCO.  We believe their procedure is not really justified.
341: 
342: \bibitem{10} W.J. Loram {\it et al}., Physica C {\bf 341--348} (2000) 831.
343: 
344: \bibitem{11} A.F. Santander-Syro {\it et al}., Phys. Rev. Lett. {\bf 88} (2002) 097005.
345: 
346: \bibitem{12} C.C. Homes, T. Timusk, R. Liang, D.A. Bonn,  and W.N. Hardy, Phys.  Rev. Lett. {\bf 71} (1993) 4210.
347: 
348: 
349: \bibitem{13} P.W. Anderson, Science {\bf 235} (1987) 1196.
350: 
351: \bibitem{14} Y. Uemura {\it et al}., Phys. Rev. lett. {\bf 66} (1991) 2665.
352: 
353: \bibitem{15} V.J. Emery, S. Kivelson, Nature {\bf 374} (1995) 434.
354: 
355: \bibitem{16} V.J. Emery, S. Kivelson, J.M. Tranquada, Proc. Ntl. Acad. Sci. USA {\bf 96} (1999) 8814.
356: 
357: \bibitem{17} S. Sachdev, cond-mat/0211005, to appear in Rev. Mod. Phys.
358: 
359: \bibitem{18} D.P. Arovas {\it et al}., Phys. Rev. Lett. {\bf 79} (1997) 2871.
360: 
361: \bibitem{19} V.F. Mitrovic {\it et al}., Nature {\bf 413} (2001) 501.
362: 
363: \bibitem{20} K. Kakuyanagi {\it et al}., Phys. Rev. Lett. {\bf 90} (2003) 197003.
364: 
365: \bibitem{21} X.-G.  Wen, P.A. Lee, Phys. Rev. Lett. {\bf 76} (1996)  503.
366: 
367: \bibitem{22} P.A. Lee, X.-G. Wen, Phys. Rev. B{\bf 63} (2001) 224517.
368: 
369: 
370: \bibitem{23} J.B. Marston, I. Affleck, Phys. Rev. B{\bf 39} (1989) 11538.
371: 
372: 
373: \bibitem{24} T. Hsu  {\it et al}., Phys. Rev. B{\bf 43}  (1991) 2866.
374: 
375: \bibitem{25} D.H. Kim and P.A.  Lee, Annals of Phys. {\bf 272}  (1999) 130.
376: 
377: \bibitem{26} N. Nagaosa, Phys. Rev. Lett. {\bf 71}  (1993) 4210; 
378: N. Nagaosa and P.A. Lee, Phys. Rev. B{\bf 61}, 9166 (2000).
379: 
380: \bibitem{27} S. Chakravarty {\it et al}., Phys. Rev. B{\bf 63} (2001) 94503.
381: 
382: \bibitem{28} P.A. Lee, J. of Phys. Chem. Solids {\bf 388-389} (2003) 7.
383: 
384: \bibitem{29} A. Paramekanti, M. Randeria, and N. Trivedi,  Phys. Rev. Lett. {\bf 87}, (2001) 217002.
385: 
386: \bibitem{30} D. Ivanov and P.A. Lee, cond-mat/0305143.
387: 
388: \bibitem{31} D.A. Ivanov, P.A. Lee, X.-G. Wen, Phys. Rev. Lett. {\bf 84} (2000) 3958.
389: 
390: \bibitem{32} P.W. Leung, Phys. Rev. B{\bf 62} (2000) 6112.
391: 
392: \bibitem{33} J. Kishine, P.A. Lee, X.-G. Wen, Phys. Rev. Lett. {\bf 86} (2001)  5365; Phys.  Rev. B{\bf 65} (2002)
393: 064526.
394: 
395: 
396: \bibitem{34} R.I. Miller {\it et al}., Phys. Rev. Lett. {\bf 88} (2002) 137002.
397: 
398: 
399: \bibitem{35} Q.-H. Wang {\it et al}., Phys. Rev. Lett. {\bf 87} (2001) 167004.
400: 
401: \bibitem{36} H. Tsuchiura, M. Ogata, Y. Tanaka, and S.Kashiwaya, cond-mat/0302030.
402: 
403: \bibitem{37} Y. Morita, J. Kishine, P.A. Lee, unpublished.
404: 
405: 
406: \bibitem{38} W. Rantner and X.-G. Wen, Phys. Rev. B{\bf 66} (2002) 144501 .
407: 
408: \bibitem{39} Y. Wang et al., Phys. Rev.  Lett. {\bf 88} (2002) 257003.
409: 
410: \end{thebibliography}
411: 
412: \end{document}
413: 
414: