1: \documentclass[12pt,showpacs,prd]{revtex4}
2: \usepackage{epsfig,amssymb,amsfonts}
3:
4: \makeatletter
5: \renewcommand{\@makefntext}[1]{\parindent=1em\noindent\hbox to 1.8em{\hss$^{\@thefnmark}$}#1}
6: \renewcommand{\@footnotemark}{\hbox{\mathsurround=0pt$^{\@thefnmark}$}}
7: \newcommand{\ftnote}[2]{\footnotemark[#1]\footnotetext[#1]{#2}}
8: \makeatother
9:
10: \begin{document}
11: \title{Mesonic states and vacuum replicas in potential quark models for
12: QCD}\author{A. V. Nefediev}
13: \affiliation{Institute of Theoretical and Experimental Physics, 117218,\\
14: B.Cheremushkinskaya 25, Moscow, Russia}
15: \author{J. E. F. T. Ribeiro}
16: \affiliation{ Centro de F\'\i sica das Interac\c c\~oes Fundamentais
17: (CFIF),Departamento de F\'\i sica, Instituto Superior T\'ecnico, Av.
18: Rovisco Pais, P-1049-001 Lisboa, Portugal}
19: \newcommand{\be}{\begin{equation}}
20: \newcommand{\bea}{\begin{eqnarray}}
21: \newcommand{\ee}{\end{equation}}
22: \newcommand{\eea}{\end{eqnarray}}
23: \newcommand{\ds}{\displaystyle}
24: \newcommand{\low}[1]{\raisebox{-1mm}{$#1$}}
25: \newcommand{\loww}[1]{\raisebox{-1.5mm}{$#1$}}
26: \newcommand{\lmn}{\mathop{\sim}\limits_{n\gg 1}}
27: \newcommand{\vpint}{\int\makebox[0mm][r]{\bf --\hspace*{0.13cm}}}
28: \newcommand{\too}{\mathop{\to}\limits_{N_C\to\infty}}
29: \newcommand{\vp}{\varphi}
30:
31: \begin{abstract}
32: In this paper, we study generic non-local NJL models for
33: four-dimensional QCD in the limit of a large number of colours. We
34: diagonalise, through a Bogoliubov-type transformation, the
35: Hamiltonian of the model completely in terms of the compound
36: operators creating/annihilating mesons --- bound states of dressed
37: quarks and antiquarks --- and demonstrate a one-to-one
38: correspondence between the Bogoliubov diagonalisation condition
39: and the Bethe--Salpeter equation for bound states. We conclude
40: that, in the leading order in $N_C$, the nontrivial contents of
41: the theory is entirely encoded in the chiral angle --- the
42: solution to the mass-gap equation for dressed quarks. As a
43: consequence, we extend the statement of existence of excited
44: solutions to the mass-gap equation, called vacuum replicas, beyond
45: the BCS level and prescribe a new index to the mesonic operators
46: --- the label of the vacuum state in which the
47: corresponding meson is created.
48: \end{abstract}
49: \pacs{12.38.Aw, 12.39.Ki, 12.39.Pn}
50: \maketitle
51:
52: \section{Introduction}
53:
54: For nearly half a century Nambu--Jona-Lasinio (NJL)\cite{NJL}
55: class of models \cite{Orsay,Orsay2,Lisbon} have been quite useful
56: tools for the on going study and understanding of low-energy
57: phenomena in QCD. Quark models with non-local quark
58: current-current microscopic interactions, parameterised by kernels
59: of particular forms, constitute prominent steps towards building
60: an effective low-energy theory for the strong interactions.
61: Besides regularising the ultraviolet divergences of the theory,
62: these nonlocal quark kernels bring in the necessary interaction
63: scale needed to make contact with the hadronic phenomenology.
64: Models with quark instantaneous interaction modeled after the
65: harmonic oscillator kernels \cite{Orsay,Orsay2,Lisbon} or kernels
66: obtained in the Coulomb-gauge QCD \cite{coulg}, as well as a
67: variety of other approaches based on the linear confinement
68: \cite{linear}, constitute as many examples of closely related
69: approaches exploiting the idea of chirally nonsymmetric solutions
70: of a nonlinear gap equation. In this paper, we continue our
71: studies of quark models \cite{Lisbon} with instantaneous
72: inter-quark interaction. An important feature of this class of
73: models is the essential covariance of results and conclusions
74: across a variety of possible forms for the confining kernel. For
75: example, the pion-pion low-energy scattering problem can be solved
76: in the general form, in terms of the pion mass and decay constant,
77: regardless of the explicit form of the quark kernel, provided it
78: is confining \cite{pipi}. Another example of such universality is
79: given by the existence of multiple chirally non-invariant
80: solutions to the mass-gap equation, predicted in
81: \cite{replica1,replica2}, found both analytically and numerically
82: for all allowed power-like confining potentials in
83: \cite{replica4}, and independently confirmed by calculations done
84: in a different approach in \cite{replica3}. The same conclusion of
85: universality holds for any low-energy phenomena considered in the
86: framework of such class of models and, therefore, we dare to
87: expect to be able to reproduce the low-energy properties of QCD.
88: An important progress was achieved in \cite{2d} where a
89: two-dimensional model for QCD \cite{tHooft} in the axial gauge was
90: considered and a second, Bogoliubov-type, transformation was
91: performed in the mesonic sector to fully diagonalise, up to
92: corrections of order $O(1/\sqrt{N_C})$, the Hamiltonian of the
93: theory in terms of the mesonic creation/annihilation operators.
94: The Hamiltonian approach to two-dimensional QCD in the light-cone
95: gauge was also developed in \cite{japan}. In the present paper, we
96: extend this result to four-dimensional QCD and establish a
97: one-to-one correspondence between the Bogoliubov condition (which
98: ensures that the Hamiltonian is diagonalised in the mesonic
99: sector) and the Bethe--Salpeter equation for bound quark-antiquark
100: states. We build the corresponding amplitudes explicitly and study
101: in detail the cases of the chiral pion and of the $\rho$-meson. We
102: conclude that solving the Bethe--Salpeter equation in the ladder
103: (rainbow) approximation is equivalent, in the leading order in
104: $N_C$, to go beyond BCS level and to diagonalise the quartic term
105: in the Hamiltonian of the theory. This feature is general and
106: holds for any confining quark kernel. We use the oscillator-type
107: quark kernel just as an example to illustrate the method. We also
108: conclude that all the information needed for the second and final
109: mesonic Bogoliubov transformation, is already contained in the
110: solutions of the mass-gap equation appearing at the BCS level.
111: Therefore, in the leading order in $N_C$, the chiral angle --- the
112: solution to the mass-gap equation
113: --- remains the only nontrivial characteristic of the theory and
114: it defines the latter completely. As a result, we discuss the
115: multiple chirally noninvariant solutions of mass-gap equations
116: --- the vacuum replicas \cite{replica1,replica2,replica4} --- and
117: conclude that, once such solutions exist, this information gets
118: carried beyond BCS and appears as an extra spectroscopic index for
119: mesonic states on top of the usual set $\{ {\vec P},n,J^{PC}\}$---
120: namely, each meson carries the following quantum numbers: the
121: total momentum ${\vec P}$, the radial excitation number $n$, the
122: $J^{PC}$ set, and, finally, the label of the vacuum state
123: in which it is created. We discuss possible consequences of this
124: generalisation.
125:
126: The paper is organised as follows. The second section is first
127: devoted to the diagonalisation of the Hamiltonian in the
128: quark sector (subsection A) and then to its subsequent
129: diagonalisation in the mesonic sector (subsection B). We end this
130: chapter with the discussion of the equivalence between the mesonic
131: Bogoliubov-Valatin transformation and the Bethe--Salpeter
132: equation. Vacuum replicas are studied in the third chapter. We
133: give the necessary details of the theory of replicas, in the first
134: subsection, and comment on the tachyon problem, in the second
135: subsection. Finally, in the third subsection, we generalise the
136: theory of replicas beyond BCS and discuss the hadronic spectrum in
137: the replica vacua. The fourth and last chapter of the paper is
138: devoted to our conclusions and outlook.
139:
140: \section{Diagonalisation of the quark Hamiltonian}
141: \subsection{Diagonalisation in the single-quark sector and the mass-gap
142: equation}
143:
144: In this section, we give the necessary details of the quark model
145: \cite{Orsay,Orsay2,Lisbon} to be used in this paper and describe
146: the Hamiltonian diagonalisation in the single-quark sector (BCS
147: level). Naively one should expect this to be the end of the story.
148: However, as we shall see, the requirement of confinement, that is,
149: no free asymptotic quarks may be allowed, will completely change
150: this picture and we shall end up with a diagonal Hamiltonian
151: written in terms of asymptotic mesons rather than asymptotic
152: quarks. How this comes about will be the subject of this paper.
153:
154: The Hamiltonian of the model,
155: \be
156: \hat{H}=\int d^3 x\bar{\psi}(\vec{x},t)\left(-i\vec{\gamma}\cdot
157: \vec{\bigtriangledown}+m\right)\psi(\vec{x},t)+ \frac12\int d^3
158: xd^3y\;J^a_\mu(\vec{x},t)K^{ab}_{\mu\nu}(\vec{x}-\vec{y})J^b_\nu(\vec{y},t),
159: \label{H}
160: \ee
161: contains the interaction of the quark currents
162: $J_{\mu}^a(\vec{x},t)=\bar{\psi}(\vec{x},t)\gamma_\mu\frac{\lambda^a}{2}
163: \psi(\vec{x},t)$ parameterised through the instantaneous quark
164: kernel,
165: \be
166: K^{ab}_{\mu\nu}(\vec{x}-\vec{y})=g_{\mu 0}g_{\nu 0}\delta^{ab}
167: V_0(|\vec{x}-\vec{y}|).
168: \label{KK}
169: \ee
170: In Eq.~(\ref{KK}) we have used the simplest form of the kernel
171: compatible with the requirement of confinement. Spatial components
172: can be also included in the kernel (\ref{KK}) \cite{Lisbon}. They
173: affect, for example, the form of the mass-gap equation for the
174: chiral angle and spin-dependent terms in the effective quark-quark
175: interaction after a Foldy--Wounthuysen transformation (see, for
176: example, papers \cite{hl} where a heavy-light quarkonium was
177: considered in the modified Fock-Schwinger gauge and an effective
178: inter-quark interaction was derived).
179:
180: But it is a fact, already mentioned in the introduction, that
181: despite all the aforementioned detail variations of possible quark
182: kernels, they all share with the kernel (\ref{KK}) the {\emph
183: same} low energy properties (Goldstone pion, $\pi -\pi$ Weinberg
184: scattering lengths, Gell-Mann--Oakes--Renner formula, and so on) so that it
185: is sufficient to stick to the form (\ref{KK}) as the mainstay of
186: this work and comment on the generalisations when it is the case.
187: Finally it is sufficient for phenomenological applications that
188: $V_0(|\vec{x}|)$ should interpolate between long range confinement
189: and short-range Coulomb behaviour. An example of such a
190: phenomenological kernel which is able to address both, heavy-quark
191: and light-quark limits of the theory, was considered in detail in
192: \cite{replica1}. A pure confining kernel of the generalised
193: power-like form,
194: \be
195: V_0(|\vec{x}|)=K_0^{\alpha+1}|\vec{x}|^{\alpha},
196: \label{potential}
197: \ee
198: with $K_0$ being the mass parameter, was studied for the first
199: time, in the series of papers \cite{Orsay} and an analysis was performed
200: concerning the general properties of the mass-gap equations for
201: such potentials. This analysis was extended in a recent paper
202: \cite{replica4} and the range of possible powers $\alpha$ was
203: identified to be $0\leqslant\alpha\leqslant 2$. In what follows,
204: we do not need to specify any particular form for $V_0(|\vec{x}|)$
205: other than requiring it to be confining. When needed, we shall use
206: the oscillator-type quark kernel corresponding to the case of
207: $\alpha=2$ in Eq.~(\ref{potential}), to exemplify some general
208: results. Convenience of such a choice follows from the fact that
209: the Fourier transform of the potential $\vec{x}^2$ is given by the
210: Laplacian of the three-dimensional delta function, so that all
211: integral equations can be transformed into differential equations
212: which are much easier for analytical and numerical studies
213: \cite{Orsay,Orsay2,Lisbon}.
214:
215: Now we proceed in the standard way by defining dressed quarks with
216: the help of a Bogoliubov-Valatin transformation having the chiral
217: angle $\vp(p)\equiv\vp_p$ \cite{Orsay,Lisbon} as a parameter:
218: \be
219: \psi_{\alpha}(\vec{x},t)=\sum_{s=\uparrow,\downarrow}\int\frac{d^3p}{(2\pi)
220: ^3}e^{i\vec{p}\vec{x}} [\hat{b}_{\alpha
221: s}(\vec{p},t)u_s(\vec{p})+\hat{d}_{\alpha s}^\dagger(-\vec{p},t)
222: v_s(-\vec{p})],
223: \label{psi}
224: \ee
225: \be \left\{
226: \begin{array}{rcl}
227: u(\vec{p})&=&\frac{1}{\sqrt{2}}\left[\sqrt{1+\sin\vp_p}+
228: \sqrt{1-\sin\vp_p}\;(\vec{\alpha}\hat{\vec{p}})\right]u(0),\\
229: v(-\vec{p})&=&\frac{1}{\sqrt{2}}\left[\sqrt{1+\sin\vp_p}-
230: \sqrt{1-\sin\vp_p}\;(\vec{\alpha}\hat{\vec{p}})\right]v(0),
231: \end{array}
232: \right.
233: \label{uandv}
234: \ee
235: \be
236: \hat{b}_{s}(\vec{p},t)=e^{iE_pt}\hat{b}_{s}(\vec{p},0),\quad
237: \hat{d}_{s}(-\vec{p},t)=e^{iE_pt}\hat{d}_{s}(-\vec{p},0),
238: \label{bandd}
239: \ee
240: where $E_p$ stands for the dispersive law of the dressed quarks;
241: $\alpha$ being the colour index, $\alpha=\overline{1,N_C}$. It is
242: convenient to define the chiral angle varying in the range
243: $-\frac{\pi}{2}<\vp_p\leqslant\frac{\pi}{2}$ with the boundary
244: conditions $\vp(0)=\frac{\pi}{2}$, $\vp(p\to\infty)\to 0$. Notice
245: that this definition is not unique and in some cases one is forced
246: to choose $\vp(0)=-\frac{\pi}{2}$, as we shall see.
247:
248: The normal ordered Hamiltonian (\ref{H}) can be split into three parts,
249: \be
250: \hat{H}=E_{\rm vac}+:\hat{H}_2:+:\hat{H}_4:,
251: \label{H3}
252: \ee
253: and the usual procedure is to demand the quadratic part
254: $:\hat{H}_2:$ to be diagonal or, equivalently, that the vacuum
255: energy $E_{\rm vac}$ should become a minimum. Then, the
256: corresponding mass-gap equation ensures the anomalous Bogoliubov
257: terms $\hat{b}^\dagger \hat{d}^\dagger-\hat{d}\hat{b}$ to be absent in $:\hat{H}_2:$,
258: \be
259: A_p\cos\vp_p-B_p\sin\vp_p=0,
260: \label{mge}
261: \ee
262: where
263: \be
264: A_p=m+\frac12C_F\int\frac{d^3k}{(2\pi)^3}V_0(\vec{p}-\vec{k})\sin\vp_k,
265: \label{A}
266: \ee
267: \be
268: B_p=p+\frac12C_F\int \frac{d^3k}{(2\pi)^3}\;
269: (\hat{\vec{p}}\hat{\vec{k}})V_0(\vec{p}-\vec{k})\cos\vp_k,
270: \label{B}
271: \ee
272: $C_F=\frac12(N_C-1/N_C)$ being the $SU(N_C)_C$ Casimir operator in
273: the fundamental representation. The large-$N_C$ limit implies that
274: the product $C_FV_0$ remains finite as $N_C\to\infty$, so that,
275: for the power-like form (\ref{potential}), an appropriate
276: rescaling of the potential strength $K_0$ is understood. In the
277: remainder of this paper we shall absorb the coefficient $C_F$ into
278: the definition of the potential, $C_FV_0(|\vec{x}|)\equiv
279: V(|\vec{x}|)$ by a trivial redefinition of $K_0$.
280:
281: As soon as the mass-gap equation is solved and a nontrivial chiral
282: angle is found, the Hamiltonian (\ref{H3}) takes a diagonal form,
283: \be
284: \hat{H}=E_{\rm vac}+\sum_\alpha\sum_{s=\uparrow,\downarrow}\int
285: \frac{d^3 p}{(2\pi)^3} E_p[\hat{b}^\dagger_{\alpha s}(\vec{p}) \hat{b}_{\alpha
286: s}(\vec{p})+\hat{d}^\dagger_{\alpha s}(-\vec{p}) \hat{d}_{\alpha
287: s}(-\vec{p})],
288: \label{H2diag}
289: \ee
290: and the contribution of the $:\hat{H}_4:$ part is suppressed as
291: $1/\sqrt{N_C}$. The dressed quark dispersive law becomes
292: \be
293: E_p=A_p\sin\vp_p+B_p\cos\vp_p,
294: \label{Ep}
295: \ee
296: and this completes the diagonalisation of the Hamiltonian in the
297: quark sector to order $1/\sqrt{N_C}$ (the BCS level).
298:
299: For further convenience we write out both the mass-gap equation
300: and the dressed quark dispersive law $E_p$ for the harmonic
301: oscillator potential, $V(|\vec{x}|)=K_0^3|\vec{x}|^2$
302: \cite{Orsay,Orsay2,Lisbon},
303: \be
304: p^3\sin\vp_p=\frac12K_0^3\left[p^2\vp''_p+2p\vp_p'+\sin2\vp_p\right]+
305: mp^2\cos\vp_p,
306: \label{diffmge}
307: \ee
308: \be
309: E_p=m\sin\vp_p+p\cos\vp_p-K_0^3\left[\frac{{\vp'_p}^2}{2}
310: +\frac{\cos^2\vp_p}{p^2}\right].
311: \label{Epharm}
312: \ee
313: In the remainder of this subsection, if not stated otherwise, we
314: consider the chiral limit $m=0$.
315:
316: We end this introductory section by stating two general properties
317: held by mass-gap equations, Eq.~(\ref{mge}). First of all,
318: notice that using the explicit form of the dressed quark
319: dispersive law $E_p$ (\ref{Epharm}), the mass-gap equation
320: (\ref{diffmge}) can be rewritten in the form of a
321: Schr{\"o}dinger-like equation with the zero eigenvalue,
322: \be
323: [-K_0^3\Delta_p+2E_p]\psi=0,
324: \label{mgho}
325: \ee
326: where $\psi\equiv\sin\vp_p$. The generic mass-gap equation (\ref{mge})
327: for an arbitrary power-like confining potential (\ref{potential})
328: admits the form similar to (\ref{mgho}) \cite{replica4}.
329:
330: The second general property concerns the asymptotic behaviour of
331: the solutions of the mass gap equation (the chiral angle) at large
332: momenta. Taking the general form of the mass-gap equation for the
333: power-like potential (\ref{potential}) \cite{replica4},
334: \be
335: p^3\sin\vp_p=K_0^{\alpha+1}\Gamma(\alpha+1)\sin\frac{\pi\alpha}{2}
336: \int_{-\infty}^{\infty}
337: \frac{dk}{2\pi}\left\{\frac{pk\sin[\vp_k-\vp_p]}{|p-k|^{\alpha+1}}+
338: \frac{\cos\vp_k\sin\vp_p}{(\alpha-1)|p-k|^{\alpha-1}}\right\},
339: \label{mg2}
340: \ee
341: and performing an expansion for $p\to\infty$, we get
342: \ftnote{1}{The case of the harmonic oscillator potential,
343: $\alpha=2$, has to be considered separately. As immediately
344: follows from Eq.~(\ref{mgho}) and the free limit of the quark
345: dispersive law, $E_p\mathop{\approx}\limits_{p\to\infty}p$, the
346: asymptotic behaviour of the chiral angle in this case is given by
347: the Airy function. Still the qualitative conclusion made below
348: holds true.}
349: \be
350: \left.\vp_p\right|_{m=0}\mathop{\approx}\limits_{p\to\infty}-
351: \frac{\pi}{N_C}\Gamma(\alpha+2)K_0^{\alpha+1}\sin\frac{\pi\alpha}{2}
352: \frac{\langle\bar{q}q\rangle}{p^{\alpha+4}}.
353: \label{asym}
354: \ee
355: From the behaviour (\ref{asym}) we see that, in the chiral limit,
356: the sign of the solution $\vp_p$ at large momenta is opposite to
357: the sign of the chiral condensate. Beyond the chiral limit the
358: expression (\ref{asym}) defines how the chiral angle approaches
359: the free large-momentum asymptote $\arctan\frac{m}{p}$. From the
360: Gell-Mann-Oakes-Renner formula \cite{GMOR} one can relate the
361: chiral condensate with the pion mass $M_\pi$,
362: \be
363: f_\pi^2M_\pi^2=-2m\langle\bar{q}q\rangle,
364: \label{GMOR}
365: \ee
366: so that, according to Eq.~(\ref{asym}), $M_\pi^2$ becomes positive
367: if the chiral angle approaches its large-momentum asymptote from
368: above, and negative if it approaches from below. In the latter
369: case the pion becomes a tachyon and this situation requires a
370: special treatment which will be discussed in detail in chapters II
371: and III.
372:
373: \subsection{Diagonalisation in the mesonic sector and the bound-state
374: equation}
375:
376: \subsubsection{Quark Hamiltonian in the space of quark-antiquark pairs}
377:
378: In this subsection, we go beyond BCS level and include the quartic
379: part $:\hat{H}_4:$ of the Hamiltonian (\ref{H3}) into
380: consideration.
381:
382: After diagonalisation at BCS level, we could have used the formal
383: invariance, under arbitrary Bogoliubov-Valatin transformations
384: (BV), of the quark field $\psi_{\alpha}(\vec{x},t)$ --- here
385: understood as an inner product between a Hilbert space, spanned
386: by the spinors $u_s(\vec{p})$ and $v_s(-\vec{p})$, and a Fock
387: space, spanned by the quark annihilation and creation operators.
388: This invariance requires that any given BV rotation in the Fock
389: space, must engender a corresponding counter-rotation in the
390: associated spinorial Hilbert space so as to maintain the inner
391: product $\psi_{\alpha}(\vec{x},t)$ invariant. Thence, the new
392: BV-rotated spinors will carry the information on the chiral angle
393: and, therefore, so does the quark propagators. Equipped with these
394: new Feynman rules, we can then proceed to evaluate a wealth of
395: physical processes among which the Bethe-Salpeter equations for
396: bound states play a central role. Formally, the Hamiltonian of
397: Eq.~(\ref{H}), when written in terms of the quark fields
398: $\psi_{\alpha}(\vec{x},t)$, is invariant under such BV
399: transformations and so it will contain --- through its $:\hat{H}_4:$
400: term --- Bogoliubov anomalous terms of various types.
401:
402: {\em It will turn out that the set of all these Bethe-Salpeter
403: equations, each one of them associated with a corresponding
404: mesonic bound state, with their right mixture of positive with
405: negative-energy amplitudes, are exactly what is required in order
406: to get rid all the Bogoliubov anomalous terms with just four quark
407: creation(annihilation) operators. The remaining Bogoliubov terms
408: (responsible for generic hadronic decays like $A\to B+C$) will be
409: suppressed as $1/\sqrt{N_C}$}. This scenario is quite plausible
410: once we realise that a group of four-quark creation(annihilation)
411: operators --- two quarks and two antiquarks --- corresponds to a
412: bilinear creation (annihilation) of a group of two mesons and, as
413: we know, these bilinear anomalous terms can be made to disappear
414: by the use of an appropriated bosonic Bogoliubov-Valatin canonical
415: transformation.
416:
417: For the implementation of the diagonalisation program we shall
418: follow the method of refs.~\cite{2d} generalising it to the
419: four-dimensional case. The basic idea of the method is the
420: following. At BCS level, we consider quarks as though they were
421: free, with confinement to be implemented at later stages, when
422: building the Bethe--Salpeter
423: equation for bound states. Instead we now start with the {\em ab-initio}
424: confinement requirement that only quark-antiquark pairs are
425: allowed to be created and annihilated, not single quarks, and that
426: the full Hamiltonian (\ref{H3}) must take a diagonal form in terms
427: of the operators creating/annihilating whole quark-antiquark
428: mesons. Therefore, every time a quark (antiquark) is created or
429: annihilated, an accompanying antiquark (quark) must be
430: created/annihilated as well. To write this {\em ab-initio} condition,
431: let us introduce a set of four colourless operators: two operators
432: which count the number of quarks and the number of antiquarks,
433: \be
434: \begin{array}{c}
435: \hat{B}_{ss'}(\vec{p},\vec{p}')=\frac{\ds
436: 1}{\ds\sqrt{N_C}}\sum_\alpha \hat{b}_{\alpha
437: s}^\dagger(\vec{p})\hat{b}_{\alpha s'}(\vec{p}'),\quad
438: \hat{D}_{ss'}(\vec{p},\vec{p}')=\frac{\ds
439: 1}{\ds\sqrt{N_C}}\sum_\alpha \hat{d}_{\alpha
440: s}^\dagger(-\vec{p})\hat{d}_{\alpha s'}(-\vec{p}'),
441: \end{array}
442: \label{operatorsBD}
443: \ee
444: and two operators which create and annihilate a quark-antiquark pair,
445: \be
446: \begin{array}{c}
447: \hat{M}^\dagger_{ss'}(\vec{p},\vec{p}')=\frac{\ds
448: 1}{\ds\sqrt{N_C}}\sum_\alpha \hat{b}^\dagger_{\alpha
449: s'}(\vec{p}')\hat{d}^\dagger_{\alpha s}(-\vec{p}),\quad
450: \hat{M}_{ss'}(\vec{p},\vec{p}')=\frac{\ds
451: 1}{\ds\sqrt{N_C}}\sum_\alpha \hat{d}_{\alpha s}(-\vec{p})\hat{b}_{\alpha
452: s'}(\vec{p}').
453: \end{array}
454: \label{operatorsMM}
455: \ee
456: The coefficients in the definitions
457: (\ref{operatorsBD}) and (\ref{operatorsMM}) are fixed in such a
458: way that, in the large-$N_C$ limit, all four operators obey the
459: standard bosonic algebra, with the only nonzero commutator being
460: \be
461: [\hat{M}_{ss'}(\vec{p},\vec{p}')\;\hat{M}_{\sigma\sigma'}^\dagger(\vec{q},\vec{q}')]=
462: (2\pi)^3\delta^{(3)}(\vec{p}-\vec{q})
463: (2\pi)^3\delta^{(3)}(\vec{p}'-\vec{q}')\delta_{s\sigma}\delta_{s'\sigma'}.
464: \label{MMcom}
465: \ee
466:
467: The pair (\ref{operatorsBD}) is sufficient to express the leading
468: part of the Hamiltonian (\ref{H2diag}),
469: \be
470: \hat{H}=E_{\rm
471: vac}+\sqrt{N_C}\sum_{s=\uparrow,\downarrow}\int\frac{d^3
472: p}{(2\pi)^3}E_p[\hat{B}_{ss}(\vec{p},\vec{p})+\hat{D}_{ss}(\vec{p},\vec{p})],
473: \label{H2diag2}
474: \ee
475: whereas the suppressed part $:\hat{H}_4:$ requires all four
476: operators.
477:
478: According to our {\em ab-initio} confinement requirement, the operators
479: $\hat{B}$ and $\hat{D}$ cannot be independent and have to be
480: expressed through the operators $\hat{M}^\dagger$ and $\hat{M}$.
481: One easily builds such relations to be \cite{2d}
482: \be
483: \left\{
484: \begin{array}{c}
485: \hat{B}_{ss'}(\vec{p},\vec{p}')=\frac{\ds
486: 1}{\ds\sqrt{N_C}}\ds\sum_{s''} \int\frac{\ds d^3p''}{\ds
487: (2\pi)^3}\hat{M}_{s''s}^\dagger(\vec{p}'',\vec{p})
488: \hat{M}_{s''s'}(\vec{p}'',\vec{p}')\\
489: \hat{D}_{ss'}(\vec{p},\vec{p}')=\frac{\ds
490: 1}{\ds\sqrt{N_C}}\ds\sum_{s''} \int\frac{\ds d^3p''}{\ds
491: (2\pi)^3}\hat{M}_{ss''}^\dagger(\vec{p},\vec{p}'')
492: \hat{M}_{s's''}(\vec{p}',\vec{p}'').
493: \end{array}
494: \right.
495: \label{anzatz}
496: \ee
497:
498: The anzatz (\ref{anzatz}) is the most crucial point of the method.
499: Formally, one can also view it as another solution to the set of
500: equations given by the commutators of the operators
501: (\ref{operatorsBD}) and (\ref{operatorsMM}). Notice that the
502: relations (\ref{anzatz}) correspond to the \lq\lq minimal
503: substitution", that is, each quark is accompanied by only one
504: antiquark, and {\em vice versa}, whereas, in the general case, a
505: whole quark-antiquark cloud should be also created together with
506: the accompanying particle. Since the creation of each next
507: quark-antiquark pair in the cloud is suppressed by an extra power
508: of $1/N_C$, then Eq.~(\ref{anzatz}) represents just the leading
509: terms of the full expressions.
510:
511: After substitution of the anzatz (\ref{anzatz}) into the
512: Hamiltonian (\ref{H3}) we have no longer a general suppression of
513: the quartic part of the Hamiltonian as compared to the quadratic
514: part. Only a number of terms in $:\hat{H}_4:$ remain suppressed
515: and will be omitted henceforth. The resulting Hamiltonian, in the
516: space of colourless quark-antiquark pairs, takes the form:
517: \be
518: \hat{H}=E_{\rm vac}'+\int \frac{d^3P}{(2\pi)^3}
519: \hat{\cal H}(\vec{P}),
520: \label{HH1}
521: \ee
522: where we have separated out the quark-antiquark cloud
523: centre-of-mass motion.
524:
525: For the sake of simplicity, let us consider the Hamiltonian
526: density $\cal H$ at rest ($\vec{P}=0$),
527: $$
528: \hat{\cal H}\equiv\hat{\cal H}(\vec{P}=0)=\sum_{s_1s_2}\int\frac{d^3
529: p}{(2\pi)^3}
530: 2E_p\hat{M}_{s_1s_2}^\dagger(\vec{p},\vec{p})\hat{M}_{s_2s_1}(\vec{p},\vec{p})
531: +\frac12\sum_{s_1s_2s_3s_4}\int\frac{d^3p}{(2\pi)^3}\frac{d^3q}{(2\pi)^3}
532: V(\vec{p}-\vec{q})
533: $$
534: \be
535: \times \left\{[v^{++}(\vec{p},\vec{q})]_{s_1s_3s_4s_2}
536: \hat{M}^\dagger_{s_2s_1}(\vec{p},\vec{p})\hat{M}_{s_4s_3}(\vec{q},\vec{q})\right.
537: +[v^{+-}(\vec{p},\vec{q})]_{s_1s_3s_4s_2}
538: \hat{M}^\dagger_{s_2s_1}(\vec{q},\vec{q})\hat{M}^\dagger_{s_3s_4}(\vec{p},\vec{p})
539: \label{HH2}
540: \ee
541: $$
542: \left.+[v^{-+}(\vec{p},\vec{q})]_{s_1s_3s_4s_2}
543: \hat{M}_{s_1s_2}(\vec{p},\vec{p})\hat{M}_{s_4s_3}(\vec{q},\vec{q})
544: +[v^{--}(\vec{p},\vec{q})]_{s_1s_3s_4s_2}
545: \hat{M}_{s_3s_4}(\vec{p},\vec{p})\hat{M}_{s_1s_2}^\dagger(\vec{q},\vec{q})\right\},
546: $$
547: with amplitudes $v$ given by
548: \be
549: \begin{array}{c}
550: [v^{++}(\vec{p},\vec{q})]_{s_1s_3s_4s_2}=
551: [u^\dagger_{s_1}(\vec{p})u_{s_3}(\vec{q})]
552: [v^\dagger_{s_4}(-\vec{q})v_{s_2}(-\vec{p})],\\[0cm]
553: [v^{+-}(\vec{p},\vec{q})]_{s_1s_3s_4s_2}=
554: [u^\dagger_{s_1}(\vec{p})v_{s_3}(-\vec{q})]
555: [u^\dagger_{s_4}(\vec{q})v_{s_2}(-\vec{p})],\\[0cm]
556: [v^{-+}(\vec{p},\vec{q})]_{s_1s_3s_4s_2}=
557: [v^\dagger_{s_1}(-\vec{p})u_{s_3}(\vec{q})]
558: [v^\dagger_{s_4}(-\vec{q})u_{s_2}(\vec{p})],\\[0cm]
559: [v^{--}(\vec{p},\vec{q})]_{s_1s_3s_4s_2}=
560: [v^\dagger_{s_1}(-\vec{q})v_{s_3}(-\vec{p})]
561: [u^\dagger_{s_4}(\vec{p})u_{s_2}(\vec{q})].
562: \end{array}
563: \label{ampls}
564: \ee
565:
566: Although there are only two independent amplitudes in
567: (\ref{ampls}), for example, $v^{++}$ and $v^{+-}$, with the two
568: remaining amplitudes, $v^{--}$ and $v^{-+}$, easily related to the
569: first two by Hermitian conjugation and renaming of spin indices,
570: we prefer to keep all four amplitudes so as to give a more
571: symmetric form to both bound-state equations and the effective
572: diagrammatic rules to be derived shortly.
573:
574: Notice that in general, the vacuum energy in the Hamiltonian
575: (\ref{HH1}), $E_{\rm vac}'$, differs from the vacuum energy
576: $E_{\rm vac}$ in the Hamiltonian (\ref{H2diag}) by terms coming
577: from the commutators of the operators $\hat{M}$ and
578: $\hat{M}^\dagger$.
579:
580: \subsubsection{The case of the chiral pion}
581:
582: Before we proceed with the diagonalisation of the Hamiltonian
583: (\ref{HH2}) in general form, let us consider the case of the
584: chiral pion as a paradigm for the general case without unnecessary
585: complications to cloud physics. The spin and spatial structure of
586: the pion
587: --- a $^1S_0$ state with $J=L=S=0$
588: --- can be chosen in matrix form so that the operator
589: $\hat{M}_{ss'}(\vec{p},\vec{p})$, creating the pion at rest, is
590: represented as
591: \be
592: \hat{M}_{ss'}(\vec{p},\vec{p})=\left[\kappa(\hat{\vec{p}})\right]_{ss'}\hat{M}(p),\quad
593: \kappa(\hat{\vec{p}})=\frac{i}{\sqrt{2}}\sigma_2Y_{00}(\hat{\vec{p}}).
594: \label{po}
595: \ee
596: Substitution of Eq.~(\ref{po}) into the Hamiltonian (\ref{HH2})
597: gives a simpler expression in terms of the radial operators
598: $\hat{M}(p)$ and $\hat{M}^\dagger(p)$,
599: $$
600: \hat{\cal H}_\pi=\int\frac{p^2d
601: p}{(2\pi)^3}2E_p\hat{M}^\dagger(p)\hat{M}(p)-\frac12\int\frac{p^2
602: dp}{(2\pi)^3}\frac{q^2dq}{(2\pi)^3}
603: \left\{T^{++}_\pi(p,q)\hat{M}^\dagger(p)M(q)\right.
604: $$
605: \be
606: \left.+T^{+-}_\pi(p,q)\hat{M}^\dagger(q)\hat{M}^\dagger(p)
607: +T^{-+}_\pi(p,q)\hat{M}(p)\hat{M}(q)+T^{--}_\pi(p,q)
608: \hat{M}^\dagger(q)\hat{M}(p)\right\},
609: \label{HHpi}
610: \ee
611: where the amplitudes $T_\pi^{\pm\pm}(p,q)$ are trivially
612: built from the amplitudes $v^{\pm\pm}(\vec{p},\vec{q})$ (see
613: Eq.~(\ref{ampls})), after inclusion of the potential
614: $V(\vec{p}-\vec{q})$, integration over angular variables, and
615: taking a spin trace against the pion spin wave function $\kappa$,
616: as it was defined in Eq.~(\ref{po}).
617:
618: To have a fully diagonalised Hamiltonian $\hat{\cal H}_\pi$ it is
619: now sufficient to perform a Bogoliubov transformation,
620: \be
621: \left\{
622: \begin{array}{l}
623: \hat{M}(p)=\hat{m}_\pi\vp_\pi^+(p)+\hat{m}_\pi^\dagger\vp_\pi^-(p)\\
624: \hat{M}^\dagger(p)=\hat{m}^\dagger_\pi\vp_\pi^+(p)+\hat{m}_\pi\vp_\pi^-(p),
625: \end{array}
626: \right.
627: \label{Mmpi}
628: \ee
629: or, inversely,
630: \ftnote{2}{Although Eqs.~(\ref{mMpi}) seem not to follow from
631: Eqs.~(\ref{Mmpi}) directly, which is a consequence of the naive
632: truncation we have made leaving only the pion instead of the whole
633: tower of mesonic states, in fact, Eqs.~(\ref{Mmpi}) and
634: (\ref{mMpi}) are self-consistent with each other at the given
635: level of truncation. See Eqs.~(\ref{Mmgen}) and (\ref{mMgen})
636: below.}
637: \be
638: \left\{
639: \begin{array}{l}
640: \hat{m}_\pi=\ds\int\frac{p^2dp}{(2\pi)^3}\left[\hat{M}(p)\vp_\pi^+(p)-
641: \hat{M}^\dagger(p)\vp_\pi^-(p)\right]\\[2mm]
642: \hat{m}_\pi^\dagger=\ds\int\frac{p^2dp}{(2\pi)^3}\left[\hat{M}^\dagger
643: (p)\vp_\pi^+(p)- \hat{M}(p)\vp_\pi^-(p)\right],
644: \end{array}
645: \right.
646: \label{mMpi}
647: \ee
648: where the operators $\hat{m}_\pi^\dagger$ and $\hat{m}_\pi$ create and annihilate pions at rest.
649:
650: Using the commutator of the radial operators,
651: \be
652: [\hat{M}(p),\;\hat{M}^\dagger
653: (q)]=\frac{(2\pi)^3}{p^2}\delta(p-q),
654: \ee
655: following from Eq.~(\ref{MMcom}), we find that
656: \be
657: [\hat{m}_\pi ,\; \hat{m}_\pi^\dagger]=
658: \int\frac{p^2dp}{(2\pi)^3}\left[\vp_\pi^{+2}(p)-\vp_\pi^{-2}(p)\right].
659: \ee
660: Therefore, requiring the canonical commutator between the pion
661: creation and annihilation operators to be $[m_\pi,m_\pi^\dagger]=1$,
662: ensures that the Bogoliubov amplitudes
663: $\vp_\pi^{\pm}$, playing the role of the pion wave functions, must
664: in turn be subject to the normalisation
665: \ftnote{3}{Notice that the wave functions $\vp^\pm_\pi$ can be
666: chosen real. This holds for all observable mesons as well. If the
667: mesonic wave function still contains an imaginary part then
668: Eqs.~(\ref{Mmpi})-(\ref{norm1}) should be changed accordingly, so
669: that, for example, $\vp_\pi^{\pm 2}$ in the normalisation
670: condition (\ref{norm1}) should be changed for $|\vp_\pi^{\pm}|^2$.
671: Below we discuss the situation with the tachyon where one is
672: forced to deal with an imaginary mass and, as a consequence, with
673: its imaginary wave functions.}
674: \be
675: \int\frac{p^2dp}{(2\pi)^3}\left[\vp_\pi^{+2}(p)-\vp_\pi^{-2}(p)\right]=1.
676: \label{norm1}
677: \ee
678:
679: To complete the program of bosonic diagonalisation of the
680: Hamiltonian (\ref{H}), we need only to look for a pair of
681: $\vp_\pi^\pm (p)$ to diagonalise the hadronic Fock subspace
682: spanned by zero momentum pions, $|\pi_{P=0}\rangle =
683: m^\dagger_\pi|\Omega\rangle$, $|\Omega\rangle$ being the mesonic
684: vacuum (see the discussion below).
685:
686: Now, if $\vp_\pi^+(p)$ and $\vp_\pi^-(p)$ are made to obey the following
687: set of equations,
688: \be
689: \left\{
690: \begin{array}{l}
691: [2E_p-M_\pi]\vp_\pi^+(p)=\ds\int\frac{\ds q^2dq}{\ds (2\pi)^3}
692: [T^{++}_\pi(p,q)\vp_\pi^+(q)+T^{+-}_\pi(p,q)\vp_\pi^-(q)]\\[0cm]
693: [2E_p+M_\pi]\vp_\pi^-(p)=\ds\int\frac{\ds q^2dq}{\ds (2\pi)^3}
694: [T^{-+}_\pi(p,q)\vp_\pi^+(q)+T^{--}_\pi(p,q)\vp_\pi^-(q)],
695: \end{array}
696: \right.
697: \label{bsp}
698: \ee
699: we can, by a simple substitution in Eq.~(\ref{HHpi}), get rid of
700: the anomalous Bogoliubov terms to have $\langle\Omega|\hat{\cal
701: H}_\pi|\pi_{P=0}\pi_{P=0}\rangle =0$ and $\langle
702: \pi_{P=0}\pi_{P=0}|\hat{\cal H}_\pi|\Omega\rangle=0$. The
703: non-anomalous terms can be readily summed to yield:
704: \be
705: \hat{\cal H}_\pi=M_\pi \hat{m}_\pi^\dagger
706: \hat{m}_\pi+\ldots,\quad \langle \pi_{P=0}|\hat{\cal
707: H}_\pi|\pi_{P=0}\rangle = M_\pi,
708: \label{diagtot}
709: \ee
710: where the ellipsis denotes terms suppressed by $N_C$. For example,
711: the quartic term in the pion creation/annihilation operators
712: responsible for the pion-pion scattering can be identified among
713: the latter.
714:
715: Notice the difference between the quark BCS vacuum $|0\rangle$,
716: annihilated by the dressed quark operators $\hat{b}$ and
717: $\hat{d}$, and the mesonic vacuum $|\Omega\rangle$, annihilated by
718: the mesonic operators, for example, by $\hat{m}_\pi$. The two
719: vacua are related by a unitary transformation,
720: $|0\rangle=U^\dagger|\Omega\rangle$, with the operator $U^\dagger$ creating pairs
721: of mesons and taking such a form that, for example,
722: $$
723: \hat{m}_\pi|\Omega\rangle=\hat{m}_\pi U^\dagger|0\rangle
724: =U^\dagger(U\hat{m}_\pi U^\dagger)|0\rangle \propto
725: U^\dagger\hat{M}(p)|0\rangle=0.
726: $$
727: Since creation of any extra quark-antiquark pair is suppressed in
728: the large-$N_C$ limit, so is the deviation of the operator $U^\dagger$
729: from unity. As a result, the chiral condensate calculated at
730: BCS level, using the BCS vacuum $|0\rangle$, coincides, up to
731: $1/N_C$ corrections, with the full quark condensate calculated in
732: the mesonic vacuum $|\Omega\rangle$.
733:
734: It is easy to recognise in Eq.~(\ref{bsp}) the Bethe--Salpeter equation for the
735: pion studied in a number papers
736: \cite{Orsay2,Lisbon}. To see this equivalence we proceed in five steps.
737:
738: {\noindent Step 1:}
739:
740: First let us define dressed quarks and derive the mass-gap
741: equation. To this end we consider the quark mass operator in the
742: rainbow approximation,
743: \be
744: i\Sigma(\vec{p})=\int\frac{d^4k}{(2\pi)^4}V(\vec{p}-\vec{k})\Gamma
745: S_F(\vec{k},k_0)\Gamma,
746: \label{Sigma01}
747: \ee
748: where $S_F({\vec p},p_0)$,
749: \be
750: S_F({\vec p},p_0)=\frac{\Lambda^{+}({\vec
751: p})\gamma_0}{p_0-E_p+i\epsilon}+ \frac{{\Lambda^{-}}({\vec
752: p})\gamma_0}{p_0+E_p-i\epsilon},
753: \label{Feynman}
754: \ee
755: $$
756: \Lambda^\pm(\vec{p})=\frac12[1\pm\gamma_0\sin\vp_p\pm(\vec{\alpha}\hat{\vec{p}})\cos\vp_p],
757: $$
758: is the Feynman propagator of the dressed quark with the dispersive
759: law $E_p$. $\Gamma$ is a generic Dirac matrix defining the Lorentz
760: nature of the confining interaction. Parameterising
761: $\Sigma(\vec{p})$ and $E_p$ as
762: \be
763: \Sigma(\vec{p})=[A_p-m]+(\vec{\gamma}\hat{\vec{p}})[B_p-p],\quad
764: E_p=A_p\sin\vp_p+B_p\cos\vp_p,
765: \ee
766: we find that
767: \be
768: S_F^{-1}(p_0,\vec{p})=\gamma_0p_0-(\vec{\gamma}\hat{\vec{p}})B_p-A_p.
769: \ee
770: Taking the integral in the energy in Eq.~(\ref{Sigma01}), we
771: arrive at the self-consistency condition,
772: \be
773: A_p\cos\vp_p-B_p\sin\vp_p=0,
774: \label{mge2}
775: \ee
776: which is the mass-gap equation for the chiral angle $\vp_p$. For
777: $\Gamma=\gamma_0$ the auxiliary functions $A_p$ and $B_p$ are
778: given by Eqs.~(\ref{A}) and (\ref{B}). Notice, however, that the
779: general form of Eq.~(\ref{mge2}) holds for any Lorentz
780: nature of confinement (see \cite{Lisbon} for a detailed
781: discussion). This completes the matching, at BCS level,
782: between the Hamiltonian and diagrammatic approaches to the theory.
783:
784: {\noindent Step 2:}
785:
786: Proceed beyond BCS level and consider the Salpeter equation
787: (in the Dirac representation) for a generic meson in the rest
788: frame,
789: \be
790: \chi({\vec p};M)=-i\int\frac{d^4q}{(2\pi)^4}V(\vec{p}-\vec{q})\;
791: \Gamma S_F({\vec q},q_0+M/2)\chi({\vec q};M)S_F({\vec q},q_0-M/2)\Gamma,
792: \label{GenericSal}
793: \ee
794: where $\chi({\vec p};M)$ stands for the mesonic Salpeter amplitude.
795: \bigskip
796:
797: {\noindent Step 3:}
798:
799: Due to the instantaneous nature of $V(\vec{p}-\vec{q})$, it is a
800: simple matter of poles logistics to see that, upon integration in
801: the energy, the only surviving combinations of the Feynman
802: projectors $\Lambda^{\pm}$ entering in Eq.~(\ref{GenericSal}) are
803: $(\Lambda^+\gamma_0)\chi(\Lambda^-\gamma_0)$ and
804: $(\Lambda^-\gamma_0)\chi(\Lambda^+\gamma_0)$, so that we can
805: decompose $\chi$ in two distinct amplitudes, $\chi^{[+]}$ and
806: $\chi^{[-]}$. To this end we get rid of the energy denominators in
807: Eq.~(\ref{GenericSal}) by taking the integral in the energy,
808: \be
809: \int_{-\infty}^{\infty}\frac{dq_0}{2\pi i}\left[\frac{1}{q_0\pm
810: M/2-E_q+i\epsilon}\right] \left[\frac{1}{q_0\mp
811: M/2+E_q-i\epsilon}\right]=-\frac{1}{2E_q\mp M},
812: \ee
813: and define:
814: $$
815: \chi^{[+]}(\vec{q};M)=\frac{\chi(\vec{q};M)}{2E_q-M},\quad
816: \chi^{[-]}(\vec{q};M)=\frac{\chi(\vec{q};M)}{2E_q+M}.
817: $$
818:
819: Then the Bethe--Salpeter equation
820: (\ref{GenericSal}) amounts to a system of two coupled equations:
821: \be
822: \left\{
823: \begin{array}{l}
824: [2E_p-M]\chi^{[+]}=-\ds\int\frac{d^3q}{(2\pi)^3}V(\vec{p}-\vec{q})\;
825: \Gamma\left[(\Lambda^+\gamma_0)\chi^{[+]}(\Lambda^-\gamma_0)
826: +(\Lambda^-\gamma_0)\chi^{[-]}(\Lambda^+\gamma_0)\right]\Gamma\\[0mm]
827: [2E_p+M]\chi^{[-]}=-\ds\int\frac{d^3q}{(2\pi)^3}V(\vec{p}-\vec{q})\;
828: \Gamma\left[(\Lambda^+\gamma_0)\chi^{[+]}(\Lambda^-\gamma_0)
829: +(\Lambda^-\gamma_0)\chi^{[-]}(\Lambda^+\gamma_0)\right]\Gamma.
830: \end{array}
831: \label{Salpeterlev3}
832: \right.
833: \ee
834: \bigskip
835:
836: {\noindent Step 4:}
837:
838: Sandwich both equations in (\ref{Salpeterlev3}) between the quark
839: spinors and use the definition of the projectors,
840: \be
841: \Lambda^+(\vec{p})=\sum_{s_1s_2}u_{s_1}({\vec p})\otimes
842: u^\dagger_{s_2} ({\vec p}),\quad
843: \Lambda^-(\vec{p})=\sum_{s_1s_2}v_{s_1}(-{\vec p})\otimes
844: v^\dagger_{s_2} (-{\vec p}),
845: \label{Lambdas}
846: \ee
847: to cast Eq.~(\ref{Salpeterlev3}) as
848: \be
849: \left\{
850: \begin{array}{r}
851: [2E_p-M]\left[\bar{u}_{s_1}\chi^{[+]}v_{s_2}\right] =-\ds\sum_{s_3s_4}\int\frac{d^3q}{(2\pi)^3}V(\vec{p}-\vec{q})
852: \left\{[\bar{u}_{s_1}\Gamma u_{s_3}][\bar{u}_{s_3}\chi^{[+]}v_{s_4}]
853: [\bar{v}_{s_4}\Gamma v_{s_2}]\right.\\
854: \left.+[\bar{u}_{s_1}\Gamma v_{s_3}]
855: [\bar{v}_{s_3}\chi^{[-]}u_{s_4}][\bar{u}_{s_4}\Gamma v_{s_2}]\right\}\hphantom{.}\\[3mm]
856: [2E_p+M]\left[\bar{v}_{s_1}\chi^{[-]}u_{s_2}\right] =-\ds\sum_{s_3s_4}\int\frac{d^3q}{(2\pi)^3}V(\vec{p}-\vec{q})
857: \left\{[\bar{v}_{s_1}\Gamma u_{s_3}][\bar{u}_{s_3}\chi^{[+]}v_{s_4}]
858: [\bar{v}_{s_4}\Gamma u_{s_2}]\right.\\
859: \left.+[\bar{v}_{s_1}\Gamma v_{s_3}]
860: [\bar{v}_{s_3}\chi^{[-]}u_{s_4}][\bar{u}_{s_4}\Gamma u_{s_2}]\right\}.
861: \label{Salpeterlev4}
862: \end{array}
863: \right.
864: \ee
865: \bigskip
866:
867: {\noindent Step 5:}
868:
869: Define $\Phi_{s_1s_2}^+=[\bar{u}_{s_1}\chi^{[+]}v_{s_2}]$ and,
870: similarly, $\Phi_{s_1s_2}^-=[\bar{v}_{s_1}\chi^{[-]}u_{s_2}]$ to get:
871: \be
872: \left\{
873: \begin{array}{l}
874: [2E_p-M]\Phi_{s_1s_2}^+=-\ds\sum_{s_3s_4}\int\frac{d^3q}{(2\pi)^3}V(\vec{p}-\vec{q})
875: \left\{[v^{++}]_{s_1s_2s_3s_4}\Phi_{s_3s_4}^+ +[v^{+-}]_{s_1s_2s_3s_4}\Phi_{s_3s_4}^-\right\}\\[1mm]
876: [2E_p+M]\Phi_{s_1s_2}^-=-\ds\sum_{s_3s_4}\int\frac{d^3q}{(2\pi)^3}V(\vec{p}-\vec{q})
877: \left\{[v^{-+}]_{s_1s_2s_3s_4}\Phi_{s_3s_4}^+ +[v^{--}]_{s_1s_2s_3s_4}\Phi_{s_3s_4}^-\right\}.
878: \label{Salpeterlev5}
879: \end{array}
880: \right.
881: \ee
882:
883: In the coefficients $v^{\pm\pm}$ one can easily recognise the amplitudes
884: which appeared in the Hamiltonian (\ref{HH2}). For the case
885: $\Gamma=\gamma_0$ they are given in Eq.~(\ref{ampls}).
886: Further details and explicit forms for the amplitudes (\ref{ampls}) in terms of the
887: chiral angle can be found in refs.~\cite{Lisbon}.
888:
889: For establishing graphical rules it is convenient to include the
890: potential into the definition of the amplitudes and to rewrite
891: them in the form:
892: \be
893: \begin{array}{l}
894: [T^{++}(\vec{p},\vec{q})]_{s_1s_2s_3s_4}=[\bar{u}_{s_1}(\vec{p})\Gamma u_{s_3}(\vec{q})]
895: [-V(\vec{p}-\vec{q})][\bar{v}_{s_4}(-\vec{q})\Gamma v_{s_2}(-\vec{p})],\\[0cm]
896: [T^{+-}(\vec{p},\vec{q})]_{s_1s_2s_3s_4}=[\bar{u}_{s_1}(\vec{p})\Gamma v_{s_3}(-\vec{q})]
897: [-V(\vec{p}-\vec{q})][\bar{u}_{s_4}(\vec{q})\Gamma v_{s_2}(-\vec{p})],\\[0cm]
898: [T^{-+}(\vec{p},\vec{q})]_{s_1s_2s_3s_4}=[\bar{v}_{s_1}(-\vec{p})\Gamma u_{s_3}(\vec{q})]
899: [-V(\vec{p}-\vec{q})][\bar{v}_{s_4}(-\vec{q})\Gamma v_{s_2}(\vec{p})],\\[0cm]
900: [T^{--}(\vec{p},\vec{q})]_{s_1s_2s_3s_4}=[\bar{v}_{s_1}(-\vec{q})\Gamma v_{s_3}(-\vec{p})]
901: [-V(\vec{p}-\vec{q})][\bar{u}_{s_4}(\vec{p})\Gamma u_{s_2}(\vec{q})],
902: \end{array}
903: \label{ampls2}
904: \ee
905: or simply,
906: \be
907: \begin{array}{c}
908: T^{++}=[\bar{u}\Gamma u][-V][\bar{v}\Gamma v],\quad T^{+-}=[\bar{u}\Gamma v][-V][\bar{u}\Gamma v],\\
909: T^{-+}=[\bar{v}\Gamma u][-V][\bar{v}\Gamma u],\quad T^{--}=[\bar{v}\Gamma v][-V][\bar{u}\Gamma u].
910: \end{array}
911: \ee
912:
913: In Fig.~1 we depict the steps leading to Eq.~(\ref{Salpeterlev5})
914: with the amplitudes (\ref{ampls2}). In this figure, the first and
915: the second lines represent the r.h.s. of the first equation in the
916: system (\ref{Salpeterlev3}), sandwiched between the quark spinors
917: $\bar{u}$ and $v$, and the r.h.s. of the first equation in the
918: system (\ref{Salpeterlev5}), respectively. Similar diagrams can be
919: drawn for the second equation of both systems --- the only
920: difference will be the ordering of the spinors $u$ and $v$.
921: Thus we introduce a diagrammatic technique relating Bethe--Salpeter
922: amplitudes and vertices to the coefficients of the Hamiltonian in the
923: representation of mesonic operators.
924:
925: \begin{figure}[t]
926: \begin{center}
927: \includegraphics[width=14.5 cm]{messalp.eps}
928: \end{center}
929: \caption{Going from the Dirac representation to the spin
930: representation: one uses the definitions of the projectors $\Lambda^{\pm}$,
931: Eq.~(\ref{Lambdas}),
932: (dash-dotted line) to reshuffle the spinors towards the $\Gamma$ vertices
933: (solid line). The
934: Salpeter amplitudes $\chi^{[\pm]}$ are vectorially multiplied
935: by the adjacent spinors in order to construct two functions --- two-by-two matrices in spins: $\Phi^+$ and
936: $\Phi^-$ (fat solid line).}
937: \end{figure}
938:
939: Now we need to project the spin-angular wave functions
940: $\Phi_{s_1s_2}^\pm(\vec{p})$ onto the wave function with the
941: definite total momentum, parity, and charge conjugation number. For
942: the pion this is quite easily done. We have
943: \be
944: \Phi_{s_1s_2}^{\pm}(\vec{p})=\left[\frac{i}{\sqrt{2}}\sigma_2\right]_{s_1s_2}
945: Y_{00}(\hat{\vec{p}})\vp_\pi^\pm(p),\quad
946: Y_{00}(\hat{\vec{p}})=\frac{1}{\sqrt{4\pi}},
947: \label{vppi}
948: \ee
949: which should be compared with Eq.~(\ref{po}). Finally, we need
950: only to perform a spin trace and to introduce the $T_\pi^{\pm\pm}$
951: amplitudes for the pion, such as,
952: \be
953: T_\pi^{++}(p,q)=-\frac{1}{2}\int d\Omega_p
954: d\Omega_qY_{00}^*(\hat{\vec{p}})V(\vec{p}-\vec{q})
955: Y_{00}(\hat{\vec{q}})Sp\left\{\sigma_2[\bar{u}(\vec{p})\Gamma
956: u(\vec{q})]\sigma_2 [\bar{v}(-\vec{q})\Gamma v(-\vec{p})]\right\},
957: \label{Tpp}
958: \ee
959: to arrive at the pion Bethe--Salpeter equation in the spin
960: representation, which coincides with the bosonic mass-gap equation
961: (\ref{bsp}). For the case of the harmonic oscillator potential it
962: can be rewritten in a matrix form,
963: \be
964: \left[ \left[-K_0^3\frac{d^2}{dp^2}+2E_p\right]\left[
965: \begin{array}{cc}
966: 1&0\\0&1
967: \end{array}
968: \right]
969: +K_0^3
970: \left[
971: \frac{\vp^{'2}_p}{2}+\frac{\cos^2\vp_p}{p^2}\right]\left[
972: \begin{array}{cc}1&1\\1&1\end{array}
973: \right]
974: -
975: M_\pi\left[
976: \begin{array}{cc}1&0\\0&-1\end{array}
977: \right]
978: \right]
979: \left[
980: \begin{array}{c}
981: \nu_\pi^+(p)\\
982: \nu_\pi^-(p)
983: \end{array}
984: \right]=0, \label{hop}
985: \ee
986: for the radial wave functions $\nu^\pm_\pi(p)=p\vp^\pm_\pi(p)$
987: which are rescaled so as to obey the one-dimensional normalisation
988: \cite{Lisbon},
989: \be
990: \int dp[\nu^{+2}_\pi(p)-\nu^{-2}_\pi(p)]=1.
991: \ee
992: Comparison of the Hamiltonian (\ref{HHpi}) with the bound-state equation
993: (\ref{Salpeterlev5}) allows us to establish a dictionary between the terms in
994: (\ref{HHpi}) and the diagrammatic rules used in the literature in order
995: to arrive at the Bethe--Salpeter equation (\ref{Salpeterlev5}). These rules are
996: easily deduced from Fig.~1 and will be exploited in what follows in order to
997: diagonalise the full Hamiltonian (\ref{HH2}).
998:
999: Notice several important properties of the bound-state equation
1000: (\ref{bsp}) and its solutions. In the chiral limit the pion mass
1001: vanishes, $M_\pi=0$, and the pion wave functions are real and are
1002: related to one another and to the chiral angle as
1003: $\vp_\pi^+(p)=-\vp_\pi^-(p)={\cal N}_\pi\sin\vp_p$, ${\cal N}_\pi$
1004: being the norm. Let us use the simple case of the harmonic
1005: oscillator kernel to illustrate this. Using the explicit form of
1006: the $T$-amplitudes taken from the papers \cite{Lisbon}, one easily
1007: finds that, for the case of the $^1S_0$ state,
1008: $T^{++}_\pi(p,q)-T^{+-}_\pi(p,q)=T^{-+}_\pi(p,q)-T^{--}_\pi(p,q)=
1009: \frac{(2\pi)^3}{p^2}\delta(p-q)K_0^3\Delta_{q}$, where the
1010: Laplacian is actually reduced to its radial part (see
1011: Eq.~(\ref{hop})). Substituting this result in the bound-state
1012: equation (\ref{bsp}) for $\vp_\pi^+(p)=-\vp_\pi^-(p)\equiv\vp_\pi$,
1013: and using $M_\pi=0$, we get
1014: \be
1015: [-K_0^3\Delta_p+2E_p]\vp_\pi=0,
1016: \ee
1017: which coincides with the mass-gap equation (\ref{mgho}) with
1018: $\vp_\pi=\psi=\sin\vp_p$. Derivation of the pion bound-state
1019: equation as a Shr{\" o}dinger-like equation for the generic form
1020: of the potential is given in Appendix A.
1021:
1022: Thus we arrive at a very important conclusion that, in the chiral
1023: limit, {\em the mass-gap equation is the bound-state equation for
1024: the Goldstone boson at rest, responsible for the spontaneous
1025: breaking of chiral symmetry. Since the chiral pion is this
1026: Goldstone boson, then we are forced to see it already at the BCS
1027: level and it will remain being the massless Goldstone boson even
1028: when we go beyond BCS.} The properties of the pion have been
1029: widely discussed in the literature--- see, for example,
1030: \cite{Orsay,Orsay2,Lisbon} for QCD in 3+1 and \cite{tHooft,2d} for
1031: QCD in 1+1.
1032:
1033: Beyond the chiral limit the solution to the bound-state equation (\ref{bsp})
1034: takes an approximate form:
1035: \be
1036: \vp_\pi^\pm(p)=\tilde{\cal N}_\pi\left[\pm\frac{1}{\sqrt{M_\pi}}\sin\vp_p+
1037: \sqrt{M_\pi}\Delta_p\right],\quad
1038: \tilde{\cal N}_\pi^2=4\int_0^\infty\frac{p^2dp}{(2\pi)^3}\Delta_p\sin\vp_p,
1039: \label{vppm}
1040: \ee
1041: where all corrections of higher order in the pion mass are neglected and the function
1042: $\Delta_p$ obeys a reduced $M_\pi$--independent equation
1043: (see, for example, \cite{Lisbon} or the papers \cite{2d} where such an
1044: equation for $\Delta_p$ is discussed in
1045: two-dimensional QCD).
1046:
1047: Notice that, although the dressed quark dispersive law and $T$-amplitudes for the pion are real
1048: (see Appendix A for the details),
1049: the bound-state equation (\ref{bsp}) admits a tachyonic solution with $M_{\rm tach}^2<0$
1050: (see also the
1051: papers \cite{Orsay2} where a tachyonic solution was found for the bound-state equation
1052: in the trivial
1053: vacuum). For $M_{\rm tach}=i|M_{\rm tach}|$, the tachyon wave functions fail to be real and
1054: acquire imaginary parts, in accordance with Eq.~(\ref{vppm}). It is easy to verify that they
1055: obey a simple transformation rule, $(\vp^\pm_{\rm tach})^*\propto \vp^\mp_{\rm tach}$.
1056:
1057: \subsubsection{The case of the $\rho$-meson}
1058:
1059: Compared to the pion, the case of the $\rho$-meson involves an
1060: extra complication, a consequence of its richer spatial structure.
1061: The Hamiltonian (\ref{HH2}) does not commute with the operators of
1062: the total spin $\vec{S}$ and the angular momentum $\vec{L}$, so it
1063: cannot be diagonalised in terms of the $^{2S+1}L_J$ states. The
1064: appropriate basis is given by the set of physical observable
1065: mesons $J^{PC}$. Then, although the lowest $0^{-+}$ state --- the
1066: chiral pion --- is represented by just only one term, a pure
1067: $^1S_0$ state, higher states are in general a mixture of angular
1068: momentum states, like, for instance, the case of $1^{--}$
1069: $\rho$-meson, which is a mixture of $^3S_1$ and $^3D_1$ terms.
1070: This problem does not appear in two-dimensional QCD and is a
1071: consequence of a richer structure of the spatial dimensions in the
1072: four-dimensional theory. Therefore, in order to formulate the
1073: eigenvalue problem for the $\rho$-meson, one should introduce a
1074: set of four wave functions,
1075: $\left\{\vp^{\pm}_{^3S_1}\equiv\vp^\pm_0\right.$,
1076: $\left.\vp^{\pm}_{^3D_1}\equiv\vp^\pm_2\right\}$,
1077: and write the system of four coupled equation,
1078: \be
1079: \left\{
1080: \begin{array}{l}
1081: [2E_p-M_\rho]\vp_0^+(p)=\ds\int\frac{\ds q^2dq}{\ds (2\pi)^3}
1082: [T^{++}_{00}\vp_0^+(q)+T^{++}_{02}\vp_2^+(q)+T^{+-}_{00}\vp_0^-(q)+T^{+-}_{
1083: 02}\vp_2^-(q)]\\[0cm]
1084: [2E_p+M_\rho]\vp_0^-(p)=\ds\int\frac{\ds q^2dq}{\ds (2\pi)^3}
1085: [T^{-+}_{00}\vp_0^+(q)+T^{-+}_{02}\vp_2^+(q)+T^{--}_{00}\vp_0^-(q)+T^{--}_{
1086: 02}\vp_2^-(q)]\\[0cm]
1087: [2E_p-M_\rho]\vp_2^+(p)=\ds\int\frac{\ds q^2dq}{\ds (2\pi)^3}
1088: [T^{++}_{22}\vp_2^+(q)+T^{++}_{20}\vp_0^+(q)+T^{+-}_{22}\vp_2^-(q)+T^{+-}_{
1089: 20}\vp_0^-(q)]\\[0cm]
1090: [2E_p+M_\rho]\vp_2^-(p)=\ds\int\frac{\ds q^2dq}{\ds (2\pi)^3}
1091: [T^{-+}_{22}\vp_2^+(q)+T^{-+}_{20}\vp_2^+(q)+T^{--}_{22}\vp_2^-(q)+T^{--}_{
1092: 20}\vp_2^-(q)],
1093: \end{array}
1094: \right.
1095: \label{bsrho}
1096: \ee
1097: where the $T$-amplitudes can be built explicitly
1098: using the graphical rules established above and with the help of
1099: the spin-angular wave functions
1100: $[\kappa_{^3S_1}(\hat{\vec{p}})]_{s_1s_2}\equiv[\kappa_0(\hat{\vec{p}})]_{s_1s_2}$
1101: and
1102: $[\kappa_{^3D_1}(\hat{\vec{p}})[_{s_1s_2}\equiv[\kappa_2(\hat{\vec{p}})]_{s_1s_2}$.
1103: For example, similarly to Eq.~(\ref{Tpp}) we have,
1104: \be
1105: T_{02}^{++}(p,q)=-\int d\Omega_p d\Omega_qV(\vec{p}-\vec{q})
1106: Sp\left\{\kappa_0(\hat{\vec{p}})[\bar{u}(\vec{p})\Gamma
1107: u(\vec{q})]\kappa_2(\hat{\vec{q}}) [\bar{v}(-\vec{q})\Gamma
1108: v(-\vec{p})]\right\},
1109: \label{Tpp2}
1110: \ee
1111: and so on.
1112:
1113: As an example, we
1114: give here the explicit form of the Bethe--Salpeter equation (\ref{bsrho})
1115: for the harmonic oscillator potential \cite{Lisbon}:
1116: $$
1117: \left[
1118: \left[-\frac{d^2}{dp^2}+2E_p\right]
1119: \left[
1120: \begin{array}{cc}1&0\\0&1\end{array}
1121: \right]
1122: +
1123: \frac{6}{p^2}\left[\begin{array}{cc}1&0\\0&0\end{array}\right]
1124: +
1125: \frac{\vp^{'2}_p}{2}\left[\begin{array}{cc}3-\sigma_1&
1126: -2\sqrt{2}\\-2\sqrt{2}&3+\sigma_1\end{array}\right]\right.
1127: +
1128: \frac{2(1-\sin\vp_p)}{3p^2}\left[\begin{array}{cc}-4&
1129: \sqrt{2}\\
1130: \sqrt{2}&-4\end{array}\right]
1131: $$
1132: \be
1133: \left.
1134: -
1135: \frac{\cos^2\vp_p}{3p^2}\left[\begin{array}{cc}2(1-\sigma_1)&
1136: \sqrt{2}(1-\sigma_1)\\\sqrt{2}(1-\sigma_1)&1-\sigma_1\end{array}\right]
1137: -
1138: M_\rho\left[\begin{array}{cc}\sigma_3&0\\0&-\sigma_3\end{array}\right]
1139: \right]
1140: \left[
1141: \begin{array}{c}
1142: \nu_2^+(p)\\
1143: \nu_2^-(p)\\
1144: \nu_0^+(p)\\
1145: \nu_0^-(p)
1146: \end{array}
1147: \right]=0,
1148: \label{horho}
1149: \ee
1150: where $\sigma$'s are the Pauli matrices and, similarly to the case of
1151: the pion, the new radial wave functions are defined as
1152: $\nu^\pm_{0,2}(p)=p\vp^\pm_{0,2}(p)$.
1153:
1154: Naively, one may conclude, from the bound-state equation
1155: (\ref{bsrho}), that the $\rho$-meson has to be described by a
1156: four-component wave function, rather than by just a two-component
1157: wave function, as it was the case with the pion. This conclusion
1158: is erroneous, since the doubling of the number of equations in
1159: (\ref{bsrho}) is a consequence of the usage of an inappropriate
1160: basis $^{2S+1}L_J$. Indeed, together with the light $\rho$-meson,
1161: we must have a second solution for the the bound-state equation
1162: (\ref{bsrho}) needed to describe the heavier vectorial partner of
1163: the $\rho$. The easiest way to show this is to neglect
1164: off-diagonal terms in Eq.~(\ref{bsrho}), thus splitting the system
1165: of four equation into two independent systems of two equations
1166: each, for $\vp_0^\pm$ and $\vp_2^\pm$, respectively. The
1167: eigenvalues found for these two independent eigenvalue problems
1168: give the masses of the pure $^3S_1$ and $^3D_1$ states, which
1169: should be mixed then as
1170: \be
1171: {\rm det}\left(
1172: \begin{array}{cc}
1173: M_{^3S_1}-M_\rho&\Delta_{SD}\\
1174: \Delta_{DS}&M_{^3D_1}-M_\rho
1175: \end{array}
1176: \right)=0,
1177: \label{MSD}
1178: \ee
1179: with $\Delta_{SD}$ and $\Delta_{DS}$
1180: being the contributions of the restored off-diagonal terms of the
1181: full Hamiltonian (\ref{HH2}). The lighter solution represents the
1182: physical $\rho$-meson in the basis $J^{PC}$.
1183: In other words, if an appropriate basis --- which
1184: diagonalises the three-dimensional spatial part of the Hamiltonian
1185: (\ref{HH2})
1186: --- is chosen from the very beginning, each physical mesonic state
1187: is fully described by just a two-component wave function and the
1188: corresponding eigenvalue --- the physical mass of the meson.
1189: Notice that the transformation (\ref{MSD}) can be viewed as yet
1190: another Bogoliubov-Valatin transformation parameterised by the
1191: $S$-wave--$D$-wave mixing angle.
1192:
1193: Now, with the $\rho$-meson included, the diagonalised Hamiltonian
1194: takes an obvious form,
1195: \be
1196: \hat{\cal H}=M_\pi m_\pi^\dagger m_\pi+M_\rho m_\rho^\dagger
1197: m_\rho+ M_{\rho'} m_{\rho'}^\dagger m_{\rho'}\ldots,
1198: \ee
1199: where $\rho'$ denotes the heavier partner of $\rho$ which is also
1200: a solution of Eq.~(\ref{MSD}) and the operators
1201: creating/annihilating the vector mesons are defined similarly to
1202: the pionic case, Eqs.~(\ref{po}) and (\ref{Mmpi}), through the
1203: $1^{--}$ wave functions $\vp^\pm_{\rho,\rho'}$. We are now ready
1204: to consider the general case.
1205:
1206: \subsubsection{The general case}
1207:
1208: Now we return to the general case of the Hamiltonian (\ref{HH2})
1209: and diagonalise it in terms of the mesonic compound operators. We work
1210: in the basis $\{n,J^{PC}\}$; $n$ being the radial quantum number.
1211: We also assume that this basis includes all states with the given
1212: $J^{PC}$, so that the set of such states is complete and has a
1213: one-to-one correspondence with the complete set $^{2S+1}L_J$. For
1214: the sake of simplicity, we denote each mesonic state as
1215: $\{n,\nu\}$, where the set of quantum numbers $\nu$ is not only
1216: based on the $J^{PC}$ classification scheme, but also identifies
1217: each meson in a given $nJ^{PC}$ multiplet, as it was discussed in
1218: the case of $\rho$ and $\rho'$. Then, for a given pair
1219: $\{n,\nu\}$, representing a physical meson, we introduce a
1220: two-component radial wave function $\vp^\pm_{n\nu}(p)$ and the
1221: spin-angular wave function $\kappa_\nu(\hat{\vec{p}})$, as it was
1222: done in the pion case --- see Eq.~(\ref{po}). The operators
1223: $\hat{M}^\dagger$ and $\hat{M}$ of Eq.(\ref{operatorsMM}),
1224: entering the Hamiltonian (\ref{HH2}), can be expanded using this
1225: basis (Eq.~(\ref{po}) for the pion follows from this relation if
1226: we take only the lowest $\nu=0^{-+}$),
1227: \be
1228: \hat{M}_{ss'}(\vec{p},\vec{p})=
1229: \sum_\nu[\kappa_\nu(\hat{\vec{p}})]_{ss'}\hat{M}_\nu(p).
1230: \ee
1231: Now we use the completeness property of the set
1232: $\{\kappa_\nu(\hat{\vec{p}})\}$ together with the fact that this
1233: set diagonalises the spin and the angular structure of the
1234: Hamiltonian (\ref{HH2}), to find:
1235: \be
1236: \sum_{s_1s_2}\int
1237: d\Omega_p[\kappa_\nu(\hat{\vec{p}})]_{s_1s_2}^*
1238: [\kappa_{\nu'}(\hat{\vec{p}})]_{s_2s_1}=\delta_{\nu\nu'},
1239: \label{59}
1240: \ee
1241: and
1242: \be
1243: \begin{array}{c}
1244: \ds \sum_{s_1s_2s_3s_4}\int d\Omega_pd\Omega_qV(\vec{p}-\vec{q})
1245: [v^{++}(\vec{p},\vec{q})]_{s_1s_3s_4s_2}[\kappa_\nu(\hat{\vec{p}})]^*_{s_2s_1
1246: }[\kappa_{\nu'}(\hat{\vec{q}})]_{s_4s_3}=
1247: -T^{++}_\nu(p,q)\delta_{\nu\nu'},\\
1248: \ds \sum_{s_1s_2s_3s_4}\int d\Omega_pd\Omega_qV(\vec{p}-\vec{q})
1249: [v^{+-}(\vec{p},\vec{q})]_{s_1s_3s_4s_2}[\kappa_\nu(\hat{\vec{q}})]^*_{s_2s_1
1250: }[\kappa_{\nu'}(\hat{\vec{p}})]_{s_3s_4}^*=
1251: -T^{+-}_\nu(p,q)\delta_{\nu\nu'},\\
1252: \ds \sum_{s_1s_2s_3s_4}\int d\Omega_pd\Omega_qV(\vec{p}-\vec{q})
1253: [v^{-+}(\vec{p},\vec{q})]_{s_1s_3s_4s_2}[\kappa_\nu(\hat{\vec{p}})]_{s_1s_2}
1254: [\kappa_{\nu'}(\hat{\vec{q}})]_{s_4s_3}=
1255: -T^{-+}_\nu(p,q)\delta_{\nu\nu'},\\
1256: \ds \sum_{s_1s_2s_3s_4}\int d\Omega_pd\Omega_qV(\vec{p}-\vec{q})
1257: [v^{--}(\vec{p},\vec{q})]_{s_1s_3s_4s_2}[\kappa_\nu(\hat{\vec{q}})]_{s_1s_2}
1258: [\kappa_{\nu'}(\hat{\vec{p}})]^*_{s_3s_4}=
1259: -T^{--}_{\nu}(p,q)\delta_{\nu\nu'}.
1260: \end{array}
1261: \ee
1262:
1263: As a result, the Hamiltonian (\ref{HH2}) takes the form, in terms
1264: of the radial operators $\{\hat{M}_\nu^\dagger(p)\}$ and
1265: $\{\hat{M}_\nu(p)\}$ as:
1266: $$
1267: \hat{\cal H}=\sum_\nu\int\frac{p^2d p}{(2\pi)^3}2E_p\hat{M}_\nu^\dagger(p)\hat{M}_\nu(p)
1268: -\frac12\sum_\nu\int\frac{p^2dp}{(2\pi)^3}\frac{q^2dq}{(2\pi)^3}\left\{
1269: T_\nu^{++}(p,q)\hat{M}_\nu^\dagger(p)\hat{M}_\nu(q)\right.
1270: $$
1271: \be
1272: \left.+T^{+-}_\nu(p,q)\hat{M}_\nu^\dagger(q)\hat{M}_\nu^\dagger(p)
1273: +T^{-+}_\nu(p,q)\hat{M}_\nu(p)\hat{M}_\nu(q)+T^{--}_\nu(p,q)
1274: \hat{M}_\nu^\dagger(q)\hat{M}_\nu(p)\right\},
1275: \label{HHgen}
1276: \ee
1277: where all radial excitations are now disentangled from one another.
1278:
1279: Then the generalisation of the relations (\ref{Mmpi}),
1280: (\ref{mMpi}) is trivial,
1281: \be
1282: \left\{
1283: \begin{array}{l}
1284: \hat{M}_\nu(p)=\sum_n[\hat{m}_{n\nu}\vp_{n\nu}^+(p)+\hat{m}_{n\nu}^\dagger\vp_{n\nu}^-(p)]\\
1285: \hat{M}_\nu^\dagger(p)=\sum_n
1286: [\hat{m}^\dagger_{n\nu}\vp_{n\nu}^+(p)+\hat{m}_{n\nu}\vp_{n\nu}^-(p)],
1287: \end{array}
1288: \right.
1289: \label{Mmgen}
1290: \ee
1291: \be
1292: \left\{
1293: \begin{array}{l}
1294: \hat{m}_{n\nu}=\ds\int\frac{p^2dp}{(2\pi)^3}\left[\hat{M}_\nu(p)\vp_{n\nu}^+(p)-
1295: \hat{M}^\dagger_\nu(p)\vp_{n\nu}^-(p)\right]\\
1296: \hat{m}_{n\nu}^\dagger=\ds\int\frac{p^2dp}{(2\pi)^3}\left[\hat{M}_\nu^\dagger
1297: (p)\vp_{n\nu}^+(p)- \hat{M}_\nu(p)\vp_{n\nu}^-(p)\right].
1298: \end{array}
1299: \right.
1300: \label{mMgen}
1301: \ee
1302:
1303: As in the pion case, we build the commutators of the operators
1304: $\hat{m}_{n\nu}$ and $\hat{m}_{m\nu}^\dagger$,
1305: \be
1306: \begin{array}{l}
1307: [\hat{m}_{n\nu},\;\hat{m}_{m\nu}^\dagger]=\ds
1308: \int\frac{p^2dp}{(2\pi)^3}\left[\vp_{n\nu}^{+}(p)\vp_{m\nu}^{+}(p)-
1309: \vp_{n\nu}^{-}(p)\vp_{m\nu}^{-}(p)\right],\\[3mm]
1310: [\hat{m}_{n\nu},\;\hat{m}_{m\nu}]=\ds
1311: \int\frac{p^2dp}{(2\pi)^3}\left[\vp_{m\nu}^{+}(p)\vp_{n\nu}^{-}(p)-
1312: \vp_{m\nu}^{-}(p)\vp_{n\nu}^{+}(p)\right],
1313: \end{array}
1314: \ee
1315: and require that they should obey the standard bosonic algebra,
1316: that is, $[\hat{m}_{n\nu},\;\hat{m}_{m\nu}^\dagger]=\delta_{mn}$,
1317: and $[\hat{m}_{n\nu},\;\hat{m}_{m\nu}]=0$. Therefore, similarly to
1318: Eq.~(\ref{norm1}), for any given $\nu$, we arrive at the
1319: normalisation and orthogonality conditions for $\vp$'s,
1320: \ftnote{4}{Strictly speaking, the orthogonality condition for the
1321: mesonic wave functions should also include their spin-angular part. For example,
1322: $$
1323: \sum_{s_1s_2}\int d\Omega_p[\kappa_\nu(\hat{\vec{p}})]^*_{s_1s_2}
1324: [\kappa_{\nu'}(\hat{\vec{p}})]_{s_2s_1}
1325: \int\frac{p^2dp}{(2\pi)^3}\left[\vp_{n\nu}^{+}(p)\vp_{m\nu'}^{+}(p)-
1326: \vp_{n\nu}^{-}(p)\vp_{m\nu'}^{-}(p)\right]=\delta_{nm}\delta_{\nu\nu'},
1327: $$
1328: where, for non-coinciding $\nu$'s, the l.h.s. vanishes due to the angular
1329: integration --- see the orthogonality condition for $\kappa$'s,
1330: Eq.~(\ref{59}). Then, for $\nu=\nu'$, one readily arrives at the first equation
1331: in (\ref{normgen}).}
1332: \be
1333: \begin{array}{l}
1334: \ds\int\frac{p^2dp}{(2\pi)^3}\left[\vp_{n\nu}^{+}(p)\vp_{m\nu}^{+}(p)-
1335: \vp_{n\nu}^{-}(p)\vp_{m\nu}^{-}(p)\right]=\delta_{nm},\\[3mm]
1336: \ds\int\frac{p^2dp}{(2\pi)^3}\left[\vp_{n\nu}^{+}(p)\vp_{m\nu}^{-}(p)-
1337: \vp_{n\nu}^{-}(p)\vp_{m\nu}^{+}(p)\right]=0.
1338: \end{array}
1339: \label{normgen}
1340: \ee
1341:
1342: The representation (\ref{Mmgen}), together with the normalisation
1343: and orthogonality conditions (\ref{normgen}), give the fully
1344: diagonalised Hamiltonian,
1345: \be
1346: \hat{\cal H}=\sum_{n,\nu}M_{n\nu}m^\dagger_{n\nu}m_{n\nu}+
1347: O\left(\frac{1}{\sqrt{N_C}}\right), \label{hdgen} \ee provided the
1348: mesonic wave functions obey the bound-state equation, \be \left\{
1349: \begin{array}{l}
1350: [2E_p-M_{n\nu}]\vp_{n\nu}^+(p)=\ds\int\frac{\ds q^2dq}{\ds (2\pi)^3}
1351: [T^{++}_\nu(p,q)\vp_{n\nu}^+(q)+T^{+-}_\nu(p,q)\vp_{n\nu}^-(q)]\\[0cm]
1352: [2E_p+M_{n\nu}]\vp_{n\nu}^-(p)=\ds\int\frac{\ds q^2dq}{\ds (2\pi)^3}
1353: [T^{-+}_\nu(p,q)\vp_{n\nu}^+(q)+T^{--}_\nu(p,q)\vp_{n\nu}^-(q)].
1354: \end{array}
1355: \right.
1356: \label{bsgen}
1357: \ee
1358:
1359: On comparing the mass-gap-like equation for the generalised
1360: Bogoliubov transformation (\ref{bsgen}) with the Bethe--Salpeter
1361: equation (\ref{Salpeterlev5}), one can easily see that they
1362: coincide, the only difference being that they are written in
1363: different representations: Eq.~(\ref{Salpeterlev5}) is written in the
1364: quark spins representation, whereas Eq.~(\ref{bsgen}) is written
1365: in the $J^{PC}$ representation. Therefore, it is a matter of
1366: rewriting the Salpeter wave function $\Phi^\pm$ in the proper
1367: representation to go from one equation to the other. Notice that
1368: the Lorentz nature of the confining interaction --- the explicit
1369: form of the matrices $\Gamma$ in Eqs.~(\ref{Sigma01}) and
1370: (\ref{GenericSal}) --- plays no role, defining only the explicit
1371: form the amplitudes $v^{\pm\pm}$.
1372: Similarly, the Hamiltonian approach to the theory can be
1373: developed, following the same lines, for an arbitrary matrix
1374: $\Gamma$.
1375:
1376: In the leading order in $N_C$ the Hamiltonian (\ref{hdgen})
1377: describes free stable mesons. The suppressed terms in
1378: (\ref{hdgen}) must involve quark exchange and correspond to the
1379: parts of the Hamiltonian responsible for hadronic decays and
1380: scattering. For example, to consider a decay $A\to B+C$, one
1381: should restore the first sub-leading term in the Hamiltonian,
1382: $\frac{1}{\sqrt{N_C}}\hat{m}_A\;\hat{m}_B^\dagger\;
1383: \hat{m}_C^\dagger$, with the coefficient giving the amplitude of
1384: the corresponding process \cite{2d}. Notice, however, that all
1385: three mesons --- $A$, $B$, and $C$ --- cannot be at rest and one
1386: encounters a problem of boosting mesonic creation/annihilation
1387: operators, which is closely related to the general problem of
1388: Lorentz boosts in such potential models. This important issue lies
1389: beyond the scope of the present paper and deserves special
1390: treatment. Notwithstanding this general problem, we are still left
1391: with an important set of sub-leading physical processes which are
1392: amenable to exact treatment in the context of non-local NJL
1393: Hamiltonians: the set of elastic hadron-hadron scattering at rest
1394: (scattering lengths) where we can define a common vacuum for all
1395: intervening hadrons \cite{pipi}. We leave the issue of vacuum
1396: boosts to future publications.
1397:
1398: This completes the matching between the Hamiltonian and the
1399: Bethe--Salpeter approaches to the non-local NJL models described
1400: in Eq.(\ref{H}). We conclude that {\em diagonalisation of the
1401: Hamiltonian of the theory in the mesonic sector in the leading
1402: order in $N_C$ is equivalent to developing the Bethe--Salpeter
1403: approach to the theory in the ladder/rainbow approximation. In the
1404: Hamiltonian approach, the mass-gap and the bound-state equations
1405: emerge to ensure the cancellation, in the Hamiltonian, of the
1406: anomalous Bogoliubov terms {\bfseries both} in the quark and in the
1407: mesonic sectors of the theory. The normalisation condition for the
1408: mesonic wave functions $\vp^\pm$ with the minus sign between the
1409: positive- and the negative-energy components follows naturally in
1410: the Hamiltonian approach as the standard normalisation of the
1411: bosonic Bogoliubov amplitudes.}
1412:
1413: \subsection{Discussion}
1414:
1415: Let us make several concluding remarks concerning the
1416: diagonalisation procedure presented in this section. We started
1417: from the Hamiltonian of a quark model with the instantaneous
1418: interaction parameterised by the quark kernel of an arbitrary, but
1419: necessarily confining, form. As an example for such like models we
1420: have, for instance, QCD in the truncated Coulomb gauge. After
1421: inclusion of self-interaction into the quark fields we obtain
1422: dressed quarks --- the so-called BCS level. Chiral symmetry is
1423: spontaneously broken at this stage, quarks acquire effective mass
1424: and the chiral condensate appears in the vacuum. The next step was
1425: to perform yet another Bogoliubov-type transformation, so as to
1426: build operators which create/annihilate $nJ^{PC}$ mesonic states
1427: in the BCS vacuum. Finally, the rules relating the second
1428: Bogoliubov transformation (which diagonalised, up to
1429: $N_C$-suppressed terms, the quartic part of the Hamiltonian) and
1430: the Bethe--Salpeter equation for bound quark-antiquark states were
1431: deduced. Crucial points of the approach necessary to carry out
1432: this program were:
1433: \begin{itemize}
1434: \item instantaneous inter-quark interaction which allows one to
1435: avoid the problem of the relative inter-quark time and to
1436: formulate a self-consistent Hamiltonian approach to the theory;
1437: \item limit of a large number of colours with the inherent
1438: suppression of all non-planar diagrams (quark exchange);
1439: \item confinement, which allows one to reformulate the theory
1440: entirely in terms of colourless bound states
1441: --- mesons (the crucial conjecture (\ref{anzatz}) will fail for
1442: nonconfining interactions);
1443: \item a nontrivial solution to the mass-gap equation defining the
1444: broken phase of the theory. This phase is characterised by two
1445: concurrencial processes, one defined by the amplitude of creation
1446: of a quark-antiquark pair by means of the $\hat{b}^\dagger
1447: \hat{d}^\dagger$ operator applied to the
1448: vacuum ($\hat{M}^\dagger_{ss'}(\vec{p},\vec{p}')$), and a second
1449: process realised through the \lq\lq borrowing" of $\bar{q} q$
1450: pairs from the chiral condensate via the operator $\hat{d}\hat{b}$
1451: ($\hat{M}_{ss'}(\vec{p},\vec{p}')$). Both processes are
1452: fundamentally important for the chiral pion to become massless due
1453: to strong cancellations between the positive- and
1454: negative-energy components of the pionic wave function.
1455: \end{itemize}
1456:
1457: The discussed approach is completely insensitive to:
1458: \begin{itemize}
1459: \item the number of spatial dimensions, provided a suitable basis
1460: is built which diagonalises the spatial part of the Hamiltonian.
1461: This is due to the fact that both the mechanism for the separation
1462: of positive from negative-energy components of the mesonic wave
1463: functions and the mechanism for the disentanglement of radial
1464: excitations are universal mechanisms for instantaneous
1465: interactions;
1466: \item the Lorentz nature of confinement, which we considered to
1467: take the simplest form --- $\gamma_0\times\gamma_0$, but which can
1468: be easily generalised to be $\gamma_\mu\times\gamma_\nu$;
1469: \item the form of the potential, which is required only to be
1470: confining and to lead to a finite mass-gap equation. All
1471: power-like potentials (\ref{potential}) with
1472: $0\leqslant\alpha\leqslant 2$ meet these conditions. Among these
1473: we have the linearly rising potential, $\alpha=1$, --- as the most
1474: natural candidate for confinement in QCD, as well as the harmonic
1475: oscillator potential, $\alpha=2$, which is the easiest example for
1476: analytical and numerical studies.
1477: \end{itemize}
1478:
1479: Finally, we arrive at the following important conclusions:
1480: \begin{enumerate}
1481:
1482: \item the Bethe--Salpeter equation for bound states of quarks and
1483: antiquarks is equivalent to diagonalisation of the quartic term in
1484: the quark Hamiltonian, expressed in terms of the dressed quark
1485: fields. Notice that no new information, at least in the leading
1486: order in $N_C$, is required/appears beyond BCS level, and then
1487: we naturally arrive at the second conclusion that \item the entire
1488: problem is completely defined as soon as the mass-gap equation is
1489: formulated and solved, the chiral angle being the only entity
1490: which one needs to know in order to solve the entire theory. In
1491: particular, \item the second, bosonic, Bogoliubov transformation
1492: is defined via the Salpeter amplitudes $\vp^+$ and $\vp^-$. In the
1493: literature of boson condensation they are usually denoted by $u$
1494: and $v$ amplitudes respectively. Therefore it is not hard to
1495: understand that $\vp^\pm$ play simultaneously the role of the
1496: mesonic wave functions and satisfy the normalisation condition
1497: which is proper of the usual Bogoliubov self-consistency condition
1498: $u^2-v^2=1$. Thence these two amplitudes can be parameterised as
1499: $u=\cosh\chi$ and $v=\sinh\chi$, with $\chi$ being the mesonic
1500: Bogoliubov angle. Therefore, for the generalised mesonic
1501: transformation, $\cosh\chi$ and $\sinh\chi$ are given by the sets
1502: of the positive- and negative-energy components $\{\vp^+_{n\nu}\}$ and
1503: $\{\vp^-_{n\nu}\}$ which are completely defined by the chiral angle
1504: (for example, for the chiral pion we have
1505: $\vp_\pi^+=-\vp_\pi^-={\cal N}_\pi\sin\vp_p$).
1506: \end{enumerate}
1507:
1508: {\em In short, solving the mesonic Bethe-Salpeter equation is
1509: tantamount to finding the bosonic Bogoliubov transformation to
1510: fully diagonalise --- in the leading order in $N_c$ --- any non-local
1511: NJL Hamiltonian}.
1512:
1513: This ends the discussion of the diagonalisation of non-local NJL
1514: models. In the next section, we generalise the results of this
1515: section to the case of multiple vacuum states --- replicas.
1516:
1517: \section{Vacuum replicas in potential quark models}
1518:
1519: \subsection{Multiple solutions to the mass-gap equation}
1520:
1521: As noticed above, the mass-gap equation is equivalent in momentum
1522: space to a Schr{\" o}dinger-like equation, with the dressed quark
1523: dispersive law playing the role of the effective potential (see
1524: the example of Eq.~(\ref{mgho})). This suggests that the solution
1525: to the mass-gap equation might not be unique. Such multiple
1526: solutions were discovered for the harmonic oscillator potential in
1527: \cite{Orsay,Lisbon}. Recently a detailed analysis was performed
1528: for the general form of the power-like confining potential,
1529: Eq.~(\ref{potential}), and the existence of an infinite tower of
1530: solutions to the corresponding mass-gap equation was proved
1531: \cite{replica4}. Let us enumerate the main statements one can make
1532: concerning the mass-gap equation:
1533: \begin{itemize}
1534: \item in order to provide a nontrivial solution to the mass-gap
1535: equation the dressed quark dispersive law must become negative in
1536: the infrared region $p\to 0$. Since the chiral symmetry is broken
1537: only for such nontrivial chiral angles, then this property of
1538: $E_p$ is absolutely necessary for chiral symmetry breaking;
1539: \item once a nontrivial solution $\vp_p$ to the mass-gap equation
1540: is found, it defines the wave function of the pion --- the
1541: Goldstone boson --- as $\psi={\cal N}_\pi\sin\vp_p$;
1542: \item the slope of the chiral angle at the origin defines the
1543: scale of the chiral symmetry breaking for the given solution;
1544: steeper solutions correspond to less broken symmetry and to a
1545: sharper behaviour of $E_p$ at the origin; \item the mass-gap
1546: equation for any single-parameter confining potential leads to an
1547: infinite number of solutions for the chiral angle.
1548: \end{itemize}
1549:
1550: To exemplify the appearance of the replicas, let us draw the
1551: following qualitative picture. For the trivial chiral angle
1552: $\vp_p=0$ there is no dressing of quarks, the quark dispersive law
1553: being just $E_p=p$ for all momenta. With such $E_p$, the effective
1554: potential in the Schr{\" o}dinger equation (\ref{mgho}) (and
1555: similarly for other forms of the confining potential), $V_{\rm
1556: eff}(p)=2E_p(p)$, does not possess a single bound state with a
1557: zero eigenvalue --- therefore, no nontrivial solutions to the
1558: mass-gap equation do exist. We encounter here the well-known
1559: problem of the constituent quark models. Indeed, as it was
1560: discussed above, the mass-gap equation plays the role of the
1561: bound-state equation for the pion at rest. Upon substituting the
1562: free quark dispersive law, $E_p=p$, into the mass-gap
1563: equation --- see for example, Eq.~(\ref{mgho}) --- we are led to the
1564: usual Schr{\" o}dinger equation for the pion,
1565: $[2p+V(r)]\psi_\pi=M_\pi\psi_\pi$, which is characteristic of
1566: potential constituent quark models. The pion mass coming out of
1567: this equation appears to be of order of the confining interaction
1568: scale, $M_\pi\sim 400\div 500MeV$, which is several times larger
1569: than the experimental value of $140MeV$ and such $M_\pi$ {\em does
1570: not} vanish in the chiral limit. In the bound-state equation, this
1571: problem is known to be solved by the presence of the
1572: negative-energy component $\vp_\pi^-$ of the pionic wave function
1573: which happens to be of the same order of magnitude as the
1574: positive-energy component $\vp_\pi^+$. Strong cancellations
1575: between the two components of the pion wave function bring the
1576: pion mass to zero. At BCS level, the solution of this problem
1577: comes from the actual form of the dressed quarks dispersive law
1578: $E_p$ which becomes negative at small momenta, and cancels the
1579: positive contribution of the confining potential allowing for the
1580: pion mass to vanish. In other words: the peculiar behaviour of the
1581: dressed quark dispersive law is {\em both a necessity and a direct
1582: consequence of the chiral symmetry breaking}. Suppose now that we
1583: can parameterise this small-$p$ negative contribution to $E_p$ by
1584: a mass scale $\mu$. For the power-like potentials
1585: (\ref{potential}), it is clear that such a contribution has to be
1586: proportional to $K_0^{1+\alpha}$ (see the definition of $E_p$
1587: through the auxiliary functions $A_p$ and $B_p$, Eq.~(\ref{Ep}),
1588: as well as Eqs.~(\ref{A}) and (\ref{B})). For dimensional reasons,
1589: $E_p(p=0)=-{\rm const \frac{K_0^{1+\alpha}}{\mu^\alpha}}$ (for
1590: example, for $\alpha=2$, from Eq.~(\ref{Epharm}) with $m=0$ and
1591: $\vp_p\mathop{\approx}\limits_{p\to 0}\frac{\pi}{2}-\frac{p}{\mu}$
1592: one readily finds that $E_p(p=0)=-\frac32\frac{K_0^3}{\mu^2}$).
1593: Thus, for a sufficiently small $\mu =\mu_0$ the effective
1594: potential $V_{\rm eff}(p|\mu)$ becomes binding enough to produce
1595: an eigenstate with a zero eigenvalue. Then $\mu_0$ defines the scale
1596: of the chiral symmetry breaking in the new vacuum and the
1597: corresponding chiral angle behaves as
1598: $\vp_0(p)\mathop{\approx}\limits_{p\to
1599: 0}\frac{\pi}{2}-\frac{p}{\mu_0}+\ldots$. This is the ground-state
1600: solution which corresponds to the BCS vacuum of the theory.
1601: Decreasing the scale $\mu$, one can reach a situation where a
1602: second zero eigenvalue bound state appears for the potential
1603: $V_{\rm eff}(p|\mu)$. According to general quantum mechanical
1604: theorems, this solution must contain one knot. This is the first
1605: replica vacuum. Continuing this procedure, we can build the whole
1606: infinite tower of replicas for a given potential. In other words,
1607: for any given value of the scale $\mu$ one has a set of orthogonal
1608: eigenstates (for the potential $V_{\rm eff}(p|\mu)$) with positive
1609: and negative eigenvalues. To search for the $n$th replica, it is
1610: sufficient to vary $\mu$, thereby shifting the whole tower of
1611: eigenstates up or down until the appearance of the sought $n$th
1612: zero eigenvalue. Therefore, different replicas are eigenstates in
1613: different potentials.
1614:
1615: The easiest way to prove such a picture is to evaluate the
1616: quasi-classical Bohr-Sommerfeld integral for such an eigenvalue
1617: problem. Since we consider a one-scale confining interaction,
1618: then, from dimensional reasons, it is clear that the kinetic and
1619: the potential energies, in momentum space, behave as
1620: $K^{1+\alpha}r^\alpha$ and $K_0^{1+\alpha}/p^\alpha$,
1621: respectively. Consequently the corresponding WKB integral depends
1622: logarithmically on the scale $\mu$, which plays the role of the
1623: cut-off, $I_{\rm WKB}\propto \ln\frac{K_0}{\mu}$ \cite{replica4}.
1624: Therefore, the quasi-classical quantisation condition,
1625: $I_{WKB}=2\pi n$, can be fulfilled for any $n$, provided the
1626: corresponding scale $\mu_n$ is small enough. For example, for the
1627: harmonic oscillator potential in $D$ dimensions, the approximate
1628: dependence of the scale $\mu$ on the index of the replica
1629: $n$ can be easily found using the aforementioned WKB method to be,
1630: \be
1631: \mu_n={\rm const}\times K_0\exp{\left(-\frac{2\pi n}{\sqrt{D(D-2)}}\right)},
1632: \ee
1633: so that, in any dimension $D>2$, the number of replicas is,
1634: indeed, infinite. The boundary case of $D=2$ contains a
1635: singularity, for $n\neq 0$, so there are no replica solutions in
1636: the harmonic oscillator potential in two dimensions, as was found
1637: in \cite{replica1} and discussed in detail in \cite{replica5}.
1638:
1639: Of course, the procedure of building replicas drawn above should
1640: be understood as a simplified qualitative method, since the scale
1641: $\mu$ is not a free parameter but it has to be defined
1642: self-consistently with the form of the chiral angle. Notice also
1643: that, for confining potentials defined through more than one
1644: parameter, the proof given above does not hold, since the WKB
1645: integral may be regularised by the second scale, instead of the
1646: parameter $\mu$. Then the number of replicas becomes finite or
1647: they do not exist at all. Notice, however, that, in the general
1648: case, it is much harder not to have replicas than to have them,
1649: and the expectation of existence of more than one solution for
1650: nonlinear mass-gap equations is not only physically attractive but
1651: also quite natural from the mathematical point of view.
1652:
1653: \subsection{The problem of the tachyon}
1654:
1655: In this subsection, we address the problem already touched upon in
1656: the paper \cite{replica1} --- namely, the problem of the tachyon
1657: which appears in the excited vacua. Indeed, it was numerically
1658: found that, in the chiral limit, the chiral condensate changes the
1659: sign from replica to replica remaining negative for even states
1660: (the ground-state BCS vacuum being the lowest representative of
1661: the even states) and becomes positive for all odd states, starting
1662: with the first replica \cite{replica1,replica4}. The universal
1663: status of this rule becomes clear from the asymptotic behaviour of
1664: the chiral angle (\ref{asym}) and the fact that for each next
1665: replica the chiral angle possesses an extra knot and, therefore,
1666: it approaches, when $p\rightarrow \infty$, the $\vp_p=0$ asymptote
1667: from the opposite half-plane, as compared to the previous replica.
1668: Then, switching on a small quark mass, we arrive at the
1669: Gell-Mann--Oakes--Renner relation (\ref{GMOR}), in which the
1670: chiral condensate on the r.h.s. can be calculated in the limit
1671: $m=0$ \ftnote{5}{Beyond the chiral limit the quark condensate
1672: acquires an infinite contribution coming from the trace of the
1673: free-particle Green's function which should be subtracted. The
1674: term of the zeroth order in $m$ in the properly regularised
1675: condensate obviously coincides with the value calculated in the
1676: chiral limit.}, and, for positive values of
1677: $\langle\bar{q}q\rangle$, the pion becomes a tachyon, $M_\pi^2<0$.
1678: As always, the presence of a tachyon means that the lowest state
1679: --- the vacuum
1680: --- is chosen improperly, and there should be a more preferable vacuum.
1681: It is easy to see that, for odd replicas, this is indeed the case.
1682:
1683: To demonstrate this, let us consider the exact chiral limit and
1684: the most general Valatin-Bogoliubov transformation from the
1685: trivial vacuum $|0\rangle_0$ to the generalised $\theta$-vacuum
1686: $|\theta\rangle$ which can be written as,
1687: \be
1688: |\theta\rangle=S_\theta|0\rangle_0,\;\;
1689: S_\theta=\exp{[Q_\theta^\dagger-Q_\theta]},\;\; Q_\theta^\dagger
1690: =\frac12\sum_{\alpha}\sum_{ss'}\int \frac{d^3p}{(2\pi)^3}\vp_p
1691: \mathfrak{M}_{ss'}[\theta]\hat{b}_{\alpha s}^\dagger(\vec{p})\hat{d}_{\alpha
1692: s'}^\dagger(-\vec{p}),
1693: \label{gBV}
1694: \ee
1695: where $\vp_p$ is the chiral angle and
1696: \be
1697: \mathfrak{M}[\theta]=\mathfrak{M}_{^3P_0}\cos\theta
1698: +i\mathfrak{M}_{^1S_0}\sin\theta.
1699: \ee
1700:
1701: The $^3P_0$ matrix
1702: $\mathfrak{M}_{^3P_0}=(\vec{\sigma}\hat{\vec{p}})i\sigma_2$ is
1703: studied in detail in \cite{Lisbon} and the $^1S_0$ matrix
1704: $\mathfrak{M}_{^1S_0}\equiv \mathfrak{M}_\pi=-i\sigma_2$. For
1705: $\theta=0$ Eq.~(\ref{gBV}) reproduces the standard definition of
1706: the BCS vacuum $|0\rangle$ \cite{Lisbon}. When
1707: $\theta=\frac{\pi}{2}$ we obtain:
1708: \be
1709: \left|\frac{\pi}{2}\right\rangle=S_{\pi/2}|0\rangle_0=
1710: \exp{\left[i\int
1711: d^3x\bar\psi(x)\gamma_0\gamma_5\psi(x)\right]}|0\rangle_0=
1712: \exp{[i\hat{Q}_5]}|0\rangle_0,
1713: \label{q5q}
1714: \ee
1715: with the operator $\hat{Q}_5$ being responsible for translations
1716: along the $\theta$ direction. The angles $\theta$ and $\vp_p$ are independent
1717: quantities. For a given solution of the mass-gap equation,
1718: $\vp_p$, when $\theta=\pi$ we get $S_\pi[\vp_p]=S_0[-\vp_p]$,
1719: which, as discussed above, is the usual pseudo-unitary $^3P_0$
1720: operator for the chiral condensation. Thus, translation from $\theta=0$
1721: to $\theta=\pi$ along the $\theta$ direction is tantamount to a
1722: chiral angle transformation of $\vp_p\to-\vp_p$. We could have
1723: started with $-\vp_p$ --- also a solution to the mass-gap
1724: solution for $m_q=0$ --- to arrive, in the end of the
1725: $\theta$-journey, at $\vp_p$. Notice that, in general, the state
1726: $|\theta\rangle$ does not have any definite parity ---
1727: we have to consider mixtures of different $\theta$-vacua in order to
1728: build a state with a given definite parity. For instance,
1729: $|\theta\rangle+|-\theta\rangle$ has the parity plus. In the
1730: special case of $\theta=\pi$, $S_\pi=S_{-\pi}$ and the state
1731: $|\theta =\pi\rangle$, as well as $|\theta =0\rangle$, has positive parity.
1732:
1733: Let us start with a given $\vp_p$ and put, for a moment,
1734: $\theta=0$. If we now plot the vacuum energy as a function of the
1735: chiral condensate $\Sigma=\langle\bar{q}q\rangle$, then $\Sigma=0$
1736: corresponds to the local maximum, whereas the stable minimum is
1737: provided by a nonzero value of $\Sigma$, easily related to the
1738: corresponding chiral angle,
1739: \be
1740: \Sigma=-\frac{N_C}{\pi^2}\int^{\infty}_0 dp\;p^2\sin\vp_p.
1741: \label{Sigma1}
1742: \ee
1743: Next, we can move along the $\theta$-valley. Since, in the chiral
1744: limit, the charge $\hat{Q}_5$ commutes with the Hamiltonian of the
1745: theory, $[\hat{Q}_5\hat{H}]=0$, then the vacuum energy is
1746: degenerate for all $\theta$'s, so that we are free to choose any
1747: value for $\theta$. Such a form of the vacuum energy as a function
1748: of two variables, $\Sigma$ and $\theta$, is known as the Mexican
1749: hat. Incidentally, we can find, starting from expression
1750: (\ref{q5q}), the chiral pion Salpeter amplitude
1751: $\vp_\pi^+(p)=-\vp_\pi^-(p)$, which is just the c-number multiplying
1752: the anomalous part of $\hat{Q}_5$ \cite{emilioResina}. It is then
1753: clear that the pion Salpeter amplitude is given by $\sin\vp_p$,
1754: independently of any consideration for a given quark kernel,
1755: provided it supports the mechanism of spontaneous chiral symmetry
1756: breaking.
1757:
1758: \begin{figure}[t]
1759: \centerline{\epsfig{file=phi0.eps,width=8cm}
1760: \epsfig{file=phi1.eps,width=8cm}}
1761: \caption{The ground-state (the left plot) and the first replica
1762: (the right plot) solutions of the mass-gap equation (\ref{diffmge})
1763: after the redefinition of the chiral angle sign for odd replicas.
1764: The current quark mass is $m=0.01K_0$.
1765: The momentum $p$ is given in the units of $K_0$.}
1766: \end{figure}
1767:
1768: If now a small quark mass is introduced, then the vacuum energy
1769: acquires a contribution of the chirally non-invariant term
1770: proportional to the quark mass $m$ and, as a result, the whole
1771: picture gets tilted
1772: --- the state with the negative sign of the chiral condensate
1773: being energetically preferable as compared to the state with the
1774: positive condensate. These two states differ by their $\theta$
1775: coordinate --- one of them corresponds to $\theta=0$, whereas the
1776: other one has $\theta=\pi$. The choice of the chiral angle adopted
1777: before, with $\vp_p(0)=\frac{\pi}{2}$, selects the vacuum
1778: $\theta=0$ as the true vacuum of the theory $|0\rangle$. This is
1779: indeed the case for the ground BCS vacuum, as well as for all even
1780: replicas, with $\Sigma<0$. The pion, which is not tachyon in these
1781: vacua, is responsible for the vacuum energy increase during the
1782: rotation in the angle $\theta$ from $|0\rangle$ to $|\pi\rangle$.
1783: For odd replicas, the chiral condensate changes the sign, so that
1784: the state with $\theta=\pi$ must acquire a lower energy than the
1785: one with $\theta=0$. As a result, in odd replicas, the pion
1786: becomes the tachyon responsible for the energy decrease during the
1787: same rotation form $|0\rangle$ to $|\pi\rangle$, the latter
1788: representing now the true vacuum. Therefore, for odd replicas, we set
1789: $\vp_p(0)= -\frac{\pi}{2}$ instead of $\frac{\pi}{2}$. This
1790: choice will ensure that the chiral angle, when the momentum $p$ goes to
1791: infinity, will approach zero from above. Then the chiral
1792: condensate (\ref{Sigma1}) changes the sign, and so does the pion
1793: mass squared
1794: --- the pion mass becomes real. As discussed before, the
1795: definition of the chiral angle admits such a change and, as a
1796: result, we arrive at two different classes of solutions to the
1797: mass-gap equation: even solutions, which start from
1798: $\frac{\pi}{2}$ at $p=0$ and approach the free limit at large
1799: $p$'s from above, and odd solutions which also approach their
1800: large-$p$ asymptote from above but, at $p=0$, start from
1801: $-\frac{\pi}{2}$. This solves the problem of the tachyon for odd
1802: replicas, making the latter normal vacuum states with the
1803: possibility of building the spectrum of hadrons above them.
1804:
1805: As an example, we give in Fig.~2, the profiles of the
1806: ground-state and the first replica solutions to the mass-gap
1807: equation (\ref{diffmge}) for the oscillator-type potential (the
1808: case of the generalised power-like confining potential
1809: (\ref{potential}) is studied in detail in \cite{replica4}). As
1810: discussed before, the sign of the chiral angle for the first
1811: replica vacuum is reversed.
1812:
1813: In Fig.~3, we represent the solutions of the bound-state problem
1814: for the pion with the harmonic oscillator confining potential,
1815: Eq.~(\ref{hop}), for the two lowest solutions to the mass-gap
1816: equation (\ref{diffmge}) depicted in Fig.~2
1817: --- for the ground BCS vacuum (the left plot in Fig.~3) and for the first replica
1818: with the reversed sign of the chiral angle (the right plot in
1819: Fig.~3). It is clearly seen from Fig.~3 that the pionic wave
1820: functions are indeed very close to $\pm\sin\vp_p$ (see Fig.~2)
1821: with the corrections given by the formula (\ref{vppm}). Notice
1822: that, after the change $\vp_p\to-\vp_p$ performed for odd
1823: replicas, the \lq\lq$+$" and the \lq\lq$-$" components of the
1824: mesonic wave function substitute each other, so that one needs
1825: either to redefine the whole set of the operator transformations
1826: used to diagonalise the Hamiltonian of the theory in the mesonic
1827: sector or, which is more economical, simply to rename $\vp^+$ to
1828: $\vp^-$ and {\em vice versa}.
1829:
1830: \begin{figure}[t]
1831: \centerline{\epsfig{file=pion0.eps,width=7.5cm}\hspace*{10mm}\epsfig{file=pion1.eps,width=7.5cm}}
1832: \caption{Solutions of the bound-state problem for the pion, Eq.~(\ref{hop}), in the BCS vacuum
1833: (the left plot) and in the first replica with the reversed sign of the chiral angle
1834: (the right plot). The overall normalisation factor is omitted, the
1835: largest component being fixed equal to unity at $p=0$.
1836: The current quark mass is $m=0.01K_0$. The momentum $p$
1837: is given in the units of $K_0$.}
1838: \end{figure}
1839:
1840: \subsection{Hadronic spectrum in the replica vacua}
1841:
1842: It is concluded, in the second section, that any NJL-type
1843: Hamiltonian model with an arbitrary confining quark kernel is
1844: completely defined at BCS level, as soon as the chiral angle
1845: is built. If the large-$N_C$ limit is assumed, then all
1846: approximations become controllable and the Hamiltonian of the
1847: theory admits complete diagonalisation in terms of the compound
1848: mesonic operators. We found that the appropriate basis for such a
1849: diagonalisation was provided by the set $\{n,\;J^{PC}\}$ and each
1850: mesonic state was described by a pair of wave functions $\vp^\pm$.
1851: In this section, following the papers
1852: \cite{replica1,replica2,replica4}, we gave evidence of existence
1853: of excited solutions to the mass-gap equation --- the vacuum
1854: replicas. Now we address the question concerning the spectrum of
1855: hadrons in the replica vacua.
1856:
1857: As argued in the previous subsection, the pion is massless in the
1858: chiral limit for any replica vacuum and, with the properly defined
1859: chiral angle, it acquires a small real mass beyond the chiral
1860: limit. As far as highly excited mesonic states are concerned, the
1861: chiral symmetry is restored in this part of the spectrum and the
1862: hadrons' properties become insensitive to the details of the
1863: vacuum in which they are created since, for high excitations, the
1864: negative-energy components $\vp^-_{n\nu}(p)$ of the wave functions
1865: are negligible and the Bethe--Salpeter bound-state equation
1866: (\ref{bsgen}) amounts to a quantum-mechanical Schr{\" o}dinger equation,
1867: \be
1868: [2E_p+V(r)]\vp^+_{n\nu}(p)=M_{n\nu}\vp^+_{n\nu}(p),\quad
1869: E_p=\sqrt{p^2+m^2}.
1870: \label{Sal}
1871: \ee
1872: Consequently, no cancellations
1873: between the positive- and negative-energy components take place
1874: anymore, as it happens to the chiral pion and leads to the small
1875: pion mass of $140MeV$, as opposed the value of about $400\div
1876: 500MeV$ which would follow from Eq.~(\ref{Sal}). As a result, in
1877: the vacua with \lq\lq less" broken chiral symmetry (higher
1878: replicas), the masses of mesons are pushed up, with the exception
1879: of the chiral pion, whose mass, in the chiral limit, is kept equal
1880: to zero by the requirement of the chiral symmetry. Beyond the
1881: chiral limit, the pion mass also increases with the index
1882: of the replica (see Fig.~3 and compare the pion mass of $0.57K_0$,
1883: in the BCS vacuum, with the value of $0.988K_0$, for the first
1884: replica).
1885:
1886: Therefore, we conclude that each mesonic state can be
1887: characterised by an extra \lq\lq quantum number" --- the index
1888: of the vacuum replica in which it was created. Then the
1889: full diagonalised Hamiltonian of the theory takes the form:
1890: \be
1891: \hat{\cal H}=\sum_{n,\nu,{\cal N}}M_{n\nu{\cal
1892: N}}m^\dagger_{n\nu{\cal N}}m_{n\nu{\cal
1893: N}}+O\left(\frac{1}{\sqrt{N_C}}\right),
1894: \label{hdgen2}
1895: \ee
1896: ${\cal N}$ being the replica label, which is the generalisation
1897: of Eq.~(\ref{hdgen}) for the multi-vacuum case.
1898:
1899: \section{Conclusions}
1900:
1901: In this paper, we have shown how the generalised NJL-type
1902: Hamiltonian model with a nonlocal confining quark kernel can be
1903: fully diagonalised in the sector of physically observable hadronic
1904: states. We extend this result to include the multiple vacuum
1905: states recently found to exist in models of such a type and which
1906: are very likely to exist in real QCD. The main result of the paper
1907: is the fully diagonalised Hamiltonian (\ref{hdgen2}). Similarly to
1908: the case of a path integral with a multi-minimum function in the
1909: exponent, the Hamiltonian (\ref{hdgen2}) sums the contributions of
1910: all minima, with the corresponding weight --- the masses of mesons
1911: in the given vacuum --- being minimal for the lowest vacuum state
1912: and increasing for higher vacua, thus suppressing the contribution
1913: of the corresponding replicas to the full partition function of
1914: the theory. We study the form of the chiral angle profiles
1915: defining the excited vacua and find that for an even (odd) replica
1916: to decay to a lower even (odd) one it is sufficient to change the
1917: chiral angle in the finite region in momentum $p$, since both
1918: asymptotes for both replicas of the same parity coincide:
1919: $\vp_p(p=0)=\frac{\pi}{2}\;\left(-\frac{\pi}{2}\right)$,
1920: $\vp_p(p\to\infty)=0$. On the contrary, for the decay of an odd
1921: replica to an even one, or {\em vice versa}, the chiral angle has
1922: to be changed dramatically in the infrared region, since the
1923: low-momentum behaviour of even and odd replicas is quite different.
1924:
1925: In the paper \cite{replica2} a quantum field theory perturbative method was
1926: developed. It used the replica-quark-antiquark vertex
1927: parameterised by a small difference between the chiral angles of
1928: the ground-state vacuum and the replica to evaluate the quark self-energy
1929: shift due to the replica presence. This method can be easily
1930: extended to the case of an infinite number of replicas, but notice
1931: that the sets of even and odd states have to be considered
1932: separately. As discussed above, only transformations of the chiral
1933: angle between states with the same parity are small and can be
1934: considered perturbatively.
1935: We conclude, therefore, that two sets of perturbative
1936: approaches to replicas should be built --- for even and for odd replicas,
1937: independently.
1938: On the contrary, a transition between
1939: states with opposite parities, for example, the decay of the first
1940: replica to the ground BCS vacuum, involves the global
1941: transformation of the chiral angle generated by the pseudoscalar
1942: pionic operator $\hat{Q}_5$ and results in a burst of pions with
1943: a huge energy release. The same transition in the opposite
1944: direction --- excitation of the replica
1945: --- requires an external global source which is not
1946: inherent to the model. Building of such a source is an important
1947: problem in the theory of replicas and it will be the subject of
1948: future publications.
1949:
1950: \begin{acknowledgments}
1951: One of the authors, A. Nefediev, is grateful to P. Bicudo and Yu.
1952: S. Kalashnikova for many fruitful discussions and would like to
1953: thank the staff of the Centro de F\'\i sica das Interac\c c\~oes
1954: Fundamentais (CFIF-IST) for cordial hospitality during his stay in
1955: Lisbon, where this work was originated, and to acknowledge the
1956: financial support of the grant NS-1774.2003.2, as well as of the
1957: Federal Programme of the Russian Ministry of Industry, Science and
1958: Technology No 40.052.1.1.1112.
1959: \end{acknowledgments}
1960:
1961: \begin{thebibliography}{99}
1962: \bibitem{NJL} Y. Nambu, G. Jona-Lasinio, Phys. Rev. {\bf 122}, 345
1963: (1961).
1964: \bibitem{Orsay} A. Amer, A. Le Yaouanc, L. Oliver, O. Pene, and J.-C.
1965: Raynal, Phys. Rev. Lett. {\bf 50}, 87 (1983); A. Le Yaouanc, L. Oliver,
1966: O. Pene, and J.-C. Raynal, Phys. Lett. {\bf 134B}, 249 (1984); Phys.
1967: Rev. D {\bf 29}, 1233 (1984).
1968: \bibitem{Orsay2} A. Le Yaouanc, L. Oliver, S. Ono, O. Pene and
1969: J.-C. Raynal, Phys. Rev. D {\bf 31}, 137 (1985).
1970: \bibitem{Lisbon} P. Bicudo and J. E. Ribeiro, Phys. Rev. D {\bf 42},
1971: 1611 (1990); {\it ibid.}, 1625 (1990); {\it ibid.}, 1635 (1990); P.
1972: Bicudo, Phys. Rev. Lett. {\bf 72}, 1600 (1994); P. Bicudo, Phys. Rev. C
1973: {\bf 60}, 035209 (1999).
1974: \bibitem{coulg} N. H. Christ and T. D. Lee, Phys. Rev. D {\bf 22}, 939
1975: (1980); see also A. P. Szczepaniak, E. S. Swanson, Phys. Rev. D {\bf
1976: 65}, 025012 (2002) and references therein.
1977: \bibitem{linear} S. L. Adler and A. C. Davis, Nucl. Phys. {\bf 244B},
1978: 469 (1984); Y. L. Kalinovsky, L. Kaschluhn, and V. N. Pervushin, Phys.
1979: Lett. {\bf 231B}, 288 (1989); P. Bicudo, J. E. Ribeiro, and J.
1980: Rodrigues, Phys. Rev. C {\bf 52}, 2144 (1995); R. Horvat, D. Kekez, D.
1981: Palle, and D. Klabucar, Z. Phys. C {\bf 68}, 303 (1995); Yu. A. Simonov,
1982: Yad. Fiz. {\bf 60}, 2252 (1997) [Phys. Atom. Nucl. {\bf 60}, 2069
1983: (1997)]; N. Brambilla and A. Vairo, Phys. Lett. {\bf 407B}, 167 (1997);
1984: Yu. A. Simonov and J. A. Tjon, Phys. Rev. D {\bf 62}, 014501 (2000); P.
1985: Bicudo, N. Brambilla, E. Ribeiro, and A. Vairo, Phys. Lett. {\bf 442B},
1986: 349 (1998); F. J. Llanes-Estrada and S. R. Cotanch, Phys. Rev. Lett.
1987: {\bf 84}, 1102 (2000).
1988: \bibitem{pipi} P. Bicudo, S. Cotanch, F. Llanes-Estrada, P. Maris, E.
1989: Ribeiro, and A. Szczepaniak, Phys. Rev. D {\bf 65}, 076008 (2002).
1990: \bibitem{replica1} P. J. A. Bicudo, A. V. Nefediev, and J. E. F. T.
1991: Ribeiro, Phys. Rev. D {\bf 65}, 085026 (2002).
1992: \bibitem{replica2} A. V. Nefediev and J. E. F. T. Ribeiro, Phys. Rev. D
1993: {\bf 67}, 034028 (2003).
1994: \bibitem{replica4} P. J. A. Bicudo and A. V. Nefediev, Phys. Rev. D {\bf
1995: 68}, 065021 (2003).
1996: \bibitem{replica3} A. A. Osipov and B. Hiller, Phys. Lett. {\bf 539B},
1997: 76 (2002).
1998: \bibitem{2d} Yu. S. Kalashnikova and A. V. Nefediev, Yad. Fiz. {\bf 62},
1999: 359 (1999) [Phys. Atom. Nucl. {\bf 62}, 323 (1999)]; Usp. Fiz. Nauk {\bf
2000: 172}, 378 (2002) [Phys. Usp. {\bf 45}, 347 (2002)]; Yu. S. Kalashnikova,
2001: A. V. Nefediev, A. V. Volodin, Yad. Fiz. {\bf 63}, 1710 (2000) [Phys.
2002: Atom. Nucl. {\bf 63}, 1623 (2000)].
2003: \bibitem{tHooft} G. 't Hooft, Nucl. Phys. {\bf B75} (1974) 461.
2004: \bibitem{japan} K. Kikkawa, Ann. Phys. {\bf 66}, 3633 (1981); A.
2005: Nakamura and K. Odaka, Phys. Lett. {\bf 105B}, 392 (1981); Nucl. Phys.
2006: {\bf 202B}, 457 (1982); S. G. Rajeev, Int. Journ. Mod. Phys. {\bf A9},
2007: 5583 (1994);A. Dhar, C. Mandal, and S. R. Wadia, Phys. Lett. {\bf
2008: 329B}, 15 (1994); A. Dhar {\it et.al.} Int. Journ. Mod. Phys. {\bf
2009: A10}, 15 (1995); M. Cavicchi, Int. Journ. Mod. Phys. {\bf A10}, 167
2010: (1995); K. Itakura, Phys. Rev. D {\bf 54}, 2853 (1996).
2011: \bibitem{hl} Yu. A. Simonov, Yad. Fiz. {\bf 60}, 2252 (1997) [Phys.
2012: At.Nucl. {\bf 60}, 2069 (1997)]; N. Brambilla and A. Vairo, Phys. Lett.
2013: {\bf 407B}, 167 (1997); Yu. S. Kalashnikova and A. V. Nefediev, Phys.
2014: Lett. {\bf 414B}, 149 (1997).
2015: \bibitem{GMOR} M. Gell-Mann, R. J. Oakes, and B. Renner, Phys. Rev. {\bf
2016: 175}, 2195 (1968).
2017: \bibitem{replica5} P. J. A. Bicudo and A. V. Nefediev, Phys. Lett. {\bf
2018: 573B}, 131 (2003).
2019: \bibitem{emilioResina} P. J. A. Bicudo, J. R. Rodrigues and
2020: J. E. F. T. Ribeiro, Phys. Rev C {\bf 52}, 2144 (1995).
2021: \bibitem{BG} I. Bars, M. B. Green, Phys. Rev. D {\bf 17}, 537 (1978).
2022: \end{thebibliography}
2023:
2024: \appendix
2025: \section{The pion Bethe-Salpeter equation}
2026:
2027: In this appendix we derive the Bethe-Salpeter equation for the pion
2028: for the generic form of the potential and
2029: for the Lorentz structure of the confinement being $\gamma_0\times\gamma_0$.
2030: We generalise the method suggested in the paper \cite{BG} for two-dimensional QCD.
2031: We start from Eq.~(\ref{GenericSal}) for the mesonic Salpeter amplitude and
2032: define a matrix mesonic wave function,
2033: \be
2034: \Psi(\vec{p};M_\pi)=\int\frac{dp_0}{2\pi}S(\vec{p},p_0+M_\pi/2)\chi(\vec{p};M_\pi)S(\vec{p},p_0-M_\pi/2).
2035: \ee
2036:
2037: We also present the Dirac projectors (\ref{Feynman}) in the form:
2038: \be
2039: \Lambda^\pm(\vec{p})=T_pP_\pm T_p^\dagger,\quad
2040: P_\pm=\frac{1\pm\gamma_0}{2},\quad
2041: T_p=\exp{\left[-\frac12(\vec{\gamma}\hat{\vec{p}})\left(\frac{\pi}{2}-\vp_p\right)\right]},
2042: \ee
2043: and introduce a modified wave function
2044: $\tilde{\Psi}(\vec{p};M_\pi)=T^\dagger_p\Psi(\vec{p};M_\pi)T^\dagger_p$.
2045: Equation for this new matrix wave function following from
2046: Eq.~(\ref{GenericSal}), with $\Gamma=\gamma_0$, reads:
2047: \be
2048: \tilde{\Psi}(\vec{p};M_\pi)=-\int\frac{d^3q}{(2\pi)^3}V(\vec{p}-\vec{q})\left[
2049: P_+\frac{T_p^\dagger
2050: T_q\tilde{\Psi}(\vec{q};M_\pi)T_qT_p^\dagger}{2E_p-M_\pi}P_-
2051: +P_-\frac{T_p^\dagger
2052: T_q\tilde{\Psi}(\vec{q};M_\pi)T_qT_p^\dagger}{2E_p+M_\pi}P_+
2053: \right].
2054: \label{PHI}
2055: \ee
2056: It is clear that a solution of Eq.~(\ref{PHI}) has the form,
2057: \be
2058: \tilde{\Psi}(\vec{p};M_\pi)=P_+AP_-+P_-BP_+,
2059: \label{AB}
2060: \ee
2061: and, due to the obvious orthogonality property of the projectors
2062: $P_\pm$, $P_+P_-=P_-P_+=0$, only matrices anti-commuting with the
2063: matrix $\gamma_0$ contribute to $A$ and $B$. The set of such
2064: matrices is
2065: $\{\gamma_5,\gamma_0\gamma_5,\vec{\gamma},\gamma_0\vec{\gamma}\}$
2066: which can be reduced even more, up to $\{\gamma_5,\vec{\gamma}\}$,
2067: since the matrix $\gamma_0$ can be always absorbed into projectors
2068: $P_\pm$. For the case of the chiral pion only $\gamma_5$
2069: contributes and one has:
2070: \be
2071: A_\pi=\gamma_5\vp^+_\pi(p),\quad
2072: B_\pi=-\gamma_5\vp_\pi^-(p),
2073: \label{AB2}
2074: \ee
2075: where the signs and the coefficients are chosen such that to comply with the
2076: definitions (\ref{Mmpi}) and (\ref{vppi}). It is an easy task now
2077: to extract the amplitudes $T_\pi^{\pm\pm}$ (see Eq.~(\ref{bsp}))
2078: from Eq.~(\ref{PHI}) using the explicit form of $\tilde{\Psi}_\pi(\vec{p};M_\pi)$
2079: and the operator $T_p$. They read:
2080: \be
2081: \begin{array}{lcr}
2082: T_\pi^{++}(p,q)=T_\pi^{-+}(p,q)&=&-\ds\int d\Omega_q V(\vec{p}-\vec{q})
2083: \left[\cos^2\frac{\vp_p-\vp_q}{2}-\frac{1-(\hat{\vec{p}}\hat{\vec{q}})}{2}\cos\vp_p\cos\vp_q\right],\\[5mm]
2084: T_\pi^{+-}(p,q)=T_\pi^{--}(p,q)&=&\ds\int d\Omega_q V(\vec{p}-\vec{q})
2085: \left[\sin^2\frac{\vp_p-\vp_q}{2}+\frac{1-(\hat{\vec{p}}\hat{\vec{q}})}{2}\cos\vp_p\cos\vp_q\right].
2086: \end{array}
2087: \label{pia}
2088: \ee
2089:
2090: For the harmonic oscillator potential, and with the amplitudes (\ref{pia}),
2091: Eq.~(\ref{bsp}) reproduces Eq.~(\ref{hop}).
2092:
2093: In the chiral limit, $\vp_\pi^+(p)=-\vp_\pi^-(p)=\vp_\pi(p)$, so that the
2094: bound-state equation (\ref{bsp}) reduces to a single equation,
2095: \be
2096: 2E_p\vp_\pi(p)=\int\frac{q^2dq}{(2\pi)^3}[T_\pi^{++}(p,q)-T_\pi^{+-}(p,q)]\vp_\pi(q)=
2097: -\int\frac{d^3q}{(2\pi)^3}V(\vec{p}-\vec{q})\vp_\pi(q),
2098: \ee
2099: or, in the coordinate space, one arrives at the Schr{\" o}dinger-like equation,
2100: \be
2101: [2E_p+V(r)]\vp_\pi=0.
2102: \ee
2103:
2104: It is also instructive to derive the bound-state equation for the pionic matrix wave function
2105: $\Psi(\vec{p};M_\pi)$. It follows directly from the representation (\ref{PHI}) after a
2106: simple algebra (see \cite{2d} for the detailed discussion of the matrix wave function formalism
2107: in two-dimensional QCD),
2108: $$
2109: M_\pi\Psi(\vec{p};M_\pi)=[(\vec{\alpha}\vec{p})+\gamma_0m]\Psi(\vec{p};M_\pi)+
2110: \Psi(\vec{p};M_\pi)[(\vec{\alpha}\vec{p})-\gamma_0m]\hspace*{5cm}
2111: $$
2112: \be
2113: +\int\frac{d^3q}{(2\pi)^3}V(\vec{p}-\vec{q})\left\{\Lambda^+(\vec{q})\Psi(\vec{p};M_\pi)\Lambda^-(-\vec{q})-
2114: \Lambda^+(\vec{p})\Psi(\vec{q};M_\pi)\Lambda^-(-\vec{p})\right.
2115: \label{matrix}
2116: \ee
2117: $$
2118: \hspace*{5cm}\left.-\Lambda^-(\vec{q})\Psi(\vec{p};M_\pi)\Lambda^+(-\vec{q})+
2119: \Lambda^-(\vec{p})\Psi(\vec{q};M_\pi)\Lambda^+(-\vec{p})\right\}.
2120: $$
2121:
2122: The explicit form of $\Psi(\vec{p};M_\pi)$ through the components $\vp_\pi^\pm$ and the
2123: chiral angle can be found easily from Eqs.~(\ref{AB}), (\ref{AB2}),
2124: \be
2125: \Psi(\vec{p};M_\pi)=T_p\left[P_+\gamma_5\vp_\pi^+-P_-\gamma_5\vp_\pi^-\right]T_p=
2126: \gamma_5G_\pi+\gamma_0\gamma_5T_p^2F_\pi,
2127: \label{exp}
2128: \ee
2129: where $G_\pi=\frac12(\vp_\pi^+-\vp_\pi^-)$, $F_\pi=\frac12(\vp_\pi^++\vp_\pi^-)$.
2130:
2131: Let us multiply the matrix bound-state equation (\ref{matrix}) by $\gamma_0\gamma_5$, integrate its
2132: both parts over $d^3p$, and, finally, take the trace. The resulting equation reads:
2133: \be
2134: M_\pi\int\frac{d^3p}{(2\pi)^3}F_\pi\sin\vp_p=2m\int\frac{d^3p}{(2\pi)^3}G_\pi.
2135: \label{GMOR2}
2136: \ee
2137:
2138: It is easy to recognise the Gell-Mann--Oakes--Renner relation (\ref{GMOR}) in
2139: Eq.~(\ref{GMOR2}). Indeed,
2140: using the explicit form of the pionic wave function beyond the chiral limit,
2141: Eq.~(\ref{vppm}), one can
2142: see that $G_\pi=(\tilde{\cal N}_\pi/\sqrt{M_\pi})\sin\vp_p$ and
2143: $F_\pi=(\tilde{\cal N}_\pi\sqrt{M_\pi})\Delta_p$ which, after substitution to
2144: Eq.~(\ref{GMOR2}), give the sought relation,
2145: \be
2146: M_\pi^2\left[\frac{N_C}{\pi^2}\int_0^\infty dp\; p^2\Delta_p\sin\vp_p\right]=-2m\langle\bar{q}q\rangle,
2147: \label{GMOR3}
2148: \ee
2149: where the definition of the chiral condensate (\ref{Sigma1}) was used.
2150: \end{document}
2151: