1: \documentstyle[aps,graphicx]{revtex}
2: %\documentstyle[aps,preprint,graphicx]{revtex}
3: \begin{document}
4:
5: \draft
6: \twocolumn[\hsize\textwidth\columnwidth\hsize\csname@twocolumnfalse\endcsname
7:
8: \title{Quantum Phase Fluctuations Responsible for Pseudogap}
9:
10: \author{V.P.~Gusynin$^1$, V.M.~Loktev$^1$\cite{e-mail},
11: R.M.~Quick$^2$ and S.G.~Sharapov$^{1,2,3}$\cite{e-mail2}}
12:
13:
14: \address{
15: $^1$Bogolyubov Institute for Theoretical Physics, 03143 Kiev, Ukraine\\
16: $^2$Department of Physics, University of Pretoria, 0002 Pretoria, South Africa\\
17: $^3$ Institut de Physique, Universite de Neuch\^atel,
18: CH-2000 Neuch\^atel, Switzerland }
19:
20: \date{27 December 2000}
21:
22:
23: \maketitle
24:
25:
26: \begin{abstract}
27: The effect of ordering field phase fluctuations on the normal and
28: superconducting properties of a simple 2D model with a local
29: four-fermion attraction is studied. Neglecting the coupling
30: between the spin and charge degrees of freedom an analytical
31: expression has been obtained for the fermion spectral function as
32: a single integral over a simple function.
33: %The analytical character of the calculation enables one to state
34: %unequivocally that, as the temperature increases through the 2D critical
35: %temperature and a nontrivial damping for a phase correlator develops,
36: %it is the quantum (or dynamical) and not the classical phase fluctuations
37: %that fill the gap in the quasiparticle spectrum.
38: From this we show that, as the temperature increases through the 2D
39: critical temperature and a nontrivial damping for a phase
40: correlator develops, quantum fluctuations fill the gap in the
41: quasiparticle spectrum.
42: Simultaneously the quasiparticle peaks broaden significantly above the critical
43: temperature, resembling the observed pseudogap behavior in
44: high-$T_c$ superconductors.
45: \end{abstract}
46:
47: \pacs{\rm PACS numbers:
48: 74.25.-q,
49: % General properties; correlation between physical properties
50: % in normal and superconducting states
51: 74.40.+k,
52: % Fluctuations
53: 74.72.-h
54: % High-T_c compounds
55: }
56: {\em Key words: pseudogap, phase fluctuations}\\
57: ]
58:
59:
60: The pseudogap, or depletion of the single particle spectral weight
61: around the Fermi level, \cite{Timusk} is the most striking
62: demonstration that cuprate superconductors are not described by
63: the BCS scenario of superconductivity. The pseudogap opens in the
64: normal state as the temperature is lowered below a crossover
65: temperature $T^{\ast}$ and extends over a wide range of
66: temperatures in the underdoped cuprates. ARPES \cite{Campuzano,Feng}
67: and the scanning tunneling spectroscopy (STS) (see Refs. in
68: \cite{Timusk}) provide particularly complete information about the
69: pseudogap behavior and show a smooth crossover from the pseudo- to
70: superconducting (SC) gap. The transition from SC to normal
71: behavior appears to be driven by phase fluctuations \cite{Corson}
72: and is well described by the Berezinskii-Kosterlitz-Thouless
73: (BKT) theory of vortex-pair unbinding.
74:
75: There are currently many possible explanations for the unusual
76: behavior of HTSC. One of these is based on the nearly
77: antiferromagnetic Fermi liquid model \cite{Pines}. Another
78: explanation, proposed by Anderson, relies on the separation of the
79: spin and charge degrees of freedom.\cite{Anderson} The third
80: approach, which we follow in this paper, relates the observed
81: anomalies to precursor SC fluctuations, and in particular
82: fluctuations in the phase of the complex ordering field, as
83: originally suggested by Emery and Kivelson \cite{Emery}.
84:
85: The phase diagram for a simple microscopic 2D model (see (\ref{Hamilton})
86: below) which formalizes the scenario of \cite{Emery} has been studied
87: in.\cite{Gusynin.JETP} The Green's function (GF) for this model was derived in
88: \cite{Gusynin.JETPLett} using the correlator
89: $\langle\exp(i\theta(x)/2)\exp(-i\theta(0)/2)\rangle$ for the phase
90: fluctuations in the classical (static) approximation and neglecting the
91: coupling between the spin and charge degrees of freedom.
92:
93: The associated spectral function (SF) $A(\omega,{\bf
94: k})=-(1/\pi){\rm Im}G(\omega+i0,{\bf k})$ has also been derived
95: analytically in \cite{Gusynin.JETPLett}. Being proportional to the
96: intensity of ARPES \cite{Campuzano}, the SF encodes information
97: about the quasiparticles and pseudogap. The result of
98: \cite{Gusynin.JETPLett} showed that, while the temperature
99: behavior of the quasiparticle peaks is in correspondence with
100: experiment \cite{Campuzano}, the gap in the spectrum remains
101: unfilled.\cite{Evans} In an earlier paper \cite{Franz} filling of
102: the gap has been achieved as a result of a Doppler shift in the
103: quasiparticle excitation spectrum. This Doppler shift originated
104: from the semiclassical coupling of the mean field $d$-wave
105: quasiparticles to the supercurrents induced by classically
106: fluctuating unbound vortex-antivortex pairs. It was shown in
107: \cite{Kwon} that the shift can be identified with the coupling of
108: the spin and charge degrees of freedom.
109:
110: The purpose of the present paper is to show analytically that
111: filling of the gap can also result from the quantum
112: (dynamical) phase fluctuations, even when the coupling between
113: charge and spin degrees of freedom is entirely neglected.
114: As pointed out in \cite{Tremblay.new} the mechanism of the gap
115: filling has to be understood yet even for the simplest attractive
116: Hubbard model. Thus we contribute into the discussion \cite{Tremblay.new}
117: as to which mechanism or mechanisms lead to pseudogap filling but within
118: the scenario based on phase fluctuations, by studying the mechanism proposed
119: in \cite{Capezzali}. The advantage of our calculation is that
120: the SF is obtained as a single integral with an analytical integrand,
121: no numerical analytical continuation was necessary resulting in far
122: greater accuracy.
123:
124: Let us consider the continuum version of the
125: %simplest
126: 2D attractive Hubbard model defined by the
127: Hamiltonian density:
128: \begin{equation}
129: \hspace{-2mm} {\cal H} = - \psi_{\sigma}^{\dagger}(x) \left(\frac{\nabla^{2}}{2
130: m} + \mu \right)
131: \psi_{\sigma}(x) - V \psi_{\uparrow}^{\dagger}(x)
132: \psi_{\downarrow}^{\dagger}(x) \psi_{\downarrow}(x)
133: \psi_{\uparrow}(x),
134: \label{Hamilton}
135: \end{equation}
136: where $x= \mbox{\bf r}, \tau$ denotes the space and imaginary time
137: variables, $\psi_{\sigma}(x)$ is a fermion field with spin $\sigma
138: =\uparrow,\downarrow$, $m$ is the effective fermion mass, $\mu$ is
139: the chemical potential, and $V$ is an effective local attraction
140: constant; we take $\hbar = k_{B} = 1$. Clearly the Hamiltonian
141: (\ref{Hamilton}) is too simple to be adequate for systems as
142: complex as cuprate HTSC. However, it has proved itself as a very
143: convenient model for both numerical, in particular Monte Carlo
144: \cite{Ran1,Ran2} simulations, and analytical approaches
145: \cite{Gusynin.JETP,Gusynin.JETPLett,Capezzali,Babaev}
146: which does exhibit gap-like behavior above $T_c$
147: (see also Refs. in the review \cite{we.review}). Moreover
148: one may use the model to obtain a fully analytic treatment of the
149: pseudogap properties, and apply such results to obtain a better
150: understanding of more complex and less tractable models.
151:
152: The calculation of the GF is performed in Nambu variables and the
153: fermions are
154: treated as composite objects, comprising both spin and charge parts:
155: \begin{equation}
156: \Psi^{\dagger}(x) = \left( \begin{array}{lr}
157: \psi_{\uparrow}^{\dagger}(x) & \psi_{\downarrow}(x)
158: \end{array} \right) =
159: \Upsilon^{\dagger} (x) \exp[-i\tau_{3}\theta(x)/2],
160: \label{Nambu.spinor}
161: \end{equation}
162: where $\Upsilon^{\dagger}(x)$ is the Nambu spinor of neutral
163: fermions. Substituting (\ref{Nambu.spinor}) into the standard
164: definition of the GF
165: $G(x) =\langle \Psi(x) \Psi^{\dagger} (0) \rangle$ one obtains
166: the GF of the charged (observed) fermions
167: \begin{equation}
168: G_{\alpha \beta}(x) = \sum_{\alpha^\prime, \beta^{\prime}} {\cal
169: G}_{\alpha^\prime\beta^\prime}(x)
170: \langle (e^{i\tau_{3}\frac{\theta(x)}{2}})_{\alpha\alpha^\prime}
171: (e^{-i\tau_{3}\frac{\theta(0)}{2}})_{\beta^\prime\beta} \rangle,
172: \label{Green.splited}
173: \end{equation}
174: as the product of the GF for the neutral fermions ${\cal
175: G}_{\alpha\beta}(x)=\langle \Upsilon_{\alpha}(x)
176: \Upsilon^{\dagger}_{\beta}(0) \rangle$ and the phase correlator
177: $\langle\exp(i\tau_{3} \theta(x)/2) \exp(-i
178: \tau_{3}\theta(0)/2)\rangle$. For the frequency-momentum
179: representation of Eq.~(\ref{Green.splited}) one has
180: \begin{eqnarray}
181: G(i \omega_{n}, \mbox{\bf k})& = & T \sum_{m = - \infty}^{\infty} \int
182: \frac{d^2 p}{(2 \pi)^{2}} \sum_{\alpha,\beta=\pm} P_\alpha {\cal G}(i
183: \omega_{m}, \mbox{\bf p}) P_\beta \nonumber \\
184: && \times D_{\alpha\beta} (i \omega_{n} - i \omega_{m}, \mbox{\bf k} -
185: \mbox{\bf p}), \label{Green.projectors.momentum}
186: \end{eqnarray}
187: where $P_{\pm}={1\over2}(\hat I\pm \tau_3)$ are the projectors;
188: $\hat I$ and $\tau_{3}$ are unit and Pauli matrices;
189: $D_{\alpha\beta}(i\Omega_m, \mbox{\bf q})$ is the Fourier
190: transformation of the phase correlator $D_{\alpha\beta}(x) =
191: \langle\exp(i\alpha \theta(x)/2) \exp(-i
192: \beta\theta(0)/2)\rangle$; $\omega_{n} = (2n+1) \pi T$ and
193: $\Omega_{n} = 2n \pi T$ are respectively odd and even Matsubara
194: frequencies. The GF of neutral fermions is taken
195: in the mean-field approximation (see \cite{Gusynin.JETP,Gusynin.JETPLett})
196: \begin{equation}
197: {\cal G}(i \omega_{n}, \mbox{\bf p}) = - \frac{ i \omega_{n} \hat{I} +
198: \tau_{3} \xi(\mbox{\bf p}) - \tau_{1} \rho} {\omega_{n}^{2} +
199: \xi^{2}(\mbox{\bf p}) + \rho^{2}} \,, \, \, \xi(\mbox{\bf p}) =
200: \frac{\mbox{\bf p}^{2}}{2m} - \mu \label{Green.neutral}
201: \end{equation}
202: with $\mbox{\bf p}$ being a 2D vector and
203: $\rho\equiv\langle\rho(x)\rangle$, where $\rho(x)$ is the modulus
204: of a complex ordering field $\Phi(x) = \rho(x) \exp [i\theta(x)]$.
205: Note that $\langle \Phi (x) \rangle = 0$ for $T \neq 0$ as it
206: should be in 2D in accordance with the
207: Coleman-Mermin-Wagner-Hohenberg theorem, while the value of $\rho$
208: is allowed to be nonzero. Note also that the GF
209: (\ref{Green.projectors.momentum}) does not contain the symmetry
210: violating term $\sim \tau_{1}$ which could originate from the
211: terms $P_{\pm} {\cal G}(i \omega_{n}, \mbox{\bf k}) P_{\mp}$ since
212: the correlators $D_{+-} = D_{-+} =0$ \cite{Gusynin.JETPLett}.
213:
214: The representations (\ref{Green.splited}), (\ref{Green.projectors.momentum})
215: for the fermion GF with decoupled spin and charged degrees of freedom
216: are appropriate when one can neglect the fluctuations of the modulus
217: $\rho(x)$ and when the energy of the phase distortions
218: is smaller than the energy gain due to nontrivial $\rho$.
219: For the present s-wave model this means that the condition $\rho \gg T$
220: should be satisfied, so that the modulus-phase representation
221: (\ref{Nambu.spinor}) should be useful even for $T$ close to $T_{\rm
222: BKT}$ and allows one to study the evolution of the SC gap to the
223: pseudogap. The region of temperatures $T$ where the condition
224: $\rho \gg T$ is satisfied depends crucially on the relative size
225: of the pseudogap region $(T^{\ast}- T_{\rm BKT})/T^{\ast}$ which
226: may be reasonably large in 2D for intermediate
227: coupling \cite{Tremblay}. The pseudogap in the current work is the
228: result of the low dimensionality of the system and does not need
229: the existence of preformed local pairs. The nonzero value of
230: $\rho$ is due to average local density of Cooper pairs which do
231: not have coherence above $T_{\rm BKT}$.
232:
233: The representation (\ref{Green.projectors.momentum}) shows
234: (see the explanation after Eq.~(\ref{Green.neutral}))
235: that the GF for the charged fermions is defined by the correlator
236: $D(x) \equiv D_{++}(x) = D_{--}(x)$.
237: The asymptotic form of the correlator at large distances,
238: describing also the temporal decay of correlations, is given by
239: the form \cite{Capezzali,Huber}
240: \begin{equation}
241: D^{R}(t, \mbox{\bf r}) = \exp (- \gamma t)
242: (r/r_{0})^{-T/8 \pi J}
243: \exp ( - r/\xi_{+}(T)).
244: \label{correlator.dynamic}
245: \end{equation}
246: Here $t$ is the real time, $\gamma$ is a decay constant, $r_0
247: \equiv (2/T) (J/K)^{1/2}$ is the scale for the algebraic decay of
248: correlations in the SC BKT phase ($T < T_{\rm BKT}$, where $T_{\rm
249: BKT}$ is the temperature of the BKT transition), $\xi_{+}(T)$ is
250: the phase coherence length for $T > T_{\rm BKT}$ ($\xi_{+}(T \to
251: T_{\rm BKT}^{+}) \to \infty$). The constants $J$ and $K$ are the
252: bare (mean-field) superfluid stiffness and compressibility, respectively
253: which have been calculated in \cite{Gusynin.JETP}.
254:
255: Previously, using the representation (\ref{Nambu.spinor}), only
256: the classical (static $\Omega_n = \gamma =0$) fluctuations have
257: been considered analytically \cite{Gusynin.JETPLett}. The Fourier
258: transform of (\ref{correlator.dynamic}) for this case is
259: \begin{equation}
260: D(i \Omega_{n}, \mbox{\bf q}) \simeq \delta_{n, 0} C
261: [\mbox{\bf q}^2 + (1/\xi_{+})^{2} ]^{-\alpha} /T \,,
262: \label{Fourier.D.final}
263: \end{equation}
264: where $C \equiv 4 \pi (\Gamma(\alpha) / \Gamma(1-\alpha) )
265: ( 2/r_{0} )^{2\alpha - 2}$ and $\alpha \equiv 1 - T / 16 \pi J$.
266: For $T \sim T_{\rm BKT}$ the value of
267: $\alpha \simeq 1 - T/32 T_{\rm BKT}$.
268: The presence of $\alpha \ne 1$ in (\ref{Fourier.D.final})
269: is related to the preexponent factor $(r/r_0)^{-T/8 \pi J}$ in
270: (\ref{correlator.dynamic}). This factor was not included in the analysis
271: of \cite{Franz,Kwon}, but the treatment of \cite{Capezzali} includes
272: this factor.
273:
274: One can now extend the analysis to the
275: case of quantum (dynamical) phase fluctuations.
276: We propose the following generalization of
277: (\ref{Fourier.D.final}):
278: \begin{equation}
279: D(i \Omega_{n}, \mbox{\bf q}) = \frac{C (v^2)^{\alpha}}
280: {T [v^2 \mbox{\bf q}^2 + (v/\xi_{+})^{2} + \Omega_{n}^{2} +
281: 2 \gamma |\Omega_{n}|]^{\alpha}},
282: \label{Fourier.D.dynamical}
283: \end{equation}
284: where $v$ is the velocity of the Bogolyubov excitations. Recall that in 2D $v
285: = v_{F}/\sqrt{2}$, where $v_{F}$ is the Fermi velocity.
286: %
287: The asymptotic form of the retarded GF corresponding to the GF
288: (\ref{Fourier.D.dynamical}) is
289: \begin{eqnarray}
290: & D^{R} (t, \mbox{\bf q}) \sim \nonumber \\
291: & \left\{
292: \begin{array}{c c}
293: t^{\alpha-1} e^{- \gamma t} \,, &
294: v^2 q^2 > \gamma^{2} - v^2 \xi_{+}^{-2} \\
295: t^{\alpha-1}
296: e^{-t (\gamma - \sqrt{\gamma^2 - v^2 \xi_{+}^{-2} - v^2 q^{2}})} \,, &
297: v^2 q^2 < \gamma^{2} - v^2 \xi_{+}^{-2}
298: \end{array} \right. \nonumber \\
299: & \qquad \qquad \qquad t \to +\infty \, .
300: \label{D.retarded.dynamical}
301: \end{eqnarray}
302: %
303: Eq.~(\ref{Fourier.D.dynamical}) can be regarded as a convenient
304: (for analytical studies)
305: generalization of the phenomenological
306: dependence (\ref{correlator.dynamic}) \cite{Capezzali} which for
307: nonzero $\gamma$ includes the decay of the phase correlations due
308: to the presence of free vortices above $T_{\rm BKT}$.
309: One can see that for $v^2 q^2 < \gamma^{2} -
310: v^2 \xi_{+}^{-2}$ the decay rate is less than $\gamma$ and it is minimal for
311: $q=0$. This means that, for large distances, phase fluctuations do not feel
312: the pair vortices which have smaller size.
313:
314: In the present work both $\gamma$ and $\xi_{+}$ are
315: phenomenological parameters which can be derived from the theory
316: of the BKT transition. Since in the SC BKT phase the vortices are
317: confined, one can state \cite{Huber} that there is a critical
318: slowing down of the phase fluctuations when the temperature
319: approaches $T_{\rm BKT}$, i.e. $\gamma(T\to T_{\rm BKT}^{+}) = 0$.
320: In fact the detailed theory of BKT transition predicts that
321: $\gamma(T) \sim \xi_{+}^{-z}(T)$, where $z$ is the dynamical exponent
322: and $\xi_{+}^{-1}(T) \simeq \xi_{0}^{-1}
323: \exp(- b/ \sqrt{T - T_{\rm BKT}})$, where $b$ is a positive
324: constant. In what follows we just restrict ourselves by comparing
325: two cases: $\gamma \neq 0$ above $T_{\rm BKT}$ and $\gamma =0$ for
326: $T < T_{\rm BKT}$.
327:
328: We note that there is now experimental evidence
329: \cite{Corson} for the vortex-pair unbinding (BKT) nature of the SC
330: transition in the Bi-cuprates. We stress also that, despite the
331: rather simple form of the GF, it takes into account the presence
332: of vortices while the self-consistent $T$-matrix approximation
333: (see e.g. \cite{Tremblay.new,we.review,Levin}) cannot describe them.
334:
335: The SF associated with GF (\ref{Green.projectors.momentum}) can readily be
336: expressed in terms of the corresponding SF for the GF (\ref{Green.neutral})
337: and (\ref{Fourier.D.dynamical}) as
338: \begin{eqnarray}
339: && A(\omega, \mbox{\bf k}) = - \frac{1}{\pi}{\rm Im} G_{11}(\omega + i0,
340: \mbox{\bf k}) = \int\limits_{-\infty}^\infty d\omega^\prime \left[\frac{1}{1 +
341: e^{\omega^\prime/T}} \right. - \nonumber \\
342: && \left. \frac{1}{1 - e^{(\omega^\prime-\omega)/T}} \right] \int\frac{d^2
343: p}{(2\pi)^{2}} a_F(\omega^\prime, \mbox{\bf p})a_B(\omega-\omega^\prime,
344: \mbox{\bf k} - \mbox{\bf p}) \,.
345: \label{spectral.1}
346: \end{eqnarray}
347: where these SF are
348: \begin{eqnarray}
349: && a_{F} (\omega, \mbox{\bf p}) = (\omega + \xi(\mbox{\bf p}))
350: \delta(\omega^{2} - \xi^{2}(\mbox{\bf p}) - \rho^2) \mbox{sgn} \omega \,,
351: \nonumber \\
352: && a_{B}(\Omega, \mbox{\bf q}) = - \frac{1}{\pi} \mbox{Im}D^{R} (\Omega,
353: \mbox{\bf q}) \,.
354: \label{spectral.input}
355: \end{eqnarray}
356: The $\delta$-function in $a_{F}$ allows one to perform the angular
357: integration
358: in (\ref{spectral.1}). Finally integrating over momentum the SF can
359: be expressed
360: as a single integral:
361: \begin{eqnarray}
362: && A(\omega,{\bf k})= - \frac{1}{2 \pi^2 T}\frac{\Gamma(\alpha)}{\Gamma(1 -
363: \alpha)} \left( \frac{2}{m r_{0}^{2}} \right)^{\alpha-1} \int
364: \limits_{-\infty}^\infty d \omega^\prime {\rm sgn}(\omega^\prime + \omega)
365: \nonumber \\ && \times \left[ \frac{1}{1-e^{\omega^\prime/T}}-
366: \frac{1}{1+e^{(\omega^\prime+\omega)/T}}\right]
367: \frac{\theta[(\omega^\prime+\omega)^2- \rho^2]}
368: {\sqrt{(\omega^\prime+\omega)^2 - \rho^2}} \nonumber \\
369: && \times \left\{ \left[ I(\omega,{\bf k},\omega^\prime)
370: \left(\omega^\prime+\omega+\sqrt{(\omega^\prime+\omega)^2- \rho^2}\right)
371: \right. \right. \nonumber \\ && \left. \times
372: \theta(\mu+\sqrt{(\omega^\prime+\omega)^2 -\rho^2}) \right]
373: \nonumber \\
374: &&\left.+\left[\sqrt{(\omega^\prime+\omega)^2-\rho^2}
375: \rightarrow -\sqrt{(\omega^\prime+\omega)^2-\rho^2}\right]\right\}\,,
376: \label{spectral.final}
377: \end{eqnarray}
378: where we have introduced a function $I(\omega,{\bf k},\omega^\prime)$:
379: \begin{eqnarray}
380: && I(\omega,{\bf k},\omega^\prime) = \nonumber \\
381: &&\pi{\rm Im} \left[(x_{-}-a- ib)^{-\alpha}
382: F\left(\frac{1}{2},\alpha;1;\frac{x_{+} -x_{-}}{a+ib-x_{- }}\right)\right]\,,
383: \nonumber \\ && a = \frac{1}{2m} \left(\frac{\omega^{\prime \: 2}}{v^2} -
384: \xi_{+}^{-2}\right)\,, \qquad b=\frac{\gamma\omega^\prime}{m v^2}\,, \nonumber \\
385: && x_{\pm}=\left(\sqrt{\frac{k^2}{2m}}\pm
386: \sqrt{\mu+\sqrt{(\omega+\omega^\prime)^2-\rho^2}}\right)^2\,.
387: \label{I}
388: \end{eqnarray}
389:
390: For $\gamma=0$ the expression (\ref{spectral.final}) can be
391: rewritten in the following form
392: \begin{eqnarray}
393: && A(\omega,{\bf k})= - \frac{1}{2 \pi^2 T}\frac{\Gamma(\alpha)} {\Gamma(1-
394: \alpha)} \left( \frac{2}{m r_{0}^{2}}\right)^{\alpha-1}
395: \int\limits_{-\infty}^\infty d\omega^{\prime} {\rm sgn}\omega^{\prime}
396: \nonumber \\
397: && \times {\rm sgn}(\omega^{\prime}+ \omega)
398: \left[\frac{1}{1-e^{\omega^{\prime}/T}}-
399: \frac{1}{1+e^{(\omega^{\prime}+\omega)/T}}\right] \nonumber \\
400: && \times \frac{\theta[(\omega^{\prime}+\omega)^2-\rho^2]}
401: {\sqrt{(\omega^{\prime}+\omega)^2-\rho^2}} \theta(\omega^{\prime
402: 2}-v^2\xi^{-2}_{+}) \nonumber \\
403: && \times \left\{\left[\pi(a-x_{-})^{-\alpha}
404: F\left({1\over2},\alpha;1;\frac{x_{+}-
405: x_{-}}{a-x_{-}}\right)\theta(a-x_{+})\right.\right.
406: \nonumber \\
407: && + (x_{+}-x_{-})^{-1/2}(a-x_{-})^{1/2-\alpha}B({1\over2}, 1-\alpha)
408: \nonumber \\ && \left. \times
409: F\left({1\over2},{1\over2};{3\over2}-\alpha;\frac{a-x_{-}}{x_{+}-x_{- }}
410: \right)\theta(x_{+}-a)
411: \theta(a-x_{- })\right] \nonumber \\
412: && \times \left(\omega^{\prime}+\omega+\sqrt{(\omega^{\prime}
413: +\omega)^2-\rho^2}\right)\theta(\mu+\sqrt{(\omega^{\prime}+\omega)^2 -\rho^2})
414: \nonumber\\
415: &&\left.+\left(\sqrt{(\omega^{\prime}+\omega)^2-\rho^2}\rightarrow-
416: \sqrt{(\omega^{\prime}+
417: \omega)^2-\rho^2}\right)\right\} \label{zero.gamma}
418: \end{eqnarray}
419: which describes the case of the non-damped dynamical phase
420: fluctuations. While Eq.~(\ref{zero.gamma}) appears to be more
421: complicated than the more general Eq.~(\ref{spectral.final}), it
422: proves useful for making a general statement about the gap
423: filling.
424:
425: Expressions (\ref{spectral.final}) and (\ref{zero.gamma}) present
426: the main result of the paper for the SF in the case of quantum
427: phase fluctuations. Before proceeding to the discussion of their
428: numerical integration, we recap briefly the case of classical
429: phase fluctuations.
430: The static SF, obtained in \cite{Gusynin.JETPLett}, is reproduced from
431: (\ref{zero.gamma}) after changing the variable
432: $\omega^\prime \to v \omega^\prime$ and taking formally the limit $v\to0$.
433: It was discussed in detail in \cite{Gusynin.JETPLett} and we would
434: like to stress here only two main points. The first is that the SF
435: is identically zero inside the gap ($A_{cl}(\omega, \mbox{\bf k})
436: =0$ for $|\omega| < \rho$) \cite{Evans,foot1}. The second is that
437: in addition to the usual quasiparticle peaks it reveals extra
438: peaks at $\omega = \pm \rho$.
439:
440: Let us now return to the discussion of the results of numerical
441: integration based on Eqs.~(\ref{spectral.final}) and
442: (\ref{zero.gamma}) which are shown in Fig.~\ref{fig:1}.
443:
444: \noindent a) For $T < T_{\rm BKT}$ there are two highly pronounced
445: quasiparticle peaks at $\omega = \pm E(\mbox{\bf k}) \equiv \pm
446: \sqrt{\xi^{2}(\mbox{\bf k}) + \rho^{2}}$. Since below $T_{\rm
447: BKT}$ both $\xi_{+}^{-1}$ and $\gamma$ are zero, the width of the
448: peaks is almost entirely controlled by $\alpha$ whose deviation
449: from 1 gives the non-Fermi liquid behavior of the GF
450: \cite{Gusynin.JETPLett}. Since the value of $\alpha$
451: is very close to 1 this non-Fermi liquid behavior is hardly
452: distinguishable from ordinary widening of the quasiparticle peaks
453: due to damping. Furthermore, because $\alpha \to 1$ as $T \to 0$
454: the width of the peaks is temperature
455: dependent so that the peaks become sharper as $T$ decreases resembling
456: the data of ARPES for the anti-node direction.\cite{Campuzano}
457:
458: We note also that $k \ne k_{F}$ in Fig.~\ref{fig:1} so that the
459: quasiparticle peaks are not equal to each other in contrast to the case
460: of symmetrized ARPES data \cite{Timusk}. The reason for choosing
461: $k \ne k_{F}$ in Fig.~\ref{fig:1} was to prove that
462: the extra peaks present in the SF calculated for the classical phase
463: fluctuations \cite{Gusynin.JETPLett,we.review} are now absent.
464:
465: The gap in the SF remains almost unfilled and has ``U''-like shape.
466: %which also fits experiment.
467: We stress, however, that in contrast
468: to the static case \cite{Gusynin.JETPLett} the SF
469: (\ref{spectral.final}) is nonzero even for $|\omega| < \rho$ and
470: this is evidently related to the presence of the quantum
471: (dynamical) phase fluctuations, i.e. the terms with $\Omega_n \neq
472: 0$. In particular, using Eq.~(\ref{zero.gamma}) one can check
473: analytically that even for $\gamma =0$ that due to these terms
474: $A(\omega = 0, {\bf k} ={\bf k}_{F}) \neq 0$.
475: We estimated the temperature $T_{cl}$ of the quantum to classical crossover
476: in the way similar to that of \cite{Randeria}.
477: For the low carrier densities this gives
478: $T_{cl} \sim T_{\rm BKT}$, while for the high carrier densities,
479: where the pseudogap phase shrinks \cite{Gusynin.JETP}
480: we obtain $T_{cl} \sim \rho(T=0) \sim T^{\ast}$.
481: Thus, the crossover temperature is always too large (in the absence of
482: dissipation) for classical fluctuations to play a significant role at
483: low $T$.
484:
485: \noindent b) For $T > T_{\rm BKT}$ the quasiparticle peaks at
486: $\omega \approx \pm E(\mbox{\bf k})$ are far less prominent. This
487: is caused by the fact that $\xi_{+}$ is now nonzero due to the
488: presence of vortices. Increasing the temperature further decreases
489: the value of $\xi_{+}$ so that the quasiparticle peaks become even
490: less pronounced. This behavior of the quasiparticle peaks
491: reproduces qualitatively the ARPES studies \cite{Campuzano,Feng}
492: for the anti-node direction which
493: show clearly that the quasiparticle SF broadens dramatically when
494: passing from the SC to normal state. The width of the
495: quasiparticle peaks is controlled primarily by $\xi_{+}$ above
496: $T_{\rm BKT}$ since this width remains practically constant as
497: $\gamma$ changes. Since $\xi_{+}$ is the phase coherence length,
498: the current model supports the premise \cite{Feng} that the
499: quasiparticle peaks are related to the phase coherence, not to
500: the energy gap. The deviation of $\alpha$ from 1 becomes, however,
501: sizable at $T \geq T_{BKT}$.
502:
503:
504: \noindent
505: c) As discussed above, for $T > T_{\rm BKT}$ the value
506: of $\gamma$ is nonzero and increases with increasing temperature.
507: This increase of $\gamma$ (typically up to values of the order
508: of $0.5 \mu$), together with the decrease of $\xi_+$, causes
509: filling of the gap and changes
510: its shape from ``U''-like to ``V''-like. In other words,
511: the quasiparticle peaks grow ``shoulders''
512: which eventually fill the gap.
513: %Thus filling of the gap is
514: %achieved by the dynamical phase fluctuations and the intensity of
515: %the filling is controlled by the value of $\gamma$ which
516: %contributes only when $\Omega_{n} \neq 0$.
517: We stress, however, that the increase of $\gamma$ in fact diminishes
518: the effect of the quantum phase fluctuations pushing
519: $T_{cl}$ down \cite{Randeria} which is seen from the
520: reducing of $A(\omega =0, {\bf k} ={\bf k}_F)$ as $\gamma$ grows
521: (see Fig.~1, where the bottom parts of the curves cross each other),
522: but the abovementioned ``shoulders'' simultaneously emerge.
523: Thus the filling of the gap at $T>T_{\rm BKT}$ in the present model is
524: caused by quantum phase fluctuations {\it in the presence} of nonzero
525: dissipation because the damping $\gamma$ contributes only when
526: $\Omega_n\neq 0$.
527:
528: \noindent d) Due to the smooth dependence of $\xi_{+}^{-1}$ and $\gamma$ on
529: $T$ as the temperature passes through $T_{\rm BKT}$ there is no sharp
530: transition at $T_{\rm BKT}$ in agreement with
531: experiment \cite{Timusk,Campuzano}.
532: A similar filling of the gap was obtained in \cite{Capezzali} for
533: $\gamma = 0.5 \mu$ where the correlator (\ref{correlator.dynamic}) was used
534: for the numerical computation of the self-energy of the fermions and the
535: subsequent extraction of the SF from the fermion GF. A recent Monte Carlo
536: simulation \cite{Tremblay} also shows similar behavior for the quasiparticle
537: peaks and the filling of the gap. However it is the analytical character of
538: the present work, which relies on the explicit introduction of the charge and
539: spin degrees of freedom (\ref{Nambu.spinor}) for the Nambu spinors, that
540: enables us to unambiguously state the correspondence between the parameters of
541: the model and the observed features of the SF (\ref{spectral.final}).
542: %More importantly, the results \cite{Capezzali,Tremblay} do not permit one to
543: %attribute the filling of the gap solely to the quantum phase
544: %fluctuations.\cite{foot1}
545:
546: In spite of some similarities between the results observed and
547: those just obtained, the latter are only illustrative since we have
548: considered a model with non-retarded $s$-wave attraction.
549: %and an isotropic fermion spectrum.
550: However, it is likely that for
551: $d$-wave pairing, the properties obtained can be used for the
552: description of the systems in the anti-node direction on the Fermi
553: surface.
554: It is however essential to consider fluctuations in the
555: modulus of the order parameter to extend the analysis to the nodal
556: directions on the Fermi surface.
557: The value of $\gamma = 0.5 \mu$
558: which results in a filled gap appears to be too large to be due to
559: the vortex-vortex interaction. This leads one to the conclusion
560: that the mechanism considered here for gap filling may well not be
561: the only possible mechanism and that other interactions which lead
562: to the filling of the gap above $T_c$, see \cite{Franz,Kwon,Tremblay.new},
563: are also important.
564:
565: In summary we studied the effect of the fluctuations
566: in the phase of the complex ordering field on the properties of a
567: 2D system with four-fermion attraction. The fermion SF has been
568: given as a single integral over a function known in closed form.
569: Through the use of analytical techniques, we have been able to
570: demonstrate that the quantum phase fluctuations in the presence of
571: dissipation lead to the filling of the pseudogap even if one
572: ignores the spin-charge coupling.
573: %Qualitative agreement with the SF of ARPES is also
574: %obtained.
575:
576: S.G.Sh thanks Prof.~H.~Beck and Dr.~M.~Capezzali for stimulating
577: discussion.
578: This work was partly supported by the NRF, South Africa (R.M.Q and S.G.Sh),
579: by SCOPES-project 7UKPJ062150.00/1 (V.P.G., V.M.L. and S.G.Sh) and
580: by the research project 2000-061901.00/1 (S.G.Sh)
581: of the Swiss National Science Foundation.
582:
583: \vspace*{-4ex}
584: \begin{references}
585: \vspace*{-11ex}
586:
587: \bibitem[\ast]{e-mail} E-mail: {\tt vloktev@bitp.kiev.ua}
588:
589: \bibitem[\dagger]{e-mail2} Sergei.Sharapov@unine.ch
590:
591: \bibitem{Timusk} For a recent review see,
592: T.~Timusk and B.~Statt, Rep.~Progr.~Phys.
593: 62 (1999) 61.
594:
595: \bibitem{Campuzano} See review, M.~Randeria and
596: J.C.~Campuzano, in Proceedings of the International School of Physics
597: ``Enrico Fermi'', Varenna, 1997 (IOS Press, Amsterdam, 1998).
598:
599: \bibitem{Feng} D.L.~Feng, D.H.~Lu, K.M.~Shen, C.~Kim, H.~Eisaki, A.~Damascelli,R.~Yoshizaki, J.-i.~Shimoyama, K.~Kishio, G.D.~Gu, S.~Oh, A.~Andrus,
600: J.~O'Donnell, J. N.~Eckstein, and Z.-X.~Shen, Science 280 (2000) 277.
601:
602: \bibitem{Corson}
603: J.~Corson, R.~Mallozzi, J.~Orenstein, J.N.~Eckstein, and I.~Bozovic,
604: Nature 398 (1999) 221.
605:
606: \bibitem{Pines} J.~Schmalian, D.~Pines, and B.~Stojkovic,
607: Phys.~Rev.~B 60 (1999) 667.
608:
609: \bibitem{Anderson} P.W.~Anderson,
610: {\em The theory of superconductivity in the high-$T_c$ cuprates},
611: Princeton Univ. Press, 1997.
612:
613: \bibitem{Emery} V.~Emery and S.A.~Kivelson,
614: Nature 374 (1995) 434;
615: Phys.~Rev.~Lett. 74 (1995) 3253.
616:
617: \bibitem{Gusynin.JETP}
618: V.P.~Gusynin, V.M.~Loktev, and S.G.~Sharapov,
619: JETP~Lett. 65 (1997) 182;
620: JETP 88 (1999) 685.
621:
622: \bibitem{Gusynin.JETPLett}
623: V.P.~Gusynin, V.M.~Loktev, and S.G.~Sharapov, JETP~Lett. 69
624: (1999) 141; JETP 90 (2000) 993.
625:
626: \bibitem{Evans} Although the calculation in
627: \cite{Gusynin.JETPLett} did not include impurities, even the
628: presence of nonmagnetic impurities will not lead to the filling of
629: the gap, see W.A.B.~Evans and G.~Rickayzen, Ann. of Physics (N.Y.)
630: 33 (1965) 275 and A.~Ghosal, M.~Randeria, and N.~Trivedi,
631: Phys.~Rev.~Lett. 81 (1998) 3940.
632:
633: \bibitem{Franz}
634: M.~Franz and A.J.~Millis,
635: Phys.~Rev.~B 58 (1998) 14572.
636:
637: \bibitem{Kwon}
638: H.J.~Kwon and A.T.~Dorsey,
639: Phys.~Rev.~B 59 (1999) 6438.
640:
641: \bibitem{Tremblay.new} B.~Kyung,
642: S.~Allen, and A.-M.S.~Tremblay,
643: preprint cond-mat/0010001.
644:
645: \bibitem{Capezzali}
646: M.~Capezzali and H.~Beck, Physica~B 259-261 (1999) 501; see also
647: M.~Capezzali, PhD thesis (1998), unpublished.
648:
649: \bibitem{Ran1} M.~Randeria, N.~Trivedi, A.~Moreo, and R.~Scalettar,
650: Phys.~Rev.~Lett. 69 (1992) 2001.
651:
652: \bibitem{Ran2} M.~Randeria and N.~Trivedi,
653: Phys.~Rev.~Lett. 75 (1995) 312.
654:
655: \bibitem{Babaev}
656: E.~Babaev and H.~Kleinert, Phys.~Rev.~B
657: 59 (1999) 12083.
658:
659: \bibitem{we.review} V.M.~Loktev, R.M.~Quick, and S.G.~Sharapov,
660: Phys. Rep. 349 (2001) 1.
661:
662: \bibitem{Tremblay}
663: S.~Allen, H.~Touchette, S.~Moukouri, Y.M.~Vilk, and A.-M.S.~Tremblay,
664: Phys.~Rev.~Lett. 83 (1999) 4128.
665:
666: \bibitem{Huber}
667: D.L.~Huber, Phys.~Lett.~A 68 (1978) 125.
668:
669: \bibitem{Levin}
670: B.~Janko, J.~Maly, and K.~Levin,
671: Phys.~Rev.~B 56 (1997) 11407.
672:
673: \bibitem{foot1} It is practically impossible to make such a statement
674: relying on numerical results. For example, using our own result
675: for the GF from \cite{Gusynin.JETPLett} calculated on the
676: imaginary axis to ten digits of precision, we tried to recover the
677: known static SF using the method of numerical analytic
678: continuation, H.J.~Vidberg and J.W.~Serene,
679: J.~Low~Temp.Phys. 29 (1977) 179. Even
680: for over 100 Matsubara frequencies the gap remained filled
681: numerically.\cite{we.review}
682:
683: \bibitem{Randeria} A.~Paramekanti,
684: M.~Randeria, T.V.~Ramakrishnan, and S.S. Mandal,
685: Phys.~Rev.~B 62 (2000) 6786.
686:
687:
688: \end{references}
689:
690: \begin{figure}
691: \centering{
692: \includegraphics[width=2.5in]{spectrum.eps}}
693: \caption{The spectral function $A(\omega, \mbox{\bf k})$ as a
694: function of $\omega$ in units of the zero temperature SC gap
695: $\Delta$ for $k > k_{F}$ taken at two different temperatures,
696: $T_{1} < T_{\rm BKT} < T_{2}$.} \label{fig:1}
697: \end{figure}
698:
699: \end{document}
700: