1: %\documentstyle[aps,prl,preprint]{revtex}
2: \documentstyle[aps,floats,epsf,twocolumn]{revtex}
3:
4:
5: \begin{document} \draft
6:
7: \title{Vortex state in a doped Mott insulator}
8:
9: \author{M. Franz and Z. Te\v{s}anovi\'c}
10: \address{Department of Physics and Astronomy, Johns Hopkins University,
11: Baltimore, MD 21218
12: \\ {\rm(\today)}
13: }
14: %\maketitle
15: %
16: %\begin{abstract}
17: \address{~
18: \parbox{14cm}{\rm
19: \medskip
20: We analyze the recent vortex core spectroscopy data on cuprate
21: superconductors and discuss what can be learned from them about the nature
22: of the ground state in these compounds. We argue that the data are
23: inconsistent with the assumption of a simple metallic
24: ground state and exhibit characteristics of a doped Mott insulator.
25: A theory of the vortex core in such a doped Mott insulator is
26: developed based on the U(1) gauge field slave boson model. In the limit of
27: vanishing gauge field stiffness such theory predicts two types of singly
28: quantized vortices: an insulating ``holon'' vortex in the underdoped
29: and metallic ``spinon'' vortex in the overdoped region of the phase diagram.
30: We argue that the holon vortex exhibits a pseudogap excitation spectrum in
31: its core qualitatively consistent with the existing experimental data on
32: Bi$_2$Sr$_2$CaCu$_2$O$_8$. As a test of this theory we propose that spinon
33: vortex with metallic core might be observed in the heavily overdoped samples.
34: }}
35: %\end{abstract}
36: \maketitle
37:
38: %\pacs{74.60.-w,74.60.Ec,74.72.-h}
39:
40: %
41: \narrowtext
42:
43: \section{Introduction}
44: Nature of the ground state as a function of doping remains one of the recurring
45: unresolved issues
46: in the theory of high-$T_c$ cuprate superconductors. The problem is
47: partly due to formidable difficulties related to the theoretical description of
48: doped Mott insulators and partly due to experimental hurdles in accessing
49: the normal state properties in the $T\to 0$ limit because of the intervening
50: superconducting order. Probes that suppress superconductivity and
51: reveal the properties of the underlying ground state are therefore of
52: considerable value. So far only pulsed
53: magnetic fields\cite{ando1} in excess of $H_{c2}$ and impurity doping
54: beyond the critical concentration\cite{lemberger1} have been used
55: towards this goal. Here we argue that the vortex core
56: spectroscopy performed using scanning tunneling microscope (STM)
57: can provide new insights into the nature of the ground state in cuprates.
58: We analyze the existing experimental data\cite{maggio1,renner1,pan1,pan2}
59: and conclude that they imply strongly correlated ``normal'' ground state,
60: presumably derivable from a doped Mott insulator. We then
61: develop a theoretical framework for the problem of
62: tunneling in the vortex state of such a doped Mott insulator.
63:
64: In the vortex core the superconducting order parameter is locally suppressed
65: to zero and the region within a coherence length $\xi$ from its center can
66: be to the first approximation thought of as normal. Spectroscopy of the
67: vortex core therefore provides information on the normal state electronic
68: excitation spectrum in the $T\to 0$ limit. More accurately, the core
69: spectroscopy reflects the spectrum in the spatially non-uniform situation
70: where the order parameter amplitude rapidly varies in response to the
71: singularity in the phase imposed by the external magnetic field. In order to
72: extract useful information regarding the underlying ground state
73: from such measurements a detailed understanding of the vortex core
74: physics is necessary. So far the problem has been addressed using the weak
75: coupling approach based on the
76: Bogoliubov-de Gennes theory generalized to the $d$-wave symmetry of
77: the order parameter\cite{soininen1,wang1,franz1,kita1}, and semiclassical
78: calculations\cite{volovik1,maki1,ichioka1}.
79: The early theoretical debate focused on the existence
80: or absence of the vortex core bound states
81: \cite{maki1,franz2,himeda1}. This debate, now resolved in favor
82: of absence of any bound states in pure $d_{x^-y^2}$ state
83: \cite{franz1,kita1,resende1}, has somewhat eclipsed the possibly
84: more important issues related to the nature of the ground state
85: in cuprates.
86:
87: The body of work based on mean field, weak coupling calculations
88: \cite{wang1,franz1,kita1,ichioka1} yields results for the
89: local density of states in the vortex core which exhibit two generic features:
90: (i) the coherence peaks (occurring at $E=\pm\Delta_0$ in the bulk) are
91: suppressed,
92: with the spectral weight transferred to a (ii) broad featureless peak
93: centered around the zero energy.
94: %Both features are easily understood on
95: %physical grounds. The smearing of the coherence peaks results from the gap
96: %not being sharp on the length-scale $\xi$ in the core and the broad peak is a
97: %remnant the bound states that would exist in the $s$-wave
98: %vortex\cite{caroli1} strongly hybridized into the continuum states
99: %present in a $d$-wave superconductor.
100: Here we wish to emphasize the heretofore little appreciated fact that
101: these features are {\em qualitatively inconsistent}
102: with the existing experimental data on cuprate superconductors. STM
103: spectroscopy on Bi$_2$Sr$_2$CaCu$_2$O$_8$
104: (BSCCO) at 4.2K indicates a ``pseudogap'' spectrum in the vortex core with the
105: spectral weight from the coherence peaks at $\pm\Delta_0\simeq 40$meV
106: transferred to {\em high energies}, and no peak whatsoever
107: around $E=0$\cite{renner1}. Recent high resolution
108: data on the same compound\cite{pan2}
109: confirmed these findings down to 200mK and found evidence for weak bound
110: states at $\pm7$meV. Experiments on YBa$_2$Cu$_3$O$_7$ (YBCO)\cite{maggio1}
111: also indicate low energy bound states, but are somewhat more difficult to
112: interpret because of the high zero-bias conductance of unknown origin
113: appearing even in the absence of magnetic field.
114:
115: The fundamental discrepancy between the theoretical predictions and
116: the experimental findings strongly suggests that models based on a simple
117: weak coupling theory break down in the vortex core. The pseudogap observed
118: in the core hints that the underlying ground state revealed
119: by local suppression of the superconducting order parameter is
120: a doped Mott insulator and not a conventional metal. Taking into account
121: the effects of strong correlations appears to be necessary to
122: consistently describe the physics of the vortex core.
123: Conversely, studying the vortex core physics could provide information
124: essential for understanding the nature of the underlying ground state
125: in cuprates.
126:
127: The first step in this direction was taken by Arovas {\em et al.}
128: \cite{arovas1} who proposed that within the framework of the SO(5) theory
129: \cite{zhang1} vortex cores could become antiferromagnetic (AF).
130: They found that such AF cores can be stabilized
131: at low $T$ but only in the close vicinity of the bulk AF phase. In contrast,
132: experimentally the pseudogap in the core is found to persist
133: into the overdoped
134: region\cite{renner1}. More recently microscopic calculations within the same
135: model\cite{andersen1} revealed electronic excitations in such AF cores with
136: behavior roughly resembling the experimental data. Quantitatively,
137: however, these spectra exhibit asymmetric
138: shifts in the coherence peaks (related to the fact that spin gap in the AF
139: core is no longer tied to the Fermi level) not observed experimentally. These
140: discrepancies suggest that generically cores will not exhibit the true AF
141: order. Finally, these previous approaches are still of the
142: Hartree-Fock-Bogoliubov type and cannot be expected to properly capture the
143: effects of strong correlations.
144:
145: Here we consider a model for the vortex core based on a version of
146: the U(1) gauge field slave boson theory formulated recently by
147: Lee\cite{dhlee1}. Originally proposed by Anderson\cite{anderson1}
148: the slave boson theory was formulated
149: to describe strongly correlated electrons in the CuO$_2$ planes of the
150: high-$T_c$ cuprates. Various versions of
151: this theory have been extensively discussed in the
152: literature \cite{baskaran1,ruckenstein1,kotliar1,affleck1,lee1}.
153: Interest in spin-charge separated
154: systems revived recently\cite{wen1,lee2,balents1,dhlee1} due to the
155: realization that it provides a natural description of the pseudogap phenomenon
156: observed in the underdoped cuprates. The common ingredient in these theories
157: is ``splintering'' of the electron into quasiparticles carrying its spin
158: and charge degrees of freedom. Within the theories based on Hubbard and
159: $t$-$J$ models this splintering is formally implemented by the
160: decomposition of the electron creation operator
161: %
162: \begin{equation}
163: c^\dagger_{i\sigma}=f^\dagger_{i\sigma} b_i
164: \label{}
165: \end{equation}
166: %
167: into a fermionic spinon $f_{i\sigma}$ and bosonic holon $b_i$. The local
168: constraint of the single occupancy $b^\dagger_ib_i+
169: f^\dagger_{i\sigma}f_{i\sigma}=1$ is enforced by a fluctuating U(1) gauge
170: field ${\bf a}$. The mean field phase diagram is known to contain
171: four phases distinguished
172: by the formation of spinon pairs, $\Delta_{ij}=\langle\epsilon_{\sigma\sigma'}
173: f^\dagger_{i\sigma}f^\dagger_{j\sigma'}\rangle$, and Bose-Einstein condensation
174: of the individual holons $b=\langle b_i\rangle$\cite{lee1}, and is illustrated
175: in Figure (\ref{fig1}).
176: %
177: \begin{figure}[t]
178: \epsfxsize=8.5cm
179: \epsffile{fig1.ps}
180: \caption[]{Schematic phase diagram of the system with spin-charge
181: separation in the doping$-$temperature plane, as applied to cuprate
182: superconductors. }
183: \label{fig1}
184: \end{figure}
185: %
186:
187: The effects of magnetic field on such spin-charge separated system is
188: most conveniently studied in the framework of an effective Ginzburg-Landau
189: (GL) theory for the condensate fields $\Delta$ and $b$. The corresponding
190: effective action can be constructed\cite{sachdev1,lee3}
191: based on the requirements of local gauge invariance with respect to
192: the physical electromagnetic vector potential ${\bf A}$ and the internal
193: gauge field ${\bf a}$:
194: %
195: \begin{eqnarray}
196: f_{\rm GL}&=&|(\nabla-2i{\bf a})\Delta|^2 +{r_\Delta}|\Delta|^2 +
197: {1\over 2}{u_\Delta}|\Delta|^4\nonumber \\
198: &+&|(\nabla-i{\bf a}-ie{\bf A})b|^2 +r_b|b|^2 +{1\over 2} u_b|b|^4
199: +v|\Delta|^2|b|^2 \nonumber \\
200: &+& {1\over 8\pi}(\nabla\times{\bf A})^2 +f_{\rm gauge}.
201: \label{fgl1}
202: \end{eqnarray}
203: %
204: The factor of 2 in the spinon gradient term reflects the fact that {\em pairs}
205: of spinons were assumed to condense. $f_{\rm gauge}$ describes the dynamics
206: of the internal gauge field ${\bf a}$. We note that unlike the physical
207: electromagnetic field ${\bf A}$ the gauge field ${\bf a}$ has no independent dynamics
208: in the underlying microscopic model since it serves only to enforce a
209: constraint. Sachdev\cite{sachdev1} and Nagaosa and Lee\cite{lee3} assumed
210: that upon integrating out the microscopic degrees of freedom a term
211: %
212: \begin{equation}
213: f_{\rm gauge}={\sigma\over2}(\nabla\times{\bf a})^2
214: \label{fgauge}
215: \end{equation}
216: %
217: is generated in the free energy. They then analyzed vortex solutions of
218: the free energy (\ref{fgl1}) and came to the conclusion that two types of
219: vortices are permissible: a ``holon vortex'' with the singularity in the
220: $b$ field and a ``spinon vortex'' with the singularity in the
221: $\Delta$ field. Because holons
222: carry electric charge $e$ the holon vortex is threaded by electronic flux
223: quantum $hc/e$, i.e. twice the conventional
224: superconducting flux quantum $\Phi_0=hc/2e$.
225: Spinons on the other hand condense in pairs, and the spinon vortex therefore
226: carries flux $\Phi_0$.
227: Stability analysis then implies that spinon vortex will be stable over the
228: most of the superconducting phase diagram, while the
229: $hc/e$ holon vortex can be stabilized
230: only in the close vicinity of the phase boundary on the underdoped side
231: \cite{sachdev1,lee3}.
232: This is a direct consequence of the fact that singly quantized vortices are
233: always energetically favorable\cite{abrikosov1,fetter1}.
234:
235: As far as the electronic excitations are concerned, the spinon vortex is
236: virtually indistinguishable from the vortex in a conventional weak coupling
237: mean field
238: theory: the spin gap $\Delta$, which gives rise to the gap in the electron
239: spectrum, vanishes in the core. Consequently, the vortex state based on the
240: results of Sachdev-Nagaosa-Lee (SNL) theory\cite{sachdev1,lee3} does not
241: exhibit the pseudogap in the core and suffers
242: from the same discrepancy with the experimental data as the weak coupling
243: theories\cite{soininen1,wang1,franz1,kita1} based on the conventional Fermi
244: liquid description. Moreover, no evidence exists at present
245: for stable doubly quantized holon vortices predicted by SNL. What is
246: needed to account for the experimental data is a {\em singly quantized
247: holon vortex} stable over the large portion of the superconducting phase in the
248: phase diagram of Figure \ref{fig1}. In the core of such a holon vortex the spin
249: gap $\Delta$ remains finite and leads naturally
250: to the pseudogap excitation spectrum. In what follows we show that under
251: certain conditions the free energy (\ref{fgl1}) permits precisely such
252: solution.
253:
254: The results of the SNL theory are predicated upon the assumption that the
255: ``stiffness'' $\sigma$ of the gauge field is relatively large and that singular
256: configurations in which $\nabla\times{\bf a}$ contains a full flux quantum
257: through an elementary plaquette are prohibited. Consider now a precisely
258: opposite physical situation, allowing unconstrained fluctuations in
259: ${\bf a}$. This amounts
260: to the assumption that the $f_{\rm gauge}$ term (\ref{fgauge})
261: can be neglected in (\ref{fgl1}), i.e. $\sigma\to 0$. Physically this
262: corresponds to the ``extreme type-I'' limit of the GL ``superconductor''
263: (\ref{fgl1}) with respect to fluctuations in ${\bf a}$. Based on Elitzur's
264: theorem\cite{elitzur1} Nayak\cite{nayak1} recently argued that
265: the exact local U(1) symmetry of the model cannot be broken, implying absence
266: of the phase stiffness term (\ref{fgauge}) at all energy scales. Our assumption
267: therefore appears reasonable and in Section III. we shall give a more thorough
268: discussion of the significance of the $f_{\rm gauge}$ term for the vortex
269: solutions of interest here. For the time being we shall assume that
270: $f_{\rm gauge}$ can be neglected and explore physical consequences of the
271: resulting theory.
272:
273: $f_{\rm GL}$ given by Eq.\ (\ref{fgl1}) is quadratic in ${\bf a}$ and with
274: the $\nabla\times{\bf a}$ term absent the gauge
275: fluctuations can be trivially integrated out. Within the closely related
276: microscopic model this procedure has been recently implemented by
277: Lee\cite{dhlee1}. The resulting effective free energy density reads
278: %
279: \begin{eqnarray}
280: f&=& f_{\rm amp} +{\rho_\Delta^2\rho_b^2\over 4\rho_\Delta^2+\rho_b^2}
281: (\nabla\phi-2\nabla\theta+2e{\bf A})^2 \nonumber \\
282: &+&{1\over 8\pi}(\nabla\times{\bf A})^2,
283: \label{feff}
284: \end{eqnarray}
285: %
286: where we have set $\Delta=\rho_\Delta e^{i\phi}$, $b=\rho_b e^{i\theta}$, and
287: %
288: \begin{eqnarray}
289: f_{\rm amp}&=&(\nabla\rho_\Delta)^2 +{r_\Delta}\rho_\Delta^2 +
290: {1\over 2}{u_\Delta}\rho_\Delta^4\nonumber \\
291: &+&(\nabla\rho_b)^2 +r_b\rho_b^2 +{1\over 2} u_b\rho_b^4
292: +v\rho_\Delta^2\rho_b^2
293: \label{famp}
294: \end{eqnarray}
295: %
296: is the amplitude piece. The most important feature of the effective free
297: energy (\ref{feff}) is that it no longer depends on the individual phases
298: $\phi$ and $\theta$ but only on their particular combination
299: %
300: \begin{equation}
301: \Omega=\phi-2\theta.
302: \label{omega}
303: \end{equation}
304: %
305: Since the physical superconducting order parameter $\Psi=\Delta^*b^2=
306: \rho_\Delta\rho_b^2 e^{-i(\phi-2\theta)}$ it is reasonable
307: to identify $\Omega$ with the phase of a {\em Cooper pair}. Physically, the
308: unconstrained fluctuations of the gauge field in Eq.\ (\ref{fgl1}) resulted in
309: partial restoration of the original electronic degrees of freedom in Eq.\
310: (\ref{feff}). In the underlying microscopic model this means that on long
311: length scales spinons and holons are always confined, in agreement with
312: Elitzur's theorem \cite{elitzur1,nayak1}. On lengthscales shorter than
313: the confinement length, such as inside the vortex core, spinons and holons
314: can still appear locally decoupled. In the present effective theory
315: this aspect is reflected by two amplitude degrees of freedom
316: present in (\ref{feff}). More detailed discussion of these issues is given in
317: Refs.\ \cite{dhlee1,nayak1}.
318:
319: We have thus arrived at an effective theory of a spin-charge separated
320: system containing one phase degree
321: of freedom $\Omega$ and two amplitudes, $\rho_\Delta$ and $\rho_b$. Deep in the
322: superconducting phase,
323: where both amplitudes are finite, the physics of (\ref{feff}) will be very
324: similar to that of a conventional GL theory. In the situations where
325: the superconducting order parameter $\Psi$ is strongly suppressed, such as
326: in the vortex core, near an impurity or a wall, the new theory has an extra
327: degree of richness, associated with the fact that it is sufficient (and
328: generally preferred by the energetics) when only {\em one}
329: of the two amplitudes
330: is suppressed. Since the two amplitudes play very different roles in the
331: electronic excitation spectrum, the effective theory (\ref{feff})
332: will lead to a number of nontrivial effects.
333:
334: To illustrate this consider what will happen in the core of a superconducting
335: vortex. Under the
336: influence of the magnetic field the phase $\Omega$ will develop a singularity
337: such that $\nabla\Omega \sim 1/r$ close to the vortex center. For
338: the free energy to remain finite the amplitude prefactor in the second
339: term of Eq.\ (\ref{feff}) must vanish for $r\to 0$.
340: This is analogous to $|\Psi|$ vanishing
341: in the core of a conventional vortex. In the present case,
342: however, it is sufficient when the product $\rho_\Delta\rho_b$ vanishes. Since
343: suppressing any of the two amplitudes costs condensation energy, in general
344: only one amplitude will be driven to zero. Which of the two is suppressed
345: will be determined by the energetics of the amplitude term (\ref{famp}).
346: On general grounds we expect that the state in the vortex core will be
347: the same as the corresponding bulk ``normal'' state obtained by raising
348: temperature above $T_c$. Thus, very crudely, we expect that holon vortex
349: will be stable in the underdoped while the spinon vortex will be stable in the
350: overdoped region of the phase diagram Figure \ref{fig1}.
351:
352: An important point by which our approach differs from the SNL theory
353: is that in the present theory {\em both} types of vortices carry the
354: {\em same} superconducting flux quantum $\Phi_0$ and thus compete
355: on equal footing. This is a direct consequence of our assumption of
356: the vanishing phase stiffness $\sigma$.
357:
358: In what follows we study in detail the vortex solutions of the free energy
359: (\ref{feff}). Our main objective is to obtain the precise estimates for
360: the energy of the two types of vortices as a function of temperature and
361: doping and deduce the corresponding phase diagram for the state inside
362: the vortex core. We show that for generic parameters in (\ref{feff})
363: the singly quantized holon vortex with a pseudogap spectrum in the core can
364: be stabilized over a large portion of the superconducting phase, as
365: required by the experimental constraints discussed above.
366:
367:
368:
369: \section{Solution for a single vortex}
370:
371: \subsection{General considerations}
372:
373: In order to provide a more quantitative discussion we now adopt some
374: assumptions
375: about the coefficients entering the free energy (\ref{feff}). We assume that
376: %
377: \begin{equation}
378: r_i=\alpha_i(T-T_i), \ \ i=b,\Delta,
379: \label{ri}
380: \end{equation}
381: %
382: where $T_i$ are corresponding ``bare'' critical temperatures, which we
383: assume depend on doping concentration $x$ in the following way:
384: %
385: \begin{equation}
386: T_\Delta = T_0(2x_m-x), \ \ \
387: T_b = T_0x.
388: \label{tc}
389: \end{equation}
390: %
391: Here $x_m$ denotes the optimal doping and $T_0$ sets the overall temperature
392: scale. We furthermore assume that $u_i$ and $v$ are all positive and
393: independent of doping and temperature. It is easy to see that such choice
394: of parameters qualitatively reproduces the bulk phase diagram of cuprates
395: in the $x$-$T$ plane shown in Figure \ref{fig1}.
396: The effect of the $v$-term is to suppress $T_c$ from
397: its bare value away from the optimal doping. In real systems fluctuations
398: will lead to additional suppression of $T_c$ which we do not consider here.
399:
400: In the absence of perturbations the bulk values of the amplitudes are
401: given by
402: %
403: \begin{eqnarray}
404: \bar{\rho}_\Delta^2 &=& -({r_\Delta} u_b-r_bv)/D, \nonumber \\
405: \bar{\rho}_b^2 &=& -(r_b{u_\Delta} -{r_\Delta} v)/D,
406: \label{rho}
407: \end{eqnarray}
408: %
409: with $D=u_b{u_\Delta}-v^2$.
410: In analogy with conventional GL theories we may define coherence
411: lengths for the two amplitudes\cite{sachdev1}
412: %
413: \begin{eqnarray}
414: \xi_\Delta^{-2} &=& -({r_\Delta}-r_bv/u_b), \nonumber \\
415: \xi_b^{-2} &=& -(r_b-{r_\Delta} v/{u_\Delta}),
416: \label{xi}
417: \end{eqnarray}
418: %
419: one of which always diverges at $T_c$ as $(T-T_c)^{-1/2}$.
420:
421: Minimization of the free energy (\ref{feff}) with respect to the vector
422: potential ${\bf A}$ yields an equation
423: %
424: \begin{equation}
425: \nabla\times\nabla\times{\bf A}=e\rho_s(\nabla\Omega-2e{\bf A}),
426: \label{lona}
427: \end{equation}
428: %
429: where
430: %
431: \begin{equation}
432: \rho_s={4\rho_\Delta^2\rho_b^2\over 4\rho_\Delta^2+\rho_b^2}
433: \label{rhos}
434: \end{equation}
435: %
436: is the effective superfluid density. The term in brackets can be
437: identified as twice the conventional superfluid velocity
438: %
439: $${\bf v}_s={1\over 2}\nabla\Omega-e{\bf A}.$$
440: %
441: Making use of the Ampere's law
442: $4\pi{\bf j}=\nabla\times{\bf B}$ we see that Eq.\ (\ref{lona}) specifies
443: the supercurrent in terms superfluid density and velocity:
444: ${\bf j}=2e\rho_s{\bf v}_s$. Minimization of (\ref{feff}) with respect
445: to $\Omega$ then implies $\nabla\cdot{\bf j}=0$; the supercurrent is
446: conserved.
447:
448: Minimizing the free energy (\ref{feff}) with respect to the amplitudes results
449: in the pair of coupled GL equations:
450: %
451: \begin{mathletters}
452: \label{gl:all}
453: \begin{equation}
454: -\nabla^2\rho_\Delta + {r_\Delta}\rho_\Delta + {u_\Delta}\rho_\Delta^3 + v\rho_b^2\rho_\Delta
455: +{4\rho_\Delta^2\rho_b^2\over (4\rho_\Delta^2+\rho_b^2)^2}{\bf v}_s^2 = 0, \label{gl:a}
456: \end{equation}
457: \begin{equation}
458: -\nabla^2\rho_b + r_b\rho_b + u_b\rho_b^3 + v\rho_\Delta^2\rho_b
459: +{16\rho_\Delta^2\rho_b^2\over (4\rho_\Delta^2+\rho_b^2)^2}{\bf v}_s^2 = 0. \label{gl:b}
460: \end{equation}
461: \end{mathletters}
462: We are interested in the behavior of the amplitudes in the vicinity of the
463: vortex center. In this region, for a strongly type-II superconductor, we
464: may neglect the vector potential {{\bf A}}
465: in the superfluid velocity ${\bf v}_s$. In
466: a singly quantized vortex $\Omega$ winds by $2\pi$ around the origin
467: leading to a singularity
468: of the form ${\bf v}_s\simeq{1\over 2}\nabla\Omega =\hat\varphi/2r$.
469: First, for the {\em holon} vortex we assume that
470: $\rho_b$ vanishes in the core as some power $\rho_b(r)\sim r^\nu$ and
471: $\rho_\Delta(r)\approx\bar\rho_\Delta$
472: remains approximately constant. Eq.\ (\ref{gl:b}) then becomes
473: %
474: \begin{equation}
475: ({1\over 4}-\nu^2)r^{\nu-2} + (r_b+v\bar\rho_\Delta^2)r^\nu +
476: u_b\bar\rho_b^2r^{3\nu}=0,
477: \label{glcore}
478: \end{equation}
479: %
480: where we have neglected $\rho_b^2(r)$ compared to $4\bar\rho_\Delta^2$
481: in the denominator of the last term in Eq.\ (\ref{gl:b}). The most singular
482: term in Eq.\ (\ref{glcore}) is the first one and we must demand that the
483: coefficient of $r^{\nu-2}$ vanishes. This implies $\nu={1\over 2}$. The
484: asymptotic short distance behavior of the holon amplitude therefore
485: can be written as
486: %
487: \begin{equation}
488: \rho_b(r)\simeq c_b\bar\rho_b \left(r\over\xi_b\right)^{1/2},
489: \label{rhobcore}
490: \end{equation}
491: %
492: where $c_b$ is a constant of order unity which may be determined
493: by the full integration of Eqs.\ (\ref{gl:all}).
494: Similar analysis of Eq.\ (\ref{gl:a}) in the vicinity of the {\em spinon}
495: vortex yields
496: %
497: \begin{equation}
498: \rho_\Delta(r)\simeq c_\Delta\bar\rho_\Delta \left(r\over\xi_d\right),
499: \label{rhodcore}
500: \end{equation}
501: %
502: with $\rho_b$ approximately constant.
503:
504: We notice the different power laws
505: in the holon and spinon results. Operationally this difference arises from
506: different numerical prefactors of the respective superfluid velocity terms
507: in Eqs.\ (\ref{gl:all}). Physically, the unusual $r$ dependence of the
508: holon amplitude in the core reflects the fact that the field $b$ describes
509: a condensate of single holons, each carrying charge $e$. Superconducting
510: vortex with the flux quantum $\Phi_0$ represents a magnetic
511: ``half-flux'' for the
512: holon field which results in non-analytic behavior of $\rho_b(r)$ at the
513: origin. Singly quantized holon vortex is therefore a peculiar object and we
514: shall discuss it more fully in Section III. Here we note that the
515: physical superconducting order parameter amplitude $|\Psi|=\rho_\Delta\rho_b^2$
516: remains analytic in the core of both the spinon and the holon
517: vortex.
518:
519:
520: \subsection{Holon vs. spinon vortex: the phase diagram}
521:
522: We are now in the position to estimate the energies of the two types
523: of vortices and deduce the phase diagram for the ``normal'' state in the
524: vortex core. To this end we consider a single isolated vortex centered
525: at the origin.
526: The total vortex line energy can be divided into electromagnetic
527: and core contributions\cite{fetter1}.
528: The electromagnetic contribution
529: consists of the energy of the supercurrents and the magnetic
530: field outside the core region. It may be estimated
531: by assuming that the amplitudes $\rho_\Delta$ and $\rho_b$
532: have reached their bulk values $\bar\rho_\Delta$ and $\bar\rho_b$ respectively.
533: Taking curl of Eq.\ (\ref{lona}) and noting
534: that $\nabla\times\nabla\Omega=2\pi\delta({\bf r})$ for a singly quantized vortex
535: we obtain the London equation for the magnetic field
536: ${\bf B}=\nabla\times{\bf A}$ of the form
537: %
538: \begin{equation}
539: B-\lambda^2\nabla^2B=\Phi_0\delta({\bf r})
540: \label{lon}
541: \end{equation}
542: %
543: where
544: %
545: \begin{equation}
546: \lambda^{-2}=8\pi e^2
547: {4\bar\rho_\Delta^2\bar\rho_b^2\over 4\bar\rho_\Delta^2+\bar\rho_b^2}.
548: \label{lam}
549: \end{equation}
550: %
551: has the meaning of the London penetration depth for the effective GL theory
552: (\ref{feff}). Aside from the unusual form of $\lambda$, Eq.\ (\ref{lon})
553: is identical to the conventional London equation.
554: The corresponding electromagnetic energy is therefore
555: the same for both types of vortices
556: and can be calculated in the usual manner\cite{abrikosov1,fetter1,sachdev1}
557: obtaining
558: %
559: \begin{equation}
560: E_{\rm EM}\simeq\left({\Phi_0\over 4\pi\lambda}\right)^2 \ln\kappa,
561: \label{em}
562: \end{equation}
563: %
564: with $\kappa=\lambda/\max(\xi_\Delta,\xi_b)$ being the generalized GL ratio.
565:
566: To estimate the core contribution to the vortex line energy we assume that
567: one of the amplitudes is suppressed to zero in the core
568: %
569: \begin{equation}
570: \rho_i(r)=0, \ \ \ r<\xi_i,
571: \label{rhor}
572: \end{equation}
573: %
574: while the other one stays constant and equal to its bulk value. This is a
575: very
576: crude approximation which we justify below by an exact numerical computation.
577: With these assumptions, the core energy is
578: %
579: \begin{equation}
580: E_{\rm core}^{(i)}\simeq\left({\Phi_0\over 4\pi\lambda_i}\right)^2,
581: \label{ecore}
582: \end{equation}
583: %
584: where $i=\Delta,b$ for spinon and holon vortex respectively and
585: %
586: \begin{equation}
587: \lambda^{-2}_i=8\pi e^2 \bar\rho_i^2.
588: \label{lami}
589: \end{equation}
590: %
591: Such a crude approximation overestimates the core energy. A more accurate
592: analysis\cite{abrikosov1,fetter1}, which we do not pursue here,
593: allows for a more realistic variation of $\rho_i(r)$ in the core and
594: indicates that the
595: value of $E_{\rm core}^{(i)}$ has the same form as Eq.\ (\ref{ecore})
596: multiplied by a numerical factor $c_1\approx 0.5$\cite{hu1,alama1}.
597: Thus, the total energy
598: of the vortex line can be written as
599: %
600: \begin{equation}
601: E^{(i)}=\left({\Phi_0\over 4\pi\lambda}\right)^2\ln\kappa
602: +c_1\left({\Phi_0\over 4\pi\lambda_i}\right)^2,
603: \label{evortex}
604: \end{equation}
605: %
606: where again $i=\Delta,b$ for spinon and holon vortex respectively.
607: Eq.\ (\ref{evortex}) parallels the Abrikosov expression for the vortex line
608: energy in a conventional GL theory\cite{abrikosov1}
609: where $\lambda$ and $\lambda_i$ are identical and equal to the ordinary
610: London penetration depth.
611: %
612: \begin{figure}[t]
613: \epsfxsize=8.5cm
614: \epsffile{fig2.ps}
615: \caption[]{Vortex core phase diagram for GL parameters chosen as follows:
616: $\alpha_\Delta=0.13$, $\alpha_b=0.10$, $T_0=200$K, $x_m=0.2$,
617: $u_\Delta=u_b=1.0$ and $v=0.5$.
618: Dashed line marks the phase boundary $T_g(x)$ obtained from
619: Eq.\ (\ref{tg}) while the solid circles correspond to the
620: numerical calculation with the same parameters. }
621: \label{fig2}
622: \end{figure}
623: %
624:
625: In the vortex state described by the free energy (\ref{feff}) the vortex with
626: lower energy $E^{(i)}$ will be stabilized. Eq.\ (\ref{evortex}) implies that
627: the difference in energy between the two types of vortices comes primarily
628: from the
629: core contribution, as expected on the basis of the physical argument presented
630: above. Condition $\lambda_\Delta=\lambda_b$ marks the
631: transition point between the two solutions. For fixed GL parameters $T_0$,
632: $x_m$, $\alpha_i$, $u_i$ and $v$ this defines a transition line in the $x$-$T$
633: plane. According to (\ref{lami}) the equation for this line is
634: %
635: \begin{equation}
636: \bar\rho_\Delta(x,T)=\bar\rho_b(x,T).
637: \label{tran}
638: \end{equation}
639: %
640: Using Eqs.\ (\ref{ri}-\ref{rho}) one can obtain an explicit expression
641: for the transition temperature $T_g$ between two types of vortices as
642: a function of doping
643: %
644: \begin{equation}
645: T_g(x)=T_0\left[{2x_m-x\over 1-\beta}+{x\over 1-\beta^{-1}}\right],
646: \label{tg}
647: \end{equation}
648: %
649: with
650: %
651: \begin{equation}
652: \beta={\alpha_b(u_\Delta+v)\over \alpha_\Delta(u_b+v)}.
653: \label{beta}
654: \end{equation}
655: %
656: Eq.\ (\ref{tg}) describes a straight line in the $x$-$T$ plane, originating
657: at $[x_m,T_0x_m]$, i.e. maximal $T_c$ at optimum doping, and terminating at
658: $[2x_m/(1+\beta),0]$. Generically, we expect that parameters $\alpha_i$ and
659: $u_i$ will be comparable in magnitude for the holon and spinon channels.
660: Parameter $\beta$ defined in Eq.\ (\ref{beta}) will therefore be of order
661: unity. The typical situation for $\beta=0.77$
662: is illustrated in Figure \ref{fig2}.
663: More generally the quartic coefficients $u_i$ and $v$ could exhibit weak
664: doping and temperature dependences leading to a curvature in the phase
665: boundary.
666:
667: The appealing feature of the present theory is that parameter
668: $\beta$ may vary from compound to compound. Thus, the experimental fact
669: that in BSCCO the pseudogap in the core persists
670: into the overdoped region is easily accounted for in the present theory.
671: It would be interesting to see if the transition
672: from holon to spinon vortex as a function of doping could be experimentally
673: observed. A good candidate for such observation would be LSCO,
674: where the transport measurements in pulsed magnetic fields\cite{ando1}
675: established a metal-insulator transition around optimal doping, i.e.\
676: $\beta\approx 1$. The current
677: theory predicts a holon vortex with the pseudogap spectrum in the
678: underdoped (insulating) region and spinon vortex with conventional metallic
679: spectrum on the overdoped side.
680:
681:
682: \subsection{Numerical results}
683:
684: In order to put the above analytical estimates on firmer ground we now pursue
685: numerical computation of the vortex line energy. For simplicity we
686: consider the strongly type-II situation $(\kappa\gg1)$ where the vector
687: potential term in ${\bf v}_s$ can be neglected to an excellent approximation,
688: as long as we focus on the behavior close to the core.
689: We are then faced with the task of numerically minimizing
690: the free energy (\ref{feff}) with respect to the two cylindrically
691: symmetric amplitudes $\rho_\Delta(r)$ and $\rho_b(r)$. As noted by Sachdev
692: \cite{sachdev1} direct numerical minimization of the free energy (\ref{feff})
693: provides a more robust solution than the numerical
694: integration of the coupled differential equations (\ref{gl:all}).
695:
696: We discretize the free energy functional (\ref{feff}) on a disk of a
697: radius $R\gg\xi_i$ in the radial coordinate $r$ with up to $N= 2000$
698: spatial points.
699: We then employ the Polak-Ribiere variant of the Conjugate Gradient Method
700: \cite{numrec} to minimize this discretized functional with respect to
701: $\rho_\Delta(r_j)$ and $\rho_b(r_j)$, initialized to suitable single vortex trial
702: functions. The procedure converges very rapidly and the results are
703: insensitive to the detailed shape of the trial functions as long as they
704: saturate to the correct bulk values outside the vortex core.
705:
706: Typical results of our numerical computations are displayed in Figure
707: (\ref{fig3}) and are in complete agreement with the analytical
708: considerations of the preceding subsections. Note in particular that
709: $\rho_b(r)$ in the holon vortex vanishes with infinite slope, consistent
710: with Eq.\ (\ref{rhobcore}). Plotting $\rho_b^2(r)$ confirms that the
711: exponent is indeed $1/2$. In the spinon vortex
712: $\rho_\Delta(r)$ is seen to vanish linearly as expected on the basis of
713: Eq.\ (\ref{rhodcore}). The nonvanishing order parameter is slightly
714: elevated in the core reflecting the effective ``repulsion'' between the
715: two amplitudes contained in the $v$-term of the free energy. The results
716: for the spinon vortex are consistent with those of Ref.\ \cite{sachdev1}.
717: %
718: \begin{figure}[t]
719: \epsfxsize=8.5cm
720: \epsffile{fig3.ps}
721: \caption[]{Order parameter amplitudes near a single isolated vortex for
722: GL parameters specified in Figure \ref{fig2}.
723: The holon vortex is plotted for $T=0$ and
724: $x=0.22$ (implying coherence lengths $\xi_\Delta=0.63$ and $\xi_b=0.70$),
725: while the spinon vortex is plotted for $T=0$ and $x=0.24$ (implying
726: $\xi_\Delta=0.75$ and $\xi_b=0.60$).}
727: \label{fig3}
728: \end{figure}
729: %
730:
731: We explored a number of other parameter configurations and obtained
732: similar results.
733: We find that Eq.\ (\ref{tran}) is a good predictor of the
734: transition line between the holon and spinon vortex, although the precise
735: numerical value of the transition temperature $T_g$ for given $x$ tends
736: to deviate slightly from the value predicted by Eq.\ (\ref{tg}). This is
737: illustrated in Figure (\ref{fig2}) where we compare the vortex core
738: phase diagrams obtained numerically and from Eq.\ (\ref{tg}). Interestingly,
739: the deviation always tends to enlarge the holon vortex sector of the phase
740: diagram at the expense of the spinon vortex sector. This is presumably
741: because the sharper $\sim\sqrt{r}$ suppression of the holon order parameter
742: in the core costs less condensation energy.
743:
744:
745:
746: \section{Gauge fluctuations and the spectral properties in the core}
747:
748: Theory of the vortex core based on the effective action (\ref{feff})
749: appears to yield results consistent with the STM data on cuprates
750: \cite{renner1,pan2} in that it implies stable holon vortex solution over
751: the large portion of the superconducting phase diagram. The state inside the
752: core of such a holon vortex is characterized by vanishing amplitude of the
753: holon condensate field,
754: $|b|=0$, and a finite spin gap $|\Delta|\approx\Delta_{\rm bulk}$. This is the
755: same state as in the pseudogap region above $T_c$. One would thus expect
756: the electronic spectrum in the core to be similar to that found in the normal
757: state of the underdoped cuprates, in agreement with the
758: data\cite{renner1,pan2}. The holon vortex
759: with this property carries conventional superconducting flux quantum
760: $\Phi_0$, in accord with experiment. This general agreement between theory
761: and experiment would suggest that the effective action (\ref{feff})
762: provides the sought for phenomenological description of the vortex core
763: physics in cuprates.
764: In what follows we amplify our argumentation that it is also
765: tenable in a broader theoretical context in that it naturally
766: follows from the U(1) slave boson models extensively studied in the
767: classic and more recent high-$T_c$ literature. We then provide a more detailed
768: discussion of the vortex core spectra and propose an explanation for the
769: experimentally observed core bound states.
770:
771:
772: \subsection{Significance of the $f_{\rm gauge}$ term}
773:
774: Derivation of the effective action (\ref{feff}) from the more general U(1)
775: action (\ref{fgl1}) hinges on our assumption that the stiffness $\sigma$
776: of the
777: gauge field ${\bf a}$ is low and that the $f_{\rm gauge}$ term (\ref{fgauge})
778: can be neglected. Assumption of large $\sigma$ by SNL leads to very different
779: vortex solutions\cite{sachdev1,lee3} which appear inconsistent with the recent
780: experimental data. We first expand on our discussion as to why is
781: $f_{\rm gauge}$ term important and then we argue why it may be permissible
782: to neglect it in the realistic models of cuprates.
783:
784: To facilitate the discussion let us rewrite Eq.\ (\ref{fgl1}) by resolving
785: the complex matter fields into amplitude and phase components:
786: %
787: \begin{eqnarray}
788: f_{\rm GL}&=& f_{\rm amp} +\rho_\Delta^2(\nabla\phi -2{\bf a})^2
789: +\rho_b^2(\nabla\theta-{\bf a}-e{\bf A})^2 \nonumber \\
790: &+&{1\over 8\pi}(\nabla\times{\bf A})^2 +
791: {\sigma\over2}(\nabla\times{\bf a})^2,
792: \label{fgl2}
793: \end{eqnarray}
794: %
795: with $f_{\rm amp}$ specified by Eq.\ (\ref{famp}). Now consider situation
796: in which the sample is subjected to uniform magnetic field
797: ${\bf B}=\nabla\times{\bf A}$. Two scenarios (discussed previously by SNL)
798: appear possible. In the first, the internal gauge field develops no net
799: flux, $\langle\nabla\times{\bf a}\rangle=0$, and the holon phase
800: $\theta$ develops
801: singularities in response to ${\bf A}$ such that
802: %
803: $$\nabla\times\nabla\theta=2\pi\sum_j\delta({\bf r}-{\bf r}_j),$$
804: %
805: where ${\bf r}_j$ denotes the vortex positions. The holon amplitude $\rho_b$
806: is driven to zero at ${\bf r}_j$, essentially to prevent the free energy from
807: diverging due to the singularity in the phase gradient.
808: Since holons carry charge $e$, each vortex is threaded by
809: flux $hc/e$, i.e.\ twice the superconducting flux quantum $\Phi_0=hc/2e$.
810: This solution represents the doubly quantized holon vortex lattice, considered
811: by SNL.
812:
813: In the second scenario
814: ${\bf a}$ develops a net flux such that ${\bf a}\approx -e{\bf A}$, which screens out the
815: ${\bf A}$ field in the holon term but produces a net flux $-2e{\bf A}$ in the
816: spinon term. In response to this flux, spinon phase $\phi$ develops
817: singularities such that
818: %
819: $$\nabla\times\nabla\phi=2\pi\sum_j\delta
820: ({\bf r}-\tilde{\bf r}_j),$$
821: %
822: corresponding to the spinon vortex lattice. $\tilde{\bf r}_j$ denotes vortex
823: positions which will be different from ${\bf r}_j$ since at the fixed field
824: $B$ there will be twice as many spinon vortices as holon vortices.
825: (Spinon vortices carry conventional superconducting quantum of flux $\Phi_0$.)
826: In this case $\rho_\Delta$ is driven to zero at the vortex centers.
827: In this scenario
828: one pays a penalty for nucleating the net flux in $\nabla\times{\bf a}$ due
829: to last term in Eq.\ (\ref{fgl2}). This energy cost can be estimated as
830: %
831: \begin{equation}
832: E_\sigma\simeq 8\pi\sigma e^2 \left({\Phi_0\over 4\pi\lambda}\right)^2
833: \label{esig}
834: \end{equation}
835: %
836: per vortex. Stiffness $\sigma$ must be small enough so that $E_\sigma$ is
837: small compared to the vortex energy (\ref{evortex}). Taking the dominant
838: $E_{\rm EM}$ term and neglecting $\ln\kappa$ this implies that
839: %
840: \begin{equation}
841: \sigma\ll {1\over 8\pi e^2},
842: \label{sig}
843: \end{equation}
844: %
845: which is the same condition as considered in Ref.\ \cite{sachdev1}.
846:
847:
848: Now consider a {\em third} scenario in which a {\em singly
849: quantized} holon vortex emerges. As a starting point consider the spinon
850: vortex solution just described. In the underdoped regime the amplitude piece
851: $f_{\rm amp}$ would favor suppressing the holon amplitude in the core
852: instead of the spinon amplitude but according to our previous considerations
853: this would ordinarily require formation of a doubly quantized vortex whose
854: magnetic energy is too large. However, if the gauge field stiffness $\sigma$
855: is sufficiently small, the system could lower its free energy by
856: setting up singularities in ${\bf a}$ which would precisely cancel the
857: singularities in $\nabla\phi$ and shift them to the holon term.
858: To arrive at this situation imagine
859: contracting the initially uniform flux $\nabla\times{\bf a}$ so that it
860: becomes localized in the individual vortex core regions. Taking this
861: procedure to the extreme, i.e. taking the limit $\sigma\to 0$,
862: the gauge field will form ``flux spikes'' of the form
863: %
864: \begin{equation}
865: 2(\nabla\times{\bf a})
866: =-\nabla\times\nabla\phi=-2\pi\sum_j\delta({\bf r}-\tilde{\bf r}_j),
867: \label{sing}
868: \end{equation}
869: %
870: completely localized at the vortex centers.
871: Gauge field of this form indeed completely cancels the singularities in the
872: spinon
873: phase gradient in Eq.\ (\ref{fgl2}) and $\rho_\Delta$ is no longer forced
874: to vanish in the core. The singularities now appear in the holon term,
875: but they stem from ${\bf a}$ rather that $\nabla\theta$ which remains
876: nonsingular. Consequently, $\rho_b$ is forced to vanish in the vortex cores.
877: By construction the vortices are located at $\tilde{\bf r}_j$ and
878: are therefore singly quantized.
879: This is the singly quantized holon vortex
880: discussed in the framework of the free energy (\ref{feff}). Based on the
881: above discussion the singly quantized holon vortex can be thought of as a
882: composite object formed by attaching half quantum $(h/2)$ of the
883: fictitious gauge flux $\nabla\times{\bf a}$ to the spinon vortex. Within
884: the full compact U(1) theory this is essentially equivalent to the Z$_2$
885: vortex discussed by Wen\cite{wen2} in the framework of topological orders
886: in spin liquids.
887:
888: In the framework of the free energy (\ref{fgl2}) one pays a penalty for such
889: a singular solution due to the gauge stiffness term. In the present
890: continuum model
891: this penalty per single vortex is actually infinite, since according to
892: Eq.\ (\ref{sing}) it involves a
893: spatial integral over $[\delta({\bf r}-\tilde{\bf r}_j)]^2$. Thus, in the
894: continuum model
895: the singular solutions of this type are prohibited. In reality, however,
896: we have to recall that our effective action (\ref{fgl1}) descended from
897: a microscopic
898: lattice model for spinons and holons in which the gauge field ${\bf a}$ lives
899: on the nearest neighbor bonds of the ionic lattice. The ionic lattice constant
900: $d$ therefore provides a natural short distance cutoff and the delta
901: function in Eq.\ (\ref{sing}) should be interpreted as a flux quantum $\Phi_0$
902: piercing an elementary plaquette of the lattice. The energy cost per vortex
903: thus becomes finite and is given by
904: %
905: \begin{equation}
906: E'_\sigma\simeq {\sigma e^2\over 2} \left({\Phi_0\over d}\right)^2.
907: \label{esigp}
908: \end{equation}
909: %
910: Again, for the solution to be stable, $E'_\sigma$ must be negligible compared
911: to the vortex energy (\ref{evortex}). This implies
912: %
913: \begin{equation}
914: \sigma\ll {1\over 8\pi^2e^2} \left({d\over \lambda}\right)^2,
915: \label{sigp}
916: \end{equation}
917: %
918: which is a much more stringent condition than (\ref{sig}) since in cuprates
919: $d\ll\lambda$.
920:
921: When condition (\ref{sigp}) is satisfied it is permissible to neglect
922: the $f_{\rm gauge}$ term in the effective action (\ref{fgl1}) and
923: it becomes fully equivalent to (\ref{feff}) as far as the vortex solutions are
924: concerned.
925: Eq.\ (\ref{sigp}) gives the precise meaning to the requirement of the weak
926: stiffness of the gauge field loosely stated when deriving the effective
927: action (\ref{feff}).
928:
929:
930:
931: \subsection{Microscopic considerations}
932:
933: As mentioned in the introduction, the gauge field ${\bf a}$ has no dynamics
934: in the original U(1) microscopic model, as it only serves
935: to enforce a constraint on spinons and holons. The stiffness term
936: (\ref{fgauge}) in the effective
937: theory was assumed to arise in the process of integrating out the
938: microscopic degrees of freedom\cite{sachdev1,lee3}. While such term is
939: certainly permitted by symmetry, assessing its strength
940: $\sigma$ is a nontrivial issue since even deep in the superconducting phase
941: neither holons nor spinons are truly gapped. Thus, in general, integrating
942: out these degrees of freedom may lead to singular and nonlocal interactions
943: between the condensate and the gauge fields. To our knowledge the procedure
944: has not been explicitly performed for the U(1) model and the precise form or
945: magnitude of the gauge stiffness term is unknown. General
946: considerations\cite{nayak1} suggest
947: that the gauge stiffness term is negligible in the class
948: of models with exact local U(1) symmetry connecting the phases of holons
949: and spinons.
950:
951: Consider now an intermediate representation of the problem where only
952: high energy microscopic degrees of freedom have been integrated out. In the
953: presence of a cutoff this is a well defined procedure even for gapless
954: excitations, as explicitly shown by Kwon and Dorsey\cite{kwon1} for a
955: simple BCS model. The corresponding effective Lagrangian density of the
956: present U(1) model can be written as
957: %
958: \begin{eqnarray}
959: {\cal L}_{\rm eff}&=&
960: {\kappa_\Delta^\mu\over 2}(\partial_\mu\phi-2a_\mu)^2
961: +{\kappa_b^\mu\over 2}(\partial_\mu\theta-a_\mu-eA_\mu)^2
962: -f_{\rm amp} \nonumber \\
963: &+&(\partial_\mu\phi-2a_\mu)J_{\rm sp}^\mu
964: +(\partial_\mu\theta-a_\mu-eA_\mu)J_h^\mu \nonumber \\
965: &+&{\cal L}_{\rm sp}[\psi_{\rm sp},\psi_{\rm sp}^\dagger;\rho_\Delta]
966: +{\cal L}_h[\psi_h,\psi_h^\dagger;\rho_b] +{\cal L}_{\rm EM}[A_\mu].
967: \label{leff}
968: \end{eqnarray}
969: %
970: The Greek index $\mu$ runs over time and two spatial dimensions,
971: $\kappa_i^0$ are compressibilities of the holon and spinon condensates,
972: while
973: %
974: \begin{equation}
975: \kappa_i^j=-2(\rho_i)^2,\ \ \ i=\Delta, b, \ \ \ j=1,2,
976: \label{kappa}
977: \end{equation}
978: %
979: are the
980: respective phase stiffnesses. $J_{\rm sp}^\mu$ and $J_h^\mu$ are spinon and
981: holon three currents respectively and ${\cal L}_{\rm sp}$ and
982: ${\cal L}_h$ are the low energy effective Lagrangians for
983: the fermionic spinon field $\psi_{\rm sp}$ and bosonic holon field $\psi_h$.
984: ${\cal L}_{\rm EM}$ is the Maxwell Lagrangian for the physical electromagnetic
985: field. Thus, ${\cal L}_{\rm eff}$ describes
986: an effective low energy theory of spinons and holons coupled to their
987: respective collective modes and a fluctuating U(1) gauge field. Similar theory
988: has been recently considered by Lee\cite{dhlee1}.
989:
990: The precise form of the microscopic Lagrangians ${\cal L}_{\rm sp}$ and
991: ${\cal L}_h$ is not important for our discussion. The salient feature which
992: we exploit here is that only the amplitude of the respective condensate field
993: enters into ${\cal L}_{\rm sp}$ and ${\cal L}_h$. Coupling to the phases and
994: the gauge field is contained entirely in the Doppler shift terms [second line
995: of Eq.\ (\ref{leff})]. Such form of the coupling is largely dictated by the
996: requirements of the gauge invariance and the particular form Eq.\ (\ref{leff})
997: can be explicitly derived by gauging away the respective phase factors
998: from the $\psi$ fields\cite{balents1,kwon1}.
999:
1000: The gauge field $a_\mu$ enters the effective Lagrangian (\ref{leff}) only via
1001: two gauge invariant terms: $(\partial_\mu\phi-2a_\mu)$ and
1002: $(\partial_\mu\theta-a_\mu-eA_\mu)$, which may be interpreted as the three
1003: velocities of the spinon and holon condensates respectively.
1004: Furthermore, the only coupling between
1005: holons and spinons arises from $a_\mu$. Therefore, if we now proceed to
1006: integrate out the remaining microscopic degrees of freedom from
1007: ${\cal L}_{\rm eff}$, the two
1008: velocity terms will not mix. This consideration suggests that upon
1009: integrating out all of the microscopic degrees of freedom, the resulting
1010: gauge stiffness term will be of the form
1011: %
1012: \begin{eqnarray}
1013: f'_{\rm gauge}&=&{\sigma_\Delta\over2}[\nabla\times(2{\bf a}-\nabla\phi)]^2
1014: \nonumber \\
1015: &+&{\sigma_b\over2}[\nabla\times({\bf a}+e{\bf A}-\nabla\theta)]^2.
1016: \label{fgaugep}
1017: \end{eqnarray}
1018: %
1019: Clearly, such term is permitted by the gauge symmetry. Furthermore, we note
1020: that for smooth (i.e.\ vortex free) configurations of phases the gradient
1021: terms will contribute nothing and we recover the gauge term considered
1022: in Ref.\ \cite{lee3}.
1023:
1024: In the presence of a vortex in $\phi$ or $\theta$ the $f'_{\rm gauge}$
1025: term will contribute formally divergent energy. Regularizing
1026: this on the lattice, as discussed above Eq.\ (\ref{esigp}),
1027: this energy will become finite and can be interpreted simply as the
1028: energy of the spinon or holon vortex core states, which have been integrated
1029: out. In the microscopic theory (\ref{leff}) such energy would arise upon
1030: solving the relevant fermionic or bosonic vortex problem.
1031:
1032: We stress that, as concluded
1033: in the preceding subsection, the main theoretical obstacle to the
1034: formation of a singly quantized holon vortex in the original SNL theory
1035: was the appearance of a formally
1036: divergent contribution in the $f_{\rm gauge}$ term (\ref{fgauge}). The
1037: argument above suggests that $f_{\rm gauge}$ in Eq.\ (\ref{fgl1})
1038: should be replaced by Eq.\
1039: (\ref{fgaugep}), in which such formally divergent contribution appears for
1040: {\em arbitrary} vortex configuration and upon regularization has a simple
1041: physical interpretation in terms of the energy of the vortex core states.
1042: Usage of the physically
1043: motivated term (\ref{fgaugep}) in place of (\ref{fgauge}) therefore
1044: removes the bias against the singly quantized holon vortex solution, which
1045: appears to be realized in real materials. With (\ref{fgaugep}) any bias
1046: between the holon and spinon vortex solutions can result only from the
1047: difference between the two stiffness constants $\sigma_\Delta$ and
1048: $\sigma_b$. It is reasonable on physical grounds to assume that
1049: constants $\sigma_\Delta$ and $\sigma_b$ are of the similar magnitudes.
1050: Furthermore, on the basis of Ref.\ \cite{nayak1} we expect these constants
1051: to be negligibly small in the physically relevant models.
1052: Consequently we
1053: expect that neglecting the $f_{\rm gauge}$ term as in our derivation
1054: of effective action (\ref{feff}) will result in accurate determination
1055: of the phase diagram for the state in the vortex core.
1056:
1057:
1058: \subsection{Vortex core states}
1059:
1060: The phenomenological theory based on the effective action (\ref{feff})
1061: does not allow us to address the interesting question of the nature of
1062: the fermionic states in the vortex core. To do this we need to consider
1063: the microscopic Lagrangian density
1064: (\ref{leff}). While the fully self consistent calculation is likely
1065: to be prohibitively difficult, one can obtain qualitative insights by first
1066: solving the GL theory (\ref{feff}) as described in Sec.\ II, and then
1067: using the order parameters $\rho_\Delta$ and $\rho_b$
1068: as an input to the fermionic and bosonic sectors
1069: of the theory specified by Eq.\ (\ref{leff}). The work on a detailed
1070: solution of this type is in progress. Here we wish to point out some
1071: interesting features of such a theory and argue that it may indeed exhibit
1072: structure in the low energy spectral density similar to that found
1073: experimentally\cite{maggio1,pan2}.
1074:
1075: It is instructive to integrate out the gauge fluctuations from the
1076: Lagrangian (\ref{leff}) as first discussed by Lee\cite{dhlee1}. Since
1077: ${\cal L}_{\rm eff}$ is quadratic in $a_\mu$ the integration can be
1078: explicitly performed resulting in the Lagrangian of the form
1079: %
1080: \begin{eqnarray}
1081: {\cal L}'_{\rm eff}&=&
1082: {1\over 2}K_\mu(v_s^\mu)^2 -f_{\rm amp} + {\cal L}_{\rm EM}\nonumber \\
1083: &-& {2\kappa_b^\mu\over 4\kappa_\Delta^\mu+\kappa_b^\mu}
1084: (v_s^\mu J_{\rm sp}^\mu)
1085: +{4\kappa_\Delta^\mu\over 4\kappa_\Delta^\mu+\kappa_b^\mu}
1086: (v_s^\mu J_h^\mu)\nonumber \\
1087: &+&{\cal L}_{\rm sp} +{\cal L}_h
1088: -{1\over 2}{1\over 4\kappa_\Delta^\mu+\kappa_b^\mu}(2J_{\rm sp}^\mu+J_h^\mu)^2,
1089: \label{leffi}
1090: \end{eqnarray}
1091: %
1092: where $K_\mu=4\kappa_\Delta^\mu\kappa_b^\mu/(4\kappa_\Delta^\mu+\kappa_b^\mu)$
1093: and
1094: %
1095: \begin{equation}
1096: v_s^\mu=(\partial_\mu\theta-{1\over2}\partial_\mu\phi-eA_\mu)
1097: \label{vs}
1098: \end{equation}
1099: %
1100: is
1101: the physical superfluid velocity. The first line reproduces the GL effective
1102: action (\ref{feff}) for the condensate fields,
1103: the second line describes the Doppler shift coupling
1104: of the superfluid velocity to the microscopic currents, and the third line
1105: contains spinon and holon pieces with additional current-current interactions
1106: generated by the gauge fluctuations\cite{dhlee1}.
1107:
1108:
1109: We now discuss the physical implications of Eq.\ (\ref{leffi}) for the two
1110: types of vortices. We focus
1111: on the static solutions (i.e. we ignore the time dependences of various
1112: quantities, e.g.\ taking $v_s^0=0$) of ${\cal L}'_{\rm eff}$ in the
1113: presence of a single isolated vortex. We are interested in the local
1114: spectral function of a physical electron. This is
1115: given by a convolution in the energy variable of the spinon and holon
1116: spectral functions. According
1117: to the analysis presented in Ref.\ \cite{lee2}, at low temperatures
1118: the electron spectral function will be essentially equal to the spinon
1119: spectral function. Convolution with the
1120: holon spectral function which is dominated by the sharp coherent peak due to
1121: the condensate merely leads
1122: to a small broadening of the order $T$. In the following we therefore focus
1123: on the behavior of spinons in the vicinity of the two types of vortices.
1124:
1125:
1126: By inspecting Eq.\ (\ref{leffi}) it is easy to see that the excitations
1127: inside the {\em spinon vortex} will be qualitatively
1128: very similar to those found in the conventional vortex described by the weak
1129: coupling $d$-wave BCS theory\cite{wang1,franz1,kita1}. In particular according
1130: to Eq. (\ref{rhodcore}) we have $\kappa_\Delta\sim r^2$,
1131: and $\kappa_b\sim$ const in the core. Recalling furthermore that
1132: $|{\bf v}_s|\sim 1/r$ we observe that the spinon current ${\bf J}_{\rm sp}$
1133: is coupled to a term that diverges as $1/r$ in the core (just as in a
1134: conventional vortex), while the holon current ${\bf J}_h$ is coupled to a
1135: nonsingular term. Thus, one may conclude that holons remain essentially
1136: unperturbed by the phase singularity in the spinon vortex while the spinons
1137: obey the essentially conventional Bogoliubov-de Gennes equations for a
1138: $d$-wave vortex.
1139:
1140: In the {\em holon vortex} the situation is quite different. According to
1141: Eq. (\ref{rhobcore}) we have $\kappa_b\sim r$
1142: and $\kappa_\Delta\sim$ const in the core. The spinon current
1143: ${\bf J}_{\rm sp}$ is now
1144: coupled to a nonsingular term ($1/r$ divergence in $v_s$ is canceled by
1145: $\kappa_b\sim r$).
1146: Therefore, there will be no topological perturbation in the spinon sector
1147: and we expect the spinon wavefunctions to be essentially unperturbed by
1148: the diverging superfluid velocity. Spinon spectral density in the core
1149: should be qualitatively similar to that far outside the core. This is our basis
1150: for expecting a pseudogap-like spectrum in the core of a holon vortex.
1151:
1152: We now address the possible origin of the experimentally observed vortex
1153: core states\cite{maggio1,pan2} within the present scenario for a holon vortex.
1154: To this end consider the effect of the last term in Eq.\ (\ref{leffi}) which
1155: we ignored so far. Upon expanding the binomial
1156: the temporal component is seen to contain a density-density
1157: interaction of the form $J_{\rm sp}^0 J_h^0$ where $J_h^0$ is the local
1158: density of {\em uncondensed} holons. Since the holon order parameter
1159: vanishes in the core and the electric neutrality dictates that
1160: the total density of holons must be approximately constant in space, we
1161: expect that uncondesed holon density will behave roughly as
1162: %
1163: \[
1164: J_h^0(r)=\bar\rho_b-\rho_b(r);
1165: \]
1166: %
1167: $J_h^0(r)$ will have a spike in the core of a holon vortex.
1168: Insofar as $J_h^0(r)$ can be viewed as a static potential acting on
1169: spinons, the uncondensed holons in the vortex core can be thought of as
1170: creating a scattering potential, akin to an
1171: impurity embedded in a $d$-wave superconductor.
1172: In fact, formally the spinon problem is identical to the problem of a
1173: fermionic quasiparticle in a $d$-wave superconductor
1174: in zero field in the presence of a localized impurity potential. It is known
1175: that such problem exhibits a pair of marginally bound impurity
1176: states\cite{balatsky1}
1177: at low energies which result in sharp resonances in the spectral density
1178: inside the gap. Such states have been extensively studied theoretically
1179: \cite{balatsky2,flatte1,shnirman1,atkinson1} and their existence was
1180: recently confirmed experimentally by Pan {\em et al.}
1181: \cite{pan1}. We propose here that, within the formalism of Eq.\ (\ref{leffi}),
1182: the same mechanism could give rise to the low energy quasiparticle states
1183: in the core of a holon vortex. Such structure, if indeed confirmed by a
1184: microscopic calculation, could explain the spectral features observed
1185: experimentally in the vortex cores of cuprate superconductors
1186: \cite{maggio1,pan2}.
1187:
1188:
1189:
1190: \section{Conclusions}
1191:
1192: Scanning tunneling spectroscopy of the vortex cores affords a unique
1193: opportunity for probing the underlying ``normal'' ground state in cuprate
1194: superconductors. The existing experimental data
1195: on YBCO and BSCCO strongly suggest that conventional mean field weak coupling
1196: theories \cite{soininen1,wang1,franz1,kita1,volovik1,maki1,ichioka1}
1197: fail to describe the physics of the vortex core.
1198: Our main objective was to develop a theoretical framework for understanding
1199: these spectra and the nature of the strongly
1200: correlated electronic system which emerges once the superconducting
1201: order is suppressed. We have shown that phenomenological model (\ref{fgl1})
1202: based on a variant of the U(1) gauge field slave boson theory
1203: \cite{dhlee1} contains the right physics, provided that the gauge field
1204: stiffness
1205: is vanishingly small. The latter assumption is consistent with the
1206: general arguments involving local gauge symmetry\cite{nayak1}. In such a
1207: theory the gauge field can be explicitly integrated out, resulting in the
1208: effective action (\ref{feff}) which contains one phase degree of freedom
1209: representing the phase of a Cooper pair and two amplitude degrees of
1210: freedom representing the holon and spinon condensates.
1211:
1212: Analysis of the effective theory (\ref{feff}) in the presence of a
1213: magnetic field
1214: establishes existence of two types of vortices, spinon and holon, with
1215: contrasting spectral properties in their core regions.
1216: Our holon vortex is singly quantized and therefore differs in a profound way
1217: from the doubly quantized holon vortex discussed by SNL\cite{sachdev1,lee3}.
1218: As indicated in Figure \ref{fig2} such a singly quantized
1219: holon vortex is expected to be
1220: stable over the large portion of the phase diagram on the underdoped side.
1221: Quasiparticle spectrum in the core of a holon vortex is predicted to exhibit
1222: a ``pseudogap'', similar to that found in the underdoped normal region
1223: above $T_c$. This is consistent with the data of Renner {\em et al.}
1224: \cite{renner1} who pointed out a remarkable similarity between the vortex
1225: core and the normal state spectra in BSCCO. Spinon vortex, on the other hand,
1226: should be virtually indistinguishable from the conventional $d$-wave BCS
1227: vortex and is expected to occur on the overdoped side of the phase diagram.
1228: Transition from the insulating holon vortex to the metallic spinon vortex
1229: as a function of doping is a concrete testable prediction of the present
1230: theory.
1231:
1232: Phenomenological theory based on the effective
1233: action (\ref{feff}) does not permit
1234: explicit evaluation of the electronic spectral function. To this end
1235: we have considered the corresponding microscopic theory
1236: (\ref{leffi}) and concluded that
1237: holon vortex will indeed exhibit a pseudogap like spectrum. Such qualitative
1238: analysis furthermore suggests a plausible mechanism for the sharp vortex
1239: core states observed in YBCO\cite{maggio1} and BSCCO
1240: \cite{pan2}. We stress that conventional mean field weak coupling theories
1241: yield neither pseudogap nor the core states.
1242: In the core of a holon vortex such states will arise
1243: as a result of spinons scattering off of the locally uncondensed holons,
1244: in a manner analogous to the quasiparticle resonant states in the vicinity
1245: of an impurity in a $d$-wave superconductor
1246: \cite{balatsky1,balatsky2,flatte1,shnirman1,atkinson1}.
1247: The latter conclusion is somewhat speculative and must be
1248: confirmed by explicitly solving the fermionic sector of the microscopic
1249: theory (\ref{leffi}).
1250:
1251: On a broader theoretical front the importance of the vortex core spectroscopy
1252: as a window to the normal state in the $T\to 0$ limit lies in its potential
1253: to discriminate between various microscopic theories of cuprates.
1254: It is reasonable to assume that the observed pseudogap in the
1255: vortex core reflects the same physics as the pseudogap observed
1256: in the normal state.
1257: This means that the mechanism responsible for the pseudogap must be operative
1258: on extremely short lengthscales, of order of several lattice spacings.
1259: The U(1) slave boson theory considered in this work apparently satisfies
1260: this requirement. Obtaining the correct vortex core spectral functions could
1261: serve as an interesting test for other theoretical approaches describing
1262: the physics of the underdoped cuprates\cite{lee2,pines1,levin1}.
1263:
1264: It will be of interest to explore
1265: the implications of the effective theories (\ref{feff}) and (\ref{leffi})
1266: in other physical situations. Of special interest are situations where
1267: the holon condensate amplitude is suppressed, locally
1268: or globally, giving rise to ``normal'' transport properties (vanishing
1269: superfluid density) but quasiparticle excitations that are characteristic
1270: of a superconducting state. These include
1271: the spectra in the vicinity of an impurity,
1272: twin boundary or a sample edge. In the latter case one might hope to observe
1273: a signature of the zero bias tunneling peak anomaly
1274: (normally seen for certain geometries deep in the superconducting phase in
1275: the optimally doped cuprates) even above $T_c$ in the underdoped samples.
1276:
1277: \acknowledgments
1278: The authors are indebted to A. J. Berlinsky,
1279: J. C. Davis, \O. Fischer, C. Kallin, D.-H. Lee, P. A. Lee,
1280: S.-H. Pan, C. Renner, J. Ye and S.C. Zhang
1281: for helpful discussions. This research was
1282: supported in part by NSF grant DMR-9415549 and by Aspen Center for Physics
1283: where part of the work was done.
1284:
1285: \vskip 10pt
1286: {\em Note added in proof.} After submission of this manuscript we learned
1287: about complementary microscopic treatments of the spin-charge separated
1288: state in the
1289: vortex core within U(1) \cite{han1} and SU(2) \cite{wen1} slave boson
1290: theories. The former agrees qualitatively with our phenomenological theory.
1291: Ref.\ \cite{wen1} proposes a new type of vortex which takes advantage of
1292: the larger symmetry group SU(2). In a related development Senthil
1293: and Fisher \cite{senthil1} discussed a Z$_2$ vortex (which is essentially
1294: equivalent to our singly quantized holon vortex) and proposed a
1295: ``vison detection'' experiment based on trapping such a vortex in the hole
1296: fabricated in a strongly
1297: underdoped superconductor. Here we wish to point out that the experiment
1298: will produce the same general outcome in a system described by the U(1)
1299: theory where the role of a vison will be played by a flux quantum of the
1300: fictitious gauge field ${\bf a}$.
1301:
1302:
1303:
1304: \begin{references}
1305: \bibitem{ando1} G. S. Boebinger, Y. Ando, A. Passer, T. Kimura,
1306: M. Okuya, J. Shimoyama, K. Kishio, K. Tamasaku, N. Ichikawa and S. Uchida,
1307: \prl {\bf 77}, 5417 (1996).
1308: \bibitem{lemberger1} K. Karpinska, M. Z. Cieplak, S. Guha, A. Malinowski,
1309: T. Skoskiewicz, W. Plesiewicz, M. Berkowski, B. Boyce, T. R. Lemberger and
1310: P. Lindenfeld, \prl {\bf 84} 155 (2000).
1311: \bibitem{maggio1} I. Maggio-Aprile, Ch. Renner, A. Erb, E. Walker, and
1312: \O. Fischer, \prl {\bf 75}, 2754 (1995).
1313: \bibitem{renner1} Ch. Renner, B. Revaz, K. Kadowaki, I. Maggio-Aprile,
1314: and \O. Fischer, \prl {\bf 80}, 3606 (1998).
1315: \bibitem{pan1} E. W. Hudson, S. H. Pan, A. K. Gupta, K.-W. Ng and
1316: J. C. Davis, {\em Science} {\bf 285}, 88 (1999).
1317: \bibitem{pan2} S. H. Pan, E. W. Hudson, A. K. Gupta, K.-W. Ng,
1318: H. Eisaki, S. Uchida and J. C. Davis,\prl {\bf 85}, 1536 (2000).
1319: \bibitem{soininen1} P.I. Soininen, C. Kallin, and A.J. Berlinsky, \prb{\bf 50},
1320: 13883 (1994).
1321: \bibitem{wang1} Y. Wang and A. H. MacDonald, \prb {\bf 52}, R3876 (1995).
1322: \bibitem{franz1} M. Franz and Z. Te\v{s}anovi\'c, \prl{\bf 60}, 4763
1323: (1998).
1324: \bibitem{kita1} K. Yasui and T. Kita, \prl {\bf 83}, 4168 (1999).
1325: \bibitem{volovik1} G.\ E.\ Volovik, JETP Lett.{\bf 58}, 469 (1993).
1326: \bibitem{maki1} Y. Morita, M. Kohmoto, and K. Maki, \prl {\bf 78}, 4841
1327: (1997).
1328: \bibitem{ichioka1} M. Ichioka, N. Hayashi, N. Enomoto, and K. Machida,
1329: \prb {\bf 53}, 15316 (1996).
1330: \bibitem{franz2} M. Franz and M. Ichioka, \prl {\bf 79}, 4513 (1997).
1331: \bibitem{himeda1} A. Himeda, M. Ogata, Y. Tanaka, and S. Kashiwaya,
1332: J.\ Phys.\ Soc.\ Japan {\bf 66}, 3367 (1997).
1333: \bibitem{resende1} X. R. Resende and A. H. MacDonald, Bull.\ Am.\ Phys.\ Soc.\
1334: {\bf 43}, 573 (1998).
1335: \bibitem{arovas1} D. P. Arovas, A. J. Berlinsky, C. Kallin, and
1336: S.-C. Zhang, \prl {\bf 79}, 2871 (1997).
1337: \bibitem{zhang1} S.-C. Zhang, {\em Science}, {\bf 275}, 1089 (1997).
1338: \bibitem{andersen1} B. M. Andersen, H. Bruus, and P. Hedeg{\aa}rd,
1339: \prb {\bf 61}, 6298 (2000).
1340: \bibitem{dhlee1} D.-H. Lee, \prl {\bf 84}, 2694 (2000).
1341: \bibitem{anderson1} P. W. Anderson, {\em Science}, {\bf 235}, 1196 (1987).
1342: \bibitem{baskaran1} G. Baskaran, Z. Zhou, and P. W. Anderson, Solid State
1343: Commun. {\bf 63}, 973 (1987).
1344: \bibitem{ruckenstein1} A. E. Ruckenstein, P. J. Hirschfeld, and J. Appel,
1345: \prb {\bf 36}, 857 (1987).
1346: \bibitem{kotliar1} G. Kotliar and J. Liu, \prb {\bf 38}, 5142 (1988).
1347: \bibitem{affleck1} I. Affleck and J. B. Marston, \prb {\bf 37}, 3774 (1988).
1348: \bibitem{lee1} P. A. Lee and N. Nagaosa, \prb {\bf 46}, 5621 (1992).
1349: \bibitem{wen1} X.-G. Wen and P. A. Lee, \prl {\bf 76}, 503 (1996);
1350: P. A. Lee and X.-G. Wen, \prl {\bf 78}, 4111 (1997).
1351: \bibitem{lee2} P. A. Lee, N. Nagaosa, T.-K. Ng, and X.-G. Wen,
1352: \prb {\bf 57}, 6003 (1998).
1353: \bibitem{balents1} L. Balents, M. P. A. Fisher and C. Nayak, Int.\ J.\ Mod.\
1354: Phys.\ B {\bf 12}, 1033 (1998); \prb {\bf 60}, 1654 (1999).
1355: \bibitem{sachdev1} S. Sachdev, \prb {\bf 45}, 389 (1992).
1356: \bibitem{lee3} N. Nagaosa and P. A. Lee, \prb {\bf 45}, 966 (1992).
1357: \bibitem{abrikosov1} A.A. Abrikosov, Zh.\ Eksp.\ Teor.\ Fiz.\ {\bf 32}, 1442
1358: (1957) [Sov.\ Phys.\ JETP {\bf 5}, 1174 (1957)].
1359: \bibitem{fetter1} A. L. Fetter and P. C. Hohenberg, in
1360: {\em Superconductivity}, edited by R. D. Parks (Marcel Dekker, New York 1969).
1361: \bibitem{elitzur1} S. Elitzur, \prd {\bf 12}, 3978 (1975).
1362: \bibitem{nayak1} C. Nayak, \prl {\bf 85}, 178 (2000).
1363: \bibitem{hu1} C.-R. Hu, \prb {\bf 6} 1756 (1972).
1364: \bibitem{alama1} S. Alama, A. J. Berlinsky, L. Bronsard, and T. Giorgi,
1365: \prb {\bf 60}, 6901 (1999).
1366: \bibitem{numrec} W. H. Press, B. P. Flannery, S. A. Teukolsky, and
1367: W. T. Vetterling: {\em Numerical Recipes in C}, (Cambridge University Press,
1368: 1988).
1369: \bibitem{wen2} X.-G. Wen, \prb {\bf 44}, 2664 (1991).
1370: \bibitem{kwon1} H.-J. Kwon and A. T. Dorsey, \prb {\bf 59}, 6438 (1999).
1371: \bibitem{balatsky1} A. V. Balatsky and M. I. Salkola, \prl {\bf 76}, 2386
1372: (1996).
1373: \bibitem{balatsky2} M. I. Salkola, A. V. Balatsky, and D. J. Scalapino,
1374: \prl {\bf 77}, 1841 (1996).
1375: \bibitem{flatte1} M. E. Flatt\'{e} and J. M. Byers, \prl {\bf 80}, 4546 (1998);
1376: Solid State Phys. {\bf 52}, 137 (1999).
1377: \bibitem{shnirman1} A. Shnirman, I. Adagideli, P. M. Goldbart, and A. Yazdani,
1378: \prb {\bf 60}, 7517 (1999).
1379: \bibitem{atkinson1} W. A. Atkinson, P. J. Hirschfeld, and A. H. MacDonald,
1380: cond-mat/9912158.
1381: \bibitem{pines1} J. Schmalian, D. Pines, B. Stojkovic, \prl {\bf 80}, 3839
1382: (1998); \prb {\bf 60}, 667 (1999).
1383: \bibitem{levin1} Q. Chen, I. Kosztin, B. Janko, K. Levin, \prl {\bf 81}, 4708
1384: (1998); \prb {\bf 59}, 7083 (1999).
1385: \bibitem{han1} J. H. Han and D.-H. Lee, \prl{\bf 85}, 1100 (2000).
1386: \bibitem{wen1} P. A. Lee and X.-G. Wen, cond-mat/0008419
1387: \bibitem{senthil1} T. Senthil and M.P.A. Fisher, \prb {\bf 62}, 7850 (2000);
1388: cond-mat/0006481.
1389:
1390: %\bibitem{} , \prl {\bf }, ().
1391: \end{references}
1392:
1393: \end{document}
1394:
1395: \bibitem{} {\em et al.}, \prl {\bf }, ().