1: % "VERSION 3''
2: % last correction: November 28, 2006 by PK"
3: \tolerance = 10000
4: %
5: %\documentclass[twocolumn,showpacs,prl,amsmath,amssymb,floatfix]{revtex4}
6: %
7: % take this for final format:
8: \documentclass[twocolumn,showpacs,prb,amsmath,amssymb,floatfix,eqsecnum]{revtex4}
9: %
10: % take this for submission in preprint style
11: %\documentclass[preprint,showpacs,prb,amsmath,amssymb,floatfix,eqsecnum]{revtex4}
12: %
13: %\documentclass[twocolumn,showpacs,superscriptaddress,prb,amsmath,amssymb,floatfix,eqsecnum]{revtex4}
14: %\documentclass[galley,showpacs,prl,amsmath,amssymb]{revtex4}
15: %
16: %\usepackage{dcolumn}
17: \usepackage{amsmath,amssymb}
18: %
19: %\usepackage{showlabels}
20: %\usepackage{drftcite}
21: %
22: \usepackage{bm}
23: \usepackage{epsfig}
24: \usepackage{psfrag}
25:
26: % set \bd to \bf or \bm
27: \newcommand{\bd}{\bm}
28:
29: \begin{document}
30:
31: \title{Fermi surface renormalization and confinement in two coupled metallic chains}
32:
33: \author{Sascha Ledowski and Peter Kopietz}
34:
35: \affiliation{Institut f\"{u}r Theoretische Physik, Universit\"{a}t
36: Frankfurt, Max-von-Laue Strasse 1, 60438 Frankfurt, Germany}
37:
38: \date{\today}
39: \date{August 4, 2006}
40:
41:
42:
43: \begin{abstract}
44: Using a non-perturbative functional
45: renormalization group approach involving both fermionic and bosonic fields
46: we calculate the interaction-induced change of the Fermi surface
47: of spinless fermions moving on two chains
48: connected by weak interchain hopping
49: $t_{\bot}$.
50: For a model containing interband backward scattering only we show that
51: the distance $ \Delta$ between the Fermi momenta
52: associated with the bonding and
53: the antibonding band can
54: be strongly reduced, corresponding to a large reduction of the
55: effective interchain hopping $t_{\bot}^{\ast} \propto \Delta $.
56: A self-consistent one-loop approximation
57: neglecting marginal vertex corrections and
58: wave-function renormalizations predicts
59: a confinement transition
60: for sufficiently large interchain backscattering, where
61: the renormalized $t_{\bot}^{\ast}$ vanishes.
62: However,
63: a more accurate calculation taking
64: vertex corrections and wave-function renormalizations
65: into account predicts only weak confinement
66: in the sense that $0< | t_{\bot}^{\ast} | \ll | t_{\bot} |$.
67: Our method can be applied to other strong-coupling problems
68: where the dominant scattering channel is known.
69:
70:
71: \end{abstract}
72:
73: \pacs{71.10.Pm, 71.27.+a,71.10.Hf}
74:
75:
76: %\preprint{}
77:
78: %\draft
79:
80: \maketitle
81:
82: \section{Introduction}
83:
84:
85:
86:
87:
88:
89: In strongly correlated Fermi systems
90: electron-electron interactions
91: can have drastic effects
92: on the geometry and the topology of the Fermi surface.
93: For example, strong
94: forward scattering can give rise to a Pomeranchuk instability,
95: where the shape of the Fermi surface spontaneously changes such that
96: it has a lower symmetry than
97: the underlying lattice \cite{Pomenanchuk58}.
98: Another example is the
99: Lifshitz transition~\cite{Lifshitz60}, where
100: the topology of the
101: Fermi surface changes discontinuously without symmetry breaking
102: as a function of some
103: external control parameter. This
104: gives rise to anomalies in thermodynamic and kinetic
105: properties of a metal.
106: Conditions on the range and the strength of the interaction leading to
107: Pomeranchuk and Lifshitz transitions have recently been derived
108: in Ref.~[\onlinecite{Quintanilla06}].
109:
110: In this work we shall focus on another type of phase transition
111: associated with the geometry of the Fermi surface, which we call
112: {\it{confinement transition}}.
113: This quantum phase transition can
114: occur in quasi one-dimensional metals with an open Fermi surface,
115: consisting of two disconnected weakly curved sheets.
116: Due to strong interactions, the curvature of the Fermi sheets
117: can be smoothed out and can eventually vanish in certain sectors.
118: In the extreme case, the
119: renormalized Fermi surface
120: consists of two completely flat parallel planes.
121: The motion of the fermions in real space is then
122: strictly one-dimensional, although
123: in the absence of interactions it is not.
124: We therefore call such a transition {\it{confinement transition}}.
125: In the confined state the low-energy properties of the system
126: resemble that of a one-dimensional Luttinger liquid.
127: Because the Fermi surface in the confined state
128: has an additional nesting symmetry, at the confinement transition the symmetry
129: of the Fermi surface increases,
130: in contrast to the Pomeranchuk instability, where interactions
131: lower the symmetry of the Fermi surface.
132: An interaction-induced flattening of the Fermi surface
133: might also play a role in
134: the Hubbard model close to half filling, where the bare Fermi surface
135: consists of four almost flat sectors.
136: Completely flat parts
137: of the Fermi surface can give rise to non-Fermi liquid
138: behavior \cite{Luther94,Zheleznyak97,Ferraz03}.
139: Evidence of an interaction-induced flattening of the Fermi surface of the
140: Hubbard model close to half filling has been found in Ref.~[\onlinecite{Honerkamp01}].
141:
142:
143: Similar to the Pomeranchuk transition,
144: the confinement transition is a strong-coupling phenomenon.
145: Hence, the usual weak coupling perturbative expansions are not sufficient to study
146: this transition.
147: Due to a lack of controlled methods to deal with strongly interacting fermions
148: in dimensions larger than one, it is very difficult to study
149: the confinement transition.
150: To shed some light on the underlying mechanism, we shall
151: in this work consider the simpler problem of just two metallic spinless chains
152: coupled by weak interchain hopping $t_{\bot}$.
153: The confined state corresponds to a vanishing
154: renormalized interchain hopping $ t_{\bot}^{\ast} =0$, so that
155: electrons cannot move from one chain to the other, in spite of the fact
156: that the bare interchain hopping is finite.
157: In a subsequent paper~\cite{Ledowski06}, we shall discuss the more difficult confinement
158: problem in an infinite array of coupled chains. It turns out that the basic
159: mechanism
160: responsible for the tendency towards confinement can
161: already be understood from the simpler two-chain problem.
162:
163: Because perturbation theory in the two-chain problem
164: is plagued by the usual infrared divergencies
165: of one-dimensional Fermi systems, even in the limit of
166: weak interactions the Fermi surface cannot be calculated
167: within renormalized
168: perturbation theory\cite{Neumayr03,Dusuel03}.
169: In dimensions $ D \geq 2$ the Fermi surface deformation has been studied
170: to all orders in perturbation theory in Ref.~[\onlinecite{Feldman96}]
171: for a general class of models.
172: Within the framework
173: of the renormalization group (RG)
174: the Fermi surface can be defined non-perturbatively from the
175: requirement that the relevant coupling constants $r_l ( {\bd{k}}_F )$ related to the
176: self-energy $ \Sigma ( {\bd{k}}_F , \omega =0)$ at the true Fermi surface
177: ${\bd{k}}_F$
178: flow into a fixed point of the RG~\cite{Kopietz01,Ledowski03}.
179: In Ref.~[\onlinecite{Ledowski05}]
180: we have calculated the shift of the
181: Fermi surface in the two-chain system
182: within the usual weak coupling
183: expansion of the RG $\beta$-functions.
184: We have shown that interchain backscattering gives rise to the dominant
185: logarithmic renormalization
186: of the distance
187: $
188: \Delta = k^{+} - k^{-}
189: $
190: between the Fermi momenta $k^{+}$ and $k^{-}$
191: associated with the bonding and the antibonding band.
192: Denoting by $\Delta_1$ the value of $\Delta$ within the Hartree-Fock
193: approximation, the self-consistency condition for the
194: true Fermi point distance in the spinless two-chain system
195: can be cast into the form
196: \begin{equation}
197: \Delta = \frac{\Delta_1 }{ 1 + 2 g_0^2 \ln ( \Lambda_0 / \Delta )}
198: \label{eq:deltaprevious}
199: \; ,
200: \end{equation}
201: where $\Lambda_0$ is an ultraviolet cutoff and
202: $g_0$ is the bare value of the dimensionless coupling constant describing
203: interchain backscattering, which will be defined more precisely in
204: Sec.~\ref{subsec:relevant}.
205: From Eq.~(\ref{eq:deltaprevious})
206: we see that sufficiently large interchain backscattering strongly reduces
207: the value of $\Delta$. But the renormalized $\Delta$ never vanishes, so that
208: there is no true confinement transition. One should keep in mind, however, that
209: Eq.~(\ref{eq:deltaprevious}) has been derived by expanding the RG $\beta$-functions to second order in the
210: coupling constants, so that it is not allowed
211: to extrapolate this expression to large values of $g_0$.
212:
213: To find out whether in the spinless two-chain system sufficiently strong
214: interchain-backscattering
215: can give rise to a confinement transition where the renormalized effective
216: interchain hopping $t_{\bot}^{\ast} \propto k^+ - k^-$ vanishes,
217: we use here a generalization of
218: the collective field functional RG approach
219: with momentum transfer cutoff
220: developed in Ref.~[\onlinecite{Schuetz05}].
221: It turns out that with this approach we can
222: analyze the regime where the dimensionless
223: interchain backscattering interaction $g_0$ is of the
224: order of unity.
225: The crucial point is that from the weak coupling analysis \cite{Ledowski05}
226: we know that the confinement transition is
227: driven by interchain backscattering,
228: so that it is natural to decouple the interaction in this
229: scattering channel via a suitable bosonic Hubbard-Stratonovich field.
230: Simple truncations in the resulting mixed boson-fermion theory
231: correspond to infinite resummations
232: in an expansion of the RG $\beta$-functions
233: in powers of $g_0$.
234:
235:
236:
237:
238: \section{Effective low-energy model}
239:
240:
241:
242: We consider spinless fermions
243: confined to two chains that are coupled by weak interchain hopping $t_{\bot}$.
244: The kinetic energy of the two-chain system is diagonalized
245: by forming symmetric (bonding band) and antisymmetric (antibonding band)
246: combinations
247: of the eigenstates associated with isolated chains.
248: Denoting by $\epsilon_k$ the energy dispersion of a
249: single chain in the absence
250: of interchain hopping,
251: the energy dispersion of the non-interacting system
252: is $\epsilon_{ k}^{\sigma} = \epsilon_k - \sigma t_{\bot}$,
253: where $\sigma = + 1$ labels the bonding band and $\sigma =-1$ labels
254: the antibonding band.
255: It is useful to think of $\sigma$ as a pseudospin label \cite{Fabrizio93},
256: in which case $t_{\bot} =h $ corresponds to a uniform magnetic field $h$
257: in the $z$-direction.
258:
259: The problem of finding the low-energy
260: properties of two coupled metallic chains has been studied
261: previously by many
262: authors \cite{Dusuel03,Ledowski05,Fabrizio93,Brazovskii85,Bourbonnais91,Kusmartsev92,Finkelstein93,Boies95,Balents96,Arrigoni98,Ledermann00,Louis01,Caron02,Bourbonnais04,Nickel06,Tsuchiizu06}.
263: However, the problem of self-consistently calculating the
264: true Fermi surface has only recently been addressed \cite{Dusuel03,Ledowski05,Louis01}.
265: At low energies all
266: possible scattering processes in the spinless two-chain system
267: can be divided into four different
268: classes~\cite{Fabrizio93}: (1) forward
269: scattering processes, parameterized in terms of
270: three different coupling constants $f^{++}$, $f^{--}$ and $f^{+-} = f^{-+}$, where the labels
271: indicate the band of the fermions involved in the scattering process;
272: (2) interband backward scattering, which in pseudospin language corresponds
273: to transverse spin-exchange, so that we shall call the corresponding
274: dimensionful coupling constant $J^{\bot}$
275: (the associated dimensionless coupling
276: $g_0$ will be introduced in Sec.~\ref{subsec:relevant});
277: (3) pair-tunneling, which can also
278: be viewed as interband Umklapp
279: scattering, parameterized in terms
280: of a coupling constant by $u$;
281: and finally (4) intraband Umklapp scattering,
282: which is expected to be important only at commensurate fillings.
283: Neglecting the latter process
284: and setting for simplicity
285: $f^{++} = f^{--}$, the low-energy interactions can
286: be expressed in terms of four marginal coupling constants
287: $f = \frac{1}{2}( f^{+-} + f^{++})$,
288: $J^{\parallel} = \frac{1}{2}( f^{+-} - f^{++} )$,
289: $ J^{\bot}$, and $u$.
290: In the bonding-antibonding basis the
291: system can then be modeled by the following
292: effective Euclidean action in pseudospin notation,
293: \begin{eqnarray}
294: S [ \bar{\psi} , \psi ] & = & \sum_{\sigma } \int_K
295: ( - i \omega + \xi_{ k}^{ \sigma } ) \bar{\psi}_{ K }^{\sigma}
296: \psi_{ K}^{ \sigma}
297: \nonumber
298: \\
299: & & \hspace{-15mm} + \frac{1}{2} \int_{\bar{K}}
300: \left[ f (\bar{k} ) \bar{\rho}_{ \bar{K} }
301: \rho_{\bar{K}}
302: - J^{\parallel} (\bar{k} )
303: \bar{m}_{ \bar{K}} m_{ \bar{K}} \right]
304: \nonumber
305: \\
306: & & \hspace{-15mm} + \int_{ \bar{K}}
307: \left[ u ( \bar{k} )\left(
308: \bar{s}_{\bar{K}} \bar{s}_{-\bar{K}} + s_{\bar{K}} s_{-\bar{K}} \right)
309: - 2 J^{\bot} ( \bar{k} ) \bar{s}_{ \bar{K} } s_{\bar{K}} \right]
310: \; ,
311: \nonumber
312: \\
313: & &
314: \label{eq:action1}
315: \end{eqnarray}
316: where $\xi_{ k}^{ \sigma} = \epsilon_k - \mu - \sigma h $, and
317: we have introduced the following composite fields,
318: \begin{subequations}
319: \begin{eqnarray}
320: \rho_{\bar{K}} & = & \sum_\sigma \int_K \bar{\psi}_{ K}^{ \sigma}
321: \psi_{ K + \bar{K}}^{ \sigma}
322: \; ,
323: \\
324: m_{\bar{K}} & = & \sum_\sigma \sigma \int_K \bar{\psi}_{ K }^{\sigma}
325: \psi_{ K + \bar{K} }^{ \sigma}
326: \; ,
327: \\
328: s_{\bar{K}} & = & \int_{{K}} \bar{\psi}_{ K }^{ -} \psi_{ K+ \bar{K} }^{+}
329: \; .
330: \end{eqnarray}
331: \end{subequations}
332: We use the
333: imaginary time formalism at zero temperature and
334: have introduced
335: collective labels $K = ( k , i \omega )$
336: for fermionic fields and
337: $\bar{K} = ( \bar{k} , i \bar{\omega} )$ for bosonic fields, with the notation
338: $\int_K = \int \frac{dk d\omega}{(2\pi)^2}$.
339: Note that the Fourier components of the density and the longitudinal
340: spin-density field satisfy
341: $\rho_{-K} = \rho_K $ and $m_{-K} = m_K$, while
342: the spin-flip field $s_K$
343: is complex and do not have this symmetry.
344: The interaction functions $f ( \bar{k} )$,
345: $J^{\parallel} ( \bar{k} ) $, $J^{\bot} ( \bar{k} )$, and $u ( \bar{k} )$
346: should be considered
347: as phenomenological low-energy couplings
348: which are only non-zero for
349: $ | \bar{k} | \leq \Lambda_0 \ll k_F$.
350: Hence, these couplings should not be directly compared with
351: the bare coupling constant in the Hubbard model \cite{footnotehubbard}.
352: The signs and normalizations in Eq.~(\ref{eq:action1}) are chosen such that
353: for $J^{\bot} = J^{\parallel}$ the model has
354: rotational invariance in pseudo-spin space, and
355: that for the Hubbard model all couplings are
356: positive \cite{footnotehubbard}.
357: However, in our effective low energy model
358: there is no reason to expect
359: rotational invariance in pseudospin-space, so that in general
360: $J^{\parallel} ( \bar{k} ) \neq J^{\bot} ( \bar{k} )$.
361:
362: The model defined in (\ref{eq:action1}) is still quite complicated and contains
363: many interaction processes which are not essential for the confinement transition.
364: In fact, from our previous weak
365: coupling analysis~\cite{Ledowski05} we know that
366: the dominant renormalization of the difference between the Fermi points is due to
367: the interchain backscattering process described
368: by the coupling $J^{\bot} ( \bar{k} )$.
369: In this work we study a minimal model describing the confinement transition
370: by simply neglecting the forward scattering
371: interactions $f ( \bar{k} )$ and $J^{\parallel} ( \bar{k} )$, as well as the
372: pair tunneling coupling $u ( \bar{k} )$ in Eq.~(\ref{eq:action1}).
373: However it is known~\cite{Brazovskii85,Boies95}
374: that sufficiently strong pair tunneling can destabilize the Luttinger liquid phase;
375: we shall come back to this point in Sec.~\ref{sec:conclusion}.
376: In pseudospin language, our model then describes a one-dimensional
377: spin $S=1/2$
378: Fermi system subject to a uniform magnetic field in $z$-direction with an attractive
379: ferromagnetic spin exchange involving only the transverse ($xy$) spin components.
380: The latter tends to align the spins in the direction perpendicular to the
381: magnetic field. The Fermi surface renormalization is essentially
382: determined by the competition between the $xy$-exchange interaction and the
383: external constraint imposed by the
384: uniform magnetic field, which tends to align the spins along the
385: $z$-axis.
386: The phase diagram of the model (\ref{eq:action1}) in the space of all couplings
387: has been discussed by Fabrizio \cite{Fabrizio93}.
388: The qualitative behavior of the weak coupling RG flow
389: in the space of couplings spanned by $J^{\parallel}$,
390: $J^{\bot}$ and $u$ is shown in Fig.~\ref{fig:flowLL}.
391: %
392: %
393: \begin{figure}[tb]
394: \centering
395: % \psfrag{g}{$J^{\bot}$}
396: % \psfrag{f}{$J^{\parallel}$}
397: % \vspace{7mm}
398: % \includegraphics{fig1.eps}
399: \epsfig{file=fig1.eps,width=75mm}
400: % \vspace{4mm}
401: \caption{%
402: (Color Online) Qualitative behavior of the weak coupling RG flow
403: of the model (\ref{eq:action1}) in the space of coupling constants $J^{\parallel}$,
404: $J^{\bot}$ and $u$. The thick black line is the line of fixed points
405: describing the stable Luttinger liquid phase.
406: }
407: \label{fig:flowLL}
408: \end{figure}
409: %
410: %
411: Obviously, there is a finite regime in coupling space where the
412: spinless two-chain system is a stable Luttinger liquid, with gapless excitations.
413: In this work we shall assume that the qualitative fixed point structure
414: suggested by the weak coupling analysis remains correct even in the strong
415: coupling regime. We can therefore choose the bare parameters such that the system
416: belongs to the basin of attraction of the Luttinger liquid fixed point manifold.
417:
418:
419: At low energies we may further simplify our model
420: (at least in the deconfined phase) by linearizing the energy
421: dispersion at the Fermi surface, which
422: for our model consists of four points $\alpha k^{ \sigma}$,
423: where the chirality index $\alpha = \pm 1$ labels the left/right
424: Fermi point.
425: Note that the true Fermi points are defined via
426: \begin{equation}
427: \epsilon_{ \alpha {k}^{ \sigma} } - \mu - \sigma h + \Sigma^{\sigma}
428: ( \alpha k^{ \sigma} , i0 ) =0
429: \; ,
430: \label{eq:FSdef}
431: \end{equation}
432: where $\Sigma^{\sigma} ( \alpha k^{ \sigma} , i0 )$
433: is the exact self-energy for vanishing
434: frequency and for momenta at the true
435: Fermi surface $\alpha k^{\sigma}$ of the interacting system.
436: To linearize the energy dispersion at the true Fermi surface, we add
437: and subtract from the
438: non-interacting energy dispersion the counter-term
439: \begin{equation}
440: \mu^{\sigma}_0 = - \Sigma^{\sigma} ( \alpha k^{ \sigma} , i0 ) \; ,
441: \label{eq:counterdef}
442: \end{equation}
443: and approximate
444: \begin{eqnarray}
445: \xi_{ \alpha k^{ \sigma} + k}^{ \sigma} & = &
446: \epsilon_{ \alpha k^{ \sigma} + k } - \mu - \sigma h
447: \nonumber
448: \\
449: & & \hspace{-16mm} =
450: \epsilon_{ \alpha k^{ \sigma} + k } - \mu - \sigma h
451: + \Sigma^{\sigma} ( \alpha k^{ \sigma} , i0 )
452: -
453: \Sigma^{\sigma} ( \alpha k^{ \sigma} , i0 )
454: \nonumber
455: \\
456: & & \hspace{-16mm} \approx
457: \alpha v^{\sigma}_0 k + \mu^{\sigma}_0
458: \; ,
459: \end{eqnarray}
460: where $v_0^{\sigma}$ is the bare Fermi velocity at the
461: true Fermi surface for pseudospin $\sigma$.
462: In analogy with the
463: definition of the couplings in the
464: Tomonaga-Luttinger model~\cite{Solyom79}, we now generalize
465: the interaction by introducing chirality
466: indices
467: $ J^{\bot} \rightarrow J^{\bot}_{\alpha \alpha^{\prime}} $.
468: We shall refer to the diagonal processes $J^{\bot}_{ \alpha \alpha }$
469: as chiral interactions (these are called $g_4$ processes in the
470: Tomonaga-Luttinger model).
471: Similarly, off-diagonal elements $J^{\bot}_{\alpha , -{\alpha}}$
472: will be called non-chiral processes (corresponding to the $g_2$-processes
473: in the Tomonaga-Luttinger model).
474: Defining new fields
475: \begin{equation}
476: \psi^{\sigma}_{ K \alpha} = \psi_{ \alpha k^{ \sigma} + k , i \omega}^{ \sigma}
477: \; ,
478: \end{equation}
479: we replace the action (\ref{eq:action1})
480: by the following effective low-energy action describing the
481: physics of the
482: confinement transition in our system of two spinless metallic chains,
483: \begin{eqnarray}
484: S [ \bar{\psi} , \psi ] & = & \sum_{\sigma, \alpha } \int_K
485: ( - i \omega + \alpha v^{\sigma}_0 k + \mu^{\sigma}_0 )
486: \bar{\psi}^{\sigma}_{ K \alpha} \psi^{\sigma}_{ K \alpha}
487: \nonumber
488: \\
489: & - & 2 \sum_{ \alpha \alpha^{\prime}}
490: \int_{\bar{K}} J^{\bot}_{ \alpha \alpha^{\prime} }
491: \bar{s}_{ \bar{K} \alpha} s_{\bar{K} \alpha^{\prime}}
492: \; ,
493: \label{eq:action2}
494: \end{eqnarray}
495: where it is understood that the $\bar{k}$-integration
496: has a momentum transfer cutoff $ |\bar{k} | \leq \Lambda_0 \ll k_F$, and
497: \begin{eqnarray}
498: s_{ \bar{K} \alpha} & = & \int_K \bar{\psi}^{-}_{ K \alpha}
499: \psi^{+}_{ K+ \bar{K} \alpha}
500: \; .
501: \end{eqnarray}
502:
503:
504:
505: \section{Exact RG flow equations}
506:
507:
508: \subsection{Hubbard Stratonovich transformation}
509:
510:
511:
512: Because the confinement transition is a strong coupling phenomenon,
513: the usual weak coupling RG approach based on the expansion
514: in powers of $J^{\bot}_{\alpha \alpha^{\prime}}$ is not suitable.
515: To develop a RG approach which does not rely on a weak coupling expansion,
516: we decouple the spin-flip interaction with the help
517: of a complex Hubbard-Stratonovich field $\chi_{\alpha}$.
518: For convenience we collect all fields in a super-field,
519: \begin{equation}
520: \Phi = ( \psi_{\alpha}^{+} , \bar{\psi}_{\alpha}^{+},
521: \psi_{\alpha}^{-} , \bar{\psi}_{\alpha}^{-}, \chi_{\alpha} , \bar{\chi}_{\alpha} )
522: \; .
523: \end{equation}
524: Taking into account that there are two chiralities $\alpha = \pm 1$,
525: our super-field has totally
526: 12 components (8 fermionic and 4 bosonic ones).
527: The ratio of the partition functions with and without interactions can then be written as
528: \begin{equation}
529: \frac{ \cal{Z}}{{\cal{Z}}_0 } = \frac{ \int{\cal{D}} [ \Phi ] e^{ - S_0 [ \Phi ] - S_1 [ \Phi ] } }{
530: \int{\cal{D}} [ \Phi ] e^{ - S_0 [ \Phi ] }}
531: \; ,
532: \end{equation}
533: with the Gaussian part of the effective action given by
534: \begin{eqnarray}
535: S_0 [ \Phi ] & = &
536: \sum_{\sigma, \alpha } \int_K
537: ( - i \omega + \alpha v^{\sigma}_0 k + \mu^{\sigma}_0 )
538: \bar{\psi}^{\sigma}_{ K \alpha} \psi^{\sigma}_{ K \alpha}
539: \nonumber
540: \\
541: &+ & \frac{1}{2} \sum_{ \alpha \alpha^{\prime} }
542: \int_{\bar{K}} [ {\bf{J}}^{\bot} ]^{-1}_{\alpha
543: \alpha^{\prime}} \bar{\chi}_{\bar{K} \alpha} \chi_{\bar{K} \alpha^{\prime} }
544: \; .
545: \label{eq:Gauss}
546: \end{eqnarray}
547: Here ${\bf{J}}^{\bot}$ is a matrix
548: in chirality space with matrix elements given by
549: ${J}^{\bot}_{\alpha \alpha^{\prime}}$, and the interaction is
550: \begin{eqnarray}
551: S_1 [ \Phi ] & = &
552: \sum_{ \alpha } \int_{\bar{K}}
553: \bigl[
554: \bar{s}_{\bar{K} \alpha} \chi_{\bar{K} \alpha} +
555: s_{\bar{K} \alpha} \bar{\chi}_{\bar{K} \alpha} \bigr]
556: \; .
557: \label{eq:S1}
558: \end{eqnarray}
559: A graphical representation of the
560: bare interaction vertices in Eq.~(\ref{eq:S1})
561: is shown in Fig.~\ref{fig:vertexS1}.
562: %
563: %
564: \begin{figure}[tb]
565: \centering
566: % \psfrag{t1}{$T >T_c$}
567: % \vspace{7mm}
568: % \includegraphics{fig1.eps}
569: \epsfig{file=fig2.eps,width=75mm}
570: % \vspace{4mm}
571: \caption{%
572: Bare interaction vertices of the action $S_1 [ \Phi ]$ given in Eq.~(\ref{eq:S1}).
573: The fermionic fields $\psi^{\sigma}$ and $\bar{\psi}^{\sigma}$ are denoted by
574: solid arrows, with the spin-projection $\sigma = \pm 1$ written next to the arrows.
575: Bosonic spin-flip fields $\chi$ and $\bar{\chi}$ are denoted by
576: wavy arrows. Incoming arrows denote $\psi^{\sigma}$ and $\chi$, while
577: outgoing arrows correspond to the conjugate fields
578: $\bar{\psi}^{\sigma}$ and $\bar{\chi}$.
579: }
580: \label{fig:vertexS1}
581: \end{figure}
582: %
583: %
584: The coupled RG flow equations for the one-line irreducible vertices
585: of the above mixed boson-fermion theory can be
586: obtained as a special case of the general flow equations given in
587: Ref.~[\onlinecite{Schuetz05}].
588:
589:
590: \subsection{Functional RG flow equations in momentum transfer cutoff scheme}
591:
592: In order to calculate the true Fermi surface, we need to know
593: the exact counter-term $\mu^{\sigma}_0
594: = - \Sigma ( \alpha k^{\sigma} , i 0 )$, which can be
595: obtained from the flowing self-energy
596: $\Sigma_{\Lambda}^{\sigma} ( K , \alpha )$
597: in the limit of vanishing infrared cutoff $\Lambda \rightarrow 0$.
598: The form of the flow equation
599: for $\Sigma_{\Lambda}^{\sigma} ( K , \alpha )$
600: depends on the
601: RG method employed. Here we use the hierarchy of functional RG
602: equations for the one-line irreducible vertices \cite{Wetterich93,Morris94}
603: of mixed boson-fermion models developed in Ref.~[\onlinecite{Schuetz05}].
604: A similar approach involving both fermionic and bosonic fields has been
605: developed in Refs.~[\onlinecite{Wetterich02,Baier04}].
606: In principle, one can also obtain the flowing self-energy
607: within the purely fermionic parameterization of the
608: hierarchy of flow equations \cite{Kopietz01,Ledowski05,Salmhofer01}.
609: However, with the usual truncations necessary
610: in this approach it is not possible to reach the strong coupling regime.
611:
612:
613: In the momentum transfer cutoff scheme \cite{Schuetz05}
614: we impose a cutoff $\Lambda$ only on the momentum $ \bar{k} $ transfered by the
615: collective bosonic field. The resulting RG flow equation
616: for the fermionic self-energy $\Sigma_{\Lambda}^{\sigma} ( K , \alpha )$ is
617: shown graphically in Fig.~\ref{fig:flowsigma}.
618: %
619: %
620: \begin{figure}[tb]
621: \centering
622: % \psfrag{a}{$\sigma$}
623: % \psfrag{b}{$- \sigma$}
624: % \psfrag{d}{$\huge{\delta_{ \sigma ,-}}$}
625: %\psfrag{u}{$\delta_{ \sigma ,+}$}
626: % \vspace{7mm}
627: \epsfig{file=fig3.eps,width=75mm}
628: % \vspace{4mm}
629: \caption{%
630: Exact flow equation for the fermionic self-energy $\Sigma^{\sigma}_{\Lambda}
631: ( K , \alpha)$
632: in the momentum transfer
633: cutoff scheme.
634: The thick solid arrow is the flowing fermionic Green function and
635: the thick wavy line with a slash is the flowing single scale
636: spin-flip propagator defined in Eq.~(\ref{eq:bossingle}).
637: The one-line irreducible vertices are represented by shaded triangles.
638: A label $(n,m)$ inside a shaded triangle means that the vertex has
639: $n$ fermionic and $m$ bosonic external legs.
640: }
641: \label{fig:flowsigma}
642: \end{figure}
643: %
644: %
645: The corresponding analytic expression is
646: \begin{eqnarray}
647: \partial_{\Lambda} \Sigma^{\sigma}_{\Lambda} ( K, \alpha) & &
648: \nonumber
649: \\
650: & & \hspace{-10mm} =
651: \int_{ \bar{K} } \dot{F}^{ \sigma \bar{\sigma}}_{\Lambda} ( \bar{K} , \alpha )
652: \Gamma^{(2,2)}_{\Lambda} ( K \sigma , -K \sigma ;
653: \bar{K} , - \bar{K} , \alpha )
654: \nonumber
655: \\
656: & & \hspace{-10mm} +
657: \int_{ \bar{K} } \dot{F}^{ \sigma \bar{\sigma}}_{\Lambda} ( \bar{K} , \alpha )
658: G^{ \bar{\sigma}}_{\Lambda} ( K + \bar{K} + \alpha \sigma \Delta, \alpha )
659: \nonumber
660: \\
661: & & \hspace{-5mm} \times
662: \Gamma^{(2,1) }_{\Lambda} ( K \sigma ; K + \bar{K}, \bar{\sigma} ; - \bar{K}, \alpha)
663: \nonumber
664: \\
665: & & \hspace{-5mm} \times
666: \Gamma^{(2,1)}_{ \Lambda} ( K + \bar{K}, \bar{\sigma} ; K, \sigma ; \bar{K} , \alpha)
667: \; .
668: \label{eq:RGselfvertex}
669: \end{eqnarray}
670: Here
671: $G_{\Lambda}^{\sigma} ( K , \alpha )$ is
672: the flowing fermionic single-particle Green function
673: for a given
674: pseudospin $\sigma$ and chirality index $\alpha$.
675: We use the notation $\bar{\sigma} = - \sigma$ and measure
676: the wave-vectors $k$
677: with respect to the true Fermi momenta $\alpha k^{\sigma}$, defining
678: \begin{equation}
679: G_{\Lambda}^{\sigma} ( K , \alpha )
680: = G_{\Lambda}^{\sigma} ( \alpha k^{\sigma} + k , i \omega )
681: \; ,
682: \end{equation}
683: and
684: \begin{equation}
685: G^{ \bar{\sigma}}_{\Lambda} ( K + \alpha \sigma \Delta, \alpha )
686: = G^{ \bar{\sigma}}_{\Lambda} ( \alpha k^{\bar{\sigma}} + k
687: + \alpha \sigma \Delta, i \omega )
688: \; .
689: \end{equation}
690: The shift
691: $ \alpha \sigma \Delta = \alpha \sigma (k^{+} - k^- )$ in the argument
692: of $G_{\Lambda}^{\bar{\sigma}} $ in Eq.~(\ref{eq:RGselfvertex})
693: is due to the fact that in $G_{\Lambda}^{\bar{\sigma}} (K , \alpha )$
694: the wave-vector $k$ is measured with respect to a different reference point than
695: in $G_{\Lambda}^{{\sigma}} ( K , \alpha )$.
696: The function $\dot{F}^{ \sigma \bar{\sigma} }_{\Lambda} ( \bar{K} \alpha )$
697: in Eq.~(\ref{eq:RGselfvertex}) is the
698: single scale bosonic spin-flip
699: propagator, which is defined by
700: \begin{equation}
701: \dot{F}^{ \sigma \bar{\sigma} }_{\Lambda} ( \bar{K} \alpha ) =
702: - \delta ( | \bar{k} | - \Lambda )
703: [ {\bf{F}}^{ \sigma \bar{\sigma}}_{\Lambda} ( \bar{K} )]_{\alpha \alpha}
704: \; ,
705: \label{eq:bossingle}
706: \end{equation}
707: where
708: $ {\bf{F}}^{ \sigma \bar{\sigma}}_{\Lambda} ( \bar{K} ) $ is a matrix
709: in chirality space whose inverse has the matrix elements
710: \begin{equation}
711: [ {\bf{F}}^{\sigma \bar{\sigma}}_{\Lambda} ( \bar{K} ) ]^{-1}_{\alpha \alpha^{\prime} } =
712: [ 2 {\bf{J}}^{\bot} ]^{-1}_{\alpha \alpha^{\prime} } -
713: \delta_{\alpha \alpha^{\prime}}
714: \Pi_{ \Lambda}^{\sigma \bar{\sigma}} ( \bar{K} , \alpha )
715: \; ,
716: \end{equation}
717: where $\Pi_{ \Lambda}^{\sigma \bar{\sigma}} ( \bar{K} , \alpha )$ is the flowing spin-flip
718: susceptibility.
719: In the momentum transfer cutoff scheme, the RG flow of
720: $\Pi_{ \Lambda}^{\sigma \bar{\sigma}} ( \bar{K} , \alpha )$
721: is driven by the one-line irreducible vertex with four external boson legs, as shown
722: in Fig.~\ref{fig:flowPi}.
723: %
724: %
725: \begin{figure}[tb]
726: \centering
727: % \psfrag{t1}{$T >T_c$}
728: % \vspace{7mm}
729: \epsfig{file=fig4.eps,width=70mm}
730: % \vspace{4mm}
731: \caption{%
732: Exact flow equation for the spin-flip susceptibility in the momentum transfer
733: cutoff scheme.
734: }
735: \label{fig:flowPi}
736: \end{figure}
737: %
738: %
739: The vertices $ \Gamma^{(2,1) }_{\Lambda}
740: ( K \sigma ; K^{\prime} \bar{\sigma} ; \bar{K} , \alpha)$ in
741: Eq.~(\ref{eq:RGselfvertex}) are the flowing spin-flip vertices with two fermionic and
742: one bosonic external legs.
743: In the momentum transfer cutoff scheme these vertices
744: satisfy the exact flow equations shown in graphically in Fig.~\ref{fig:flowvert}, with initial
745: condition
746: \begin{equation}
747: \Gamma^{(2,1)}_{ \Lambda_0 } ( K \sigma ;
748: K^{\prime} \bar{\sigma} ; \bar{K}, \alpha ) = 1
749: \; .
750: \label{eq:Gammainitial}
751: \end{equation}
752: %
753: %
754: \begin{figure}[tb]
755: \centering
756: % \psfrag{t1}{$T >T_c$}
757: % \vspace{7mm}
758: \epsfig{file=fig5.eps,width=60mm}
759: % \vspace{4mm}
760: \caption{%
761: Exact flow equation for the spin-flip vertices
762: in the momentum transfer
763: cutoff scheme.
764: }
765: \label{fig:flowvert}
766: \end{figure}
767: %
768: %
769: Finally, the vertex $\Gamma^{(2,2)}_{\Lambda}$
770: on the right-hand side of Eq.~(\ref{eq:RGselfvertex}) is the one-line irreducible
771: vertex with two fermionic and two bosonic external legs.
772: We do not give the flow equation for this vertex, because
773: purely bosonic vertices with more than two external legs
774: and mixed vertices with two fermionic and more than one bosonic external leg
775: have negative scaling dimensions and are irrelevant~\cite{Schuetz05}.
776: We expect that their effect can be implicitly taken into account by re-defining
777: the numerical values of the relevant and marginal couplings \cite{Polchinski84}.
778: % Of course, the purely fermionic vertex
779: % with four external legs as also marginal, but in
780: % the momentum transfer cutoff scheme this vertex
781: % does not couple to the flow of the fermionic
782: % self-energy.
783:
784:
785: The initial condition for
786: the fermionic self-energy at scale $\Lambda = \Lambda_0$ is simply
787: \begin{equation}
788: \Sigma^{\sigma}_{\Lambda_0} ( K , \alpha ) = 0
789: \; .
790: \label{eq:sigmainitial}
791: \end{equation}
792: Similarly, the vertices with two fermion legs and more than one
793: boson leg
794: appearing on the right-hand side of the flow equation for the
795: spin-flip vertex shown in
796: Fig.~\ref{fig:flowvert} also vanish at the initial scale.
797: However,
798: the price we pay for introducing a cutoff only in the momentum transfer
799: are non-trivial initial conditions for the purely bosonic vertices,
800: which are initially given by closed fermion loops \cite{Schuetz05}.
801: In particular,
802: the loop with two external boson legs
803: is initially given by the non-interacting spin-flip susceptibility,
804: \begin{eqnarray}
805: \Pi^{\sigma \bar{\sigma}}_{ \Lambda_0} ( \bar{K} , \alpha ) & = &
806: - \int_K
807: G_{\Lambda_0}^{\sigma} ( K , \alpha ) G_{\Lambda_0}^{\bar{\sigma}}
808: ( K + \bar{K} + \alpha \sigma \Delta, \alpha )
809: \; ,
810: \nonumber
811: \\
812: & &
813: \label{eq:Piinitial}
814: \end{eqnarray}
815: where for our model with linear energy dispersion,
816: \begin{equation}
817: G_{\Lambda_0}^{\sigma} ( K , \alpha ) = \frac{ 1 }{ i \omega - \alpha v^{\sigma}_0 k
818: - \mu^{\sigma}_0 }
819: \; .
820: \label{eq:G0lin}
821: \end{equation}
822: Denoting by
823: \begin{equation}
824: \Delta_{0} =
825: k^{+}_{0} - k^{-}_{0}
826: \label{eq:delta0def}
827: \end{equation}
828: the distance between the Fermi momenta $k_0^{+}$ and $k_{0}^-$ in the absence
829: of interactions, the relation between the true distance
830: $\Delta = k^{+} - k^{-}$
831: and $\Delta_{0}$ can be expressed in terms of the counter-terms
832: $\mu^{\sigma}_0 =
833: - \Sigma ( \alpha k^{\sigma} , i0 , \alpha )$
834: as follows,
835: \begin{equation}
836: \Delta =
837: \Delta_{0}
838: + \left[ \frac{ \mu^{+}_0}{ v^+_0}
839: - \frac{ \mu^{- }_0 }{ v^-_0} \right]
840: \; ,
841: \label{eq:DeltaDelta0}
842: \end{equation}
843: see also Eq.~(\ref{eq:FSdef2}) below.
844: Using Eqs.~(\ref{eq:G0lin}) and (\ref{eq:DeltaDelta0})
845: we can explicitly evaluate Eq.~(\ref{eq:Piinitial}),
846: \begin{eqnarray}
847: \Pi^{\sigma \bar{\sigma}}_{ \Lambda_0} ( \bar{K} , \alpha ) & = &
848: \frac{ 1 }{ 2 \pi v^{\sigma}_0 }
849: \frac{ v^{\bar{\sigma}}_0 ( \sigma \Delta_{0} + \alpha \bar{k}) }{
850: v^{\bar{\sigma}}_0 ( \sigma \Delta_{0} + \alpha \bar{k} )
851: - i \bar{\omega} }
852: \; .
853: \label{eq:Piinitial2}
854: \end{eqnarray}
855:
856:
857:
858:
859: \subsection{Rescaled flow equations and classification of vertices}
860:
861: To classify the various vertices according to their relevance,
862: it is useful to make them dimensionless by multiplying them
863: with a suitable power of the running cutoff $\Lambda$.
864: Following Ref.~[\onlinecite{Schuetz05}],
865: we define dimensionless fermionic labels
866: $Q = ( q , i \epsilon ) = ( k/ \Lambda , i \omega / \Omega_{\Lambda} )$,
867: and bosonic ones
868: $\bar{Q} = ( \bar{q} , i \bar{\epsilon} ) = ( \bar{k} / \Lambda, i \bar{\omega} / \Omega_{ \Lambda} )$. Here
869: $\Omega_{\Lambda} = v_F \Lambda$, where
870: $v_F $ is some average Fermi velocity.
871: For simplicity we shall write the above
872: relations as $Q = K / \Lambda$ and $\bar{Q} = \bar{K} / \Lambda$.
873: We consider all rescaled quantities as functions of the
874: logarithmic flow parameter $l = \ln ( \Lambda_0 / \Lambda)$.
875:
876: In order to define the Fermi surface within the framework
877: of the RG, we subtract the counter-term
878: $\Sigma^{\sigma} ( \alpha k^{ \sigma} , i 0 , \alpha)
879: = - \mu^{\sigma}_0$ from the flowing self-energy
880: and then rescale~\cite{Kopietz01},
881: \begin{eqnarray}
882: \tilde{\Sigma}_{l}^{ \sigma} ( Q , \alpha )
883: & = & \frac{ Z^{\sigma}_{ l}}{ \Omega_{\Lambda} }
884: \left[ \Sigma_{\Lambda}^{ \sigma}
885: ( K , \alpha ) -
886: \Sigma^{\sigma} ( \alpha k^{ \sigma}, i 0 ) \right]
887: \nonumber
888: \\
889: & = &
890: \frac{ Z^{\sigma}_{ l}}{ \Omega_{ \Lambda} }
891: \left[ \Sigma_{\Lambda}^{ \sigma}
892: ( \Lambda Q , \alpha ) + \mu^{\sigma}_0 \right]
893: \; .
894: \label{eq:sigmasub}
895: \end{eqnarray}
896: Here $Z^{\sigma}_l$ is the flowing wave-function renormalization factor,
897: which is defined in terms of the flowing self-energy as follows,
898: \begin{equation}
899: Z^{\sigma }_{ l} = 1 + \left. \frac{ \partial
900: \tilde{\Sigma}_l^{\sigma} ( 0 , i \epsilon , \alpha )}{
901: \partial ( i \epsilon ) } \right|_{ \epsilon =0}
902: \; .
903: \end{equation}
904: The corresponding rescaled fermionic propagator is
905: \begin{equation}
906: \tilde{G}_l^{\sigma} ( Q, \alpha ) = \frac{ \Omega_{\Lambda}}{Z^{ \sigma }_{ l} }
907: G^{ \sigma}_{\Lambda} ( \Lambda Q , \alpha )
908: \; .
909: \label{eq:Gscale}
910: \end{equation}
911: The rescaled self-energy satisfies
912: the exact RG flow equation\cite{Kopietz01,Ledowski03}
913: \begin{equation}
914: \partial_l \tilde{\Sigma}_{l}^{ \sigma} ( Q , \alpha ) =
915: ( 1 - \eta^{\sigma }_{ l} - q \partial_q - \epsilon \partial_{\epsilon} )
916: \tilde{\Sigma}_{l}^{ \sigma} ( Q , \alpha )
917: + \dot{{\Gamma}}_{ l }^{\sigma} ( Q , \alpha)
918: \; ,
919: \label{eq:flowsigma}
920: \end{equation}
921: where the flowing anomalous dimension of the Fermi fields is
922: \begin{equation}
923: \eta_l^{\sigma} = - \partial_l \ln Z^{\sigma}_l =
924: \left. - \frac{ \partial \dot{{\Gamma}}^{\sigma}_{l } ( 0, i \epsilon , \alpha)}{
925: \partial ( i \epsilon) } \right|_{ \epsilon =0}
926: \; ,
927: \label{eq:etadef}
928: \end{equation}
929: and the function $ \dot{{\Gamma}}_{ l }^{\sigma} ( Q , \alpha) $
930: follows from Eq.~(\ref{eq:RGselfvertex}),
931: \begin{eqnarray}
932: \dot{{\Gamma}}_{ l }^{\sigma} ( Q , \alpha)
933: & = & \frac{ Z^{ \sigma }_{ l}}{\Omega_{\Lambda}}[ - \Lambda \partial_{\Lambda}
934: \Sigma_{\Lambda}^{\sigma} ( K, \alpha ) ]
935: \nonumber
936: \\
937: & = &
938: \int_{\bar{Q}}
939: \dot{\tilde{F}}^{ \sigma \bar{\sigma}}_l ( \bar{Q} , \alpha )
940: \tilde{\Gamma}^{(2,2)}_l ( Q , \sigma ; -Q \sigma ; \bar{Q} , - \bar{Q} , \alpha )
941: \nonumber
942: \\
943: & + &\int_{\bar{Q}}
944: \dot{\tilde{F}}^{ \sigma \bar{\sigma}}_l ( \bar{Q} , \alpha )
945: \tilde{G}^{ \bar{\sigma}}_{l} ( Q + \bar{Q} + \alpha \sigma \tilde{\Delta}_l^{\ast}, \alpha )
946: \nonumber
947: \\
948: & & \times
949: \tilde{\Gamma}^{(2,1) }_{ l} ( Q , \sigma ; Q + \bar{Q}, \bar{\sigma} ; - \bar{Q} , \alpha)
950: \nonumber
951: \\
952: & & \times
953: \tilde{\Gamma}^{(2,1)}_{ l } ( Q + \bar{Q}, \bar{\sigma} ; Q \sigma ; \bar{Q} , \alpha)
954: \; ,
955: \label{eq:dotgamma2}
956: \end{eqnarray}
957: where
958: \begin{equation}
959: \tilde{\Delta}_l^{\ast} = \frac{ \Delta}{\Lambda} = \frac{ \Delta}{\Lambda_0} e^l
960: \label{eq:deltastardef}
961: \end{equation}
962: is the rescaled true difference between the Fermi points.
963: The rescaled bosonic
964: single scale propagator follows from Eq.~(\ref{eq:bossingle}),
965: \begin{eqnarray}
966: \dot{\tilde{F}}^{ \sigma \bar{\sigma}}_l ( \bar{Q} , \alpha ) & = &
967: - \frac{ \nu_0}{\bar{Z}_l } \Lambda \dot{F}^{\sigma \bar{\sigma}}_{\Lambda}
968: ( \Lambda \bar{Q} , \alpha )
969: \nonumber
970: \\
971: & = & \delta ( | \bar{q} | -1 )
972: [ \tilde{{\bf{F}}}^{ \sigma \bar{\sigma}}_{l} ( \bar{Q} )]_{\alpha \alpha}
973: \; ,
974: \label{eq:bossinglescale}
975: \end{eqnarray}
976: with
977: \begin{equation}
978: [ \tilde{{\bf{F}}}^{\sigma \bar{\sigma}}_{l} ( \bar{Q} ) ]^{-1}_{\alpha \alpha^{\prime} } =
979: \bar{Z}_l [ 2 \nu_0 {\bf{J}}^{\bot} ]^{-1}_{\alpha \alpha^{\prime} } -
980: \delta_{\alpha \alpha^{\prime}}
981: \tilde{\Pi}_{ l }^{\sigma \bar{\sigma}} ( \bar{Q} , \alpha )
982: \; ,
983: \label{eq:tildeFpropdef}
984: \end{equation}
985: and the rescaled spin-flip susceptibility
986: \begin{equation}
987: \tilde{\Pi}_l^{ \sigma \bar{\sigma}} ( \bar{Q}, \alpha ) =
988: \frac{ \bar{Z}_l}{ \nu_0 } \Pi_{\Lambda}^{\sigma \bar{\sigma}} ( \Lambda \bar{Q} , \alpha )
989: \; .
990: \label{eq:Pisfrescale}
991: \end{equation}
992: Here $\bar{Z}_l$ is the
993: wave-function renormalization factor associated with the
994: bosonic spin-flip field $\chi_{\alpha}$, and the constant
995: $\nu_0$ with units of inverse velocity has been introduced
996: to make all rescaled vertices dimensionless.
997: The rescaled spin-flip vertex in Eq.~(\ref{eq:dotgamma2}) is
998: \begin{eqnarray}
999: \tilde{\Gamma}^{(2,1)}_{ l } ( Q, \sigma ;
1000: Q^{\prime}, \bar{\sigma}; \bar{Q} , \alpha ) & = &
1001: \left[ { \bar{Z}_l }/{ \nu_0 }
1002: \right]^{1/2} \left[ { \Lambda Z_{l}^{+} Z_{l}^- }/{\Omega_{\Lambda} }
1003: \right]^{1/2}
1004: \nonumber
1005: \\
1006: & & \hspace{-15mm} \times
1007: {\Gamma}^{(2,1)}_{ \Lambda } ( \Lambda Q , \sigma ;
1008: \Lambda Q^{\prime}, \bar{\sigma} ; \Lambda \bar{Q}, \alpha )
1009: \; ,
1010: \label{eq:vertexrescale}
1011: \end{eqnarray}
1012: and the rescaled vertex with two external fermion and two boson legs is
1013: \begin{eqnarray}
1014: \tilde{\Gamma}^{(2,2) }_l ( Q^{\prime}, \sigma ; Q , \sigma ; \bar{Q}^{\prime} ,
1015: \bar{Q} , \alpha )
1016: & = & \Lambda Z_l^{\sigma} ( \bar{Z}_l / \nu_0 )
1017: \nonumber
1018: \\
1019: & & \hspace{-38mm} \times
1020: {\Gamma}^{(2,2) }_{ \Lambda } ( \Lambda Q^{\prime} , \sigma;
1021: \Lambda Q, \sigma ; \Lambda \bar{Q}^{\prime}, \Lambda \bar{Q}, \alpha )
1022: \; .
1023: \label{eq:vertexrescale22}
1024: \end{eqnarray}
1025: For convenience we now choose
1026: $\nu_0 = \Lambda / \Omega_{\Lambda} = 1/v_F$ so that the
1027: prefactor on the right-hand side
1028: of Eq.~(\ref{eq:vertexrescale}) reduces to
1029: $[ \bar{Z}_l Z^{+}_l Z^-_l ]^{1/2}$.
1030:
1031:
1032: Let us now classify the various vertices according to their relevance.
1033: First of all, the key quantity to
1034: obtain the counter-terms $\mu^{\sigma}_0$ is
1035: the momentum- and frequency independent part of the rescaled self-energy
1036: $ \tilde{\Sigma}_{l}^{ \sigma} ( Q , \alpha ) $ defined in Eq.~(\ref{eq:sigmasub}),
1037: which we call
1038: \begin{equation}
1039: r^{\sigma }_{ l} = \tilde{\Sigma}_{l}^{ \sigma} ( 0 , \alpha ) =
1040: \frac{ Z^{\sigma}_{ l}}{ \Omega_\Lambda }
1041: \left[ \Sigma_{\Lambda}^{ \sigma}
1042: ( 0 , \alpha ) + \mu^{\sigma}_0 \right]
1043: \; .
1044: \end{equation}
1045: The couplings $r_l^{\sigma}$ satisfy the exact flow equation
1046: \begin{equation}
1047: \partial_l r_{ l }^{ \sigma} = ( 1 - \eta_{l}^{ \sigma} ) r_{ l }^{ \sigma}
1048: + \dot{{\Gamma}}^{\sigma}_{l } ( 0, \alpha)
1049: \; ,
1050: \label{eq:rlflow}
1051: \end{equation}
1052: with initial condition
1053: \begin{equation}
1054: r_0^{\sigma} = \frac{\mu^{\sigma}_0 }{ \Omega_{\Lambda_0}}
1055: = - \frac{ \Sigma^{\sigma} ( \alpha k^{\sigma} , i0 )}{ v_F \Lambda_0 }
1056: \; .
1057: \label{eq:initialr0}
1058: \end{equation}
1059: There are two marginal couplings related to the self-energy. The first is the
1060: wave-function renormalization factor $Z_l^{\sigma}$, which according to
1061: Eq.~(\ref{eq:etadef}) is related to the flowing anomalous dimension via
1062: \begin{equation}
1063: \partial_l Z^{\sigma}_{ l} = - \eta^{\sigma }_{ l} Z^{\sigma }_{ l}
1064: \; .
1065: \label{eq:Zflow}
1066: \end{equation}
1067: The second is
1068: the dimensionless Fermi velocity renormalization factor
1069: \begin{equation}
1070: \tilde{v}_l^{\sigma} = Z_l^{\sigma} + \left. \frac{ \partial
1071: \tilde{\Sigma}_l^{\sigma} ( q , i 0 , \alpha )}{
1072: \partial ( \alpha q) } \right|_{ q =0}
1073: \; ,
1074: \end{equation}
1075: which satisfies the exact flow equation
1076: \begin{equation}
1077: \partial_l \tilde{v}_l^{\sigma} = - \eta_l^{\sigma} \tilde{v}_l^{\sigma} +
1078: \left. \frac{ \partial \dot{\Gamma}^{\sigma}_l ( q , i 0, \alpha)
1079: }{\partial (\alpha q) }
1080: \right|_{ q=0}
1081: \label{eq:vlflow}
1082: \; .
1083: \end{equation}
1084: If we retain only relevant and marginal couplings, the
1085: rescaled fermionic propagator with energy dispersion linearized at the
1086: true Fermi surface is given by
1087: \begin{equation}
1088: \tilde{G}_l^{\sigma} ( q , i \epsilon, \alpha ) \approx \frac{1}{
1089: i \epsilon - \alpha \tilde{v}_l^{\sigma} q - r_l^{\sigma}}
1090: \; .
1091: \label{eq:Gscalerel1}
1092: \end{equation}
1093: Apart from $Z_l^{\sigma}$ and $\tilde{v}_l^{\sigma}$,
1094: the third marginal coupling of our model is
1095: the momentum- and frequency-independent part
1096: of the rescaled spin-flip vertex defined in Eq.~(\ref{eq:vertexrescale}),
1097: \begin{equation}
1098: \gamma_l = \tilde{\Gamma}^{(2,1)}_{l} ( 0, \sigma;0, \bar{\sigma} ;0 , \alpha)
1099: \; .
1100: \end{equation}
1101: It satisfies a flow equation of the form
1102: \begin{equation}
1103: \partial_l \gamma_l = - \frac{ \bar{\eta}_l + \eta_l^+ + \eta_l^- }{2}
1104: \gamma_l + \dot{\Gamma}^{(2,1)}_l
1105: \; ,
1106: \label{eq:gammabotflow}
1107: \end{equation}
1108: where $\bar{\eta}_l = - \partial_l \ln \bar{Z}_l$ is the flowing anomalous
1109: dimension of the spin-flip field, and
1110: the inhomogeneity $\dot{\Gamma}^{(2,1)}_l$ depends on the
1111: irrelevant higher interaction vertices involving more
1112: than one external boson leg shown in Fig.~\ref{fig:flowvert}.
1113: In particular,
1114: from the right-hand side of Eq.~(\ref{eq:vertexrescale22})
1115: it is clear that the vertex $\tilde{\Gamma}_l^{(2,2) }$
1116: with two external fermion and two boson legs
1117: is irrelevant with scaling dimension $-1$.
1118: This and the higher order irrelevant vertices vanish at the initial scale $\Lambda_0$
1119: and we shall set them equal to zero, expecting that their effect
1120: can be implicitly taken into account
1121: via a redefinition of the numerical values of the
1122: relevant and marginal couplings \cite{Polchinski84}.
1123: An exception is the vertex $\Gamma^{(0,4)}_{\Lambda}$
1124: involving four external bosonic legs, which
1125: according to Fig.~\ref{fig:flowPi}
1126: drives the flow of the spin-flip susceptibility
1127: in the momentum transfer cutoff scheme.
1128: In contrast to the other irrelevant vertices, the vertex
1129: $\Gamma^{(0,4)}_{\Lambda}$ is finite at the initial
1130: scale $\Lambda = \Lambda_0$, where it
1131: reduces to a symmetrized closed fermion loop \cite{Schuetz05}.
1132: Below we shall propose a simple approximate procedure to take
1133: the renormalization of the spin-flip susceptibility generated by this vertex
1134: into account.
1135: Finally, we note that
1136: the vertex with four external fermionic legs is also marginal, but
1137: in the momentum transfer cutoff scheme it does not directly couple
1138: to the flow of the fermionic self-energy.
1139:
1140:
1141:
1142: \subsection{Defining the Fermi surface within the functional RG}
1143: \label{subsec:FS}
1144:
1145:
1146: The general method to obtain the counter-terms necessary to construct the
1147: true Fermi surface within the framework of the
1148: functional RG has been
1149: developed in Refs.~[\onlinecite{Kopietz01,Ledowski03,Ledowski05}].
1150: Let us briefly recall the main idea.
1151: As long as the flowing anomalous dimensions
1152: $\eta_l^{\sigma}$ of the Fermi fields remains smaller than unity
1153: for $l \rightarrow \infty$, we may
1154: define the true Fermi surface self-consistently from the
1155: requirement that the relevant couplings $r^{ \sigma }_{ l}$ associated with the
1156: fermionic self-energy approach finite limits
1157: for $l \rightarrow \infty$.
1158: This requires fine tuning of the
1159: initial values $r^{\sigma}_0$, which defines a relation between
1160: $r^{\sigma}_0$ and the flowing couplings
1161: on the entire RG trajectory.
1162: In higher dimensions, where the Fermi surface is a continuum, infinitely
1163: many relevant couplings $r_l ( {\bd{k}}_F )$ have to be fine tuned
1164: to define the Fermi surface. In the usual classification of critical fixed points, the
1165: Fermi surface thus corresponds to a multicritical point of infinite order.
1166: Once the proper initial values $r_0^{\sigma}$ are known,
1167: the exact self-energy $\Sigma^{\sigma} ( \alpha k^{\sigma} , i0)$ can
1168: be constructed using Eq.~(\ref{eq:initialr0}),
1169: \begin{equation}
1170: \Sigma^{\sigma} ( \alpha k^{\sigma} , i 0 ) = - \mu^{\sigma}_0 = - v_F \Lambda_0 r_0^{\sigma}
1171: \; .
1172: \end{equation}
1173: The requirement that $r_l^{\sigma}$ flows into a RG fixed point
1174: implies for the initial values\cite{Ledowski03},
1175: \begin{equation}
1176: r_{0}^{ \sigma} = - \int_{0}^{\infty} dl e^{ - ( 1 -
1177: \bar{\eta}^{\sigma}_l ) l }
1178: \dot{{\Gamma}}^{\sigma}_{l } (0)
1179: \; ,
1180: \label{eq:selfcon}
1181: \end{equation}
1182: where
1183: \begin{equation}
1184: \bar{\eta}_l^{\sigma} = \frac{1}{l} \int_0^{l} d t \eta_{t}^{\sigma}
1185: \end{equation}
1186: is the average of the flowing anomalous dimension along the RG trajectory,
1187: and we have written $\dot{{\Gamma}}^{\sigma}_{l } (0) =
1188: \dot{{\Gamma}}^{\sigma}_{l } (0, i0 , \alpha)$ to emphasize
1189: that this quantity is actually independent of the chirality index $\alpha$.
1190: For our effective model with linear energy dispersion we obtain
1191: for the Fermi point distance at constant chemical potential
1192: [see also Eq.~(\ref{eq:FSdef})],
1193: \begin{eqnarray}
1194: \tilde{\Delta} & = & \tilde{\Delta}_0
1195: +
1196: \left[ \frac{r_0^+}{\tilde{v}_0^{+}} - \frac{ r_0^-}{\tilde{v}_0^-} \right]
1197: \nonumber
1198: \\
1199: & & \hspace{-10mm} = \tilde{\Delta}_0
1200: - \sum_{\sigma} \frac{\sigma}{ \tilde{v}_0^{\sigma} }
1201: \int_{0}^{\infty} dl e^{ - ( 1 -
1202: \bar{\eta}^{\sigma}_l ) l }
1203: \dot{{\Gamma}}^{\sigma}_{l } (0, \alpha)
1204: \; ,
1205: \label{eq:FSdef2}
1206: \end{eqnarray}
1207: where we have defined
1208: \begin{equation}
1209: \tilde{\Delta} = \frac{ k^{+} - k^{-}}{\Lambda_0} \; , \;
1210: \tilde{\Delta}_0 = \frac{ k^{+}_0 - k^{-}_0}{\Lambda_0}
1211: \; .
1212: \end{equation}
1213:
1214:
1215: \section{Calculation of the true Fermi surface}
1216: \label{sec:Calculation}
1217:
1218: \subsection{Truncation based on relevance}
1219:
1220: Because a possible confinement transition is expected to be a strong-coupling
1221: phenomenon,
1222: the usual perturbative weak coupling RG \cite{Ledowski05} is not
1223: sufficient.
1224: We therefore propose an alternative truncation scheme
1225: based on the truncation according
1226: to relevance in the RG sense.
1227: In our model we have to keep track of the RG flow of the
1228: two relevant couplings $r_l^{\sigma}$, $\sigma = \pm 1$, and the
1229: marginal couplings $Z_l^{\sigma}$, $\tilde{v}_l^{\sigma}$, and $\gamma_l$.
1230: The couplings $r_l^{\sigma}$, $Z_l^{\sigma}$ and
1231: $\tilde{v}_l^{\sigma}$ associated with the fermionic Green function
1232: satisfy the flow equations
1233: given in Eqs.~(\ref{eq:rlflow}, \ref{eq:Zflow}, \ref{eq:vlflow}).
1234: The function
1235: $ \dot{{\Gamma}}_{ l }^{\sigma} ( Q , \alpha) $ appearing on the
1236: right-hand side of these equations
1237: is in general given in Eq.~(\ref{eq:dotgamma2}); approximating
1238: the fermionic Green function by Eq.~(\ref{eq:Gscalerel1}) and
1239: the spin-flip vertex by its momentum- and frequency independent part
1240: $\gamma_l$, we obtain
1241: \begin{eqnarray}
1242: \dot{{\Gamma}}_{ l }^{\sigma} (q , i \epsilon , \alpha)
1243: & = & \int \frac{ d \bar{q} d \bar{\epsilon} }{( 2 \pi )^2} \delta ( | \bar{q} | -1 )
1244: \nonumber
1245: \\
1246: & & \hspace{-23mm} \times \frac{ \gamma_l^2
1247: [ {\bf{\tilde{F}}}_l^{\sigma \bar{\sigma}} ( \bar{q}, i \bar{\epsilon} ) ]_{ \alpha \alpha }
1248: e^{ i \bar{\epsilon} 0^+} }{
1249: i ( \bar{\epsilon} + \epsilon ) - \alpha \tilde{v}_l^{\bar{\sigma}} (
1250: \bar{q} +q ) - \sigma \tilde{v}^{\bar{\sigma}}_l \tilde{\Delta}^{\ast}_l
1251: -r^{\bar{\sigma}}_l
1252: }
1253: \; .
1254: \label{eq:dotgamma3}
1255: \end{eqnarray}
1256: Here $\tilde{\Delta}_{l}^{\ast}$
1257: is the rescaled true difference between the Fermi
1258: points defined in Eq.~(\ref{eq:deltastardef}), and
1259: the rescaled bosonic spin-flip propagator
1260: ${\bf{\tilde{F}}}_l^{\sigma \bar{\sigma}} ( \bar{Q} )$
1261: is defined in Eq.~(\ref{eq:tildeFpropdef}).
1262:
1263: To calculate the Fermi surface, we need
1264: additional flow equations for the
1265: marginal part of the spin-flip vertex $\gamma_l$ and for the
1266: flowing
1267: spin-flip susceptibility
1268: $\tilde{\Pi}^{ \sigma \bar{\sigma}}_l ( \bar{Q}, \alpha ) $.
1269: As far as $\gamma_l$ is concerned,
1270: we note from Fig.~\ref{fig:flowvert} that the inhomogeneity
1271: $\dot{\Gamma}^{(2,1)}_l$
1272: in Eq.~ (\ref{eq:gammabotflow}) which drives the flow of
1273: $\gamma_l$
1274: involves vertices with two fermionic and more than one bosonic external legs.
1275: These vertices are
1276: irrelevant and vanish at the initial scale $\Lambda_0$, so that it is
1277: reasonable to neglect them. We therefore set $\dot{\Gamma}^{(2,1)}_l = 0$.
1278: We shall also neglect
1279: the bosonic wave-function renormalization, setting
1280: $\bar{Z}_l =1$. In this approximation the flow of the rescaled spin-flip vertex
1281: is driven by the fermionic wave-function renormalization,
1282: \begin{equation}
1283: \partial_l \gamma_l = - \frac{\eta_l^+ + \eta_l^-}{2} \gamma_l
1284: \label{eq:gammalflow}
1285: \; .
1286: \end{equation}
1287:
1288: Before discussing the spin-flip susceptibility
1289: $\tilde{\Pi}^{ \sigma \bar{\sigma}}_l ( \bar{Q}, \alpha ) $, note that
1290: Eqs.~(\ref{eq:vlflow}) and (\ref{eq:dotgamma3}) imply for the Fermi velocity
1291: renormalization factor
1292: \begin{equation}
1293: \partial_l \tilde{v}_l^{\sigma} = - \eta_l^{\sigma} ( \tilde{v}_l^{\sigma}
1294: - \tilde{v}_l^{\bar{\sigma}} )
1295: \; ,
1296: \end{equation}
1297: which yields for the difference
1298: \begin{equation}
1299: \partial_l ( \tilde{v}_l^+ - \tilde{v}_l^-) =
1300: - ( {\eta}_l^+ + \eta_l^- ) ( \tilde{v}_l^+ - \tilde{v}_l^-)
1301: \; .
1302: \label{eq:vdifflow}
1303: \end{equation}
1304: Keeping in mind that
1305: $\eta_l^{\sigma} \geq 0$,
1306: Eq.~(\ref{eq:vdifflow}) implies that
1307: a small initial difference between the Fermi velocities
1308: decreases under renormalization.
1309: Thus, if the initial difference $v_0^+ - v_0^-$ is small
1310: and negligible, it becomes even smaller as we iterate the RG.
1311: Since the flow of the other couplings
1312: is not sensitive to a small difference in the Fermi velocities,
1313: it is consistent to approximate $v_0^{\sigma} \approx v_F$, so that
1314: from now on we shall set $\tilde{v}_l^{\sigma} = 1$.
1315:
1316: To close our system of flow equations, we need an equation
1317: for the
1318: rescaled spin-flip susceptibility
1319: $\tilde{\Pi}_l^{ \sigma \bar{\sigma}} ( \bar{Q} , \alpha )$,
1320: which in turn determines the flow of the spin-flip propagator as given
1321: in Eq.~(\ref{eq:Pisfrescale}).
1322: In the momentum transfer cutoff scheme, the flow
1323: of $\tilde{\Pi}_l^{ \sigma \bar{\sigma}} ( \bar{Q} , \alpha )$
1324: is driven by the one-line irreducible vertex $\Gamma^{(0,4)}_{\Lambda}$
1325: with four external bosonic legs, as shown in Fig.~\ref{fig:flowPi}.
1326: Although this vertex is irrelevant, it is finite at the initial scale
1327: $\Lambda_0$, in contrast to the higher order vertices
1328: that drive the flow of the spin-flip vertex $\gamma_l$ shown
1329: in Fig.~\ref{fig:flowvert}.
1330: It is therefore important to take the renormalizations
1331: of $\tilde{\Pi}_l^{ \sigma \bar{\sigma}} ( \bar{Q} , \alpha )$
1332: due to $\Gamma^{(0,4)}_{\Lambda}$ at least approximately into account.
1333: Guided by the initial condition (\ref{eq:Piinitial2})
1334: for the spin-flip susceptibility, we
1335: propose the following {\it{adiabatic approximation}},
1336: \begin{eqnarray}
1337: \tilde{\Pi}_l^{ \sigma \bar{\sigma}} ( \bar{Q} , \alpha ) & \approx &
1338: \frac{ \gamma_l^2 }{ 2 \pi }
1339: \frac{ \sigma \tilde{\Delta}_{l} + \alpha \bar{q} }{
1340: \sigma \tilde{\Delta}_{l} + \alpha \bar{q}
1341: - i \bar{\epsilon} }
1342: \; ,
1343: \label{eq:Pisfapprox}
1344: \end{eqnarray}
1345: where
1346: \begin{equation}
1347: \tilde{\Delta}_{l} = \tilde{\Delta}_l^{\ast} - ( r_l^{+} - r_l^- )
1348: \; .
1349: \label{eq:tildeDeltaldef}
1350: \end{equation}
1351: Note that
1352: Eq.~(\ref{eq:Pisfapprox}) preserves the
1353: initial form of the spin-flip susceptibility
1354: given in Eq.~(\ref{eq:Piinitial2}), but with the
1355: initial gap $\tilde{\Delta}_0 = ( k_0^{+} - k_0^-)/\Lambda_0$
1356: is replaced by the flowing gap $\tilde{\Delta}_l$ at scale $l$,
1357: and an overall reduction of the amplitude
1358: by the vertex correction $\gamma_l^2$.
1359: Indeed, using Eq.~(\ref{eq:FSdef2}) we find
1360: $\tilde{\Delta}_{ l=0} = \tilde{\Delta}_0 = ( k_0^{+} - k_0^-)/\Lambda_0$,
1361: so that
1362: for $l=0$
1363: we recover from Eq.~(\ref{eq:Pisfapprox}) the rescaled version
1364: of the initial condition (\ref{eq:Piinitial2}).
1365: On the other hand,
1366: using the fact that the $r_l^{\sigma}$ are fine tuned
1367: to reach a finite limit for $l \rightarrow \infty$, we see that
1368: $\Delta_l \rightarrow \Delta^{\ast}_l = e^l (k^+ - k^- ) / \Lambda_0 = e^l \tilde{\Delta}$
1369: for $l \rightarrow \infty$.
1370: % We may therefore define the flowing Fermi point difference
1371: % $k_l^{+} - k_l^-$ via
1372: % \begin{equation}
1373: % \frac{ k_l^{+} - k_l^-}{\Lambda} = \tilde{\Delta}_l =
1374: % \frac{ k^{+} - k^-}{\Lambda} -( r^+_l - r^-_l )
1375: % \; .
1376: % \end{equation}
1377: A justification for the adiabatic
1378: approximation
1379: (\ref{eq:Pisfapprox}) is given in the Appendix.
1380:
1381:
1382: To simplify the integrals, it is convenient
1383: to slightly modify the denominator in the expression for
1384: $\dot{{\Gamma}}_{ l }^{\sigma} (q , i \epsilon , \alpha)$
1385: in Eq.~(\ref{eq:dotgamma3}),
1386: \begin{equation}
1387: \sigma \tilde{\Delta}_l^{\ast} + r_l^{\bar{\sigma}} =
1388: \sigma \tilde{\Delta}_l + r_l^{{\sigma}}
1389: \approx \sigma \tilde{\Delta}_l
1390: \; .
1391: \label{eq:rneglect}
1392: \end{equation}
1393: We have checked numerically
1394: from the solution of the resulting equations that
1395: this approximation is self-consistent
1396: by verifying the neglected term $r_l^{\sigma}$ is indeed small.
1397: We thus arrive at the following approximation for the
1398: inhomogeneity $ \dot{{\Gamma}}_{ l }^{\sigma} (q , i \epsilon , \alpha)$
1399: that controls the flow of the fermionic self-energy,
1400: \begin{eqnarray}
1401: \dot{{\Gamma}}_{ l }^{\sigma} (q , i \epsilon , \alpha)
1402: & = & \int \frac{ d \bar{q} d \bar{\epsilon} }{( 2 \pi )^2} \delta ( | \bar{q} | -1 )
1403: \nonumber
1404: \\
1405: & & \hspace{-23mm} \times \frac{ \gamma_l^2
1406: [ {\bf{\tilde{F}}}_l^{\sigma \bar{\sigma}} ( \bar{q}, i \bar{\epsilon} ) ]_{ \alpha \alpha }
1407: e^{ i \bar{\epsilon} 0^+} }{
1408: i ( \bar{\epsilon} + \epsilon ) - \alpha (
1409: \bar{q} +q ) - \sigma \tilde{\Delta}_l
1410: }
1411: \; .
1412: \label{eq:dotgamma4}
1413: \end{eqnarray}
1414: To be consistent the approximation (\ref{eq:rneglect})
1415: we should also neglect the
1416: flowing anomalous dimension $\eta_l^{\sigma}$ in
1417: the flow equation (\ref{eq:rlflow}) for $r_l^{\sigma}$, because
1418: an expansion of Eq.~(\ref{eq:dotgamma3})
1419: in powers of $r_l^{\sigma}$ leads to a cancellation
1420: of the term $\eta_l^{\sigma} r_l^{\sigma}$.
1421: The flow equation for $r_l^{\sigma}$ then reduces to
1422: \begin{equation}
1423: \partial_l r_{ l }^{\sigma} = r_{ l }^{\sigma}
1424: + \dot{{\Gamma}}_{l }^{\sigma} (0)
1425: \; ,
1426: \label{eq:rflow4}
1427: \end{equation}
1428: where
1429: \begin{eqnarray}
1430: \dot{{\Gamma}}_{l }^{\sigma} ( 0)
1431: & = & \int \frac{ d \bar{q} d \bar{\epsilon} }{( 2 \pi )^2} \delta ( | \bar{q} | -1 )
1432: \nonumber
1433: \\
1434: & \times & \frac{ \gamma_l^2
1435: [ {\bf{\tilde{F}}}_l^{\sigma \bar{\sigma}} ( \bar{q}, i \bar{\epsilon} ) ]_{ \alpha \alpha }
1436: e^{ i \bar{\epsilon} 0^+} }{
1437: i \bar{\epsilon} - \alpha
1438: \bar{q} - \sigma \tilde{\Delta}_l
1439: }
1440: \; .
1441: \label{eq:dotgamma04}
1442: \end{eqnarray}
1443: Our general self-consistency equation (\ref{eq:FSdef2})
1444: for the true Fermi point distance
1445: can then we written as an integral involving
1446: the flow of the couplings on the entire RG trajectory,
1447: \begin{eqnarray}
1448: \tilde{\Delta} & = & \tilde{\Delta}_0
1449: - \int_{0}^{\infty} dl e^{-l}
1450: \sum_{\sigma} \sigma
1451: \dot{{\Gamma}}^{\sigma}_{l } (0)
1452: \; .
1453: \label{eq:FSapprox1}
1454: \end{eqnarray}
1455: Anticipating that within our approximations $\eta_l^{\sigma}$ is
1456: independent of $\sigma$, we may write
1457: $\eta^{\sigma}_l = \eta_l$.
1458: The flow equation (\ref{eq:gammalflow})
1459: for the spin-flip vertex then reduces to
1460: \begin{equation}
1461: \partial_l \gamma_l = - \eta_l \gamma_l
1462: \; ,
1463: \label{eq:gammabotflow2}
1464: \end{equation}
1465: where from Eqs.~(\ref{eq:etadef}) and (\ref{eq:dotgamma4})
1466: we find
1467: \begin{equation}
1468: \eta_l =
1469: \int \frac{ d \bar{q} d \bar{\epsilon} }{( 2 \pi )^2} \delta ( | \bar{q} | -1 ) \frac{ \gamma_l^2
1470: [ {\bf{\tilde{F}}}_l^{\sigma \bar{\sigma}} ( \bar{q}, i \bar{\epsilon} ) ]_{ \alpha \alpha }
1471: }{ [
1472: i \bar{\epsilon} - \alpha \bar{q} - \sigma \tilde{\Delta}_l
1473: ]^2 }
1474: \; .
1475: \label{eq:etaapprox2}
1476: \end{equation}
1477: We thus arrive at
1478: a closed system of flow equations
1479: for the two relevant couplings $r_l^{+}$ and $r_l^{-}$ and the two marginal
1480: couplings $g_{n,l}$ and $g_{c,l}$.
1481: We emphasize that our truncation does not rely
1482: on a weak coupling expansion, which enables us to
1483: study a possible confinement transition.
1484:
1485: To give an explicit expression for the
1486: bosonic spin-flip propagator,
1487: we define dimensionless bare couplings
1488: (keeping in mind that we have chosen $\nu_0 = 1/v_F$),
1489: \begin{equation}
1490: 2 \nu_0 J^{\bot}_{ \alpha \alpha} = 2 \pi g_{c,0}
1491: \; , \;
1492: 2 \nu_0 J^{\bot}_{ \alpha, - \alpha} = 2 \pi g_{n,0}
1493: \; ,
1494: \end{equation}
1495: and the flowing couplings
1496: \begin{equation}
1497: g_{c,l} = \gamma_l^2 g_{c,0}
1498: \; , \;
1499: g_{n,l} = \gamma_l^2 g_{n,0}
1500: \label{eq:gnl}
1501: \; ,
1502: \end{equation}
1503: which according to Eq.(\ref{eq:gammabotflow2}) satisfy the flow equations
1504: \begin{equation}
1505: \partial_l g_{c,l} = - 2 \eta_l g_{c,l }
1506: \; , \;
1507: \partial_l g_{n,l} = - 2 \eta_l g_{n,l }
1508: \; .
1509: \label{eq:gflow}
1510: \end{equation}
1511: The rescaled spin-flip propagator can then be written as
1512: \begin{widetext}
1513: \begin{eqnarray}
1514: \gamma_l^2
1515: [ {\bf{\tilde{F}}}_l^{\sigma \bar{\sigma}} ( \bar{q}, i \bar{\epsilon} ) ]_{ \alpha \alpha }
1516: & = &
1517: \frac{ \gamma_l^2 2 \pi [ g_{c,0} -
1518: (g_{c,0}^2 - g_{n,0}^2) 2 \pi
1519: \tilde{\Pi}_l^{\sigma \bar{\sigma}} ( \bar{Q} , -\alpha )]}{1 - g_{c,0}
1520: 2 \pi [ \tilde{\Pi}_l^{\sigma \bar{\sigma}} ( \bar{Q} ,+ )
1521: + \tilde{\Pi}_l^{\sigma \bar{\sigma}} ( \bar{Q} , - ) ]
1522: + (g_{c,0}^2 - g_{n,0}^2)
1523: (2 \pi )^2 \tilde{\Pi}_l^{\sigma \bar{\sigma}} ( \bar{Q} , + )
1524: \tilde{\Pi}_l^{\sigma \bar{\sigma}} ( \bar{Q} , - )
1525: }
1526: \nonumber
1527: % \\
1528: % & = & 2 \pi
1529: % \frac{ g_{c,l}
1530: % [ ( i \bar{\epsilon} - \sigma \tilde{\Delta}_l )^2 - \bar{q}^2 ]
1531: % + (g_{c,l}^2 - g_{n,l}^2) [ \tilde{\Delta}_l^2 - \bar{q}^2 - i \bar{\epsilon}
1532: % ( \sigma \tilde{\Delta}_l - \alpha \bar{q} )] }{
1533: % [ i \bar{\epsilon} - \omega^+_l ( \bar{q}, \sigma \tilde{\Delta}_l ) ]
1534: % [ i \bar{\epsilon} - \omega^-_l ( \bar{q} , \sigma \tilde{\Delta}_l ) ]
1535: % }
1536: % \nonumber
1537: \\ & = &
1538: 2 \pi
1539: ( i \bar{\epsilon} - \alpha \bar{q} - \sigma \tilde{\Delta}_l )
1540: \frac{ g_{c,l}
1541: ( i \bar{\epsilon} + \alpha \bar{q} - \sigma \tilde{\Delta}_l )
1542: + (g_{c,l}^2 - g_{n,l}^2) ( \alpha \bar{q} - \sigma \tilde{\Delta}_l )
1543: }{
1544: [ i \bar{\epsilon} - \omega^+_l ( \bar{q}, \sigma \tilde{\Delta}_l ) ]
1545: [ i \bar{\epsilon} - \omega^-_l ( \bar{q} , \sigma \tilde{\Delta}_l ) ]
1546: }
1547: \; ,
1548: \end{eqnarray}
1549: where
1550: \begin{eqnarray}
1551: \omega^{\pm}_l ( \bar{q} , x ) & = & x ( 1 - g_{c,l} ) \pm
1552: \sqrt{ x^2 g_{n,l}^2 + \bar{q}^2 [ ( 1 - g_{c,l} )^2 - g_{n,l}^2 ] }
1553: \nonumber
1554: \\
1555: & = &
1556: x ( 1 - g_{c,l} ) \pm
1557: \sqrt{ x^2 ( 1 - g_{c,l} )^2 + ( \bar{q}^2 - x^2) [ ( 1 - g_{c,l})^2 -
1558: g_{n,l}^2 ] }
1559: \; .
1560: \label{eq:omegapmdef}
1561: \end{eqnarray}
1562: Noting that $\omega^{\pm}_l ( 0,x ) = x ( 1 - g_{c,l} \pm g_{n,l} )$,
1563: we see that
1564: for small interaction strength
1565: both modes $\omega^+_l$ and $\omega^-_l$ are gapped.
1566: However, the gap of the mode $\omega^{-}_l ( 0,x )$ vanishes for
1567: $ g_{c,l} + g_{n,l} =1$, signaling a possible quantum phase transition
1568: to a confined state.
1569: In the present work we do not attempt to extend the RG beyond this
1570: point, but focus
1571: on the regime $ g_{c,l} + g_{n,l} \leq 1$ were both modes are gapped.
1572: The integrals in Eqs.~(\ref{eq:dotgamma04}) and (\ref{eq:etaapprox2})
1573: can be carried out analytically using the residue theorem, with the result
1574: \begin{eqnarray}
1575: \dot{\Gamma}_l^{\sigma} ( 0 ) & = &
1576: 2 \delta_{ \sigma ,-1} \Theta( \tilde{\Delta}_l -1 ) g_{c,l}
1577: + \Theta (1- \tilde{\Delta}_l)
1578: % \nonumber
1579: % \\
1580: %& & \hspace{-13mm}
1581: %\times
1582: \left[ g_{c,l} + \frac{ \sigma \tilde{\Delta}_l g_{n,l}^2}{ \sqrt{
1583: (1-g_{c,l})^2 - g_{n,l}^2 (1 - \tilde{\Delta}_l^2 ) } }
1584: \right]
1585: \; ,
1586: \end{eqnarray}
1587: and
1588: \begin{eqnarray}
1589: \eta_l & = &
1590: \frac{ \Theta ( 1 - \tilde{\Delta}_l )
1591: g_{n,l}^2}{ \sqrt{ (1-g_{c,l})^2 - g_{n,l}^2 (1 - \tilde{\Delta}_l^2 ) }
1592: \left[ 1- g_{c,l} + \sqrt{ (1-g_{c,l})^2 - g_{n,l}^2 (1 - \tilde{\Delta}_l^2 ) } \right]}
1593: \; .
1594: \label{eq:etares1}
1595: \end{eqnarray}
1596: \end{widetext}
1597:
1598:
1599:
1600: \subsection{Self-consistent one-loop approximation}
1601: \label{subsec:relevant}
1602:
1603: The above system of coupled equations can only be solved numerically.
1604: However, if we neglect the flow of the coupling constants on the
1605: right-hand sides of these equations, we can obtain an approximate analytical solution,
1606: which is useful to get a rough idea about the mechanism responsible
1607: for the confinement transition. In this subsection we therefore
1608: set
1609: \begin{eqnarray}
1610: g_{ i, l} & \approx & g_{i ,0}
1611: \; ,
1612: \label{eq:gapprox}
1613: \\
1614: \tilde{\Delta}_l & \approx & \tilde{\Delta}_l^{\ast} \equiv
1615: \tilde{\Delta} e^l
1616: \; .
1617: \label{eq:Deltaapprox}
1618: \end{eqnarray}
1619: We expect that these approximations over-estimate
1620: the tendency towards confinement, because we
1621: know from Eq.~(\ref{eq:gflow}) that the flowing couplings $g_{c,l}$ and $g_{n,l}$
1622: are smaller than the bare ones.
1623: The second approximation (\ref{eq:Deltaapprox})
1624: is justified provided the trajectory integral
1625: (\ref{eq:FSapprox1}) is dominated by $l \gtrsim 1$
1626: where $\tilde{\Delta}_l^{\ast}
1627: \gg | r_l^{\sigma} |$.
1628: Substituting $x =
1629: \tilde{\Delta} e^{l}$ on the right-hand side of
1630: Eq.~(\ref{eq:FSapprox1}) we obtain
1631: \begin{eqnarray}
1632: \tilde{\Delta}
1633: & = & \tilde{\Delta}_0 - \tilde{\Delta}
1634: \int_{\tilde{\Delta}}^{\infty} dx I ( x )
1635: \; ,
1636: \label{eq:FSapprox2}
1637: \end{eqnarray}
1638: with
1639: \begin{eqnarray}
1640: I ( x ) & =& - \Theta ( x-1 ) \frac{2 g_{c,0}}{x^2}
1641: \nonumber
1642: \\
1643: & & \hspace{-20mm} + \Theta ( 1-x ) \frac{ 2 g_{n,0}^2}{x
1644: \sqrt{
1645: (1-g_{c,0})^2 - g_{n,0}^2 (1- x^2) }
1646: }
1647: \; .
1648: \end{eqnarray}
1649: The $x$-integration is elementary and
1650: we finally obtain
1651: \begin{equation}
1652: \tilde{\Delta} = \frac{ \tilde{\Delta}_0 }{ 1 + R ( \tilde{\Delta} ) }
1653: \; ,
1654: \label{eq:Rdef}
1655: \end{equation}
1656: with
1657: \begin{eqnarray}
1658: R ( \tilde{\Delta} ) & = &
1659: \int_{ {\tilde{\Delta}} }^{\infty} dx I ( x )
1660: = - 2 g_{c,0}
1661: \nonumber
1662: \\
1663: & & \hspace{-15mm} + \frac{ 2 g_{n,0}^2}{ \sqrt{ (1 - g_{c,0}) ^2 - g_{n,0}^2} }
1664: \ln \left[ \frac{1 + \sqrt{ 1 + \frac{ \tilde{\Delta}^2 g_{n,0}^2}{ (1 - g_{c,0})^2 -g_{n,0}^2}}}{
1665: \tilde{\Delta}
1666: \left( 1 + \sqrt{ \frac{ ( 1 - g_{c,0})^2}{ (1 - g_{c,0})^2 - g_{n,0}^2}} \right) } \right]
1667: \; .
1668: \nonumber
1669: \\
1670: & &
1671: \label{eq:Rres1}
1672: \end{eqnarray}
1673: This expression diverges for $g_{c,0} + g_{n,0} \rightarrow 1$,
1674: corresponding
1675: to a confinement transition with $\tilde{\Delta} \rightarrow 0$.
1676: Expanding
1677: $R ( \tilde{\Delta} )$ to second order in the couplings we obtain
1678: \begin{equation}
1679: R ( \tilde{\Delta} ) =
1680: - 2 g_{c,0} + 2 g_{n,0}^2 \ln (1 / \tilde{\Delta} ) + O (g_{i,0}^3 )
1681: \; .
1682: \label{eq:Rexpand}
1683: \end{equation}
1684: Keeping in mind that in Ref.~[\onlinecite{Ledowski05}]
1685: we have neglected the chiral couplings and that here we have retained only
1686: interchain backscattering,
1687: Eq.~(\ref{eq:Rexpand}) is consistent with
1688: our previous weak-coupling result given in
1689: Eq.~(\ref{eq:deltaprevious}).
1690: From Eq.~(\ref{eq:Rexpand}) it is clear that
1691: there is a competition between the chiral part $g_{c,0}$ and the
1692: non-chiral part $g_{n,0}$ of the interaction.
1693: The chiral part $g_{c,0}$
1694: leads to a repulsion of the Fermi points, while the non-chiral part
1695: $g_{n,0}$ generates an attraction and eventually triggers the confinement
1696: transition, because
1697: for sufficiently small $\tilde{\Delta}$ the logarithm
1698: overwhelms the term linear in $g_{c,0}$.
1699:
1700: To simplify the following analysis we shall from now on restrict ourselves
1701: to the special case $g_{c,0} =0$. This is sufficient
1702: to study the confinement transition, which is driven by the
1703: non-chiral part of the interaction.
1704: Setting $g_0 = g_{ n , 0}$ the confinement
1705: transition occurs within the approximations in this section
1706: at $g_0 =1$.
1707: A numerical solution of Eqs.~(\ref{eq:Rdef}) and (\ref{eq:Rres1}) for the true
1708: $\tilde{\Delta}$ as function of $g_0$ is shown
1709: in Fig.~\ref{fig:deltachiral}.
1710: %
1711: %
1712: \begin{figure}[tb]
1713: \centering
1714: % \psfrag{t1}{$T >T_c$}
1715: % \vspace{7mm}
1716: % \includegraphics{fig5.eps}
1717: \epsfig{file=fig6.eps,width=75mm}
1718: % \vspace{4mm}
1719: \caption{%
1720: Numerical solution of Eq.~(\ref{eq:Rdef}) for $g_{c,0} =0$ as function of
1721: $g_0 = g_{n,0}$ for different values of $\tilde{\Delta}_0 = ( k^{+}_0 - k^{-}_0 )/\Lambda_0$.
1722: }
1723: \label{fig:deltachiral}
1724: \end{figure}
1725: %
1726: %
1727: The confinement transition for $g_0 \rightarrow 1$ is clearly visible.
1728: In fact, the behavior of $\tilde{\Delta}$ for $g_0 \rightarrow 1$ can
1729: be obtained analytically.
1730: In this case
1731: $\tilde{\Delta} \ll \sqrt{ 1 - g_0^2}$, so that we may approximate
1732: \begin{eqnarray}
1733: R ( \tilde{\Delta} ) & \approx &
1734: \frac{2}{ \sqrt{ 1 - g_0^2 } }
1735: \ln \left[ \frac{2 \sqrt{ 1 - g_0^2} }{ \tilde{\Delta} }\right]
1736: \; .
1737: \label{eq:Rdiv}
1738: \end{eqnarray}
1739: The self-consistency condition (\ref{eq:Rdef}) for $\tilde{\Delta}$
1740: then reduces to
1741: \begin{equation}
1742: \tilde{\Delta} \approx
1743: \sqrt{ 1 - g_0^2}
1744: \left[ \frac{ \tilde{\Delta}_0}{ 2 \ln
1745: \left( \frac{ 2 \sqrt{1 - g_0^2} }{ {\tilde{\Delta}} } \right) } \right]
1746: \; .
1747: \label{eq:confinement}
1748: \end{equation}
1749: For $\tilde{\Delta} \ll \sqrt{ 1 - g_0^2}$ the second factor in the square braces
1750: of Eq.~(\ref{eq:confinement}) is small compared with unity, so that
1751: it is consistent to take the limit $g_0 \rightarrow 1$ in this expression.
1752: If we identify $\Delta$ with the order parameter of the confinement transition
1753: transition (with $\Delta \neq 0$ corresponding to the deconfined
1754: phase), then Eq.~(\ref{eq:confinement}) predicts mean-field behavior with
1755: logarithmic corrections.
1756:
1757:
1758:
1759:
1760:
1761:
1762:
1763:
1764:
1765: Our simple one-loop approximation thus predicts
1766: that for $g_0 \rightarrow 1$
1767: there is a confinement transition where the true
1768: Fermi point distance $\Delta$ collapses, corresponding to vanishing effective
1769: interchain hoping $t_{\bot}^{\ast} =0$.
1770: In pseudo-spin language,
1771: the quantum critical point $g_0=1$ corresponds
1772: to a vanishing magnetization in $z$-direction, in spite of the
1773: fact that there is a uniform magnetic field.
1774: For $g_0 > 1$ our one-loop approximation
1775: suggests that there is long-range ferromagnetic order in
1776: $xy$-direction. However, in one dimension we do not expect true long-range order,
1777: so that fluctuations beyond the one-loop approximation should be important.
1778: It is therefore important to go beyond this approximation, which we shall do
1779: in the following subsection.
1780:
1781:
1782: \subsection{Including the renormalization of the
1783: effective interaction}
1784: \label{subsec:including}
1785:
1786:
1787: We now improve the above calculation by taking the flow of the
1788: effective interaction into account.
1789: For simplicity, we focus again on the special case without
1790: chiral interactions, so that we need to keep track only
1791: of the flowing non-chiral interaction $g_{ n , l} = g_l$.
1792: Furthermore, in the absence of chiral couplings $r_l^{\sigma} = \sigma r_l$.
1793: The self-consistency equation (\ref{eq:FSapprox1})
1794: for the Fermi point distance then reduces to
1795: \begin{equation}
1796: \tilde{\Delta} = \tilde{\Delta}_0 -
1797: \int_0^{\infty} dl e^{-l}
1798: \frac{ 2 \Theta ( 1 - \tilde{\Delta}_l ) \tilde{\Delta}_l g_l^2 }{ \sqrt{ 1 - g_l^2
1799: ( 1 - \tilde{\Delta}_l )^2 } }
1800: \; ,
1801: \label{eq:DeltaDelta}
1802: \end{equation}
1803: where
1804: \begin{equation}
1805: \tilde{\Delta}_l = \tilde{\Delta}_l^{\ast} - 2 r_l =
1806: \tilde{\Delta} e^l - 2 r_l
1807: \label{eq:tildedeltalr}
1808: \; ,
1809: \end{equation}
1810: and the flow of $r_l$ and $g_l$ is determined by
1811: \begin{eqnarray}
1812: \partial_l r_l & = & r_l + A ( g_l, \tilde{\Delta}_l )
1813: \; ,
1814: \label{eq:Arl}
1815: \\
1816: \partial_l g_l & = & B ( g_l , \tilde{\Delta}_l )
1817: \; ,
1818: \label{eq:Brl}
1819: \end{eqnarray}
1820: with
1821: \begin{equation}
1822: A ( g_l , \tilde{\Delta}_l ) =
1823: \frac{ \Theta (1- \tilde{\Delta}_l) \tilde{\Delta}_l g_{l}^2}{ \sqrt{
1824: 1 - g_{l}^2 (1 - \tilde{\Delta}_l^2 ) } }
1825: \; ,
1826: \label{eq:Adef}
1827: \end{equation}
1828: and
1829: \begin{eqnarray}
1830: B ( g_l , \tilde{\Delta}_l ) & = & - 2 \eta_l g_l
1831: \nonumber
1832: \\
1833: & & \hspace{-25mm}
1834: =
1835: \frac{ - 2 \Theta ( 1 - \tilde{\Delta}_l )
1836: g_{l}^3}{ \sqrt{ 1 - g_{l}^2 (1 - \tilde{\Delta}_l^2 ) }
1837: \left[ 1 + \sqrt{ 1 - g_{l}^2 (1 - \tilde{\Delta}_l^2 ) } \right]}
1838: \; .
1839: \label{eq:Bdef}
1840: \end{eqnarray}
1841: The initial value $r_0$ has to be fine tuned such that
1842: $\lim_{ l \rightarrow \infty} r_l $ remains finite. This leads to the
1843: self-consistency equation (\ref{eq:DeltaDelta}) for the true Fermi point distance.
1844: We emphasize again that
1845: our approximation scheme is not based on a weak coupling expansion,
1846: so that the $\beta$-function given in
1847: Eq.~(\ref{eq:Bdef}) is non-perturbative in the coupling $g_l$.
1848: Instead of Eq.~(\ref{eq:Rres1}) we now obtain for the dimensionless renormalization
1849: factor $R ( \tilde{\Delta} )$ defined in Eq.~(\ref{eq:Rdef}),
1850: \begin{eqnarray}
1851: R ( \tilde{\Delta} ) & = & \frac{2}{\tilde{\Delta}} \int_0^{\infty} d l e^{ - l }
1852: A ( g_l , \tilde{\Delta}_l )
1853: \nonumber
1854: \\
1855: & & \hspace{-17mm} = 2
1856: \int_0^{\infty} d l
1857: ( 1 - 2 e^{-l} r_l / \tilde{\Delta} )
1858: \frac{ \Theta ( 1 - \tilde{\Delta}_l ) g_l^2 }{
1859: \sqrt{ 1 - g_l^2 ( 1 - \tilde{\Delta}_l^2 )} }
1860: \; .
1861: \label{eq:Rres2}
1862: \end{eqnarray}
1863: Note that formally the
1864: confinement transition
1865: manifests itself via a divergence of
1866: the function $R (\tilde{\Delta} )$
1867: for $\tilde{\Delta} \rightarrow 0$.
1868: However, as shown in Fig.~\ref{fig:Rres2},
1869: the renormalization factor remains now finite so that
1870: also the self-consistent $\tilde{\Delta}$
1871: is finite for $g_0 =1$, as shown in Fig.~\ref{fig:deltatrueg0}.
1872: %
1873: %
1874: \begin{figure}[tb]
1875: \centering
1876: % \psfrag{t1}{$T >T_c$}
1877: % \vspace{7mm}
1878: %\includegraphics{fig6.eps}
1879: \epsfig{file=fig7.eps,width=75mm}
1880: % \vspace{4mm}
1881: \caption{%
1882: Numerical solution of the renormalization factor $R ( \tilde{\Delta} )$
1883: defined in Eq.~(\ref{eq:Rres2})
1884: as a function of the bare coupling $g_0$ for
1885: different values of $\tilde{\Delta}_0$.
1886: }
1887: \label{fig:Rres2}
1888: \end{figure}
1889: %
1890: %
1891: \begin{figure}[tb]
1892: \centering
1893: % \psfrag{t1}{$T >T_c$}
1894: % \vspace{7mm}
1895: \epsfig{file=fig8.eps,width=75mm}
1896: % \vspace{4mm}
1897: \caption{%
1898: Numerical solution of the true Fermi point distance $\tilde{\Delta}
1899: = ( k^{+} - k^- )/\Lambda_0$
1900: as a function of the bare coupling $g_0$ for
1901: different values of the bare distance $\tilde{\Delta}_0 = ( k^{+}_0 - k^-_0 )/\Lambda_0 $.
1902: }
1903: \label{fig:deltatrueg0}
1904: \end{figure}
1905: %
1906: %
1907: We conclude that
1908: the confinement transition obtained in
1909: the previous subsection is an artefact of
1910: the approximations (\ref{eq:gapprox}) and (\ref{eq:Deltaapprox}).
1911:
1912: Let us take a closer look at the point $g_0 =1$ where the one-loop
1913: approximation
1914: predicts a confinement transition.
1915: The RG flow of the couplings $g_l$ and $r_l$ as well as the flowing
1916: anomalous dimension $\eta_l$ for this case is shown in Fig.~\ref{fig:RGflow}.
1917: %
1918: %
1919: \begin{figure}[tb]
1920: \centering
1921: % \psfrag{t1}{$T >T_c$}
1922: % \vspace{7mm}
1923: \epsfig{file=fig9.eps,width=80mm}
1924: % \vspace{4mm}
1925: \caption{%
1926: RG flow of the couplings $g_l$, $r_l$ and the anomalous dimension $\eta_l$
1927: as a function of the logarithmic flow parameter $l$ for
1928: $\tilde{\Delta}_0 =0.1$ and $g_0 =1$.
1929: }
1930: \label{fig:RGflow}
1931: \end{figure}
1932: %
1933: %
1934: One clearly sees that the initially large flowing anomalous dimension
1935: drives the running coupling $g_l$ towards smaller values;
1936: eventually $g_l$ approaches a finite limit for large $l$. Furthermore, the running
1937: coupling $r_l$
1938: approaches its asymptotic value for large $l$ non-monotonously.
1939: The true Fermi point distance $\Delta$ as a function of the
1940: bare one for $g_0 =1$ is shown in Fig.~\ref{fig:deltatrue_g1}.
1941: %
1942: %
1943: \begin{figure}[tb]
1944: \centering
1945: % \psfrag{t1}{$T >T_c$}
1946: % \vspace{7mm}
1947: \epsfig{file=fig10.eps,width=75mm}
1948: % \vspace{4mm}
1949: \caption{%
1950: Numerical solution of the true Fermi point distance $\tilde{\Delta}$
1951: for $g_0=1$ as a function of the bare distance $\tilde{\Delta}_0$.
1952: }
1953: \label{fig:deltatrue_g1}
1954: \end{figure}
1955: We see that large interchain backscattering
1956: strongly reduces the Fermi point distance, although the Fermi surface never collapses,
1957: in agreement with scenario suggested by the weak coupling analysis \cite{Ledowski05}.
1958:
1959:
1960: \section{Conclusions}
1961: \label{sec:conclusion}
1962:
1963: In this work we have used a functional RG approach
1964: to calculate self-consistently the true distance
1965: $\Delta = k^{+} - k^{-}$ between the Fermi points
1966: of the bonding and the antibonding band in
1967: a system consisting of two chains of spinless fermions
1968: connected by weak interchain hopping $t_{\bot}$.
1969: Using the insight from our earlier weak coupling analysis \cite{Ledowski05}
1970: that the renormalization of the Fermi surface is essentially
1971: determined by interchain backscattering,
1972: we have treated this scattering
1973: process non-perturbatively by representing it in terms of
1974: a collective bosonic field $\chi$.
1975: In pseudospin language,
1976: where $t_{\bot} = h$ corresponds to a uniform magnetic
1977: field in $z$-direction and
1978: the interchain backscattering interaction
1979: corresponds to a ferromagnetic $xy$-interaction,
1980: the field $\chi$ can be viewed as a fluctuating transverse magnetic field,
1981: which competes with the uniform field $h$ in $z$-direction.
1982: A self-consistent one-loop approximation
1983: predicts that for sufficiently strong interchain backscattering
1984: there is indeed a quantum critical point were
1985: the renormalized Fermi point distance $\Delta \propto t_{\bot}^{\ast} $ vanishes.
1986: However, a more accurate calculation taking vertex corrections
1987: and wave-function renormalizations into account
1988: shows that the renormalized $\Delta \propto t_{\bot}^{\ast}$
1989: remains finite.
1990: This is in agreement with the expectation that a ferromagnetic $xy$-interaction
1991: in a one-dimensional itinerant electron gas cannot
1992: give rise to long-range ferromagnetic order.
1993: Previous studies of the spinless two-chain problem \cite{Bourbonnais91,Caron02}
1994: came to the conclusion that the system exhibits a confinement transition
1995: if the anomalous dimension $\eta_0$ of the Luttinger liquid for $t_{\bot}=0$
1996: is unity. The important difference between these earlier works and our calculations is that
1997: we have completely neglected pair tunneling. In a subsequent article~\cite{Ledowski06}
1998: we shall show how
1999: the inclusion of this process stabilizes again the flow of the interchain backward
2000: scattering and enhances the tendency towards confinement.
2001: In the same article we shall
2002: also show that our approach can be
2003: generalized to study the
2004: more interesting and physically more relevant
2005: confinement problem in an infinite array of
2006: weakly coupled metallic chains. In this case
2007: the Fermi surface consists of two disconnected sheets, which
2008: self-consistently develop completely flat sectors
2009: at the confinement transition. We have preliminary evidence
2010: that in this case there exists a confined phase
2011: where the renormalized interchain hopping vanishes.
2012: The essential scattering process driving this transition is
2013: the non-chiral part of the density-density interaction which transfers momentum
2014: within a given sheet of the Fermi surface.
2015:
2016:
2017: Finally we point out that this work
2018: describes also some technical progress:
2019: we have been able to
2020: find a sensible extrapolation of the weak coupling
2021: functional RG approach to the strong coupling regime.
2022: Our truncation strategy of the formally exact hierarchy
2023: of functional RG equations relies on the classification
2024: of the vertices according to their relevance.
2025: An obvious disadvantage of this approach is that
2026: we cannot give reliable error estimates, which
2027: is a common feature of most truncations of the
2028: coupled functional RG flow equations for the vertex functions.
2029: Note, however, that a similar truncation of the
2030: functional RG equations for the interacting Bose gas
2031: gave quite accurate results for the shift in the critical temperature\cite{Ledowski04}.
2032: In the present
2033: problem we know a priori from the weak coupling analysis that
2034: the physics is dominated by
2035: a single scattering channel, the inter-chain backscattering.
2036: However, an extension of our approach
2037: to problems where several scattering
2038: channels compete seems to be possible.
2039:
2040:
2041:
2042:
2043:
2044: \section*{ACKNOWLEDGMENTS}
2045: We thank Florian Sch\"{u}tz for sharing his insights
2046: on the subtleties of the collective field functional
2047: RG approach with us.
2048:
2049:
2050: \appendix
2051: %\setcounter{section}{0}
2052: %\setcounter{equation}{0}
2053:
2054:
2055: %\renewcommand{\theequation}{\Roman{section}.\arabic{equation}}
2056: \renewcommand{\theequation}{A.\arabic{equation}}
2057: \renewcommand{\thesubsection}{A.\arabic{subsection}}
2058:
2059:
2060: \section*{Appendix: Justification of the adiabatic approximation}
2061:
2062: We give here a justification for the adiabatic approximation
2063: for the rescaled spin-flip
2064: susceptibility given in Eq.~(\ref{eq:Pisfapprox}).
2065: % The perhaps simplest way to arrive at this expression
2066: % is based on the
2067: % Dyson-Schwinger
2068: % equation~\cite{Schuetz05,ZinnJustin02}
2069: % for the spin-flip susceptibility, which is shown diagrammatically
2070: % in Fig.~\ref{fig:DS}.
2071: % %
2072: % %
2073: % \begin{figure}[tb]
2074: % \centering
2075: % % \psfrag{t1}{$T >T_c$}
2076: % % \vspace{7mm}
2077: % \epsfig{file=fig10.eps,width=60mm}
2078: % % \vspace{4mm}
2079: % \caption{%
2080: % Dyson-Schwinger equation for the spin-flip susceptibility.
2081: % }
2082: % \label{fig:DS}
2083: % \end{figure}
2084: % %
2085: % %
2086: % Replacing in the Dyson-Schwinger equation
2087: % the spin-flip vertex by unity,
2088: % approximating the rescaled Green functions by Eq.~(\ref{eq:Gscalerel1}),
2089: % and using the fact that within our approximations
2090: % $Z_l^{+} Z_l^{-} = \gamma_l^2$,
2091: % we arrive at Eq.~(\ref{eq:Pisfapprox}).
2092: Let us therefore use a more general two-cutoff procedure
2093: where we impose a
2094: band-width cutoff $\Lambda^F_{\tau} = \Lambda_{0}^F e^{- \tau }$
2095: on the fermionic propagator in addition to
2096: the bosonic momentum transfer cutoff $ \Lambda_l = \Lambda_0 e^{- l}$.
2097: All vertices and coupling constants then depend on both logarithmic
2098: flow parameters $l $ and $\tau$.
2099: Instead of Eq.~(\ref{eq:Gscalerel1})
2100: the rescaled fermionic propagator can then be approximated by
2101: \begin{equation}
2102: \tilde{G}_{l, \tau }^{\sigma} ( q , i \epsilon , \alpha ) \approx
2103: \frac{ \Theta ( 1 < | q| < e^{\tau} )}{
2104: i \epsilon - \alpha \tilde{v}_{l,\tau}^{\sigma} q - r_{l, \tau}^{\sigma}}
2105: \; ,
2106: \label{eq:Gscalerel2}
2107: \end{equation}
2108: and the corresponding single-scale propagator is
2109: \begin{equation}
2110: \dot{ \tilde{G}}_{l, \tau }^{\sigma} ( q , i \epsilon , \alpha ) \approx
2111: \frac{ \delta ( | q| -1 )}{
2112: i \epsilon - \alpha \tilde{v}_{l, \tau}^{\sigma} q - r_{l, \tau}^{\sigma}}
2113: \; .
2114: \label{eq:Gsinglescalerel2}
2115: \end{equation}
2116: We recover the vertices of the momentum
2117: transfer cutoff scheme by taking first the limit $ \tau \rightarrow \infty$.
2118: Of course,
2119: the result of the RG should be independent of how we reach a
2120: certain point in two-dimensional cutoff space
2121: spanned by $\Lambda_l$ and $\Lambda_{\tau}^F$.
2122: Suppose we first fix $\Lambda^F_{\tau} = \Lambda^F_0$
2123: and perform the reduction of the momentum transfer cutoff
2124: $\Lambda_0 \rightarrow \Lambda_l$. As a second step we reduce the
2125: fermionic cutoff
2126: $\Lambda^F \rightarrow 0$.
2127: On the right-hand side of the flow equation for the
2128: spin-flip susceptibility there are then two additional
2129: diagrams~\cite{Schuetz05} involving the fermionic single-scale
2130: propagator, which are
2131: shown in Fig.~\ref{fig:flowsuscepfermi}.
2132: %
2133: %
2134: \begin{figure}[tb]
2135: \centering
2136: % \psfrag{t1}{$T >T_c$}
2137: % \vspace{7mm}
2138: \epsfig{file=fig11.eps,width=70mm}
2139: % \vspace{4mm}
2140: \caption{%
2141: Additional diagrams contributing to the flow of the spin-flip susceptibility
2142: in a cutoff scheme with a fermionic band-width cutoff.
2143: The solid arrow with slash is the fermionic single-scale propagator.
2144: }
2145: \label{fig:flowsuscepfermi}
2146: \end{figure}
2147: %
2148: %
2149: For a fixed scale $l$ the corresponding flow equation is
2150: (for simplicity we set $\alpha=1$ and omit the chirality label)
2151: \begin{eqnarray}
2152: \partial_\tau \tilde{\Pi}^{\sigma \bar{\sigma}}_{l,\tau}
2153: ( \bar{Q}) &=&
2154: - [ \bar{q} \partial_{\bar{q}} + \bar{\epsilon} \partial_{\bar{\epsilon}} ]
2155: \tilde{\Pi}^{\sigma \bar{\sigma}}_{l,\tau} ( \bar{Q})
2156: + \dot{{\Pi}}^{\sigma \bar{\sigma}}_{l,\tau} (\bar{Q})
2157: ,
2158: \label{FlowPi}
2159: \end{eqnarray}
2160: with
2161: \begin{eqnarray}
2162: \dot{\Pi}^{\sigma \bar{\sigma}}_{l,\tau} (\bar{Q}) &=&
2163: - {\gamma}_{l,\tau}^2 \int_{Q} \Bigl[
2164: \dot{\tilde{G}}_{l, \tau}^{\sigma} ({Q})
2165: {\tilde{G}}^{\bar{\sigma}}_{l,\tau} (Q+\bar{Q}+ \sigma \tilde{\Delta}^{\ast}_{\tau})
2166: \nonumber
2167: \\
2168: & & + {\tilde{G}}_{l,\tau}^{\sigma}
2169: ({Q}) \dot{\tilde{G}}^{\bar{\sigma}}_{l,\tau}
2170: (Q+\bar{Q}+ \sigma \tilde{\Delta}_{\tau}^{\ast} )
2171: \Bigr]
2172: \; ,
2173: \label{DotPi}
2174: \end{eqnarray}
2175: where $\tilde{\Delta}_{\tau}^{\ast} = \Delta / \Lambda_{\tau}^F$.
2176: The flow equation for the spin-flip vertex is now of the form
2177: \begin{equation}
2178: \partial_\tau {\gamma}_{l,\tau} = \Theta( \lambda + l - \tau )
2179: C (l,\tau )
2180: \; ,
2181: \end{equation}
2182: where $C ( l , \tau )$ is some function of the flow parameters
2183: and of the running coupling constants, and
2184: $\lambda = \ln ( \Lambda_0^F / \Lambda_0 )$.
2185: For simplicity we choose $\Lambda_0 = \Lambda_0^F$, so that $\lambda =0$.
2186: The $\Theta$-function is due to the fact that
2187: the internal loop momenta are restricted by the momentum transfer cutoff
2188: $\Lambda_0$. Obviously, $ \partial_\tau {\gamma}_{l,\tau} =0$ for $\tau > l$,
2189: so that $\gamma_l \equiv \gamma_{ l , \tau > l }$ is independent of $\tau$.
2190: Similarly, the
2191: flow equation for $r_{l , \tau }^{\sigma}$ is of the form
2192: \begin{equation}
2193: \partial_\tau {r}^\sigma_{l,\tau}
2194: = {r}^{\sigma}_{l, \tau} + \Theta( l - \tau ) A ( l , \tau )
2195: \; ,
2196: \end{equation}
2197: with some other function $A ( l , \tau )$. This
2198: implies $r_{l , \tau }^{\sigma} = e^{ \tau - l } r^{\sigma}_{ l , \tau = l}$
2199: for $ \tau > l$,
2200: so that the flowing Fermi point distance
2201: $\tilde{\Delta}_{ l , \tau } $, defined analogous to
2202: Eq.~(\ref{eq:tildeDeltaldef}) via
2203: \begin{equation}
2204: \tilde{\Delta}_{ l , \tau } =
2205: \frac{ \Delta }{ \Lambda_\tau^{F} }
2206: - ( r^{+}_{ l , \tau }
2207: - r^-_{l , \tau } )
2208: \; ,
2209: \label{eq:Fermidistflow}
2210: \end{equation}
2211: is for $ \tau > l$ of the form
2212: \begin{eqnarray}
2213: \tilde{\Delta}_{ l , \tau > l } & = &
2214: \frac{ \Delta }{ \Lambda_\tau^{F} }
2215: - e^{( \tau - l) } ( r^{+}_{ l , l }
2216: - r^-_{l , l } )
2217: \nonumber
2218: \\
2219: & = &
2220: e^{\tau - l }
2221: \left[
2222: \frac{ \Delta }{ \Lambda_{ l} }
2223: - ( r^{+}_{ l }
2224: - r^-_{l } ) \right]
2225: \nonumber
2226: \\
2227: & = &
2228: e^{\tau - l } \tilde{\Delta}_{ l}
2229: \; ,
2230: \end{eqnarray}
2231: where we have defined $r_l^{\sigma} = r_{ l , \tau = l}^{\sigma}$ and
2232: $ \tilde{\Delta}_{ l}
2233: = \Delta / \Lambda_{ l } - ( r_l^{+} - r_l^- )$, see
2234: Eq.~(\ref{eq:tildeDeltaldef}).
2235: For $ \tau \rightarrow \infty$
2236: the solution of Eq.~(\ref{FlowPi}) can therefore
2237: be written as
2238: \begin{eqnarray}
2239: \tilde{\Pi}^{\sigma \bar{\sigma}}_{l} ( \bar{Q} )
2240: &=& \int_0^\infty \!\!\! d\tau \:
2241: \dot{\Pi}^{\sigma \bar{\sigma} }_{l, \tau}( \bar{Q}
2242: e^{\tau- l } ; \tilde{\Delta}_{l, \tau})
2243: \nonumber
2244: \\
2245: &=& \int_0^{ {l}} \!\!\! d\tau \:
2246: \dot{\Pi}^{\sigma \bar{\sigma} }_{l, \tau}( \bar{Q}
2247: e^{\tau- l } ; \tilde{\Delta}_{l, \tau})
2248: \nonumber
2249: \\
2250: & + &
2251: \int_{ {l}}^\infty \!\!\! d\tau \;
2252: \dot{\Pi}^{\sigma \bar{\sigma} }_{l, \tau}( \bar{Q}
2253: e^{\tau- l } ; \tilde{\Delta}_{ l} e^{\tau - l } )
2254: \label{SolutionPi}
2255: \; .
2256: \end{eqnarray}
2257: Using the fact that
2258: in the integral of the last term we may pull out a factor
2259: of $\gamma_l^2$ and
2260: approximating the fermionic Green functions
2261: by Eqs.~(\ref{eq:Gscalerel2}) and
2262: (\ref{eq:Gsinglescalerel2}) with $\tilde{v}^{\sigma}_{ l, \tau}$ and
2263: $r_{l , \tau }$ set equal to zero,
2264: we recover from the last term
2265: the adiabatic approximation (\ref{eq:Pisfapprox}).
2266: Actually, to calculate the flow of the fermionic self energy, we
2267: only need the bosonic Green function at $q= \pm 1$.
2268: For $\tilde{\Delta}_l < 1$ the non-adiabatic
2269: contribution to the polarization then vanishes,
2270: leaving us just with the adiabatic part. For larger $\tilde{\Delta}_l$
2271: there is a crossover to the more general expression. However,
2272: the leading contribution to the flow of $\tilde{\Delta}_l$ itself stems from the region where the adiabatic approximation is valid.
2273:
2274:
2275: Finally we note that the adiabatic approximation (\ref{eq:Pisfapprox})
2276: is also consistent with the
2277: flow equation for the spin-flip susceptibility
2278: in the momentum transfer cutoff scheme
2279: shown in Fig.~\ref{fig:flowPi}: taking
2280: the derivative of the right-hand side
2281: of Eq.~(\ref{eq:Pisfapprox}) with respect to the flow parameter
2282: and inserting for the derivative of the fermionic self-energy
2283: its flow equation (\ref{eq:RGselfvertex}), we find
2284: that the right-hand side can be written in terms of
2285: a symmetrized closed fermion loop with four bosonic legs and
2286: renormalized propagators. This corresponds to
2287: the adiabatic approximation
2288: for the vertex $\Gamma^{(0,4)}_{\Lambda}$
2289: which according to Fig.~\ref{fig:flowPi} drives
2290: the flow of the spin-flip susceptibility.
2291:
2292:
2293: %\vspace{-4mm}
2294: \begin{thebibliography}{99}
2295: %\vspace{-4mm}
2296: %
2297: \bibitem{Pomenanchuk58}
2298: I. J. Pomeranchuk, Zh. Eksp. Teor. Fiz. {\bf{35}}, 524 (1958)
2299: [Sov. Phys. JETP {\bf{8}}, 361 (1958)].
2300: %
2301: \bibitem{Lifshitz60}
2302: I. M. Lifshitz, Zh. Eksp. Teor. Fiz. {\bf{38}}, 1569 (1960)
2303: [Sov. Phys. JETP {\bf{11}}, 1130 (1960)].
2304: %
2305: \bibitem{Quintanilla06}
2306: J. Quintanilla and A. J. Schofield, Phys. Rev. B {\bf{74}}, 115126 (2006).
2307: %
2308: \bibitem{Luther94}
2309: A. Luther, Phys. Rev. B {\bf{50}}, 11446 (1994).
2310: %
2311: \bibitem{Zheleznyak97}
2312: A. T. Zheleznyak, V. M. Yakovenko, and I. E. Dzyaloshinskii,
2313: Phys. Rev. B {\bf{55}}, 3200 (1997).
2314: %
2315: \bibitem{Ferraz03}
2316: A. Ferraz, Phys. Rev. B {\bf{68}}, 075115 (2003);
2317: H. Freire, E. Correa, and A. Ferraz, Phys. Rev. B {\bf{71}}, 165113 (2005).
2318: %
2319: \bibitem{Honerkamp01}
2320: C. Honerkamp, M. Salmhofer, N. Furukawa, and T. M. Rice,
2321: Phys. Rev. B {\bf{63}}, 035109 (2001).
2322: %
2323: \bibitem{Ledowski06}
2324: S. Ledowski and P. Kopietz, unpublished.
2325: %
2326: \bibitem{Neumayr03}
2327: A. Neumayr and W. Metzner, Phys. Rev. B {\bf{67}}, 035112 (2003).
2328: %
2329: \bibitem{Dusuel03}
2330: S. Dusuel and B. Dou\c{c}ot, Phys. Rev. B {\bf{67}}, 205111 (2003).
2331: %
2332: \bibitem{Feldman96}
2333: J. Feldman, M. Salmhofer, and E. Trubowitz, J. Stat. Phys. {\bf{84}}, 1209 (1996);
2334: Commun. Pure Appl. Math. {\bf{51}}, 1133 (1998);
2335: {\it{ibid.}} {\bf{52}}, 273 (1999).
2336: %
2337: \bibitem{Kopietz01}
2338: P. Kopietz and T. Busche, Phys. Rev. B {\bf{64}}, 155101 (2001).
2339: %
2340: \bibitem{Ledowski03}
2341: S. Ledowski and P. Kopietz,
2342: J. Phys.: Condens. Matter {\bf{15}}, 4779 (2003).
2343: %
2344: \bibitem{Ledowski05}
2345: S. Ledowski, P. Kopietz, and A. Ferraz, Phys. Rev. B {\bf{71}}, 235106 (2005).
2346: %
2347: \bibitem{Schuetz05}
2348: F. Sch\"{u}tz, L. Bartosch, and P. Kopietz, Phys. Rev. B {\bf{72}},
2349: 035107 (2005);
2350: %
2351: %\bibitem{Schuetz06}
2352: F. Sch\"{u}tz and P. Kopietz, J. Phys. A: Math. Gen. {\bf{39}}, 8205 (2006).
2353: %
2354: \bibitem{Fabrizio93}
2355: M. Fabrizio, Phys. Rev. B {\bf{48}}, 15838 (1993).
2356: %
2357: \bibitem{Brazovskii85}
2358: S. A. Brazovskii and V. M. Yakovenko,
2359: Zh. Eksp. Teor. Fiz. {\bf{89}}, 2318 (1985) [Sov. Phys. JETP {\bf{62}}, 1340 (1985)].
2360: %
2361: \bibitem{Bourbonnais91}
2362: C. Bourbonnais and L. G. Caron, Int. J. Mod. Phys. B {\bf{5}}, 1033 (1991).
2363: %
2364: \bibitem{Kusmartsev92}
2365: F. Kusmartsev, A. Luther, and A. A. Nersesyan,
2366: Pis'ma Zh. Eksp. Teor. Fiz. {\bf{55}}, 692 (1992)
2367: [JETP Lett. {\bf{55}}, 724 (1992)]; V. M. Yakovenko, {\it{ibid.}} {\bf{56}}, 523 (1992)
2368: [{\bf{56}}, 510 (1992)].
2369: %
2370: \bibitem{Finkelstein93}
2371: A. M. Finkelstein and A. I. Larkin, Phys. Rev. B {\bf{47}}, 10461 (1993).
2372: %
2373: \bibitem{Boies95}
2374: D. Boies, C. Bourbonnais and A.-M. S. Tremblay, Phys. Rev. Lett. {\bf{74}}, 968 (1995).
2375: %
2376: \bibitem{Balents96}
2377: L. Balents and M. P. A. Fisher, Phys. Rev. B {\bf{53}}, 12133 (1996);
2378: H.-H. Lin, L. Balents, and M. P. A. Fisher, Phys. Rev. B {\bf{56}}, 6569 (1997).
2379: %
2380: \bibitem{Arrigoni98}
2381: E. Arrigoni, Phys. Rev. Lett. {\bf{80}}, 790 (1998); Phys. Rev. B {\bf{61}}, 7909 (2000).
2382: %
2383: \bibitem{Ledermann00}
2384: U. Ledermann and K. Le Hur, Phys. Rev. B {\bf{61}}, 2497 (2000).
2385: %
2386: \bibitem{Louis01}
2387: K. Louis, J. V. Alvarez, and C. Gros, Phys. Rev. B {\bf{64}}, 113106 (2001);
2388: K. Hamacher, C. Gros, and W. Wenzel, Phys. Rev. Lett. {\bf{88}}, 217203 (2002).
2389: %
2390: \bibitem{Caron02}
2391: L. G. Caron and C. Bourbonnais, Phys. Rev. B {\bf{66}}, 045101 (2002).
2392: %
2393: \bibitem{Bourbonnais04}
2394: C. Bourbonnais, B. Guay, and R. Wortis, in
2395: {\it{Theoretical Methods for Strongly Correlated Electrons}},
2396: ed. D. S\'{e}n\'{e}chal, A. M. Tremblay and C. Bourbonnais
2397: (Springer-Verlag, New York, 2004).
2398: %
2399: \bibitem{Nickel06}
2400: J. C. Nickel, R. Duprat, C. Bourbonnais, and N. Dupuis,
2401: Phys. Rev. B {\bf{73}}, 165126 (2006).
2402: %
2403: \bibitem{Tsuchiizu06}
2404: M. Tsuchiizu, Phys. Rev. B {\bf{74}}, 155109 (2006).
2405: %
2406: \bibitem{footnotehubbard}
2407: There are several possibilities to
2408: write the Hubbard interaction $U \bar{\psi}^+ \psi^+ \bar{\psi}^- \psi^-$
2409: in the form (\ref{eq:action1}).
2410: One possibility is to set
2411: $J^{\bot} ( \bar{k} ) = U a /2 $ and $f ( \bar{k} ) = u ( \bar{k} ) =
2412: J^{\parallel} ( \bar{k} ) =0$, where $a$ is the lattice spacing.
2413: An alternative is
2414: $f ( \bar{k} ) = J^{\parallel} ( \bar{k} ) = Ua/2$ and
2415: $J^{\bot} ( \bar{k} ) = u ( \bar{k} ) =0$. This freedom
2416: translates into similar ambiguities in the choice of
2417: decoupling of the Hubbard interaction in
2418: terms of bosonic collective fields.
2419: For a discussion of this point see
2420: C. A. Mac\^{e}do and M. D. Coutinho-Filho,
2421: Phys. B {\bf{43}}, 13515 (1991), and
2422: S. De Palo, C. Castellani, C. Di Castro, and B. K. Chakraverty,
2423: Phys. Rev. B {\bf{60}}, 564 (1999).
2424: %
2425: \bibitem{Solyom79}
2426: J. Solyom, Adv. Phys. {\bf{28}}, 201 (1979).
2427: %
2428: \bibitem{Wetterich93}
2429: C. Wetterich, Phys. Lett. B {\bf{301}}, 90 (1993).
2430: %
2431: \bibitem{Morris94}
2432: T. R. Morris, Int. J. Mod. Phys. A {\bf{9}}, 2411 (1994).
2433: %
2434: \bibitem{Wetterich02}
2435: C. Wetterich, cond-mat/0208361v3.
2436: %
2437: \bibitem{Baier04}
2438: T. Baier, E. Bick, and C. Wetterich, Phys. Rev. B {\bf{70}}, 125111 (2004).
2439: %
2440: \bibitem{Salmhofer01}
2441: M. Salmhofer and C. Honerkamp, Prog. Theor. Physics {\bf{105}}, 1 (2001).
2442: %
2443: \bibitem{Polchinski84}
2444: J. Polchinski, Nucl. Phys. B {\bf{231}}, 269 (1984).
2445: %
2446: \bibitem{Ledowski04}
2447: S. Ledowski, N. Hasselmann, and P. Kopietz,
2448: Phys. Rev. A {\bf{69}}, 061601(R) (2004);
2449: N. Hasselmann, S. Ledowski, and P. Kopietz,
2450: {\it{ibid.}} {\bf{70}}, 063621 (2004).
2451: %
2452: %\bibitem{ZinnJustin02}
2453: %J. Zinn-Justin, {\it{Quantum Field Theory and Critical Phenomena}},
2454: %4th ed. (Clarendon, Oxford, 2002).
2455: %
2456: \end{thebibliography}
2457:
2458:
2459:
2460:
2461:
2462: \end{document}
2463:
2464:
2465:
2466:
2467:
2468:
2469:
2470:
2471: