cond-mat0204380/pap.tex
1: %%%%%%%%%%%%%%%%%%%%%%% file template.tex %%%%%%%%%%%%%%%%%%%%%%%%%
2: %    
3: % This is a template file for The European Physical Journal
4: %
5: % Copy it to a new file with a new name and use it as the basis
6: % for your article
7: %
8: %%%%%%%%%%%%%%%%%%%%%%%% Springer-Verlag %%%%%%%%%%%%%%%%%%%%%%%%%%
9: %
10: \begin{filecontents}{leer.eps}
11: %!PS-Adobe-2.0 EPSF-2.0
12: %%CreationDate: Mon Jul 13 16:51:17 1992
13: %%DocumentFonts: (atend)
14: %%Pages: 0 1
15: %%BoundingBox: 72 31 601 342
16: %%EndComments
17: 
18: gsave
19: 72 31 moveto
20: 72 342 lineto
21: 601 342 lineto
22: 601 31 lineto
23: 72 31 lineto
24: showpage
25: grestore
26: %%Trailer
27: %%DocumentFonts: Helvetica
28: \end{filecontents}
29: 
30: \documentclass[epj]{svjour}
31: \usepackage{graphicx,axodraw}
32: \begin{document}
33: \title{Landau-Fermi liquid analysis of the  2D $t$-$t'$ Hubbard model}
34: \author{P. A. Frigeri\inst{1}\fnmsep \thanks{e-mail:
35:     \texttt{pfrigeri@itp.phys.ethz.ch}} \and C.
36:   Honerkamp\inst{1,2}\and T. M. Rice\inst{1}}
37: 
38: \institute{ Theoretische Physik,
39:  ETH-H\"onggerberg, CH-8093 Z\"urich, Switzerland \and
40:  Department of Physics, Massachusetts
41:  Institute of Technology, Cambridge MA 02139, USA \thanks{present address}
42: }
43: 
44: \date{\today}
45: 
46: \abstract{We calculate the Landau interaction function 
47: $f(\vec{k},\vec{k}')$ for the two-dimensional $t$-$t'$ Hubbard model on the
48: square lattice using second and higher order perturbation theory. Within the
49: Landau-Fermi liquid framework we discuss the behavior of spin and charge
50: susceptibilities as function of the onsite interaction and band filling. In
51: particular we analyze the role of elastic umklapp processes as driving force for the anisotropic reduction of the compressibility on parts of the Fermi surface.}
52: 
53: \PACS{
54:       {71.10.Fd}{Lattice fermion models (Hubbard model, etc.)}   \and
55: {74.72.Jt}{Other cuprates}
56:      } 
57: 
58: \authorrunning{Frigeri, Honerkamp, and Rice} 
59: \titlerunning{Landau-Fermi liquid analysis of the 2D $t$-$t'$ Hubbard model}
60: \maketitle
61:  
62: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
63: \section{Introduction}
64: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65: %
66: The  Hubbard model remains one of the most studied
67: theoretical models in the efforts to understand the high-tem\-perature
68: superconducting cuprates. Despite their  exceptionally high transition temperatures into
69: the superconducting state, much of the interest in these materials arises from 
70: their anomalous normal state properties over a large part of the
71: tem\-perature-doping phase diagram. These effects occur unequivocally and most
72: strongly  in the underdoped samples with electron densities close to  half
73: band filling where the system is a stoichiometric antiferromagnetic Mott insulator. 
74: In the underdoped regime, the transition into the superconducting state 
75: occurs out of the so-called pseudogap phase, 
76: which shows significant deviations from conventional metallic
77: Fermi-liquid behavior\cite{anderson,timusk}. This  is in 
78: contrast with sufficiently
79: overdoped samples where many observations point toward a conventional
80: transition from a Fermi-liquid into the superconducting state\cite{proust}.
81: There have been experimental and theoretical studies suggesting that a
82: quantum critical point hidden underneath the superconducting
83: dome\cite{tallon,sachdev,chakravarty}  or the onset of Fermi surface
84: truncation\cite{norman,furukawa,honerkamp} give rise to these phenomena. 
85: On the other hand one may try to take a more conservative route 
86: and attempt to understand the
87: experimental findings within a conventional Fermi-liquid framework. Typically, 
88: rather specific assumptions on the Landau interaction function 
89: have to be made\cite{millis} in order to obtain agreement 
90: with the experimental data.
91: Thus it appears useful to simply calculate within perturbation theory the Fermi-liquid properties of the
92: 2D Hubbard model and to see whether these assumptions can be justified within
93: this transparent theoretical framework.
94: 
95: Another motivation for this work is the question whether 
96: as the density is increased towards half filling 
97: some kind of drastic change of the low energy properties, 
98: e.g. corresponding to Fermi surface truncation as indicated by one-loop renormalization group 
99: calculations\cite{furukawa,honerkamp}, is foreshadowed already 
100: in a perturbative Fermi liquid picture. Such a partial 
101: truncation of the Fermi surface has been argued  to be 
102: seen in angular resolved photoemission
103: experiments \cite{norman}. 
104: It is a promising candidate to understand the many anomalies 
105: of the underdoped normal state of the high-$T_c$ cuprates, as it naturally
106: combines insulator-like features such as the pseudogap, the observed c-axis
107: resistivity and the strongly reduced superfluid weight, with residual metallic
108: properties as evidenced  by the in-plane transport and superconductivity itself.  
109: Since both the renormalization group 
110: approaches and the present Fermi liquid analysis are perturbative techniques, 
111: we will not obtain an accurate description of possible phases
112: with a truncated Fermi surface,
113: but we will carefully look for  insulator-like tendencies in these 
114: calculations as the interaction gets stronger. In particular we analyze 
115: the influence of umklapp processes between $\vec{k}$-space regions in the
116: vicinity of the saddle points at $(\pi,0)$ 
117: and $(0,\pi)$. Elastic umklapp scattering with momentum transfer of
118: $(\pi, \pi)$ across the Fermi surface is allowed when the Fermi surface 
119: extends to the Brillouin zone boundary for a certain density range 
120: close to half filling. 
121: 
122: A Fermi liquid analysis of the 2D Hubbard model close to half filling  
123: has been presented by Fuseya et
124: al.\cite{fuseya}. These authors studied the case with only nearest neighbor
125: hopping, $t$, and found a suppression of the uniform spin susceptibility for
126: sufficiently strong interaction close to half-filling, which maybe interpreted
127: as indicating the opening of a spin gap. 
128: Unfortunately this promising indication turns out to be an
129: artifact of the calculation scheme used by Fuseya et al. \cite{fuseya}, as
130: will be discussed below. From renormalization group studies\cite{halboth,honerkamp} 
131: we know
132: that nonzero values of the next nearest neighbor hopping, $t'$, can change the
133: weak coupling physics significantly. Therefore, in this paper we extend the analysis
134: of the Landau interaction function and related quantities to the case of
135: $t'\not= 0$.  Particular attention is paid to the anisotropy of the
136: interaction function for different parts of the Fermi surface.
137: 
138: In the following we first apply a second order perturbative method and discuss the
139: $\vec{k}$-space integrated and $\vec{k}$-space resolved Fermi liquid
140: properties of the 2D $t$-$t'$ Hubbard model. Next we describe how ladder summations in the
141: particle-particle and particle-hole channels, which are necessary to remove
142: unphysical divergences, modify the tendencies observed
143: in the second order approach. 
144: Finally we conclude and compare our results to
145: other approaches and experimental observations for the high-$T_c$ cuprates.
146: 
147: 
148: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
149: \section{Second order calculation of the quasiparticle interaction} 
150: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
151: %
152: The kinetic energy for  the $t$-$t'$ Hubbard model on the 2D square lattice
153: is given by 
154: %
155: \begin{equation}  
156: \epsilon_{\vec{k}}=-2t(\cos k_x + \cos k_y )+4t'\cos k_x \cos k_y \, ,  \label{eofk}
157: \end{equation}  
158: where $t$ and $t'$ denote the amplitudes for hopping between nearest and next-nearest neighbors, respectively.
159: %
160: %
161: The short range Coulomb interaction between the electrons is taken into
162: account by the usual onsite repulsion 
163: %
164: \begin{equation}  
165: H_I =  U \sum_{i} n_{i, \uparrow} n_{i, \downarrow} \, .  \label{U}
166: \end{equation} 
167: The particle number $N_e$ is controlled by a chemical potential $\mu$.
168: 
169: The Landau interaction function $f_{s,s'}(\vec{k}_F,\vec{k}'_F)$ can be
170: obtained \cite{theory-of-FL1} from the two-particle vertex 
171: $\Gamma_{ss'} (\vec{K},\vec{K}',\vec{K}+\vec{Q})$. Here $\vec{K}=(\vec{k},\nu)$
172: is a short notation for wavevectors and frequencies. $\Gamma$ is a function of
173: the interaction strength, $U$. To first order in $U$, $\Gamma$ is
174: wavevector independent, but higher orders lead to a pronounced wavevector dependence.
175: $\vec{K}$, $\vec{K}'$ and spin indices $s$,$s'$ characterize the two initial particles
176: and $\vec{K}+\vec{Q}$, $\vec{K}'-\vec{Q}$ and $s$,$s'$ their final
177: states, and $\vec{Q}=(\vec{q},\omega)$ denotes wavevector and frequency
178: transfer in the scattering process. The Landau quasiparticle interaction
179: $f(\vec{k}_F,\vec{k}'_F)$ corresponds to the vertex function in the limit
180: $\vec{Q} \rightarrow 0$ with $\frac{q}{\omega} \rightarrow 0$ with
181: $\vec{K}=(\vec{k}_F,0)$ and $\vec{K}'=(\vec{k}'_F,0)$. More precisely  
182: %
183: \begin{equation} \label{rel-vertex-int}  
184:   f_{s,s'}(\vec{k}_F,\vec{k}'_F)=  z_{k} z_{k'} 
185:   \lim_{\omega \rightarrow 0 } \lim_{\vec{q} \rightarrow 0} 
186:   \Gamma_{s,s'}(\vec{K},\vec{K}',\vec{K}+\vec{Q}),  \label{ffromGamma}
187: \end{equation}
188: %
189: %
190: where $\vec{k}_F$, $\vec{k}'_F$ are two wavevectors on the Fermi surface
191: and $z_k$ is the residue of the pole of the interacting one-particle Green's
192: function $G(\vec{k}_F,\nu)$, i.e. $z_k=(1-\delta \Sigma / \delta
193: \nu)^{-1}$ with self-energy $\Sigma(\vec{k},\nu)$.
194: The two-particle vertex is determined by the coupling function $V_{s
195:  s'}(\vec{K},\vec{K}',\vec{K}+\vec{Q})=V( \vec{K},s; \vec{K}',s'; \vec{K}+\vec{Q},s)$ via 
196: %
197: %
198: {\setlength \arraycolsep{2pt} 
199:   \begin{eqnarray}  \label{coupling}
200:     \Gamma_{s,s'}(\vec{K},\vec{K}',\vec{K}+\vec{Q})&=& \delta_{s,s'}[V_{s
201:       s'}(\vec{K},\vec{K}',\vec{K}+\vec{Q})\nonumber \\
202:     &-&V_{ss'}(\vec{K}',\vec{K},\vec{K}+\vec{Q})] \nonumber \\
203:     &+&\delta_{s,-s'}V_{ss'}(\vec{K},\vec{K}',\vec{K}+\vec{Q}),
204:   \end{eqnarray}  
205: %
206: which expresses the fact that $\Gamma$ has to be antisymmetric with respect to the exchange of two particles with the same spin orientation. 
207: Then the Landau interaction function can be obtained via
208: (\ref{coupling}) by taking the  limit (\ref{rel-vertex-int}) in the
209: perturbation expansion for $V_{ss'}(\vec{K},\vec{K}',\vec{K}+\vec{Q})$. 
210: The first and second order graphs contributing are shown
211: in Fig. \ref{graph1-vertex}. 
212: %
213: %
214: \begin{figure}[h] 
215: \begin{center} \begin{picture}(370,180)(-20,-100)
216: \SetScale{0.6}
217: \ArrowLine(20,20)(35,35)
218: \Text(10,5)[r]{\scriptsize{$\vec{K}$}}
219: \ArrowLine(35,35)(50,20)
220: \Text(30,5)[l]{\scriptsize{$\vec{K}+\vec{Q}$}}
221: \ArrowLine(20,100)(35,85)
222: \Text(30,70)[l]{\scriptsize{$\vec{K}'-\vec{Q}$}}
223: \ArrowLine(35,85)(50,100)
224: \Text(10,70)[r]{\scriptsize{$\vec{K}'$}}
225: \Photon(35,35)(35,85){4}{5}
226: \Text(27,37)[l]{\scriptsize{$(a)$}}
227: \Vertex(35,35){1.5}
228: \Vertex(35,85){1.5}
229: %
230: \ArrowLine(270,20)(285,35) %(250,0)
231: \Text(160,5)[r]{\scriptsize{$\vec{K}^{\bar \sharp}$}} %(150,)
232: \ArrowLine(285,35)(300,20)
233: \Text(180,5)[l]{\scriptsize{$\vec{K}^{\bar \sharp}+\vec{Q}$}}
234: \Photon(260,110)(310,110){4}{5.5}
235: \ArrowLine(255,115)(285,85)
236: \Text(180,77)[l]{\scriptsize{$\vec{K}^{\sharp}-\vec{Q}$}}
237: \ArrowLine(285,85)(315,115)
238: \Text(160,77)[r]{\scriptsize{$\vec{K}^{\sharp}$}}
239: \Photon(285,35)(285,85){4}{5}
240: \Text(177,37)[l]{\scriptsize{$(c)$}}
241: \Vertex(285,35){1.5}
242: \Vertex(285,85){1.5}
243: %
244: \ArrowLine(35,-105)(20,-120)
245: \Text(10,-78)[r]{\scriptsize{$\vec{K}+\vec{Q}$}}
246: \ArrowLine(120,-120)(105,-105)
247: \Text(75,-78)[l]{\scriptsize{$\vec{K}$}}
248: \ArrowLine(105,-55)(120,-40)
249: \Text(75,-18)[l]{\scriptsize{$\vec{K}'-\vec{Q}$}}
250: \ArrowLine(20,-40)(35,-55) 
251: \Text(10,-18)[r]{\scriptsize{$\vec{K}'$}}
252: \ArrowLine(105,-105)(35,-105)
253: \ArrowLine(35,-55)(105,-55)
254: \Text(42,-75)[c]{\scriptsize{$(d)$}}
255: \Photon(35,-105)(35,-55){4}{5}
256: \Photon(105,-105)(105,-55){4}{5}
257: \Vertex(35,-105){1.5}
258: \Vertex(35,-55){1.5}
259: \Vertex(105,-105){1.5}
260: \Vertex(105,-55){1.5}
261: %
262: \ArrowLine(210,-120)(225,-105)
263: \Text(125,-78)[r]{\scriptsize{$\vec{K}$}}
264: \ArrowLine(295,-105)(310,-120)
265: \Text(190,-78)[l]{\scriptsize{$\vec{K}+\vec{Q}$}}
266: \ArrowLine(295,-55)(310,-40)
267: \Text(190,-18)[l]{\scriptsize{$\vec{K}'-\vec{Q}$}}
268: \ArrowLine(210,-40)(225,-55)
269: \Text(125,-18)[r]{\scriptsize{$\vec{K}'$}}
270: \ArrowLine(225,-105)(295,-105)
271: \Text(157,-73)[c]{\scriptsize{$(e)$}}
272: \ArrowLine(225,-55)(295,-55)
273: \Photon(225,-105)(225,-55){4}{5}
274: \Photon(295,-105)(295,-55){4}{5}
275: \Vertex(225,-105){1.5}
276: \Vertex(225,-55){1.5}
277: \Vertex(295,-105){1.5}
278: \Vertex(295,-55){1.5}
279: %
280: \ArrowLine(150,20)(165,35)  %(120,160)
281: \Text(90,5)[r]{\scriptsize{$\vec{K}$}}
282: \ArrowLine(165,35)(180,20)
283: \Text(110,5)[l]{\scriptsize{$\vec{K}+\vec{Q}$}}
284: \ArrowLine(150,120)(165,105)
285: \Text(110,77)[l]{\scriptsize{$\vec{K}'-\vec{Q}$}}
286: \ArrowLine(165,105)(180,120)
287: \Text(90,77)[r]{\scriptsize{$\vec{K}'$}} %(80,95)
288: \Photon(165,35)(165,48.7){4}{1}
289: \Photon(165,105)(165,91.15){4}{1}
290: \Text(110,37)[l]{\scriptsize{$(b)$}}
291: \Vertex(165,35){1.5}
292: \Vertex(165,105){1.5}
293: \CArc(143.75,70)(30,-45,45)
294: \LongArrowArc(143.75,70)(30,-45,5)
295: \CArc(186.25,70)(30,135,225)
296: \LongArrowArc(186.25,70)(30,135,185)
297: \Vertex(165,92.15){1.5}
298: \Vertex(165,48.7){1.5}
299: \end{picture}
300: \end{center}
301: \caption{\footnotesize{ First and second order graphs contributing to 
302: $\Gamma$. We use the notations $\sharp=\{ \; ,'\}$ and 
303: $\bar \sharp=\{',\, \}$.}}
304: \label{graph1-vertex} 
305: \end{figure}
306: % 
307: First the Landau function will be evaluated to second order in $U$. 
308: Corrections from the $z_k$ factors are at least second order in $U$, 
309: therefore setting
310: $z_k=1$ is fully consistent up to second order. The same argument justifies
311: the neglect of effective mass corrections. Finally the second order
312: quasiparticle interaction function reads
313: %
314: {\setlength \arraycolsep{2pt} 
315:   \begin{eqnarray}  \label{f_second_order_upup}
316:     f_{\uparrow \uparrow}(\vec{k}_F,\vec{k}'_F)
317:     &=&f_{\downarrow \downarrow}(\vec{k}_F,\vec{k}'_F) \nonumber \\
318:     &=&U^2 \chi_{PH}(\vec{k}_F-\vec{k}'_F), \\
319:     \nonumber \\
320:     f_{\uparrow \downarrow}(\vec{k}_F,\vec{k}'_F)
321:     &=&f_{\downarrow \uparrow}(\vec{k}_F,\vec{k}'_F) \nonumber \\
322:     &=&U+U^2
323: (\chi_{PP}(\vec{k}_F+\vec{k}'_F)+\chi_{PH}(\vec{k}_F-\vec{k}'_F)),\nonumber \\
324:   \end{eqnarray}
325: %
326: with 
327: %    
328: {\setlength \arraycolsep{2pt} 
329:   \begin{eqnarray} 
330:     \label{chiPH}
331:     \chi_{PH}(\vec{q}_0)&=& \frac{1}{( 2\pi )^2} \int d^2p \,
332: \frac{n_{\vec{p}}-n_{\vec{q}_0+\vec{p}}}
333:     {\epsilon_{\vec{q}_0+\vec{p}}-\epsilon_{\vec{p}}}.
334:   \end{eqnarray} 
335: %
336: and
337: %
338: {\setlength \arraycolsep{2pt} 
339:   \begin{eqnarray} 
340:     \chi_{PP}(\vec{q}_0)&=& \frac{1}{( 2\pi )^2} \int d^2p \,
341: \frac{1-n_{\vec{p}}-n_{\vec{q}_0-\vec{p}}}{2\mu -
342:       \epsilon_{\vec{p}}-\epsilon_{\vec{q_0}-\vec{p}}}. \nonumber \\
343:   \end{eqnarray} 
344: Here $n_p=1/(1 + \exp(\epsilon_p-\mu)/T)$ is the Fermi
345: distribution at temperature $T$.
346: 
347: For the analysis of the physical properties of the interacting Fermi liquid 
348: it is convenient to introduce symmetric ($f^s$) and  antisymmetric parts 
349: ($f^a$) of the Landau quasiparticle interaction by
350: %
351:   \begin{equation}  
352:     f^{s,a} = \frac{1}{2}(f_{\uparrow \uparrow}\pm f_{\downarrow \uparrow}).
353:   \end{equation}  
354: %
355: 
356: $f^s(\vec{k}_F, \vec{k}'_F)$ and $f^a(\vec{k}_F, \vec{k}'_F)$ are
357: computed for a fixed value of the chemical potential $\mu$.
358: $\vec{k}_F$ and $\vec{k}'_F$ are parame\-trized by the two angles $\theta$ and
359: $\theta'$, measured from the symmetry axis $\Gamma X$, as shown in the inset 
360: of Fig. \ref{loc_suc_2o}. Fig.\ref{Land_int_grap} shows the results
361: obtained for the parameters $t'/t = 0.3$, $\mu/t'=-3.83$, 
362: $T/t=0.002$ and $U/t=1$. Note that for this choice of the chemical potential
363: $\mu/t'>-4$, the Fermi surface consists of four arcs which lead to the breaks
364: in the curves. The most prominent feature is the divergence along a diagonal
365: line with $\pi < \theta' < 3\pi/2$ in both functions. This reflects simply the
366: cooper divergence in second order perturbation theory and it is removed by
367: summing the ladder graphs in the particle-particle channel as discussed below.
368: % 
369: \begin{figure}[h] 
370:   \begin{center}
371:     \includegraphics[width=.5 \textwidth]{fs_115.eps} %Landau115s.m
372:     \includegraphics[width=.5 \textwidth]{fa_115.eps} %Landau115a.m
373:     \caption{\footnotesize{ Numerical results for the symmetric 
374:  $f^s(\theta,\theta')$
375:     and antisymmetric part  $f^a(\theta,\theta')$  of the Landau
376:     interaction obtained by the second order perturbation approximation  for
377:     the parameters $t'=0.3t$, $\mu=-3.83t'$, $U=1t$ and $T=0.002t$. To the second order $f^{a,s}(\vec{k}_F,-\vec{k}_F)$ show unphysical
378:     divergences (sharp diagonal features for $\pi < \theta' < 3\pi/2$).}}   
379:   \label{Land_int_grap}
380:   \end{center}
381: \end{figure} 
382: 
383: 
384: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
385: \section{Total charge and spin susceptibility for the almost isotropic case}
386: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
387: %
388: %
389: In a Fermi liquid the static uniform charge and spin susceptibilities
390: $X_c$ and $X_s$ involve only excitations in the neighborhood
391: of the Fermi surface. Therefore they can be calculated as Fermi surface 
392: integrals over
393: $\vec{k}$-space local quantities $\chi_c(\vec{k}_F)$ and $\chi_s(\vec{k}_F)$
394: which define local susceptibilities for Fermi surface points described by the
395: wavevector $\vec{k_F}$,
396: %
397:   \begin{eqnarray}  
398:     \rho^2 X_c &=& \frac{1}{\Omega} \frac{dN_e}{d\mu}= 
399:     \int dl_F 
400:     \, \chi_c(\vec{k}_F){} \label{chic}, \\
401:     \frac{X_s}{\mu_B^2} &=& \frac{1}{\mu_B^2 \Omega} \frac{dM}{dH}= \int
402:     dl_F \, \chi_s(\vec{k}_F).  \label{chis} 
403:   \end{eqnarray}  
404: %
405: Here, $\int  dl_{F}$ denotes the line
406: integral along the Fermi surface and $\Omega$ is the total volume. 
407: $N_e$ denotes the number of electrons and $\rho=N_e/\Omega$ denotes their density. $\mu_B$ is the Bohr magneton.
408: The $\vec{k}$-space local susceptibilities are solutions of two
409: inhomogeneous linear equations,
410: 
411: \begin{equation} \label{dkfdmu}  
412:   L^{s}( \chi_{c})=2, \qquad  L^{a}(\chi_{s})=2,
413: \end{equation}
414: % 
415: which express via the two linear integral operators
416: $L^{s,a}$,
417: %
418: %{\setlength \arraycolsep{2pt} 
419:   \begin{eqnarray}
420:     \frac{L^{s,a}(\chi)}{(2\pi)^2}&=& v_F \chi(\vec{k}_F) \,  
421:     + \int \frac{dl_F'}{2\pi^2} \; f^{s,a}(\vec{k}_F,\vec{k}'_F)\,
422:     \chi(\vec{k}'_F), \nonumber  \label{lineq} \\ 
423:   \end{eqnarray}  
424: %
425: how the local susceptibilities are renormalized by the Landau
426: $f$-function \cite{theory-of-FL1}.
427: Here, $v_F \equiv v_{\vec{k}_F}$ denotes the magnitude of the Fermi velocity   
428: $\vec{v}_{\vec{k}_F}=\partial \epsilon_{\vec{k}}/ \partial \vec{k}|_{\vec{k}=\vec{k}_F}$    
429: of a quasiparticle with wavevector $\vec{k}$.  
430: %
431: For an isotropic Fermi surface and isotropic interactions, 
432: $\chi_c(\vec{k})$ and $\chi_s(\vec{k})$ are obviously two constants 
433: independent of $\vec{k}$. This
434: allows one to express $X_c$ and  $X_s$ using
435: the conventional Landau parameters $F_{0}^{s,a}$, 
436: which are defined as  zeroth order coefficients in the expansion 
437: of the Landau interaction functions ${f^{s,a}}$ in normalized 
438: Legendre polynomials. 
439: In the anisotropic case the Landau interaction functions can 
440: be expanded in a similar way,
441: %
442: \begin{equation}  
443:   f^{s,a}(\vec{k}_F,\vec{k}'_F)=\frac{1}{N_F}\sum_{n=0}^{\infty}
444:   \sum_{n'=0}^{\infty} F_{n,n'}^{s,a} \psi_n(\vec{k}_F)  \psi_{n'}(\vec{k}'_F)
445:   \, , 
446: \end{equation}       
447: %
448: where $\psi_0(\vec{k}_F)=1$, $\psi_n$ are orthogonal functions on the Fermi 
449: surface and $N_F$ is the density of states at the Fermi surface. Note that
450:  in general
451: the off-diagonal terms $F_{n,n'}^{s,a}= F_{n',n}^{s,a}$ do not vanish when 
452: both $\psi_n$ and $\psi_n'$ belong to the same irreducible representation 
453: of the crystal symmetry group.
454: 
455: 
456: Initially, let us assume that the isotropic components of $f^{s,a}$,
457: %
458: \begin{equation}  
459:   \label{Landau_par}
460:   F_0^{s,a}=\frac{ \int \int \frac{dl_{F}}{v_F} 
461:     \frac{dl'_{F}}{v'_F}
462:     f^{s,a}(\vec{k}_F, \vec{k'}_F)} {2\pi^2 \int 
463:     \frac{dl_F}{v_F}} \equiv F_{0,0}^{s,a},
464: \end{equation}    
465: %
466: %
467: are dominant. This assumption was also made
468: in Ref. \cite{fuseya}. In this case the compressibility (\ref{chic}) 
469: simplifies to
470: %
471: %
472: {\setlength \arraycolsep{2pt} 
473:   \begin{eqnarray}  
474:     \rho^2 X_c &=& N_F -  F^s_0 
475:     \int dl'_{F} \chi_c(\vec{k}'_F){} \nonumber \\
476:     &=& N_F -  F^s_0 \rho^2 X_c{} \, . \nonumber  \\ 
477:   \end{eqnarray}  
478: %
479: %                                
480: Thus we arrive at the well-known result for an isotropic Fermi liquid,
481:  \begin{equation}    \rho^2 X_c(\mu) = \frac{ N_F}{1+F^s_0(\mu)} \, . \label{comp_isot} \end{equation}
482: Using the same simplification we obtain for the spin susceptibility 
483: %
484: {\setlength \arraycolsep{2pt} 
485:   \begin{eqnarray}  
486:     \frac{X_s(\mu)}{\mu_B^2} &=& \frac{ N_F }{1+F^a_0(\mu)}.  \label{susc_isot}
487:   \end{eqnarray}  
488: Fuseya et al. \cite{fuseya} evaluated charge and spin susceptibilities 
489: for the Hubbard model with  $t'=0$ assuming that 
490: the isotropic components of $f^{s,a}$ were dominant.  
491: Their results showed a reduction of the uniform spin susceptibility 
492: for $U>2t$ at density near to half-filling. It turns out that this effect 
493: is tied the van Hove band filling where the density of states at the 
494: Fermi surface diverges due to van Hove singularities at the 
495: saddle points $(\pi, 0 )$ and $(0,\pi)$.
496: In the special case $t'=0$ the reduction of the spin susceptibility
497:  does not have a simple physical interpretation. 
498: In this case the van Hove filling coincides 
499: with half-filling, where the system is believed to be an antiferromagnetic 
500: insulator and the Fermi liquid approach must break down at low energy scales. 
501: But for finite next-nearest neighbor hopping $t'\not= 0$ 
502: the van Hove filling does not coincide anymore with
503: the half-filled or perfectly nested case and in the vicinity of this
504: filling the system may be in a paramagnetic state. In this case a
505: reduction of the uniform spin susceptibility 
506: for $U>2t$ could be interpreted as an indication for an opening of a
507: spin gap in the Hubbard model close to half filling. 
508: %
509: \begin{figure}[h] 
510:   \begin{center}
511:     \includegraphics[width=.5 \textwidth]{t_susc_00.eps} %exact_SC00.m%
512:     \includegraphics[width=.5 \textwidth]{t_susc_03.eps} %exact_SC03.m% 
513:     \caption{\footnotesize{Comparison between the uniform spin 
514:         susceptibility $X_s$ as function of the chemical potential $\mu$ 
515:         obtained from the approximate relations (\ref{susc_isot}) 
516:         (plots on the left) and 
517:        from the solution of the full integral equations 
518:         (\ref{dkfdmu},\ref{chis}) (plots on the right),
519:         for different values of interaction $U$: $*$ $U/t=1$, $\circ$ $U/t=2$, 
520:         $\bigtriangledown$ $U/t=3$, $\bigtriangleup$
521:         $U/t=4$. For all plots $T/t=0.002$. The susceptibility for the 
522: non-interacting system ($U/t=0$) is shown as dashed line. 
523:         }}
524:     \label{comparison}
525: \end{center}
526: \end{figure}   
527: %
528: However, our results obtained by 
529: solving the full integral equations (\ref{dkfdmu}), without the
530: assumption  of a dominance of the $n=0$ terms,  unfortunately remove this
531: promising finding. 
532: This is shown in Fig. \ref{comparison}, where the data assuming that 
533: the isotropic terms are dominant (left panels) are compared 
534: with those obtained by
535: solving (\ref{chis}, \ref{dkfdmu}) (right panels). In particular in 
536: the two figures to the right 
537: the spin susceptibility does not decrease in any significant way as the van
538: Hove filling ($\mu=0$ for $t'=0$ and $\mu/t'=-4$ for $t'=0.25t$) is 
539: approached. For the charge susceptibility the qualitative behavior of the
540: results obtained by solving (\ref{chic}, \ref{dkfdmu}) is compatible 
541: with the one obtained by Fuseya, even if a quantitative difference is 
542: present, see Fig.
543: \ref{comparison1}.
544: %
545: \begin{figure}[h] 
546:   \begin{center}
547:     \includegraphics[width=.5 \textwidth]{t_comp_00.eps} %exact_CC00.m%
548:     \includegraphics[width=.5 \textwidth]{t_comp_03.eps} %exact_CC03.m% 
549:     \caption{\footnotesize{Same as in Fig. \ref{comparison}, but for
550:         the charge susceptibility $X_c$.
551:         }}
552:     \label{comparison1}
553: \end{center}
554: \end{figure}   
555: 
556: Thus the conclusion of Fuseya et al.\cite{fuseya} that indications
557: for a spin gap in the 2D Hubbard model can already be observed in second
558: order perturbation theory seems to be an artifact of their
559: approximation. The solution of the full integral equation reveals that 
560: the assumption of a dominant $l=0$ Landau parameter is not 
561: always justified.
562: %
563: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
564: \section{Local spin and charge susceptibilities in $\vec{k}$-space}
565: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
566: %
567: Since the Landau interaction functions are very anisotropic,
568: it is interesting to analyze how the local values of the static
569: susceptibilities change with the interaction $U$. As mentioned in the introduction, we are particularly interested in the question whether 
570: there are tendencies 
571: towards a Fermi surface truncation, i.e. the formation of incompressible 
572: $\vec{k}$-space regions at the saddle points, as suggested by one-loop 
573: renormalization group studies, already within a low order Fermi liquid treatment.
574: For the non-interacting system, 
575: $\chi_c(\vec{k}_F)$ and $\chi_s(\vec{k}_F)$ are proportional 
576: to  the inverse velocity $1/v_{\vec{k}_F}$.
577: Since $v_{\vec{k}_F}$ vanishes at the saddle points, 
578: a divergence appears in the local spin and charge 
579: susceptibilities around these points at the
580: van Hove filling - a behavior opposite to incompressibility.
581: The  results for the interacting system obtained with the second order 
582: Landau function show a general 
583: reduction of the local values of the susceptibilities with increasing 
584: interaction strength $U$.
585: %
586: \begin{figure}[h] 
587:   \begin{center}
588:     \includegraphics[width=.5 \textwidth]{loc_sus_2o.eps}
589:     \caption{\footnotesize{$\vec{k}$-space local values of $\chi_s(\vec{k}_F)$
590:         and $\chi_c(\vec{k}_F)$ for $\vec{k}_F$ near a saddle point
591:         for different values of  $U$: $U/t=1$ thick line, $U/t=2$  thick dashed
592:         line, $U/t=3$ thin line, $U/t=4$ thin dashed line. The inset
593:         shows how the angle $\theta$ is used to parameterize the points 
594:         of the Fermi surface ($\vec{k}_F(\theta)$). Results computed for
595:         $t'=0.3$, $\mu/t'=-3.99$, and $T/t=0.002$.}}
596:     \label{loc_suc_2o}
597: \end{center}
598: \end{figure}        
599: %
600: Fig. \ref{loc_suc_2o} shows $\chi_s(\vec{k}_F)$ and $\chi_c(\vec{k}_F)$ 
601: when $\vec{k}_F$ is close to the saddle point. The inset in the spin
602: susceptibility graph indicates how the angle $\theta$ is used to 
603: parameterize the points of the Fermi surface, $\vec{k}_F(\theta)$.
604: Particularly interesting is the fact that near the saddle point the reduction
605: of the local spin susceptibility is pronounced. 
606: The charge susceptibility shows a similar behavior, but in a less marked way.
607: %
608: \begin{figure}[h] 
609:   \begin{center}
610:     \includegraphics[width=.48 \textwidth]{susc_rel197.eps} %susc_rel197.m
611:     \caption{\footnotesize{Relative variations of the $\vec{k}$-space 
612: local spin and charge susceptibilities ($\chi_s(U)$ and  $\chi_c(U)$) near the van Hove 
613:         filling for three different values of the angle $\theta$ as 
614:         function of the interaction $U$. $\theta=\theta_{min}$ thick line,
615:         $\theta \approx \pi/8$ thin dashed line, $\theta=\pi/4$
616:         thin line. T/t=0.002. }}
617:     \label{susc_rel197}
618: \end{center}
619: \end{figure}   
620: %
621: In fact the comparison between
622: the two graphs of the Fig. \ref{susc_rel197} shows that the relative value of
623: $\chi_s(\vec{k}_F)$ decreases more rapidly near the saddle points
624: ($\theta=\theta_{min}$) than at other points on the
625: Fermi surface (e.g. $\theta \approx \pi/8$ and $\theta=\pi/4$), while for $\chi_c(\vec{k}_F)$ the situation is reversed. 
626: %
627: \begin{figure}[h] 
628:   \begin{center}
629:     \includegraphics[width=.48 \textwidth]{susc_rel.eps}  %susc_rel.eps
630:     \caption{\footnotesize{Left plots: relative 
631:         variation of the local spin and charge  
632:         susceptibilities ($\chi_s(U)$,  $\chi_c(U)$) for three different 
633:         values of the angle $\theta$ as 
634:         function of the interaction $U$. $\theta=\theta_{min}$ thick line,
635:         $\theta \approx \pi/8$ thin dashed line, $\theta=\pi/4$
636:         thin line. $T/t=0.002$. The band filling is larger than the van Hove filling and elastic umklapp processes at the Fermi surface are possible.
637:  Right plots: results obtained excluding 
638: all umklapp processes from the Landau interaction
639:         functions. }}
640:     \label{susc_rel}
641:   \end{center}
642: \end{figure}    
643: %
644: 
645: The behavior of the susceptibilities for values of $\mu$ further away from the van Hove filling is noteworthy, too. 
646: In particular the qualitative difference 
647: of the susceptibilities between the case $\mu > -4t'$ (density larger than van Hove filling) and $\mu < -4t'$ (density smaller than van Hove filling), deserves to be noticed.
648: In the first case the two graphs on the left in Fig. \ref{susc_rel} show
649: that the relative reduction of the local susceptibilities 
650: is stronger for Fermi surface points near to $X M$ ($\theta=\theta_{min}$), compared to Fermi surface points closer to the diagonal $\Gamma M$ ($\theta \approx \pi/8$, 
651: $\theta=\pi/4$). The situation
652: changes if the calculations are done excluding umklapp processes of all kinds 
653: from the second order contributions, see the two graphs on the right 
654: in Fig. \ref{susc_rel}. This last fact already indicates the importance of elastic umklapp scattering at the Fermi surface that is allowed for densities larger than the van Hove density.
655: %
656: \begin{figure}[h] 
657:   \begin{center}
658:     \includegraphics[width=.48 \textwidth]{susc_rel1.eps} %susc_rel1.m
659:     \caption{\footnotesize{Same as in Fig. \ref{susc_rel} for band filling less than the van Hove filling such that elastic umklapp processes at the Fermi surface are absent. }}
660:  \label{susc_rel1}
661:  \end{center}
662: \end{figure}
663: %
664: The second case is when $\mu < -4t'$ 
665: and elastic umklapp scattering does not enter the low energy physics. 
666: The two graphs on the left in
667: Fig. \ref{susc_rel1} show a qualitative behavior of the local spin
668: susceptibility that is similar to that observed for $\mu > -4t'$. 
669: But the strongest reduction of the local charge susceptibility is now shifted
670: to $\theta \approx \pi/8$ and $\theta=\pi/4$ and the role of the umklapp processes becomes marginal at these densities.
671: 
672: %
673: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%   
674: \section{ Ladder summations}
675: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
676: %
677: The results in the previous sections were 
678: obtained with the Landau interaction function
679: calculated to second order in $U$. Although we observed a general 
680: suppression of the charge and spin susceptibility with respect to 
681: the non-interacting values, the effects were weak and no clear 
682: tendencies towards partial incompressibility could be found. 
683: In this section we extend the perturbative calculation of the 
684: Landau function to summations of infinite numbers of diagrams. 
685: Our motivation for doing this is twofold: {\em a)}
686: A severe defect of the second order approach is that
687: the contribution of the $s$-wave pairing channel to the interaction
688: between the quasiparticles is typically overestimated to this order. 
689: In fact $f_{a}(\vec{k}_F,\vec{k}'_F)= -U^2\chi_{PP}
690: (\vec{k}_F+\vec{k}'_F)$ diverges at $T=0$ when $\vec{k}'_F 
691: \rightarrow -\vec{k}_F$ due to the singular behavior
692: of the Cooper graph in systems with parity $\epsilon(\vec{k}) =
693: \epsilon(-\vec{k})$, that is clearly visible from the Fig. 
694: \ref{Land_int_grap}. In the following we will remedy this defect by summing up the ladder diagrams generated by the particle-particle 
695: graphs, the result is shown in the Fig. \ref{LAD_int_grap}. 
696: In this way for $U>0$ the net contribution of the Cooper channel 
697: at $T=0$ will be zero.
698: %
699: \begin{figure}[h] 
700:   \begin{center}
701:     \includegraphics[width=.499 \textwidth]{LADfs_115.eps}  %LADLandau115.m
702:     \includegraphics[width=.5 \textwidth]{LADfa_115.eps}  %LADLandau115a.m
703:     \caption{\footnotesize{Numerical solution of the 
704:         symmetric $f^s(\theta,\theta')$ and antisymmetric 
705:         part  $f^a(\theta,\theta')$ of the Landau
706:         interaction obtained by the ladder and bubble summation for 
707:         $t'=0.3t$, $\mu=-3.83t'$,  $U=1.5t$ and $T=0.04t$. 
708: The unphysical
709:     divergences of $f^{a,s}(\vec{k}_F,-\vec{k}_F)$ found with the second order
710:     approximation in Fig. \ref{Land_int_grap} are suppressed by the ladder summation.}}   
711:     \label{LAD_int_grap}
712:   \end{center}
713: \end{figure}
714: %
715: {\em b)}
716: Furthermore we know that the Hubbard model close to half filling is close to
717: an antiferromagnetic instability. This effect can be modeled in a perturbative
718: way by summing up an infinite number of bubble and ladder diagrams in the
719: particle-hole channel, as described below. By this we can enhance the
720: proximity to the AF instability and simultaneously study whether the local charge and spin susceptibilities exhibit significant changes. 
721: We should add that our calculation does not aim at a quantitative determination of the phase diagram of the model. Therefore we do not worry too much about the question whether we neglect certain contributions by selecting this particular class of diagrams. Our goal is rather to include and enhance a specific type of scattering processes and to determine the effects of these processes on the Fermi liquid interaction and derived quantities. In the same spirit we neglect possible self energy corrections in these calculations. Within the Fermi liquid framework this concerns mainly the effective mass of the quasiparticles, but a consistent treatment should also take into account  the Fermi surface shift caused by the interactions and lifetime effects on the virtual excitations in the ladder diagrams. However, since generally in spatial dimensions higher than one the self energy does not develop singularities without clear signs in the interactions and since we seek for qualitative results that do not depend on details of the model, we ignore the selfenergy effects and focus on the interaction function. This has the advantage that the analysis remains relatively simple and lucid. 
722: 
723: The graphs contributing to the particle-particle 
724: and particle-hole ladder and bubble summations are shown
725: in Fig. \ref{graph2-Ladder}. The only graph
726: contributing to $f_{\uparrow,\uparrow}(\vec{k}_F,\vec{k}'_F)$ to the second
727: order is the one-loop particle-hole bubble. This implies that the 
728: ladder summation for
729: $f_{\uparrow,\uparrow}$ involves only odd powers of the bubble graph.    
730: %
731: \begin{figure}[h] 
732: \begin{center} \begin{picture}(370,200)(-20,-60)
733: \SetScale{0.6}
734: %
735: \ArrowLine(35,150)(35,165)  
736: \Text(9,99)[r]{\scriptsize{$f_{\uparrow \uparrow}=$}}
737: \Text(21,115)[c]{\scriptsize{$\vec{k}_F \uparrow$}}
738: \ArrowLine(35,165)(35,180)
739: \Text(21,85)[c]{\scriptsize{$\vec{k}'_F \uparrow$}}
740: \ArrowLine(105,150)(105,165)
741: \Text(63,115)[c]{\scriptsize{$\vec{k}'_F \uparrow$}}
742: \ArrowLine(105,165)(105,180)
743: \Text(63,85)[c]{\scriptsize{$\vec{k}_F \uparrow$}} 
744: \Photon(35,165)(48.7,165){4}{1}
745: \Photon(105,165)(91.15,165){4}{1}
746: \Text(42,110)[c]{\scriptsize{$\downarrow$}}
747: \Vertex(35,165){1.5}
748: \Vertex(105,165){1.5}
749: \CArc(70,143.75)(30,45,135)
750: \LongArrowArc(70,143.75)(30,45,95)
751: \CArc(70,186.25)(30,225,315)
752: \LongArrowArc(70,186.25)(30,225,275)
753: \Vertex(92.15,165){1.5}
754: \Vertex(48.7,165){1.5}
755: \Text(75,99)[c]{\scriptsize{$+$}}
756: %
757: %(100)(100)
758: \ArrowLine(145,150)(145,165)
759: \Text(87,115)[c]{\scriptsize{$\vec{k}_F \uparrow$}}
760: \ArrowLine(145,165)(145,180)
761: \Text(87,85)[c]{\scriptsize{$\vec{k}'_F \uparrow$}}
762: \Text(197.2,115)[c]{\scriptsize{$\vec{k}'_F \uparrow$}}
763: \Text(197.2,85)[c]{\scriptsize{$\vec{k}_F \uparrow$}} 
764: \Photon(145,165)(158.7,165){4}{1}
765: \Photon(215,165)(201.15,165){4}{1}
766: \Text(108,110)[c]{\scriptsize{$\downarrow$}}
767: \Text(141.6,110)[c]{\scriptsize{$\uparrow$}}
768: \Text(175.2,110)[c]{\scriptsize{$\downarrow$}}
769: \Vertex(145,165){1.5}
770: \Vertex(215,165){1.5}
771: \CArc(180,143.75)(30,45,135)
772: \LongArrowArc(180,143.75)(30,45,95)
773: \CArc(180,186.25)(30,225,315)
774: \LongArrowArc(180,186.25)(30,225,275)
775: \Vertex(201.15,165){1.5}
776: \Vertex(157.7,165){1.5}
777: \Photon(271,165)(257.15,165){4}{1}
778: \Vertex(271,165){1.5}
779: \CArc(236,143.75)(30,45,135)
780: \LongArrowArc(236,143.75)(30,45,95)
781: \CArc(236,186.25)(30,225,315)
782: \LongArrowArc(236,186.25)(30,225,275)
783: \Vertex(257.15,165){1.5}
784: \Photon(327,165)(313.15,165){4}{1}
785: \Vertex(327,165){1.5}
786: \CArc(292,143.75)(30,45,135)
787: \LongArrowArc(292,143.75)(30,45,95)
788: \CArc(292,186.25)(30,225,315)
789: \LongArrowArc(292,186.25)(30,225,275)
790: \Vertex(313.15,165){1.5}
791: \ArrowLine(327,165)(327,180)
792: \ArrowLine(327,150)(327,165)
793: \Text(201,99)[l]{\scriptsize{$+\cdots$}}
794: %
795: \ArrowLine(35,50)(35,65)  
796: \Text(9,39)[r]{\scriptsize{$f_{\uparrow \downarrow}=$}}
797: \Text(21,55)[c]{\scriptsize{$\vec{k}_F \uparrow$}}
798: \ArrowLine(35,65)(35,80)
799: \Text(21,25)[c]{\scriptsize{$\vec{k}_F \uparrow$}}
800: \ArrowLine(105,50)(105,65)
801: \Text(63,55)[c]{\scriptsize{$\vec{k}'_F \downarrow$}}
802: \ArrowLine(105,65)(105,80)
803: \Text(63,25)[c]{\scriptsize{$\vec{k}'_F \downarrow$}} 
804: \Photon(35,65)(105,65){4}{5}
805: \Vertex(35,65){1.5}
806: \Vertex(105,65){1.5}
807: \Text(75,39)[c]{\scriptsize{$+$}}
808: %
809: \Line(145,50)(145,80)
810: \ArrowLine(145,80)(145,95)
811: \ArrowLine(145,35)(145,50)
812: \Text(87,65)[c]{\scriptsize{$\vec{k}_F \uparrow$}}
813: \Text(87,15)[c]{\scriptsize{$\vec{k}_F \uparrow$}}
814: \Line(215,50)(215,80)
815: \ArrowLine(215,80)(215,95)
816: \ArrowLine(215,35)(215,50)
817: \Text(129,65)[c]{\scriptsize{$\vec{k}'_F \downarrow$}}
818: \Text(129,15)[c]{\scriptsize{$\vec{k}'_F \downarrow$}} 
819: \Photon(145,50)(215,50){4}{5}
820: \Photon(145,80)(215,80){4}{5}
821: \Vertex(145,50){1.5}
822: \Vertex(215,50){1.5}
823: \Vertex(145,80){1.5}
824: \Vertex(215,80){1.5}
825: \Text(141,39)[c]{\scriptsize{$+$}}
826: %
827: \Line(257,50)(257,80)
828: \ArrowLine(257,80)(257,95)
829: \ArrowLine(257,35)(257,50)
830: \Text(154.2,65)[c]{\scriptsize{$\vec{k}_F \uparrow$}}
831: \Text(154.2,15)[c]{\scriptsize{$\vec{k}_F \uparrow$}}
832: \ArrowLine(327,50)(327,80)
833: \ArrowLine(327,80)(327,95)
834: \ArrowLine(327,35)(327,50)
835: \Text(196.2,65)[c]{\scriptsize{$\vec{k}'_F \downarrow$}}
836: \Text(196.2,15)[c]{\scriptsize{$\vec{k}'_F \downarrow$}} 
837: \Photon(257,50)(327,50){4}{5}
838: \Photon(257,65)(327,65){4}{5}
839: \Photon(257,80)(327,80){4}{5}
840: \Vertex(257,65){1.5}
841: \Vertex(327,65){1.5}
842: \Vertex(257,50){1.5}
843: \Vertex(327,50){1.5}
844: \Vertex(257,80){1.5}
845: \Vertex(327,80){1.5}
846: \Text(141,39)[c]{\scriptsize{$+$}}
847: \Text(201,39)[l]{\scriptsize{$+\cdots$}}
848: %
849: \SetScaledOffset(0,-110)
850: \Line(145,50)(145,80)
851: \ArrowLine(145,95)(145,80)
852: \ArrowLine(145,50)(145,35)
853: \Line(215,50)(215,80)
854: \ArrowLine(215,80)(215,95)
855: \ArrowLine(215,35)(215,50)
856: \Photon(145,50)(215,50){4}{5}
857: \Photon(145,80)(215,80){4}{5}
858: \Vertex(145,50){1.5}
859: \Vertex(215,50){1.5}
860: \Vertex(145,80){1.5}
861: \Vertex(215,80){1.5}
862: 
863: \Line(327,50)(327,80)
864: \Line(257,50)(257,80)
865: \ArrowLine(257,95)(257,80)
866: \ArrowLine(257,50)(257,35)
867: \ArrowLine(327,80)(327,95)
868: \ArrowLine(327,35)(327,50)
869: \Photon(257,50)(327,50){4}{5}
870: \Photon(257,65)(327,65){4}{5}
871: \Photon(257,80)(327,80){4}{5}
872: \Vertex(257,65){1.5}
873: \Vertex(327,65){1.5}
874: \Vertex(257,50){1.5}
875: \Vertex(327,50){1.5}
876: \Vertex(257,80){1.5}
877: \Vertex(327,80){1.5}
878: %
879: \SetOffset(0,-66)
880: \Text(75,39)[c]{\scriptsize{$+$}}
881: \Text(87,65)[c]{\scriptsize{$\vec{k}_F \uparrow$}}
882: \Text(87,15)[c]{\scriptsize{$\vec{k}_F \uparrow$}}
883: \Text(129,65)[c]{\scriptsize{$\vec{k}'_F \downarrow$}}
884: \Text(129,15)[c]{\scriptsize{$\vec{k}'_F \downarrow$}} 
885: \Text(154.2,65)[c]{\scriptsize{$\vec{k}_F \uparrow$}}
886: \Text(154.2,15)[c]{\scriptsize{$\vec{k}_F \uparrow$}}
887: \Text(196.2,65)[c]{\scriptsize{$\vec{k}'_F \downarrow$}}
888: \Text(196.2,15)[c]{\scriptsize{$\vec{k}'_F \downarrow$}} 
889: \Text(141,39)[c]{\scriptsize{$+$}}
890: \Text(201,39)[l]{\scriptsize{$+\cdots$}}
891: %
892: \end{picture}
893: \end{center}
894: \caption{\footnotesize{Graphs contributing to the Landau interaction
895:  functions generated by the ladder summations of the first 
896:  and second order graphs.}}
897: \label{graph2-Ladder} 
898: \end{figure}
899: 
900: Thus the Landau interaction functions can be expressed as 
901: %
902: {\setlength \arraycolsep{2pt} 
903:   \begin{eqnarray}  \label{f_ladd_sum_uu}
904:     f_{\uparrow,\uparrow}
905:     &=&\frac{U^2 \chi_{PH}(\vec{k}_F-\vec{k}'_F)}{1-U^2 \chi_{PH}^2(\vec{k}_F-\vec{k}'_F)}, \\
906:     \nonumber \\
907:     f_{\uparrow,\downarrow}
908:     &=&\frac{U}{1-U\chi_{PP}(\vec{k}_F+\vec{k}'_F)}+\frac{U^2 \chi_{PH}
909:       (\vec{k}_F-\vec{k}'_F)}{1-U\chi_{PH}
910:       (\vec{k}_F-\vec{k}'_F)}.\nonumber \\
911:   \label{f_ladd_sum_ud}
912:   \end{eqnarray}
913: % 
914: With the Landau interaction function obtained above Eqns. 
915: (\ref{dkfdmu}) and (\ref{lineq}) are solved again. Contrary to
916: the second order approximation the local values of the spin 
917: susceptibility diverge to $\infty$ for $U>U_c$ with a 
918: critical $U_c(\mu)$ indicating a
919: ferromagnetic instability of the system near the van Hove 
920: filling. This increased tendency towards ferromagnetism 
921: can be explained by the suppression of singlet pairing   
922: by the ladder summation. In fact the same unphysical divergence 
923: of the second order particle-particle diagram was the main 
924: reason for the suppression of the spin susceptibility in 
925: the second order results. However, a more sophisticated ladder 
926: summation using a renormalized pair scattering\cite{scalapino} or an 
927: renormalization group approach \cite{honerkamp,halboth} will tend to suppress the spin susceptibility again. This is because singlet pairing tendencies 
928: will be generated in higher angular momentum channels, especially in 
929: the $d_{x^2-y^2}$-channel.  
930: 
931: The behavior of the local charge susceptibility when the ladder summation is
932: included is a consistent 
933: extension of the second order calculation with a stronger 
934: suppression of $\chi_c(\vec{k}_F)$ 
935: near the saddle points. The graphs on the left in Fig. \ref{loc_comp_lad} 
936: show $\chi_c(\vec{k}_F)$ with $\vec{k}_F$ close to a
937: saddle point for four different values of $U$. At $U=1.62t$ 
938: ( thin dashed line), the local charge susceptibility goes to zero. 
939: This does not occur if the umklapp processes are excluded from the perturbation expansion, as shown
940: in the right panel of Fig. \ref{loc_comp_lad}. The comparison
941: between these two last graphs confirms the fundamental role 
942: of umklapp processes for the suppression of the charge susceptibility 
943: near the saddle points. In that sense the ladder summation results 
944: for the Landau function lead to a similar behavior of the charge 
945: compressibility as the one observed in one-loop renormalization group treatments.         
946: %
947: \begin{figure}[h] 
948:   \begin{center}
949:     \includegraphics[width=.5 \textwidth]{loc_comp_lad04.eps} %LAD14_comp_unklapp_197
950:     \caption{\footnotesize{Left: $\vec{k}$-space local 
951:         values of $\chi_c(\vec{k}_F)$ for $\vec{k}_F$ near a saddle 
952:         point for different values of  $U$: $U/t=0.5$ thick line, 
953:         $U/t=1.2$  thick dashed line, $U/t=1.91$ thin line, 
954:         $U/t=2.05$ thin dashed line. Right: results 
955:         obtained excluding umklapp processes 
956: from the perturbation expansion. For all data 
957:         $t'=0.3$, $\mu=-3.99t'$ and $T/t=0.04$. }}
958:     \label{loc_comp_lad}
959: \end{center}
960: \end{figure}
961: %
962: 
963: Another similarity with the renormalization group calculations are the tendencies towards
964: deformations of the Fermi surface, also known as Pomeranchuk instabilities and
965: originally suggested for the $t$-$t'$ Hubbard model by Halboth and
966: Metzner\cite{halboth}. Such an instability spontaneously breaks the fourfold
967: symmetry of the electronic dispersion. Close to the van Hove filling the 
968: main energy gain comes from lowering the kinetic energy by pushing one saddle point deeper below the Fermi level.
969: The Landau energy functional for this process can be written, as 
970: quadratic form in the Fermi surface shifts $\delta s(\vec{k}_F)$,
971: % 
972: \begin{equation} \delta E =\frac{1}{(2\pi)^4} \int dl_F L^s(\delta s) \delta s(\vec{k}_F) \, \label{deformationenergy} \end{equation} 
973: %
974: where $L^s$ is defined by (\ref{lineq}).
975: Negative eigenvalues of the linear integral operator
976: $L^s$ imply a lowering of the energy (\ref{deformationenergy}) by a suitable deformation of the Fermi surface.
977: For the specific case $t'/t=0.3$, $\mu/t'=-3.99$ and $T/t=0.04$ 
978: a negative eigenvalue appears
979: already at $U/t=1.91$, slightly before 
980: $\chi_c(\vec{k}_F)$ is suppressed down to zero close to the saddle points at $U/t=2.05$. 
981: The corresponding eigenvector is
982: shown in the graph to the left of Fig. \ref{pomeranchuk} and signals a
983: deformation $\delta s(\theta)$ of the Fermi surface breaking the point
984: group symmetry of the square lattice, as shown schematically in the plot on the right of the same figure. No attempt was made to estimate the actual magnitude of the Fermi surface deformation.
985: 
986: %
987: \begin{figure}[h] 
988:   \begin{center}
989:     \includegraphics[width=.5 \textwidth]{pomeranchuk04.eps} %LAD17eigenvalp.m
990:     \caption{\footnotesize{Left: unstable eigenvector $\delta
991:         s(\theta)$ of the operator $T_{\Xi_c^{-1}}$ with 
992:         eigenvalue $\lambda=-0.01t < 0$ obtained for $t'=0.3t$, $\mu=-3.99t'$, and $U/t=1.91$. Right: 
993:         Pomeranchuk deformation of the Fermi 
994:         surface $\propto \vec{v} (\theta) \delta s(\theta)$.}}
995:     \label{pomeranchuk}
996: \end{center}
997: \end{figure}
998: 
999: In the case $t'= 0.3t$, the above scenario is observed for a very large range
1000: of chemical potential values, i.e. between $\mu / t' = -5.66$ and $\mu /t' =-3.33$. Note however that this instability is only one example out of many potential instabilities of the 2D Hubbard model close to half filling. A direct comparison of several types of instabilities can be found in Ref. \cite{honerkamp2}.
1001: 
1002: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1003: \section{Current carried by the quasiparticle and the Drude weight}
1004: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1005: 
1006: According to Landau-Fermi liquid theory the current $\vec{J}_{\vec{k}_F}$ carried by the quasiparticle $\vec{k}_F$ is given by\cite{theory-of-FL1}
1007:   
1008: \begin{equation}  
1009:   \label{current}
1010:   \vec{J}_{\vec{k}_F}=\vec{v}_{\vec{k}_F}+ \int \frac{dl_F'}{4\pi^2 v_F'} \;
1011:   f^s(\vec{k}_F,\vec{k}'_F)\, \vec{v}_{\vec{k}'_F}.
1012: \end{equation} 
1013: %
1014: The first term of (\ref{current}) describes the current of the bare 
1015: quasiparticle and the second one the ``backflow'' of the 
1016: medium around it. 
1017: The Drude weight is connected to the quasiparticle current via \cite{theory-of-FL1} 
1018: 
1019: \begin{equation}  
1020:   \label{Drude}
1021:   D^{\alpha \beta}=\pi e^2 \int \frac{dl_F}{4\pi^2 v_F} \;
1022:   v^{\alpha}_{\vec{k}_F} J^{\beta}_{\vec{k}_F}.
1023: \end{equation}
1024: % 
1025: With a simple symmetry argument it is easy to show that the off-diagonal
1026: components vanish. This remains true even when a
1027: Pomeranchuk instability breaks the point group symmetry of the 
1028: square lattice, inducing a tetragonal symmetry. The Pomeranchuk instability
1029: merely modifies the diagonal elements 
1030: $D^{aa}$ and $D^{bb}$, which would take different values. 
1031: In this section we will assume that no 
1032: Pomeranchuk instability affects the system, so that 
1033: $D^{aa}$=$D^{bb}=D$.
1034: 
1035: Note that in a Galilean invariant system
1036: the total current is a constant of motion and does not 
1037: change when the interaction is switched on adiabatically. But as a lattice model the Hubbard model does not belong to this class of systems, 
1038: thus the magnitude of $D$ is related to the strength of the 
1039: interaction $U$. Fig. \ref{currtot} shows the change of the Drude weight
1040: when the Landau interaction function that is used to solve
1041: (\ref{current}) and (\ref{Drude}) is approximated by the ladder and bubble 
1042: summations (\ref{f_ladd_sum_uu}) and (\ref{f_ladd_sum_ud}). The thick lines 
1043: show $D(U)/(e^2 \pi)$ for three different values of the chemical potential $\mu$. With increasing interaction $U$, 
1044: the Drude weight is reduced and goes to zero for a finite
1045: $U$. The reduction occurs for smaller interactions the closer the electron density is to the van Hove filling. 
1046: %
1047: \begin{figure}[h] 
1048:   \begin{center}
1049:     \includegraphics[width=.4 \textwidth]{currtotT04.eps} %ccourtotT04.m
1050:     \caption{\footnotesize{The thick lines show $D(U)$ obtained 
1051:         by the ladder summation for three different values of 
1052:         $\mu$: $\mu/t'=-3.33$ dashed
1053:         line, $\mu/t'=-3.99$ dash-dotted line, 
1054:         $\mu/t'=-5.66$ continuous line and T=0.04t. 
1055:         The thin lines show the corresponding results obtained excluding 
1056:         the umklapp process. 
1057: }}
1058:     \label{currtot}
1059: \end{center}
1060: \end{figure}
1061: %
1062: We note that in our calculation the Drude weight vanishes only at interaction strengths that are larger than the critical value necessary to induce 
1063: the Pomeranchuk instability ($U_p$: $U/t=2.075$ for $\mu=-3.33t$, $U/t=1.91$ for $\mu=-3.99t$ and $U/t=2.6$ for
1064:  $\mu=-5.66t$). Thus the above results have no clear physical meaning for
1065: $U>U_p$. The behavior of the Drude weight for $U < U_p$, however, provides
1066: some useful information: as shown in Fig. \ref{currtot}, its decrease is
1067: strongest at the van Hove filling, $\mu = -4t'$. 
1068: In addition, the results confirm again the important role played by the
1069: umklapp processes. In fact, if they are excluded from the calculation (thin
1070: lines of the Fig. \ref{currtot}), the magnitude of the Drude weight even
1071: increases with the strength of $U$.
1072: 
1073: It is interesting to determine the scattering processes
1074: responsible for the decrease of $D$. The continuous lines plotted in
1075: Fig. \ref{LADlandau100} show the symmetric Landau interaction function
1076: $f(\theta,\theta')$ for  $\mu/t'=-3.33$ that has been used to solve (\ref{current}) for
1077: different values of $\theta$ and as function of $\theta'$. The thick vertical 
1078: lines mark the different values of $\theta$ and the thin lines are plotted to emphasize the relevant peaks of $f(\theta,\theta')$ for the given $\theta$. 
1079: The insets shows the locations of $\vec{k}_F(\theta)$ (thick line) 
1080: and $\vec{k}'_F(\theta')$ (thin lines) where the peaks in $f(\theta,\theta')$ occur.         
1081: 
1082: 
1083: \begin{figure}[h] 
1084:   \begin{center}
1085:     \includegraphics[width=.5 \textwidth]{LADlandausym100.eps} %ccourtotT04.m
1086:     \caption{\footnotesize{The continuous lines show $f(\theta,\theta')$ for 
1087:         $\mu/t'=-3.33$, $U=2.075t$ and $T=0.04t$,
1088:         obtained for three different values of $\theta$, as a function of
1089:         $\theta'$. The thick vertical 
1090:         lines mark the different values of $\theta$ and the thin vertical
1091:         ones are plotted to emphasize the relevant peaks of 
1092:         $f(\theta,\theta')$. The insets show the locations of the Fermi surface points marked by the thick and thin lines. 
1093:         The dashed lines show the corresponding results of 
1094:         $f(\theta,\theta')$ obtained excluding 
1095:         the umklapp process.
1096: }}
1097:     \label{LADlandau100}
1098: \end{center}
1099: \end{figure}  
1100: 
1101: Fig.\ref{LADlandau100} shows that the peaks in the Landau
1102: function for fillings larger than the van Hove filling are tied to the Fermi surface 
1103: points $\vec{k}_F$, $\vec{k}'_F$ which can be connected by wavevectors close 
1104: to  $(\pi, \pi)$, i.e.  
1105: $\vec{q}_0=\vec{k}_F-\vec{k'}_F \simeq (\pm \pi, \pm \pi)$. These processes
1106:  are enhanced in the ladder and bubble summation of the terms involving
1107: $\chi_{PH}(\vec{q}_0)$ (\ref{chiPH}). For fillings smaller than the 
1108: van Hove filling and $-6.6 <\mu/t'<-4$ the Fermi surface is 
1109: partially nested and the peaks appear when $\vec{q}_0=\vec{k}_F-\vec{k'}_F$
1110: is equal to the corresponding nesting vectors 
1111: (that are smaller than $(\pi,\pi)$). For $\mu/t'< -6.6$ the 
1112: Fermi surface is strictly convex and the 
1113: Drude weight increases with the strength of $U$. 
1114: 
1115: Without umklapp processes the relevant peaks of $f^s$ disappear, the corresponding Landau interaction (thin dash\-ed lines of
1116: the Fig. \ref{LADlandau100}) causes a ''backflow'' of the medium in the
1117: direction of the bare quasiparticle current implying the increase of
1118: $\vec{J}_{\vec{k}_F}$ for anyvalue of $\vec{k}_F$. Thus the Drude weight will increase.
1119: 
1120: \section{Conclusions}
1121: 
1122: We have presented a detailed study of the Landau interaction function $f(\vec{k},\vec{k}')$ and wave-vector resolved compressibilities within Fermi liquid theory for the two-dimensional Hubbard model. This model is often considered as minimal model for the description of high-$T_c$ superconductivity in the copper-oxide materials. 
1123: 
1124: A less surprising conclusion of this work is that a  Fermi liquid picture
1125: alone can hardly account for the observed anomalies that occur in the underdoped high-$T_c$  materials when the doping is reduced towards half-filling: all 
1126: effects are rather weak and one has to resort to stronger interactions $U$ or
1127: move closer to a magnetic instability (e.g. by ladder summations) to obtain sizable effects such as a significant reduction of the compressibility on parts of the Fermi surface. No rapid decrease of the spin susceptibility reminiscent of spin gap formation\cite{fuseya} can be found. Similar conclusions regarding the high-$T_c$ cuprates were reached by Kim and Coffey\cite{kim} who analyzed the spectral function of the 2D Hubbard model within random phase approximation.
1128:  
1129: On the other hand, our weak coupling analysis provides  
1130: insights into the physical processes which lead to deviations 
1131: from the conventional picture. We find that many of the observed 
1132: tendencies towards partially insulating behavior - as we believe 
1133: an essential ingredient of the underdoped cuprate superconductors - are 
1134: related to elastic umklapp processes. These processes start to affect 
1135: the low energy physics upon reducing the doping when the Fermi surface 
1136: touches the Brillouin zone boundaries. As we know from earlier 
1137: renormalization group studies\cite{furukawa,honerkamp} 
1138: of the same model, strong umklapp processes with momentum transfer $(\pi,
1139: \pi)$ between the saddle point 
1140: regions lead to an interesting mutual reinforcement of antiferromagnetic 
1141: and $d$-wave pairing correlations. Furthermore the renormalization group 
1142: treatment yielded 
1143: a regime with a strong suppression of the charge compressibility near 
1144: the saddle points. A similar effect was observed by Otsuka et 
1145: al.\cite{otsuka} in quantum Monte Carlo calculations for $U=4t$. 
1146: In our Fermi liquid analysis we clearly identify the elastic umklapp 
1147: processes as being responsible for a strong reduction of the Drude weight 
1148: and the angle resolved charge compressibility on parts of the Fermi surface. 
1149: In one-dimensional systems it is well known that elastic umklapp processes 
1150: imply these effects. Our calculations shows that in the 2D Hubbard model 
1151: there are observable tendencies in the same direction. Although our 
1152: perturbative analysis does not allow to conclude
1153:  that umklapp processes are the ultimate cause for the 
1154: pseudogap features at stronger interactions, they will at least contribute 
1155: to a certain extent. Furthermore there 
1156: is no obvious reason why the importance of the umklapp scattering should 
1157: decrease again for stronger interactions.
1158: 
1159: Our perturbative calculations show tendencies which can be interpreted 
1160: as precursors to stronger reductions in the compressibility and Drude 
1161: weight at stronger interactions, features which appear in some
1162: phenomenological descriptions of the socalled pseudogap region of the 
1163: phase diagram of the underdoped cuprates. In this way they support the
1164: proposal that the critical doping of the cuprates that separates the 
1165: underdoped and overdoped regimes \cite{tallon}, is determined by the
1166: appearance of elastic umklapp scattering with momentum transfer 
1167: $(\pi,\pi)$ connecting the Fermi surface regions near to the saddle
1168: points.
1169: 
1170: We thank M. Sigrist and M. Troyer for helpful discussions. We acknowledge
1171: financial support by Swiss National Foundation. C.H. also acknowledges
1172: financial support by the German Science Foundation (DFG). Computations were 
1173: carried out on the Beowulfcluster Asgard at ETHZ.  
1174: 
1175: \begin{thebibliography}{99}
1176: \bibitem{anderson} P.W. Anderson, {\em The Theory of Superconductivity in the
1177: High-$T_c$ Cuprates}, Princeton University Press (1997).
1178: \bibitem{timusk} T. Timusk and B. Statt, Rep.Prog.Phys. {\bf
1179: 62}, 61 (1999).
1180: \bibitem{proust} C. Proust, E. Boaknin, R.W. Hill, L. Taillefer, and A.P. Mackenzie, cond-mat/0202101.
1181: \bibitem{tallon} J.L. Tallon and J.W. Loram, Physica C {\bf 349}, 53 (2001).  
1182: \bibitem{sachdev} M. Vojta, Y. Zhang, and S. Sachdev, Phys. Rev. B{\bf 62},
1183: 6721 (2000).
1184:  \bibitem{chakravarty} S. Chakravarty, R.B. Laughlin, D.K. Morr, and C. Nayak,
1185: Phys. Rev. B {\bf 63}, 094503; S. Chakravarty, Hae-Young Kee, and C. Nayak, Int. J. Mod. Phys. B {\bf 15}, 2901 (2001).
1186: \bibitem{norman} M.R.Norman et al., Nature {\bf 392}, 157 (1998).
1187: \bibitem{furukawa} N. Furukawa, T.M. Rice, and M. Salmhofer, Phys.Rev.Lett. {\bf
1188: 81}, 3195 (1998).
1189: \bibitem{honerkamp}
1190: C.Honerkamp, M.Salmhofer, N.Furukawa, and T.M. Rice,
1191: Phys. Rev. B {\bf 63} 35109 (2001); C. Honerkamp, Euro. Phys. J. B {\bf 21},
1192: 81 (2001).
1193: \bibitem{millis} A.J. Millis, S.M.  Girvin, L.B.  Ioffe, A.I. Larkin, J. Phys. Chem. Sol. {\bf
1194: 59}, 1742 (1998).
1195: \bibitem{fuseya} Y. Fuseya, H. Maebashi, S. Yotsuhashi and Kazumasa
1196:   Miyake, J.  Phys. Soc. of Japan Vol. {\bf 69} 2158 (2000).
1197: \bibitem{halboth} C.J. Halboth and W. Metzner, Phys. Rev. B {\bf 61}, 7364 (2000); Phys. Rev. Lett. {\bf 85}, 5162 (2000).
1198: \bibitem{theory-of-FL1} Philippe Nozieres: Theory of Interacting Fermi Systems, Advanced book classics ('97).
1199: \bibitem{scalapino} D.J. Scalapino, E. Loh Jr., and J.E. Hirsch, Phys. Rev. 
1200: B{\bf 35} 6694 (1987).
1201: \bibitem{honerkamp2} C. Honerkamp, M. Salmhofer, and T.M. Rice, submitted to EPJB. 
1202: %\bibitem{ioffe} L.B. Ioffe, A.J. Millis, Phys. Rev. B {\bf 58}, 11631
1203:  %(1998). 
1204: \bibitem{kim} J. Kim and D. Coffey, Phys. Rev. B{\bf 62}, 4288 (2000).
1205: \bibitem{otsuka}  Y. Otsuka, Y. Morita, Y. Hatsugai, cond-mat/0106420.
1206: %\bibitem{abrikosov} A. A. Abrikosov and I. M. Khalatnikov: Sov. Phys. JETP 6 ('58) 888.
1207: \end{thebibliography}
1208: \end{document}