1: % last correction: August 21 by Florian Schuetz
2: %
3: \tolerance = 10000
4: %%%%%%%%%%%For two-column take this:%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5: \documentclass[twocolumn,showpacs,superscriptaddress,prb,amsmath,amssymb,floatfix,eqsecnum]{revtex4}
6: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7: %
8: %
9: %%%%%%%%%%%For one-column take this:%%%%%%%%%%%%%%%%%%%%%%%%%%
10: %\documentclass[galley,showpacs,superscriptaddress,prb,amsmath,amssymb,floatfix,eqsecnum]{revtex4}
11: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
12: %
13: %
14: \usepackage{amsmath,amssymb}
15: \usepackage{bm}
16: \usepackage{epsfig,psfrag}
17:
18: % set \bd to \bf or \bm
19: \newcommand{\bd}{\bm}
20:
21: \newcommand{\G}{\mathcal{G}}
22: \newcommand{\Z}{\mathcal{Z}}
23: \newcommand{\LL}{\mathcal{L}}
24: \newcommand{\idf}{\int\!\!D\Phi\,\,}
25: \newcommand{\idp}{\int\!\!D\psi\,\,}
26: \newcommand{\idpf}{\int\!\!D\psi\,D\varphi\,\,}
27: \newcommand{\Go}{\mathbf{G}_0}
28: \newcommand{\p}[1]{\frac{\delta}{\delta #1}}
29: \newcommand{\pp}[2]{\frac{\delta #1}{\delta #2}}
30: \newcommand{\ppt}[3]{\frac{\delta^{(2)} #1}{\delta #2 \delta #3}}
31: \newcommand{\Tr}{\mathrm{Tr}}
32:
33: \begin{document}
34:
35: \title{Collective fields in the functional renormalization group for
36: fermions, Ward identities, and the exact solution of the
37: Tomonaga-Luttinger model}
38:
39: \author{Florian Sch\"{u}tz}
40: \affiliation{Institut f\"{u}r Theoretische
41: Physik, Universit\"{a}t Frankfurt, Max-von-Laue-Strasse 1, 60054
42: Frankfurt, Germany}
43:
44: \author{Lorenz Bartosch}
45: \affiliation{Institut f\"{u}r Theoretische
46: Physik, Universit\"{a}t Frankfurt, Max-von-Laue-Strasse 1, 60054
47: Frankfurt, Germany}
48:
49: \affiliation{Department of Physics, Yale University, P.O.Box 208120,
50: New Haven, CT 06520-8120, USA}
51:
52: \author{Peter Kopietz}
53: \affiliation{Institut f\"{u}r Theoretische
54: Physik, Universit\"{a}t Frankfurt, Max-von-Laue-Strasse 1, 60054
55: Frankfurt, Germany}
56:
57:
58: \date{September 17, 2004}
59: %\date{\today}
60:
61: \begin{abstract}
62: We develop a new formulation of the functional renormalization group
63: (RG) for interacting fermions. Our approach unifies the purely
64: fermionic formulation based on the Grassmannian functional integral,
65: which has been used in recent years by many authors, with the
66: traditional Wilsonian RG approach to quantum systems pioneered by
67: Hertz [Phys. Rev. B {\bf 14}, 1165 (1976)], which attempts to
68: describe the infrared behavior of the system in terms of an
69: effective bosonic theory associated with the soft modes of the
70: underlying fermionic problem. In our approach, we decouple the
71: interaction by means of a suitable Hubbard-Stratonovich
72: transformation (following the Hertz-approach), but do not eliminate
73: the fermions; instead, we derive an exact hierarchy of RG flow
74: equations for the irreducible vertices of the resulting coupled
75: field theory involving both fermionic and bosonic fields. The
76: freedom of choosing a momentum transfer cutoff for the bosonic soft
77: modes in addition to the usual band cutoff for the fermions opens
78: the possibility of new RG schemes. In particular, we show how the
79: exact solution of the Tomonaga-Luttinger model (i.e.,
80: one-dimensional fermions with linear energy dispersion and
81: interactions involving only small momentum transfers) emerges from
82: the functional RG if one works with a momentum transfer cutoff.
83: Then the Ward identities associated with the local particle
84: conservation at each Fermi point are valid at every stage of the RG
85: flow and provide a solution of an infinite hierarchy of flow
86: equations for the irreducible vertices. The RG flow equation for
87: the irreducible single-particle self-energy can then be closed and
88: can be reduced to a linear integro-differential equation, the
89: solution of which yields the result familiar from bosonization. We
90: suggest new truncation schemes of the exact hierarchy of flow
91: equations, which might be useful even outside the weak coupling
92: regime.
93: \end{abstract}
94:
95: \pacs{%
96: %71.10.-w,
97: 71.10.Pm., 71.10.Hf}
98:
99: \maketitle
100:
101: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
102: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
103: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
104: \section{Introduction}
105:
106: In condensed matter and statistical physics the renormalization group
107: (RG) in the form developed by Wilson and coauthors
108: \cite{Wilson72,Wilson74,Fisher98} has been very successful. At the
109: heart of this intuitively appealing formulation of the RG lies the
110: concept of an effective action describing the physical properties of a
111: system at a coarse grained scale. Conceptually, the derivation of
112: this effective action is rather simple provided the theory can be
113: formulated in terms of a functional integral: one simply integrates
114: out the degrees of freedom describing short-wavelength fluctuations in
115: a certain regime, and subsequently rescales the remaining degrees of
116: freedom in order to compare the new effective action with the original
117: one. Of course, in practice, the necessary functional integration can
118: almost never be performed analytically, so that one has to resort to
119: some approximate procedure.\cite{Ma76} However, if the elimination of
120: the short wavelength modes is performed in infinitesimal steps, one
121: can write down a formally exact RG flow equation describing the change
122: of the effective action due to mode elimination and rescaling. The
123: earliest version of such a functional RG equation has been derived by
124: Wegner and Houghton.\cite{Wegner73} Subsequently, many authors have
125: derived alternative versions of the functional RG with the same
126: physical content, using different types of generating functionals. In
127: particular, the advantages of working with the generating functional
128: of the one-particle irreducible vertices have been realized early on
129: by Di Castro, Jona-Lasinio, and Peliti,\cite{DiCastro74} and by
130: Nicoll, Chang, and Stanley.\cite{Nicoll76,Chang92} The focus of the
131: above works was an accurate description of second-order phase
132: transitions at finite temperatures, where quantum mechanics is
133: irrelevant.
134:
135: In recent years, there has been much interest in quantum systems
136: exhibiting phase transitions as a function of some nonthermal control
137: parameter, such as pressure or density. In a pioneering paper, Hertz
138: \cite{Hertz76} showed how the powerful machinery of the Wilsonian RG
139: can be generalized to study quantum critical phenomena in Fermi
140: systems. Technically, this is achieved with the help of so-called
141: Hubbard-Stratonovich transformations, which replace the fermionic
142: two-particle interaction by a suitable bosonic field that couples to a
143: quadratic form in the fermion operators.\cite{Negele88} The fermions
144: can then be integrated out in a formally exact way, resulting in an
145: effective action for the bosonic field. Of course, there are many
146: possible ways of decoupling fermionic two-body interactions by means
147: of Hubbard-Stratonovich transformations. In the spirit of the usual
148: Ginzburg-Landau-Wilson approach to classical critical phenomena, one
149: tries to construct the effective bosonic theory such that the field
150: can be identified with the fluctuating order parameter, or its field
151: conjugate. However, if the system supports additional soft modes that
152: couple to the order parameter,\cite{Kirkpatrick96,Rosch01} an attempt
153: to construct an effective field theory in terms of the order parameter
154: alone leads in general to an effective action with singular and
155: nonlocal vertices, so that the usual RG methods developed for
156: classical phase transitions cannot be applied. In this case, it is
157: better to construct an effective action involving all soft modes
158: explicitly. However, for a given problem the nature of the soft modes
159: is not known {\it a priori}, so that the explicit introduction of the
160: corresponding degrees of freedom by means of a suitable
161: Hubbard-Stratonovich transformation is always based on some prejudice
162: about the nature of the ground state and the low-lying excitations of
163: the system. Recently, the breakdown of simple Ginzburg-Landau-Wilson
164: theory has also been discussed in the context of quantum
165: antiferromagnets by Senthil {\em et~al.}\cite{Senthil04}
166:
167: In order to have a completely unbiased RG approach for interacting
168: fermions, one can also apply the Wilsonian RG directly to the
169: Grassmannian functional integral representation of the partition
170: function or the Green's functions of an interacting Fermi
171: system.\cite{Shankar94} In the past ten years, several groups have
172: further developed this
173: method.\cite{Zanchi96,Halboth00,Honerkamp01,Honerkamp01b,Salmhofer01,Kopietz01,Busche02,Ledowski03,Tsai01,Binz02,Meden02,Kampf03,Katanin04}
174: In particular, the mode elimination step of the RG transformation has
175: been elegantly cast into formally exact differential equations for
176: suitably defined generating functionals. These functional equations
177: then translate to an infinite hierarchy of integro\-differential
178: equations for the vertices. On a technical level, it is often
179: advantageous to work with the generating functional of the
180: one-particle irreducible vertices.\cite{DiCastro74,Nicoll76,Chang92}
181: For classical field theories the exact RG flow equation for this
182: generating functional has been obtained by
183: Wetterich,\cite{Wetterich93} and by Morris.\cite{Morris94} For
184: nonrelativistic fermions, the corresponding flow equation has been
185: derived by Salmhofer and Honerkamp \cite{Salmhofer01} and,
186: independently and in more explicit form, by Kopietz and
187: Busche.\cite{Kopietz01}
188:
189: However, the purely fermionic formulation of the Wilsonian RG has
190: several disadvantages. In order to obtain at least an approximate
191: solution of the formally exact hierarchy of RG flow equations for the
192: vertices, severe truncations have to be made, which can only be
193: justified as long as the fermionic four-point vertex (i.e., the
194: effective interaction) remains small. Hence, in practice the
195: fermionic functional RG is restricted to the weak coupling regime, so
196: that possible strong coupling fixed points are not accessible within
197: this method. Usually, one-loop truncated RG equations are iterated
198: until at least one of the marginal interaction constants becomes
199: large, which is then interpreted as a weak-coupling instability of the
200: Fermi system in the corresponding
201: channel.\cite{Halboth00,Honerkamp01,Honerkamp01b}
202:
203: Within the framework of the Wilsonian functional RG, a consistent
204: two-loop calculation has not been performed so far due to the immense
205: technical difficulties involved. Such a calculation should also take
206: into account the frequency dependence of the effective interaction and
207: the self-energy corrections to the internal Green's functions. Note
208: that in the work by Katanin and Kampf \cite{Kampf03} the frequency
209: dependence of the effective interaction has been ignored, so that the
210: effect of possible bosonic collective modes is not included in these
211: calculations. It is well known that two-loop calculations are more
212: conveniently performed using the field-theoretical RG method, provided
213: the physical problem of interest can be mapped onto a renormalizable
214: field theory. Ferraz and coauthors \cite{Ferraz03,Ferraz03b}
215: recently used the field-theoretical RG to calculate the
216: single-particle Green function at the two loop level for a special
217: two-dimensional Fermi system with a flat Fermi surface. Interestingly,
218: they found a new non-Fermi liquid fixed point characterized by a
219: finite renormalized effective interaction and a vanishing
220: wave-function renormalization. The running interaction without
221: wave-function renormalization diverges in this model, but the
222: divergence is canceled by the vanishing wave-function renormalization
223: such that the renormalized interaction remains finite.
224:
225: Another difficulty inherent in any RG approach to Fermi systems at
226: finite densities arises from the fact that the true Fermi surface of
227: the interacting many-body system is not known {\it a priori}. In fact, in
228: dimensions $D > 1$ interactions can even change the symmetry of the
229: Fermi surface.\cite{Pomeranchuk58,Metzner04} In a perturbative
230: approach, finding the renormalized Fermi surface is a delicate
231: self-consistency problem \cite{Nozieres64}; if one starts from the
232: wrong Fermi surface one usually encounters unphysical
233: singularities.\cite{Kohn60} So far, the self-consistent
234: renormalization of the Fermi surface has not been included in the
235: numerical analysis of the one-loop truncated fermionic functional RG.
236: As shown in Refs.~\onlinecite{Kopietz01}, \onlinecite{Ledowski03}, and
237: \onlinecite{Ferraz03b}, the renormalized Fermi surface can be defined
238: as a fixed point of the RG and can in principle be calculated
239: self-consistently entirely within the RG framework.
240:
241: The progress in overcoming the above difficulties inherent in the RG
242: approach to fermions using a purely fermionic parametrization has
243: been rather slow. In our opinion, this has a simple physical reason:
244: the low lying excitations (i.e., soft modes) of an interacting Fermi
245: system consist not only of fermionic quasiparticles, but also of
246: bosonic collective excitations.\cite{Pines89} The latter are rather
247: difficult to describe within the purely fermionic parametrizations
248: used in
249: Refs.~\onlinecite{Zanchi96,Halboth00,Honerkamp01,Honerkamp01b,Salmhofer01,Kopietz01,Busche02,Ledowski03,Tsai01,Binz02,Meden02,Kampf03,Katanin04}.
250: Naturally, a formulation of the functional RG where both fermionic and
251: bosonic excitations are treated on equal footing should lead to a more
252: convenient parametrization. Such a strategy seems also natural in
253: light of the observation by Kirkpatrick, Belitz, and co-workers
254: \cite{Kirkpatrick96} (see also Ref.~\onlinecite{Rosch01}) that an
255: effective low-energy and long-wavelength action with well-behaved
256: vertices is only obtained if the soft modes are not integrated out,
257: but appear explicitly as quantum fields.
258:
259:
260: Our aim in this work is to set up a functional renormalization group
261: scheme that allows for a simultaneous treatment of fermionic as well
262: as collective bosonic degrees of freedom. This is done by explicitly
263: decoupling the interaction via a Hubbard-Stratonovich transformation
264: in the spirit of Hertz \cite{Hertz76} and then considering the
265: functional renormalization group equations for the mixed field theory
266: involving both fermionic and bosonic fields. This type of approach
267: has been suggested previously by Correia, Polonyi, and
268: Richert,\cite{Correia01} who studied the homogeneous electron gas by
269: means of a gradient expansion of a functional version of a
270: Callan-Symanzik equation. Here, we follow the more standard approach
271: and derive a hierarchy of flow equations for the vertex functions of
272: our coupled Fermi-Bose theory. A related approach has also been
273: developed by Wetterich and coauthors.\cite{Wetterich04,Baier03}
274: However, they discussed only a simple truncation of the exact
275: hierarchy of RG flow equations involving an effective potential in a
276: bosonic sector and a momentum-independent Yukawa coupling. They did
277: not pay any attention to the problem of the compatibility of the RG
278: flow with Ward identities, which play a crucial role for interacting
279: fermions with dominant forward scattering. Our procedure bridges the
280: gap between the purely bosonic approach to quantum critical phenomena
281: by Hertz \cite{Hertz76} and the purely fermionic functional RG method
282: developed in the past decade by several
283: authors.\cite{Zanchi96,Halboth00,Honerkamp01,Honerkamp01b,Salmhofer01,Kopietz01,Busche02,Ledowski03,Tsai01,Binz02,Meden02,Kampf03,Katanin04}
284: Contrary to the approach by Hertz,\cite{Hertz76} in our approach the
285: fermionic degrees of freedom are still present, so that the feedback
286: of the collective bosonic modes on the one-particle spectral
287: properties of the fermions can be studied.
288:
289:
290: The parametrization of the low-lying excitations of an interacting
291: Fermi system in terms of collective bosonic fields is most natural in
292: one spatial dimension,\cite{Giamarchi04,Haldane81} where Fermi-liquid
293: theory breaks down and is replaced by the Luttinger-liquid
294: concept.\cite{Haldane81} An exactly solvable paradigm of a Luttinger
295: liquid is the Tomonaga-Luttinger model
296: (TLM),\cite{Giamarchi04,Haldane81,Stone94} consisting of fermions with
297: exactly linear energy dispersion and interactions involving only small
298: momentum transfers. A description in terms of bosonic variables is
299: well known to provide an exact solution for thermodynamic quantities
300: as well as correlation functions.\cite{Giamarchi04,Haldane81,Stone94}
301: One might then wonder if it is also possible to obtain the complete
302: form of the correlation functions of the TLM entirely within the
303: framework of the functional RG. An attempt \cite{Busche02} to
304: calculate the momentum- and frequency-dependent single-particle
305: spectral function $A ( k , \omega )$ of the TLM by means of an
306: approximate iterative two-loop solution of the functional RG equations
307: at weak coupling yields the correct behavior of $A ( \pm k_F ,
308: \omega)$ known from bosonization (where $k_F$ is the Fermi momentum),
309: but yields incorrect threshold singularities for momenta away from
310: $\pm k_F$. In this work we go considerably beyond Ref.
311: \onlinecite{Busche02} and show how the TLM can be solved
312: {\it{exactly}} using our mixed fermionic-bosonic functional RG. A
313: crucial point of our method is that the RG can be set up in such a way
314: that the Ward identities underlying the exact solubility of the TLM
315: \cite{Dzyaloshinskii74,Bohr81,Metzner98,Kopietz97} are preserved by
316: the RG. This is not the case in the purely fermionic formulation of
317: the functional RG.\cite{Katanin04}
318:
319: Very recently, Benfatto and Mastropietro\cite{Benfatto04} also
320: developed an implementation of the RG for the TLM which takes the
321: asymptotic Ward identities into account. However, these authors did
322: not introduce bosonic collective fields and did not attempt to
323: calculate the exact single-particle Green's function.
324:
325: The rest of the paper is organized as follows. In
326: Sec.~\ref{sec:preliminaries}, we introduce the model, carry out the
327: decoupling of the interaction and set up a compact notation which
328: turns out to facilitate the bookkeeping in the derivation of the
329: functional RG equations that is presented in detail in
330: Sec.~\ref{sec:flow_eq}. In Sec.~\ref{sec:1dflow}, we
331: introduce a new RG scheme which uses the momentum transfer of the
332: effective interaction as a cutoff. We show that for linearized energy
333: dispersion the resulting infinite hierarchy of flow equations for the
334: irreducible vertices involving two external fermion legs and an
335: arbitrary number of external boson legs can be solved exactly by means
336: of an infinite set of Ward identities. Using these identities, the
337: flow equation for the irreducible self-energy can then be reduced to a
338: closed linear integro\-differential equation, which can be solved
339: exactly. In one dimension, we recover in Sec.~\ref{subsec:exactTLM}
340: the exact solution of the Tomonaga-Luttinger model in the form
341: familiar from functional bosonization \cite{Kopietz97,Kopietz95}.
342: Finally, in Sec.~\ref{sec:summary}, we summarize our results and give
343: a brief outlook on possible further applications of our method. There
344: are three appendices where we present some more technical details. In
345: Appendix A, we use our compact notation introduced in
346: Sec.~\ref{sec:preliminaries} to discuss the structure of the tree
347: expansion in our coupled Fermi-Bose theory. Appendix B contains a
348: derivation of the skeleton diagrams for the first few irreducible
349: vertices of our theory using the Dyson-Schwinger equations of motion,
350: which follow from the invariance of the functional integral with
351: respect to infinitesimal shift transformations. Finally, in Appendix C
352: we use the gauge invariance of the mixed Fermi-Bose action to derive a
353: cascade of infinitely many Ward identities involving vertices with two
354: fermion legs and an arbitrary number of boson legs.
355:
356: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
357: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
358: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
359: \section{Interacting Fermions as coupled Fermi-Bose systems}
360: \label{sec:preliminaries}
361:
362: In this section we discuss the Hubbard-Stratonovich transformation and
363: set up a condensed notation to treat fermionic and bosonic fields on
364: the same footing. This will allow us to keep track of the rather
365: complicated diagrammatic structure of the flow equations associated
366: with our coupled Fermi-Bose system in a very efficient way. A similar
367: notation has been used previously in Refs.~\onlinecite{Salmhofer01}
368: and \onlinecite{Baier03}.
369:
370: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
371: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
372: \subsection{Hubbard-Stratonovich transformation}
373:
374: We consider a normal fermionic many-body system with two-particle
375: density-density interactions. In the usual Grassmannian functional
376: integral approach \cite{Negele88} the grand-canonical partition
377: function and all (imaginary)-time-ordered Green's functions can be
378: represented as functional averages involving the following Euclidean
379: action:
380: \begin{eqnarray}
381: S[\bar{\psi} , \psi]&=& S_0 [\bar{\psi}, \psi] + S_{\mathrm{int}} [\bar{\psi} ,\psi]
382: \label{eq:Spsidef}
383: \; ,
384: \\ S_0 [\bar{\psi}, \psi]
385: && = \sum_{\sigma}\int_{K} \bar{\psi}_{K\sigma}
386: [-i\omega + \xi_{{\bf k}\sigma}]\psi_{K\sigma}
387: \label{eq:S0psidef}
388: \; ,
389: \\
390: S_{\mathrm{int}} [\bar{\psi} ,\psi]
391: & = &
392: \frac12\sum_{\sigma\sigma'}\int_{\bar{K}}
393: f_{\bar{\bf{k}}}^{\sigma\sigma'}
394: \bar{\rho}_{\bar{K}\sigma}\rho_{\bar{K}\sigma'}\,,
395: \label{eq:Sintpsidef}
396: \end{eqnarray}
397: where the composite index $K=(i\omega,{\bf k})$ contains a fermionic
398: Matsubara frequency $i \omega$ as well as an ordinary wave vector
399: ${\bf{k}}$. Here the energy dispersion
400: $\xi_{\mathbf{k}\sigma}=\epsilon_{\mathbf{k}\sigma}-\mu$ is measured
401: relative to the chemical potential $\mu$, and
402: $f_{\bar{\bf{k}}}^{\sigma\sigma'}$ are some momentum-dependent
403: interaction parameters. Throughout this paper, labels with an overbar
404: refer to bosonic frequencies and momenta, while labels without an overbar
405: refer to fermionic ones. We have normalized the Grassmann fields
406: $\psi_{K \sigma}$ and $\bar{\psi}_{K \sigma}$ such that the
407: integration measure in Eq.~(\ref{eq:Spsidef}) is
408: \begin{equation}
409: \int_K = \frac{1}{\beta V}\sum_{\omega, {\bf{k}}}
410: \quad\stackrel{\beta,V\to\infty}{\longrightarrow}\quad
411: \int\frac{d\omega}{2\pi}\frac{d^Dk}{(2\pi)^D}\,,
412: \end{equation}
413: where $\beta$ is the inverse temperature and $V$ is the volume of the
414: system. The Fourier components of the density are represented by the
415: following composite field:
416: \begin{equation}
417: \rho_{\bar{K}\sigma}=\int_K
418: \bar{\psi}_{K \sigma}\psi_{K+ \bar{K},\sigma}\,,
419: \end{equation}
420: which implies $\bar{\rho}_{\bar{K}\sigma}=\rho_{-\bar{K}\sigma}$. The
421: discrete index $\sigma$ is formally written as a spin projection, but
422: will later on also serve to distinguish right and left moving fields
423: in the Tomonaga-Luttinger model. This is why a dependence of the
424: dispersion $\xi_{{\bf{k}} \sigma} $ on $\sigma$ has been kept.
425:
426: The interaction is bilinear in the densities and can be decoupled by
427: means of a Hubbard-Stratonovich transformation.\cite{Kopietz97} The
428: interaction is then mediated by a real field $\varphi$ and the
429: resulting action reads as
430: \begin{eqnarray}
431: S[\bar{\psi}, \psi,\varphi]&=& S_0 [\bar{\psi}, \psi] + S_0 [\varphi]+
432: S_{1}[\bar{\psi}, \psi,\varphi]
433: \; ,
434: \label{eq:Spsiphidef}
435: \end{eqnarray}
436: where the free bosonic part is given by
437: \begin{eqnarray}
438: S_0 [\varphi] &=&
439: \frac12\sum_{\sigma\sigma'}\int_{\bar{K}}
440: [f^{-1}_{\bar{\bf{k}}}]^{\sigma\sigma'}
441: \varphi^*_{\bar{K}\sigma} \varphi_{\bar{K}\sigma'}
442: \label{eq:S0phi}
443: \; ,
444: \end{eqnarray}
445: and the coupling between Fermi and Bose fields is
446: \begin{eqnarray}
447: S_{1}[\bar{\psi} , \psi,\varphi] &=&
448: i\sum_{\sigma}\int_{\bar{K}}
449: \bar{\rho}_{\bar{K}\sigma}\varphi_{\bar{K}\sigma}
450: \nonumber \\
451: & = &
452: i\sum_{\sigma}\int_{{K}} \int_{\bar{K}}
453: \bar{\psi}_{ K + \bar{K}, \sigma} \psi_{K \sigma} \varphi_{\bar{K}\sigma}
454: \,.
455: \label{eq:Spsiphi}
456: \end{eqnarray}
457: The Fourier components of a real field satisfy
458: $\varphi^*_{\bar{K}\sigma}=\varphi_{-\bar{K}\sigma}$. For the
459: manipulations in the next section it will prove advantageous to
460: further condense the notation and collect the fields in a vector $\Phi
461: = (\psi,\bar{\psi},\varphi)$. The quadratic part of the action can
462: then be written in the symmetric form
463: \begin{eqnarray}
464: S_0[\Phi]& = & S_0[\bar{\psi}, \psi]+S_0 [\varphi] =
465: -\frac12\left(\Phi,\left[{\bf G}_0\right]^{-1}\Phi\right)
466: \nonumber
467: \\
468: & = &
469: - \frac12 \int_{\alpha} \int_{\alpha^{\prime}} \Phi_{\alpha}
470: \left[{\bf G}_0\right]^{-1}_{\alpha \alpha^{\prime}} \Phi_{ \alpha^{\prime}}
471: \,,
472: \end{eqnarray}
473: where ${\bf G}_0$ is now a matrix in frequency, momentum, spin, and
474: field-type indices, and $\alpha$ is a ``super label'' for all of these
475: indices. The symbol $\int_{\alpha}$ denotes integration over the
476: continuous components and summation over the discrete components of
477: $\alpha$. The matrix ${\bf{G}}_0^{-1}$ has the block structure
478: \begin{equation}
479: \mathbf{G}_0^{-1} = \left(\begin{array}{ccc}
480: 0&\zeta [\hat{G}_0^{-1}]^T&0\\
481: \hat{G}_0^{-1}&0&0\\
482: 0&0&-\hat{F}_0^{-1}
483: \end{array}\right)\,,
484: \label{eq:G0matrixinv}
485: \end{equation}
486: where~\cite{footnotezeta} $\zeta = -1$ and $\hat{G}_0$ and $\hat{F}_0$
487: are infinite matrices in frequency, momentum, and spin space, with
488: matrix elements
489: \begin{eqnarray}
490: [ \hat{G}_0 ]_{ K\sigma, K'\sigma'} &=& \delta_{K,K'} \delta_{\sigma \sigma'}
491: G _{0, \sigma} ( K ) \; ,
492: \label{eq:G0matrixdef}\\
493: {}[\hat{F}_0 ]_{ \bar{K} \sigma, \bar{K}^{\prime} \sigma^{\prime}} &=&
494: \delta_{\bar{K} + \bar{K}', 0 }
495: F_{0, \sigma \sigma^{\prime} } ( \bar{K} )
496: \label{eq:F0matrixdef}
497: \; ,
498: \end{eqnarray}
499: where
500: \begin{eqnarray}
501: G_{0 , \sigma} ( K ) &=&
502: [i\omega - \xi_{{\bf{k}}\sigma}]^{-1} \,,
503: \label{eq:G0def}\\
504: F_{0, \sigma \sigma^{\prime}} (\bar{K})
505: &=&
506: f_{\bar{\bf{k}}}^{\sigma \sigma'} \, .
507: \label{eq:F0def}
508: \end{eqnarray}
509: The Kronecker $\delta_{K,K'}=\beta V
510: \delta_{\omega,\omega'}\delta_{\mathbf{k},\mathbf{k}'}$ appearing in
511: Eqs.~(\ref{eq:G0matrixdef},\ref{eq:F0matrixdef}) is normalized such
512: that it reduces to Dirac $\delta$ functions
513: $\delta_{K,K'}\to(2\pi)^{D+1}\delta(\omega-\omega')\delta^{(D)}(\mathbf{k}-\mathbf{k}')$
514: in the limit $\beta,V\to\infty$. Note that the bare interaction plays
515: the role of a free bosonic Green's function. For later reference, we
516: note that the inverse of Eq.~(\ref{eq:G0matrixinv}) is
517: \begin{equation}
518: \mathbf{G}_0 = \left(\begin{array}{ccc}
519: 0& \hat{G}_0 &0\\
520: \zeta \hat{G}_0^{T}&0&0\\
521: 0&0&-\hat{F}_0
522: \end{array}\right)\,,
523: \label{eq:G0matrix}
524: \end{equation}
525: and that the transpose of $\mathbf{G}_0$ satisfies
526: \begin{equation}
527: \mathbf{G}_0^T = \mathbf{Z} \mathbf{G}_0 = \mathbf{G}_0 \mathbf{Z}
528: \; ,
529: \label{eq:G0transpose}
530: \end{equation}
531: where the ``statistics matrix'' $\mathbf{Z}$ is defined by
532: \begin{equation}
533: [ \mathbf{Z} ]_{\alpha \alpha^{\prime} } = \delta_{\alpha \alpha^{\prime} }
534: \zeta_{\alpha}
535: \; .
536: \end{equation}
537: Here, $\zeta_{\alpha} = -1$ if the superindex $\alpha$ refers to a
538: Fermi field, and $\zeta_{\alpha} = 1$ if $\alpha$ labels a Bose field.
539:
540: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
541: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
542: \subsection{Generating functionals}
543:
544: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
545: \subsubsection{Generating functional of connected Green's functions}
546:
547: We now introduce sources $J_{\alpha}$ and define the generating
548: functional $\G [J]$ of the Green's functions as follows:
549: \begin{equation}
550: \G[J] = e^{\G_c[J]} = \frac{1}{\Z_0}\idf e^{-S_0-S_1+(J,\Phi)}
551: \; .
552: \label{eq:Ggen}
553: \end{equation}
554: Here, $\G_c [J]$ is the generating functional for connected Green
555: functions and the partition function $\Z_0$ of the noninteracting
556: system can be written as the Gaussian integral
557: \begin{equation}
558: \Z_0 = \idf e^{-S_0}\,.
559: \end{equation}
560: Let us use the compact notation
561: \begin{equation}
562: (J,\Phi ) = \int_{\alpha} J_{\alpha} \Phi_{\alpha}
563: \; .
564: \label{eq:JPhiscalarproduct}
565: \end{equation}
566: Conventionally, the source terms for fields of different types are
567: written out explicitly in the form\cite{Negele88}
568: \begin{eqnarray}
569: &&(J,\Phi)=(\bar{\jmath},\psi)+(\bar{\psi},j)+(J^*,\varphi)=
570: \nonumber\\
571: &&\sum_{\sigma}\int_K \bar{\jmath}_{K\sigma}\psi_{K\sigma}
572: + \sum_{\sigma}\int_K \bar{\psi}_{K\sigma}j_{K\sigma}
573: + \sum_{\sigma}\int_{\bar{K}} J^*_{\bar{K}\sigma}\varphi_{\bar{K}\sigma}\; .\nonumber\\
574: \label{eq:sourcesstandard}
575: \end{eqnarray}
576: A comparison between Eq.~(\ref{eq:JPhiscalarproduct}) and
577: Eq.~(\ref{eq:sourcesstandard}) shows that the sources in the compact
578: notation are related to the standard ones by $J=(\bar{\jmath},\zeta
579: j,J^*)$. The connected $n$-line Green's functions $\G^{(n)}_{c,
580: \alpha_1 \ldots \alpha_n} $ are then defined via the functional
581: Taylor expansion
582: \begin{equation}
583: \G_c[J]=\sum_{n=0}^{\infty}\frac1{n!}\int_{\alpha_1}\dots
584: \int_{\alpha_n} \G^{(n)}_{c, \alpha_1 \dots \alpha_n} J_{\alpha_1}
585: \cdot\ldots\cdot J_{\alpha_n}\,,
586: \label{eq:Gcexpansion}
587: \end{equation}
588: implying
589: \begin{equation}
590: \G^{(n)}_{c, \alpha_1 \dots \alpha_n} =
591: \left.
592: \frac{ \delta^{(n)} \G_c[J] }{ \delta J_{ \alpha_n} \ldots \delta J_{\alpha_1} }
593: \right|_{ J = 0 }
594: \label{eq:Gcndef}
595: \; .
596: \end{equation}
597: In particular, the exact Green's function of our interacting system is
598: given by
599: \begin{equation}
600: [ \mathbf{G} ]_{\alpha \alpha^{\prime}} =-\left.
601: \ppt{\G_c}{J_{\alpha}}{J_{\alpha^{\prime}}}\right|_{J=0}
602: = - \G^{(2)}_{c , \alpha^{\prime} \alpha }
603: \,,
604: \label{eq:Gmatrix}
605: \end{equation}
606: which we shall write in compact matrix notation as
607: \begin{equation}
608: \mathbf{G}=-\left.\ppt{\G_c}{J}{J}\right|_{J=0}
609: =
610: \left(\begin{array}{ccc}
611: 0& \hat{G} &0\\
612: \zeta \hat{G}^{T}&0&0\\
613: 0&0&-\hat{F}
614: \end{array}\right)\, .
615: \label{eq:Gmatrixcompact}
616: \end{equation}
617: For the last equality, it has been assumed that no symmetry breaking
618: occurs. Thus, $\mathbf{G}$ has the same block structure as the
619: noninteracting $\mathbf{G}_0$ in Eq.~(\ref{eq:G0matrix}) so that,
620: similarly to Eq.~(\ref{eq:G0transpose}),
621: \begin{equation}
622: \mathbf{G}^T = \mathbf{Z} \mathbf{G} = \mathbf{G} \mathbf{Z}
623: \; .
624: \label{eq:Gtranspose}
625: \end{equation}
626: In the noninteracting limit ($S_{1} \rightarrow 0$) one easily
627: verifies by elementary Gaussian integration that the matrix
628: $\mathbf{G}$ given in Eq.~(\ref{eq:Gmatrix}) reduces to $\mathbf{G}_0$,
629: as defined in Eq.~(\ref{eq:G0matrix}). The self-energy matrix also
630: has the same block structure as the inverse free propagator. Dyson's
631: equation then reads as
632: \begin{equation}
633: \mathbf{G}^{-1}=\mathbf{G}_0^{-1}-\mathbf{\Sigma}\,,
634: \label{eq:Dyson}
635: \end{equation}
636: where the matrix ${\bf \Sigma}$ contains the one-fermion-line
637: irreducible self-energy $\Sigma_{\sigma} ( K ) $ and the
638: one-interaction-line irreducible polarization $\Pi_{\sigma } ( \bar{K}
639: )$ in the following blocks:
640: \begin{equation}
641: \mathbf{\Sigma} = \left(\begin{array}{ccc}
642: 0&\zeta [\hat{\Sigma}]^T&0\\
643: \hat{\Sigma}&0&0\\
644: 0&0&\hat{\Pi}
645: \end{array}\right)\,,
646: \end{equation}
647: where
648: \begin{eqnarray}
649: [\hat{\Sigma} ]_{K\sigma,K'\sigma'}&=& \delta_{K,K'}\delta_{\sigma \sigma'}
650: \Sigma_{\sigma} ( K' ) \,,
651: \label{eq:Sigmairdef} \\{}
652: [\hat{\Pi}]_{\bar{K}\sigma,\bar{K}'\sigma'}&=&
653: \delta_{\bar{K} + \bar{K}', 0 } \delta_{\sigma \sigma'} \,\, \Pi_{\sigma}
654: ( \bar{K}' ) \,.
655: \label{eq:Piirdef}
656: \end{eqnarray}
657: These matrices are spin-diagonal because the bare coupling $S_1
658: [\bar{\psi}, \psi , \varphi ]$ between Fermi and Bose fields in
659: Eq.~(\ref{eq:Spsiphi}) is diagonal in the spin index. The blocks of
660: the full Green's function matrix $\mathbf{G}$ in
661: Eq.~(\ref{eq:Gmatrixcompact}) contain the exact single-particle Green's
662: function and the effective (screened) interaction,
663: \begin{eqnarray}
664: [\hat{G} ]_{K\sigma,K'\sigma'}&=& \delta_{K,K'}\delta_{\sigma \sigma'}
665: G_{\sigma} ( K ) \,,
666: \label{eq:Gfulldef} \\{}
667: [\hat{F}]_{\bar{K}\sigma,\bar{K}'\sigma'}&=&
668: \delta_{\bar{K} + \bar{K}', 0 } \,\, F_{\sigma \sigma^{\prime}}
669: ( \bar{K} ) \,,
670: \label{eq:Ffulldef}
671: \end{eqnarray}
672: with
673: \begin{eqnarray}
674: G_{\sigma} ( K ) & = & [ G_{0, \sigma}^{-1} ( K ) - \Sigma_{\sigma} ( K ) ]^{-1}
675: \; ,
676: \label{eq:Gdiagdef}
677: \\
678: F_{\sigma \sigma^{\prime}} ( \bar{K} ) & = &
679: \left[ \hat{F}_{0}^{-1} + \hat{\Pi} \right]^{-1}_{\bar{K} \sigma , -\bar{K} \sigma^{\prime} }
680: \; .
681: \label{eq:Fdiagdef}
682: \end{eqnarray}
683:
684: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
685: \subsubsection{Generating functional of one-line irreducible vertices}
686:
687: Below, we shall derive exact functional RG equations for the one-line
688: irreducible vertices of our coupled Fermi-Bose theory. The
689: diagrammatic perturbation theory consists of both fermion and boson
690: lines. We require irreducibility with respect to both types of lines
691: and call this one-line irreducibility. One should keep in mind that a
692: boson line represents the two-body electron-electron interaction which
693: is screened by zero-sound bubbles for small momentum transfers. This
694: means that in fermionic language our vertices are not only
695: one-particle irreducible but are also approximately two-particle
696: irreducible in the zero-sound channel in the sense that particle-hole
697: bubbles are eliminated in favor of the effective bosonic propagator.
698: In order to obtain the generating functional of the corresponding
699: irreducible vertices, we perform a Legendre transformation with
700: respect to all field components, introducing the classical field
701: \cite{footnotefield}
702: \begin{equation}
703: \Phi_{\alpha}=\pp{\G_c}{J_{\alpha}}
704: \label{eq:def_phi}\,.
705: \end{equation}
706: After inverting this relation for $J = J [ \Phi ]$ we may calculate
707: the Legendre effective action
708: \begin{equation}
709: \LL[\Phi] = (J[\Phi],\Phi)-\G_c[J[\Phi]]\,.
710: \label{eq:Ldef}
711: \end{equation}
712: From this we obtain
713: \begin{equation}
714: J_{\alpha} = \zeta_{\alpha}\pp{\LL}{\Phi_{\alpha}}\,,
715: \label{eq:def_J}
716: \end{equation}
717: which we may write in compact matrix notation as
718: \begin{equation}
719: J = \mathbf{Z} \pp{\LL}{\Phi}
720: \; .
721: \label{eq:def_Jcompact}
722: \end{equation}
723: In this notation the chain rule simply reads as
724: \begin{equation}
725: \frac{\delta}{\delta \Phi} =
726: \ppt{\LL}{\Phi}{\Phi}\mathbf{Z} \frac{\delta}{\delta J }
727: \; .
728: \label{eq:chain}
729: \end{equation}
730: Applying this to both sides of Eq.~(\ref{eq:def_phi}) we obtain
731: \begin{equation}
732: \mathbf{1}=\pp{\Phi}{\Phi}=\ppt{\LL}{\Phi}{\Phi}\mathbf{Z}\ppt{\G_c}{J}{J}\,.
733: \label{eq:quad_rel}
734: \end{equation}
735: For vanishing fields $\Phi$ and $J$ this yields
736: \begin{equation}
737: \left.\ppt{\LL}{\Phi}{\Phi}\right|_{\Phi=0}=-\mathbf{Z}\mathbf{G}^{-1}
738: =- [\mathbf{G}^{-1}]^T
739: \; .
740: \label{eq:LGcrelation}
741: \end{equation}
742: The advantage of our compact notation is now obvious: the minus signs
743: associated with the Grassmann fields can be neatly collected in the
744: ``statistics matrix'' $\mathbf{Z}$. If the Grassmann sources are
745: introduced in the conventional way,\cite{Negele88} the minus signs
746: generated by commuting two Grassmann fields are distributed in a more
747: complicated manner in the matrices of
748: second derivatives.\cite{Kopietz01,Correia01}
749:
750: From Eq.~(\ref{eq:LGcrelation}) it is evident that we need to subtract
751: the free action from $\LL [\Phi ]$ to obtain the generating functional
752: for the irreducible vertex functions,
753: \begin{equation}
754: \Gamma[\Phi]= \LL[\Phi]-S_0[\Phi]=\LL[\Phi]+\frac12(\Phi,[\Go^{-1}]\Phi)\,.
755: \label{eq:Gammadef}
756: \end{equation}
757: Then we have, using the Dyson equation (\ref{eq:Dyson}),
758: \begin{eqnarray}
759: \left.\ppt{\Gamma}{\Phi}{\Phi}\right|_{\Phi=0} & = &
760: \left.\ppt{\LL}{\Phi}{\Phi}\right|_{\Phi=0}
761: + [ \mathbf{G}_0^{-1} ]^T
762: \nonumber
763: \\
764: & = &
765: - [ \mathbf{G}^{-1} ]^T + [ \mathbf{G}_0^{-1} ]^T
766: = \mathbf{\Sigma}^T
767: \; .
768: \label{eq:Sigmarelation}
769: \end{eqnarray}
770: In general, the one-line irreducible vertices are defined as
771: coefficients in an expansion of $\Gamma [ \Phi ]$ with respect to the
772: fields,
773: \begin{equation}
774: \Gamma[\Phi]=\sum_{n=0}^{\infty}\frac1{n!}\int_{\alpha_1}\dots
775: \int_{\alpha_n}\Gamma^{(n)}_{\alpha_1,\dots,\alpha_n}\Phi_{\alpha_1}
776: \cdot\ldots\cdot\Phi_{\alpha_n}\,.
777: \label{eq:Gammaexpansion}
778: \end{equation}
779: The vertices $\Gamma^{(n)}$ have the same symmetry with respect to
780: interchange of the indices as the monomial in the fields, i.e., the
781: interchange of two neighboring Fermi fields yields a minus sign.
782: Graphically, we represent the vertices $\Gamma^{(n)}$
783: %
784: %
785: \begin{figure}
786: \begin{center}
787: \epsfig{file=generalvertex.eps,width=0.7\hsize}
788: \end{center}
789: \caption{Graphical representation of the symmetrized one-line irreducible
790: $n$-point vertex defined via Eq.~(\ref{eq:Gammaexpansion}).
791: Because for fermions the order of the indices is important
792: (exchanging two neighboring fermion legs will generate a minus
793: sign) the circles representing the irreducible vertices have an
794: arrow that points to the leg corresponding to the first index;
795: subsequent indices are arranged in the order indicated by the
796: arrow. External legs denote either outgoing fermions
797: ($\bar{\psi}$), incoming fermions ($\psi$), or bosons ($\varphi$).
798: }
799: \label{fig:generalvertex}
800: \end{figure}
801: %
802: %
803: by an oriented circle with $n$ external legs, as shown in
804: Fig.~\ref{fig:generalvertex}. With the
805: definition~(\ref{eq:Gammaexpansion}) and Eq.~(\ref{eq:Sigmarelation})
806: we have $\Gamma^{(2)}_{\alpha\alpha'}= [ \mathbf{\Sigma}]_{\alpha'
807: \alpha}$. The fact that also the higher-order vertices
808: $\Gamma^{(n)}$ defined in Eq.~(\ref{eq:Gammaexpansion}) are indeed
809: one-line irreducible (i.e., cannot be separated into two parts by
810: cutting a single fermion line or a single interaction line) can be
811: shown iteratively by generating a tree expansion from higher-order
812: derivatives of Eq.~(\ref{eq:quad_rel}). We show this explicitly in
813: Appendix~A.
814:
815: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
816: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
817: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
818: \section{Functional RG flow equations for one-line irreducible vertices}
819: \label{sec:flow_eq}
820:
821: In this section we derive exact RG flow equations for the generating
822: functional of the one-line irreducible vertices of our coupled
823: Fermi-Bose theory. We also classify the various vertices of the theory
824: based on their scaling dimensions and propose a new truncation scheme
825: involving the building blocks of the skeleton diagrams for fermionic
826: and bosonic two-point functions.
827:
828: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
829: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
830: \subsection{Cutoff schemes}
831:
832: Since the interaction now appears as a propagator of the field
833: $\varphi$, it is possible to introduce a momentum-transfer cutoff in
834: the interaction on the same footing as a bandwidth cutoff. A bandwidth
835: cutoff restricts the relevant fermionic degrees of freedom to the
836: vicinity of the Fermi surface, and appears to be most natural in the
837: RG approach to fermions in one spatial dimension.\cite{Solyom79} In
838: higher dimensions, the Wilsonian idea of eliminating the degrees of
839: freedom in the vicinity of the Fermi surface is implemented by
840: defining for each momentum ${\bf{k}}$ an associated ${\bf{k}}_F$ by
841: means of a suitable projection onto the Fermi surface \cite{Kopietz01}
842: and then integrating over fields with momenta in the energy shell $
843: v_0 \Lambda < | \epsilon_{ {\bf{k}} } - \epsilon_{ {\bf{k}}_F } | <
844: v_0 \Lambda_0$, where $\epsilon_{\bf{k}} $ is the energy dispersion in
845: the absence of interactions. Here $v_0$ is some suitably defined
846: velocity (for example, some average Fermi velocity), which we introduce
847: to give $\Lambda$ units of momentum. We shall refer to $v_0 \Lambda$
848: as bandwidth cutoff. Formally, we introduce such a cutoff into our
849: theory by substituting for the free fermionic Green's function in
850: Eq.~(\ref{eq:G0def})
851: \begin{eqnarray}
852: G_{0, \sigma} (K ) &\longrightarrow& \Theta( \Lambda< D_K < \Lambda_0)\,
853: G_{0, \sigma} (K )\nonumber\\
854: &&= \frac{\Theta(\Lambda< D_K < \Lambda_0)}{i\omega - \xi_{{\bf k}\sigma}}
855: \label{eq:bandcutoff}
856: \; ,
857: \end{eqnarray}
858: with
859: \begin{equation}
860: D_K = |\epsilon_{\bf{k}} - \epsilon_{{\bf{k}}_F}| /v_0
861: \; .
862: \label{eq:OmegaKdef}
863: \end{equation}
864: Here $ \Theta ( \Lambda < x < \Lambda_0 ) = 1$ if the logical
865: expression in the brackets is true, and $ \Theta ( \Lambda < x <
866: \Lambda_0 ) = 0$ if the logical expression is false. Ambiguities
867: associated with the sharp $\Theta$-function cutoff can be avoided by
868: smoothing out the $\Theta$ functions and taking the sharp cutoff limit
869: at the end of the calculation.\cite{Morris94} In order to construct
870: a consistent scaling theory, the ${\bf{k} }_F$ in
871: Eq.~(\ref{eq:OmegaKdef}) should refer to the true Fermi surface of the
872: interacting system, which can in principle be obtained
873: self-consistently from the condition that the RG flows into a fixed
874: point.\cite{Kopietz01,Ledowski03}
875:
876: The above bandwidth-cutoff procedure has several disadvantages. On
877: the one hand, for any finite value of the cutoff parameter
878: $v_0\Lambda$ the Ward identities are violated.\cite{Katanin04}
879: Moreover, the RG flow of two-particle response functions probing the
880: response at small momentum transfers (such as the compressibility or
881: the uniform magnetic susceptibility) is artificially suppressed by the
882: bandwidth cutoff. To cure the latter problem, various other
883: parameters have been proposed to serve as a cutoff for the RG, such as
884: the temperature \cite{Honerkamp01b} or even the strength of the
885: interaction.\cite{Honerkamp04} While for practical calculations these
886: new cutoff schemes may have their advantages, the intuitively
887: appealing RG picture that the coarse grained parameters of the
888: renormalized theory contain the effect of the degrees of freedom at
889: shorter length scales and higher energies gets somewhat blurred (if
890: not completely lost) by these new schemes.
891:
892: The above mentioned problems can be elegantly avoided in our mixed
893: Fermi-Bose theory if we work with a momentum cutoff in the bosonic
894: sector of our theory, which amounts to replacing in
895: Eq.~(\ref{eq:F0def}),
896: \begin{eqnarray}
897: F_{0, \sigma \sigma^{\prime} } (\bar{K})
898: &\longrightarrow& \Theta(\Lambda< \bar{D}_{\bar{K}}
899: <\Lambda_0)\,F_{0 , \sigma \sigma^{\prime}} (\bar{K})
900: \nonumber\\
901: &&=\Theta(\Lambda< \bar{D}_{\bar{K}}
902: <\Lambda_0)\,f^{\sigma\sigma^{\prime}}_{\bar{\bf k}}
903: \; ,
904: \label{eq:momentumtransfercutoff}
905: \end{eqnarray}
906: where
907: \begin{equation}
908: \bar{D}_{\bar{K}} = |\bar{\bf{k}}|
909: \; .
910: \label{eq:DKbardef}
911: \end{equation}
912: Keeping in mind that the bosonic field mediates the effective
913: interaction, it is clear that $\Lambda$ is a cutoff for the momentum
914: transfer of the interaction. This is precisely the same cutoff scheme
915: employed in the seminal work by Hertz,\cite{Hertz76} who discussed
916: also more general frequency-dependent cutoffs for the labels of the
917: bosonic Hubbard-Stratonovich fields, corresponding to more complicated
918: functions $\bar{D}_{\bar{K}}$ than the one given in
919: Eq.~(\ref{eq:DKbardef}). Moreover, in the exact solution of the
920: one-dimensional Tomonaga-Luttinger model (abbreviated here as TLM, as
921: already defined above) by means of a careful application of the
922: bosonization method \cite{Schoenhammer03} the maximal momentum
923: transfered by the interaction appears as the natural cutoff scale.
924:
925: In our RG approach we have the freedom of choosing both the bandwidth
926: cutoff $v_0 \Lambda $ and the momentum-transfer cutoff $\Lambda$
927: independently. In particular, we may even choose to get rid of the
928: bandwidth cutoff completely and work with a momentum-transfer cutoff
929: only. In this work, we shall show that if the interaction involves
930: only small momentum transfers, then the pure momentum-transfer cutoff
931: scheme indeed regularizes all infrared singularities in one dimension.
932: Moreover and most importantly, introducing a cutoff only in the
933: momentum transfer leads to exact RG flow equations that do not
934: violate the Ward identities responsible for the exact solubility of
935: the TLM. Given this fact, it is not surprising that we can solve the
936: infinite hierarchy of RG flow equations exactly and obtain the exact
937: single-particle Green's function of the TLM within the framework of the
938: functional RG.
939:
940: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
941: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
942: \subsection{Flow equations for completely symmetrized vertices}
943:
944: With the substitutions (\ref{eq:bandcutoff}) and
945: (\ref{eq:momentumtransfercutoff}) the noninteracting Green's function
946: $\mathbf{G}_0$, and hence all generating functionals, depend on the
947: cutoff parameter $\Lambda$. We can now follow the evolution of the
948: generating functionals as we change the cutoff. The differentiation of
949: Eq.~(\ref{eq:Ggen}) with respect to $\Lambda$ yields for the
950: generating functional of the Green's functions,
951: \begin{equation}
952: \partial_{\Lambda}\G = \left\{
953: \frac12\left(
954: \p{J},\partial_{\Lambda}[\Go^{-1}]\p{J}\right)
955: -\partial_{\Lambda}\ln\Z_0
956: \right\}\G\,.
957: \end{equation}
958: For the connected version $\G_c [ J ] = \ln \G [ J ]$, we obtain
959: \begin{eqnarray}
960: \partial_{\Lambda}\G_c =
961: \frac12\left(
962: \pp{\G_c}{J},\partial_{\Lambda}[\Go^{-1}]\pp{\G_c}{J}\right)
963: ~~~~~~~~~~~~~~~~~~~~~~~&&\nonumber\\
964: +\frac12\Tr\left(\partial_{\Lambda}[\Go^{-1}]\left[\ppt{\G_c}{J}{J}\right]^T\right)
965: -\partial_{\Lambda}\ln\Z_0\,.&&
966: \label{eq:Gcflow}
967: \end{eqnarray}
968: In the derivation of flow equations for $\LL$ or $\Gamma$ [see
969: Eqs.~(\ref{eq:Ldef}) and (\ref{eq:Gammadef})], we should keep in mind
970: that in these functionals the fields $\Phi$ are held constant rather
971: than the sources $J$. Hence, Eq.~(\ref{eq:Ldef}) implies
972: \begin{equation}
973: \partial_{\Lambda}\LL[\Phi]=-\left.\partial_{\Lambda}\G_c[J]\right|_{J=J_{\Lambda}[\Phi]}
974: \; .
975: \end{equation}
976: Using this and Eq.~(\ref{eq:Gcflow}) we obtain for the functional
977: $\Gamma [\Phi ] = \LL [\Phi] - S_0 [\Phi]$,
978: \begin{eqnarray}
979: \partial_{\Lambda} \Gamma =
980: & - & \frac12\Tr\left(\partial_{\Lambda}[\Go^{-1}]\left[\ppt{\G_c}{J}{J}\right]^T\right)
981: + \partial_{\Lambda}\ln\Z_0 \; .
982: \nonumber
983: \\
984: & &
985: \label{eq:Gammaflowprelim}
986: \end{eqnarray}
987: To derive a closed equation for $\Gamma$, we express the matrix
988: $\ppt{\G_c}{J}{J}$ in terms of derivatives of $\Gamma$ using
989: Eq.~(\ref{eq:GcJJexpansion}). After some rearrangements we obtain the
990: exact flow equation for the generating functional $\Gamma [\Phi ] $ of
991: the one-line irreducible vertices,
992: \begin{eqnarray}
993: \partial_{\Lambda}\Gamma&=&
994: -\frac12\Tr\left[\mathbf{Z}\dot{\mathbf{G}}^T\mathbf{U}^T
995: \left\{
996: \mathbf{1}-\mathbf{G}^T\mathbf{U}^T
997: \right\}^{-1}\right]\nonumber\\[0.2cm]
998: &&-\frac12\Tr\left[\mathbf{Z}\dot{\mathbf{G}}_0^T\mathbf{\Sigma}^T
999: \left\{
1000: \mathbf{1}-\mathbf{G}_0^T\mathbf{\Sigma}^T
1001: \right\}^{-1}\right]\,,
1002: \label{eq:Gammaflow}
1003: \end{eqnarray}
1004: %
1005: \begin{figure}[t]
1006: \begin{center}
1007: \epsfig{file=twopoint.eps,width=1.0\hsize}
1008: \end{center}
1009: \caption{Graphical representation of Eq.~(\ref{eq:flow_vert}) for $n=2$, describing
1010: the flow of the totally symmetric two-point vertex. Empty circles
1011: with ${\bf G}$ and $\dot{\bf G}$ denote the exact matrix propagator
1012: ${\bf G}$ and the single-scale propagator $\dot{\bf G}$ defined in
1013: Eqs.~(\ref{eq:Gmatrix}) and (\ref{eq:Gdotmatrix}), respectively.}
1014: \label{fig:flowGamma2}
1015: \end{figure}%
1016: %
1017: where the matrix $\mathbf{U} [ \Phi ] $ is the field-dependent part of
1018: the second functional derivative of $\Gamma [\Phi]$, as defined in
1019: Eq.~(\ref{eq:Udef}). For convenience we have introduced the
1020: single-scale propagator $\dot{\mathbf{G}}$ as
1021: \begin{equation}
1022: \dot{\mathbf{G}} = - \mathbf{G}\partial_{\Lambda}[\mathbf{G}_0^{-1}]\mathbf{G}
1023: =[\mathbf{1}-\mathbf{G}_0\mathbf{\Sigma}]^{-1}
1024: \left(\partial_{\Lambda}\mathbf{G}_0
1025: \right)
1026: [\mathbf{1}-\mathbf{\Sigma}\mathbf{G}_0]^{-1}
1027: \; ,
1028: \label{eq:Gdotmatrix}
1029: \end{equation}%
1030: which reduces to $\dot{\mathbf{G}}_0 = \partial_{\Lambda}
1031: \mathbf{G}_0$ in the absence of interactions. The matrix
1032: $\dot{\mathbf{G}}$ has the same block structure as the matrix
1033: $\mathbf{G}$ in Eq.~(\ref{eq:Gmatrixcompact}). We denote the
1034: corresponding blocks by $\dot{\hat{G}}$ and $\dot{\hat{F}}$. In the
1035: limit of a sharp $\Theta$-function cutoff~\cite{Morris94} the blocks
1036: of the single scale propagator are explicitly given by
1037: \begin{eqnarray}
1038: [\dot{\hat{G} }]_{K\sigma,K'\sigma'}&=& \delta_{K,K'}\delta_{\sigma \sigma'}
1039: \dot{G}_{\sigma} ( K ) \,,
1040: \label{eq:Gsinglescale} \\{}
1041: [\dot{\hat{F}}]_{\bar{K}\sigma,\bar{K}'\sigma'}&=&
1042: \delta_{\bar{K} + \bar{K}', 0 } \,\, \dot{F}_{\sigma \sigma^{\prime}}
1043: ( \bar{K} ) \,,
1044: \label{eq:Fsinglescale}
1045: \end{eqnarray}
1046: with
1047: \begin{eqnarray}
1048: \hspace{-4mm} \dot{G}_{\sigma} ( K ) & = & - \frac{ \delta ( \Lambda - D_K ) }{i \omega
1049: - \xi_{ {\bf{k}} \sigma} - \Sigma_{\sigma} ( K ) }
1050: \; ,
1051: \label{eq:dotGdiagdef}
1052: \\
1053: \hspace{-4mm} \dot{F}_{\sigma \sigma^{\prime}} ( \bar{K} ) & = & -
1054: \delta ( \Lambda - \bar{D}_{\bar{K}} )
1055: \left[ \hat{F}_{0}^{-1} + \hat{\Pi} \right]^{-1}_{ \bar{K} \sigma , -\bar{K} \sigma^{\prime} }
1056: \; ,
1057: \label{eq:dotFdiagdef}
1058: \end{eqnarray}
1059: where on the right-hand side of Eq.~(\ref{eq:dotFdiagdef}) it is
1060: understood that the $\Theta$-function cutoff should be omitted from
1061: the matrix elements of $\hat{F}_0$.
1062:
1063: The second line in Eq.~(\ref{eq:Gammaflow}) does not depend on the
1064: fields any longer and therefore represents the flow of the interaction
1065: correction $\Gamma^{(0)}$ to the grand-canonical potential,
1066: \begin{equation}
1067: \partial_{\Lambda} \Gamma^{(0)} =
1068: -\frac12\Tr\left[\mathbf{Z}\dot{\mathbf{G}}_0^T\mathbf{\Sigma}^T
1069: \left\{
1070: \mathbf{1}-\mathbf{G}_0^T\mathbf{\Sigma}^T
1071: \right\}^{-1}\right]\,.
1072: \label{eq:Gamma0flow}
1073: \end{equation}
1074: Since we have already dropped constant parts of the action in the
1075: Hubbard-Stratonovich transformation, we will not keep track of
1076: $\Gamma^{(0)}$ in the following.
1077:
1078: The first line on the right-hand side of Eq.~(\ref{eq:Gammaflow})
1079: gives the flow of one-line irreducible vertices. We can generate a
1080: hierarchy of flow equations for the vertices by expanding both sides
1081: in powers of the fields. On the left-hand side, we simply insert the
1082: functional Taylor expansion (\ref{eq:Gammaexpansion}) of $\Gamma [
1083: \Phi ]$, while on the right-hand side we substitute the expansion of
1084: $\mathbf{U} [ \Phi ]$ given in Eq.~(\ref{eq:Uexpansion}). For a
1085: comparison of the coefficients on both sides, the right-hand side has
1086: to be symmetrized with respect to external lines on different
1087: vertices. We can write down the resulting infinite system of flow
1088: equations for the one-line irreducible vertices $\Gamma^{(n)}$ with $n
1089: \geq 1$ in the following closed form:
1090: \begin{widetext}
1091: \begin{eqnarray} \partial_{\Lambda} {\Gamma}^{(n)}_{\alpha_1,\dots,\alpha_n} =
1092: -\frac12\sum\limits_{l=1}^{\infty}
1093: \sum\limits_{m_1,\dots,m_l=1}^{\infty}\delta_{n,m_1+\ldots+m_l}\,
1094: {\cal{S}}_{\alpha_1,\dots,\alpha_{m_1};\alpha_{m_1+1},\dots,\alpha_{m_1+m_2};\dots;\alpha_{m_1+\ldots+m_{l-1}+1},\dots,\alpha_{n}}
1095: \Big\{~~~~~~~~&&\nonumber\\ \times \Tr\left[
1096: \mathbf{Z}\dot{\mathbf{G}}^T \mathbf{\Gamma}^{(m_1+2)\,T}_{\alpha_1,\dots,\alpha_{m_1}}\mathbf{G}^T
1097: \mathbf{\Gamma}^{(m_2+2)\,T}_{\alpha_{m_1+1},\dots,\alpha_{m_1+m_2}}\mathbf{G}^T
1098: \dots
1099: \mathbf{\Gamma}^{(m_l+2)\,T}_{\alpha_{m_1+\ldots+m_{l-1}+1},\dots,\alpha_{n}}
1100: \right] \Big\}\,.
1101: \label{eq:flow_vert}
1102: \end{eqnarray}
1103: \end{widetext}
1104: Here the matrices $\mathbf{\Gamma}^{(m+2)}_{\alpha_1 \ldots \alpha_m}$
1105: are given in Eq.~(\ref{eq:Gammamatrix}) and the symmetrization
1106: operator ${\cal{S}}$ is defined in Eq.~(\ref{eq:symmopdef}). The
1107: effect of ${\cal{S}}$ is rather simple: it acts on an expression
1108: already symmetric in the index groups separated by semicolons to
1109: generate an expression symmetric also with respect to the exchange of
1110: indices between different groups. From one summand in
1111: Eq.~(\ref{eq:flow_vert}) the symmetrization operator ${\cal{S}}$ thus
1112: creates $n!/(m_1!\cdot\ldots\cdot m_l!)$ terms.
1113: %
1114: \begin{figure}[t]
1115: \begin{center}
1116: \epsfig{file=threepoint.eps,width=1.0\hsize}
1117: \end{center}
1118: \caption{Graphical representation of Eq.~(\ref{eq:flow_vert}) for $n=3$, describing
1119: the flow of the totally symmetric three-point vertex.}
1120: \label{fig:flowGamma3}
1121: \end{figure}
1122: %
1123:
1124: Figures.~\ref{fig:flowGamma2} and \ref{fig:flowGamma3} show a graphical
1125: representation of the flow of the vertices $\Gamma^{(2)}$ and
1126: $\Gamma^{(3)}$. With the graphical notation for the totally symmetric
1127: vertices introduced in Fig.~\ref{fig:generalvertex} all the signs and
1128: combinatorics have a graphical representation. In the next section
1129: we will leave the shorthand notation and go back to more physical
1130: vertices, explicitly exhibiting the different types of fields. All
1131: this can be done on a graphical level and involves only
1132: straightforward combinatorics. In this sense the derivation of higher
1133: flow equations is at the same level of complexity as ordinary Feynman
1134: graph expansions.
1135:
1136:
1137: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1138: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1139: \subsection{Flow equations for physical vertices}
1140: \label{subsec:flowphysical}
1141:
1142: %
1143: %
1144: %
1145: \begin{figure}[b]
1146: \begin{center}
1147: \epsfig{file=dictionary.eps,width=0.8\hsize}
1148: \vspace{0.2cm}
1149: \caption{Pictorial
1150: dictionary to translate graphs involving totally symmetrized
1151: vertices to ones involving physical vertices, which are only
1152: symmetrized within fields of the same type. The diagrams on the
1153: right-hand sides of the last four lines represent $G$,
1154: $\dot{G}$, $F$ and $\dot{F}$ respectively. }
1155: \label{fig:picdic}
1156: \end{center}
1157: \end{figure}
1158: %
1159: %
1160: %
1161: Usually, the generating functional $\Gamma [ \bar{\psi}, \psi ,
1162: \varphi ]$ is expanded in terms of correlation functions that are not
1163: symmetrized with respect to the exchange of legs involving different
1164: types of fields. If we explicitly display momentum and frequency
1165: conservation, such an expansion reads as
1166: \begin{widetext}
1167: \begin{eqnarray}
1168: \Gamma[\bar{\psi}, \psi,\varphi] & =& \sum_{n=0}^{\infty}\sum_{m=0}^{\infty}
1169: \frac{1}{(n!)^2m!}
1170: \int_{K_1'\sigma_1'}\dots\int_{K_n'\sigma_n'}
1171: \int_{K_1\sigma_1}\dots\int_{K_n\sigma_n}
1172: \int_{\bar{K}_1\bar{\sigma}_1}\dots\int_{\bar{K}_m\bar{\sigma}_m}
1173: \delta_{ K'_1 + \ldots + K'_n , K_1 + \ldots + K_n + \bar{K}_1 + \ldots + \bar{K}_m }
1174: \nonumber \\
1175: & & \hspace{20mm} \times
1176: \Gamma^{(2n,m)}(K'_1\sigma_1',\dots,K'_n\sigma_n';
1177: K_1\sigma_1,\dots,
1178: K_n\sigma_n;\bar{K}_1\bar{\sigma}_1,\dots,\bar{K}_m\bar{\sigma}_m) \nonumber\\
1179: & & \hspace{20mm} \times
1180: \bar{\psi}_{K'_1\sigma_1'}\cdot\ldots\cdot\bar{\psi}_{K'_n\sigma_n'}
1181: \psi_{K_1\sigma_1}\cdot\ldots\cdot\psi_{K_n\sigma_n}
1182: \varphi_{\bar{K}_1\bar{\sigma}_1}\cdot\ldots\cdot\varphi_{\bar{K}_m\bar{\sigma}_m}\,.
1183: \label{eq:expansion2}
1184: \end{eqnarray}
1185: \end{widetext}
1186: Diagrammatically, we represent a physical vertex $\Gamma^{(2n,m)}$
1187: involving $2n$ external fermion legs and $m$ external boson legs by a
1188: triangle to emphasize that our theory contains three types of fields,
1189: see Fig.~\ref{fig:picdic}.
1190: We represent a leg associated with a $\bar{\psi}$ field by an arrow
1191: pointing outward, a leg for $\psi$ by an arrow pointing inward, and
1192: a leg for $\varphi$ with a wiggly line without an arrow. Recall that
1193: our Bose field is real because it couples to the density, so that it
1194: should be represented graphically by an undirected line. On the
1195: contrary, for a propagator $\mathbf{G}$ or $\dot{\mathbf{G}}$, the
1196: field $\bar{\psi}$ is represented by an arrow pointing inward and
1197: $\psi$ by an arrow pointing outward. Apart from the energy- and
1198: momentum-conserving delta function, the totally symmetric vertices
1199: defined by the expansion (\ref{eq:Gammaexpansion}) coincide with the
1200: nonsymmetric ones in Eq.~(\ref{eq:expansion2}) for the same order of
1201: the indices. We can therefore obtain the flow equations for the
1202: nonsymmetric vertices by choosing a definite realization of the
1203: external legs and by carrying out the intermediate sums over the
1204: different field species, i.e., by drawing all possible lines in the
1205: intermediate loop (two possible orientations of solid lines or one
1206: wiggly line). On the right-hand side one then has to appropriately
1207: order all the legs on the vertices, keeping track of signs for the
1208: interchange of two neighboring fermion legs. Having done so, we can
1209: use the pictorial dictionary in Fig.~\ref{fig:picdic} to obtain
1210: diagrams involving the physical correlation functions. In this way,
1211: we obtain from the diagram for the completely symmetric two-point
1212: vertex shown in Fig.~\ref{fig:flowGamma2} the diagram for the
1213: fermionic self-energy in Fig.~\ref{fig:flowSigma} as well as the
1214: diagram for the irreducible polarization shown in
1215: Fig.~\ref{fig:flowPi}.
1216: %
1217: %
1218: \begin{figure}
1219: \begin{center}
1220: \epsfig{file=Sigmaflow.eps,width=0.9\hsize}
1221: \end{center}
1222: \caption{Flow of the irreducible fermionic self-energy.
1223: The diagrams are obtained from the diagrams shown in
1224: Fig.~\ref{fig:flowGamma2} by specifying the external legs to be
1225: one outgoing and one incoming fermion leg.}
1226: \label{fig:flowSigma}
1227: \end{figure}
1228: %
1229: %
1230: \begin{figure}
1231: \begin{center}
1232: \epsfig{file=Piflow.eps,width=0.9\hsize}
1233: \end{center}
1234: \caption{Flow of the irreducible polarization, obtained from the
1235: totally symmetric diagram in Fig.~\ref{fig:flowGamma2} by setting
1236: both external legs equal to boson legs. Note that each closed
1237: fermion loop gives rise to an additional factor of $\zeta=-1$.}
1238: \label{fig:flowPi}
1239: \end{figure}
1240: %
1241: %
1242: %
1243: %
1244: \begin{figure}
1245: \begin{center}
1246: \epsfig{file=vcorrflow.eps,width=0.9\hsize}
1247: \end{center}
1248: \caption{Flow of three-legged vertex with two fermion legs and one boson leg, obtained as a special case
1249: of the diagram in Fig.~\ref{fig:flowGamma3}. }
1250: \label{fig:flowVCorr}
1251: \end{figure}
1252: %
1253: %
1254: Moreover, if we specify the external legs in the diagram for the
1255: completely symmetric three-legged vertex shown in
1256: Fig.~\ref{fig:flowGamma3} to be two fermion legs and one boson leg, we
1257: obtain the flow equation for the three-legged vertex shown in
1258: Fig.~\ref{fig:flowVCorr}.
1259:
1260: The flow equation for the vertex correction in
1261: Fig.~\ref{fig:flowVCorr} looks very complicated, so that at this point
1262: the reader might wonder how in one dimension we will be able to obtain
1263: the exact solution of the TLM using our approach. We shall explain
1264: this in detail in Sec.~\ref{sec:1dflow}, but let us anticipate here
1265: the crucial step: obviously all diagrams shown in
1266: Figs.~\ref{fig:flowSigma}, \ref{fig:flowPi}, and \ref{fig:flowVCorr}
1267: can be subdivided into two classes: those involving a fermionic
1268: single-scale propagator (the slash appears on an internal fermion
1269: line), and those with a bosonic single-scale propagator (with a slash
1270: on an internal boson line). At this point we have not specified the
1271: cutoff procedure, but as mentioned in the Introduction, in
1272: Sec.~\ref{sec:1dflow} we shall work with a bosonic cutoff only. In
1273: this case all diagrams with a slash attached to a fermion line should
1274: simply be omitted. This is the crucial simplification which will allow
1275: us to solve the hierarchy of flow equations exactly. Let us proceed
1276: in this section without specifying a particular cutoff procedure.
1277:
1278:
1279: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1280: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1281: \subsection{Rescaling and classification of vertices}
1282: \label{subsec:rescale}
1283:
1284: In order to assign scaling dimensions to the vertices, we have to
1285: define how we rescale momenta, frequencies, and fields under the RG
1286: transformation. The rescaling is not unique but depends on the nature
1287: of the fixed point we are looking for. Let us be general here and
1288: assume that in the bosonic sector the relation between momentum and
1289: frequency is characterized by a bosonic dynamic exponent $z_\varphi$
1290: (this is the exponent $z$ introduced by Hertz\cite{Hertz76}), while
1291: in the fermionic sector the corresponding dynamic exponent is
1292: $z_{\psi}$. Rescaled dimensionless bosonic momenta $\bar{\bf{q}}$ and
1293: frequencies $\bar{\epsilon}$ are then introduced as
1294: usual\cite{Hertz76}
1295: \begin{equation}
1296: \bar{\bf{q}} = \bar{\bf{k}} / \Lambda \; \; , \; \; \bar{\epsilon} = \bar{\omega} /
1297: \bar{\Omega}_{\Lambda} \; \; ,
1298: \; \; \bar{\Omega}_{\Lambda} \propto \Lambda^{z_{\varphi}}
1299: \; .
1300: \label{eq:bosrescale}
1301: \end{equation}
1302: For convenience, we choose the factor $\bar{\Omega}_{\Lambda}$ such
1303: that it has units of energy; $\bar{\epsilon}$ is then dimensionless.
1304:
1305: The proper rescaling of the fermionic momenta is not so obvious.
1306: Certainly, all momenta should be measured with respect to suitable
1307: points ${\bf{k}}_F$ on the Fermi surface. One possibility is to
1308: rescale only the component $k_{\parallel} = ( {\bf{k}} - {\bf{k}}_F )
1309: \cdot \hat{\bf{v}}_F$ of a given momentum that is parallel to the local
1310: Fermi surface velocity ${\bf{v}}_F$ (and hence perpendicular to the Fermi
1311: surface),\cite{Shankar94,Kopietz01} where $\hat{\bf v}_F$ is a unit vector
1312: in the direction of ${\bf v}_F$. Unfortunately, in dimensions $D >
1313: 1$ this leads to rather complicated geometric constructions, because
1314: in a fixed reference frame the component $k_{\parallel}$ to be
1315: rescaled varies for different points on the Fermi surface. However,
1316: if the initial momentum-transfer cutoff $\Lambda_0$ in
1317: Eq.~(\ref{eq:momentumtransfercutoff}) is small compared with the
1318: typical radius of the Fermi surface, the initial and final momenta
1319: associated with a scattering process lie both on nearby points on the
1320: Fermi surface. In this case it seems natural to pick one fixed
1321: reference point ${\bf{k}}_{F, \sigma}$ on the Fermi surface, and then
1322: measure all fermionic momentum labels ${\bf{k}}_i$ and
1323: ${\bf{k}}_i^{\prime}$ in $\Gamma^{(2n,m)}(K'_1,\dots,K'_n; K_1,\dots,
1324: K_n;\bar{K}_1,\dots,\bar{K}_m) $ relative to this ${\bf{k}}_{F,
1325: \sigma}$. Here, the index $\sigma$ labels the different points on
1326: the Fermi surface, for example, in one dimension $\sigma = \pm 1$, with
1327: $k_{F,\pm 1} = \pm k_F$. We then define rescaled fermionic momenta
1328: ${\bf{q}}$ and frequencies ${\epsilon}$ as follows:
1329: \begin{equation}
1330: {\bf{q}} = ( {\bf{k}} - {\bf{k}}_{F, \sigma} )/ \Lambda \; \; , \; \; {\epsilon} = {\omega} /
1331: \Omega_{\Lambda} \; \; ,
1332: \; \; \Omega_{\Lambda} \propto \Lambda^{z_{\psi}}
1333: \; .
1334: \label{eq:fermirescale}
1335: \end{equation}
1336: The factor $\Omega_{\Lambda}$ should again have units of energy such
1337: that $\epsilon$ is dimensionless. Iterating the usual RG steps
1338: consisting of mode elimination and rescaling, we then coarse grain the
1339: degrees of freedom in a sphere around the chosen point ${\bf{k}}_{F ,
1340: \sigma}$. Because by assumption the maximal momentum transfer
1341: mediated by the interaction is small compared with $ | {\bf{k}}_{F,
1342: \sigma} |$, the fermionic momenta appearing in
1343: $\Gamma^{(2n,m)}(K'_1,\dots,K'_n; K_1,\dots,
1344: K_n;\bar{K}_1,\dots,\bar{K}_m) $ are all in the vicinity of the chosen
1345: ${\bf{k}}_{F , \sigma}$. This property is also responsible for the
1346: approximate validity of the {\it{closed loop theorem}} for interacting
1347: fermions with dominant forward scattering in arbitrary
1348: dimensions.\cite{Kopietz95,Kopietz97,Metzner98}
1349:
1350: Apart from the rescaling of momenta and frequencies, we have to
1351: specify the rescaling of the fields. As usual, we require that the
1352: Gaussian part $S_0 [ \Phi ] = S_0 [ \bar\psi , \psi ] + S_0 [\varphi]$
1353: of our effective action is invariant under rescaling. For the
1354: fermionic part this is achieved by defining renormalized fields
1355: $\tilde{\psi}_{Q\sigma}$ in $D$ dimensions via
1356: \begin{equation}
1357: \psi_{K \sigma} =
1358: \left(
1359: \frac{ Z }{ \Lambda^D \Omega_{\Lambda}^2 } \right)^{1/2} \tilde{\psi}_{ Q \sigma }
1360: \; ,
1361: \label{eq:Fermifieldrescale}
1362: \end{equation}
1363: where $Z$ is the fermionic wave-function renormalization factor and $Q
1364: = ( {\bf{q}} , i \epsilon )$ denotes the rescaled fermionic momenta
1365: and Matsubara frequencies as defined in Eq.~(\ref{eq:fermirescale}).
1366: With this rescaling the wave-function renormalization and the Fermi
1367: velocity have a vanishing scaling dimension (corresponding to marginal
1368: couplings), while the momentum- and frequency-independent part of the
1369: self-energy is relevant with scaling dimension $+1$; see
1370: Ref.~\onlinecite{Kopietz01}. Analogously, we find that the bosonic Gaussian
1371: part of the action is invariant under rescaling if we express it in
1372: terms of the renormalized bosonic field $\tilde{\varphi}_{\bar{Q}
1373: \sigma }$ defined by
1374: \begin{equation}
1375: \varphi_{\bar{K} \sigma} =
1376: \left(
1377: \frac{ \bar{Z} }{ \Lambda^D \bar{\Omega}_{\Lambda} \nu_0 } \right)^{1/2}
1378: \tilde{\varphi}_{ \bar{Q} \sigma }
1379: \; ,
1380: \label{eq:Bosefieldrescale}
1381: \end{equation}
1382: where $\bar{Z}$ is the bosonic wave-function renormalization factor,
1383: $\bar{Q} = ( \bar{\bf{q}} , i \bar{\epsilon} )$ denotes the rescaled
1384: bosonic momenta and Matsubara frequencies defined in
1385: Eq.~(\ref{eq:bosrescale}), and $\nu_0$ denotes the noninteracting
1386: density of states at the Fermi surface. We introduce the factor of
1387: $\nu_0$ for convenience to make all rescaled vertices dimensionless.
1388: By construction Eq.~(\ref{eq:Bosefieldrescale}) assigns vanishing
1389: scaling dimensions to the bare interaction parameters $ f_{
1390: \bar{\bf{k}} }^{\sigma \sigma^{\prime} }$, corresponding to marginal
1391: Landau interaction parameters.
1392:
1393: Expressing each term in the expansion of the generating functional
1394: $\Gamma [ \bar{\psi} , \psi , \varphi ]$ given in
1395: Eq.~(\ref{eq:expansion2}) in terms of the rescaled variables defined
1396: above and using the fact that $\Gamma$ is dimensionless, we obtain the
1397: scaling form of the vertices. Omitting for simplicity the degeneracy
1398: labels $\sigma$, and assuming $z_{\psi}\leq z_{\varphi}$,\cite{zfootnote}
1399: we define the rescaled vertices,
1400: \begin{eqnarray}
1401: \tilde{\Gamma}_l^{(2n,m)}(Q_1',\dots,Q_n';Q_1,\dots,Q_n;\bar{Q}_1,\dots,\bar{Q}_m)
1402: = & &
1403: \nonumber\\
1404: & & \hspace{-76mm}
1405: \nu_0^{-m/2} \Lambda^{D ( n -1 + m/2) }
1406: \Omega_{\Lambda}^{-1} \bar{\Omega}_{\Lambda}^{{m}/{2}}
1407: Z^n \bar{Z}^{{m}/2}
1408: \nonumber
1409: \\
1410: & & \hspace{-79mm} \times
1411: \Gamma_{\Lambda}^{(2n,m)}(K_1',\dots,K_n';K_1,\dots,K_n;\bar{K}_1,\dots,\bar{K}_m)\,,
1412: \label{eq:Gammarescaledef}
1413: \end{eqnarray}
1414: where we have to exclude the cases of purely bosonic vertices ($n=0$)
1415: as well as the fermionic two-point vertex (i.e., the rescaled
1416: irreducible self-energy, corresponding to $n=1$ and $m=0$), which both
1417: need separate definitions. For the purely bosonic vertices ($n=0$) we
1418: set
1419: \begin{eqnarray}
1420: \tilde{\Gamma}^{(0,m)}_{l} ( \bar{Q}_1,\dots,\bar{Q}_m ) & = &
1421: \nu_0^{-m/2} ( \Lambda^{D} \bar{\Omega}_{\Lambda} )^{ -1 + {m}/{2} }
1422: \bar{Z}^{{m}/2}
1423: \nonumber
1424: \\
1425: & \times &
1426: \Gamma_{ \Lambda }^{(0,m)} ( \bar{K}_1,\dots,\bar{K}_m)
1427: \; ,
1428: \label{eq:Gammarescaledefbos}
1429: \end{eqnarray}
1430: while for the fermionic two-point vertex we should subtract the exact
1431: fixed point self-energy $ \Sigma_{\ast} ( {\bf{k}}_{F ,\sigma} , i0 )$
1432: at the Fermi-surface reference-point ${\bf{k}}_{F, \sigma}$ and for
1433: vanishing frequency as a counterterm,\cite{Kopietz01,Ledowski03}
1434: \begin{equation}
1435: \tilde{\Gamma}^{(2,0)}_l ( Q ; Q )
1436: \equiv \tilde{\Sigma}_l (Q ) = \frac{Z}{\Omega_{\Lambda}}
1437: \left[ \Sigma ( K ) - \Sigma_{\ast} ( {\bf{k}}_{F, \sigma} , i0 ) \right]
1438: \; .
1439: \label{eq:Sigmarescaledef}
1440: \end{equation}
1441: If necessary, the counterterm $ \Sigma_{\ast} ( {\bf{k}}_{F ,\sigma}
1442: , i0 )$ can be reconstructed from the condition that the constant part
1443: $\tilde{r}_l = \tilde{\Sigma}_l ( 0 )$ of the self-energy flows into
1444: an RG fixed point.\cite{Kopietz01,Ledowski03} We consider the
1445: rescaled vertices to be functions of the logarithmic flow parameter $l
1446: = - \ln ( \Lambda / \Lambda_0 )$. Introducing the flowing anomalous
1447: dimensions associated with the fermionic and bosonic fields,
1448: \begin{equation}
1449: \eta_l = -\partial_l \ln Z \; \; \; , \; \; \;
1450: \bar{\eta}_l = -\partial_l \ln \bar{Z} \,,
1451: \label{eq:etadef}
1452: \end{equation}
1453: we can then write down the flow equations for the rescaled vertices.
1454: Omitting the arguments, we obtain for $n \geq 1$ the flow equation\cite{zfootnote}
1455: \begin{widetext}
1456: \begin{eqnarray}
1457: \partial_l \tilde{\Gamma}^{(2n,m)}_l &=&
1458: \left[ (1-n) D + z_{\text{min}} - \frac{m}{2} ( D + z_\varphi ) - n \eta_l -\frac{m}2 \bar{\eta}_l
1459: - \sum_{i=1}^n (Q_i' \cdot \frac{\partial} {\partial Q_i'} +
1460: Q_i \cdot \frac{\partial}{\partial Q_i})
1461: - \sum_{i=1}^m \bar{Q}_i \cdot \frac{\partial}{\partial \bar{Q}_i}
1462: \right]
1463: \tilde{\Gamma}_l^{(2n,m)}
1464: \nonumber\\
1465: &&+\;\dot{\tilde{\Gamma}}_l^{(2n,m)}\,,
1466: \label{eq:flowGammarescale1}
1467: \end{eqnarray}
1468: where $z_{\text{min}}=\text{min}\{z_{\varphi},z_{\psi}\}$. For $n=0$
1469: we obtain from Eq.~(\ref{eq:Gammarescaledefbos}),
1470: \begin{equation}
1471: \partial_l \tilde{\Gamma}^{(0,m)}_l =
1472: \left[ (1 - \frac{m}{2}) ( D + z_\varphi ) -\frac{m}2 \bar{\eta}_l
1473: - \sum_{i=1}^m \bar{Q}_i \cdot \frac{\partial}{\partial \bar{Q}_i}
1474: \right]
1475: \tilde{\Gamma}_l^{(0,m)}
1476: \,+\,\dot{\tilde{\Gamma}}_l^{(0,m)}\,,
1477: \label{eq:flowGammarescale2}
1478: \end{equation}
1479: \end{widetext}
1480: where we have introduced the notation
1481: \begin{eqnarray}
1482: Q \cdot \frac{\partial }{ \partial Q} & \equiv & {\bf{q}} \cdot \mathbf{\nabla}_{\bf{q}} + z_{\psi} \; \epsilon
1483: \frac{\partial}{\partial \epsilon}
1484: \label{eq:QdQdef}
1485: \; ,
1486: \\
1487: \bar{Q} \cdot \frac{\partial }{ \partial \bar{Q}} & \equiv & \bar{\bf{q}} \cdot \mathbf{\nabla}_{\bar{\bf{q}}}
1488: + z_{\varphi} \; \bar{\epsilon}
1489: \frac{\partial}{\partial \bar{\epsilon}}
1490: \label{eq:barQdQdef}
1491: \; .
1492: \end{eqnarray}
1493: The inhomogeneities in Eqs.~(\ref{eq:flowGammarescale1}) and
1494: (\ref{eq:flowGammarescale2}) are given by the rescaled version of the
1495: right-hand sides of the flow equations for the unrescaled vertices,
1496: i.e., for $n \geq 1$, and $z_{\psi}\leq z_{\varphi}$,\cite{zfootnote}
1497: \begin{eqnarray}
1498: \dot{\tilde{\Gamma}}_l^{(2n,m)}(Q_1',\dots,Q_n';Q_1,\dots,Q_n';\bar{Q}_1,\dots,\bar{Q}_m)
1499: & = &
1500: \nonumber
1501: \\
1502: & & \hspace{-70mm}
1503: \nu_0^{-m/2} \Lambda^{D ( n -1 + m/2) }
1504: \Omega_{\Lambda}^{-1} \bar{\Omega}_{\Lambda}^{{m}/{2}}
1505: Z^n \bar{Z}^{{m}/2}
1506: \nonumber
1507: \\
1508: & & \hspace{-70mm} \times
1509: [ - \Lambda \partial_{\Lambda} \Gamma_{\Lambda}^{(2n,m)}( \left\{ K_i' ;K_i;\bar{K}_i\right\}) ]
1510: \label{eq:inhrescale1}
1511: \; ,
1512: \end{eqnarray}
1513: and for $n=0$,
1514: \begin{eqnarray}
1515: \dot{\tilde{\Gamma}}^{(0,m)}_{l} ( \bar{Q}_1,\dots,\bar{Q}_m ) & = &
1516: \nu_0^{-m/2} ( \Lambda^{D} \bar{\Omega}_{\Lambda} )^{ -1 + {m}/{2} }
1517: \bar{Z}^{{m}/2}
1518: \nonumber
1519: \\
1520: & & \hspace{-15mm} \times
1521: [ - \Lambda \partial_{\Lambda}
1522: \Gamma_{ \Lambda }^{(0,m)} ( \bar{K}_1,\dots,\bar{K}_m) ] \, .
1523: \label{eq:inhrescale2}
1524: \end{eqnarray}
1525: By properly counting all factors it is then not difficult to see that
1526: the explicit expressions for the inhomogeneities in
1527: Eqs.~(\ref{eq:inhrescale1}) and (\ref{eq:inhrescale2}) can be simply
1528: obtained from their unrescaled counterparts by replacing all vertices
1529: and propagators with their rescaled analogs, where the rescaled
1530: propagators are defined by
1531: \begin{equation}
1532: {G}(K)=\frac{Z}{\Omega_{\Lambda} } \tilde{G}(Q)\,,
1533: \; \; \;
1534: {F}(\bar{K})=\frac{\bar{Z}}{ \nu_0 }\tilde{F} (\bar{Q})\,,
1535: \label{eq:rescaleprop}
1536: \end{equation}
1537: and the corresponding rescaled single-scale propagators are defined
1538: via
1539: \begin{equation}
1540: \Lambda \dot{G}(K)=-\frac{Z}{\Omega_{\Lambda} } \dot{\tilde{G}} (Q)\,,
1541: \; \; \;
1542: \Lambda \dot{F}(\bar{K})=-\frac{\bar{Z}}{ \nu_0 }\dot{\tilde{F}} (\bar{Q})\,.
1543: \label{eq:rescalesingleprop}
1544: \end{equation}
1545: From Eqs.~(\ref{eq:flowGammarescale1}) and
1546: (\ref{eq:flowGammarescale2}) we can read off the scaling dimensions of
1547: the vertices: the scaling dimension of $\tilde{\Gamma}^{(2n,m)}$ in
1548: $D$ dimensions is
1549: \begin{eqnarray}
1550: D^{(2n,m)} =~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~&& \nonumber\\[1mm]
1551: \left\{
1552: \begin{array}{ll}
1553: (1-n) D + z_{\text{min}} -
1554: ( D + z_{\varphi} ) m/2 & \mbox{for $ n \geq 1$} \\
1555: ( D + z_{\varphi} ) ( 1 - m/2) &\mbox{for $ n =0$}
1556: \end{array}
1557: \right.
1558: \,.&&\nonumber\\
1559: &&\label{eq:scaledim}
1560: \end{eqnarray}
1561: In the particular case of the Tomonaga-Luttinger model, where $D=1$
1562: and $z_{\psi} = z_{\varphi} = 1$, we have $D^{(2n,m)} = 2 -n - m$.
1563: Hence, in this case $\tilde{\Gamma}^{(2,0)} ( Q =0)$ and
1564: $\tilde{\Gamma}^{(0,1)} $ are relevant with scaling dimension $+1$,
1565: while $\tilde{\Gamma}^{ (4,0 )} (Q_i =0)$ and $\tilde{\Gamma}^{(2,1)}
1566: ( Q_i= \bar{Q}_i =0)$ are marginal. All other vertices are irrelevant.
1567: Of course, the linear terms in the expansion of
1568: $\tilde{\Gamma}^{(2,0)} ( Q;Q )$ for small $Q$ are also marginal.
1569: These terms determine the wave-function renormalization factor $Z$ and
1570: the Fermi velocity renormalization $\tilde{v}_l$, see
1571: Eqs.~(\ref{eq:Zdefexplicit}) and (\ref{eq:vtildedef}) below. Note
1572: that for short-range interactions the dispersion of the zero-sound
1573: mode is linear in any dimension.\cite{Kopietz97} Hence, as long as the
1574: density response is dominated by the zero sound mode,
1575: Eq.~(\ref{eq:scaledim}) remains valid for $D > 1$ with $z_{\psi} =
1576: z_{\psi} = 1$. In this case the scaling dimension of the purely
1577: fermionic four-point vertex is $D^{(4,0)} = 1 - D$ and the scaling
1578: dimension of the three-legged vertex with two fermion legs and one
1579: boson leg is $D^{(2,1)} = (1 - D )/2$. Both vertices become
1580: irrelevant in $D > 1$. As discussed in the following section, this
1581: means that the random-phase approximation (RPA) for the effective
1582: interaction, as well as the so-called GW approximation~\cite{Hedin65}
1583: for the fermionic self-energy, are qualitatively correct in $D>1$.
1584:
1585: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1586: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1587: \subsection{A simple truncation scheme: Keeping only the skeleton elements for
1588: two-point functions}
1589: \label{subsec:truncation}
1590:
1591: In order to solve the flow equations explicitly, one is forced to
1592: truncate the infinite hierarchy of flow equations. In the
1593: one-particle irreducible version of the purely fermionic functional RG
1594: it is common
1595: practice\cite{Zanchi96,Halboth00,Honerkamp01,Honerkamp01b,Kampf03,Katanin04}
1596: to retain only vertices up to the four-point vertex and set all higher
1597: order vertices equal to zero. Our approach offers new possibilities
1598: for truncation schemes. Consider the skeleton graphs
1599: \cite{Nozieres64} for the one-particle irreducible fermionic
1600: self-energy and the one-interaction-line irreducible polarization
1601: shown in Figs.~\ref{fig:skeletonsigmapi}(a) and (b).
1602: %
1603: \begin{figure}
1604: \begin{center}
1605: \epsfig{file=skeletonsigmapi.eps,width=1.0\hsize}
1606: \end{center}
1607: \caption{Skeleton diagrams for (a) the
1608: one-particle irreducible fermionic self-energy; (b) the
1609: one-interaction-line irreducible polarization; and (c) the
1610: three-legged vertex with two fermion legs and one boson leg.
1611: The small black circle denotes the bare three-legged vertex. Thin
1612: lines denote external legs. The other graphical elements are the
1613: same as in Fig.~\ref{fig:picdic}. }
1614: \label{fig:skeletonsigmapi}
1615: \end{figure}
1616: %
1617: The skeleton graphs contain three basic elements: the exact fermionic
1618: Green's function, the exact bosonic Green's function (i.e., the effective
1619: screened interaction), and the three-legged vertex with two fermion
1620: legs and one boson leg. A systematic derivation of the skeleton
1621: expansion for the vertices in our coupled Fermi-Bose theory is
1622: presented in Appendix~B. One advantage of our RG approach (as
1623: compared with more conventional methods involving only fermionic
1624: fields) is that it yields directly the flow equations for basic
1625: elements appearing in the skeleton graphs for the self-energy and the
1626: polarization shown in Fig.~\ref{fig:skeletonsigmapi}. Of course, in
1627: principle the three-legged vertex can be obtained from the vertex with
1628: four fermion legs with the help of the skeleton graph shown in
1629: Fig.~\ref{fig:skeletonsigmapi}(c). However, calculating the
1630: three-legged vertex from the four-legged vertex in this way involves
1631: an intermediate integration, which requires the knowledge of the
1632: momentum and frequency dependence of the four-legged vertex.
1633: Unfortunately, in practice the purely fermionic functional RG
1634: equations have to be severely truncated so that up to now it has not
1635: been possible to keep track of the frequency dependence of the
1636: four-legged fermion vertex within the purely fermionic functional RG.
1637:
1638: To obtain a closed system of RG equations involving only the skeleton
1639: elements, let us retain only the vertices $\Sigma_{\sigma} ( K )$,
1640: $\Pi_{\sigma} ( {\bar{K}} )$ and $\Gamma^{(2,1)} ( K + \bar{K} \sigma;
1641: K \sigma ; \bar{K} \sigma)$ on the right-hand sides of the exact flow
1642: equations for these quantities shown in Figs.~\ref{fig:flowSigma},
1643: \ref{fig:flowPi}, and \ref{fig:flowVCorr}, and set all other vertices
1644: equal to zero. The resulting closed system of flow equations is shown
1645: graphically in Fig.~\ref{fig:flowtrunc}.
1646: \begin{figure}
1647: \begin{center}
1648: \epsfig{file=flowtrunc.eps,width=1.0\hsize}
1649: \end{center}
1650: \caption{Truncation of the flow equations for
1651: (a) fermionic self-energy, (b) irreducible polarization, and (c)
1652: three-legged vertex which sets all other vertices equal to zero.
1653: The internal lines are full propagators, which depend on the
1654: self-energies $\Gamma^{(2,0)} = \Sigma $ and $\Gamma^{(0,2)} =
1655: \Pi$. }
1656: \label{fig:flowtrunc}
1657: \end{figure}
1658: %
1659: Explicitly, the flow equations are
1660: \begin{eqnarray}
1661: \partial_{\Lambda} {\Sigma}_{\sigma} ( K ) & = &
1662: \nonumber
1663: \\
1664: & & \hspace{-20mm}
1665: \int_{\bar{K}}
1666: \left[ \dot{F}_{\sigma \sigma} ( \bar{K} ) G_{\sigma} ( K + \bar{K} )
1667: + {F}_{\sigma \sigma} ( \bar{K} ) \dot{G}_{\sigma} ( K + \bar{K} )
1668: \right]
1669: \nonumber
1670: \\
1671: & & \hspace{-20mm} \times
1672: \Gamma^{(2,1)} ( K + \bar{K} \sigma; K \sigma; \bar{K} \sigma )
1673: \Gamma^{(2,1)} ( K \sigma; K + \bar{K} \sigma; - \bar{K} \sigma )
1674: \; ,
1675: \nonumber
1676: \\
1677: \label{eq:flowSigmatrunc}
1678: \end{eqnarray}
1679: %
1680: \begin{eqnarray}
1681: \partial_{\Lambda} {\Pi}_{\sigma} ( \bar{K} ) & = &
1682: \nonumber
1683: \\
1684: & & \hspace{-20mm} -\zeta \int_K
1685: \left[ \dot{ G}_{\sigma} ( K ) G_{\sigma} ( K + \bar{K} ) + { G}_{\sigma} ( K )
1686: \dot{G}_{\sigma} ( K + \bar{K} ) \right]
1687: \nonumber
1688: \\
1689: & & \hspace{-20mm} \times
1690: \Gamma^{(2,1)} ( K + \bar{K} \sigma; K \sigma; \bar{K} \sigma )
1691: \Gamma^{(2,1)} ( K \sigma; K + \bar{K} \sigma; - \bar{K} \sigma )
1692: \; ,
1693: \nonumber
1694: \\
1695: \label{eq:flowPitrunc}
1696: \end{eqnarray}
1697: %
1698: \begin{eqnarray}
1699: \partial_{\Lambda} {\Gamma}^{(2,1)} ( K + \bar{K} \sigma ; K \sigma ;
1700: \bar{K} \sigma ) & = &
1701: \nonumber
1702: \\
1703: & & \hspace{-40mm}
1704: \int_{\bar{K}^{\prime}}
1705: \Bigl[ \dot{ F}_{\sigma \sigma} ( \bar{K}^{\prime} ) G_{\sigma} ( K + \bar{K}^{\prime} ) G_{\sigma} ( K + \bar{K} + \bar{K}^{\prime} )
1706: \nonumber
1707: \\
1708: & & \hspace{-35mm} +
1709: { F}_{\sigma \sigma} ( \bar{K}^{\prime} ) \dot{G}_{\sigma} ( K + \bar{K}^{\prime} ) G_{\sigma} ( K + \bar{K} + \bar{K}^{\prime} )
1710: \nonumber
1711: \\
1712: & & \hspace{-35mm} +
1713: { F}_{\sigma \sigma} ( \bar{K}^{\prime} ) G_{\sigma} ( K + \bar{K}^{\prime} ) \dot{G}_{\sigma} ( K + \bar{K} + \bar{K}^{\prime} )
1714: \Bigr]
1715: \nonumber
1716: \\
1717: & & \hspace{-35mm} \times
1718: \Gamma^{(2,1)} ( K + \bar{K} \sigma ; K + \bar{K} + \bar{K}^{\prime} \sigma ; - \bar{K}^{\prime} \sigma )
1719: \nonumber
1720: \\
1721: & & \hspace{-35mm} \times
1722: \Gamma^{(2,1)} ( K + \bar{K} + \bar{K}^{\prime} \sigma ; K + \bar{K}^{\prime} \sigma ; \bar{K} \sigma )
1723: \nonumber
1724: \\
1725: & & \hspace{-35mm} \times
1726: \Gamma^{(2,1)} ( K + \bar{K}^{\prime} \sigma ; K \sigma ; \bar{K}^{\prime} \sigma )
1727: \; .
1728: \label{eq:flowGammatrunc}
1729: \end{eqnarray}
1730: These equations form a closed system of integrodifferential equations
1731: that can in principle be solved numerically. If the initial momentum
1732: transfer cutoff $\Lambda_0$ is chosen larger than the maximal momentum
1733: transferred by the bare interaction, and if the initial bandwidth
1734: cutoff $v_0 \Lambda_0$ is larger than the bandwidth of the bare energy
1735: dispersion, then the initial conditions are ${\Sigma}_{\sigma} ( K
1736: )_{\Lambda_0} =0$, ${\Pi}_{\sigma} ( \bar{K} )_{\Lambda_0} =0$, and
1737: ${\Gamma}^{(2,1)} ( K + \bar{K} \sigma ; K \sigma ; \bar{K} \sigma
1738: )_{\Lambda_0} = i$. A numerical solution of these coupled equations
1739: seems to be a difficult task, which we shall not attempt in this work.
1740: Note, however, that in Sec.~\ref{subsec:rescale} we have argued that
1741: for regular interactions in dimensions $D > 1$ the three-legged vertex
1742: is actually irrelevant in the RG sense. Hence, we expect that the
1743: qualitatively correct behavior of the fermionic self-energy and of the
1744: polarization can be obtained by ignoring the flow of the three-legged
1745: vertex, setting $\Gamma^{(2,1)} \rightarrow i$. If we further ignore
1746: interaction corrections to the internal propagators in the flow
1747: equation (\ref{eq:flowPitrunc}) for the polarization, it is easy to
1748: see that the solution of this equation is nothing but the
1749: noninteracting polarization. This is equivalent with the RPA for the
1750: effective interaction. Substituting this into the flow equation
1751: (\ref{eq:flowSigmatrunc}) for the self-energy and ignoring again
1752: self-energy corrections to the internal Green's functions, we obtain the
1753: non-self-consistent GW approximation\cite{Hedin65} for the fermionic
1754: self-energy. For regular interactions in $D > 1$ we therefore expect
1755: that the RPA and the GW approximation are qualitatively correct.
1756: However, for strong bare interactions quantitatively accurate results
1757: can only be expected if the vertex corrections described by
1758: Eq.~(\ref{eq:flowGammatrunc}) are at least approximately taken into
1759: account.
1760:
1761: We shall consider this problem again in Sec.~\ref{subsec:truncrel},
1762: where we discuss truncations of an expansion based on relevance. To
1763: lowest order, this approximation will agree with
1764: Eqs.~(\ref{eq:flowSigmatrunc})--(\ref{eq:flowGammatrunc}) when the
1765: dependence of the vertex $\Gamma^{(2,1)}$ on momenta and frequencies
1766: is ignored. There, we use the resulting equations to calculate an
1767: approximation to the electronic Green's function of the one-dimensional
1768: Tomonaga-Luttinger model. Amazingly, this simple truncation is
1769: sufficient to reproduce the correct anomalous dimension known from
1770: bosonization even for large values of the bare coupling.
1771:
1772:
1773: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1774: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1775: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1776: \section{The momentum-transfer cutoff as flow parameter}
1777: \label{sec:1dflow}
1778:
1779: The truncation discussed in Sec.~\ref{subsec:truncation} violates the
1780: Ward identities relating vertices with different numbers of external
1781: legs (for a self-contained derivation of the Ward identities within
1782: the framework of our functional integral approach; see
1783: Appendix~\ref{sec:ward}). Moreover, even if we do not truncate the
1784: exact hierarchy of flow equations shown in Figs.~\ref{fig:flowSigma},
1785: \ref{fig:flowPi}, and \ref{fig:flowVCorr}, the Ward identities are
1786: violated for any finite value of the bandwidth-cutoff $v_0 \Lambda$,
1787: because the cutoff leads to a violation of the underlying gauge
1788: symmetry. We can thus only expect the Ward identities to be restored
1789: in the limit $v_0 \Lambda \rightarrow 0$. Recall that in the
1790: Tomonaga-Luttinger model the Ward identities are valid in the strict
1791: sense only in the presence of the Dirac sea, implying that the
1792: ultraviolet cutoff $v_0\Lambda_0$ has been removed. The Ward
1793: identities and the underlying asymptotic conservation laws are crucial
1794: for the exact solubility of the TLM~\cite{Dzyaloshinskii74,Bohr81} and
1795: its higher-dimensional
1796: generalization.\cite{Metzner98,Kopietz97,Haldane92,Bartosch99} In
1797: order to reproduce the exact solution of the TLM known from
1798: bosonization within the functional RG, it is very important to have RG
1799: flow equations which are consistent with the Ward identities, even for
1800: finite values of the cutoff. In this section we show that this
1801: requirement is fulfilled if we work in our mixed Fermi-Bose RG with a
1802: momentum-transfer cutoff $\Lambda$ only and take the limit $v_0
1803: \Lambda \rightarrow 0$ of a vanishing bandwidth cutoff.
1804:
1805: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1806: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1807: \subsection{Exact flow equations for momentum-transfer cutoff}
1808:
1809: %
1810: %
1811: \begin{figure}
1812: \begin{center}
1813: \epsfig{file=flowsigmapitrans.eps,width=1.0\hsize}
1814: \end{center}
1815: \caption{
1816: Exact flow equations for (a) the fermionic self-energy and (b) the
1817: irreducible polarization in the momentum-transfer cutoff scheme.
1818: }
1819: \label{fig:flowsigmapitrans}
1820: \end{figure}
1821: %
1822: %
1823: \begin{figure}
1824: \begin{center}
1825: \epsfig{file=flowGammatrans.eps,width=0.8\hsize}
1826: \end{center}
1827: \caption{
1828: Exact flow equations for the three-legged vertex with two
1829: fermion legs and one boson leg in the momentum-transfer cutoff
1830: scheme. }
1831: \label{fig:flowGammatrans}
1832: \end{figure}
1833: %
1834: As already briefly mentioned at the end of
1835: Sec.~\ref{subsec:flowphysical}, if we work with a momentum transfer
1836: cutoff $\Lambda$ only, then all diagrams with a slash on an internal
1837: fermionic Green's function [corresponding to the fermionic component of
1838: the single-scale propagator given in Eq.~(\ref{eq:dotGdiagdef})] on
1839: the right-hand sides of the exact flow equations shown in
1840: Figs.~\ref{fig:flowSigma}, \ref{fig:flowPi}, and \ref{fig:flowVCorr}
1841: should be omitted. The exact flow equations for the electronic
1842: self-energy and the irreducible polarization then reduce to
1843: \begin{eqnarray}
1844: \partial_{\Lambda} \Sigma_{\sigma}(K) &=& \frac12 \int_{\bar{K}}
1845: \dot{F}_{\sigma\sigma}(\bar{K})
1846: \Gamma^{(2,2)}(K\sigma;K\sigma;\bar{K}\sigma,-\bar{K}\sigma)
1847: \nonumber\\
1848: && \hspace{-20mm} + \int_{\bar{K}}
1849: \dot{F}_{\sigma\sigma}(\bar{K})G_{\sigma}(K+\bar{K})
1850: \Gamma^{(2,1)} ( K + \bar{K} \sigma; K \sigma; \bar{K} \sigma )
1851: \nonumber
1852: \\
1853: & & \times
1854: \Gamma^{(2,1)} ( K \sigma; K + \bar{K} \sigma; - \bar{K} \sigma )
1855: \; ,
1856: \label{eq:flowsigmatransfer}
1857: \end{eqnarray}
1858: %
1859: %
1860: \begin{eqnarray}
1861: \partial_{\Lambda} \Pi_{\sigma}(\bar{K}) &=& \frac12 \int_{\bar{K}'}
1862: \dot{F}_{\sigma\sigma}(\bar{K})
1863: \Gamma^{(0,4)}(\bar{K}'\sigma,-\bar{K}'\sigma,\bar{K}\sigma,-\bar{K}\sigma)
1864: \nonumber\\
1865: &&\hspace{-22mm}
1866: -
1867: \int_{\bar{K}'}
1868: \dot{F}_{\sigma\sigma}(\bar{K}')F_{\sigma\sigma}(\bar{K}+\bar{K}')
1869: \Gamma^{(0,3)}(-\bar{K}\sigma,\bar{K}+\bar{K}'\sigma,-\bar{K}'\sigma)
1870: \nonumber\\
1871: &&\times
1872: \Gamma^{(0,3)}(\bar{K}'\sigma,-\bar{K}-\bar{K}'\sigma,\bar{K}\sigma) \; .
1873: \label{eq:flowPitransfer}
1874: \end{eqnarray}
1875: These equations are shown graphically in
1876: Fig.~\ref{fig:flowsigmapitrans}. A graphical representation of the
1877: corresponding exact flow equation for the three-legged vertex is shown
1878: in Fig.~\ref{fig:flowGammatrans}.
1879: Still, these flow equations look rather complicated. Since we have
1880: imposed a cutoff only in the momentum transfered by the bosons, the
1881: initial conditions at scale $\Lambda_0$ are nontrivial. The initial
1882: value of the three-legged vertex is still $\Gamma^{(2,1)}_{\Lambda_0}
1883: = i$, but the pure boson vertices $\Gamma^{(0,m)}$ with $m$ external
1884: legs are initially given by the symmetrized closed fermion loops shown
1885: in Fig.~\ref{fig:closedloop}. All other vertices vanish at the initial
1886: scale $\Lambda_0$.
1887: %
1888: \begin{figure}
1889: \begin{center}
1890: \epsfig{file=closedloop.eps,width=0.8\hsize}
1891: \end{center}
1892: \caption{
1893: Initial condition for the pure boson vertices in the momentum
1894: transfer cutoff scheme. The sum is taken over the $m!$
1895: permutations of the labels of the external legs. For linearized
1896: energy dispersion, all symmetrized closed fermion loops with more
1897: than two external legs vanish. }
1898: \label{fig:closedloop}
1899: \end{figure}
1900: %
1901: An essential simplification occurs now if we linearize the energy
1902: dispersion relative to the Fermi surface. If the initial momentum
1903: transfer cutoff $\Lambda_0$ is small compared with the typical Fermi
1904: momentum, then we may set all pure boson vertices $\Gamma^{(0,m)}$
1905: with more than two external boson legs ($ m \geq 2$) equal to zero.
1906: This is nothing but the closed loop
1907: theorem,\cite{Dzyaloshinskii74,Bohr81,Kopietz97,Kopietz95,Metzner98}
1908: which is valid exactly for the one-dimensional TLM (where the energy
1909: dispersion is linear by definition). In higher dimensions, the closed
1910: loop theorem is valid to a very good approximation as long as the
1911: linearization of the energy dispersion is justified within a given
1912: sectorization of the Fermi surface and scattering processes that
1913: transfer momentum between different sectors of the Fermi surface can
1914: be neglected.\cite{Kopietz97,Kopietz95} Note that the closed loop
1915: theorem is consistent with the momentum-transfer cutoff flow, because
1916: pure boson vertices $\Gamma^{(0,m )}$ with $ m \geq 3$ are not
1917: generated if they initially vanish.
1918:
1919: Assuming the validity of the closed loop theorem, the right-hand side
1920: of the flow equation (\ref{eq:flowPitransfer}) for the polarization
1921: vanishes identically, because it depends only on boson vertices with
1922: more than two external legs. Physically, this means that there are no
1923: corrections to the noninteracting polarization, so that the RPA for
1924: the effective interaction is exact. This is of course well known since
1925: the pioneering work by Dzyaloshinskii and
1926: Larkin.\cite{Dzyaloshinskii74} Moreover, the last three diagrams in
1927: the flow equation for $\Gamma^{(2,1)}$ shown in
1928: Fig.~\ref{fig:flowGammatrans} also vanish, because they contain the
1929: vertex $\Gamma^{(0,3)}$. However, the remaining diagrams in
1930: Fig.~\ref{fig:flowsigmapitrans}(a) and Fig.~\ref{fig:flowGammatrans}
1931: still look quite complicated, so that we still have to solve an
1932: infinite hierarchy of coupled flow equations. In the next subsection
1933: we show how this infinite system of coupled integro\-differential
1934: equations can be solved exactly.
1935:
1936:
1937: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1938: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1939: \subsection{Ward identities as solutions of the infinite hierarchy
1940: of flow equations}
1941:
1942: %
1943: %
1944: \begin{figure*}
1945: \epsfig{file=flow_2Fermi_mBoson.eps,width=0.9\hsize}
1946: \caption{Diagrammatic representation
1947: of the flow equation~(\ref{eq:flowdisord}) of vertices with two
1948: fermion legs and a general number of boson legs provided the pure
1949: boson vertices with more than two external legs vanish, as implied
1950: by the closed loop theorem.}
1951: \label{fig:flow_2Fermi_mBoson}
1952: \end{figure*}
1953: %
1954: %
1955:
1956: Let us consider the terms on the right-hand sides of the flow
1957: equations for the vertices $\Gamma^{(2,m)}$ with two external fermion
1958: legs and an arbitrary number of boson legs. Assuming the validity of
1959: the closed loop theorem, all pure boson vertices $\Gamma^{(0,m)}$ with
1960: $m \geq 2$ vanish. From Fig.~\ref{fig:flowsigmapitrans}(a) and
1961: Fig.~\ref{fig:flowGammatrans} it is clear that in general the
1962: right-hand side of the flow equation for $\partial_{\Lambda}
1963: \Gamma^{(2,m)}$ depends on $\Gamma^{(2, m+2 )}$ and on all
1964: $\Gamma^{(2,m^{\prime})}$ with $m^{\prime} \leq m$. In fact, from our
1965: general expression for the flow of the totally symmetrized vertices
1966: given in Eq.~(\ref{eq:flow_vert}), we can derive the flow equations for
1967: the vertices $\Gamma^{(2,m)}$ with arbitrary $m$ in closed form (we
1968: omit for simplicity the degeneracy index $\sigma$),
1969: \begin{widetext}
1970: \begin{eqnarray}
1971: \partial_{\Lambda}\Gamma^{(2,m)}(K';K;\bar{K}_1,\dots,\bar{K}_m)
1972: = \frac12\int_{\bar{K}}\dot{F}_{\sigma\sigma}(\bar{K})
1973: \Gamma^{(2,m+2)}(K';K;-\bar{K},\bar{K},\bar{K}_1,\dots,\bar{K}_m)
1974: +\sum_{l=2}^{\infty}\sum_{m_1,\dots,m_l=1}^{\infty}\!\!\frac{\delta_{m,\sum_im_i}}{\prod _i m_i!}
1975: &&
1976: \nonumber\\
1977: \times\sum_{P}\int_{\bar{K}}
1978: \dot{F}(\bar{K})
1979: \Gamma^{(2,m_1+1)}\left(K';\tilde{K}_1;\bar{K}_{P(1)},\dots,\bar{K}_{P(m_1)},-\bar{K}\right)
1980: G(\tilde{K}_1)
1981: \Gamma^{(2,m_2)}\left(\tilde{K}_1;\tilde{K}_2;\bar{K}_{P(m_1+1)},\dots,\bar{K}_{P(m_1+m_2)}\right)
1982: &&
1983: \nonumber\\
1984: \times G(\tilde{K}_2)
1985: \cdot\ldots\cdot
1986: G(\tilde{K}_{l-1})
1987: \Gamma^{(2,m_l+1)}\left(\tilde{K}_{l-1};K;\bar{K},\bar{K}_{P(m-m_l+1)},\dots,\bar{K}_{P(m)}\right)
1988: \; ,~~~~~~~~~~~~~~~~~~~~~~~&&
1989: \label{eq:flowdisord}
1990: \end{eqnarray}
1991: \end{widetext}
1992: where we have defined
1993: \begin{equation}
1994: K' = K + \sum_{i=1}^m \bar{K}_i\,,
1995: \qquad
1996: \tilde{K}_i = K' + \bar{K} - \sum_{j=1}^{m_1+\ldots+m_i} \bar{K}_{P(j)}\,,
1997: \end{equation}
1998: and $P$ denotes a permutation of $\{1,\dots,m\}$. A graphical
1999: representation of Eq.~(\ref{eq:flowdisord}) is shown in
2000: Fig.~\ref{fig:flow_2Fermi_mBoson}.
2001: Note that the flow equation (\ref{eq:flowsigmatransfer}) for the
2002: irreducible self-energy is a special case of Eq.~(\ref{eq:flowdisord})
2003: for $m=0$.
2004:
2005: We are now facing the problem of solving the infinite hierarchy of
2006: coupled flow equations given by Eq.~(\ref{eq:flowdisord}). In view of
2007: the fact that these equations are exact and that in one dimension the
2008: single-particle Green's function of the TLM can be calculated exactly
2009: via bosonization, we expect that this infinite hierarchy of flow
2010: equations can also be solved exactly. Indeed, the solutions of these
2011: equations are nothing but infinitely many Ward identities relating the
2012: vertex $\Gamma^{(2,m)}$ with two fermion legs and $m$ boson legs to
2013: the vertex $\Gamma^{(2, m-1)}$ with one boson leg less. We derive
2014: these Ward identities within the framework of our functional integral
2015: approach in Appendix~C. For $m =1$ the Ward identity is well
2016: known~\cite{Dzyaloshinskii74,Bohr81,Metzner98,Kopietz97}
2017: \begin{eqnarray}
2018: G ( K + \bar{K} ) \Gamma^{(2,1)} ( K + \bar{K} ; K ; \bar{K} ) G ( K )
2019: & = &
2020: \nonumber
2021: \\
2022: & & \hspace{-55mm} =
2023: \frac{-i}{i\bar{\omega} - {\bf{v}}_{F, \sigma} \cdot \bar{\bf{k}}}
2024: \Big[ G ( K + \bar{K} ) - G ( K ) \Big]
2025: \; .
2026: \label{eq:WI1}
2027: \end{eqnarray}
2028: Here ${\bf{v}}_{F, \sigma}$ is the Fermi velocity associated with the
2029: independent fermionic label $K = ({\bf{k}} , i \omega )$, where $|
2030: {\bf{k}} - {\bf{k}}_{ F , \sigma} | \ll | {\bf{k}}_{ F , \sigma} | $.
2031: The Ward identity (\ref{eq:WI1}) has been used in
2032: Refs.~\onlinecite{Dzyaloshinskii74} and \onlinecite{Metzner98} to
2033: close the skeleton equation for the self-energy and thus obtain the
2034: exact Green's function of the TLM without invoking the machinery of
2035: bosonization. A Ward identity for $\Gamma^{(4,1)}$ has also been used
2036: to prove the vanishing of the renormalization group $\beta$ function
2037: for the TLM.\cite{DiCastro91} However, for solving the TLM exactly
2038: within the framework of the functional RG, we need the Ward identities
2039: for all vertices $\Gamma^{(2,m)}$ with $m \geq 1$. As shown in
2040: Appendix C, for linear energy dispersion we have
2041: \begin{widetext}
2042: \begin{eqnarray}
2043: \Gamma^{(2,m)} \Big(K';K;\bar{K}_1,\dots,\bar{K}_m\Big)
2044: =
2045: \frac{-i}{i\bar{\omega}_l - {\bf{v}}_{F,\sigma} \cdot \bar{\bf{k}}_l}
2046: \Bigg[
2047: \Gamma^{(2,m-1)} \Big(K';K+\bar{K}_l;
2048: \bar{K}_1,\dots,\bar{K}_{l-1},\bar{K}_{l+1},\dots,\bar{K}_m\Big)
2049: &&\nonumber\\
2050: -
2051: \Gamma^{(2,m-1)} \Big(K'-\bar{K}_l;K;
2052: \bar{K}_1,\dots,\bar{K}_{l-1},\bar{K}_{l+1},\dots,\bar{K}_m\Big)
2053: \Bigg] \; , &&
2054: \label{eq:WIm}
2055: \end{eqnarray}
2056: where $1 \leq l \leq m$. For clarity let us write down here the
2057: special case $m=2$,
2058: \begin{eqnarray}
2059: \Gamma^{(2,2)} \Big(K + \bar{K}_1 + \bar{K}_2 ;K;\bar{K}_1,\bar{K}_2\Big)
2060: =
2061: \frac{-i}{i\bar{\omega}_1 - {\bf{v}}_{F , \sigma} \cdot \bar{\bf{k}}_1}
2062: \Big[
2063: \Gamma^{(2,1)} \Big(K + \bar{K}_1 + \bar{K}_2 ; K+\bar{K}_1;
2064: \bar{K}_2\Big)
2065: -
2066: \Gamma^{(2,1)} \Big(K + \bar{K}_2;K;
2067: \bar{K}_2\Big)
2068: \Big]
2069: &&
2070: \nonumber
2071: \\
2072: && \hspace{-140mm} = \frac{-i}{i\bar{\omega}_2 -
2073: {\bf{v}}_{F, \sigma} \cdot \bar{\bf{k}}_2}
2074: \Big[
2075: \Gamma^{(2,1)} \Big(K + \bar{K}_1 + \bar{K}_2 ; K+\bar{K}_2;
2076: \bar{K}_1\Big)
2077: % &&\nonumber\\
2078: -
2079: \Gamma^{(2,1)} \Big(K + \bar{K}_1;K;
2080: \bar{K}_1\Big)
2081: \Big]
2082: \; .
2083: \label{eq:WI2}
2084: \end{eqnarray}
2085: \end{widetext}
2086: Diagrammatic representations of the Ward identities given in
2087: Eqs.~(\ref{eq:WI1}) and (\ref{eq:WIm}) are shown in
2088: Fig.~\ref{fig:ward}.
2089: \begin{figure}[b]
2090: \begin{center}
2091: \epsfig{file=ward.eps,width=1.0\hsize}
2092: \end{center}
2093: \caption{
2094: (a) Diagrammatic representation of the Ward identity (\ref{eq:WI1})
2095: for the three-legged vertex and (b) of the Ward identity
2096: (\ref{eq:WIm}) for the vertex with two fermion legs and $m>1$
2097: boson legs. The small arrow indicates the place in the diagram
2098: where the external bosonic energy-momentum enters. A double slash
2099: to the right of an arrow means that the bosonic momentum is added
2100: before the corresponding Green's function, while a double slash to
2101: the left of an arrow means that the momentum is added after the
2102: Green's function. }
2103: \label{fig:ward}
2104: \end{figure}
2105: %
2106: To prove that these Ward identities indeed solve our infinite system
2107: of flow equations given by Eqs.~(\ref{eq:flowsigmatransfer}) and
2108: (\ref{eq:flowdisord}), we start from the flow equation for
2109: $\Gamma^{(2, m+1)}$. Substituting on both sides of this exact flow
2110: equation the Ward identities, we can reduce it to a new flow equation
2111: involving only vertices where the number of boson legs is reduced by
2112: one, but with an external bosonic momentum entering the vertices at
2113: various places. Graphically, we indicate the place where the bosonic
2114: momentum enters the vertex by a double slash, as shown in
2115: Fig.~\ref{fig:ward}. The important point is now that all diagrams
2116: with double slashes attached to intermediate Green's functions cancel
2117: due to the fact that all vertices $\Gamma^{(2,m)}$ can be expressed in
2118: terms of a difference of vertices $\Gamma^{(2,m-1)}$, with a same
2119: prefactor that is independent of $m$. Graphically, only the diagrams
2120: with a double slash attached to the leftmost or rightmost Green's
2121: function survive. Canceling the common prefactor, it is then easy to
2122: see that the RG equation derived in this way from the functional RG
2123: equation for $\Gamma^{(2,m+1)}$ is nothing but the exact RG equation
2124: for $\Gamma^{(2,m)}$. Hence, the Ward identities provide relations
2125: between the vertices $\Gamma^{(2,m)}$ that are consistent with the
2126: relations implied by the exact hierarchy of RG flow equations in the
2127: momentum-transfer cutoff scheme. In other words, the Ward identities
2128: are the solutions of the infinite hierarchy of flow equations!
2129:
2130: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2131: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2132: \subsection{Exact solution of the Tomonaga-Luttinger model via the exact
2133: RG}
2134: \label{subsec:exactTLM}
2135:
2136: Given the cascade of Ward identities (\ref{eq:WI1}) and (\ref{eq:WIm})
2137: we can close the integro\-differential equation
2138: (\ref{eq:flowsigmatransfer}) for the irreducible self-energy. Note
2139: that this equation involves both the three-legged vertex and the
2140: four-legged vertex with two fermion legs and two boson legs, so that
2141: the Ward identity (\ref{eq:WI1}) is not sufficient to close the flow
2142: equation. Of course, if one is only interested in calculating the
2143: Green's function of the TLM, it is simpler to start from the skeleton
2144: equation for the self-energy shown in Fig.~\ref{fig:skeletonsigmapi},
2145: which can be closed by means of the Ward identity (\ref{eq:WI1}) for
2146: the three-legged vertex only. Nevertheless, it is instructive to see
2147: how the exact solution emerges within the framework of the functional
2148: RG. Substituting Eqs.~(\ref{eq:WI1}) and (\ref{eq:WI2}) into
2149: Eq.~(\ref{eq:flowsigmatransfer}), we obtain the following
2150: integro\-differential equation for the electronic self-energy:
2151: \begin{eqnarray}
2152: \partial_{\Lambda}\Sigma_{\sigma} (K) & = & G_{\sigma}^{-2}(K)
2153: \int_{\bar{K}}\frac{\dot{F}_{\sigma \sigma}(\bar{K})}{
2154: (i\bar{\omega}- {\bf{v}}_{F, \sigma} \cdot \bar{\bf{k}} )^2}
2155: \nonumber
2156: \\
2157: & & \times
2158: \left[G_{\sigma}(K)-G_{\sigma}(K+\bar{K})\right]\,.
2159: \label{eq:sigmaflowTLM}
2160: \end{eqnarray}
2161: Here the index $\sigma$ labels not only the different spin species,
2162: but also the different patches of the sectorized Fermi
2163: surface.\cite{Kopietz97} For example, for the spinless case $\sigma =
2164: \pm k_F$. Using the fact that in the momentum-transfer cutoff scheme
2165: $G^2 \partial_{\Lambda} \Sigma = \partial_{\Lambda} G$ we can
2166: alternatively write Eq.~(\ref{eq:sigmaflowTLM}) as a {\it{linear}}
2167: integro\-differential equation for the fermionic Green's function,
2168: \begin{eqnarray}
2169: \partial_{\Lambda}G_{\sigma}(K)
2170: & = &
2171: \int_{\bar{K}}\frac{\dot{F}_{\sigma\sigma}(\bar{K})}{(i\bar{\omega}-
2172: {\bf{v}}_{F ,\sigma} \cdot \bar{\bf{k}} )^2}
2173: \nonumber
2174: \\
2175: & & \times
2176: \left[G_{\sigma}(K)-G_{\sigma}(K+\bar{K})\right]\,.
2177: \label{eq:Gflow_ex}
2178: \end{eqnarray}
2179: If we had simply set the vertex $\Gamma^{(2,2)}$ equal to zero in
2180: Eq~(\ref{eq:flowsigmatransfer}) and had then closed this equation by
2181: means of the Ward identity (\ref{eq:WI1}), we would have obtained a
2182: nonlinear equation. Thus, the linearity of Eq.~(\ref{eq:Gflow_ex}) is
2183: the result of a cancellation of nonlinear terms arising from both
2184: Ward identities (\ref{eq:WI1}) and (\ref{eq:WI2}). Because the second
2185: term on the right-hand side of Eq.~(\ref{eq:Gflow_ex}) is a
2186: convolution, we can easily solve this equation by means of a Fourier
2187: transformation to imaginary time and real space. Defining
2188: \begin{eqnarray}
2189: G_{\sigma} ( X ) & = &
2190: \int_K e^{ i ( {\bf{k}} \cdot {\bf{r}} - \omega \tau )}
2191: G_{\sigma }(K)\,,
2192: \label{eq:FTGdef}
2193: \\
2194: H_{ \Lambda , \sigma } ( X )
2195: & = &
2196: \int_{\bar{K}} e^{ i ( \bar{\bf{k}} \cdot {\bf{r}} - \bar{\omega} \tau )}
2197: \frac{\dot{F}_{\sigma\sigma}(\bar{K})}{(i\bar{\omega}-
2198: {\bf{v}}_{F ,\sigma} \cdot \bar{\bf{k}} )^2}
2199: \; ,
2200: \label{eq:Hsigmadef}
2201: \end{eqnarray}
2202: where $ X = (\tau , {\bf{r}} )$, the flow equation (\ref{eq:Gflow_ex})
2203: is transformed to
2204: \begin{equation}
2205: \big[\partial_{\Lambda}+H_{ \Lambda , \sigma }(X) - H_{ \Lambda , \sigma }(0)\Big]G_{\sigma}(X) = 0\,.
2206: \end{equation}
2207: This implies the conservation law
2208: \begin{equation}
2209: \partial_{\Lambda}\left[
2210: \exp\left\{
2211: \int_0^{\Lambda}d\Lambda'\left[
2212: H_{ \Lambda' , \sigma }(X)-H_{ \Lambda' , \sigma }(0)
2213: \right]
2214: \right\}
2215: G_{\sigma}(X)
2216: \right] = 0\,.
2217: \end{equation}
2218: Integrating from $\Lambda=0$ to $\Lambda=\Lambda_0$, we obtain
2219: \begin{equation}
2220: G_{\sigma}(X)=G_{0 , \sigma}(X)\,\exp\left[Q_{\sigma}(X)\right]\,,
2221: \label{eq:ex_sol}
2222: \end{equation}
2223: with
2224: \begin{equation}
2225: Q_{\sigma}(X) = S_{\sigma}(0) - S_{\sigma}(X)\,,
2226: \label{eq:Qdef}
2227: \end{equation}
2228: and
2229: \begin{eqnarray}
2230: S_{\sigma}(X)& = & - \int_0^{\Lambda_0}d\Lambda'H_{\Lambda', \sigma }(X)
2231: \nonumber
2232: \\
2233: & & \hspace{-15mm} = \int_{\bar{K}}\frac{\Theta(\Lambda_0 - |\bar{\bf{k}}| )
2234: F_{\sigma\sigma}(\bar{K})}
2235: {(i\bar{\omega}- {\bf{v}}_{F ,\sigma} \cdot \bar{\bf{k}} )^2}
2236: \cos( \bar{\bf{k}} \cdot {\bf{r}} - \bar{\omega} \tau)\,,
2237: \label{eq:DW}
2238: \end{eqnarray}
2239: where we have used the invariance of the RPA interaction $F ( \bar{K}
2240: )$ under $\bar{K}\to-\bar{K}$. The solution in
2241: Eqs.~(\ref{eq:ex_sol})--(\ref{eq:DW}) is well known from the
2242: functional integral approach to bosonization
2243: \cite{Kopietz97,Kopietz95,Fogedby76,Lee88} where $Q_{\sigma}(X)$
2244: arises as a Debye-Waller factor from Gaussian averaging over the
2245: distribution of the Hubbard-Stratonovich field. In one dimension,
2246: Eqs.~(\ref{eq:ex_sol})--(\ref{eq:DW}) can be shown \cite{Kopietz97} to
2247: be equivalent to the exact solution for the Green's function of the
2248: Tomonaga-Luttinger model obtained via conventional bosonization.
2249:
2250: Once the exact single-particle Green's function is known, the Ward
2251: identities in Eqs.~(\ref{eq:WI1}) and (\ref{eq:WIm}) iteratively yield
2252: expressions for the vertices $\Gamma^{(2,m)}$ that solve the whole
2253: hierarchy of flow equations for the vertices with two fermion and an
2254: arbitrary number of boson legs. In principle, the method described in
2255: this section can be applied also to vertices with more than two
2256: fermion legs. For example, the right-hand sides of the flow equations for the
2257: vertices $\Gamma^{(4,m)}$ contain only vertices with no more than four
2258: fermion legs. Ward identities for these vertices would again yield a
2259: solution of this complete hierarchy, once the vertices
2260: $\Gamma^{(2,m)}$ are known. This procedure can be iterated to obtain
2261: vertices with an arbitrary number of external legs using at each step
2262: the complete flow of vertices with two fewer fermion legs obtained in
2263: the previous step. We have thus devised a method to obtain all
2264: correlation functions of the TLM entirely within the framework of the
2265: functional RG.
2266:
2267:
2268: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2269: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2270: \subsection{Truncation scheme based on relevance}
2271: \label{subsec:truncrel}
2272:
2273: The structure of the exact Green's function of the TLM and the
2274: corresponding spectral function $A ( k , \omega )=-\pi^{-1}\text{Im}\,
2275: G(k,\omega+i0)$ depend crucially on the Ward identities
2276: discussed above, which in turn are only valid if the energy dispersion
2277: is strictly linear. In order to assess the validity of the
2278: linearization of the energy dispersion, it is important to develop
2279: truncations of the exact hierarchy of flow equations that do not
2280: explicitly make use of the validity of the asymptotic Ward identities.
2281: We now propose such a truncation scheme.
2282:
2283: The coefficients generated in the expansion of a given vertex
2284: $\Gamma^{(2n,m)}(K'_1,\dots,K'_n; K_1, \dots,
2285: K_n;\bar{K}_1,\dots,\bar{K}_m) $ in powers of frequencies and momenta
2286: have decreasing scaling dimensions, so that the most relevant part of
2287: any vertex is obtained by setting all momenta and frequencies equal to
2288: zero. This classification leads to a simple truncation scheme: We
2289: retain only those vertices whose leading (momentum- and
2290: frequency-independent) part has a positive or vanishing scaling
2291: dimension, corresponding to relevant or marginal couplings in the
2292: usual RG jargon. In the context of calculating the critical
2293: temperature of the weakly interacting Bose gas in three dimensions,
2294: such a truncation procedure has recently been shown to give very
2295: accurate results.\cite{Ledowski04}
2296:
2297: To begin with, let us classify all couplings according to their
2298: relevance. With the rescaling defined in Sec.~\ref{subsec:rescale},
2299: for $D = z_{\psi} = z_{\varphi} = 1$, the scaling dimensions of the
2300: vertices $\tilde{\Gamma}^{(2n,m)}$ are $D^{(2n,m)} = 2 -n -m$, see
2301: Eq.~(\ref{eq:scaledim}). Hence the vertex $\tilde{\Gamma}^{(2,2)}$ as
2302: well as the vertices $\tilde{\Gamma}^{(0,3)}$ and
2303: $\tilde\Gamma^{(0,4)}$, whose unrescaled versions appear on the
2304: right-hand sides of Eqs.~(\ref{eq:flowsigmatransfer}) and
2305: (\ref{eq:flowPitransfer}), are irrelevant in the RG sense. In
2306: contrast, the momentum- and frequency-independent part of the
2307: three-legged vertex,
2308: \begin{equation}
2309: \tilde{\gamma}_l = \tilde{\Gamma}^{(2,1)} ( 0; 0;0) = \left( \frac{\Lambda}{\nu_0 \Omega_{\Lambda} }
2310: \right)^{1/2} Z_l \Gamma^{(2,1)} ( K_F ; K_F ; 0)
2311: \; ,
2312: \label{eq:tildegdef}
2313: \end{equation}
2314: is marginal.\cite{Ueda84} Here $K_F = ( \pm k_F , \omega =0)$. From
2315: the general flow equations (\ref{eq:flowGammarescale1}) and
2316: (\ref{eq:inhrescale1}) for the rescaled vertices we obtain the
2317: following exact flow equation for the rescaled self-energy defined in
2318: Eq.~(\ref{eq:Sigmarescaledef}):
2319: \begin{equation}
2320: \partial_l \tilde{\Sigma}_l ( Q ) = \left( 1 - \eta_l + Q \cdot \frac{ \partial}{\partial Q} \right)
2321: \tilde{\Sigma}_l ( Q ) + \dot{\tilde{\Gamma}}^{(2,0)}_l ( Q )
2322: \label{eq:flowsigma}
2323: \; ,
2324: \end{equation}
2325: with [see Eq.~(\ref{eq:inhrescale1})]
2326: \begin{equation}
2327: \dot{\tilde{\Gamma}}^{(2,0)}_l ( Q ) = - \frac{Z_l}{\Omega_{\Lambda}} \Lambda \partial_{\Lambda}
2328: \Gamma_{\Lambda}^{(2,0)} ( K )
2329: \; .
2330: \end{equation}
2331: We restrict ourselves to spinless fermions here and choose
2332: $\Omega_{\Lambda} = \bar{\Omega}_{\Lambda} = v_F \Lambda$, so that
2333: with $\nu_0=(\pi v_F)^{-1}$ the prefactor in Eq.~(\ref{eq:tildegdef})
2334: turns out to be $ ( \frac{\Lambda}{\nu_0 \Omega_{\Lambda} } )^{1/2} =
2335: \pi^{1/2}$. As usual, the fermionic wave-function renormalization
2336: factor $Z_l$ is defined via
2337: \begin{equation}
2338: Z_l = \left[ 1 - \left. \frac{\partial \Sigma ( K ) }{\partial ( i \omega ) } \right|_{ K =0} \right]^{-1}
2339: = 1 + \left. \frac{\partial \tilde{\Sigma}_l ( Q ) }{\partial ( i \epsilon ) } \right|_{ Q =0}
2340: \label{eq:Zdefexplicit}
2341: \; .
2342: \end{equation}
2343: According to Eq.~(\ref{eq:etadef}) the wave-function renormalization
2344: $Z_l$ satisfies the flow equation
2345: \begin{equation}
2346: \partial_lZ_l= -\eta_l Z_l\;,
2347: \end{equation}
2348: where the flowing anomalous dimension of the fermion fields is given
2349: by
2350: \begin{equation}
2351: \eta_l = - \left. \frac{\partial \dot{\tilde{\Gamma}}^{(2,0)}_l ( Q ) }{\partial ( i \epsilon ) } \right|_{ Q =0}
2352: \; .
2353: \label{eq:etaexplicit}
2354: \end{equation}
2355: According to Eq.~(\ref{eq:flowsigma}) the constant part of the
2356: self-energy,
2357: \begin{equation}
2358: \tilde{r}_l = \tilde{\Sigma}_l ( 0 )
2359: \; ,
2360: \label{eq:rldef}
2361: \end{equation}
2362: is relevant and satisfies
2363: \begin{equation}
2364: \partial_l \tilde{r}_l = \left( 1 - \eta_l \right)
2365: \tilde{r}_l + \dot{\tilde{\Gamma}}^{(2,0)}_l ( 0 )
2366: \label{eq:flowr}
2367: \; .
2368: \end{equation}
2369: In general, $\tilde{r}_l$ will only flow into the fixed point if the
2370: initial coupling $\tilde{r}_0$ is properly fine-tuned. Apart from
2371: $Z_l$, there are two more marginal couplings. The first is the Fermi
2372: velocity renormalization factor \cite{Busche02}
2373: \begin{equation}
2374: \tilde{v}_l = Z_l + \left.
2375: \frac{\partial \tilde{\Sigma}_l ( Q ) }{\partial q } \right|_{ Q =0}
2376: \; ,
2377: \label{eq:vtildedef}
2378: \end{equation}
2379: and the second marginal coupling is the momentum- and
2380: frequency-independent part $\tilde{\gamma}_l$ of the rescaled
2381: three-legged vertex given in Eq.~(\ref{eq:tildegdef}). The exact flow
2382: equations for $\tilde{v}_l$ and $\tilde{\gamma}_l$ are
2383: \begin{equation}
2384: \partial_l \tilde{v}_l = - \eta_l \tilde{v}_l
2385: + \left. \frac{ \partial \dot{\tilde{\Gamma}}^{(2,0)}_l ( Q )}{\partial q }
2386: \right|_{ Q=0}
2387: \label{eq:flowv}
2388: \; ,
2389: \end{equation}
2390: and
2391: \begin{equation}
2392: \partial_l \tilde{\gamma}_l = - \eta_l \tilde{\gamma}_l
2393: + \dot{\tilde{\Gamma}}^{(2,1)}_l ( 0 ; 0;0)
2394: \label{eq:flowg}
2395: \; .
2396: \end{equation}
2397: If we retain only relevant and marginal couplings, then in the
2398: momentum-transfer cutoff scheme the rescaled fermionic Green's function
2399: defined in Eq.~(\ref{eq:rescaleprop}) is in $D=1$ simply approximated
2400: by
2401: \begin{equation}
2402: \tilde{G} ( Q ) \approx \frac{1}{i \epsilon - \tilde{v}_l q - \tilde{r}_l }
2403: \; .
2404: \label{eq:Gtildeapprox}
2405: \end{equation}
2406: In order to make progress, we have to approximate the inhomogeneities
2407: $\dot{\tilde{\Gamma}}^{(2,0)}_l ( Q )$ and
2408: $\dot{\tilde{\Gamma}}^{(2,1)}_l ( 0 ; 0;0)$. In
2409: Sec.~\ref{subsec:truncation} we have proposed an approximation scheme
2410: which retains only the skeleton elements of the two-point functions.
2411: In the momentum-transfer cutoff scheme, the corresponding flow
2412: equations (\ref{eq:flowSigmatrunc},\ref{eq:flowPitrunc},
2413: \ref{eq:flowGammatrunc}) further simplify because we should omit all
2414: terms involving the fermionic single-scale propagator. Unfortunately,
2415: the resulting nonlinear integro\-differential equations still cannot
2416: be solved analytically. In order to simplify these equations further,
2417: let us replace the three-legged vertex on the right-hand sides of
2418: these equations by its marginal part. In this approximation we obtain,
2419: from Eq.~(\ref{eq:flowSigmatrunc}),
2420: \begin{equation}
2421: \dot{\tilde{\Gamma}}^{(2,0)}_l ( Q ) \approx %-
2422: \tilde{\gamma}_l^2 \int_{ \bar{Q}}
2423: \dot{ \tilde{F}} ( \bar{Q} ) \tilde{G} ( Q + \bar{Q} )
2424: \label{eq:dotGtrunc}
2425: \; ,
2426: \end{equation}
2427: and from Eq.~(\ref{eq:flowGammatrunc}),
2428: \begin{equation}
2429: \dot{\tilde{\Gamma}}^{(2,1)}_l ( 0;0;0 ) \approx %-
2430: \tilde{\gamma}_l^3 \int_{ \bar{Q}}
2431: \dot{ \tilde{F}} ( \bar{Q} ) \tilde{G}^2 ( \bar{Q} )
2432: \label{eq:dotGammatrunc}
2433: \; .
2434: \end{equation}
2435: In order to be consistent, we should approximate $\tilde{G} ( Q )$ in
2436: Eqs.~(\ref{eq:dotGtrunc}) and (\ref{eq:dotGammatrunc}) by
2437: Eq.~(\ref{eq:Gtildeapprox}). Then it is easy to see that the second
2438: term on the right-hand sides of the flow equations (\ref{eq:flowv})
2439: and (\ref{eq:flowg}) exactly cancels the contribution from the
2440: anomalous dimension, so that
2441: \begin{equation}
2442: \partial_l \tilde{\gamma}_l = 0
2443: \;, \; \; \; \partial_l \tilde{v}_l = 0
2444: \; .
2445: \label{eq:flowgzero}
2446: \end{equation}
2447: For explicit calculations, let us assume that the usual couplings of
2448: the TLM\cite{Solyom79} are $g_2 = g_4 = f_0$, so that
2449: \begin{equation}
2450: \dot{ \tilde{F}} ( \bar{Q} ) = %-
2451: \delta ( 1 - | \bar{q} | )
2452: \frac{ \tilde{f}_0 (\bar{q}^2 + \bar{\epsilon}^2)}{ ( 1 + \tilde{f}_0) \bar{q}^2 + \bar{\epsilon}^2}
2453: \; ,
2454: \label{eq:frpatlm}
2455: \end{equation}
2456: where $\tilde{f}_0 = \nu_0 f_0$. From Eqs.~(\ref{eq:etaexplicit}) and
2457: (\ref{eq:dotGtrunc}) we then find that the anomalous dimension
2458: $\eta=\eta_l$ does not flow and is given by
2459: \cite{footnoteeta}
2460: \begin{equation}
2461: \eta = \frac{ \tilde{f}_0^2}{ 2 \sqrt{ 1 + \tilde{f}_0}
2462: \left[ \sqrt{ 1 + \tilde{f}_0} + 1 \right]^2}
2463: \; ,
2464: \label{eq:etares}
2465: \end{equation}
2466: which agrees exactly with the bosonization
2467: result.\cite{Kopietz97,footnoteVolker} We emphasize that
2468: Eq.~(\ref{eq:etares}) is the correct anomalous dimension of the TLM,
2469: even for $\tilde{f}_0 \gg 1$, so that, at least as far as the
2470: calculation of $\eta$ is concerned, the validity of our simple
2471: truncation is not restricted to the weak coupling regime. Recall that
2472: the restriction to weak coupling is one of the shortcomings of the
2473: conventional fermionic functional
2474: RG,\cite{Zanchi96,Halboth00,Honerkamp01,Honerkamp01b,Kopietz01,Busche02,Ledowski03,Tsai01,Binz02,Meden02,Kampf03,Katanin04}
2475: which was implemented for the TLM in Ref.~\onlinecite{Busche02}.
2476: Because $\eta$ is finite, the running vertex $ \Gamma^{(2,1)} ( K_F ;
2477: K_F ; 0)$ without wave-function renormalization actually diverges for
2478: $\Lambda \rightarrow 0$. However, the properly renormalized vertex
2479: $\tilde{\gamma}_l \propto Z_l \Gamma^{(2,1)} ( K_F ; K_F ; 0 )$
2480: remains finite due to the vanishing wave-function renormalization
2481: \begin{equation}
2482: Z_l=e^{-\eta l}=\left(\frac{\Lambda}{\Lambda_0}\right)^{\eta}
2483: \;
2484: \end{equation}
2485: for $l\to\infty$. Integrating the flow equation (\ref{eq:flowsigma})
2486: for the self-energy with the inhomogeneity approximated by
2487: Eqs.~(\ref{eq:dotGtrunc}) and (\ref{eq:Gtildeapprox}), we obtain,
2488: after going back to physical variables \cite{footnoteeta}
2489: \begin{eqnarray}
2490: \Sigma ( k_F + k , i \omega ) = ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~&&
2491: \nonumber\\
2492: - \int_{ - \Lambda_0}^{\Lambda_0}
2493: \frac{ d \bar{k}}{ 2 \pi} \int_{ - \infty}^{\infty}
2494: \frac{ d \bar{\omega}}{ 2 \pi}
2495: \left( \frac{\Lambda_0}{ | \bar{k} | } \right)^{\eta}
2496: \frac{ f^{\rm RPA} ( \bar{k} , i \bar{\omega} )}{ i ( \omega + \bar{\omega} )
2497: - v_F ( k + \bar{k} ) }\; ,&&
2498: \nonumber\\
2499: \label{eq:sigmatruncres}
2500: \end{eqnarray}
2501: where the RPA screened interaction is
2502: \begin{equation}
2503: f^{\rm RPA} ( \bar{k} , i \bar{\omega} ) = f_0
2504: \frac{ v_F^2 \bar{k}^2 + \bar{\omega}^2}{
2505: v_c^2 \bar{k}^2 + \bar{\omega}^2}
2506: \; .
2507: \end{equation}
2508: Here $v_c = v_F \sqrt{ 1 + \tilde{f}_0}$ is the velocity of collective
2509: charge excitations. Equation (\ref{eq:sigmatruncres}) resembles the GW
2510: approximation,\cite{Hedin65} but with the RPA interaction multiplied
2511: by an additional singular vertex correction $ ( \Lambda_0 / |
2512: {\bar{k}} |)^{ \eta}$. The explicit evaluation of
2513: Eq.~(\ref{eq:sigmatruncres}) is rather tedious and will not be further
2514: discussed in this work. The resulting spectral function $A ( k ,
2515: \omega )$ agrees at $k = k_F$ with the bosonization result (even at
2516: strong coupling), but has the wrong threshold singularities for $ |
2517: \omega | \rightarrow v_c | ( k \pm k_F) | $. So far we have not been
2518: able to find a reasonably simple truncation of the exact flow
2519: equations which completely produces the spectral line shape of $A ( k
2520: , \omega )$, as predicted by bosonization or by our exact solution
2521: presented in the previous section. Whether a self-consistent numerical
2522: solution of the truncation discussed in Sec.~\ref{subsec:truncation}
2523: [see Eqs.~(\ref{eq:flowSigmatrunc})-(\ref{eq:flowGammatrunc})] would
2524: reproduce the correct spectral line shape or not remains an open
2525: problem. The numerical solution of these equations seems to be rather
2526: difficult and is beyond the scope of this work.
2527:
2528: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2529: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2530: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2531: \section{Summary and outlook}
2532: \label{sec:summary}
2533:
2534: In this work we have developed a new formulation of the functional RG
2535: for interacting fermions, which is based on the explicit introduction
2536: of collective bosonic degrees of freedom via a suitable
2537: Hubbard-Stratonovich transformation. A similar strategy has been used
2538: previously in Refs.~\onlinecite{Correia01}, \onlinecite{Wetterich04},
2539: and \onlinecite{Baier03}. However, on the technical level the
2540: practical implementation of this method presented here differs
2541: considerable from previous works. We have payed particular attention
2542: to asymptotic Ward identities, which play a crucial role if the
2543: interaction is dominated by small momentum transfers. In one
2544: dimension, this is the key to obtain the exact solution of the
2545: Tomonaga-Luttinger model entirely within the functional RG. By using
2546: the momentum transfer associated with the bosonic field as a cutoff
2547: parameter, we have formulated the functional RG in such a way that the
2548: RG flow does not violate the Ward identities. In fact, we have shown
2549: that Ward identities emerge as the solution of the infinite hierarchy
2550: of coupled RG flow equations for the one-line irreducible vertices
2551: involving two external fermion legs and an arbitrary number of boson
2552: legs. In principle this method can be iterated to obtain all
2553: correlation functions of the TLM entirely within the framework of the
2554: functional RG.
2555:
2556: Here we have mainly laid the theoretical foundation of our approach
2557: and developed an efficient method to keep track of all terms. In
2558: future work, we are planning to apply our technique to other
2559: physically interesting problems. Let us mention some problems where
2560: it might be advantageous to use our approach:
2561:
2562: (a) {\it{Strong coupling fixed points.}} One of the big drawbacks of
2563: the conventional (purely fermionic) functional RG used by many authors
2564: \cite{Zanchi96,Halboth00,Honerkamp01,Honerkamp01b,Tsai01,Binz02,Meden02,Kampf03,Katanin04}
2565: is that in practice the frequency dependence of the four-point vertex
2566: $\Gamma^{(4)}$ has to be neglected, so that the wave-function
2567: renormalization factor is $Z=1$. Although the resulting runaway flow
2568: of the vertices to strong coupling at a finite scale can be
2569: interpreted in terms of corresponding instabilities, there is the
2570: possibility that for small $Z$ the renormalized effective interaction
2571: $Z^2\Gamma^{(4)}$ remains finite even though the vertex $\Gamma^{(4)}$
2572: without wave-function renormalization seems to diverge.\cite{Ferraz03}
2573: In our approach, the effective interaction acquires a
2574: frequency dependence, even within the lowest-order approximation. In
2575: fact, if we ignore vertex corrections, the effective interaction is
2576: simply given by the RPA. Hence, strong coupling fixed points might be
2577: accessible within our approach. Recall that the rather simple
2578: truncation of Sec.~\ref{subsec:truncrel} gave the exact anomalous
2579: dimension of the TLM for arbitrary strength of the interaction.
2580: Possibly, more elaborate truncations of the exact hierarchy of RG flow
2581: equations (for example, the truncation based on retaining skeleton
2582: elements of the two-point functions discussed in
2583: Sec.~\ref{subsec:truncation}, see Fig.~\ref{fig:flowtrunc}) will give
2584: accurate results for the spectral properties.
2585:
2586:
2587: (b) {\it{Nonuniversal effects in one-dimensional metals.}} If we do
2588: not linearize the energy dispersion in one dimension, there should be
2589: a finite momentum scale $k_c$ (depending on the interaction and the
2590: band curvature) below which typical scaling behavior predicted by the
2591: TLM emerges. The calculation of $k_c$ as well as the associated
2592: nonuniversal spectral line shape are difficult within
2593: bosonization.\cite{Busche01} On the other hand, within the framework
2594: of the functional RG the inclusion of irrelevant couplings is
2595: certainly possible, so that with our method it might be possible to
2596: shed some new light onto this old problem. For an explicit
2597: calculation of an entire crossover scaling function between the
2598: critical regime and the short-wavelength regime of interacting bosons
2599: in $D=3$, see Ref.~\onlinecite{Ledowski04}. An analogous calculation
2600: of the dynamic scaling functions for interacting fermions in one
2601: dimension remains to be done.
2602:
2603:
2604: (c) {\it{Itinerant ferromagnetism}.} Spontaneous ferromagnetism in
2605: Fermi systems is driven by sufficiently strong interactions involving
2606: small momentum transfers. Assuming a given form of the ferromagnetic
2607: susceptibility, Altshuler, Ioffe, and Millis \cite{Altshuler94}
2608: concluded on the basis of an elaborate diagrammatic analysis that in
2609: the vicinity of the paramagnetic-ferromagnetic quantum-critical point
2610: in dimensions $D = 2$ a simple one-loop calculation already yields the
2611: correct qualitative behavior of the electronic self-energy. If this is
2612: true, then in this problem vertex corrections are irrelevant.
2613: Unfortunately, due to the peculiar momentum and frequency dependence
2614: of the ferromagnetic susceptibility $\chi ( \bar{{\bf{k}}} ,
2615: \bar{\omega})$ at the quantum critical point, the assumption of asymptotic
2616: velocity conservation [see Eq.~(\ref{eq:linearization}) in Appendix C]
2617: leading to the Ward identity (\ref{eq:WI1})-(\ref{eq:WI2}) is not
2618: justified. Note, however, that in Ref.~\onlinecite{Altshuler94} the
2619: form of the susceptibility is assumed to be given. The feedback of
2620: the collective ferromagnetic fluctuations on the non-Fermi-liquid form
2621: of the electronic properties has not been discussed. The fact that in
2622: $D < 3$ the leading interaction corrections to the inverse
2623: susceptibility $\chi^{-1} ( \bar{{\bf{k}}} , \bar{\omega})$ generate a
2624: nonanalytic momentum dependence\cite{Belitz97,Chubukov03} suggests
2625: that the problem should be reconsidered taking the interplay between
2626: fermionic single-particle excitations and collective magnetic
2627: fluctuations into account. The formalism developed in this work might
2628: be suitable to shed some new light also onto this problem. The
2629: electronic properties of a two-dimensional Fermi system in the
2630: vicinity of a ferromagnetic instability have recently been studied in
2631: Ref.~\onlinecite{Katanin04b}. However, these authors focused on the
2632: finite-temperature properties of the phase transition; they did not
2633: attempt to calculate the fermionic single-particle Green's function in
2634: the vicinity of the zero-temperature phase transition. Note also that
2635: for sufficiently strong interactions even one-dimensional fermions can
2636: in principle have a ferromagnetic instability if the energy dispersion
2637: is nonlinear.\cite{Bartosch03,Yang04}
2638:
2639: (d) {\it{Quantum phase transitions and symmetry breaking.}} Our
2640: method unifies the traditional approach to quantum-phase transitions
2641: pioneered by Hertz\cite{Hertz76} with the modern developments in the
2642: field of fermionic functional RG, so that it might simplify the
2643: theoretical description of quantum phase transitions in situations
2644: where the fermions cannot be completely integrated out. In order to
2645: describe quantum phase transitions in interacting Fermi systems within
2646: the framework of the traditional Ginzburg-Landau-Wilson approach, all
2647: soft modes in the system should be explicitly
2648: retained.\cite{Kirkpatrick96} In the purely fermionic functional RG
2649: \cite{Zanchi96,Halboth00,Honerkamp01,Honerkamp01b,Salmhofer01,Kopietz01,Busche02,Ledowski03,Tsai01,Binz02,Meden02,Kampf03,Katanin04}
2650: symmetry breaking manifests itself via the divergence of the relevant
2651: order-parameter susceptibility; the symmetry broken phase is difficult
2652: to describe within this approach. On the other hand, in our approach
2653: the order parameter can be introduced explicitly as a bosonic field,
2654: which acquires a vacuum expectation value in the symmetry broken
2655: phase. Previously, a similar approach has been developed in
2656: Ref.~\onlinecite{Baier03} to study antiferromagnetism in the
2657: two-dimensional Hubbard model.
2658:
2659:
2660: In summary, we believe that the formulation of the exact functional RG
2661: presented in this work will be quite useful in many different physical
2662: contexts.
2663:
2664:
2665:
2666: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2667: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2668: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2669: \section*{ACKNOWLEDGMENTS}
2670:
2671: This work was completed while two of us (F. S. and P. K.) participated
2672: at the {\it{Winter School on Renormalization Group Methods for
2673: Interacting Electrons}} at {\it{The International Center of
2674: Condensed Matter Physics (ICCMP)}} of the University of
2675: Bras\'\i{}lia, Brazil. This gave us the opportunity to discuss the
2676: subtleties of the functional RG for Fermi systems with many colleagues
2677: -- we thank all of them. We also thank Alvaro Ferraz and the very
2678: friendly staff of the ICCMP for their hospitality. This work was
2679: financially supported by the DFG, Grant No. KO 1442/5-3 (F. S. and
2680: P. K.) and Grant No. BA 2263/1-1 (L. B.).
2681:
2682:
2683: % ++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++
2684: % ++++++++++++++++++++++++++ APPENDIX ++++++++++++++++++++++++++++++++++++
2685: % ++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++
2686:
2687:
2688: \begin{appendix}
2689:
2690: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2691: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2692: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2693: \section{Tree expansion of connected Green's functions
2694: in terms of one-line irreducible vertices}
2695: \label{sec:tree}
2696:
2697: In this Appendix we show explicitly that the vertices $\Gamma^{(n)}_{
2698: \alpha_1 \ldots \alpha_n}$ defined in terms of the functional Taylor
2699: expansion of $\Gamma [ \Phi ]$ in Eq.~ (\ref{eq:Gammaexpansion}) are
2700: indeed one-line irreducible. This is usually \cite{Negele88} done
2701: graphically by taking higher-order derivatives of the relation
2702: (\ref{eq:quad_rel}) between the second functional derivatives of $ \LL
2703: [ \Phi ]$ and $\G_c [ J ]$. With the help of our compact notation we
2704: can even give the tree expansion of the connected Green's function in
2705: terms of one-line irreducible vertices in closed form. To do so, it
2706: is advantageous to define the functional
2707: \begin{equation}
2708: \mathbf{U} =\left[\ppt{\Gamma}{\Phi}{\Phi}-\left.\ppt{\Gamma}{\Phi}{\Phi}\right|_{\Phi=0}\right]^T
2709: =\left[\ppt{\Gamma}{\Phi}{\Phi}\right]^T-\mathbf{\Sigma}
2710: \; ,
2711: \label{eq:Udef}
2712: \end{equation}
2713: which is a matrix in superindex space. With this definition, we have
2714: \begin{equation}
2715: \ppt{\LL}{\Phi}{\Phi} = \mathbf{U}^T - [ \mathbf{G}^{-1} ]^T
2716: \label{eq:LUrelation}
2717: \; ,
2718: \end{equation}
2719: so that
2720: \begin{eqnarray}
2721: \left[\ppt{\LL}{\Phi}{\Phi}\right]^{-1} & = &
2722: -\mathbf{G}^T[\mathbf{1}-\mathbf{U}^T\mathbf{G}^T]^{-1}
2723: \nonumber
2724: \\
2725: & = & -\sum_{l=0}^{\infty}\mathbf{G}^T(\mathbf{U}^T\mathbf{G}^T)^l
2726: \; .
2727: \label{eq:Gc2}
2728: \end{eqnarray}
2729: From Eq.~(\ref{eq:quad_rel}) we then obtain
2730: \begin{eqnarray}
2731: \ppt{\G_c}{J}{J} & = & \mathbf{Z}
2732: \left[ \ppt{\LL}{\Phi}{\Phi} \right]^{-1} =
2733: - \mathbf{Z} \mathbf{G}^T[\mathbf{1}-\mathbf{U}^T\mathbf{G}^T]^{-1}
2734: \nonumber
2735: \\
2736: & = & - \sum_{l=0}^{\infty} \mathbf{Z} \mathbf{G}^T(\mathbf{U}^T\mathbf{G}^T)^l
2737: \label{eq:GcJJexpansion}
2738: \; .
2739: \end{eqnarray}
2740: We now expand both sides of Eq.~(\ref{eq:GcJJexpansion}) in powers of
2741: the sources $J$ and compare coefficients. For the matrix on the
2742: left-hand side we obtain, from Eq.~(\ref{eq:Gcexpansion}),
2743: \begin{equation}
2744: \ppt{\G_c}{J}{J}
2745: =\sum_{n=0}^{\infty}\frac1{n!}\int_{\alpha_1}\dots
2746: \int_{\alpha_n} \left[\mathbf{G}^{(n+2)}_{c, \alpha_1 \dots \alpha_n}\right]^T J_{\alpha_1}
2747: \cdot\ldots\cdot J_{\alpha_n}\,,
2748: \label{eq:Gc2expansion}
2749: \end{equation}
2750: where the matrix $\mathbf{G}^{(n+2)}_{c, \alpha_1 \dots \alpha_n}$ is
2751: defined by
2752: \begin{equation}
2753: [ \mathbf{G}^{(n+2)}_{c, \alpha_1 \dots \alpha_n} ]_{\alpha \alpha^{\prime}}
2754: = \G^{(n+2)}_{c, \alpha \alpha^{\prime} \alpha_1 \dots \alpha_n }
2755: \; .
2756: \label{eq:Gcn2def}
2757: \end{equation}
2758: On the right-hand side we use Eqs.~(\ref{eq:Udef}) and
2759: (\ref{eq:Gammaexpansion}) to write
2760: \begin{equation}
2761: \mathbf{U} =\sum_{n=1}^{\infty}\frac1{n!}\int_{\alpha_1}\dots
2762: \int_{\alpha_n} \mathbf{\Gamma}^{(n+2)}_{\alpha_1,\dots,\alpha_n}\Phi_{\alpha_1}
2763: \cdot\ldots\cdot\Phi_{\alpha_n}\,,
2764: \label{eq:Uexpansion}
2765: \end{equation}
2766: where
2767: \begin{equation}
2768: [ \mathbf{\Gamma}^{(n+2)}_{\alpha_1,\dots,
2769: \alpha_n} ]_{\alpha \alpha^{\prime} }
2770: = \Gamma^{(n+2)}_{ \alpha \alpha^{\prime} \alpha_1 \dots \alpha_n }
2771: \; .
2772: \label{eq:Gammamatrix}
2773: \end{equation}
2774: To compare terms with the same powers of the sources $J$ on both sides
2775: of Eq.~(\ref{eq:GcJJexpansion}), we need to express the fields
2776: $\Phi_\alpha$ on the right-hand side of Eq.~(\ref{eq:Uexpansion}) in
2777: terms of the sources, using Eqs.~(\ref{eq:Gcexpansion}) and
2778: (\ref{eq:def_phi}),
2779: \begin{equation}
2780: \Phi_{\alpha}=\pp{\G_c}{J_{\alpha}}
2781: =
2782: \sum_{m=0}^{\infty}\frac1{m!}\int_{\beta_1}\dots
2783: \int_{\beta_m} {\G}^{(m+1)}_{c, \alpha \beta_1 \dots \beta_m} J_{\beta_1}
2784: \dots J_{\beta_m}
2785: \label{eq:PhiJexpansion}\,.
2786: \end{equation}
2787: Substituting Eqs.~(\ref{eq:Gc2expansion}), (\ref{eq:Uexpansion}), and
2788: (\ref{eq:PhiJexpansion}) into Eq.~(\ref{eq:GcJJexpansion}) and comparing terms
2789: with the same powers of the sources (after symmetrization), we obtain
2790: a general relation between the connected and the one-line irreducible
2791: correlation functions,
2792: \begin{equation}
2793: \begin{array}{rcl}
2794: \mathbf{G}_{c,\beta_1,\dots,\beta_n}^{(n+2)}&=& -\sum\limits_{l=0}^{\infty}\sum\limits_{n_1,\dots,n_l=1}^{\infty}
2795: \frac{1}{n_1!\cdot\ldots\cdot n_l!}
2796: \\[0.4cm]
2797: & & \hspace{-12mm} \times
2798: \int_{\alpha_1^1}\dots\int_{\alpha_{n_1}^1}\dots\int_{\alpha_1^l}\dots\int_{\alpha_{n_l}^l}
2799: \\[0.4cm]
2800: && \hspace{-12mm} \times \sum\limits_{m_1^1,\dots,m_{n_1}^1=1}^{\infty}\dots\sum\limits_{m_1^l,\dots,m_{n_l}^l=1}^{\infty}
2801: \delta_{n,\sum_{i=1}^l\sum_{j=1}^{n_i}m^i_j}\\[0.8cm]
2802: && \hspace{-12mm} \times \left[\mathbf{Z}\mathbf{G}^T \mathbf{\Gamma}^{(n_1+2)\,T}_{\alpha_1^1,
2803: \dots,\alpha_{n_1}^1}\mathbf{G}^T\cdot\ldots\cdot\mathbf{G}^T
2804: \mathbf{\Gamma}^{(n_l+2)\,T}_{\alpha_1^l,\dots,\alpha_{n_l}^l}\mathbf{G}^T\right]^T
2805: \,\,\\
2806: && \hspace{-12mm} \times {\cal{S}}_{\beta_1,\dots,\beta_{m_1^1};\dots;\beta_{n-m^l_{n_l}+1},\dots,\beta_n}
2807: \Big\{ \G^{(m_1^1+1)}_{c,\alpha_1^1,\beta_1, \dots,\beta_{m_1^1}}
2808: \\[0.4cm]
2809: && \hspace{-9mm}\cdot\ldots\cdot
2810: \G^{(m_{n_l}^l+1)}_{c,\alpha_{n_l}^l,\beta_{n-m^l_{n_l}+1},\dots,\beta_n}
2811: \Big\}
2812: \; .
2813: \end{array}
2814: \label{eq:GcGamma}
2815: \end{equation}
2816: On the right-hand side of this rather cumbersome expression, only
2817: connected correlation functions with a degree smaller than on the
2818: left-hand side appear. One can therefore recursively express all
2819: connected correlation functions via their one-line irreducible
2820: counterparts. Only a finite number of terms contribute on the
2821: right-hand side. The operator ${\cal{S}}$ symmetrizes the expression
2822: in curly brackets with respect to indices on different correlation
2823: functions, i.e., it generates all permutations of the indices with
2824: appropriate signs, counting expressions only once that are generated
2825: by permutations of indices on the same vertex. More precisely the
2826: action of ${\cal{S}}$ is given by ($m=\sum_{i=1}^lm_i$)
2827: \begin{eqnarray}
2828: {\cal{S}}_{\alpha_1,\dots,\alpha_{m_1};\dots;\alpha_{m-m_l+1},\dots,\alpha_{m}}
2829: \{ A_{\alpha_1,\dots,\alpha_m} \}
2830: & = &
2831: \nonumber
2832: \\
2833: & & \hspace{-65mm} \frac{1}{\prod_i m_i!}\sum_P
2834: \mathrm{sgn}_{\zeta}(P) \,
2835: A_{\alpha_{P(1)},\dots,\alpha_{P(m)}}\,,
2836: \label{eq:symmopdef}
2837: \end{eqnarray}
2838: where $P$ denotes a permutation of $\{1,\dots,m\}$ and
2839: $\mathrm{sgn}_{\zeta}$ is the sign created by permuting field
2840: variables according to the permutation $P$, i.e.,
2841: \begin{equation}
2842: \Phi_{\alpha_1}\cdot\ldots\cdot\Phi_{\alpha_m} =
2843: \mathrm{sgn}_{\zeta}(P) \,
2844: \Phi_{\alpha_{P(1)}}\cdot\ldots\cdot\Phi_{\alpha_{P(m)}}
2845: \; .
2846: \end{equation}
2847: %
2848: \begin{figure}
2849: \begin{center}
2850: \epsfig{file=GcGamma_new.eps,width=0.9\hsize}
2851: \end{center}
2852: \caption{Graphical representation of the relation between connected
2853: Green's functions and one-line irreducible vertices up to the
2854: four-point functions. The irreducible vertices are represented by
2855: shaded oriented circles with the appropriate number of legs; see
2856: Fig.~\ref{fig:generalvertex}. The connected Green's functions are
2857: drawn as empty oriented circles with a number indicating the
2858: number of external legs. }
2859: \label{fig:GcGamma}
2860: \end{figure}
2861: %
2862: A diagrammatic representation of the first few terms of the
2863: tree expansion generated by Eq.~(\ref{eq:GcGamma}) is given in
2864: Fig.~\ref{fig:GcGamma}. Let us give the corresponding analytic
2865: expressions: If we set $n=0$ in Eq.~(\ref{eq:GcGamma}), then only the
2866: term with $l=0$ contributes, and we obtain
2867: \begin{equation}
2868: \mathbf{G}_c^{(2)} = - \mathbf{Z} \mathbf{G} = - \mathbf{G}^T
2869: \; ,
2870: \label{eq:Gtree2}
2871: \end{equation}
2872: which is Eq.~(\ref{eq:Gmatrix}) in matrix form. For $n=1$ the single
2873: term with $l=1$, $n_1 =1$, $m_1^1=1$ contributes on the right-hand
2874: side of Eq.~(\ref{eq:GcGamma}). Using $\mathbf{Z} \mathbf{G} =
2875: \mathbf{G}^T$ the tree expansion of the connected Green's function with
2876: three external legs can be written as
2877: \begin{equation}
2878: \G^{(3)}_{c, \beta_1 \beta_2 \beta_3 }
2879: = \int_{\alpha_1} \int_{\alpha_2} \int_{\alpha_3}
2880: [\mathbf{G}]_{ \beta_1 \alpha_1}
2881: [\mathbf{G}]_{ \beta_2 \alpha_2}
2882: [\mathbf{G}]_{ \beta_3 \alpha_3}
2883: \Gamma^{(3)}_{ \alpha_1 \alpha_2 \alpha_3}
2884: \; .
2885: \end{equation}
2886: Finally, consider the connected Green's function with four external
2887: legs, corresponding to $n=2$ in Eq.~(\ref{eq:GcGamma}). In this case
2888: the following three terms contribute:
2889: \begin{eqnarray*}
2890: \nonumber
2891: \begin{array}{l|c|c|c}
2892: {\rm{term}} & l & n_i & m^i_j \\
2893: \colrule
2894: 1.) &1 &n_1 =1 & m_1^1 =2 \\
2895: 2.) & 1 & n_1 =2 & m_1^1 =m_2^1 = 1 \\
2896: 3.) & 2 & n_1 = n_2 =1 & m_1^1 = m_1^2 =1
2897: \end{array}
2898: \end{eqnarray*}
2899: The corresponding analytic expression is
2900: \begin{eqnarray}
2901: \G^{(4)}_{c, \beta_1 \beta_2 \beta_3 \beta_4}
2902: & = &
2903: \nonumber
2904: \\
2905: & & \hspace{-20mm} - \int_{\alpha_1} \ldots \int_{\alpha_4}
2906: [\mathbf{G}]_{ \beta_1 \alpha_1}
2907: [\mathbf{G}]_{ \beta_2 \alpha_2}
2908: [\mathbf{G}]_{ \beta_3 \alpha_3}
2909: [\mathbf{G}]_{ \beta_4 \alpha_4}
2910: \Gamma^{(4)}_{ \alpha_1 \alpha_2 \alpha_3 \alpha_4}
2911: \nonumber
2912: \\
2913: & & \hspace{-20mm}
2914: - \int_{\alpha_1} \ldots \int_{\alpha_6}
2915: [\mathbf{G}]_{ \beta_1 \alpha_1}
2916: [\mathbf{G}]_{ \beta_2 \alpha_2}
2917: [\mathbf{G}]_{ \beta_3 \alpha_3}
2918: [\mathbf{G}]_{ \beta_4 \alpha_4}
2919: \nonumber
2920: \\
2921: & & \times \Gamma^{(3)}_{ \alpha_1 \alpha_2 \alpha_5}
2922: [\mathbf{G}]_{ \alpha_5 \alpha_6}
2923: \Gamma^{(3)}_{ \alpha_6 \alpha_3 \alpha_4}
2924: \nonumber
2925: \\
2926: & & \hspace{-20mm}
2927: - \int_{\alpha_1} \ldots \int_{\alpha_6}
2928: S_{\beta_3;\beta_4} \Big\{
2929: [\mathbf{G}]_{ \beta_1 \alpha_1}
2930: [\mathbf{G}]_{ \beta_2 \alpha_2}
2931: [\mathbf{G}]_{ \beta_3 \alpha_3}
2932: [\mathbf{G}]_{ \beta_4 \alpha_4}
2933: \nonumber
2934: \\
2935: & & \times
2936: \Gamma^{(3)}_{ \alpha_1 \alpha_5 \alpha_4}
2937: [\mathbf{G}]_{ \alpha_5 \alpha_6}
2938: \Gamma^{(3)}_{ \alpha_6 \alpha_2 \alpha_3}
2939: \Big\}
2940: \; .
2941: \label{eq:Gtree4}
2942: \end{eqnarray}
2943:
2944: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2945: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2946: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2947: \section{Dyson-Schwinger equations and skeleton diagrams}
2948: \label{sec:skeleton}
2949:
2950: In this appendix we show how the skeleton diagrams for the two-point
2951: functions and the three-legged vertex in
2952: Fig.~\ref{fig:skeletonsigmapi} can be formally derived from the
2953: Dyson-Schwinger equations of motion. Although the skeleton graphs are
2954: usually written down directly from topological considerations of the
2955: structure of diagrammatic perturbation theory,\cite{Nozieres64} it is
2956: instructive to see how the skeleton expansion of the irreducible
2957: vertices can be derived formally within our functional integral
2958: approach.
2959:
2960: The invariance of the generating functional $\G [ J] $ of the Green's
2961: functions defined in Eq.~(\ref{eq:Ggen}) with respect to infinitesimal
2962: shifts in the integration variables $\Phi_{\alpha}$ implies the
2963: Dyson-Schwinger equations of motion \cite{ZinnJustin98}
2964: \begin{equation}
2965: \left( \zeta_{\alpha} J_{\alpha} -
2966: \frac{ \delta S}{\delta \Phi_{\alpha}} \left[ \frac{ \delta}{\delta J_{\alpha} } \right]
2967: \right) \G [ J_{\alpha} ] = 0
2968: \; .
2969: \label{eq:DysonSchwinger}
2970: \end{equation}
2971: For our coupled Fermi-Bose system with Euclidean action $S [
2972: \bar{\psi} , \psi , \varphi ]$ given by Eqs.~(\ref{eq:S0psidef},
2973: \ref{eq:Spsiphidef}, \ref{eq:S0phi}, \ref{eq:Spsiphi}) involving three
2974: types of fields, Eq.~(\ref{eq:DysonSchwinger}) is actually equivalent
2975: with the following three equations:
2976: \begin{eqnarray}
2977: & &
2978: \Biggl( J_{ - \bar{K} \sigma} -
2979: \sum_{\sigma^{\prime}} [ f_{ \bar{\bf{k}} }^{-1} ]^{\sigma^{\prime} \sigma}
2980: \frac{ \delta}{ \delta J_{\bar{K} \sigma^{\prime}} }
2981: \Biggr) \G
2982: \nonumber
2983: \\
2984: & & \hspace{10mm}
2985: - i \zeta \int_K \frac{ \delta^{(2)} \G }{
2986: \delta j_{ K + \bar{K} \sigma} \delta \bar{\jmath}_{ K \sigma} } =0
2987: \; ,
2988: \label{eq:DysonSchwinger1}
2989: \\
2990: & &
2991: \Biggl( \zeta \bar{\jmath}_{ K \sigma}
2992: + [ i \omega - \xi_{ {\bf{k}} \sigma } ]
2993: \frac{ \delta}{\delta j_{ K \sigma}}
2994: \Biggr) \G
2995: \nonumber
2996: \\
2997: & & \hspace{10mm}
2998: - i \int_{\bar{K}} \frac{ \delta^{(2)} \G }{
2999: \delta j_{ K + \bar{K} \sigma} \delta J_{ -\bar{K} \sigma} } =0
3000: \; ,
3001: \label{eq:DysonSchwinger2}
3002: \\
3003: & &
3004: \Biggl( j_{ K \sigma}
3005: + [ i \omega - \xi_{ {\bf{k}} \sigma } ]
3006: \frac{ \delta}{\delta \bar{\jmath}_{ K \sigma}}
3007: \Biggr) \G
3008: \nonumber
3009: \\
3010: & & \hspace{10mm}
3011: - i \int_{\bar{K}} \frac{ \delta^{(2)} \G }{
3012: \delta \bar{\jmath}_{ K - \bar{K} \sigma} \delta J_{ -\bar{K} \sigma} } =0
3013: \; .
3014: \label{eq:DysonSchwinger3}
3015: \end{eqnarray}
3016: Expressing these equations in terms of the generating functionals
3017: $\G_c [ \bar{\jmath} , j , J]$ of the connected Green's functions and
3018: the corresponding generating functional $\Gamma [ \bar{\psi} , \psi ,
3019: \varphi ]$ of the irreducible vertices defined in Eq.~(\ref
3020: {eq:Gammadef}), we obtain the Dyson-Schwinger equations of motion in
3021: the following form:
3022: \begin{eqnarray}
3023: & &
3024: \frac{ \delta \Gamma}{\delta \varphi_{ \bar{K} \sigma} }
3025: - i \int_K \left[
3026: \bar{\psi}_{ K + \bar{K}, \sigma} \psi_{ K \sigma} +
3027: \frac{ \delta^{(2)} \G_c}{ \delta \bar{\jmath}_{ K \sigma} \delta j_{ K + \bar{K} , \sigma} }
3028: \right] = 0 \; ,
3029: \nonumber
3030: \\
3031: \label{eq:DysonSchwinger4}
3032: \\
3033: & &
3034: \frac{ \delta \Gamma}{\delta \psi_{ {K} \sigma} }
3035: - i \int_{\bar{K}} \left[ \zeta
3036: \bar{\psi}_{ K + \bar{K}, \sigma} \varphi_{ \bar{K} \sigma} +
3037: \frac{ \delta^{(2)} \G_c}{ \delta j_{ K + \bar{K} , \sigma} \delta J_{ - \bar{K} \sigma}}
3038: \right] = 0 \; ,
3039: \nonumber
3040: \\
3041: \label{eq:DysonSchwinger5}
3042: \\
3043: & &
3044: \frac{ \delta \Gamma}{\delta \bar{\psi}_{ {K} \sigma} }
3045: - i \int_{\bar{K}} \left[
3046: {\psi}_{ K - \bar{K}, \sigma} \varphi_{ \bar{K} \sigma} +
3047: \frac{ \delta^{(2)} \G_c}{ \delta \bar{\jmath}_{ K - \bar{K} , \sigma} \delta J_{ - \bar{K} \sigma}}
3048: \right] = 0 \; .
3049: \nonumber
3050: \\
3051: \label{eq:DysonSchwinger6}
3052: \end{eqnarray}
3053: The second functional derivatives of $\G_c$ can be expressed in terms
3054: of the irreducible vertices using Eq.~(\ref{eq:GcJJexpansion}).
3055: Taking derivatives of
3056: Eqs.~(\ref{eq:DysonSchwinger4}--\ref{eq:DysonSchwinger6}) with respect
3057: to the fields and then setting the fields equal to zero we obtain the
3058: desired skeleton expansions of the irreducible vertices. Let us start
3059: with the skeleton diagram for the self-energy shown in
3060: Fig.~\ref{fig:skeletonsigmapi}(a). To derive this,we simply
3061: differentiate Eq.~(\ref{eq:DysonSchwinger6}) with respect to $\psi_{
3062: K^{\prime} \sigma}$. Using
3063: \begin{equation}
3064: \left. \frac{ \delta^{(2)} \Gamma}{ \delta \psi_{ K^{\prime} \sigma} \delta
3065: \bar{\psi}_{ K \sigma} }
3066: \right|_{ \rm{fields} = 0} = \delta_{ K , K^{\prime}} \Sigma_{\sigma} (K)
3067: \label{eq:Gamma2Sigma}
3068: \; ,
3069: \end{equation}
3070: we obtain
3071: \begin{equation}
3072: \delta_{ K , K^{\prime}} \Sigma_{\sigma} (K) = i
3073: \int_{ \bar{K}}
3074: \left. \frac{ \delta^{(3)} \G_c}{ \delta \psi_{ K^{\prime} \sigma}
3075: \delta \bar{\jmath}_{ K - \bar{K} , \sigma} \delta J_{ - \bar{K} \sigma}}
3076: \right|_{ \rm{fields} = 0}
3077: \; .
3078: \end{equation}
3079: From the $l=1$ term in the expansion~(\ref{eq:GcJJexpansion}) it is easy to show that
3080: \begin{eqnarray}
3081: \left. \frac{ \delta^{(3)} \G_c}{ \delta \psi_{ K^{\prime} \sigma}
3082: \delta \bar{\jmath}_{ K - \bar{K} , \sigma} \delta J_{ - \bar{K} \sigma}}
3083: \right|_{ \rm{fields} = 0} = \delta_{ K , K^{\prime}} F_{ \sigma \sigma} ( \bar{K} )
3084: &&
3085: \nonumber
3086: \\
3087: \times G_{\sigma} ( K+\bar{K} )
3088: \Gamma^{(2,1)} ( K + \bar{K} \sigma ; K \sigma ; \bar{K} \sigma )
3089: \; ,&&
3090: \end{eqnarray}
3091: so that
3092: \begin{eqnarray}
3093: \Sigma_{\sigma} (K) &=& i \int_{ \bar{K}}
3094: F_{ \sigma \sigma} ( \bar{K} ) G_{\sigma} ( K + \bar{K} )
3095: \nonumber\\
3096: &&~~~~\times\Gamma^{(2,1)} ( K + \bar{K} \sigma ; K \sigma ; \bar{K} \sigma )
3097: \; ,
3098: \end{eqnarray}
3099: which is the analytic expression for the skeleton graph shown in
3100: Fig.~\ref{fig:skeletonsigmapi}(a). Similarly, we obtain the skeleton
3101: expansion of the irreducible polarization by differentiating
3102: Eq.~(\ref{eq:DysonSchwinger4}) with respect to $\varphi_{ - \bar{K}
3103: \sigma}$,
3104: \begin{eqnarray}
3105: \Pi_{\sigma} ( \bar{K} ) & = &
3106: i
3107: \int_{ {K}}
3108: \left. \frac{ \delta^{(3)} \G_c}{ \delta \varphi_{ -\bar{K} \sigma}
3109: \delta \bar{\jmath}_{ K , \sigma} \delta j_{ K + \bar{K} \sigma}}
3110: \right|_{ \rm{fields} = 0}
3111: \nonumber
3112: \\
3113: & & \hspace{-16mm} = - i \zeta \int_K
3114: G_{\sigma} ( K ) G_{\sigma} ( K + \bar{K} )
3115: \Gamma^{(2,1)} ( K + \bar{K} \sigma ; K \sigma ; \bar{K} \sigma )
3116: \; ,
3117: \nonumber
3118: \\
3119: & &
3120: \end{eqnarray}
3121: which is shown diagrammatically in Fig.~\ref{fig:skeletonsigmapi} (b).
3122: Finally, applying the operator $\frac{\delta^{(2)}}{ \delta
3123: \bar{\psi}_{ K + \bar{K} \sigma} \delta \psi_{ K \sigma}}$ to
3124: Eq.~(\ref{eq:DysonSchwinger4}) and subsequently setting the fields
3125: equal to zero we obtain the skeleton expansion of the three-legged
3126: vertex shown in Fig.~\ref{fig:skeletonsigmapi}(c),
3127: \begin{eqnarray}
3128: \Gamma^{(2,1)} ( K + \bar{K} \sigma ; K \sigma ; \bar{K} \sigma )
3129: & = & i
3130: \nonumber
3131: \\
3132: & & \hspace{-45mm} - i \zeta \int_{ K^{\prime}}
3133: G_{\sigma} ( K^{\prime} ) G_{\sigma} ( K^{\prime} + \bar{K} )
3134: \nonumber
3135: \\
3136: & & \hspace{-37mm} \times
3137: \Gamma^{(4,0)} ( K + \bar{K} \sigma , K^{\prime} \sigma ;
3138: K^{\prime} + \bar{K} \sigma , K \sigma )
3139: \; .
3140: \label{eq:skeletongamma}
3141: \end{eqnarray}
3142: Skeleton expansions for higher-order vertices can be obtained
3143: analogously from the appropriate functional derivatives of
3144: Eqs.~(\ref{eq:DysonSchwinger4})--(\ref{eq:DysonSchwinger6}).
3145:
3146:
3147: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3148: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3149: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3150: \section{Ward identities}
3151: \label{sec:ward}
3152:
3153: In this appendix we give a self-contained derivation of the Ward
3154: identities in Eqs.~(\ref{eq:WI1})--(\ref{eq:WI2}) within the framework
3155: of our functional integral approach. Although the Ward identity
3156: (\ref{eq:WI1}) for the three-legged vertex is well
3157: known,\cite{Dzyaloshinskii74,Bohr81,Metzner98,Kopietz97} it seems that
3158: the higher-order Ward identities given in Eqs.~(\ref{eq:WIm}) and
3159: (\ref{eq:WI2}) cannot be found anywhere in the literature. Since we
3160: are interested in deriving infinitely many Ward identities involving
3161: the vertices $\Gamma^{(2,m)}$ with two fermion legs and an arbitrary
3162: number $m$ of boson legs, it is convenient to derive first a ``master
3163: Ward identity'' for the generating functional for the irreducible
3164: vertices, from which we can obtain all desired Ward identities for the
3165: vertices by taking appropriate functional derivatives.
3166:
3167: Consider the generating functional of the Green's function of our mixed
3168: Fermi-Bose theory defined in Eq.~(\ref{eq:Ggen}), which in explicit
3169: notation is given by
3170: \begin{equation}
3171: \G[\bar{\jmath}, j , J] = \frac{1}{\Z_0}
3172: \int D [ \bar{\psi} , \psi, \varphi ]
3173: e^{-S [ \bar{\psi} , \psi , \varphi ] +(\bar{\jmath} ,\psi) + ( \bar{\psi} , j )
3174: + ( J^*, \varphi )}
3175: \; .
3176: \label{eq:Ggenexplicit}
3177: \end{equation}
3178: If we rewrite the parts of the action involving the fermionic fields
3179: $\bar{\psi}$ and $\psi$ in real space and imaginary time, the
3180: Euclidean action reads [we use again the notation $X =
3181: (\tau,{\bf{r}})$ introduced in Sec.~\ref{subsec:exactTLM}] as
3182: \begin{eqnarray}
3183: S[ \bar{\psi} , \psi , \varphi ]
3184: &=& S_0[\bar{\psi},\psi] + S_0[\varphi] + S_1[\bar{\psi},\psi,\varphi]
3185: \nonumber\\
3186: S_0[\bar{\psi},\psi]
3187: &=& \sum_{\sigma} \int_X \bar{\psi}_{\sigma}(X)\partial_{\tau}\psi_{\sigma}(X)
3188: \nonumber\\
3189: &+&\sum_{\sigma}\int\!d\tau\int\!d^Dr\int\!d^Dr'\,\bar{\psi}_{\sigma}(\tau,\mathbf{r})
3190: \nonumber\\\
3191: &&~~~~~~~~~~\times
3192: \xi_{\sigma}(\mathbf{r}-\mathbf{r}')
3193: \psi_{\sigma}(\tau,\mathbf{r}')
3194: \;,\\
3195: S_1[\bar{\psi},\psi,\varphi]
3196: &=& i\sum_{\sigma}\int_X \bar{\psi}_{\sigma}(X)\psi_{\sigma}(X)\varphi_{\sigma}(X)\;,
3197: \label{eq:Slocal}
3198: \end{eqnarray}
3199: where we have defined the Fourier transform of the dispersion
3200: \begin{equation}
3201: \xi_{\sigma}(\mathbf{r})=\int\frac{d^Dk}{(2\pi)^D}\,\xi_{\mathbf{k}\sigma}\,e^{i\mathbf{k}\cdot\mathbf{r}}
3202: \;.
3203: \end{equation}
3204: Suppose now that we perform a local gauge transformation on the fermion
3205: fields, defining new fields $\psi^{\prime}$ and $\bar{\psi}^{\prime}$
3206: via
3207: \begin{equation}
3208: \psi_{\sigma} ( X ) = e^{ i \alpha_{\sigma} ( X ) }
3209: \psi^{\prime}_{\sigma} ( X )
3210: \; \; , \; \;
3211: \bar{\psi}_{\sigma} ( X ) = e^{ - i \alpha_{\sigma} ( X ) }
3212: \bar{\psi}^{\prime}_{\sigma} ( X )
3213: \; ,
3214: \label{eq:gaugepsi}
3215: \end{equation}
3216: where $\alpha_{\sigma} ( X )$ is an arbitrary real function. It is easy
3217: to show that, to linear order in $\alpha_{\sigma} ( X )$, the action
3218: (\ref{eq:Slocal}) transforms as follows:
3219: \begin{eqnarray}
3220: S [ e^{ - i \alpha} \bar{\psi}^{\prime} , e^{i \alpha} \psi^{\prime} ,
3221: \varphi ] &=&
3222: S [ \bar{\psi}^{\prime} , \psi^{\prime} , \varphi]
3223: \nonumber\\
3224: &&\hspace{-4cm}+\;i\sum_{\sigma}\int_X \bar{\psi}^{\prime}_{\sigma}(X)
3225: [\partial_{\tau}\alpha_{\sigma}(X)]\psi^{\prime}_{\sigma}(X)
3226: \nonumber\\
3227: &&
3228: \hspace{-4cm}-\;i\sum_{\sigma}\int\!d\tau\int\!d^Dr\int\!d^Dr'\,
3229: \bar{\psi}^{\prime}_{\sigma}(\tau,\mathbf{r})
3230: [\alpha_{\sigma}(\tau,\mathbf{r})-\alpha_{\sigma}(\tau,\mathbf{r}')]
3231: \nonumber\\
3232: &&
3233: \times
3234: \xi_{\sigma}(\mathbf{r}-\mathbf{r}')\psi^{\prime}_{\sigma}(\tau,\mathbf{r}')
3235: \; .
3236: \label{eq:Sinvariance}
3237: \end{eqnarray}
3238: Using this relation, we see that the invariance of the generating
3239: functional in Eq.~(\ref{eq:Ggenexplicit}) with respect to the change
3240: of integration variables defined by Eq.~(\ref{eq:gaugepsi}) implies, to
3241: linear order in $\alpha_{\sigma} ( X )$,
3242: \begin{eqnarray}
3243: 0 & = &
3244: \frac{1}{\Z_0}
3245: \int D [ \bar{\psi} , \psi, \varphi ]
3246: e^{-S [ \bar{\psi} , \psi , \varphi ] +(\bar{\jmath} ,\psi) + ( \bar{\psi} , j ) n+ ( J , \varphi )}
3247: \nonumber
3248: \\
3249: & \times & \Biggl\{
3250: -\;\sum_{\sigma}\int_X \bar{\psi}_{\sigma}(X)
3251: [\partial_{\tau}\alpha_{\sigma}(X)]\psi_{\sigma}(X)
3252: \nonumber\\
3253: &&
3254: ~~+\;\sum_{\sigma}\int\!d\tau\int\!d^Dr\int\!d^Dr'\,
3255: \bar{\psi}_{\sigma}(\tau,\mathbf{r})
3256: \nonumber\\
3257: &&
3258: ~~~~~~~\times
3259: [\alpha_{\sigma}(\tau,\mathbf{r})-\alpha_{\sigma}(\tau,\mathbf{r}')]
3260: \xi_{\sigma}(\mathbf{r}-\mathbf{r}')\psi_{\sigma}(\tau,\mathbf{r}')
3261: \nonumber\\
3262: &&
3263: ~~+\;(\bar{\jmath},\alpha\psi)-(\bar{\psi}\alpha,j)
3264: \;\;
3265: \Biggr\} \; .
3266: \label{eq:invarianceG}
3267: \end{eqnarray}
3268: Taking the functional derivative of this equation with respect to
3269: $\alpha_{\sigma}(X)$, this implies in Fourier space,
3270: \begin{eqnarray}
3271: 0 & = &
3272: \int_K \Biggl\{
3273: \left[ i \bar{\omega} - \xi_{\mathbf{k}+\bar{\mathbf{k}},\sigma} + \xi_{\mathbf{k}\sigma}
3274: \right]
3275: \frac{ \delta^{(2)} \G }{ \delta \bar{\jmath}_{K \sigma} \delta j_{ K + \bar{K} \sigma} }
3276: \nonumber
3277: \\
3278: & & +
3279: \bar{\jmath}_{ K + \bar{K} \sigma}
3280: \frac{ \delta \G }{ \delta \bar{\jmath}_{ K \sigma} }
3281: -
3282: j_{ K \sigma} \frac{ \delta \G }{ \delta j_{ K + \bar{K} \sigma}}
3283: \Biggr\}
3284: \label{eq:WIG2}
3285: \; .
3286: \end{eqnarray}
3287: Expressing this equation in terms of the generating functional $\G_c =
3288: \ln \G $ of the connected Green's functions and the generating
3289: functional and $\Gamma [ \bar{\psi} , \psi , \varphi ]$ of the
3290: irreducible vertices as defined in Eq.~(\ref{eq:Gammadef}), we obtain
3291: \begin{eqnarray}
3292: 0 & = &
3293: \int_K \Biggl\{
3294: \left[ i \bar{\omega} -
3295: \xi_{\mathbf{k}+\bar{\mathbf{k}},\sigma} + \xi_{\mathbf{k}\sigma}
3296: \right]
3297: \frac{ \delta^{(2)} \G_c }{
3298: \delta \bar{\jmath}_{K \sigma} \delta j_{ K + \bar{K} \sigma} }
3299: \nonumber
3300: \\
3301: & & + \psi_{ K \sigma}
3302: \frac{ \delta \Gamma }{ \delta {\psi}_{ K + \bar{K} \sigma} }
3303: - \bar{\psi}_{ K + \bar{K} \sigma}
3304: \frac{ \delta \Gamma }{ \delta \bar{\psi}_{ K \sigma} }
3305: \Biggr\}
3306: \label{eq:masterWI}
3307: \; .
3308: \end{eqnarray}
3309: Alternatively, using the Dyson-Schwinger
3310: equation~(\ref{eq:DysonSchwinger4}), we may rewrite this as
3311: \begin{eqnarray}
3312: 0 & = &
3313: i \bar{\omega} \left[
3314: \frac{ \delta \Gamma}{ \delta \varphi_{ \bar{K} \sigma} }
3315: - i \int_K \bar{\psi}_{ K + \bar{K} \sigma} \psi_{K \sigma} \right]
3316: \nonumber
3317: \\
3318: &- & i
3319: \int_K
3320: (\xi_{\mathbf{k}+\bar{\mathbf{k}},\sigma}-\xi_{\mathbf{k}\sigma})
3321: \frac{ \delta^{(2)} \G_c }{
3322: \delta \bar{\jmath}_{K \sigma} \delta j_{ K + \bar{K} \sigma} }
3323: \nonumber
3324: \\
3325: & + & i \int_K \Biggl[ \psi_{ K \sigma}
3326: \frac{ \delta \Gamma }{ \delta {\psi}_{ K + \bar{K} \sigma} }
3327: - \bar{\psi}_{ K + \bar{K} \sigma}
3328: \frac{ \delta \Gamma }{ \delta \bar{\psi}_{ K \sigma} }
3329: \Biggr]
3330: \label{eq:masterWI2}
3331: \; .
3332: \end{eqnarray}
3333: Equations.~(\ref{eq:masterWI}) and (\ref{eq:masterWI2}) are our ``master
3334: Ward identities'' from which we can now obtain Ward identities for the
3335: vertices by differentiation. For example, taking the derivative
3336: $\frac{\delta}{\delta \varphi_{ - \bar{K} \sigma}}$ of
3337: Eq.~(\ref{eq:masterWI2}) we obtain
3338: \begin{equation}
3339: i \bar{\omega} \Pi_{\sigma} ( \bar{K} ) -
3340: \Pi^c_{\sigma} ( \bar{K} )
3341: = 0
3342: \label{eq:continuity}
3343: \; ,
3344: \end{equation}
3345: where we have defined
3346: \begin{eqnarray}
3347: \Pi^c_{\sigma} ( \bar{K} ) & = &
3348: - i \zeta \int_K
3349: (\xi_{\mathbf{k}+\bar{\mathbf{k}},\sigma}-\xi_{\mathbf{k}\sigma})
3350: G_{\sigma} ( K ) G_{\sigma} ( K + \bar{K} )
3351: \nonumber
3352: \\
3353: & & \times
3354: \Gamma^{(2,1)} ( K + \bar{K} \sigma ; K \sigma ; \bar{K} \sigma )
3355: \; .
3356: \end{eqnarray}
3357: Equation~(\ref{eq:continuity}) is a relation between response functions,
3358: which follows more directly from the equation of continuity. If we
3359: are interested in vertices involving at least one fermionic momentum
3360: and if the momentum transferred by the interaction is small, our master
3361: Ward identities can be further simplified. Then all fermionic momenta
3362: lie close to a given point ${\bf{k}}_{ F , \sigma}$ on the Fermi
3363: surface so that Eqs.~(\ref{eq:masterWI}) and (\ref{eq:masterWI2})
3364: become simpler if we assume asymptotic velocity conservation. This
3365: means that we replace under the integral sign
3366: \begin{equation}
3367: \xi_{\mathbf{k}+\bar{\mathbf{k}},\sigma} - \xi_{\mathbf{k}\sigma}
3368: \rightarrow {\bf{v}}_{F, \sigma} \cdot \bar{\mathbf{k}}
3369: \; .
3370: \label{eq:linearization}
3371: \end{equation}
3372: This approximation amounts to the linearization of the energy
3373: dispersion relative to the point ${\bf{k}}_{ F , \sigma}$ on the Fermi
3374: surface. Using again Eq.~(\ref{eq:DysonSchwinger4}), our master Ward
3375: identity becomes
3376: \begin{eqnarray}
3377: 0 & = &
3378: ( i \bar{\omega} - {\bf{v}}_{F, \sigma} \cdot \bar{\bf{k}} ) \left[
3379: \frac{ \delta \Gamma}{ \delta \varphi_{ \bar{K} \sigma} }
3380: - i \int_K \bar{\psi}_{ K + \bar{K} \sigma} \psi_{K \sigma} \right]
3381: \nonumber
3382: \\
3383: & + & i \int_K \Biggl[ \psi_{ K \sigma}
3384: \frac{ \delta \Gamma }{ \delta {\psi}_{ K + \bar{K} \sigma} }
3385: - \bar{\psi}_{ K + \bar{K} \sigma}
3386: \frac{ \delta \Gamma }{ \delta \bar{\psi}_{ K \sigma} }
3387: \Biggr]
3388: \label{eq:masterWI3}
3389: \; .
3390: \end{eqnarray}
3391: Differentiating this simplified master Ward identity with respect to
3392: the fields using the relation (\ref{eq:Gamma2Sigma}) as well as
3393: \begin{eqnarray}
3394: \left .\frac{ \delta^{(3)} \Gamma}{
3395: \delta \varphi_{ \bar{K} \sigma } \delta \psi_{ K \sigma}
3396: \delta \bar{\psi}_{ K + \bar{K} \sigma } }
3397: \right|_{ \rm{fields} =0} & &
3398: \nonumber \\
3399: & & \hspace{-40mm} =
3400: \Gamma^{(2,1)} ( K + \bar{K}\sigma ; K \sigma ; \bar{K}\sigma )
3401: \; ,
3402: \end{eqnarray}
3403: \begin{eqnarray}
3404: \left .\frac{ \delta^{(4)} \Gamma}{
3405: \delta \varphi_{ \bar{K}_1 \sigma }
3406: \delta \varphi_{ \bar{K}_2 \sigma } \delta \psi_{ K \sigma}
3407: \delta \bar{\psi}_{ K + \bar{K}_1 + \bar{K}_2 \sigma } }
3408: \right|_{ \rm{fields} =0} & &
3409: \nonumber \\
3410: & & \hspace{-60mm} =
3411: \Gamma^{(2,2)} ( K + \bar{K}_1 + \bar{K}_2\sigma ; K \sigma;
3412: \bar{K}_1 \sigma , \bar{K}_2 \sigma)
3413: \; ,
3414: \end{eqnarray}
3415: and so on, we obtain the Ward identities for the irreducible vertices
3416: given in Eqs.~(\ref{eq:WI1}, \ref{eq:WIm}, \ref{eq:WI2}).
3417:
3418: Of course, other Ward identities, e.g., the Ward identity for
3419: $\Gamma^{(4,1)}$ discussed in Ref.~\onlinecite{Benfatto04}, can also
3420: be obtained from Eq.~(\ref{eq:masterWI3}). Note that if the
3421: approximation (\ref{eq:linearization}) is not made, the master Ward
3422: identity (\ref{eq:masterWI3}) should be replaced by the more general
3423: master Ward identity (\ref{eq:masterWI2}), so that the Ward identities
3424: ~(\ref{eq:WI1}, \ref{eq:WIm}, \ref{eq:WI2}) for the vertices acquire
3425: correction terms. The effect of these correction terms on the Ward
3426: identities for $\Gamma^{(2,1)}$ and $\Gamma^{(4,1)}$ has very recently
3427: been studied in a mathematically rigorous way by Benfatto and
3428: Mastropietro.\cite{Benfatto04}
3429:
3430: \end{appendix}
3431:
3432:
3433: \begin{thebibliography}{99}
3434:
3435: \bibitem{Wilson72}
3436: K. G. Wilson, Phys. Rev. Lett. {\bf{28}}, 548 (1972).
3437: %
3438: \bibitem{Wilson74}
3439: K. G. Wilson and J. G. Kogut, Phys. Reports {\bf{12C}}, 75 (1974).
3440: %
3441: \bibitem{Fisher98}
3442: For a recent review with historical remarks
3443: see M. E. Fisher, Rev. Mod. Phys. {\bf{70}}, 653 (1998).
3444: %
3445: \bibitem{Ma76}
3446: S. K. Ma, {\it{Modern Theory of Critical Phenomena}}
3447: (Benjamin/Cummings, Reading, MA, 1976).
3448: %
3449: \bibitem{Wegner73}
3450: F. J. Wegner and A. Houghton, Phys. Rev. A {\bf{8}}, 401 (1973).
3451: %
3452: \bibitem{DiCastro74}
3453: C. Di Castro, G. Jona-Lasinio, and L. Peliti, Ann. Phys. (New York) {\bf{87}}, 327 (1974).
3454: %
3455: \bibitem{Nicoll76}
3456: J. F. Nicoll, T. S. Chang, and H. E. Stanley, Phys. Lett.
3457: A {\bf{57}}, 7 (1976);
3458: J. F. Nicoll and T. S. Chang, {\it{ibid.}} {\bf{62}}, 287 (1977).
3459: %
3460: \bibitem{Chang92}
3461: T. S. Chang, D. D. Vvedensky, and J. F. Nicoll,
3462: Phys. Rep. {\bf{217}}, 279 (1992).
3463: %
3464: \bibitem{Hertz76}
3465: J. A. Hertz, Phys. Rev. B {\bf{14}}, 1165 (1976).
3466: %
3467: \bibitem{Negele88}
3468: J.W.~Negele and H.~Orland, {\it{ Quantum Many-Particle Systems}}
3469: (Addison-Wesley, Redwood City, CA, 1988).
3470: %
3471: \bibitem{Kirkpatrick96}
3472: T. R. Kirkpatrick and D. Belitz, Phys. Rev. B {\bf{53}}, 14364 (1996);
3473: D. Belitz, T. R. Kirkpatrick, R. Narayanan, and T. Vojta,
3474: Phys. Rev. Lett. {\bf{85}}, 4602 (2000);
3475: D. Belitz, T. R. Kirkpatrick, M. T. Mercaldo, and S. L. Sessions,
3476: Phys. Rev. B {\bf{63}}, 174427 (2001); {\it{ibid.}} {\bf{63}}, 174428 (2001);
3477: D. Belitz, T. R. Kirkpatrick, and J. Rollb\"{u}hler, Phys. Rev. Lett. {\bf{93}},
3478: 155701 (2004).
3479: %
3480: \bibitem{Rosch01}
3481: A. Rosch, Phys. Rev. B {\bf{64}}, 174407 (2001).
3482: %
3483: \bibitem{Senthil04}
3484: T.~Senthil, A.~Vishwanath, L.~Balents, S.~Sachdev, and M.~P.~A.~Fisher,
3485: Science {\bf 303}, 1490 (2004);
3486: T.~Senthil, L.~Balents, S.~Sachdev, A.~Vishwanath, and
3487: M.~P.~A.~Fisher, Phys. Rev. B {\bf{70}}, 144407 (2004).
3488: %
3489: \bibitem{Shankar94}
3490: R. Shankar, Rev. Mod. Phys. {\bf{66}}, 129 (1994).
3491: %
3492: \bibitem{Zanchi96}
3493: D. Zanchi and H. J. Schulz, Phys. Rev. B {\bf{54}}, 9509 (1996);
3494: Europhys. Lett. {\bf{44}}, 235 (1998);
3495: Phys. Rev. B {\bf{61}}, 13609 (2000);
3496: Europhys. Lett. {\bf{55}}, 376 (2001).
3497: %
3498: \bibitem{Halboth00}
3499: C. J. Halboth and W. Metzner, Phys. Rev. B {\bf{61}}, 7364 (2000);
3500: Phys. Rev. Lett. {\bf{85}}, 5162 (2000).
3501: %
3502: \bibitem{Honerkamp01}
3503: C. Honerkamp, M. Salmhofer, N. Furukawa, and T. M. Rice,
3504: Phys. Rev. B {\bf{63}}, 35109 (2001);
3505: C. Honerkamp, Euro. Phys. J. B {\bf{21}}, 81 (2001).
3506: %
3507: \bibitem{Honerkamp01b}
3508: C. Honerkamp and M. Salmhofer, Phys. Rev. B {\bf{64}}, 184516 (2001);
3509: Phys. Rev. Lett. {\bf{87}} 187004 (2001).
3510: %
3511: \bibitem{Salmhofer01}
3512: M. Salmhofer and C. Honerkamp, Prog. Theor. Physics
3513: {\bf{105}}, 1 (2001).
3514: %
3515: \bibitem{Kopietz01}
3516: P. Kopietz and T.Busche, Phys. Rev. B {\bf{64}}, 155101 (2001).
3517: %
3518: \bibitem{Busche02}
3519: T.\ Busche, L.\ Bartosch, and P.\ Kopietz,
3520: J. Phys.: Cond. Mat. {\bf{14}}, 8513 (2002).
3521: %
3522: \bibitem{Ledowski03}
3523: S.~Ledowski and P.~Kopietz,
3524: J. Phys.: Condens. Matter {\bf 15}, 4779 (2003).
3525: %
3526: \bibitem{Tsai01}
3527: S. W. Tsai and B. Marston, Can. J. Phys. {\bf{79}}, 1463 (2001).
3528: %
3529: \bibitem{Binz02}
3530: B. Binz, D. Baeriswyl, and B. Dou\c{c}ot, Eur. Phys. J. B {\bf 25}, 69 (2002);
3531: Ann. Phys. (Leipzig) {\bf{12}}, 704 (2004).
3532: %
3533: \bibitem{Meden02}
3534: V. Meden, W. Metzner, U. Schollw\"{o}ck, and K. Sch\"{o}nhammer,
3535: Phys. Rev. B {\bf{65}}, 045318 (2002); S.~Andergassen, T.~Enss, V.~Meden,
3536: W.~Metzner, U.~Schollw\"{o}ck, and K.~Sch\"{o}nhammer,
3537: Phys. Rev. B {\bf 70}, 075102 (2004).
3538: %
3539: \bibitem{Kampf03}
3540: A. P. Kampf and A. A. Katanin, Phys. Rev. B {\bf{67}}, 125104 (2003);
3541: A. A. Katanin and A. P. Kampf, Phys. Rev. B {\bf{68}}, 195101 (2003);
3542: Phys. Rev. Lett. {\bf{93}}, 106406 (2004); cond-mat/0408246.
3543: %
3544: \bibitem{Katanin04}
3545: A. A. Katanin, Phys. Rev. Lett. {\bf{70}}, 115109 (2004).
3546: %
3547: \bibitem{Wetterich93}
3548: C. Wetterich, Phys. Lett. B {\bf{301}}, 90 (1993).
3549: %
3550: \bibitem{Morris94}
3551: T. R. Morris, Int. J. Mod. Phys. A {\bf{9}}, 2411 (1994).
3552: %
3553: \bibitem{Ferraz03}
3554: A. Ferraz, Phys. Rev. B {\bf{68}}, 075115 (2003);
3555: H. Freire, E. Correa, and A. Ferraz, Phys. Rev. B {\bf 71}, 165113 (2005).
3556: %
3557: \bibitem{Ferraz03b}
3558: A. Ferraz, Europhys. Lett. {\bf{61}}, 228 (2003).
3559: %
3560: \bibitem{Pomeranchuk58}
3561: I. Ia. Pomeranchuk,
3562: Zh. Eksp. Teor. Fiz. {\bf{35}}, 524 (1958)
3563: [Sov. Phys. JETP {\bf{8}}, 361 (1958)].
3564: %
3565: \bibitem{Metzner04}
3566: W. Metzner, D. Rohe, and S. Andergassen,
3567: Phys. Rev. Lett. {\bf{91}}, 066402 (2003).
3568: %
3569: \bibitem{Nozieres64}
3570: P.\ Nozi\`{e}res, {\it{Theory of Interacting Fermi Systems}}
3571: (Benjamin, New York, 1964).
3572: %
3573: \bibitem{Kohn60}
3574: W. Kohn and J. M. Luttinger, Phys. Rev. {\bf{118}}, 41 (1960);
3575: J. M. Luttinger, Phys. Rev. {\bf{119}}, 1153 (1960).
3576: %
3577: \bibitem{Pines89}
3578: D. Pines and P. Nozi\`{e}res, {\it{The Theory of Quantum Liquids}}, Volume I,
3579: (Addison-Wesley Advanced Book Classics, Redwood City, CA, 1989).
3580: %
3581: \bibitem{Correia01}
3582: S.~Correia, J.~Polonyi, and J.~Richert,
3583: %{\it Functional Callan-Symanzik Eqution for the Coulomb Gas},
3584: Ann. Phys. (New York) {\bf{296}}, 214 (2002).
3585: %
3586: \bibitem{Wetterich04}
3587: C. Wetterich, cond-mat/0208361v3.
3588: %
3589: \bibitem{Baier03}
3590: T. Baier, E. Bick, and C. Wetterich, Phys. Lett. B {\bf 605}, 144 (2005)
3591: and Phys. Rev. B {\bf 70}, 125111 (2004).
3592: %cond-mat/0309715.
3593: %
3594: \bibitem{Giamarchi04}
3595: For an up to date review on one-dimensional Fermi systems see the recent
3596: book by T. Giamarchi, {\it{Quantum Physics in One Dimension}}
3597: (Clarendon Press, Oxford, 2004).
3598: %
3599: \bibitem{Haldane81}
3600: F. D. M. Haldane, J. Phys. C: Solid State Phys. {\bf{14}}, 2585 (1981).
3601: %
3602: \bibitem{Stone94}
3603: M. Stone, {\it{Bosonization}} (World Scientific, Singapore, 1994).
3604: %
3605: \bibitem{Dzyaloshinskii74}
3606: I. E. Dzyaloshinskii and A. I. Larkin, Zh. Eksp. Teor. Fiz. {\bf{65}}, 411 (1973)
3607: [Sov. Phys. JETP {\bf{38}}, 202 (1974)].
3608: %
3609: \bibitem{Bohr81}
3610: T. Bohr, Nordita preprint 81/4, {\it{Lectures on the Luttinger Model}},
3611: 1981 (unpublished).
3612: %
3613: \bibitem{Metzner98}
3614: W. Metzner, C. Castellani, and C. Di Castro, Adv. Phys. {\bf{47}}, 317 (1998).
3615: %
3616: \bibitem{DiCastro91}
3617: C. Di Castro and W. Metzner, Phys. Rev. Lett. {\bf{67}}, 3852 (1991);
3618: W. Metzner and C. Di Castro, Phys. Rev. B {\bf{47}}, 16107 (1992).
3619:
3620: \bibitem{Kopietz97}
3621: P. Kopietz, {\it{Bosonization of Interacting Fermions in Arbitrary Dimensions}}
3622: (Springer, Berlin, 1997).
3623: %
3624: \bibitem{Benfatto04}
3625: G. Benfatto and V. Mastropietro, cond-mat/0409049.
3626: %
3627: \bibitem{Kopietz95}
3628: P. Kopietz, J. Hermisson, and K. Sch\"{o}nhammer, Phys. Rev. B {\bf{52}}, 10877 (1995);
3629: P. Kopietz and K. Sch\"{o}nhammer, Z. Phys. B {\bf{100}}, 259 (1996).
3630: %
3631: \bibitem{footnotezeta}
3632: The factor $\zeta $ arises from the antisymmetry of the Grassmann fields.
3633: Although throughout this work it is understood that $\zeta =-1$,
3634: all expressions involving the factor $\zeta$
3635: remain valid for bosonic fields $\psi$ and $\bar{\psi}$ if we set
3636: $\zeta =+1$. We adopt here the notation of Ref.~\onlinecite{Negele88}.
3637: %
3638: \bibitem{footnotefield}
3639: More precisely, we should distinguish the
3640: quantum field $\Phi_{\alpha}$ from its expectation value, i.e., we should
3641: write
3642: $\langle \Phi_{\alpha} \rangle = \delta \G_c / \delta J_{\alpha}$.
3643: For notational simplicity we redefine
3644: $\langle \Phi_{\alpha} \rangle \rightarrow \Phi_\alpha$.
3645: Whenever we work with the Legendre effective action it is understood
3646: that $\Phi_{\alpha}$ denotes the expectation value of the corresponding quantum field.
3647: %
3648: \bibitem{Solyom79}
3649: J. Solyom, Adv. Phys. {\bf{28}}, 201 (1979).
3650: %
3651: \bibitem{Honerkamp04}
3652: C. Honerkamp, D. Rohe, S. Andergassen, and T. Enss,
3653: Phys. Rev. B {\bf 70}, 235115 (2004).
3654: %cond-mat/0403633.
3655: %
3656: \bibitem{Schoenhammer03}
3657: K. Sch\"{o}nhammer, in {\it{Strong Interactions in Low Dimensions}},
3658: edited by D. Baeriswyl and L. Degiorgi (Dordrecht, Kluwer, 2003);
3659: cond-mat/9710330.
3660: %
3661: \bibitem{zfootnote} The factor of $\Omega_{\Lambda}^{-1}$ on the
3662: right-hand sides of Eqs.~(\ref{eq:Gammarescaledef}) and
3663: (\ref{eq:inhrescale1}) arises from the elimination of a fermionic
3664: frequency using the rescaled version of the energy-momentum
3665: conserving $\delta$ function
3666: $\delta_{K'_1+\ldots+K'_n,K_1+\ldots+K_n+\bar{K}_1+\ldots+\bar{K}_m}$.
3667: Such a procedure is only meaningful if $z_{\psi}\leq z_{\varphi}$,
3668: so that bosonic frequencies are not more relevant than fermionic
3669: ones. For $z_{\varphi}<z_{\psi}$ we should use the
3670: $\delta$ function to eliminate a bosonic frequency, which amounts
3671: to the replacement
3672: $\Omega_{\Lambda}^{-1}\rightarrow\bar{\Omega}_{\Lambda}^{-1}$ on
3673: the right-hand side of Eqs.~(\ref{eq:Gammarescaledef}) and
3674: (\ref{eq:inhrescale1}). This is the reason for the factor
3675: $z_{\text{min}}=\text{min}\{z_{\varphi},z_{\psi}\}$ in Eqs.~(\ref{eq:flowGammarescale1})
3676: and (\ref{eq:scaledim}).
3677: %
3678: \bibitem{Hedin65}
3679: L. Hedin, Phys. Rev. {\bf{139}}A 796 (1965);
3680: L. Hedin and S. Lundquist, {\it{Effects of Electron-Electron and Electron-Phonon
3681: Interactions on the One-Electron States of Solids}},
3682: in {\it{Solid State Physics}}, Vol. 23, edited by F. Seitz, D. Turnbull, and
3683: H. Ehrenreich (Academic Press, New York, 1969).
3684: %
3685: \bibitem{Haldane92}
3686: F.~D.~M.~Haldane, Helv. Phys. Acta {\bf 65}, 152
3687: (1992); {\it Luttinger's Theorem and Bosonization of the Fermi
3688: surface}, in {\it Proceedings of the International School of
3689: Physics ``Enrico Fermi''}, Course 121, 1992, edited by
3690: R.~Schrieffer and R.~A.~Broglia (North Holland, New York, 1994).
3691: %
3692: \bibitem{Bartosch99}
3693: L.~Bartosch and P.~Kopietz,
3694: Phys. Rev. B {\bf 59}, 5377 (1999).
3695: %
3696: \bibitem{Fogedby76}
3697: H. C. Fogedby, J. Phys. C {\bf{9}}, 3757 (1976).
3698: %
3699: \bibitem{Lee88}
3700: D. K. K. Lee and Y. Chen, J. Phys. A {\bf{21}}, 4155 (1988).
3701: %
3702: \bibitem{Ledowski04}
3703: S. Ledowski, N. Hasselmann, and P. Kopietz,
3704: Phys. Rev. A {\bf{69}}, 061601(R) (2004);
3705: N. Hasselmann, S. Ledowski, and P. Kopietz,
3706: Phys. Rev. A {\bf 70}, 63621 (2004).
3707: %cond-mat/0409167.
3708: %
3709: \bibitem{Ueda84}
3710: For $D > 1$ the coupling $\tilde{\gamma}_l$
3711: becomes irrelevant with scaling dimension $-(D - 1)/2$.
3712: If we formally treat $D-1$ as a small parameter,
3713: it might be possible to
3714: study the stability of the Luttinger liquid fixed point
3715: by means of an expansion in powers of
3716: $\epsilon = D-1$. Previously, K. Ueda and T. M. Rice, Phys. Rev. B {\bf{29}}, R1514 (1984),
3717: attempted to construct such an expansion within the
3718: field theoretical RG method.
3719: %
3720: \bibitem{footnoteeta}
3721: In deriving Eqs.~(\ref{eq:etares}) and (\ref{eq:sigmatruncres}), we have
3722: set the fixed point value $\tilde{r}_{\ast}$ of the relevant
3723: coupling $\tilde{r}_l$ equal to zero, which amounts to a redefinition
3724: of the bare coupling constant.
3725: Otherwise, $\tilde{r}_{\ast}$ would explicitly appear in
3726: Eqs.~(\ref{eq:etares}) and (\ref{eq:sigmatruncres}).
3727: Setting $\tilde{r}_{\ast} =0$ seems to correspond to the
3728: usual normal ordering with respect to the filled Dirac sea
3729: in the solution of the TLM via bosonization.
3730: In general, the
3731: functional dependence of $\eta$ on the interaction
3732: beyond the leading term in the weak coupling expansion
3733: depends on the
3734: regularization procedure; see H. J. Schulz and
3735: B.~S.~Shastry, Phys. Rev. Lett. {\bf{80}}, 1924 (1998).
3736: %
3737: \bibitem{footnoteVolker} V. Meden pointed out to us that this perfect
3738: agreement of Eq.~(\ref{eq:etares}) with bosonization might no longer
3739: occur in the more general TLM with $g_2\neq g_4$.
3740: %
3741: \bibitem{Busche01}
3742: T. Busche and P. Kopietz, Int. J. Mod. Phys. B {\bf{14}}, 1481 (2000).
3743: %
3744: \bibitem{Altshuler94}
3745: B. L. Altshuler, L. B. Ioffe, and A. J. Millis, Phys. Rev. B {\bf{50}}, 14048 (1994).
3746: %
3747: \bibitem{Belitz97}
3748: D. Belitz, T. R. Kirkpatrick, and T. Vojta, Phys. Rev. B {\bf{55}}, 9452 (1997).
3749: %
3750: \bibitem{Chubukov03}
3751: A. V. Chubukov and D. L. Maslov, Phys. Rev. B {\bf{68}}, 155113 (2003);
3752: A. V. Chubukov, C. Pepin, and J. Rech, Phys. Rev. Lett. {\bf{92}}, 147003 (2004).
3753: %
3754: \bibitem{Katanin04b}
3755: A. A. Katanin, A. P. Kampf, and V. Yu. Irkhin,
3756: Phys. Rev. B {\bf 71}, 085105 (2005).
3757: %cond-mat/0407473.
3758: %
3759: \bibitem{Bartosch03}
3760: L. Bartosch, M. Kollar, and P. Kopietz, Phys. Rev. B {\bf{67}}, 092403 (2003).
3761: %
3762: \bibitem{Yang04}
3763: K. Yang, Phys. Rev. Lett. {\bf{93}}, 066401 (2004).
3764: %
3765: \bibitem{ZinnJustin98}
3766: J.~ Zinn-Justin,
3767: {\it Quantum Field Theory and Critical Phenomena},
3768: (Clarendon Press, Oxford, 4th edition, 2002), Chap. 7.3.
3769:
3770: \end{thebibliography}
3771:
3772: \end{document}
3773:
3774:
3775:
3776: