0707.3840/GF.tex
1: \documentclass[12pt,twoside]{article}
2: 
3: \usepackage{amssymb}
4: \usepackage{amsmath}
5: \usepackage{wrapfig}
6: \usepackage{graphicx}
7: \usepackage[eqsecnum]{cmp2e}
8: \usepackage{citesort}
9: 
10: \newcommand{\Tr}{\operatorname{Tr}}
11: 
12: \hyphenation{Post-Script}
13: 
14: \title[GF Formalism for Highly Correlated Systems]%
15: {Green's Function Formalism for \protect\\
16: Highly Correlated Systems}
17: 
18: \author{F. Mancini and A. Avella}
19: \address{Dipartimento di Fisica ``E.R. Caianiello'' - Unit\`{a}
20: CNISM di Salerno \\
21: Universit\`{a} degli Studi di Salerno, I-84081 Baronissi (SA),
22: Italy}
23: 
24: \begin{document}
25: 
26: \maketitle
27: 
28: \begin{abstract}
29: We present the Composite Operator Method (COM) as a modern approach
30: to the study of strongly correlated electronic systems, based on the
31: equation of motion and Green's function method. COM uses propagators
32: of composite operators as building blocks at the basis of
33: approximate calculations and algebra constrains to fix the
34: representation of Green's functions in order to maintain the
35: algebraic and symmetry properties.
36: %
37: \keywords Strongly Correlated Systems, Green's Function Formalism,
38: Equations of Motion Approach, Composite Operator Method
39: %
40: \pacs 71.10.-w; 71.27.+a; 71.10.Fd
41: \end{abstract}
42: 
43: \section{Introduction}
44: 
45: The Green's function method is a very convenient formalism in
46: condensed matter physics, and many progresses have been achieved in
47: the last fifty years. When applied to interacting systems, such an
48: approach is usually based on the hypothesis that the interaction
49: among the particles is weak and can be treated in the framework of
50: some perturbation schemes. In this line of thinking a consolidated
51: scheme has been constructed, mostly based on diagrammatic
52: expansions, Wick's theorem, Dyson equation, and so on. However, in
53: the last few decades new materials with unconventional properties
54: have been discovered. It is believed that the origin of such
55: anomalous behaviors is generally due to strong electronic
56: correlations in narrow conduction bands \cite{Anderson_87}. In this
57: line of thinking many analytical methods have been developed for the
58: study of strongly correlated electron systems \cite{Fulde_95}. The
59: main difficulties are connected with the absence of any obvious
60: small parameter in the strong coupling regime and with the
61: simultaneous presence of itinerant and atomic aspects. The concept
62: that breaks down is the existence of the electrons as particles with
63: some well-defined and intrinsic properties. The presence of
64: interaction modifies the properties of the particles: what are
65: observed are new particles with new peculiar properties entirely
66: determined by the dynamics and by the boundary conditions. These new
67: objects appear as the final result of the modifications imposed by
68: the interactions on the original particles and contain, by the very
69: beginning, the effects of correlations. The choice of new
70: fundamental particles, whose properties have to be self-consistently
71: determined by dynamics, symmetries and boundary conditions, becomes
72: relevant.
73: 
74: As a simple example, let us consider an atomic system described by
75: the Hamiltonian
76: \begin{equation}\label{1.1}
77: H=-\mu \sum_\sigma   \varphi _\sigma ^\dag \varphi _\sigma +V\varphi
78: _\uparrow ^\dag \varphi _\downarrow ^\dag \varphi _\downarrow
79: \varphi _\uparrow
80: \end{equation}
81: $\varphi _\sigma$ denotes an Heisenberg electronic field with spin
82: $\sigma =\uparrow,\,\downarrow$, satisfying canonical
83: anticommutation relations; $\mu $ is the chemical potential and $V$
84: is the strength of the interaction. This model is exactly solvable
85: in terms of the operators
86: \begin{equation}\label{1.2}
87: \xi _\sigma =\varphi _\sigma \varphi _{-\sigma }\varphi _{-\sigma
88: }^\dag \quad \quad \quad \eta _\sigma =\varphi _\sigma \varphi
89: _{-\sigma }^\dag \varphi _{-\sigma }
90: \end{equation}
91: which are eigenoperators of the Hamiltonian
92: \begin{equation}\label{1.3}
93: \mathrm{i} {\partial \over {\partial t}}\xi =[\xi ,H]=-\mu \xi \quad
94: \quad \quad \mathrm{i}{\partial  \over {\partial t}}\eta =[\eta
95: ,H]=-(\mu -V)\eta
96: \end{equation}
97: 
98: Due to the presence of the interaction, the original electrons
99: $\varphi _\sigma$ are no more observables and new stable elementary
100: excitations, described by the field operators $\xi $ and $\eta $,
101: appear. Due to the $V$-interaction, two sharp features develop in
102: the band structure: the energy level $E=-\mu$ of the bare electron
103: splits in the two levels $E_1=-\mu $ and $E_2=V-\mu $. The bare
104: electron reveals itself to be precisely the wrong place to start. A
105: perturbative solution will never give the band splitting.
106: 
107: On the basis of this evidence one can be induced to move the
108: attention from the original fields to the new fields generated by
109: the interaction. The operators describing these excitations, once
110: they have been found, can be written in terms of the original ones
111: and are known as composite operators.
112: 
113: The convenience of developing a formulation to treat composite
114: excitations as fundamental objects has been noticed for the
115: many-body problem of condensed matter physics since long time.
116: Recent years have seen remarkable developments in many-body theory
117: in the form of an assortment of techniques that may be termed
118: composite particle methods. The beginnings of these types of
119: techniques may be traced back to the work of Bogolubov
120: \cite{Bogoliubov_47} and later to that of Dancoff \cite{Dancoff_50}.
121: The work of Zwanzig \cite{Zwanzig_61}, Mori \cite{Mori_65} and
122: Umezawa \cite{Umezawa_93} has to be mentioned. Closely related to
123: this work is that of Hubbard
124: \cite{Hubbard_63,Hubbard_64,Hubbard_64a}, Rowe \cite{Rowe_68}, Roth
125: \cite{Roth_69} and Tserkovnikov
126: \cite{Tserkovnikov_81,Tserkovnikov_81a}. The slave boson method
127: \cite{Barnes_76,Coleman_84,Kotliar_86}, the spectral density
128: approach \cite{Kalashnikov_69,Nolting_72} and the composite operator
129: method (COM)
130: \cite{Ishihara_94,Mancini_94,Mancini_95,Mancini_95a,Mancini_95b,Avella_98,Mancini_00,Mancini_04}
131: are also along similar lines. This large class of theories is
132: founded on the conviction that an analysis in terms of elementary
133: fields might be inadequate for a system dominated by strong
134: interactions.
135: 
136: All these approaches are very promising because all the different
137: approximation schemes are constructed on the basis of interacting
138: particles: some amount of the interaction is already present in the
139: chosen basis and permits to overcome the problem of finding an
140: appropriate expansion parameter. However, one price must be paid. In
141: general, the composite fields are neither Fermi nor Bose operators,
142: since they do not satisfy canonical (anti)commutation relations, and
143: their properties, because of the inherent definition, must be
144: self-consistently determined. They can only be recognized as
145: fermionic or bosonic operators according to the number, odd or even,
146: of the constituting original electronic fields. New techniques of
147: calculus have to be developed in order to treat with composite
148: fields. In developing perturbation calculations, where the building
149: blocks are now the propagators of composite fields, the consolidated
150: scheme (diagrammatic expansions, Wick's theorem, \ldots) gives rise
151: to very complicated approaches \cite{Izyumov_90} whose application
152: is far from being easy. The formulation of the Green's function
153: method must be revisited and new frameworks of calculations have to
154: be formulated.
155: 
156: \section{Green's Function and Equation of Motion Formalism}
157: 
158: Let us consider a system of $N_e$ interacting Wannier-electrons
159: residing on a Bravais lattice of $N$ sites, spanned by the vectors
160: $R_\mathbf{i}=\mathbf{i}$. We ignore the presence of magnetic
161: impurities and restrict the analysis to single-band electron models.
162: The generalization of the formalism to more complex systems is
163: straightforward [see for example
164: \cite{Matsumoto_94,Mancini_00e,Villani_00,Avella_02,Bak_02}]. The
165: system is enclosed in a finite but macroscopically large volume V
166: and is supposed to be in a thermodynamical equilibrium state at a
167: finite temperature T. In a second quantization scheme this system is
168: described by a certain Hamiltonian
169: \begin{equation}\label{2.1}
170: H=H[\varphi (i)]
171: \end{equation}
172: describing, in complete generality, the free propagation of the
173: electrons and all the interactions among them and with external
174: fields. $\varphi (i)$ denotes an Heisenberg electronic field [
175: $i=(\mathbf{i},t)$] satisfying canonical anticommutation relations
176: derived from the Pauli principle
177: \begin{equation}\label{2.2}
178: \{ \varphi _\sigma (\mathbf{i},t),\varphi _{\sigma }^\dag
179: (\mathbf{j},t)\} =\delta _{\mathbf{ij}}\delta _{\sigma \sigma }\quad
180: \quad \quad \{ \varphi _\sigma (\mathbf{i},t),\varphi _{\sigma
181: }(\mathbf{j},t)\} =0
182: \end{equation}
183: 
184: Any physical property of the system can be connected to the
185: expectation value of a specific operator. The expectation value of
186: an arbitrary operator $A=A[\varphi(i)]$ can be computed, for the
187: grand canonical ensemble, by means of
188: \begin{equation}\label{2.3}
189: \langle A \rangle = {{Tr[e^{-\beta (H-\mu \hat N)}A]} \over
190: {Tr[e^{-\beta (H-\mu \hat N)}]}}
191: \end{equation}
192: where the trace implies a sum over a complete set of states in the
193: Hilbert space. $\hat N=\sum_{i,\sigma }  \varphi _\sigma ^\dag
194: (i)\varphi _\sigma (i)$ is the total number operator, $\beta
195: =(k_BT)^{-1}$ is the inverse temperature, $\mu $ is the chemical
196: potential which is fixed in order to give the chosen average number
197: of particles $N_e= \langle \hat N \rangle$.
198: 
199: To evaluate the expectation value $\langle A\rangle $, it is
200: possible to use the equation of motion
201: \begin{equation}\label{2.4}
202: \mathrm{i}{\partial  \over {\partial t}}\varphi (i)=[\varphi (i),H]
203: \end{equation}
204: in order to derive one or more equations for this quantity or,
205: better, for the corresponding Green's functions, as explained below.
206: However, the equation of motion (\ref{2.4}) generates higher-order
207: operators and more and more complex equations are needed. The
208: traditional approximation schemes, often based on perturbative
209: calculations, use as building blocks the noninteracting propagators.
210: The mean-field formulation, which corresponds to a linearization of
211: the equation of motion (\ref{2.4}), also belongs to this category.
212: 
213: On the hypothesis that the original fields are not a good basis, we
214: choose a set of composite fields $\{ \psi (i)\}$ in terms of which a
215: perturbation scheme will be constructed. Firstly, we choose the set
216: $\psi (i)$ according to the physical properties we want to study.
217: Roughly, the properties of electronic systems can be classified in
218: two large classes: single particle properties, described in terms of
219: fermionic propagators, and response functions, described in terms of
220: bosonic propagators. These two sectors, fermionic and bosonic, are
221: not independent but interplay each other, and a fully
222: self-consistent solution usually requires that both sectors are
223: simultaneously solved. Once the sector, fermionic or bosonic, has
224: been fixed, we have several criteria for the choice of the new
225: basis. In constructing the composite fields no recipe can be given
226: without thinking to its drawbacks, but many recipes can assure a
227: correct and controlled description of relevant aspects of the
228: dynamics. One can choose: the higher order fields emerging from the
229: equations of motion (i.e., the conservation of some spectral moments
230: is assured), the eigenoperators of some relevant interacting terms
231: (i.e., the relevant interactions are treated exactly), the
232: eigenoperators of the problem reduced to a small cluster, \ldots
233: 
234: Let $\psi (i)$ be a $n$-component field
235: \begin{equation}\label{2.5}
236: \psi (i)= \left( \begin{array}{c}
237: \psi _1(i) \\
238: \vdots     \\
239: \psi _n(i)
240: \end{array} \right)
241: \end{equation}
242: We do not specify the nature, fermionic or bosonic, of the set $\{
243: \psi (i)\}$. In the case of fermionic operators it is intended that
244: we use the spinorial representation
245: \begin{equation}\label{2.6}
246: \psi _m(i)=\left(\begin{array}{c}
247: \psi _{\uparrow m}(i) \\
248: \psi _{\downarrow m}(i)
249: \end{array} \right)
250: \quad \quad \quad \psi _m^\dag (i)=\left(\psi _{\uparrow m}^\dag
251: (i),\,\psi _{\downarrow m}^\dag (i)\right)
252: \end{equation}
253: 
254: The dynamics of these operators is governed by the given Hamiltonian
255: $H=H[\varphi (i)]$ and can be written as
256: \begin{equation}\label{2.7}
257: \mathrm{i}{\partial  \over {\partial t}}\psi (i)=[\psi (i),H]=J(i)
258: \end{equation}
259: 
260: It is always possible to decompose the source $J(i)$ under the form
261: \begin{equation}\label{2.8}
262: J(i)=\varepsilon (-i\nabla )\psi (i)+\delta J(i)
263: \end{equation}
264: where the linear term represents the projection of the source on the
265: basis $\{ \psi \}$ and is calculated by means of the equation
266: \begin{equation}\label{2.9}
267: \langle [\delta J(\mathbf{i},t),\psi ^\dag (\mathbf{j},t)]_\eta
268: \rangle =0
269: \end{equation}
270: Here $\eta =\pm 1$; usually, it is convenient to take $\eta =1$
271: ($\eta =-1$ ) for a  fermionic (bosonic) set $\psi (i)$ (i.e., for a
272: composite field constituted of an odd (even) number of original
273: fields) in order to exploit the canonical anticommutation relations
274: of $\{ \psi (i)\}$; but, in principle, both choices are possible.
275: Accordingly, we define
276: \begin{equation}\label{2.10}
277: \left[ A , B \right]_\eta = \left \{
278: \begin{array}{l}
279: \left\{ A , B \right\} =AB+BA \quad for \quad \eta =1 \\
280: \left[ A , B \right] = AB-BA \quad for \quad \eta =-1
281: \end{array} \right.
282: \end{equation}
283: $\langle \cdots \rangle $ denotes the quantum statistical average
284: over the grand canonical ensemble, according to Eq.~(\ref{2.3}).
285: Hereafter, unless otherwise specified, time and space translational
286: invariance will be considered. The action of the derivative operator
287: $\varepsilon (-\mathrm{i}\nabla )$ on $\psi (i)$ is defined in
288: momentum space
289: \begin{equation}\label{2.11}
290: \varepsilon (-\mathrm{i}\nabla )\psi (i)=\varepsilon
291: (-\mathrm{i}\nabla ){1 \over {\sqrt N}}\sum_\mathbf{k}
292:  e^{\mathrm{i} \mathbf{k} \cdot \mathbf{R_i}}\psi (\mathbf{k},t)={1
293: \over {\sqrt N}}\sum_\mathbf{k}  e^{\mathrm{i}\mathbf{k}\cdot
294: \mathbf{R_i}}\varepsilon (\mathbf{k})\psi (\mathbf{k},t)
295: \end{equation}
296: where $\mathbf{k}$ runs over the first Brillouin zone. The
297: constraint (\ref{2.9}) gives
298: \begin{equation}\label{2.12}
299: m(\mathbf{k})=\varepsilon (\mathbf{k})I(\mathbf{k})
300: \end{equation}
301: after defining the normalization matrix
302: \begin{equation}\label{2.13}
303: I(\mathbf{i},\mathbf{j})=\langle [\psi (\mathbf{i},t),\psi ^\dag
304: (\mathbf{j},t)]_\eta \rangle ={1 \over N}\sum_\mathbf{k}
305: e^{\mathrm{i}\mathbf{k}\cdot
306: (\mathbf{R_i}-\mathbf{R_j})}I(\mathbf{k})
307: \end{equation}
308: and the m-matrix
309: \begin{equation}\label{2.14}
310: m(\mathbf{i},\mathbf{j})=\langle [J(\mathbf{i},t),\psi ^\dag
311: (\mathbf{j},t)]_\eta \rangle ={1 \over N}\sum_\mathbf{k}
312: e^{\mathrm{i}\mathbf{k}\cdot
313: (\mathbf{R_i}-\mathbf{R_j})}m(\mathbf{k})
314: \end{equation}
315: 
316: Since the components of $\psi (i)$ contain composite operators, the
317: normalization matrix $I(\mathbf{k})$ is not the identity matrix and
318: defines the spectral content of the excitations In fact, the use of
319: composite operators has the advantage of describing crossover
320: phenomena as the phenomena in which the weight of some operator is
321: shifted to another one.
322: 
323: It is worth noting that the normalization matrix $I(\mathbf{k})$ and
324: the $m(\mathbf{k})$ matrix are the lowest order generalized spectral
325: moments  $M^{(p)}(\mathbf{k})$ which are defined as
326: \begin{equation}\label{2.15}
327: M^{(p)}(\mathbf{k})=F.T.\left\langle {\left[ {\left(
328: {\mathrm{i}\partial /\partial t} \right)^p\psi (\mathbf{i},t),\psi
329: ^\dag (\mathbf{j},t)} \right]_\eta } \right\rangle
330: \end{equation}
331: where $F.T.$ stands for the Fourier transform. The generalized
332: spectral moments $M^{(p)}(k)$ have the Hermiticity property
333: \begin{equation}\label{2.16}
334: M_{ab}^{(p)}(\mathbf{k})=\left( {M_{ba}^{(p)}(\mathbf{k})} \right)*
335: \end{equation}
336: since at the equilibrium $M^{(p)}(\mathbf{k})$ is time-independent.
337: 
338: Coming back to our original problem, the evaluation of the
339: expectation value $\langle A\rangle $, it is possible to use the
340: equation of motion $\mathrm{i}{\partial  \over {\partial t}}\psi
341: (i)=[\psi (i),H]$ in order to derive equations for $\langle A\rangle
342: $. However, the correlation functions satisfy homogeneous equations.
343: A convenient generalization of the concept of correlation functions
344: is furnished by the Green's functions (GF) which have some
345: advantages in the construction and solution of the equations
346: determining them. In particular, the two-time Green's functions
347: contain most of the relevant information on the properties of the
348: system: expectation values of the observables, excitation spectrum,
349: response to external perturbation, and so on. Different types of GF
350: can be defined; for statistical systems it is better to consider the
351: real time thermodynamic Green's functions where the averaging
352: process of the Heisenberg operators is performed over the grand
353: canonical ensemble.
354: 
355: By considering the two-time thermodynamic Green's functions
356: \cite{Bogoliubov_59,Zubarev_60,Zubarev_74}, let us define the causal
357: function
358: \begin{equation}\label{2.17a}
359: G^C(i,j)=\langle C[\psi (i)\psi ^\dag (j)]\rangle =\theta
360: (t_i-t_j)\langle \psi (i)\psi ^\dag (j)\rangle -\eta \theta
361: (t_j-t_i)\langle \psi ^\dag (j)\psi (i)\rangle
362: \end{equation}
363: the retarded and advanced functions
364: \begin{equation}\label{2.17b}
365: G^{R,A}(i,j)=\langle R,A[\psi (i)\psi ^\dag (j)]\rangle =\pm \theta
366: [\pm (t_i-t_j)]\langle [\psi (i),\psi ^\dag (j)]_\eta
367: \rangle
368: \end{equation}
369: 
370: By means of the Heisenberg equation (\ref{2.7}) and using the
371: decomposition (\ref{2.8}), the Green's function $G^Q(i,j)=\langle Q[
372: \psi (i) \psi ^\dag (j)]\rangle$, where $Q=C,\,R,\,A$, satisfies the
373: equation
374: \begin{equation}\label{2.18}
375: \Lambda (\partial _i)G^Q(i,j)\Lambda ^\dag
376: (\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\leftarrow$}}\over
377: \partial } _j)=\Lambda (\partial _i)G_0^Q(i,j)\Lambda ^\dag
378: (\mathord{\buildrel{\lower3pt\hbox{$\scriptscriptstyle\leftarrow$}}\over
379: \partial } _j)+\langle Q[\delta J(i)\delta J^\dag (j)]\rangle
380: \end{equation}
381: where the derivative operator $\Lambda (\partial _i)$ is defined as
382: \begin{equation}\label{2.19}
383: \Lambda (\partial _i)=\mathrm{i}{\partial  \over {\partial
384: t_i}}-\varepsilon (-\mathrm{i}\nabla _i)
385: \end{equation}
386: and the propagator $G_0^Q(i,j)$ is defined by the equation
387: \begin{equation}\label{2.20}
388: \Lambda (\partial _i)G_0^Q(i,j)=i\delta
389: (t_i-t_j)I(\mathbf{i},\mathbf{j})
390: \end{equation}
391: 
392: By introducing the Fourier transform
393: \begin{equation}\label{2.21}
394: G^Q(i,j)={1 \over N}\sum_\mathbf{k}{\mathrm{i} \over {(2\pi
395: )}}\int_{-\infty }^{+\infty }  d\omega \kern 1pt
396: e^{\mathrm{i}\mathbf{k}\cdot
397: (\mathbf{R_i}-\mathbf{R_j})-\mathrm{i}\omega
398: (t_i-t_j)}G^Q(\mathbf{k},\omega )
399: \end{equation}
400: equation (\ref{2.18}) in momentum space is written as
401: \begin{equation}\label{2.22}
402: G^Q(\mathbf{k},\omega )=G_0^Q(\mathbf{k},\omega
403: )+G_0^Q(\mathbf{k},\omega )\Sigma^{Q*}(\mathbf{k},\omega
404: )G_0^Q(\mathbf{k},\omega )
405: \end{equation}
406: where the self-energy $\Sigma ^{Q*}(\mathbf{k},\omega )$ has the
407: expression
408: \begin{equation}\label{2.23}
409: \Sigma ^{Q*}(\mathbf{k},\omega
410: )=I^{-1}(\mathbf{k})B^Q(\mathbf{k},\omega )I^{-1}(\mathbf{k})
411: \end{equation}
412: with
413: \begin{equation}\label{2.24}
414: B^Q(\mathbf{k},\omega )=F.T.\langle Q[\delta J(i)\delta J^\dag
415: (j)]\rangle
416: \end{equation}
417: 
418: Next, we introduce the irreducible self-energy $\Sigma
419: ^Q(\mathbf{k},\omega )$ by means of the definition
420: \begin{equation}\label{2.25}
421: \Sigma ^Q(\mathbf{k},\omega )G^Q(\mathbf{k},\omega
422: )=I(\mathbf{k})\Sigma ^{Q*}(\mathbf{k},\omega
423: )G_0^Q(\mathbf{k},\omega )
424: \end{equation}
425: Equation (\ref{2.22}) takes the form
426: \begin{equation}\label{2.26}
427: G^Q(\mathbf{k},\omega )=G_0^Q(\mathbf{k},\omega
428: )+G_0^Q(\mathbf{k},\omega )I^{-1}(\mathbf{k})\Sigma
429: ^Q(\mathbf{k},\omega )G^Q(\mathbf{k},\omega )
430: \end{equation}
431: and can be formally solved as
432: \begin{equation}\label{2.27}
433: G^Q(\mathbf{k},\omega )={1 \over {\omega -\varepsilon
434: (\mathbf{k})-\Sigma ^Q(\mathbf{k},\omega )}}I(\mathbf{k})
435: \end{equation}
436: 
437: The formal definition (\ref{2.23}) of self-energy must be
438: manipulated to avoid any flowing on tautology. By noting that
439: $G^{Q-1}(\mathbf{k},\omega )=I^{-1}(\mathbf{k})[\omega -\varepsilon
440: (\mathbf{k})-\Sigma ^Q(\mathbf{k},\omega )]$ we can express $\Sigma
441: ^Q(\mathbf{k},\omega )$ as
442: \begin{equation}\label{2.28}
443: \Sigma ^Q(\mathbf{k},\omega )=B_{irr}^Q(\mathbf{k},\omega
444: )I^{-1}(\mathbf{k})
445: \end{equation}
446: where $B_{irr}^Q(\mathbf{k},\omega )$ indicates the irreducible part
447: of the propagator $B^Q(\mathbf{k},\omega )$, given by
448: \begin{equation}\label{2.29}
449: B_{irr}^Q(\mathbf{k},\omega )={1 \over {B^{Q-1}(\mathbf{k},\omega
450: )+I^{-1}(\mathbf{k})G_0^Q(\mathbf{k},\omega )I^{-1}(\mathbf{k})}}
451: \end{equation}
452: 
453: We have constructed a generalized perturbative approach designed for
454: formulations using composite fields. Equation (\ref{2.26}) is a
455: Dyson-like equation and may represents the starting point for a
456: perturbative calculation in terms of the propagator
457: $G_0^Q(\mathbf{k},\omega )$. Contrarily to the usual perturbation
458: schemes, the calculation of the "free propagator"
459: $G_0^Q(\mathbf{k},\omega )$ is not an easy task and large part of
460: this article will be dedicated to this problem. Then, the attention
461: will be given to the calculation of the self-energy $\Sigma
462: ^Q(\mathbf{k},\omega )$, and some approximate methods will be
463: presented. It should be noted that the computation of the two
464: quantities $G_0^Q(\mathbf{k},\omega )$ and $\Sigma
465: ^Q(\mathbf{k},\omega )$ are intimately related. The total weight of
466: the self-energy corrections is bounded by the weight of the residual
467: source operator $\delta J(i)$. According to this, it can be made
468: smaller and smaller by increasing the components of the basis $\psi
469: (i)$ [e.g. by including higher-order composite operators appearing
470: in $\delta J(i)$]. The result of such a procedure will be the
471: inclusion in the energy matrix of part of the self-energy as an
472: expansion in terms of coupling constants multiplied by the weights
473: of the newly includes basis operators. In general, the enlargement
474: of the basis leads to a new self-energy with a smaller total weight.
475: However, it is necessary pointing out that this process can be quite
476: cumbersome and the inclusion of fully momentum and frequency
477: dependent self-energy corrections can be necessary to effectively
478: take into account low-energy and virtual processes. According to
479: this, one can chose a reasonable number of components for the basic
480: set and then use another approximation method to evaluate the
481: residual dynamical corrections.
482: 
483: \section{GF properties, spectral representation, zero-frequency
484: functions}
485: 
486: \subsection{Equations of motion}
487: 
488: In the previous Section we have constructed a generalized
489: perturbative approach based on a Dyson equation designed for
490: formulations using composite fields. Two quantities appear in the
491: Dyson equation (\ref{2.26}): the "free propagator"
492: $G_0^Q(\mathbf{k},\omega )$ and the self-energy $\Sigma
493: ^Q(\mathbf{k},\omega )$. By postponing to next Sections the problem
494: of computing the self-energy, in this Section we concentrate on the
495: calculation of the Green's functions $G_0^Q(\mathbf{k},\omega )$
496: which constitute the building blocks of the perturbation scheme we
497: are trying to formulate. For the sake of simplicity we will drop the
498: sub index $0$ in the definition of $G_0^Q(k,\omega )$.
499: 
500: One fundamental aspect in a Green's function formulation is the
501: choice of the representation. The knowledge of the Hamiltonian and
502: of the operatorial algebra is not sufficient to completely specify
503: the GF. The GF refer to a specific representation (i.e., to a
504: specific choice of the Hilbert space) and this information must be
505: supplied as a boundary condition to the equations of motion that
506: alone are not sufficient to completely determine the GF. As well
507: known, the same system can exist in different phases according to
508: the external conditions; the existence of infinite inequivalent
509: representations \cite{Umezawa_65,Umezawa_66,Leplae_74} where the
510: equations of motions can be realized, allows us to pick up, among
511: the many possible choices, the right Hilbert space appropriate to
512: the physical situation under study. The use of composite operators
513: leads to an enlargement of the Hilbert space by the inclusion of
514: some unphysical states. As a consequence of this, it is difficult to
515: satisfy a priori all the sum rules and, in general, the symmetry
516: properties enjoined by the system under study. In addition, since
517: the representation where the operators are realized has to be
518: dynamically determined, the method clearly requires a process of
519: self-consistency.
520: 
521: From this discussion it is clear that fixing the representation is
522: not an easy task and requires special attention. In the literature
523: the properties of the GF are usually determined by starting from the
524: knowledge of the representation. Owing to the difficulties above
525: discussed we cannot proceed in this way. Therefore, we will derive
526: the general properties of the GF on the basis of the two elements we
527: have: the dynamics, fixed by the choice of the Hamiltonian
528: (\ref{2.1}), and the algebra, fixed by the choice of the basic set
529: (\ref{2.5}). The problem of fixing the representation will be
530: considered in the next Sections.
531: 
532: Let $\psi (i)$ be a n-component field satisfying linear equations of
533: motion
534: \begin{equation}\label{3.1}
535: \mathrm{i}{\partial  \over {\partial t}}\psi
536: _m(\mathbf{i},t)=\sum_\mathbf{j} \sum_{l=1}^n \varepsilon
537: _{ml}(\mathbf{i},\mathbf{j})\psi _l(\mathbf{j},t)
538: \end{equation}
539: with the energy matrix $\varepsilon (\mathbf{i},\mathbf{j})$ defined
540: by (\ref{2.12})-(\ref{2.14}). If the fields $\psi (i)$ are
541: eigenoperators of the total Hamiltonian, the equations of motion
542: (\ref{3.1}) are exact. There are many non-trivial realistic systems
543: for which it is possible to obtain a complete set of eigenoperators
544: of the Hamiltonian (for instance see
545: \cite{Mancini_05,Mancini_05b,Mancini_05a,Avella_05}). If the fields
546: $\psi (i)$ are not eigenoperators of $H$, the equations are
547: approximated; they correspond to neglecting the residual source
548: operator $\delta J(i)$ in the full equation of motion [see
549: (\ref{2.7}) and (\ref{2.8})] and all the formalism is developed with
550: the intention of using the propagators of these fields as a basis to
551: set up a perturbative scheme of calculations on the ground of the
552: Dyson equation (\ref{2.26}) derived in the previous Section.
553: 
554: By means of the field equation (\ref{3.1}) the Fourier transforms of
555: the various Green's functions defined by (\ref{2.16}), (\ref{2.17a})
556: and (\ref{2.17b}) satisfy the following equation
557: \begin{equation}\label{3.2}
558: [\omega -\varepsilon (\mathbf{k})]G^{Q(\eta )}(\mathbf{k},\omega
559: )=I^{(\eta )}(\mathbf{k})
560: \end{equation}
561: where the dependence on the parameter $\eta $ has been explicitly
562: introduced. As mentioned in Section 2, the set $\psi(i)$ can be
563: fermionic or bosonic and the parameter $\eta$ generally takes the
564: value $\eta =1$ ($\eta =-1$) for a  fermionic (bosonic) set $\psi
565: (i)$. The three Green's functions $G^C$, $G^R$ and $G^A$ satisfy the
566: same equation of motion which alone is not sufficient and must be
567: supplemented by other equations. Indeed, the GF are determined by
568: solving a first order differential equation of motion, thereby the
569: GF are given only within an arbitrary constant of integration. The
570: retarded and advanced GF can be completely determined because the
571: factor $\theta [\pm (t_i-t_j)]$ provides the boundary condition:
572: $G^{R,A}(i,j)=0\quad for\;t_i=t_j\mp \delta $. The determination of
573: the causal GF is not so immediate. The most general solution of
574: equation (3.2) is
575: \begin{equation}\label{3.3}
576: G^{Q(\eta )}(\mathbf{k},\omega )=\sum_{l=1}^n  \left\{ {P\left(
577: {{{\sigma ^{(l,\eta )}(k)} \over {\omega -\omega _l(\mathbf{k})}}}
578: \right)-\mathrm{i}\pi \delta [\omega -\omega
579: _l(\mathbf{k})]g^{(l,\eta )Q}(\mathbf{k})} \right\}
580: \end{equation}
581: where $\omega _l(\mathbf{k})$ are the eigenvalues of the
582: $\varepsilon (\mathbf{k})$, $\sigma ^{(l)}(\mathbf{k})$ are defined
583: by
584: \begin{equation}\label{3.4}
585: \sigma _{ab}^{(l)}(\mathbf{k})=\sum_{c=1}^n  \Omega
586: _{al}(\mathbf{k})\Omega
587: _{lc}^{-1}(\mathbf{k})I_{cb}(\mathbf{k})\quad \quad \quad
588: a,b=1,\ldots,n
589: \end{equation}
590: $\Omega (\mathbf{k})$ is the $n\times n$ matrix, whose columns are
591: the eigenvectors of $\varepsilon (\mathbf{k})$. We note that while
592: the spectral density matrix $\sigma ^{(l)}(\mathbf{k})$ is
593: completely determined by the energy $\varepsilon (\mathbf{k})$ and
594: normalization $I^{(\eta )}(\mathbf{k})$ matrices, the matrix
595: $g^{(l,\eta )Q}(\mathbf{k})$ is not fixed by the equations of motion
596: and must be determined by means of the boundary conditions. $P$
597: represents the principal value.
598: 
599: By recalling the retarded and advanced nature of $G^{R,A(\eta
600: )}(i,j)$ it is immediate to see that
601: \begin{equation}\label{3.5}
602: g^{(l,\eta )R}(\mathbf{k})=-g^{(l,\eta )A}(\mathbf{k})=\sigma
603: ^{(l,\eta )}(\mathbf{k})
604: \end{equation}
605: Then, the retarded and advanced functions are completely determined
606: in terms of the matrices $\varepsilon (\mathbf{k})$ and $I^{(\eta
607: )}(\mathbf{k})$. As well known, as functions of $\omega$ the
608: $G^{R,A(\eta )}(\mathbf{k},\omega )$ are analytic in the upper and
609: lower half-planes, respectively.
610: 
611: The determination of $g^{(l,\eta )C}(k)$ requires more work. From
612: the definitions (\ref{2.16}), (\ref{2.17a}) and  (\ref{2.17b}) we
613: can derive the following exact relations
614: \begin{equation}\label{3.6a}
615: G^{R(\eta )}(i,j)+G^{A(\eta )}(i,j)=2G^{C(\eta )}(i,j)-\langle [\psi
616: (i),\psi ^\dag (j)]_{-\eta }\rangle
617: \end{equation}
618: \begin{equation}\label{3.6b}
619: G^{R(\eta )}(i,j)-G^{A(\eta )}(i,j)=\langle [\psi (i),\psi ^\dag
620: (j)]_\eta \rangle
621: \end{equation}
622: where there appear the two correlation functions
623: \begin{equation}\label{3.7}
624: C_{\psi \psi ^\dag }(i,j)=\langle \psi (i)\psi ^\dag (j)\rangle
625: \quad \quad \quad C_{\psi ^\dag \psi }(i,j)=\langle \psi ^\dag
626: (j)\psi (i)\rangle
627: \end{equation}
628: The definition of $C_{\psi ^\dag \psi }(i,j)$ must be interpreted at
629: level of matrix elements: \\ \noindent $C_{\psi ^\dag \psi
630: ;ba}(i,j)=\langle \psi _b^\dag (j)\psi _a(i)\rangle $; no scalar
631: product is intended, neither in the spin space. These functions are
632: defined for all real times and the equations of motion they obey
633: contain no inhomogeneous terms involving a delta function. Indeed,
634: by means of the equations of motion (\ref{3.1}) the Fourier
635: transform of these correlation functions satisfy the homogeneous
636: equations
637: \begin{equation}\label{3.8}
638: [\omega -\varepsilon (\mathbf{k})]C_{\psi \psi ^\dag
639: }(\mathbf{k},\omega )=0\quad \quad \quad [\omega -\varepsilon
640: (\mathbf{k})]C_{\psi ^\dag \psi }(\mathbf{k},\omega )=0
641: \end{equation}
642: These equations tell us that the Fourier transforms of the
643: correlation functions are zero unless the frequency $\omega $ is
644: equal to one of  the energy levels $\omega _l(\mathbf{k})$ of the
645: system. The solutions of (\ref{3.8}) have the general form
646: \begin{equation}\label{3.9a}
647: C_{\psi \psi ^\dag }(\mathbf{k},\omega )=\sum_{l=1}^n  \delta
648: [\omega -\omega _l(\mathbf{k})]c_{_{\psi \psi ^\dag
649: }}^{(l)}(\mathbf{k})
650: \end{equation}
651: \begin{equation}\label{3.9b}
652: C_{\psi ^\dag \psi }(\mathbf{k},\omega )=\sum_{l=1}^n  \delta
653: [\omega -\omega _l(\mathbf{k})]c_{_{\psi ^\dag \psi
654: }}^{(l)}(\mathbf{k})
655: \end{equation}
656: with the momentum-dependent Fourier components $c_{_{\psi \psi ^\dag
657: }}^{(l)}(\mathbf{k})$ and $c_{_{\psi ^\dag \psi
658: }}^{(l)}(\mathbf{k})$ to be determined.
659: 
660: We now recall the Kubo-Martin-Schwinger (KMS) relation
661: \begin{equation}\label{3.10}
662: \langle A(t)B(t)\rangle =\langle B(t)A(t+i\beta )\rangle
663: \end{equation}
664: where $A(t)$ and $B(t)$ are Heisenberg operators at time $t$. This
665: relation implies that the Fourier transforms $C_{\psi \psi ^\dag
666: }(\mathbf{k},\omega )$ and $C_{\psi ^\dag \psi }(\mathbf{k},\omega
667: )$ are related
668: \begin{equation}\label{3.11}
669: C_{\psi ^\dag \psi }(\mathbf{k},\omega )=e^{-\beta \omega }C_{\psi
670: \psi ^\dag }(\mathbf{k},\omega )
671: \end{equation}
672: and the $\eta-$commutator $\langle [\psi (i),\psi ^\dag (j)]_\eta
673: \rangle $ can be expressed in terms of the correlation function as
674: \begin{equation}\label{3.12}
675: \langle [\psi (i),\psi ^\dag (j)]_\eta \rangle ={1 \over
676: N}\sum_\mathbf{k} {1 \over {2\pi }}\int  \kern 1pt d\omega
677: \,e^{\mathrm{i}\mathbf{k}\cdot
678: (\mathbf{i}-\mathbf{j})-\mathrm{i}\omega (t_i-t_j)}[1+\eta e^{-\beta
679: \omega }]C_{\psi \psi ^\dag }(\mathbf{k},\omega )
680: \end{equation}
681: 
682: By putting (\ref{3.12}) into (\ref{3.6a}) and (\ref{3.6b})  and by
683: taking into account Eqs.~(\ref{3.3}), (\ref{3.5}) and (\ref{3.9a})
684: we obtain
685: \begin{equation}\label{3.13a}
686: \sum_{l=1}^n  \delta [\omega -\omega _l(\mathbf{k})]\left\{
687: {g^{(l,\eta )C}(\mathbf{k})-{1 \over {2\pi }}[1-\eta e^{-\beta
688: \omega }]c_{_{\psi \psi ^\dag }}^{(l)}(\mathbf{k})} \right\}=0
689: \end{equation}
690: \begin{equation}\label{3.13b}
691: \sum_{l=1}^n  \delta [\omega -\omega _l(\mathbf{k})]\left\{ {\sigma
692: ^{(l,\eta )}(\mathbf{k})-{1 \over {2\pi }}[1+\eta e^{-\beta \omega
693: }]c_{_{\psi \psi ^\dag }}^{(l)}(\mathbf{k})} \right\}=0
694: \end{equation}
695: 
696: The solution of Eqs.~(\ref{3.13a}) and (\ref{3.13b}) is remarkably
697: different according to the value of the parameter $\eta $ and we
698: shall treat separately the two cases.
699: 
700: \subsection{Fermionic fields}
701: 
702: For the case of fermionic fields it is convenient to choose $\eta
703: =1$. Then, the solution of Eqs.~(\ref{3.13a}) and (\ref{3.13b}) is
704: \begin{align}\label{3.14}
705: &c^{(l)}(\mathbf{k})=\pi \left[ {1+\tanh \left( {{{\beta \omega
706: _l(\mathbf{k})} \over 2}} \right)} \right]\sigma
707: ^{(l,+1)}(\mathbf{k})\\
708: &g^{(l,+1)C}(\mathbf{k})=\tanh \left( {{{\beta \omega
709: _l(\mathbf{k})} \over 2}} \right)\sigma ^{(l,+1)}(\mathbf{k})
710: \end{align}
711: 
712: By putting (\ref{3.5}) and (\ref{3.14}) into (\ref{3.3}),
713: (\ref{3.9a}) and (\ref{3.9b}) we get the following general
714: expressions for the Green's functions and correlation functions
715: \begin{equation}\label{3.15a}
716: G^{R,A(+1)}(\mathbf{k},\omega )=\sum_{l=1}^n  {{\sigma
717: ^{(l,+1)}(\mathbf{k})} \over {\omega -\omega _l(\mathbf{k})\pm
718: \mathrm{i}\delta }}
719: \end{equation}
720: \begin{equation}\label{3.15b}
721: G^{C(+1)}(\mathbf{k},\omega )=\sum_{l=1}^n  \sigma
722: ^{(l,+1)}(\mathbf{k})\left[ {{{1-f_\mathrm{F}[\omega
723: _l(\mathbf{k})]} \over {\omega -\omega
724: _l(\mathbf{k})+\mathrm{i}\delta }}+{{f_\mathrm{F}[\omega
725: _l(\mathbf{k})]} \over {\omega -\omega
726: _l(\mathbf{k})-\mathrm{i}\delta }}} \right]
727: \end{equation}
728: \begin{equation}\label{3.15c}
729: C_{\psi \psi ^\dag }(\mathbf{k},\omega )=\pi \sum_{l=1}^n  \delta
730: [\omega -\omega _l(\mathbf{k})]\left[ {1+\tanh \left( {{{\beta
731: \omega _l(\mathbf{k})} \over 2}} \right)} \right]\sigma
732: ^{(l,+1)}(\mathbf{k})
733: \end{equation}
734: \begin{equation}\label{3.15d}
735: C_{\psi ^\dag \psi }(\mathbf{k},\omega )=\pi \sum_{l=1}^n  \delta
736: [\omega -\omega _l(\mathbf{k})]\left[ {1-\tanh \left( {{{\beta
737: \omega _l(\mathbf{k})} \over 2}} \right)} \right]\sigma
738: ^{(l,+1)}(\mathbf{k})
739: \end{equation}
740: where $f_\mathrm{F}(\omega )$ is the Fermi distribution function:
741: $f_\mathrm{F}(\omega )={1 \over {e^{\beta \omega }+1}}$.
742: 
743: By recalling that all the fermionic energies are shifted by the
744: chemical potential, the locus in the $k$-space, defined by $\omega
745: _l(k)=0$, will define the Fermi surface. By looking at (3.15b) we
746: can see that the imaginary part of the causal GF vanishes on the
747: Fermi surface. In the fermionic case the right procedure of
748: calculation is to start from the retarded (advanced) GF, and then to
749: compute the other Green's functions and correlation functions by
750: means of the relations
751: \begin{equation}\label{3.17a}
752: \Re [G^{C(+1)}(\mathbf{k},\omega )]=\Re
753: [G^{R,A(+1)}(\mathbf{k},\omega )]
754: \end{equation}
755: \begin{equation}\label{3.17b}
756: \Im [G^{C(+1)}(\mathbf{k},\omega )]=\pm \tanh \left( {{{\beta \omega
757: } \over 2}} \right)\Im [G^{R,A(+1)}(\mathbf{k},\omega )]
758: \end{equation}
759: \begin{equation}\label{3.17c}
760: C_{\psi \psi ^\dag }(\mathbf{k},\omega )=\mp \left[ {1+\tanh \left(
761: {{{\beta \omega } \over 2}} \right)} \right]\Im
762: [G^{R,A(+1)}(\mathbf{k},\omega )]
763: \end{equation}
764: 
765: We note the dispersion relations
766: \begin{align}\label{3.18}
767: &\Re [G^{R,A(+1)}(\mathbf{k},\omega )]=\mp {1 \over \pi
768: }P\int_{-\infty }^{+\infty }  d\omega \;{1 \over {\omega -\omega
769: }}\Im [G^{R,A(+1)}(\mathbf{k},\omega )]\\
770: &\Re [G^{C(+1)}(\mathbf{k},\omega )]=-{1 \over \pi }P\int_{-\infty
771: }^{+\infty }  d\omega \;{1 \over {\omega -\omega }}\coth \left(
772: {{{\beta \omega } \over 2}} \right)\Im [G^{C(+1)}(\mathbf{k},\omega
773: )]
774: \end{align}
775: By introducing the spectral function
776: \begin{equation}\label{3.19}
777: \rho ^{(+1)}(\mathbf{k},\omega )=\sum_{l=1}^n  \delta [\omega
778: -\omega _l(\mathbf{k})]\sigma ^{(l,+1)}(\mathbf{k})=\mp {1 \over \pi
779: }\Im [G^{R,A(+1)}(\mathbf{k},\omega )]
780: \end{equation}
781: we can establish the spectral representation
782: \begin{equation}\label{3.20a}
783: G^{R,A(+1)}(\mathbf{k},\omega )=\int_{-\infty }^{+\infty }  d\omega'
784: \;{{\rho ^{(+1)}(\mathbf{k},\omega' )} \over {\omega -\omega' \pm
785: \mathrm{i}\delta}}
786: \end{equation}
787: \begin{equation}\label{3.21b}
788: G^{C(+1)}(\mathbf{k},\omega )=\int_{-\infty }^{+\infty }  d\omega'
789: \;\rho ^{(+1)}(\mathbf{k},\omega' )\left[ {{{1-f_\mathrm{F}(\omega'
790: )} \over {\omega -\omega' +\mathrm{i}\delta
791: }}+{{f_\mathrm{F}(\omega' )} \over {\omega -\omega'
792: -\mathrm{i}\delta }}} \right]
793: \end{equation}
794: 
795: \subsection{Bosonic fields}
796: 
797: For the case of bosonic fields it is convenient to choose $\eta
798: =-1$. For any given momentum $\mathbf{k}$ we can always write
799: \begin{equation}\label{3.22}
800: \omega _l(\mathbf{k})=\left\{
801: \begin{array}{l}
802: =0 \quad \text{for} \;l \in A(\mathbf{k})\subseteq N=\{ 1,\ldots,n\}\\
803: \ne 0 \quad \text{for} \; l \in B(\mathbf{k})=N-A(\mathbf{k})
804: \end{array}
805: \right.
806: \end{equation}
807: 
808: Obviously, $A(\mathbf{k})$ can also be the empty set (i.e.,
809: $A(\mathbf{k})=\emptyset $ and $B(\mathbf{k})=N$). For $l\in
810: B(\mathbf{k})$ the solution of (\ref{3.13a}) and (\ref{3.13b}) is
811: \begin{align}\label{3.23}
812: &c_{\psi \psi ^\dag}^{(l)}(\mathbf{k})=\pi \left[ {1+\coth \left(
813: {{{\beta \omega _l(\mathbf{k})} \over 2}} \right)} \right]\sigma
814: ^{(l,-1)}(\mathbf{k})\quad \quad \forall l\in
815: B(\mathbf{k})\\
816: &g^{(l,-1)C}(\mathbf{k})=\coth \left( {{{\beta \omega
817: _l(\mathbf{k})} \over 2}} \right)\sigma ^{(l,-1)}(\mathbf{k})\quad
818: \quad \quad \quad \forall l\in B(\mathbf{k})
819: \end{align}
820: For $l\in A(\mathbf{k})$ the Fourier coefficients $c_{_{\psi \psi
821: ^\dag }}^{(l)}(\mathbf{k})$ and  $g^{(l,-1)C}(\mathbf{k})$ cannot be
822: determined from Eqs.~(\ref{3.13b}). It is convenient to introduce
823: the function $\Gamma (\mathbf{k})$
824: \begin{equation}\label{3.24}
825: \Gamma (\mathbf{k})={1 \over {2\pi }}\sum_{l\in A(\mathbf{k})}
826: c_{_{\psi \psi ^\dag }}^{(l)}(\mathbf{k})={1 \over 2}\sum_{l\in
827: A(\mathbf{k})} g^{(l,-1)C}(\mathbf{k})
828: \end{equation}
829: By considering that from (\ref{3.13b}) and (\ref{3.9a})
830: \begin{equation}\label{3.25}
831: \mathop {\lim }_{\omega \to 0}[1-e^{-\beta \omega }]C_{\psi \psi
832: ^\dag }(\mathbf{k},\omega )=2\pi \delta (\omega )\sum_{l\in
833: A(\mathbf{k})} \sigma ^{(l,-1)}(\mathbf{k})
834: \end{equation}
835: we must distinguish two situations:
836: \begin{enumerate}
837:  \item If  $\sum_{l\in A(\mathbf{k})} \sigma ^{(l,-1)}(\mathbf{k})=0$, then
838: \begin{equation}\label{3.26a}
839: \sum_{l=1}^n  \delta [\omega -\omega _l(\mathbf{k})]c_{_{\psi \psi
840: ^\dag }}^{(l)}(\mathbf{k})=2\pi \delta (\omega )\Gamma
841: (\mathbf{k})+2\pi \sum_{l\in B(\mathbf{k})} \delta [\omega -\omega
842: _l(\mathbf{k})]{{e^{\beta \omega _l(\mathbf{k})}} \over {e^{\beta
843: \omega _l(\mathbf{k})}-1}}\sigma ^{(l,-1)}(\mathbf{k})
844: \end{equation}
845: \begin{equation}\label{3.26b}
846: \sum_{l=1}^n  \delta [\omega -\omega _l(\mathbf{k})]c_{_{\psi ^\dag
847: \psi }}^{(l)}(\mathbf{k})=2\pi \delta (\omega )\Gamma
848: (\mathbf{k})+2\pi \sum_{l\in B(\mathbf{k})} \delta [\omega -\omega
849: _l(\mathbf{k})]{1 \over {e^{\beta \omega _l(\mathbf{k})}-1}}\sigma
850: ^{(l,-1)}(\mathbf{k})
851: \end{equation}
852: and
853: \begin{equation}\label{3.27}
854: C_{\psi \psi ^\dag }(\mathbf{k},0)=C_{\psi ^\dag \psi
855: }(\mathbf{k},0)
856: \end{equation}
857:  \item If $\sum_{l\in A(\mathbf{k})} \sigma ^{(l,-1)}(\mathbf{k})\ne 0$ but finite,
858: then in order to satisfy (\ref{3.25}) $C_{\psi \psi ^\dag
859: }(\mathbf{k},\omega )$ must have a  singularity of the type ${1
860: \over \omega }$ in the limit $\omega \to 0$. In fact
861: \begin{equation}\label{3.28}
862: (1-e^{-\beta \omega }){1 \over {\beta \omega }}=(\beta \omega -{1
863: \over 2}\beta ^2\omega ^2+\ldots){1 \over {\beta \omega }}=1-{1
864: \over 2}\beta \omega +\ldots
865: \end{equation}
866: Then
867: \begin{equation}\label{3.29a}
868: \sum_{l=1}^n  \delta [\omega -\omega _l(\mathbf{k})]c_{_{\psi \psi
869: ^\dag }}^{(l)}(\mathbf{k})=2\pi \delta (\omega )\Gamma
870: (\mathbf{k})+2\pi \sum_{l=1}^n \delta [\omega -\omega
871: _l(\mathbf{k})]{{e^{\beta \omega _l(\mathbf{k})}} \over {e^{\beta
872: \omega _l(\mathbf{k})}-1}}\sigma ^{(l,-1)}(\mathbf{k})
873: \end{equation}
874: \begin{equation}\label{3.29b}
875: \sum_{l=1}^n  \delta [\omega -\omega _l(\mathbf{k})]c_{_{\psi ^\dag
876: \psi }}^{(l)}(\mathbf{k})=2\pi \delta (\omega )\Gamma
877: (\mathbf{k})+2\pi \sum_{l=1}^n \delta [\omega -\omega
878: _l(\mathbf{k})]{1 \over {e^{\beta \omega _l(\mathbf{k})}-1}}\sigma
879: ^{(l,-1)}(\mathbf{k})
880: \end{equation}
881: and
882: \begin{equation}\label{3.30}
883: C_{\psi \psi ^\dag }(\mathbf{k},0)-C_{\psi ^\dag \psi
884: }(\mathbf{k},0)=\sum_{l\in A(\mathbf{k})} \sigma
885: ^{(l,-1)}(\mathbf{k})
886: \end{equation}
887: \end{enumerate}
888: 
889: It is clear from (\ref{3.29a}) and (\ref{3.29b}) that the situation
890: where $\sum_{l\in A(\mathbf{k})} \sigma ^{(l,-1)}(\mathbf{k})\ne 0$
891: leads to a situation in which for $l\in A(\mathbf{k})$ the Fourier
892: coefficients $c_{_{\psi \psi ^\dag }}^{(l)}(\mathbf{k})$ and
893: $c_{_{\psi ^\dag \psi }}^{(l)}(\mathbf{k})$ diverge as $[\beta
894: \omega _l(\mathbf{k})]^{-1}$. Since the correlation function in
895: direct space must be finite, at finite temperature this is
896: admissible only in the thermodynamic limit and if the dispersion
897: relation $\omega _l(\mathbf{k})$ is such that the divergence in
898: momentum space is integrable and the corresponding correlation
899: function in real space remains finite. For finite systems and for
900: infinite systems where the divergence is not integrable we must have
901: $\sum_{l\in A(\mathbf{k})}\sigma ^{(l,-1)}(\mathbf{k})=0$. The
902: calculation of the spectral density matrices $\sigma
903: ^{(l,-1)}(\mathbf{k})$ it not a simple dynamical problem, but
904: requires the self-consistent calculation of some expectation values,
905: where the boundary condition and the choice of the representation
906: play a crucial role. A finite value of $\sum_{l\in A(\mathbf{k})}
907: \sigma ^{(l,-1)}(\mathbf{k})$ is generally related to the presence
908: of long-range order and the previous statement is nothing but the
909: Mermin-Wagner theorem \cite{Mermin_66}.
910: 
911: Summarizing, by using (\ref{3.5}) and by putting (\ref{3.26a}) and
912: (\ref{3.26b}) into (\ref{3.13a}), for finite systems and $T\ne 0$,
913: we get the following general expressions for the GF and correlations
914: function
915: \begin{equation}\label{3.31a}
916: G^{R,A(-1)}(\mathbf{k},\omega )=\sum_{l=1}^n {{\sigma
917: ^{(l,-1)}(\mathbf{k})} \over {\omega -\omega _l(\mathbf{k})\pm
918: \mathrm{i}\delta }}
919: \end{equation}
920: \begin{align}\label{3.31b}
921: &G^{C(-1)}(\mathbf{k},\omega )=\Gamma (\mathbf{k})\left[ {{1 \over
922: {\omega +\mathrm{i}\delta }}-{1 \over {\omega -\mathrm{i}\delta }}}
923: \right]\nonumber\\
924: &+\sum_{l\in B(\mathbf{k})} \sigma ^{(l,-1)}(\mathbf{k})\left[
925: {{{1+f_\mathrm{B}(\omega )} \over {\omega -\omega
926: _l(\mathbf{k})+\mathrm{i}\delta }}-{{f_\mathrm{B}(\omega )} \over
927: {\omega -\omega _l(\mathbf{k})-\mathrm{i}\delta }}} \right]
928: \end{align}
929: \begin{equation}\label{3.31c}
930: C_{\psi \psi ^\dag }(\mathbf{k},\omega )=2\pi \Gamma
931: (\mathbf{k})\delta (\omega )+\pi \sum_{l\in B(\mathbf{k})}  \delta
932: [\omega -\omega _l(\mathbf{k})]\left[ {1+\coth \left( {{{\beta
933: \omega _l(\mathbf{k})} \over 2}} \right)} \right]\sigma
934: ^{(l,-1)}(\mathbf{k})
935: \end{equation}
936: \begin{equation}\label{3.31d}
937: C_{\psi ^\dag \psi }(\mathbf{k},\omega )=2\pi \Gamma
938: (\mathbf{k})\delta (\omega )-\pi \sum_{l\in B(\mathbf{k})}  \delta
939: [\omega -\omega _l(\mathbf{k})]\left[ {1-\coth \left( {{{\beta
940: \omega _l(\mathbf{k})} \over 2}} \right)} \right]\sigma
941: ^{(l,-1)}(\mathbf{k})
942: \end{equation}
943: with the condition that $\sum_{l\in A(\mathbf{k})}  \sigma
944: ^{(l,-1)}(\mathbf{k})=0$. $f_\mathrm{B}(\omega )$ is the Bose
945: distribution function: $f_\mathrm{B}(\omega )={1 \over {e^{\beta
946: \omega }-1}}$. It is possible to have $\sigma
947: ^{(l,-1)}(\mathbf{k})\ne 0\quad for\quad l\in A(\mathbf{k})$ only
948: for infinite systems and if the divergence is integrable.
949: 
950: The previous formulas show that when zero-energy modes are present a
951: zero-frequency singularity appears in the correlation function
952: $C_{\psi \psi ^\dag }(\mathbf{k},\omega )$ and in the imaginary part
953: of the causal function $G^{C(-1)}(\mathbf{k},\omega )$. Such
954: singularity does not contribute to the retarded and advanced GF.
955: Then, in the bosonic case the right procedure of calculation is to
956: start from the causal GF, and compute the other GF by means of the
957: relations
958: \begin{equation}\label{3.34a}
959: \Re [G^{R,A(-1)}(\mathbf{k},\omega )]=\Re
960: [G^{C(-1)}(\mathbf{k},\omega )]
961: \end{equation}
962: \begin{equation}\label{3.34b}
963: \Im [G^{R,A(-1)}(\mathbf{k},\omega )]=\pm \tanh \left( {{{\beta
964: \omega } \over 2}} \right)\Im [G^{C(-1)}(\mathbf{k},\omega )]
965: \end{equation}
966: \begin{equation}\label{3.34c}
967: C_{\psi \psi ^\dag }(\mathbf{k},\omega )=-\left[ {1+\tanh \left(
968: {{{\beta \omega } \over 2}} \right)} \right]\Im
969: [G^{C(-1)}(\mathbf{k},\omega )]
970: \end{equation}
971: 
972: We note the dispersion relations
973: \begin{align}\label{3.35}
974: &\Re [G^{R,A(-1)}(\mathbf{k},\omega )]=\mp {1 \over \pi
975: }P\int_{-\infty }^{+\infty }  d\omega \;{1 \over {\omega -\omega
976: }}\Im [G^{R,A(-1)}(\mathbf{k},\omega )]\\
977: &\Re [G^{C(-1)}(\mathbf{k},\omega )]=-{1 \over \pi }P\int_{-\infty
978: }^{+\infty }  d\omega \;{1 \over {\omega -\omega }}\tanh \left(
979: {{{\beta \omega } \over 2}} \right)\Im [G^{C(-1)}(\mathbf{k},\omega
980: )]
981: \end{align}
982: 
983: Also in the bosonic case we can introduce a spectral function
984: \begin{equation}\label{3.36}
985: \rho ^{(-1)}(\mathbf{k},\omega )=\sum_{l=1}^n  \delta [\omega
986: -\omega _l(\mathbf{k})]\sigma ^{(l,-1)}(\mathbf{k})=\mp {1 \over \pi
987: }\Im [G^{R,A(-1)}(\mathbf{k},\omega )]
988: \end{equation}
989: 
990: However, the zero-frequency function $\Gamma (\mathbf{k})$ does not
991: contribute to $\rho ^{(-1)}(\mathbf{k},\omega )$ and a spectral
992: representation can be established only for the retarded (advanced)
993: GF
994: \begin{equation}\label{3.37}
995: G^{R,A(-1)}(\mathbf{k},\omega )=\int_{-\infty }^{+\infty }  d\omega'
996: \;{{\rho ^{(-1)}(\mathbf{k},\omega' )} \over {\omega -\omega' \pm
997: \mathrm{i}\delta }}
998: \end{equation}
999: 
1000: For the bosonic causal GF a spectral representation exists only when
1001: $\Gamma (\mathbf{k})=0$:
1002: \begin{equation}\label{3.38}
1003: G^{C(-1)}(\mathbf{k},\omega )=\int_{-\infty }^{+\infty } d\omega'
1004: \;\rho ^{(-1)}(\mathbf{k},\omega' )\left[ {{{1+f_\mathrm{B}(\omega'
1005: )} \over {\omega -\omega' +\mathrm{i}\delta
1006: }}-{{f_\mathrm{B}(\omega' )} \over {\omega -\omega'
1007: -\mathrm{i}\delta }}} \right]
1008: \end{equation}
1009: 
1010: \subsection{Sum rules and some useful relations}
1011: 
1012: Coming back to a generic value of $\Gamma (\mathbf{k})$, we note
1013: that from the definition (\ref{3.4}) the following sum rule can be
1014: derived
1015: \begin{equation}\label{3.39}
1016: \int_{-\infty }^{+\infty }  d\omega \kern 1pt \rho ^{(\eta
1017: )}(\mathbf{k},\omega )=\sum_{l=1}^n  \sigma ^{(l,\eta
1018: )}(\mathbf{k})=I^{(\eta )}(\mathbf{k})
1019: \end{equation}
1020: 
1021: This is a particular case of a general sum rule. From the results
1022: (\ref{3.15c}), (\ref{3.15d}), (\ref{3.31c}) and (\ref{3.31d}) we
1023: obtain
1024: \begin{equation}\label{3.40}
1025: \langle [\psi (i),\psi ^\dag (j)]_\eta \rangle =\sum_{l=1}^n  {1
1026: \over N}\sum_\mathbf{k}  \,e^{\mathrm{i}\mathbf{k}\cdot
1027: (\mathbf{i}-\mathbf{j})-\mathrm{i}\omega
1028: _l(\mathbf{k})(t_i-t_j)}\sigma ^{(l,\eta )}(\mathbf{k})
1029: \end{equation}
1030: Recalling the expression (\ref{2.15}) of the generalized spectral
1031: moment $M^{(p,\eta )}(\mathbf{k})$ we immediately have
1032: \begin{equation}\label{3.41}
1033: \int_{-\infty }^{+\infty }  d\omega \kern 1pt \omega ^p\rho ^{(\eta
1034: )}(\mathbf{k},\omega )=\sum_{l=1}^n  \,\omega _l(\mathbf{k})^p\sigma
1035: ^{(l,\eta )}(\mathbf{k})=M^{(p,\eta )}(\mathbf{k})
1036: \end{equation}
1037: 
1038: Some interesting results can be obtained by noting that the
1039: correlation function $C_{\psi \psi ^\dag }(i,j)=\langle \psi (i)\psi
1040: ^\dag (j)\rangle $ and the energy matrix $\varepsilon
1041: (\mathbf{i},\mathbf{j})$ do not depend on $\eta $. As mentioned
1042: above, once we have chosen a basic set $\{ \psi (i)\} $, fermionic
1043: or bosonic, it is a only a question of convenience to choose$\eta
1044: =1$ or $\eta =-1$. Let us consider the case of a bosonic set $\{
1045: \psi (i)\} $ and let us suppose to perform two series of
1046: calculations: one with $\eta =-1$ and another with $\eta =1$. Then,
1047: it is immediate to obtain the following relations:
1048: \begin{enumerate}
1049:  \item by equating (\ref{3.15c}) and (\ref{3.31c}) we obtain
1050: \begin{equation}\label{3.42a}
1051: \Gamma (\mathbf{k})={1 \over 2}\sum_{l\in A(\mathbf{k})}  \sigma
1052: ^{(l,+1)}(\mathbf{k})
1053: \end{equation}
1054: \begin{equation}\label{3.42b}
1055: \sigma ^{(l,-1)}(\mathbf{k})=\tanh \left( {{{\beta \omega
1056: _l(\mathbf{k})} \over 2}} \right)\sigma ^{(l,+1)}(\mathbf{k})\quad
1057: \quad \forall \;l\in B(\mathbf{k})
1058: \end{equation}
1059:  \item by equating (\ref{3.17c}) and (\ref{3.34c}) we obtain
1060: \begin{equation}\label{3.43}
1061: \Im [G^{C(-1)}(\mathbf{k},\omega )]=\pm \Im
1062: [G^{R,A(+1)}(\mathbf{k},\omega )]
1063: \end{equation}
1064:  \item from the sum rule (\ref{3.39}) and by means of (\ref{3.42b}) and
1065: (\ref{3.4}) we obtain
1066: \begin{equation}\label{3.44}
1067: I_{ab}^{(-1)}(\mathbf{k})=\sum_{l=1}^n  \tanh \left( {{{\beta \omega
1068: _l(\mathbf{k})} \over 2}} \right)\sum_{c=1}^n  \Omega
1069: _{al}(\mathbf{k})\Omega
1070: _{lc}^{-1}(\mathbf{k})I_{cb}^{(+1)}(\mathbf{k})
1071: \end{equation}
1072: \end{enumerate}
1073: 
1074: We see that the general structure of the GF is remarkably different
1075: according to the statistics. For fermionic composite fields (i.e.,
1076: when it is natural to choose $\eta =1$) all the Green's functions
1077: and correlation functions are completely determined. The
1078: zero-frequency function $\Gamma (\mathbf{k})$, defined on the Fermi
1079: surface $\omega _l(\mathbf{k})=\mu$, contributes to the spectral
1080: function $\rho ^{(+1)}(\mathbf{k},\omega )$ (see \ref{3.19}), it is
1081: directly related to the spectral density functions $\sigma
1082: ^{(l,+1)}(\mathbf{k})$ by means of equation (\ref{3.42a}), and its
1083: calculation does not require more information. Also, it does not
1084: contribute to the imaginary part of the causal GF. For bosonic
1085: composite fields (i.e., when it is natural to choose $\eta=-1$) the
1086: retarded and advanced GF are completely determined, but the causal
1087: GF and the correlation function depend on the zero-frequency
1088: function $\Gamma (\mathbf{k})$, defined on the surface  $\omega
1089: _l(\mathbf{k})=0$. It is now clear that the causal and retarded
1090: (advanced) GF contain different information and that the right
1091: procedure of calculation is controlled by the statistics. In
1092: particular, in the case of bosonic fields (i.e., for $\eta =-1$) one
1093: must start from the causal function and then use (3.34) to compute
1094: the other GF. On the contrary, for fermionic fields (i.e., for $\eta
1095: =1$) the right procedure for computing the correlation function
1096: requires first the calculation of the retarded (advanced) function
1097: and then the use (\ref{3.17a}), (\ref{3.17b}) and (\ref{3.17c}) to
1098: compute the other GF. Moreover, it is worth noting that $\Gamma
1099: (\mathbf{k})$ is undetermined within the bosonic sector (i.e., $\eta
1100: =-1$). It is true that $\Gamma (\mathbf{k})$ could be computed by
1101: considering an anticommutating algebra: remaining in the bosonic
1102: sector we make the choice $\eta =1$ and $\Gamma (\mathbf{k})$ can be
1103: calculated by (\ref{3.42a}) or equivalently by means of the
1104: following relation $\Gamma (\mathbf{k})={1 \over 2}\mathop {\lim
1105: }_{\omega \to 0}\omega G^{C(+1)}(\mathbf{k},\omega )$ which can be
1106: easily obtained from (\ref{3.31b}). However, the calculation of the
1107: $\sigma ^{(l,+1)}(\mathbf{k})$ requires the calculation of the
1108: normalization matrix $I^{(+1)}(\mathbf{k})$ that, for bosonic
1109: fields, generates unknown momentum dependent correlation functions
1110: whose determination can be very cumbersome as requires, at least in
1111: principle, the self-consistent solution of the integral equations
1112: connecting them to the corresponding Green's functions. In practice,
1113: also for simple, but anyway composite, bosonic fields the $\Gamma
1114: (\mathbf{k})$ remains undetermined and other methods should be used.
1115: Similar methods, like the use of the relaxation function
1116: \cite{Kubo_57}, would lead to the same problem.
1117: 
1118: Actually, all issues related to $\Gamma (\mathbf{k})$ have a natural
1119: playground in dealing with the ergodicity of the dynamics under
1120: investigation. More detail on this topic can be found in a
1121: manuscript in this same issue \cite{Avella_06}.
1122: 
1123: The formulation given in this Section needs some modifications in
1124: the case of zero temperature. In particular, Eqs.~(\ref{3.13a}) and
1125: (\ref{3.13b}) are not applicable and we must proceed in a different
1126: way. After a straightforward derivation \cite{Mancini_00}, it is
1127: immediate to see that the limit $T\to 0$ of the expressions
1128: (\ref{3.15c}), (\ref{3.15d}), (\ref{3.31c}) and (\ref{3.31d}) gives
1129: the right result.
1130: 
1131: \section{A self-consistent scheme}
1132: 
1133: As stressed in Section 1, in the study of highly interacting
1134: systems, where traditional perturbative calculations in terms of the
1135: noninteracting fields fail, a way to reconcile the powerful
1136: perturbation theory with the presence of complex and/or strong
1137: interactions is to describe the system in terms of a new set of
1138: fields, composite operators, generated by the interactions
1139: themselves. These field operators undoubtedly constitute a better
1140: starting point: they appear as the final result of the modifications
1141: imposed by the interactions on the original particles and contain,
1142: by the very beginning, the effects of the correlations.   Once a
1143: choice of composite fields has been made, the relative Green's
1144: function formalism can be set up, as illustrated in Section 2, where
1145: a generalized Dyson equation has been derived. On this basis one can
1146: construct a non-standard perturbation formalism where the basic
1147: ingredients are the propagators of a subset of the fundamental
1148: basis, satisfying linear equations of motion [cfr.~(\ref{3.1})]. By
1149: means of the equations of motion and by using the boundary
1150: conditions related to the definitions of the various Green's
1151: functions we have been able to derive explicit expressions for these
1152: latter [cfr.~(\ref{3.15c}), (\ref{3.15d}), (\ref{3.31c}) and
1153: (\ref{3.31d})]. However, these expressions can only determine the
1154: functional dependence; the knowledge of the GF is not fully achieved
1155: yet. The reason is that the algebra of the field is not canonical.
1156: As a consequence, the inhomogeneous terms $I^{(\eta )}(\mathbf{k})$
1157: in the equations of motion (\ref{3.2}) and the energy matrix
1158: $\varepsilon (\mathbf{k})$ contain some unknown static correlation
1159: functions, correlators, that have to be self-consistently
1160: calculated. Three serious problems arise with the study of the
1161: Green's functions: (a) the calculation of some parameters expressed
1162: as correlation functions of field operators not belonging the chosen
1163: basis; (b) the appearance of some zero-frequency constants (ZFC) and
1164: their determination; (c) the problem of fixing the representation
1165: where the Green's functions are formulated.
1166: 
1167: In the Composite Operator Method \cite{Mancini_00,Mancini_04} (COM)
1168: the three problems (a), (b) and (c) are not considered separately
1169: but they are all connected in one self-consistent scheme. The main
1170: idea is that fixing the values of the unknown parameters and of the
1171: ZFC implies to put some constraints on the representation where the
1172: GF are realized. As the determination of this representation is not
1173: arbitrary, it is clear that there is no freedom in fixing these
1174: quantities. They must assume values compatible with the dynamics and
1175: with the right representation. Which is the right representation?
1176: 
1177: From the algebra it is possible to derive several relations among
1178: the operators. We will call algebra constraints (AC) all possible
1179: relations among the operators dictated by the algebra. This set of
1180: relations valid at microscopic level must be satisfied also at
1181: macroscopic level, when expectations values are considered. Also, we
1182: note that, in general, the Hamiltonian has some symmetry properties
1183: (i.e. rotational invariance in coordinate and spin space, phase
1184: invariance, gauge invariance,\ldots). These symmetries generate a
1185: set of relations among the matrix elements: the Ward-Takahashi
1186: identities \cite{Ward_50,Takahashi_57} (WT).
1187: 
1188: Now, certainly the right representation must be the one where the
1189: relations among the operators and the conservation laws are
1190: maintained when expectation values are taken; in other words, all
1191: the AC and WT are preserved. By imposing these conditions we obtain
1192: a set of self-consistent equations that will fix the unknown
1193: correlators, the ZFC and the right representation at the same time.
1194: Several equations can be written down, according to the different
1195: symmetries we want to preserve. A large class of self-consistent
1196: equations is given by the following equation
1197: \begin{equation}\label{4.1}
1198: \langle \psi (i)\psi ^\dag (i)\rangle ={1 \over N}\sum_\mathbf{k} {1
1199: \over {2\pi }}\int_{-\infty }^{+\infty }  d\omega \,C_{\psi \psi
1200: ^\dag }(\mathbf{k},\omega )
1201: \end{equation}
1202: where the l.h.s. is fixed by the AC, the WT and the boundary
1203: conditions compatible with the phase under investigation and in the
1204: r.h.s. the correlation function $C_{\psi \psi ^\dag
1205: }(\mathbf{k},\omega )$ is computed by means of the equation of
1206: motion, as illustrated in Section 3. Equations (\ref{4.1}) generate
1207: a set of self-consistent equations which determine the unknown
1208: parameters (i.e., ZFC and unknown correlators) and, consequently,
1209: the proper representation \cite{Mancini_00,Mancini_04}, avoiding the
1210: problem of uncontrolled and uncontrollable decoupling.
1211: 
1212: \section{Approximation schemes}
1213: 
1214: 
1215: The generalized Dyson equation (\ref{2.26}) is an exact equation and
1216: permits, in principle, once the normalization matrix
1217: $I(\mathbf{i},\mathbf{j})$ [cfr. Eq. (2.13)], the m-matrix
1218: $m(\mathbf{i},\mathbf{j})$ [cfr. Eq. (2.14)] and the propagator
1219: $B(i,j)$ [cfr. Eq. (2.24)] are known, in the framework of the
1220: self-consistent scheme outlined in Sections 3 and 4, the calculation
1221: of the various Green's functions. However, for most of the physical
1222: systems of interest the calculation of the propagator $B(i,j)$ is a
1223: very difficult task and some approximations are needed. Various
1224: approximate schemes have been proposed. We will summarize some of
1225: them.
1226: 
1227: \subsection{The n-pole approximation}
1228: 
1229: The simplest approximation is based on completely neglecting the
1230: dynamical part $\Sigma (\mathbf{k},\omega )$ [cfr. (2.28)] of the
1231: self-energy. This approximation is largely used in the literature
1232: \cite{Hubbard_63,Hubbard_64,Hubbard_64a,Becker_90,Mori_65,Mori_65a,Fedro_92,Fulde_95,Plakida_89,Mehlig_95,Rowe_68,Roth_69,Beenen_95,Kalashnikov_73,Nolting_72,%
1233: Geipel_88,Nolting_89,Lonke_71,Tserkovnikov_81,Tserkovnikov_81a,Ishihara_94,Mancini_94,Mancini_95,Mancini_95a,Mancini_95b,Avella_98,Shimahara_91,Kruger_94,%
1234: Kondo_72,Yamaji_73} and is called pole-approximation. In this
1235: approximation we only need the knowledge of the normalization matrix
1236: and the $m$-matrix. The constraint (\ref{2.9}) produces a physics of
1237: the solution totally extraneous to the complementary physical space,
1238: orthogonal to that spanned by the multiplet $\psi(i)$. This
1239: approximation, or assumption to a larger extent, consists in
1240: retaining that one can neglect finite life-time effects (i.e., the
1241: dynamical part of the self-energy) paying attention to the choice of
1242: a proper extended operatorial basis, with respect to which the
1243: self-energy corrections have a small total weight. Indeed, the total
1244: weight of the corrections is bounded by the thermal average
1245: (\ref{2.24}) involving the residual source $\delta J (i)$. It is
1246: worth noting \cite{Mancini_98b} that the $n$-pole structure of the
1247: various GF corresponds to a Dyson-like equation
1248: \begin{equation}\label{5.1}
1249: G_{ab}^Q(\mathbf{k},\omega )={{I_{ab}(\mathbf{k})} \over {\omega
1250: -\Sigma _{ab}^Q(\mathbf{k},\omega )}}
1251: \end{equation}
1252: where the self-energy components $\Sigma _{ab}^Q(\mathbf{k},\omega
1253: )$ have a $(n-1)$-pole structure.
1254: 
1255: \subsection{Self-consistent Born approximation}
1256: 
1257: In order to improve the approximation one needs to take into account
1258: self-energy corrections by developing some methods to calculate the
1259: effects of $\Sigma (\mathbf{k},\omega )$. In the self-consistent
1260: Born approximation (SCBA), or non-crossing approximation, the
1261: many-particle Green's functions, appearing in the expression of
1262: $\Sigma (\mathbf{k},\omega )$ [see (\ref{2.28})], are calculated by
1263: assuming that the fermionic and bosonic modes propagate
1264: independently.
1265: 
1266: By recalling the results given in Section 2, the knowledge of the
1267: self-energy requires the calculation of the higher-order propagator
1268: $B^Q(i,j)=\langle Q[\delta J(i)\delta J^\dag (j)]\rangle $. In order
1269: to illustrate the approximation, let us consider the case where the
1270: basic set $\{ \psi (i)\} $ is of a fermionic type. Then, typically
1271: we have to calculate GF of the form $H^R(i,j)=\langle
1272: R[B(i)F(i)F^\dag (j)B^\dag (j)]\rangle $ where $F(i)$ and $B(i)$ are
1273: fermionic and bosonic field operators, respectively. By means of the
1274: spectral representation (\ref{3.20a}) we can write
1275: \begin{equation}\label{5.3}
1276: H^R(\mathbf{k},\omega )=-{1 \over \pi }\int_{-\infty }^{+\infty }
1277: d\omega' {1 \over {\omega -\omega +\mathrm{i}\varepsilon }}\coth
1278: {{\beta \omega' } \over 2}\Im [H^c(\mathbf{k},\omega' )]
1279: \end{equation}
1280: where $H^C(i,j)=\langle T[B(i)F(i)F^\dag (j)B^\dag (j)]\rangle $ is
1281: the causal function. In the SCBA we approximate $H^C(i,j) \approx
1282: f^C(i,j)b^C(i,j)$ where
1283: \begin{equation}\label{5.5}
1284: f^C(i,j)=\langle T[F(i)F^\dag (j)]\rangle \quad \quad \quad
1285: b^C(i,j)=\langle T[B(i)B^\dag (j)]\rangle
1286: \end{equation}
1287: 
1288: This approximation has been used in many works (for instance see
1289: \cite{Plakida_99,Plakida_01,Avella_03c,Krivenko_04}. By assuming
1290: that the system is ergodic we can use the spectral representations
1291: (\ref{3.21b}) and (\ref{3.38}) to obtain
1292: \begin{align}\label{5.6}
1293: &f^C(\mathbf{k},\omega )=-{1 \over \pi }\int_{-\infty }^{+\infty }
1294: d\omega' [{{1-f_\mathrm{F}(\beta \omega' )} \over {\omega -\omega'
1295: +\mathrm{i}\delta }}+{{f_\mathrm{F}(\beta \omega' )} \over {\omega
1296: -\omega' -\mathrm{i}\delta }}]\Im [f^R(\mathbf{k},\omega )]\\
1297: &b^C(\mathbf{k},\omega )=-{1 \over \pi }\int_{-\infty }^{+\infty }
1298: d\omega' [{{1+f_\mathrm{B}(\beta \omega' )} \over {\omega -\omega
1299: +\mathrm{i}\delta }}-{{f_\mathrm{B}(\beta \omega' )} \over {\omega
1300: -\omega' -\mathrm{i}\delta }}]\Im [b^R(\mathbf{k},\omega )]
1301: \end{align}
1302: 
1303: Use of (\ref{5.3})-(\ref{5.6}) leads to
1304: \begin{align}\label{5.7}
1305: &H^R(\mathbf{k},\omega )={1 \over \pi }\int_{-\infty }^{+\infty }
1306: d\omega' {1 \over {\omega -\omega' +\mathrm{i}\delta }}{{a^d} \over
1307: {(2\pi )^{d+1}}}\int_{\Omega _B}  d^dp d\Omega \Im
1308: [f^R(\mathbf{p},\Omega
1309: )]\nonumber\\
1310: &\Im [b^R(\mathbf{k}-\mathbf{p},\omega' -\Omega )][\tanh {{\beta
1311: \Omega } \over 2}+\coth {{\beta (\omega' -\Omega )} \over 2}]
1312: \end{align}
1313: where $d$ is the dimensionality of the system, $a$ is the lattice
1314: constant and $\Omega_\mathrm{B}$ is the volume of the Brillouin
1315: zone.
1316: 
1317: It should be noted that the SCBA can also be applied to the
1318: correlation function. We start from the expression
1319: \begin{equation}\label{5.8}
1320: H^R(\mathbf{k},\omega )={1 \over {2\pi }}\int_{-\infty }^{+\infty }
1321: d\omega' {{1+e^{-\beta \omega' }} \over {\omega -\omega'
1322: +\mathrm{i}\varepsilon }}H(\mathbf{k},\omega' )
1323: \end{equation}
1324: where $H(i,j)=\langle B(i)F(i)F^\dag (j)B^\dag (j)\rangle $ is the
1325: correlation function. In the SCBA we approximate
1326: \begin{equation}\label{5.9}
1327: H(i,j)=\langle B(i)F(i)F^\dag (j)B^\dag (j)\rangle \approx \langle
1328: F(i)F^\dag (j)\rangle \langle B(i)B^\dag (j)\rangle
1329: \end{equation}
1330: Then, by proceeding in the similar way we arrive to the same
1331: expression (\ref{5.7}).
1332: 
1333: \subsection{Two-site resolvent approach}
1334: 
1335: In this subsection we will consider an approximation scheme
1336: \cite{Matsumoto_96,Matsumoto_97}, where the dynamical part $\Sigma
1337: (\mathbf{k},\omega )$ [cfr.~(\ref{2.28})] of the self-energy is
1338: estimated by a two-site approximation in combined use with the
1339: resolvent method \cite{Kuramoto_83}. In this approximation the
1340: higher order propagator (\ref{2.24}) is written as
1341: \begin{equation}\label{5.10}
1342: B^Q(\mathbf{k},\omega )=F.T.\langle Q[\delta J(i)\delta J^\dag
1343: (j)]\rangle \approx B_0^Q(\omega )+\alpha (\mathbf{k})B_1^Q(\omega )
1344: \end{equation}
1345: where $B_0^Q(\omega )$ is related to level transitions on equal site
1346: \begin{equation}\label{5.11}
1347: B_0^Q(\omega )={1 \over {2d}}F.T.\langle R[\delta
1348: J(\mathbf{i},t_i)\delta J^\dag (\mathbf{i},t_j)]\rangle
1349: \end{equation}
1350: while $B_1^Q(\omega )$ is related to transitions across two sites
1351: \begin{equation}\label{5.12}
1352: B_1^Q(\omega )={1 \over {2d}}F.T.\langle R[\delta
1353: J(\mathbf{i},t_i)\delta J^{\dag\alpha} (\mathbf{i},t_j)]\rangle
1354: \end{equation}
1355: 
1356: The Green's function (\ref{2.27}) takes the form
1357: \begin{equation}\label{5.13}
1358: G^Q(\mathbf{k},\omega )={1 \over {\omega -\varepsilon
1359: (\mathbf{k})+t^2V(\omega )\alpha (\mathbf{k})}}I(\mathbf{k})
1360: \end{equation}
1361: where $V(\omega )$ has to be calculated from the definition
1362: (\ref{2.28}) by making use of approximation (\ref{5.10}). We give
1363: now a brief sketch of the calculation.
1364: 
1365: $B_0^Q(\omega )$ and $B_1^Q(\omega )$ are computed by expressing
1366: $\delta J(i)$ in terms of transitions among the two-site levels
1367: \begin{equation}\label{5.14}
1368: \delta J=\sum_{nm}  a_{nm}\Phi _n^\dag \Phi _m
1369: \end{equation}
1370: where $\{ \Phi _m\} $ is the complete set of operators for two-site
1371: levels. By means of the non crossing approximation
1372: \cite{Kuramoto_83,Grewe_83}, the propagator $\langle Q[\Phi _n^\dag
1373: (t_i)\Phi _m(t_i)\Phi _{n}^\dag (t_j)\Phi _m(t_j)]\rangle $ is
1374: expressed in terms of the resolvent $R_{nm}(t_i-t_j)=\langle Q[\Phi
1375: _n(t_i)\Phi _m^\dag (t_j)]\rangle _R$ where  the subscript $R$
1376: indicates the reservoir system, i.e., the part $H_R$ of the
1377: Hamiltonian other than that concerned with two sites. The
1378: calculation of the resolvent brings to a modification of the
1379: original two-site levels by the surroundings. In this scheme effects
1380: of time delay in the local correlations are treated trough the
1381: time-dependent modifications of the two-site level transitions,
1382: which are included as time-dependent local effects in the electron
1383: self-energy.
1384: 
1385: \section{Conclusions}
1386: 
1387: In this article we have illustrated an approach to the study of
1388: highly correlated electronic systems, based on the equation of
1389: motion and Green's function method.  Such an approach is based on
1390: two main ideas: (i) propagators of composite operators as building
1391: blocks at the basis of approximate calculations; (ii) use of algebra
1392: constrains to fix the representation of the GF in order to maintain
1393: the algebraic and symmetry properties. This formalism has been
1394: applied to the study of several models of highly interacting
1395: systems, and we refer the interested readers to
1396: Ref.~\cite{Mancini_04} for an exhaustive bibliography.
1397: 
1398: %\bibliographystyle{prsty}
1399: %\bibliography{biblio}
1400: 
1401: \begin{thebibliography}{10}
1402: 
1403: \bibitem{Anderson_87}
1404: P.~W. Anderson, Science {\bf 235},  1196  (1987).
1405: 
1406: \bibitem{Fulde_95}
1407: P. Fulde, {\em Electron {C}orrelations in {M}olecules and {S}olids},
1408: 3rd  ed.
1409:   (Springer-Verlag, Berlin, 1995).
1410: 
1411: \bibitem{Bogoliubov_47}
1412: N.~N. Bogoliubov, J.~Phys.~USSR {\bf 11},  23  (1947).
1413: 
1414: \bibitem{Dancoff_50}
1415: S.~M. Dancoff, Phys.~Rev. {\bf 78},  382  (1950).
1416: 
1417: \bibitem{Zwanzig_61}
1418: R. Zwanzig,  in {\em Lectures in {T}heoretical {P}hysics}, edited by
1419: W.
1420:   Britton, B. Downs, and J. Downs (Interscience, New York, 1961), Vol.~3, p.\
1421:   106.
1422: 
1423: \bibitem{Mori_65}
1424: H. Mori, Prog.~Theor.~Phys. {\bf 33},  423  (1965).
1425: 
1426: \bibitem{Umezawa_93}
1427: H. Umezawa, {\em Advanced {F}ield {T}heory: {M}icro, {M}acro and
1428: {T}hermal
1429:   {P}hysics} (AIP, New York, 1993), and references therein.
1430: 
1431: \bibitem{Hubbard_63}
1432: J. Hubbard, Proc.~Roy.~Soc.~A {\bf 276},  238  (1963).
1433: 
1434: \bibitem{Hubbard_64}
1435: J. Hubbard, Proc.~Roy.~Soc.~A {\bf 277},  237  (1964).
1436: 
1437: \bibitem{Hubbard_64a}
1438: J. Hubbard, Proc.~Roy.~Soc.~A {\bf 281},  401  (1964).
1439: 
1440: \bibitem{Rowe_68}
1441: D.~J. Rowe, Rev.~Mod.~Phys. {\bf 40},  153  (1968).
1442: 
1443: \bibitem{Roth_69}
1444: L.~M. Roth, Phys.~Rev. {\bf 184},  451  (1969).
1445: 
1446: \bibitem{Tserkovnikov_81}
1447: Y.~A. Tserkovnikov, Teor.~Mat.~Fiz. {\bf 49},  219  (1981).
1448: 
1449: \bibitem{Tserkovnikov_81a}
1450: Y.~A. Tserkovnikov, Teor.~Mat.~Fiz. {\bf 50},  261  (1981).
1451: 
1452: \bibitem{Barnes_76}
1453: S.~E. Barnes, J.~Phys.~F {\bf 6},  1375  (1976).
1454: 
1455: \bibitem{Coleman_84}
1456: P. Coleman, Phys.~Rev.~B {\bf 29},  3035  (1984).
1457: 
1458: \bibitem{Kotliar_86}
1459: G. Kotliar and A.~E. Ruckenstein, Phys.~Rev.~Lett. {\bf 57},  1362
1460: (1986).
1461: 
1462: \bibitem{Kalashnikov_69}
1463: O.~K. Kalashnikov and E.~S. Fradkin, Sov.~Phys.~JETP {\bf 28},  317
1464: (1969).
1465: 
1466: \bibitem{Nolting_72}
1467: W. Nolting, Z.~Phys. {\bf 255},  25  (1972).
1468: 
1469: \bibitem{Ishihara_94}
1470: S. Ishihara, H. Matsumoto, S. Odashima, M. Tachiki, and F. Mancini,
1471:   Phys.~Rev.~B {\bf 49},  1350  (1994).
1472: 
1473: \bibitem{Mancini_94}
1474: F. Mancini, S. Marra, A.~M. Allega, and H. Matsumoto, Physica~C {\bf
1475: 235},
1476:   2253  (1994).
1477: 
1478: \bibitem{Mancini_95}
1479: F. Mancini, S. Marra, and H. Matsumoto, Physica~C {\bf 244},  49
1480: (1995).
1481: 
1482: \bibitem{Mancini_95a}
1483: F. Mancini, S. Marra, and H. Matsumoto, Physica~C {\bf 250},  184
1484: (1995).
1485: 
1486: \bibitem{Mancini_95b}
1487: F. Mancini, S. Marra, and H. Matsumoto, Physica~C {\bf 252},  361
1488: (1995).
1489: 
1490: \bibitem{Avella_98}
1491: A. Avella, F. Mancini, D. Villani, L. Siurakshina, and V.~Y.
1492: Yushankhai,
1493:   Int.~J.~Mod.~Phys.~B {\bf 12},  81  (1998).
1494: 
1495: \bibitem{Mancini_00}
1496: F. Mancini and A. Avella, Eur.~Phys.~J.~B {\bf 36},  37  (2003).
1497: 
1498: \bibitem{Mancini_04}
1499: F. Mancini and A. Avella, Adv.~Phys. {\bf 53},  537  (2004).
1500: 
1501: \bibitem{Izyumov_90}
1502: Y.~A. Izyumov and B.~M. Letfulov, J.~Phys.:~Condens.~Matter {\bf 2},
1503: 8905
1504:   (1990).
1505: 
1506: \bibitem{Matsumoto_94}
1507: H. Matsumoto, A. Allega, S. Odashima, and F. Mancini, Physica~C {\bf
1508: 235},
1509:   2227  (1994).
1510: 
1511: \bibitem{Mancini_00e}
1512: F. Mancini, N.~B. Perkins, and N.~M. Plakida, Phys.~Lett.~A {\bf
1513: 284},  286
1514:   (2001).
1515: 
1516: \bibitem{Villani_00}
1517: D. Villani, E. Lange, A. Avella, and G. Kotliar, Phys.~Rev.~Lett.
1518: {\bf 85},
1519:   804  (2000).
1520: 
1521: \bibitem{Avella_02}
1522: A. Avella, R. Hayn, and F. Mancini, Eur.~Phys.~J.~B {\bf 37},  465
1523: (2004).
1524: 
1525: \bibitem{Bak_02}
1526: M. Bak and F. Mancini, Physica~B {\bf 312},  732  (2002).
1527: 
1528: \bibitem{Bogoliubov_59}
1529: N.~N. Bogoliubov and S.~V. Tyablikov, Dokl.~Akad.~Nauk.~USSR {\bf
1530: 126},  53
1531:   (1959).
1532: 
1533: \bibitem{Zubarev_60}
1534: D.~N. Zubarev, Sov.~Phys.~Uspekhi {\bf 3},  320  (1960).
1535: 
1536: \bibitem{Zubarev_74}
1537: D.~N. Zubarev, {\em Non {E}quilibrium {S}tatistical
1538: {T}hermodynamics}
1539:   (Consultants Bureau, New York, 1974).
1540: 
1541: \bibitem{Umezawa_65}
1542: H. Umezawa, Acta~Phys.~Hung. {\bf 19},  9  (1965).
1543: 
1544: \bibitem{Umezawa_66}
1545: H. Umezawa, Suppl.~of~Progr.~Theor.~Phys. {\bf 37-38},  585  (1966).
1546: 
1547: \bibitem{Leplae_74}
1548: L. Leplae, F. Mancini, and H. Umezawa, Phys.~Rep. {\bf 10},  151
1549: (1974).
1550: 
1551: \bibitem{Mancini_05}
1552: F. Mancini, Europhys.~Lett. {\bf 70},  485  (2005).
1553: 
1554: \bibitem{Mancini_05b}
1555: F. Mancini, Eur.~Phys.~J.~B {\bf 45},  497  (2005).
1556: 
1557: \bibitem{Mancini_05a}
1558: F. Mancini, Eur.~Phys.~J.~B {\bf 47},  527  (2005).
1559: 
1560: \bibitem{Avella_05}
1561: A. Avella and F. Mancini, Eur.~Phys.~J.~B {\bf 50},  527  (2006).
1562: 
1563: \bibitem{Mermin_66}
1564: N.~D. Mermin and H. Wagner, Phys.~Rev.~Lett. {\bf 17},  1133
1565: (1966).
1566: 
1567: \bibitem{Kubo_57}
1568: R. Kubo, J.~Phys.~Soc.~Jpn. {\bf 12},  570  (1957).
1569: 
1570: \bibitem{Avella_06}
1571: A. Avella, F. Mancini, and E. Plekhanoff, Condens.~Matter~Phys. {\bf
1572: 9},  ???
1573:   (2006).
1574: 
1575: \bibitem{Ward_50}
1576: J.~C. Ward, Phys.~Rev. {\bf 78},  182  (1950).
1577: 
1578: \bibitem{Takahashi_57}
1579: Y. Takahashi, Nuovo Cimento {\bf 6},  370  (1957).
1580: 
1581: \bibitem{Becker_90}
1582: K.~W. Becker, W. Brenig, and P. Fulde, Z.~Phys.~B {\bf 81},  165
1583: (1990).
1584: 
1585: \bibitem{Mori_65a}
1586: H. Mori, Prog.~Theor.~Phys. {\bf 34},  399  (1965).
1587: 
1588: \bibitem{Fedro_92}
1589: A.~J. Fedro, Y. Zhou, T.~C. Leung, B.~N. Harmon, and S.~K. Sinha,
1590: Phys.~Rev.~B
1591:   {\bf 46},  14785  (1992).
1592: 
1593: \bibitem{Plakida_89}
1594: N.~M. Plakida, V.~Y. Yushankhai, and I.~V. Stasyuk, Physica~C {\bf
1595: 162-164},
1596:   787  (1989).
1597: 
1598: \bibitem{Mehlig_95}
1599: B. Mehlig, H. Eskes, R. Hayn, and M.~B.~J. Meinders, Phys.~Rev.~B
1600: {\bf 52},
1601:   2463  (1995).
1602: 
1603: \bibitem{Beenen_95}
1604: J. Beenen and D.~M. Edwards, Phys.~Rev.~B {\bf 52},  13636  (1995).
1605: 
1606: \bibitem{Kalashnikov_73}
1607: O.~K. Kalashnikov and E.~S. Fradkin, Phys.~Stat.~Sol.~(b) {\bf 59},
1608: 9  (1973).
1609: 
1610: \bibitem{Geipel_88}
1611: G. Geipel and W. Nolting, Phys.~Rev.~B {\bf 38},  2608  (1988).
1612: 
1613: \bibitem{Nolting_89}
1614: W. Nolting and W. Borgiel, Phys.~Rev.~B {\bf 39},  6962  (1989).
1615: 
1616: \bibitem{Lonke_71}
1617: A. Lonke, J.~Math.~Phys. {\bf 12},  2422  (1971).
1618: 
1619: \bibitem{Shimahara_91}
1620: H. Shimahara and S. Takada, J.~Phys.~Soc.~Jpn. {\bf 60},  2394
1621: (1991).
1622: 
1623: \bibitem{Kruger_94}
1624: P. Kr{\"u}ger and P. Schuck, Europhys.~Lett. {\bf 27},  395  (1994).
1625: 
1626: \bibitem{Kondo_72}
1627: J. Kondo and K. Yamaji, Prog.~Theor.~Phys. {\bf 47},  807  (1972).
1628: 
1629: \bibitem{Yamaji_73}
1630: K. Yamaji and J. Kondo, Phys.~Lett.~A {\bf 45},  317  (1973).
1631: 
1632: \bibitem{Mancini_98b}
1633: F. Mancini, Phys.~Lett.~A {\bf 249},  231  (1998).
1634: 
1635: \bibitem{Plakida_99}
1636: N. Plakida and V. Oudovenko, Phys.~Rev.~B {\bf 59},  11949  (1999).
1637: 
1638: \bibitem{Plakida_01}
1639: N.~M. Plakida, L. Anton, S. Adam, and G. Adam, JETP {\bf 97},  331
1640: (2003).
1641: 
1642: \bibitem{Avella_03c}
1643: A. Avella, S. Krivenko, F. Mancini, and N.~M. Plakida,
1644: J.~Magn.~Magn.~Mat. {\bf
1645:   272},  456  (2004).
1646: 
1647: \bibitem{Krivenko_04}
1648: S. Krivenko, A. Avella, F. Mancini, and N. Plakida, Physica~B {\bf
1649: 359-361},
1650:   666  (2005).
1651: 
1652: \bibitem{Matsumoto_96}
1653: H. Matsumoto, T. Saikawa, and F. Mancini, Phys.~Rev.~B {\bf 54},
1654: 14445
1655:   (1996).
1656: 
1657: \bibitem{Matsumoto_97}
1658: H. Matsumoto and F. Mancini, Phys.~Rev.~B {\bf 55},  2095  (1997).
1659: 
1660: \bibitem{Kuramoto_83}
1661: Y. Kuramoto, Z.~Phys.~B {\bf 53},  37  (1983).
1662: 
1663: \bibitem{Grewe_83}
1664: N. Grewe, Z.~Phys.~B {\bf 53},  271  (1983).
1665: 
1666: \end{thebibliography}
1667: 
1668: 
1669: \label{last@page}
1670: \end{document}
1671: