cond-mat0408246/BS6.tex
1: \documentstyle[aps,floats,psfig,epsf,twocolumn]{revtex}
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: \draft
4: 
5: \begin{document}
6: \title{Order parameter symmetries for magnetic and superconducting
7: instabilities: Bethe-Salpeter analysis of functional renormalization-group
8: solutions}
9: \author{A. A. Katanin$^{a,b}$ and A. P. Kampf$^c$}
10: \address{$^a$Max-Planck-Institut f\"ur Festk\"orperforschung, D-70569
11: Stuttgart, Germany\\
12: $^b$Institute of Metal Physics, 620219 Ekaterinburg, Russia\\
13: $^c$Theoretische Physik III, Elektronische Korrelationen und Magnetismus,\\
14: Institut f\"ur Physik, Universit\"at Augsburg, D-86135 Augsburg, Germany}
15: \address{~\\
16: \parbox{14cm}{\rm \medskip \vskip0.2cm The Bethe-Salpeter equation
17: is combined with the temperature-cutoff functional renormalization
18: group approach to analyze the order parameter structure for the
19: leading instabilities of the 2D $t$-$t^{\prime}$ Hubbard model. We find
20: significant deviations from the conventional s-, p-, or d-wave forms, which is
21: due to the frustration of antiferromagnetism at small and intermediate
22: $t^{\prime}$. With adding a direct antiferromagnetic spin-exchange
23: coupling the eigenfunctions in the particle-hole channel have extended
24: s-wave form, while in the particle-particle singlet pairing channel a higher
25: angular momentum component arises besides the standard d$_{x^2-y^2}$-wave
26: component, which flattens the angular dependence of the gap. For $t^{\prime}$
27: closer to $t/2$ we find a delicate competition of ferromagnetism and triplet
28: pairing with a nontrivial pair-wavefunction.
29: \vskip0.05cm\medskip PACS Numbers:
30: 71.10.Fd; 71.27.+a;74.25.Dw }}
31: \maketitle
32: 
33: \tighten
34: 
35: It is by now well established that the superconducting order parameter in
36: high-T$_c$ compounds is well described by a ($\cos k_x-\cos k_y$)-momentum
37: dependence: it is largest at the Fermi surface (FS) points close to $(\pi,0)$
38: and $(0,\pi)$ and vanishes at the FS crossings on the Brillouin zone (BZ)
39: diagonals. Accurate measurements of the gap function, however, revealed a
40: slight deviation from this d$_{x^2-y^2}$-wave momentum dependence \cite{Mesot},
41: with a flatter angular dependence near the nodal
42: points. The symmetry and the details of the momentum dependence of the
43: superconducting order parameter are closely related to the structure of the
44: effective attractive interaction between the electrons. The precise momentum
45: dependence of the energy gap therefore contains valuable information about
46: the underlying pairing mechanism, for which antiferromagnetic (AFM) spin
47: fluctuations are a viable candidate for cuprates \cite
48: {Scalapino,BSW,Pines,Chubukov}.
49: 
50: Another type of unconventional superconductor is the layered ruthenate
51: Sr$_2$RuO$_4$ \cite{PTReview}, which most likely has triplet pairing with
52: p-wave symmetry \cite{SrRuGap}. It was proposed that the pairing in this
53: material results from ferromagnetic (FM) spin fluctuations
54: \cite{Mazin,Murakami}. Although inelastic neutron scattering has so far been
55: unsuccessful to detect significant low-energy FM spin fluctuations
56: \cite{Maeno}, this idea finds support from the enhanced tendencies towards
57: ferromagnetism in the electron doped compound Sr$_{2-x}$La$_x$RuO$_4$
58: \cite{ED} and in isoelectronic Ca$_2$RuO$_4$ under hydrostatic pressure
59: \cite{Maeno1}.
60: 
61: The important role of AFM and FM fluctuations as a possible driving source
62: of singlet and triplet superconductivity, respectively, was emphasized
63: early on in the pioneering work in Refs. \cite{KL,BS,SLH}. Recently, the
64: interplay of antiferromagnetism and d-wave superconductivity (dSC) and
65: ferromagnetism and p-wave superconductivity (pSC), respectively, was
66: reconsidered within the $t$-$t^{\prime }$ Hubbard model using functional
67: renormalization-group (fRG) techniques
68: \cite{Zanchi,Metzner,SalmHon,SalmHon1,KK}. Early versions of fRG
69: \cite{Zanchi,Metzner,SalmHon}, which used the momentum cutoff procedure, were
70: unable to search for ferromagnetism. This drawback is overcome in the
71: temperature-cutoff fRG approach (TCRG) \cite{SalmHon1}, which proved
72: successful in describing both, AFM and FM instabilities together with
73: singlet- and triplet superconducting pairing in the weak-coupling
74: $t$-$t^{\prime}$ Hubbard model and its extensions in a broad parameter range
75: \cite{SalmHon1,KK,KK1}. In the previous fRG analyses it has been a common
76: practice to assume order parameter structures, which have the form of square
77: lattice basis functions with s-, p- or d-wave symmetry
78: \cite{Zanchi,Metzner,SalmHon,SalmHon1,KK,KK1}. Although it was recognized that
79: in the RG flow of the order parameter susceptibilities also the momentum
80: dependence acquires specific corrections to their initial form \cite
81: {Metzner,SalmHon1,KK,KK1,Neumayr}, these corrections were so far not
82: analyzed in detail. Indeed, actual order parameter structures may have
83: admixtures of different symmetry components, and it is necessary to develop
84: an unbiased method to determine their precise momentum dependence.
85: 
86: In the present paper we use the Bethe-Salpeter equations to extract
87: eigenfunctions and eigenvalues of the effective interaction in the
88: particle-particle (pp) or the particle-hole (ph) channel -- similarly as in
89: previous quantum Monte Carlo studies \cite{BetheSal}. Here we consider a
90: combination of the Bethe-Salpeter equations and the fRG approach. We choose
91: the TCRG version \cite{SalmHon1} as the most suitable tool, because the
92: effective model obtained within this scheme does not contain any
93: unintegrated degrees of freedom even at the intermediate stages of the RG
94: flow.
95: 
96: We apply this procedure to identify the order parameter structure for the
97: leading instabilities of the 2D $t$-$t^{\prime }$ extended Hubbard model
98: $H_\mu=H-(\mu-4t^{\prime})N$ with
99: \begin{equation}
100: H=-\sum_{ij\sigma }t_{ij}c_{i\sigma }^{\dagger }c_{j\sigma
101: }+U\sum_in_{i\uparrow }n_{i\downarrow }+J\sum_{\langle ij\rangle }{\bf S}
102: _i\cdot {\bf S}_j\,,  \label{H}
103: \end{equation}
104: where $t_{ij}=t$ for nearest neighbor (nn) sites $i$ and $j$ and
105: $t_{ij}=-t^{\prime }$ for next-nn sites ($t,t^{\prime }>0$) on a square
106: lattice; for convenience we have shifted the chemical potential $\mu$ by
107: $4t^{\prime}$. In Eq. (\ref{H}) we have included a direct nn spin exchange
108: interaction $J$; ${\bf S}_i=c_{i\alpha}^{\dagger}\mbox{\boldmath$\sigma$}_{
109: \alpha \beta }c_{i\beta }/2,$ and $\mbox{\boldmath$\sigma$}$ denotes the
110: Pauli matrices. While such an interaction is generated from the on-site
111: Coulomb repulsion at strong-coupling, we add it here as an independent
112: interaction in the weak-coupling regime, where the RG scheme is applicable.
113: 
114: We follow the many-patch fRG version for one-particle irreducible Green
115: functions as proposed in Ref. \cite{SalmHon1}. This TCRG scheme uses the
116: temperature as a natural cutoff parameter and accounts for excitations with
117: momenta far from and close to the FS, which is necessary for the description
118: of instabilities arising from zero-momentum ph scattering, e.g.
119: ferromagnetism. Neglecting the frequency dependence of the vertices, which is
120: expected to have minor relevance in the weak-coupling regime, the RG
121: differential equation for the interaction vertex has the form
122: \cite{SalmHon1}
123: \begin{eqnarray}
124: \frac{{\rm d}V_T}{{\rm d}T}=-V_T\circ \frac{{\rm d}L_{{\rm pp}}}{{\rm d}T}%
125: \circ V_T+V_T\circ \frac{{\rm d}L_{{\rm ph}}}{{\rm d}T}\circ V_T\,,
126: \label{dV}
127: \end{eqnarray}
128: where $\circ $ is a short notation for summations over intermediate momenta
129: and spin,
130: \begin{equation}
131: L_{\text{ph,pp}}({\bf k},{\bf k}^{\prime })=\frac{f_T(\varepsilon _{{\bf k}})
132: -f_T(\pm \varepsilon _{{\bf k}^{\prime }})}{\varepsilon _{{\bf k}}\mp
133: \varepsilon _{{\bf k}^{\prime }}},
134: \label{Lphpp}
135: \end{equation}
136: and $f_T(\varepsilon )$ is the Fermi function. The upper sign in Eq.
137: (\ref{Lphpp}) is for $L_{ph}$ and the lower sign for $L_{pp}$, respectively.
138: Eq. (\ref{dV}) has to be solved with the initial condition
139: $V_{T_0}({\bf k}_1,{\bf k}_2,{\bf k}_3,{\bf k}_4)=U$; the initial temperature
140: is chosen as large as $T_0=400t$.
141: 
142: We discretize the momentum space in $N_p=48$ patches using the same patching
143: scheme as in Ref. \cite{SalmHon1}. This reduces the integro-differential
144: equations (\ref{dV}) and (\ref{dH}) to a set of 5824 differential equations,
145: which were solved numerically. The evolution of the vertices with decreasing
146: temperature determines the temperature dependence of the susceptibilities
147: according to \cite{SalmHon1,SalmHon02}
148: \begin{eqnarray}
149: \displaystyle{\frac{{\rm d}\chi _{m,r}}{{\rm d}T}} &=&\sum_{{\bf k}^{\prime
150: }}{\cal R}_{{\bf k}^{\prime }}^{m,r}{\cal R}_{\mp {\bf k}^{\prime }+{\bf q}%
151: _m}^{m,r}\displaystyle{\frac{{\rm d}L_{\text{pp},\text{ph}}({\bf k}^{\prime
152: },\mp {\bf k}^{\prime }+{\bf q}_m)}{{\rm d}T}},  \label{dH} \\
153: \displaystyle{\frac{{\rm d}{\cal R}_{{\bf k}}^{m,r}}{{\rm d}T}} &=&\mp \sum_{%
154: {\bf k}^{\prime }}{\cal R}_{{\bf k}^{\prime }}^{m,r}\Gamma _m^T({\bf k},{\bf %
155: k}^{\prime })\displaystyle{\frac{{\rm d}L_{\text{pp},\text{ph}}({\bf k}%
156: ^{\prime },\mp {\bf k}^{\prime }+{\bf q}_m)}{{\rm d}T}},  \nonumber
157: \end{eqnarray}
158: where ($\Gamma _m^T\equiv \Gamma _m^T({\bf k},{\bf k}^{\prime })$)
159: \begin{equation}
160: \Gamma _m^T=\left\{
161: \begin{array}{cl}
162: \begin{array}{c}
163: V_T({\bf k},{\bf k}^{\prime },{\bf k}^{\prime }+{\bf q}_m) \\
164: -2V_T({\bf k},{\bf k}^{\prime },{\bf k}+{\bf q}_m)
165: \end{array}
166: & \text{phc (PI, DDW)}, \\
167: V_T({\bf k},{\bf k}^{\prime },{\bf k}^{\prime }+{\bf q}_m) & \text{phs (AFM,
168: FM),} \\
169: V_T({\bf k},-{\bf k+q}_m,{\bf k}^{\prime }) & \text{pp (dSC, pSC).}
170: \end{array}
171: \right.   \label{Gamma}
172: \end{equation}
173: $m$ denotes one of the possible channels: ph spin (phs), which traces FM and
174: AFM instabilities, ph charge (phc) for the Pomeranchuk instability (PI) \cite
175: {Pomeranchuk} or d-density wave (DDW) \cite{DDW}, or pp for superconducting
176: singlet and triplet instabilities; ${\bf q}_m={\bf Q=}(\pi ,\pi )$ for AFM
177: and DDW and ${\bf q}_m={\bf 0}$ otherwise. In the following we use the
178: notation $m{\bf q}_m$ to denote the specific channel. The index $r$ in Eq.
179: (\ref{dH}) represents the symmetry of the corresponding channel. The upper
180: signs and pp-indices in Eq. (\ref{dH}) refer to the superconducting
181: instabilities (dSC and pSC), and the lower signs and ph-indices refer to the
182: charge and magnetic instabilities.
183: 
184: The three-point vertices ${\cal R}_{{\bf k}}^{m,r}$ can be considered
185: as vertices describing the propagation of the electron in a static external
186: field; their diagrammatic representation
187: is shown in Fig. 1. The initial conditions at $T_0$ for Eqs. (\ref{dH}) are
188: ${\cal R}_{{\bf k}}^{m,r}=f_{{\bf k}}^{(r)}$ and $\chi_{m,r}=0$, where the
189: function $f_{{\bf k}}^{(r)}$ belongs to one of the irreducible representations
190: of the point group of the square lattice ($A_{1g}$ or $B_{1,2g}$), e.g.
191: 
192: \begin{figure}[t!]
193: \psfig{file=Fig1.eps,width=90mm,silent=} \vspace{2mm}
194: \caption{Diagrammatic representation of the three-point vertex
195: ${\cal R}_{{\bf k}}^{m,r}$ (solid triangle), which has the structure of a
196: vertex diagram in a static external field acting in the channel m with the
197: r-th initial vertex function $f_{{\bf k}}^{(r)}$ defined in Eq. (\ref{ff}) and
198: shown as an open triangle. The specific example shows the phs{\bf Q} channel.
199: The circle represents the reducible electron-electron interaction vertex.}
200: \label{fig:Fig1}
201: \end{figure}
202: 
203: \begin{equation}
204: f_{{\bf k}}^{(r)}=A^{-1}\left\{
205: \begin{array}{cll}
206: \cos k_x-\cos k_y & \text{d$_{x^2-y^2}$-wave } & (B_{1g}), \\
207: \sin k_{x(y)} & \text{p-wave} & (B_{2g}), \\
208: 1 & \text{s-wave} & (A_{1g}) \label{ff}
209: \end{array}
210: \right.
211: \end{equation}
212: with a normalization coefficient $A=(1/N)\sum_{{\bf k}}f_{{\bf k}}^2$. The
213: momentum dependence of the vertices ${\cal R}_{{\bf k}}^{m,r}$ changes during
214: the RG flow and at low temperatures it is expected to reflect the momentum
215: dependence of the ground-state order parameter of the selected symmetry.
216: However, these vertex functions depend on the initial choice of the functions
217: $f_{{\bf k}}^{(r)}$, and therefore they can not serve as an unbiased tool to
218: obtain the structure of the ground-state order parameters.
219: 
220: To perform an alternative analysis, which is not based on the knowledge of
221: the wave functions at high temperatures, we consider the solution of the
222: Bethe-Salpeter equations \cite{BetheSal}
223: \begin{eqnarray}
224: \sum_{{\bf p}}\overline{\Gamma }_{ph}({\bf k},{\bf p};{\bf p},{\bf k})L_{ph}(
225: {\bf p},{\bf p}+{\bf q}_m)\phi _{{\bf p}}^{ph} &=&\lambda _{ph}\phi _{{\bf k}
226: }^{ph}\,,  \nonumber \\
227: -\sum_{{\bf p}}\overline{\Gamma }_{pp}({\bf k},{\bf p};{\bf p},{\bf k}
228: )L_{pp}({\bf p},{\bf p})\phi _{{\bf p}}^{pp} &=&\lambda _{pp}\phi _{{\bf k}
229: }^{pp}\, ,
230: \label{BSirr}
231: \end{eqnarray}
232: where $\overline{\Gamma }_{ph}$ and $\overline{\Gamma }_{pp}$ denote
233: irreducible vertices in ph and pp channels, respectively. Exploiting the
234: connection to reducible vertices, Eqs. (\ref{BSirr}) can be rewritten as
235: \cite{BetheSal}
236: \begin{equation}
237: \sum_{{\bf p}}\Gamma _m^T({\bf k},{\bf p})L_{ph,pp}^T({\bf p},\pm {\bf p}+
238: {\bf q}_m)\phi _{{\bf p}}^{m,l}=\displaystyle{\frac{\lambda _{m,l}\phi_{{\bf
239: k}}^{m,l}}{1-\lambda _{m,l}}}\, ,
240: \label{BSr}
241: \end{equation}
242: where $l$ enumerates eigenvalues and -functions for a given channel $m$. The
243: reducible vertices $\Gamma _m^T({\bf k},{\bf p})$ can be directly extracted
244: from the fRG flow according to Eq. (\ref{Gamma}).
245: 
246: The value $\lambda _{m,l}=1$ corresponds to an ordering instability with the
247: symmetry of the eigenfunction $\phi _{{\bf k}}^{m,l}$. Therefore, tracing
248: the temperature dependence of eigenvalues and -functions allows to identify
249: both, the leading instabilities {\it and} their concomitant order parameter
250: structure. We stop the fRG flow at the temperature $T_X$, where the maximum
251: interaction vertex $V_{\max}\equiv\max\{V({\bf k}_1,{\bf k}_2;{\bf k}_3,
252: {\bf k}_4)\}$ reaches the value $20t$.
253: $T_X$ should be understood as a crossover temperature into a renormalized
254: classical regime with exponentially large correlation length \cite{KK1},
255: we have verified that the following results for the eigenfunctions are
256: only weakly dependent on the choice of $T_X$.
257: 
258: Eigenfunctions and eigenvalues of the Bethe-Salpeter equations provide more
259: detailed information than can be obtained from the momentum dependence of the
260: three-point vertices ${\cal R}_{{\bf k}}^{m,r}$, which determine the
261: order-parameter susceptibilities according to Eq. (\ref{dH}). Unlike the
262: vertices ${\cal R}_{{\bf k}}^{m,r}$, the solutions of the Bethe-Salpeter
263: equations do not depend on the initial choice of the functions $f_{{\bf k}}^{
264: (r)}$ and, therefore, do not require {\it a priori} the knowledge of the
265: symmetry of the leading instability.
266: 
267: \begin{figure}[t!]
268: \psfig{file=Fig2.eps,width=90mm,silent=} \vspace{2mm}
269: \caption{Eigenvalues (a) and angular $\theta_{{\bf k}_F}$-dependence on the
270: FS of the eigenfunctions $\phi _{{\bf p}}^{m,l}=\phi _l$ of the
271: Bethe-Salpeter equation and the normalized three-point vertices ${\cal R}_{
272: {\bf k}}^{m,r}$ (denoted as ${\cal R}$-$m,r$) at $T=T_X$ (b,c,d) for
273: $t^{\prime }=0.1t,$ $U=2t,$ $J=\mu=0$. $T_X$ is the lowest temperature
274: reached in the fRG flow (see (a)).}
275: \label{fig:Fig2}
276: \end{figure}
277: 
278: The three-point vertices
279: \begin{equation}
280: \widetilde{{\cal R}}_{{\bf k}}^{m,r}=f_{{\bf k}}^{(r)}+\sum_{{\bf p}}\Gamma
281: _m^T({\bf k},{\bf p})L_{ph,pp}^T({\bf p},\pm {\bf p}+{\bf q}_m)f_{{\bf p}}^{
282: (r)}
283: \end{equation}
284: described by the diagram of Fig. 1 can be obtained directly from the solutions
285: of the Bethe-Salpeter equations. Expanding the functions $f_{{\bf k}}^{(r)}$
286: in terms of eigenfunctions of the Bethe-Salpeter equation and using Eqs.
287: (\ref{BSr}), we obtain
288: \begin{equation}
289: \widetilde{{\cal R}}_{{\bf k}}^{m,r}=\sum_l\frac{\phi _{{\bf k}}^{m,l}}{%
290: 1-\lambda _{m,l}}\sum_{{\bf k}^{\prime }}f_{{\bf k}^{\prime }}^{(r)}\phi _{%
291: {\bf k}^{\prime }}^{m,l}\, .
292: \label{RR}
293: \end{equation}
294: The vertices $\widetilde{{\cal R}}_{{\bf k}}^{m,r}$ found in this way,
295: however, do not necessarily coincide with those obtained directly from the
296: RG procedure, Eq. (\ref{dH}), because generally the one-loop approximation
297: does not exactly reproduce the solution of the corresponding Bethe-Salpeter
298: equation. The only case when the three-point vertices ${\cal R}_{{\bf k}}^{m,
299: r}$ and $\widetilde{{\cal R}}_{{\bf k}}^{m,r}$ coincide is the ladder
300: approximation when either only the pp-channel or one of the ph-channels in
301: Eq. (\ref{dV}) is retained. Nevertheless, as we will see below, the results
302: for the vertex ${\cal R}_{{\bf k}}^{m,r}$ from Eqs. (\ref{dH}) are
303: qualitatively similar to those found from Eq. (\ref{RR}).
304: 
305: It is clearly established from Eq. (\ref{RR}) that $\widetilde{{\cal R}}_{{\bf
306: k}}^{m,r}$ (and similarly ${\cal R}_{{\bf k}}^{m,r}$) is actually a mixture of
307: different eigenfunctions $\phi_{{\bf k}}^{m,l}$, whose weights are
308: proportional to $1/(1-\lambda _{m,l})$ with coefficients determined by the
309: overlap of the eigenfunction with $f_{{\bf k}^{\prime}}^{(r)}$. If one of the
310: eigenvalues $\lambda _{m,l}$ is much closer to unity than all the others, the
311: corresponding term in the sum over $l$ in Eq. (\ref{RR}) is expected to
312: dominate and $\widetilde{{\cal R}}_{{\bf k}}^{m,r}$ essentially coincides with
313: the corresponding eigenfunction $\phi_{{\bf k}}^{m,l}$.
314: 
315: Below we discuss the results of the numerical solution of the Bethe-Salpeter
316: equations and compare them to the results for the three-point vertices
317: ${\cal R}_{{\bf k}}^{m,r}$ for different parameter regimes. We start in Fig.
318: 2 with results at the van Hove (vH) band filling ($\mu =0$) for small
319: $t^{\prime }=0.1t$, $J=0$, and $U=2t$. The AFM phs{\bf Q} instability has the
320: largest eigenvalue at $T_X$ in agreement with previous fRG work based on the
321: analysis of order parameter susceptibilities \cite{SalmHon1}. The
322: corresponding eigenfunction (Fig. 2c) has a slight variation around the FS
323: with an enhancement near ($\pi ,0$) and ($0,\pi$), which most likely
324: originates from the vH singularity nature of these points. The
325: eigenfunctions in the subleading phc{\bf 0} and pp{\bf 0} channels are
326: sizable near ($\pi ,0$) and ($0,\pi ,$) only. Although the eigenvalue of the
327: zero-momentum ph instability in the charge channel (phc{\bf 0}) is
328: relatively small at $T=T_X$ ($\lambda \simeq 0.5$), it rapidly increases at
329: low temperatures. The corresponding eigenfunction (Fig. 2d) is antisymmetric
330: with respect to 90$^{\circ }$ rotation, thus providing the possibility for
331: the Pomeranchuk instability with a spontaneous d-wave like deformation of
332: the FS \cite{Yamase,Metzner1}. The eigenvalues for the PI at $t^{\prime
333: }=0.1t$ are larger than those at $t^{\prime}=0$ (not shown), which supports
334: the conclusion of Ref. \cite{Metzner1}, that a finite value of $t^{\prime}$
335: at vH band fillings enhances the tendency towards a Pomeranchuk instability.
336: 
337: The momentum dependence of the three-point vertices ${\cal R}_{{\bf k}}^{m,r}$
338: is qualitatively similar to the leading eigenfunctions in the pp{\bf 0} and
339: phs{\bf Q }channels, however with a smaller variation around the FS. In the
340: phc{\bf 0} channel we observe a stronger difference of the vertex
341: ${\cal R}_{{\bf k}}^{phc,d\text{-wave}}$, which has almost the
342: d$_{x^2-y^2}$-wave form, from the corresponding eigenfunction of the
343: Bethe-Salpeter equation.
344: 
345: \begin{figure}[t!]
346: \psfig{file=Fig3.eps,width=90mm,silent=} \vspace{2mm}
347: \caption{Same as in Fig. 1 for $U=1.5t,$ $J=0.3t.$}
348: \label{fig:Fig3}
349: \end{figure}
350: 
351: To investigate the role of $t^{\prime }$-induced frustration of
352: antiferromagnetism, we turn on a small direct exchange interaction $J=0.3t$
353: to strengthen the AFM correlations, and decrease simultaneously the value of
354: $U$ to $1.5t$ in order to remain in the weak-coupling regime. Fig. 3 shows
355: the resulting changes. At finite $J$ the eigenfunctions in the pp{\bf 0} and
356: phc{\bf 0} channels are essentially nonzero all around the FS, however they
357: are flatter than the $d_{x^2-y^2}$-wave function. The eigenfunction in the
358: phs{\bf Q} channel is of extended s-wave form. The eigenvalue for the
359: phc{\bf 0} channel is smaller than at $J=0$, implying that stronger
360: antiferromagnetism weakens the tendency towards a Pomeranchuk instability.
361: Similarly to the $J=0$ case, we observe a smaller variation of the vertices
362: ${\cal R}_{{\bf k}}^{m,r}$ around the FS, than in the momentum dependence of
363: the corresponding eigenfunctions of the Bethe-Salpeter equation.
364: 
365: With increasing $t^{\prime }$ to $0.3t$ (Fig. 4) the largest eigenvalue
366: occurs in the singlet dSC channel. The pp{\bf 0} pair wavefunction maintains
367: its shape with a slight deviation from the ($\cos k_x - \cos k_y$)-form,
368: i.e. a flattening near the BZ diagonal. In the phs{\bf Q} and phc{\bf 0}
369: channels we observe again a weaker momentum variation of ${\cal R}_{{\bf
370: k}}^{m,r}$. The vertex ${\cal R}_{{\bf k}}^{pp{\bf 0},d\text{-wave}}$ instead
371: is very close to the shape of the corresponding Bethe-Salpeter eigenfunctions
372: in the pp{\bf 0} channel.
373: 
374: \begin{figure}[t!]
375: \psfig{file=Fig4.eps,width=90mm,silent=} \vspace{2mm}
376: \caption{Same as in Fig. 1 for $t^{\prime }=0.3t,$ $U=2t$, $J=0.3t$. }
377: \label{fig:Fig4}
378: \end{figure}
379: 
380: For $t^{\prime }=0.45t$ a ferromagnetic instability is expected
381: \cite{SalmHon1,KK}, and we start again with the vH band filling case for
382: $\mu =0$. We increase the interaction strength to $U=3t$, since the
383: corresponding crossover temperatures for the FM instability are lower. The
384: largest Bethe-Salpeter eigenvalues arise in the FM (phs{\bf 0}), pSC (pp{\bf 0}
385: triplet) and AFM channels (phs{\bf Q}). Remarkably, the wavefunction in the
386: {\it triplet} pp{\bf 0} channel deviates significantly from the $\sin k_x$
387: form (see Fig. 5b), but its eigenvalue remains smaller than the eigenvalue
388: for ferromagnetism. This deviation is also well reproduced by the
389: corresponding momentum dependence of the vertex ${\cal R}_{{\bf k}}^{pp{\bf
390: 0},p\text{-wave}}$.
391: 
392: On moving away from the vH band filling at $t^{\prime }=0.45t$, we observe a
393: further increase of the eigenvalue in the triplet pairing channel (Fig. 6),
394: which however remains smaller than the eigenvalue of the ferromagnetic
395: instability at the lowest temperature we can safely reach in the fRG flow.
396: Simultaneously,
397: the eigenfunction in the triplet superconducting channel slightly distorts
398: towards the wave function
399: $f_{{\bf k}}^{(p\text{-wave})}$, but strong deviations persist in both, the
400: eigenfunction of the Bethe-Salpeter equation and the vertex
401: ${\cal R}_{{\bf k}}^{pp{\bf 0},p\text{-wave}}$. Singlet superconductivity,
402: which was not among the dominant instabilities at $\mu =0$, is also
403: significantly enhanced for $\mu >0$. There are two eigenfunctions, which are
404: symmetric and antisymmetric with respect to reflection at the BZ diagonal with
405: almost equal eigenvalues in the pp singlet channel. Both eigenfunctions are
406: essentially nonzero at ($\pi,0$) and (0,$\pi$) only, and therefore not
407: expected to be stabilized thermodynamically. Because of the presence of two
408: nearly-degenerate eigenfunctions, the three-point vertex ${\cal R}_{{\bf
409: k}}^{pp{\bf 0},d\text{-wave}}$ in this case substantially deviates from the
410: momentum dependence of the leading eigenfunction.
411: 
412: \begin{figure}[t!]
413: \psfig{file=Fig5.eps,width=90mm,silent=} \vspace{2mm}
414: \caption{Same as in Fig. 1 for $t^{\prime }=0.45t,$ $U=3t$}
415: \label{fig:Fig5}
416: \end{figure}
417: 
418: In conclusion, we have investigated the symmetry of the leading
419: instabilities of the 2D $t$-$t^{\prime }$ Hubbard model using as a novel
420: tool the combination of the Bethe-Salpeter equation and the fRG approach.
421: Although in most cases the leading instabilities coincide with those from
422: susceptibility based analyses, the true shape of the eigenfunctions
423: significantly differs from the s-, p- or d-wave basis functions. At small
424: $t^{\prime }$ we find in particular a flattening of the eigenfunction in the
425: pairing pp{\bf 0} channel near the nodes -- in qualitative agreement with
426: experimental data for cuprates \cite{Mesot}. The addition of a direct spin
427: exchange interaction to the Hubbard model at weak-coupling was essential to
428: reproduce a ($\cos k_x-\cos k_y$)-like form of the superconducting gap at low
429: and intermediate $t^{\prime}$. The instability towards triplet pairing at
430: larger $t^{\prime }$ also shows a substantial deviation from the standard
431: p-wave $\sin k_x$-form. In many
432: cases we have investigated, the eigenfunctions of the Bethe-Salpeter equations
433: strongly deviate from the corresponding three-point vertices ${\cal R}_{{\bf
434: k}}^{m,r}$, which enter the order parameter susceptibilities, due to either a
435: qualitative difference of the eigenfunctions from the basis functions
436: $f_{{\bf k}}^{(s,p,d)\text{-wave}}$ or the presence of several almost
437: degenerate eigenfunctions. The proposed new technique may prove most useful in
438: future studies of magnetic and superconducting instabilities without
439: presupposing a special momentum dependence of the candidate order parameters.
440: 
441: This work was supported by the Deutsche Forschungsgemeinschaft through SFB
442: 484. We gratefully acknowledge discussions with D. J. Scalapino, W. Metzner,
443: C. Honerkamp, and A. Castro-Neto.
444: 
445: \begin{figure}[t!]
446: \psfig{file=Fig6.eps,width=90mm,silent=} \vspace{2mm}
447: \caption{Same as in Fig. 1 for $t^{\prime }=0.45t,$ $U=3t$, and $\mu =0.035t$.}
448: \label{fig:Fig6}
449: \end{figure}
450: 
451: \begin{references}
452: \bibitem{Mesot}  J. Mesot, M. R. Norman, H. Ding, M. Randeria, J. C.
453: Campuzano, A. Paramekanti, H. M. Fretwell, A. Kaminski, T. Takeuchi, T.
454: Yokoya, T. Sato, T. Takahashi, T. Mochiku, and K. Kadowaki, Phys. Rev. Lett.
455: {\bf 83}, 840 (1999).
456: 
457: \bibitem{Scalapino}  D. J. Scalapino, J. Low Temp. Phys. {\bf 117,} 179
458: (1999).
459: 
460: \bibitem{BSW}  N. E. Bickers, D. J. Scalapino, and S. R. White, Phys. Rev.
461: Lett. {\bf 62,} 961 (1989).
462: 
463: %\bibitem{Zhang}  S. C. Zhang, Science {\bf 275,} 1089 (1997); E. Demler and
464: %S. C. Zhang, Nature (London) {\bf 396,} 733 (1998).
465: 
466: \bibitem{Pines}  J. Schmalian, D. Pines, and B. Stojkovic, Phys. Rev. Lett.
467: {\bf 80,} 3839 (1998).
468: 
469: \bibitem{Chubukov}  A. Chubukov and D. Morr, Phys. Rep. {\bf 288}, 355
470: (1997); A. Abanov and A. Chubukov, Phys. Rev. Lett. {\bf 84}, 5608 (2000).
471: 
472: \bibitem{PTReview}  Y. Maeno, T. M. Rice, and M. Sigrist, Physics Today {\bf %
473: 54}, 42 (2001).
474: 
475: \bibitem{SrRuGap}  K. Deguchi, Z. Q. Mao, and Y. Maeno, J. Phys. Soc. Jpn.
476: {\bf 73}, 1313 (2004).
477: 
478: \bibitem{Mazin}  I. I. Mazin and D. J. Singh, Phys. Rev. Lett. {\bf 79}, 733
479: (1997); ibid. {\bf 82}, 4324 (1999).
480: 
481: \bibitem{Murakami}  S. Murakami, N. Nagaosa, and M. Sigrist, Phys. Rev.
482: Lett. {\bf 82}, 2939 (1999).
483: 
484: \bibitem{Maeno}  Y. Sidis, M. Braden, P. Bourges, B. Hennion, S. Nishizaki,
485: Y. Maeno, and Y. Mori, Phys. Rev. Lett. {\bf 83}, 3320 (1999).
486: 
487: \bibitem{ED}  N. Kikugawa and Y. Maeno, cond-mat/0211248 (unpublished).
488: 
489: \bibitem{Maeno1}  F. Nakamura, T. Goko, M. Ito, T. Fujita, S. Nakatsuji, H.
490: Fukazawa, Y. Maeno, P. Alireza, D. Forsythe, and S. R. Julian, Phys. Rev. B
491: {\bf 65}, 220402(R) (2002).
492: 
493: %\bibitem{ARPES}  A. Ino, C. Kim, M. Nakamura, T. Yoshida, T. Mizokawa, A.
494: %Fujimori, Z.-X. Shen, T. Kakeshita, H. Eisaki, and S. Uchida, Phys. Rev. B
495: %{\bf 65}, 094504 (2002).
496: 
497: %\bibitem{ARPES1}  P.V. Bogdanov, A. Lanzara, X.J. Zhou, S.A. Kellar, D.L.
498: %Feng, E.D. Lu, H. Eisaki, J.-I. Shimoyama, K. Kishio, Z. Hussain, and Z. X.
499: %Shen, Phys. Rev. B {\bf 64}, 180505 (2001).
500: 
501: %\bibitem{ARPES2}  D.L. Feng, C. Kim, H. Eisaki, D.H. Lu, A. Damascelli, K.M.
502: %Shen, F. Ronning, N.P. Armitage, N. Kaneko, M. Greven, J. Shimoyama, K.
503: %Kishio, R. Yoshizaki, G.D. Gu, and Z.-X. Shen, Phys. Rev. B {\bf 65},
504: %220501(R) (2002).
505: 
506: %\bibitem{Damascelli}  A. Damascelli, D. H. Lu, K. M. Shen, N. P. Armitage,
507: %F. Ronning, D. L. Feng, C. Kim, Z.-X. Shen, T. Kimura, Y. Tokura, T.
508: %Tsukuba, Q. Mao, and Y. Maeno, Phys. Rev. Lett. {\bf 85}, 5194 (2000).
509: 
510: %\bibitem{SrBand}  T. Oguchi, Phys. Rev. B {\bf 51}, 1385 (1995); D. J.
511: %Singh, Phys. Rev. B {\bf 52}, 1358 (1995).
512: 
513: %\bibitem{ARS}  D. F. Agterberg, T. M. Rice, and M. Sigrist, Phys. Rev. Lett.
514: %{\bf 78}, 3374 (1997).
515: 
516: \bibitem{KL}  W. Kohn and J. M. Luttinger, Phys. Rev. Lett. {\bf 15}, 524
517: (1965).
518: 
519: \bibitem{BS}  N. F. Berk and J. R. Schrieffer, Phys. Rev. Lett. {\bf 17},
520: 433 (1966).
521: 
522: \bibitem{SLH}  D. J. Scalapino, E. Loh, Jr., and J. E. Hirsch, Phys. Rev. B
523: {\bf 34}, 8190 (1986); ibid. {\bf 35}, 6694 (1987).
524: 
525: %\bibitem{White}  S. R. White et al., Phys. Rev. B {\bf 39}, 839 (1989);
526: %N. Bulut et al., Phys. Rev. B {\bf 47}, 2742 (1993)
527: 
528: %\bibitem{Santos}  R. R. Santos, Phys. Rev. B {\bf 39}, 7259 (1989).
529: 
530: \bibitem{Zanchi}  D. Zanchi and H.J. Schulz, Phys. Rev. B {\bf 54}, 9509
531: (1996); ibid. {\bf 61}, 13609 (2000).
532: 
533: \bibitem{Metzner}  C. J. Halboth and W. Metzner, Phys. Rev. B {\bf 61}, 7364
534: (2000).
535: 
536: \bibitem{SalmHon}  C. Honerkamp, M. Salmhofer, N. Furukawa, and T. M. Rice,
537: Phys. Rev. B {\bf 63}, 035109 (2001).
538: 
539: \bibitem{SalmHon1}  C. Honerkamp and M. Salmhofer, Phys. Rev. Lett. {\bf 87}%
540: , 187004 (2001); Phys. Rev. B{\bf \ 64}, 184516 (2001).
541: 
542: \bibitem{KK}  A. A. Katanin and A. P. Kampf, Phys. Rev. B {\bf 68}, 195101
543: (2003).
544: 
545: \bibitem{KK1}  A. P. Kampf and A. A. Katanin, Phys. Rev. B {\bf 67}, 125104
546: (2003).
547: 
548: \bibitem{Neumayr}  See also the renormalized perturbation theory analysis of
549: A. Neumayr and W. Metzner, Phys. Rev. B {\bf 67}, 035112 (2003).
550: 
551: \bibitem{BetheSal}  N. Bulut, D. J. Scalapino, and S. R. White, Phys. Rev. B
552: {\bf 47}, 6157 (1993); ibid. {\bf 47}, 14599 (1993).
553: 
554: \bibitem{SalmHon02}  C. Honerkamp, M. Salmhofer, and T. M. Rice, Eur. Phys.
555: J. B {\bf 27}, 127 (2002).
556: 
557: \bibitem{Pomeranchuk}  I. J. Pomeranchuk, Sov. Phys. JETP {\bf 8}, 361
558: (1958).
559: 
560: \bibitem{DDW}  S. Chakravarty, R. B. Laughlin, D. K. Morr, and C. Nayak,
561: Phys. Rev. B {\bf 63}, 094503 (2001).
562: 
563: \bibitem{Yamase}  H. Yamase and H. Kohno, J. Phys. Soc. Jpn. {\bf 69}, 332
564: (2000); ibid. {\bf 69}, 2151 (2000).
565: 
566: \bibitem{Metzner1}  C. J. Halboth and W. Metzner, Phys. Rev. Lett. {\bf 85},
567: 5162 (2000).
568: \end{references}
569: 
570: \end{document}