hep-ph0605290/sg.tex
1: \documentclass[prd,eqsecnum,twocolumn,amsfonts,showpacs]{revtex4}
2: 
3: \usepackage{graphicx}
4: 
5: \usepackage{bm}
6: 
7: \setlength{\unitlength}{1cm}
8: 
9: \def\fsl#1{\setbox0=\hbox{$#1$}           % set a box for #1
10:    \dimen0=\wd0                                 % and get its size
11:    \setbox1=\hbox{/} \dimen1=\wd1               % get size of /
12:    \ifdim\dimen0>\dimen1                        % #1 is bigger
13:       \rlap{\hbox to \dimen0{\hfil/\hfil}}      % so center / in box
14:       #1                                        % and print #1
15:    \else                                        % / is bigger
16:       \rlap{\hbox to \dimen1{\hfil$#1$\hfil}}   % so center #1
17:       /                                         % and print /
18:    \fi}                                         %
19: 
20: \newcommand{\beq}{\begin{equation}}
21: \newcommand{\eeq}{\end{equation}}
22: \newcommand{\beqs}{\begin{eqnarray}}
23: \newcommand{\eeqs}{\end{eqnarray}}
24: \newcommand{\lsim}{\mathrel{\raisebox{-
25: .6ex}{$\stackrel{\textstyle<}{\sim}$}}}
26: \newcommand{\gsim}{\mathrel{\raisebox{-
27: .6ex}{$\stackrel{\textstyle>}{\sim}$}}}
28: \newcommand{\pslash}{p\hspace{-0.067in}\slash}
29: \newcommand{\qslash}{q\hspace{-0.067in}\slash}
30: \newcommand{\kslash}{k\hspace{-0.067in}\slash}
31: \newcommand{\drawsquare}[2]{\hbox{%
32: \rule{#2pt}{#1pt}\hskip-#2pt%  left vertical
33: \rule{#1pt}{#2pt}\hskip-#1pt%  lower horizontal
34: \rule[#1pt]{#1pt}{#2pt}}\rule[#1pt]{#2pt}{#2pt}\hskip-#2pt%  upper horizontal
35: \rule{#2pt}{#1pt}}% right vertical
36: \newcommand{\fund}{\raisebox{-.5pt}{\drawsquare{6.5}{0.4}}}%  fund
37: \newcommand{\sym}{\raisebox{-.5pt}{\drawsquare{6.5}{0.4}}\hskip-0.4pt%
38:         \raisebox{-.5pt}{\drawsquare{6.5}{0.4}}}%  symmetric second rank
39: \newcommand{\asym}{\raisebox{-3.5pt}{\drawsquare{6.5}{0.4}}\hskip-6.9pt%
40:         \raisebox{3pt}{\drawsquare{6.5}{0.4}}}%  antisymmetric second rank
41: 
42: \begin{document}
43: 
44: \title{Study of the Change from Walking to Non-Walking Behavior \\
45: in a Vectorial Gauge Theory as a Function of $N_f$} 
46: 
47: \author{Masafumi Kurachi}
48: %\thanks{masafumi.kurachi@sunysb.edu}
49: 
50: \author{Robert Shrock}
51: %\thanks{robert.shrock@sunysb.edu}
52: 
53: \affiliation{
54: C.N. Yang Institute for Theoretical Physics \\
55: State University of New York \\
56: Stony Brook, NY 11794}
57: 
58: \begin{abstract}
59: 
60: We study a vectorial gauge theory with gauge group ${\rm SU}(N_c)$ and a
61: variable number, $N_f$, of massless fermions in the fundamental representation
62: of this group.  Using approximate solutions of Schwinger-Dyson and
63: Bethe-Salpeter equations, we calculate meson masses and investigate how these
64: depend on $N_f$.  We focus on the range of $N_f$ extending from near the
65: boundary with a non-Abelian Coulomb phase, where the theory exhibits a slowly
66: running (``walking'') gauge coupling, toward smaller values where the theory
67: has non-walking behavior. Our results include determinations of the masses of
68: the lowest-lying flavor-adjoint mesons with $J^{PC}=0^{-+}$, $1^{--}$,
69: $0^{++}$, and $1^{++}$ (the generalized $\pi$, $\rho$, $a_0$, and $a_1$).
70: Related results are given for flavor-singlet mesons and for the generalization
71: of $f_\pi$.  These results give insight into the change from walking to
72: non-walking behavior in a gauge theory, as a function of $N_f$. 
73: 
74: \end{abstract}
75: 
76: \pacs{11.10.St, 12.38.Aw, 12.40.Yx, 14.40-n} 
77: 
78: %\keywords{} 
79: 
80: \maketitle
81: 
82: \vspace{16mm}
83: 
84: \newpage
85: \pagestyle{plain}
86: \pagenumbering{arabic}
87: 
88: \section{Introduction}
89: 
90: We consider a $(3+1)$-dimensional vectorial gauge theory (at zero temperature
91: and chemical potential) with the gauge group SU($N_c$) and $N_f$ massless
92: fermions transforming according to the fundamental representation of this
93: group.  For $N_c=3$, if one took $N_f=2$, this would be an approximation to
94: actual quantum chromodynamics (QCD) with just the $u$ and $d$ quarks, since
95: their current quark masses are small compared with the scale $\Lambda_{QCD}
96: \simeq 400$ MeV.  We restrict here to the range $N_f < (11/2)N_c$ for which the
97: theory is asymptotically free.  An analysis using the two-loop beta function
98: and Schwinger-Dyson equation (reviewed below) leads to the inference that for
99: $N_f$ in this range, the theory includes two phases: (i) for $0 \le N_f \le
100: N_{f,cr}$ a phase with confinement and spontaneous chiral symmetry breaking
101: (S$\chi$SB); and (ii) for $N_{f,cr} \le N_f \le (11/2)N_c$ a non-Abelian
102: Coulomb phase with no confinement or spontaneous chiral symmetry breaking. We
103: shall refer to $N_{f,cr}$, the critical value of $N_f$, as the boundary of the
104: non-Abelian Coulomb (conformal) phase \cite{bz}.  Here we take electroweak
105: interactions to be turned off.  We denote the fermions as $f_i^a$ with
106: $a=1,...,N_c$ and $i=1,...,N_f$.  The theory has an ${\rm SU}(N_f)_L \times
107: {\rm SU}(N_f)_R \times {\rm U}(1)_V$ global symmetry (the U(1)$_A$ being
108: explicitly broken by instantons), which is spontaneously broken to ${\rm
109: SU}(N_f)_V \times {\rm U}(1)_V$ by the formation of a bilinear fermion
110: condensate.
111: 
112: For $N_f$ slightly less than $N_{f,cr}$, the theory exhibits an approximate
113: infrared (IR) fixed point with resultant walking behavior.  That is, as the
114: energy scale $\mu$ decreases from large values, $\alpha=g^2/(4\pi)$ ($g$ being
115: the SU($N_c$) gauge coupling) grows to be O(1) at a scale $\Lambda$, but
116: increases only rather slowly as $\mu$ decreases below this scale, so that there
117: is an extended interval in energy below $\Lambda$ where $\alpha$ is large, but
118: slowly varying.  Associated with this slowly running behavior, the resultant
119: dynamically generated fermion mass, $\Sigma$, is much smaller than $\Lambda$.
120: In addition to its intrinsic field-theoretic interest, this walking behavior
121: has played an important role in theories of dynamical electroweak symmetry
122: breaking \cite{wtc1}-\cite{chipt3}.  As $N_f$ approaches $N_{f,cr}$ from below,
123: quantities with dimensions of mass vanish continuously; i.e., the chiral phase
124: transition separating phases (i) and (ii) is continous.  Recently, meson masses
125: and other quantities such as the generalized pseudoscalar decay constant
126: $f_\pi$ were calculated in the walking limit of an SU($N_c$) gauge theory
127: \cite{mm}.
128: 
129: In the present paper we shall investigate how meson masses and other quantities
130: change as one decreases $N_f$ below $N_{f,cr}$, moving away from the boundary,
131: as a function of $N_f$, between phases (i) and (ii), deeper into the confined
132: phase.  Our paper is thus a study of the change (crossover) between the walking
133: behavior that occurs near to this boundary, and the non-walking behavior that
134: occurs for smaller $N_f$.  In a non-walking (asymptotically free, confining)
135: theory such as real QCD, as the energy scale $\mu$ decreases through $\Lambda$,
136: $\alpha$ increases rapidly through values of order unity, triggering
137: spontaneous chiral symmetry breaking on this scale, so that $\Sigma \sim
138: \Lambda$.  This is quite different from a theory with walking, in which $\Sigma
139: \ll \Lambda$.  Our basic calculational methods are essentially the same as
140: those employed in Ref. \cite{mm}, i.e., we use the Schwinger-Dyson (SD)
141: equation to compute the dynamical fermion mass $\Sigma$ (generalized
142: constituent quark mass) and then insert this into the Bethe-Salpeter (BS)
143: equation to obtain the masses of the low-lying mesons.  We restrict to an
144: interval of $N_f$ values for which the theory has an infrared fixed point (as
145: calculated from the beta function, to be discussed further below).  The reason
146: for this is that it makes our calculations more robust since for our interval
147: of $N_f$ we can avoid having to introduce a cutoff on the growth of $\alpha$ in
148: the infrared.  If one uses Schwinger-Dyson and Bethe-Salpeter equations to
149: explore a region of $N_f$ where the beta function does not have an infrared
150: fixed point, one must use such an IR cutoff, which leads to cutoff-dependence
151: of the results.  For definiteness, we shall take $N_c=3$; however, as will be
152: seen, $N_c$ only enters indirectly, via the dependence of the value of the
153: infrared fixed point $\alpha_*$ (eq. (\ref{alfcrit}) below) on $N_c$.  Hence,
154: our findings may also be applied in a straightforward way, with appropriate
155: changes in the value of $\alpha_*$, to an SU($N_c$) gauge theory with a
156: different value of $N_c$.
157: 
158: 
159: We mention some background and related work.  Many studies have investigated
160: the hadron mass spectrum for QCD with $N_f=2$ light quarks.  Among the earliest
161: were static quark models \cite{qm}, and bag models\cite{mitbag,mitbag_pwave}.
162: Lattice gauge theory has provided an especially powerful method \cite{lat}. The
163: Schwinger-Dyson and Bethe-Salpeter equations have been used for many years to
164: study spontaneous chiral symmetry breaking and relativistic bound states in
165: field theories (a partial list of papers and reviews includes \cite{wtc2}-
166: \cite{mm}, \cite{bs}-\cite{HY}).  In particular, the
167: Bethe-Salpeter equation has been used to calculate meson masses in QCD
168: \cite{abkmn}-\cite{marisroberts03}.  For the walking limit, in addition to
169: Refs. \cite{mm}, these methods have also been used in connection with spectral
170: function sum rules to study the $\pi^+ - \pi^0$ mass difference \cite{pimdif}
171: and the $S$ parameter (equivalently, the chiral Lagrangian coefficient
172: $L_{10}$) \cite{HKY}.
173: 
174: This paper is organized as follows. In Section II we review some background
175: material concerning the beta function, approximate infrared fixed point, and
176: walking behavior.  Section III includes a discussion of the Schwinger-Dyson
177: equation and our solution of it, as well as our calculation of the pseudoscalar
178: decay constant $f_P$, the generalization of $f_\pi$.  In Section IV we present
179: our calculation of meson masses using the Bethe-Salpeter equation.  Section V
180: contains some further remarks and our conclusions.  
181: 
182: 
183: \section{Preliminaries}
184: 
185: In order to study meson masses and other quantities as one moves away from the
186: boundary between the confined phase with spontaneous chiral symmetry breaking
187: and the non-Abelian Coulomb phase, it is first necessary to know as accurately
188: as possible where this boundary lies, as a function of $N_f$, i.e., to know the
189: value of $N_{f,cr}$.  We first review the estimate \cite{chipt3} based on using
190: the two-loop SU($N_c$) beta function \cite{b0,b1}
191: %
192: \beq
193: \beta = \frac{d \alpha}{dt} = - \frac{\alpha^2}{2\pi}\left ( b_0 +
194: \frac{b_1}{4\pi}\alpha + O(\alpha^2) \right ) \ ,
195: \label{beta}
196: \eeq
197: %
198: where $t=\ln \mu$, with $\mu$ the energy scale. The two terms listed are
199: scheme-independent.  (Two higher-order terms in $\beta$ have been calculated
200: but are scheme-dependent; inclusion of these does not significantly affect our
201: results.) For the relevant case of an asymptotically free theory, $b_0 > 0$ so
202: that an infrared fixed point exists if $b_1 < 0$.  This coefficient $b_1$ is
203: positive for $0 \le N_f \le N_{f,IR}$, where
204: %
205: \beq
206: N_{f,IR}=\frac{34N_c^3}{13N_c^2-3}
207: \label{nfir}
208: \eeq
209: %
210: and negative for larger $N_f$.  For $N_c=3$, $N_{f,IR} \simeq 8.1$
211: \cite{integer}.  The value of $\alpha$ at this IR fixed point, denoted
212: $\alpha_*$, is given by $\alpha_* = -4\pi b_0/b_1$.  Substituting the known
213: values of these terms, one has 
214: %
215: \beq
216: \alpha_* = \frac{-4\pi(11N_c -2N_f)}{34N_c^2-13N_cN_f+3N_c^{-1}N_f} \ . 
217: \label{alpha_irfp}
218: \eeq
219: %
220: 
221: Solving eq. (\ref{alpha_irfp}) for $N_f$ in terms of $\alpha_*$ yields 
222: %
223: \beq
224: N_f=\frac{2N_c^2[17N_c(\alpha_*/\pi)+22]}{(13N_c^2-3)(\alpha_*/\pi)+8N_c} \ . 
225: \label{nfsol}
226: \eeq
227: %
228: In the one-gluon exchange approximation, the Schwinger-Dyson gap equation for
229: the inverse propagator of a fermion transforming according to the
230: representation $R$ of SU($N_c$) has a nonzero solution for the dynamically
231: generated fermion mass, which is an order parameter for spontaneous chiral
232: symmetry breaking, if $\alpha \ge \alpha_{cr}$, where $\alpha_{cr}$ is given by
233: %
234: \beq
235: \frac{3 \alpha_{cr} C_2(R)}{\pi} = 1,
236: \label{alfcritcondition}
237: \eeq
238: %
239: and $C_2(R)$ denotes the quadratic Casimir invariant for the representation $R$
240: \cite{casimir}.  Using $C_2(fund.) \equiv C_{2f}=(N_c^2-1)/(2N_c)$ for the
241: fundamental representation yields
242: %
243: \beq
244: \alpha_{cr} = \frac{2\pi N_c}{3(N_c^2-1)} \ . 
245: \label{alfcrit}
246: \eeq
247: %
248: For the case $N_c=3$ that we use for definiteness here, eq. (\ref{alfcrit})
249: gives $\alpha_{cr} = \pi/4 \simeq 0.79$.  To estimate $N_{f,cr}$, one solves
250: the equation $\alpha_* = \alpha_{cr}$, yielding the result \cite{chipt3}
251: %
252: \beq
253: N_{f,cr} = \frac{2N_c(50N_c^2-33)}{5(5N_c^2-3)} \ . 
254: \label{nfcr}
255: \eeq
256: %
257: For $N_c=3$ this gives $N_{f,cr} \simeq 11.9$.  These estimates are only rough,
258: in view of the strongly coupled nature of the physics.  Effects of higher-order
259: gluon exchanges have been studied in Ref. \cite{alm}.  These calculations are
260: semi-perturbative and do not include instanton effects.  It is known that
261: instantons enhance the formation of the bilinear fermion condensates
262: \cite{instantons}, which suggests that their inclusion would expand the phase
263: with confinement and spontaneous chiral symmetry breaking, i.e., would increase
264: $N_{f,cr}$ somewhat relative to the value obtained from the two-loop beta
265: function and gap equation.  In principle, lattice gauge simulations provide a
266: way to determine $N_{f,cr}$, but the groups that have studied this have
267: not reached a consensus \cite{iwasaki}-\cite{mawhinney}.  
268: 
269: In our analysis, what we actually vary is the value of the approximate IR fixed
270: point $\alpha_*$, which depends parametrically on $N_f$.  Thus, although our SD
271: and BS equations are semi-perturbative, the analysis is self-consistent in the
272: sense that our $\alpha_{cr}$ really is the value at which, in our
273: approximation, one passes from the confinement phase to the non-Abelian Coulomb
274: phase, and our values of $\alpha$ do span the interval over which there is a
275: crossover from walking to QCD-like (i.e., non-walking) behavior. 
276: 
277: We elaborate here on the origin of the walking behavior.  Since the theory is
278: asymptotically freee, it follows that as the energy scale $\mu$ decreases from
279: values $\gg \Lambda$, $\alpha$ increases.  If $N_f < N_{f,IR}$, there is no
280: perturbative IR fixed point.  If $N_{f,IR} < N_f < N_{f,cr}$, as the energy
281: scale decreases toward zero, the coupling $\alpha$ approaches $\alpha_*$, which
282: is larger than $\alpha_{cr}$.  The coupling $\alpha_*$ is only an approximate
283: IR fixed point since, as $\alpha$ increases past $\alpha_{cr}$ and the fermion
284: condensate forms, the fermions gain a dynamical mass $\Sigma$ so that in the
285: low-energy effective theory applicable for energy scales $\mu < \Sigma$, one
286: integrates out these fermions and is left with a pure gluonic SU($N_c$) theory
287: with a different beta function, such that $\alpha$ increases further. If $N_f >
288: N_{f,cr}$ (and smaller than $(11/2)N_c$), the theory is in the non-Abelian
289: Coulomb phase and $\alpha_*$ is an exact IR fixed point.  In the case that
290: $N_f$ is only slightly less than $N_{f,cr}$, or equivalently, $\alpha_*$ is
291: only slightly greater than $\alpha_{cr}$, it follows that as the energy scale
292: decreases and $\alpha$ approaches $\alpha_*$ from below, the rate of increase
293: of $\alpha$, i.e., $|\beta|$, decreases, so that the theory has a large
294: coupling $\alpha \sim O(1)$ which, however, runs very slowly.  
295: 
296: 
297: As is evident from the above results, decreasing $N_f$ below $N_{f,cr}$ has the
298: effect of increasing $\alpha_*$ and thus moving the theory deeper in the phase
299: with confinement and spontaneous chiral symmetry breaking, away from the
300: boundary with the non-Abelian Coulomb phase.  This is the key parametric
301: dependence that we shall use for our study. Our aim is to investigate how meson
302: masses and other observable quantities depend on $N_f$ in the crossover region;
303: operationally, what we actually vary is $\alpha_*$.  In Ref. \cite{mm} the
304: range of $\alpha_*$ used for the calculation of meson masses was chosen to be
305: $0.89 \le \alpha_* \le 1.0$, an interval where there is pronounced walking
306: behavior.  For the case $N_c=3$ considered in Ref. \cite{mm} and here, given
307: the above-mentioned value, $\alpha_{cr}=\pi/4$, it follows that this lower
308: limit, $\alpha_*=0.89$, is about 12 \% greater than this critical coupling.
309: The reason for this choice of lower limit on $\alpha_*$ was that the hadron
310: masses become exponentially small relative to the scale $\Lambda$ as
311: $\alpha_* - \alpha_{cr} \to 0^+$, rendering numerical evaluations of the
312: relevant integrals increasingly difficult in this limit.  For our study of the
313: shift away from walking behavior we consider an interval extending to larger
314: couplings, from $\alpha_*=1.0$ to $\alpha_*=2.5$.  Our upper limit is chosen in
315: order for the ladder approximation used in our solutions of the Schwinger-Dyson
316: and Bethe-Salpeter equations to have reasonable reliability.  From
317: eq. (\ref{nfsol}) it follows that $\alpha_*=0.89$ corresponds to $N_f=11.65$,
318: about 2 \% less than $N_{f,cr}$.  For a coupling as large as $\alpha_* = 2.5$,
319: the semi-perturbative methods used to derive eqs. (\ref{alpha_irfp}) and
320: (\ref{nfsol}) are subject to large corrections from higher-order perturbative,
321: and from nonperturbative, contributions; recognizing this, the above upper
322: limit of $\alpha_*$ corresponds formally to $N_f \simeq 9.8$, a roughly 20 \%
323: reduction from $N_{f,cr}=11.9$.
324: 
325: Since the chiral transition which occurs as $N_f$ increases through $N_{f,cr}$
326: is second-order (continuous), and since there is no spontaneous chiral symmetry
327: breaking in the non-Abelian Coulomb phase, it follows that as $N_f \nearrow
328: N_{f,cr}$, (i) the masses of all hadron states vanish continuously; and (ii)
329: hadron states that are related to each other by a parity reflection become
330: degenerate. 
331: 
332: 
333: \section{Schwinger-Dyson Equation}
334: 
335: We first use the Schwinger-Dyson equation for the fermion propagator to
336: calculate the dynamically generated mass $\Sigma$ of this fermion.  This
337: extends the calculation of these quantities in Ref. \cite{mm} to smaller $N_f$
338: and, accordingly, larger $\alpha_*$.  The inverse fermion propagator is
339: $S_f(p)^{-1} = A(p^2) \fsl{p} - B(p^2)$.  We approximate the full
340: Schwinger-Dyson equation by using an effective running coupling and the
341: lowest-order gluon propagator:
342: %
343: \beq
344: S_f(p)^{-1} - \fsl{p} = -i C_{2f} \int \frac{d^4 q}{(2\pi)^4}
345: \bar{g}^2(p,q) \, D_{\mu\nu}(p-q) \,  \gamma^\mu \, S_f(q) \, \gamma^\nu
346: \label{sdeqgen}
347: \eeq
348: %
349: We use the Landau gauge for the gluon propagator $D_{\mu\nu}(k)$, i.e.,
350: $D_{\mu\nu} = (-g_{\mu\nu}+k_\mu k_\nu/k^2)/k^2$ because this simplifies the
351: calculation.  The physical results are, of course, gauge-invariant (e.g.,
352: \cite{alm}).  Equation (\ref{sdeqgen}) yields two separate equations for
353: $A(p^2)$ and $B(p^2)$.  As in Ref. \cite{mm}, we make the ansatz for the
354: running coupling, after Euclidean rotation,
355: %
356: \beq
357: \alpha(p_E,q_E) = \alpha(p_E^2+q_E^2) \ , 
358: \label{gsqform}
359: \eeq
360: %
361: where the subscript denotes Euclidean.  Since $\alpha$ would naturally depend
362: on the gluon momentum squared, $(p-q)^2 = p^2+q^2-2p \cdot q$, the
363: functional form (\ref{gsqform}) amounts to dropping the scalar product term,
364: $-2p \cdot q$.  This is a particularly reasonable approximation in the case of
365: a walking gauge theory because most of the contribution to the integral
366: (\ref{sdeqgen}) comes from a region of Euclidean momenta where $\alpha$ is
367: nearly constant.  Hence, the shift upward or downward due to the $-2p \cdot q$
368: term in the argument of $\alpha$ has very little effect on the value of this
369: coupling for the range of momenta that make the most important contribution to
370: the integral. The approximation (\ref{gsqform}) enables one to carry out the
371: angular integration, obtaining the results $A(p_E^2)=1$ and, for $B(p_E^2)
372: \equiv \Sigma(p_E^2)$, setting $x \equiv p_E^2$ and $y \equiv q_E^2$, 
373: %
374: \beq
375: \Sigma(x) = \frac{3 C_{2f}}{4 \pi} \int_0^\infty y \, dy
376: \frac{\alpha(x+y) \, \Sigma(y)}{\max(x,y) \, [y + \Sigma^2(y)]} \ . 
377: \label{sdeq}
378: \eeq
379: %
380: In terms of the momentum-scale-dependent fermion mass $\Sigma(p_E^2)$, we 
381: define the dynamical mass $\Sigma$ as
382: %
383: \beq
384: \Sigma \equiv \Sigma (p_E^2=\Sigma^2) \ . 
385: \label{sigdef}
386: \eeq
387: %
388: 
389: As noted above, in the walking region, over most of the range of integration
390: over $q_E$ in eq. (\ref{sdeq}) below $\Lambda$, the running coupling $\alpha$
391: is approximately constant and equal to its fixed-point value, $\alpha_*$ (see
392: Fig. 2 of Ref. \cite{mm}).  This means that in the walking region one does not
393: have to introduce any infrared cutoff on the growth of $\alpha$, as was
394: necessary in earlier studies of the Schwinger-Dyson and Bethe-Salpeter
395: equations for regular QCD \cite{higashijima}, \cite{miranskyrev}, \cite{abkm}.
396: In the interval $q_E \lsim \Sigma \ll \Lambda$, the fermions decouple, having
397: gained dynamical masses $\Sigma$, and in this low-energy theory with the
398: fermions integrated out, the resultant $\alpha$ would evolve away from
399: $\alpha_*$ as calculated via the perturbative beta function.  However, since
400: this dynamical mass scale is much smaller than $\Lambda$ in a walking theory,
401: it follows that this lowest range of the integration over $q_E$ makes a
402: negligibly small contribution to the entire integral.  One can thus employ the
403: approximation of using the same functional form for $\alpha$ down to $q_E=0$ in
404: the integral.  This convenient feature does not hold if $N_f$ decreases very
405: far below $N_{f,cr}$, i.e., $\alpha_*$ increases too far above $\alpha_{cr}$.
406: 
407: 
408: Having made the approximation of using the same functional form for $\alpha$
409: for $k_E$ in the range $0 \le k_E \le \Sigma$ as in the range $k_E >
410: \Sigma$, and solving for $\alpha(k_E)$ from the two-loop beta function, one
411: finds that, in terms of the variable $\ln(k_E/\Lambda)$, it increases rather
412: quickly from small values to values of O(1) as $k_E$ decreases below $\Lambda$.
413: This motivates an additional simplification, namely approximating $\alpha(k_E)$
414: as the step function,
415: %
416: \beq
417: \alpha(k_E) = \alpha_*\theta(\Lambda - k_E) \ . 
418: \label{stepfun}
419: \eeq
420: %
421: 
422: Then as $\alpha_* \searrow \alpha_{cr}$, if one also approximates the
423: denominator of the fermion propagator in eq. (\ref{sdeq}) as 
424: $(q_E^2+\Sigma^2)$, i.e., one sets $\Sigma(q_E^2) = \Sigma$ in this
425: denominator, then the solution is \cite{chipt2}-\cite{chipt3}
426: %
427: \beq
428: \Sigma = c \Lambda \, \exp
429: \bigg [ -\pi \Big ( \frac{\alpha_*}{\alpha_{cr}} - 1 \Big )^{-1/2} \bigg ] \ , 
430: \label{sigsol}
431: \eeq
432: %
433: where $c$ is a constant. 
434: 
435: One next discretizes the Schwinger-Dyson equation and solves it using iterative
436: numerical methods, as described in Ref. \cite{mm}.  In Fig. \ref{Sigma_comp} we
437: show the solution for the dynamical fermion mass $\Sigma$ as a function of
438: $\alpha_*$.  A fit to the numerical solution in the walking region $0.89 \le
439: \alpha_* \le 1.0$ in Ref. \cite{mm} found agreement with the functional form
440: (\ref{sigsol}) with $c = 4.0$.  Our calculations for larger $\alpha_*$ show the
441: expected shift away from walking behavior.  This shift is evident in Fig.
442: \ref{Sigma_comp} for $\alpha_*$ larger than about 1.2.  Note that our solution
443: of the full Schwinger-Dyson equation does not make the approximation of setting
444: $\Sigma(q_E^2)=\Sigma$ in the fermion propagator denominator but instead
445: incorporates the full functional dependence of $\Sigma(q_E^2)$.  In real QCD,
446: precision fits to deep inelastic lepton scattering data, hadronic decays of the
447: $Z$, etc. probe the theory in momentum regions where $N_f = 4$ or $N_f=5$, and
448: yield, for the effective $N_f$-dependent scale $\Lambda^{(5)}_{QCD} \simeq 200$
449: MeV and $\Lambda^{(4)}_{QCD} \simeq 280$ MeV, with larger values for
450: $\Lambda^{(N_f)}_{QCD}$ with $N_f=3,2$. In actual QCD one thus has
451: $\Sigma/\Lambda^{(N_f)} \simeq O(1)$ for these low values of $N_f$.  These
452: contrast with the limiting walking behavior, in which $\Sigma \ll \Lambda$, as
453: indicated in eq. (\ref{sigsol}).  Our calculation of $\Sigma$, shown in
454: Fig. \ref{Sigma_comp}, shows that $\Sigma/\Lambda$ increases substantially, by
455: about a factor of 30, from a value of about 0.01 at $\alpha_* = 1.0$ to 0.32 at
456: $\alpha_*=2.5$, much closer to the value of O(1) for this ratio in QCD.
457: 
458: %
459: \begin{figure}
460:   \begin{center}
461:     \includegraphics[height=6cm]{Sigma_comp.eps}
462:   \end{center}
463: \caption{Numerical solutions for $\Sigma$, for several values of $\alpha_*$
464: (indicated by $\Diamond$). By way of comparison, we show, as the dotted curve,
465: the solution (\ref{sigsol}) with $c=4.0$ derived from a fit to the results in
466: the interval $0.89 \le \alpha_* \le 1.0$.  See text for further discussion.}
467: \label{Sigma_comp}
468: \end{figure}
469: %
470: 
471: Another quantity of interest is the pseudoscalar decay constant $f_P$, the
472: $N_f$-flavor generalization of the pion decay constant.  For $N_f=2$ QCD this
473: is defined as $\langle 0 | J^j_\mu | \pi^k(q) \rangle = i f_\pi q_\mu
474: \delta^{jk}$ where $1 \le j,k \le 3$ are SU(2) isospin indices. Here, we use a
475: generalization of this definition, with the symbol $f_\pi$ replaced by $f_P$
476: and the SU($N_f$) isospin indices in the range $1 \le j,k \le N_f^2-1$.  In
477: QCD, one rough measure of the dynamical (constituent) quark mass is $\Sigma
478: \simeq M_N/N_c \simeq 313$ MeV, where $M_N$ is the nucleon mass.  An alternate
479: definition sets $\Sigma \simeq M_\rho/2$; this would yield a somewhat larger
480: value.  Here we use the estimate $\Sigma \simeq 330$ MeV.  Using the measured
481: value $f_\pi \simeq 92.4 \pm 0.3$ MeV \cite{pdg}, one thus has
482: %
483: \beq
484: \bigg (\frac{\Sigma}{f_\pi} \bigg )_{QCD} \simeq 3.6 \ . 
485: \label{sig_over_fpi_qcd}
486: \eeq
487: %
488: 
489: An approximate relation connecting $\Sigma$ and $f_P$ is \cite{psrel} (with $y
490: \equiv k_E^2$)
491: %
492: \beq
493:  f_P^2 \, = \, \frac{N_c}{4\pi^2} \int_0^\infty y \, dy \,
494:        \frac{\Sigma^2(y) \, - \,  \frac{y}{4} \,
495: \frac{d}{dy} \left[ \Sigma^2(y) \right] }
496:        { [ y \ +\  \Sigma^2(y) \ ]^2}  \ .
497: \label{psrel}
498: \eeq
499: %
500: The integration is rendered finite by the softness of the dynamical
501: mass $\Sigma(k_E^2)$, which behaves for $k \gg \Sigma$ as
502: %
503: \beq
504: \Sigma(k_E^2) \propto \Sigma \bigg ( \frac{\Sigma}{k_E} \bigg )^{2-\gamma}
505: \label{sigmak}
506: \eeq
507: %
508: where $\gamma$ is the anomalous dimension of the bilinear operator $\bar f f$,
509: having the value $\gamma \simeq 1$ in the walking regime and decreasing toward
510: zero at very large energy scales $\mu >> \Lambda$ (since $\gamma$ is
511: a power series in $\alpha$ and $\alpha \to 0$ in this limit due to the
512: asymptotic freedom of the theory).  Thus, the relation (\ref{psrel}) suggests
513: that for QCD 
514: %
515: \beq
516: f_\pi^2 \simeq \frac{N_c \Sigma^2}{4\pi^2}  \ . 
517: \label{fsigrel}
518: \eeq
519: %
520: For $N_c=3$, this is $\Sigma/f_\pi \simeq 2\pi/\sqrt{3} \simeq 3.6$, which
521: agrees, to within the theoretical uncertainties, with experiment.  In QCD,
522: with $\Lambda^{(2)} \simeq 400$ MeV, one has 
523: %
524: \beq
525: \frac{f_\pi}{\Lambda^{(2)}_{QCD}} \simeq 0.25 \ . 
526: \label{fpi_over_lam_qcd}
527: \eeq
528: %
529: 
530: 
531: In Fig. \ref{fp_comp} we show our results for $f_P$ calculated from
532: substituting our solution for $\Sigma(k^2)$ into eq. (\ref{psrel}).  In the
533: walking limit, $f_P$ has been shown to satisfy a relation similar to
534: eq. (\ref{sigsol}), i.e., it is exponentially smaller than the scale $\Lambda$.
535: We display, as the dotted curve, the fit from Ref. \cite{mm} for the walking
536: interval $0.89 \le \alpha_* \le 1.0$, given by eq. (\ref{sigsol}) with $c=1.5$.
537: Our results show the change from this walking type of behavior as $\alpha_*$
538: increases above 1.2; as $\alpha_*$ increases from 1.0 to 2.5, $f_P/\Lambda$
539: increases substantially, from about $ 3 \times 10^{-3}$ to about 0.08.  This is
540: similar to the factor by which we found that $\Sigma\Lambda$ increased as
541: $\alpha_*$ increased through this interval.
542: 
543: 
544: %
545: \begin{figure}
546:   \begin{center}
547:     \includegraphics[height=6cm]{fp_comp.eps}
548:   \end{center}
549: \caption{Values of $f_P$ calculated from eq. (\ref{psrel}) for several
550:   values of $\alpha_*$ (indicated by $\Diamond$).  For comparison, we show, as
551:   the dotted line, the analytic solution given by eq. (\ref{sigsol}) with
552:   $\Sigma$ replaced by $f_P$ and $c=1.5$ derived from a fit to the
553:   calculations for $0.89 \le \alpha_* \le 2.5$.  See text for further
554:  discussion.}
555: \label{fp_comp}
556: \end{figure}
557: %
558: 
559: 
560: The strong increase in $\Sigma/\Lambda$ and $f_P/\Lambda$ as $\alpha_*$ ascends
561: from the value 0.89 near the walking limit to the value 2.5 deeper within the
562: confinement phase is easily understood as reflecting the removal of the extreme
563: exponential suppression evident in eq. (\ref{sigsol}) and its analogue for
564: $f_P$ for $\alpha_* - \alpha_{cr} \to 0^+$.  One does not expect such a
565: dramatic change in the ratio $\Sigma/f_P$ over this interval, and this
566: expectation is borne out by our calculations. In Fig. \ref{Sigoverf} we show
567: the ratio of $\Sigma/f_P$, which increases gradually from about 2.6 to 3.9.
568: The fact that we find a ratio comparable to the observed one in actual QCD,
569: given by eq. (\ref{sig_over_fpi_qcd}), can be understood as a consequence of
570: the property that much of the strong dependence on $N_f$ divides out in this
571: ratio.
572: 
573: %
574: \begin{figure}
575:   \begin{center}
576:     \includegraphics[height=6cm]{Sigoverf.eps}
577:   \end{center}
578: \caption{Plot of the ratio $\Sigma/f_P$ for $1 \le \alpha_* \le 2.5$.}
579: \label{Sigoverf}
580: \end{figure}
581: %
582: 
583: 
584: \section{Calculation of Meson Masses via the Bethe-Salpeter Equation}
585: 
586: 
587: \subsection{General Discussion} 
588: 
589: 
590: We denote the wavefunction for a hadron with a given flavor combination for the
591: generalized $\pi$, $\rho$, etc. as follows.  Define the flavor vector $f^a
592: \equiv (f^{a1},...,f^{aN_f})$.  Recall that in the confined phase the global
593: symmetry ${\rm SU}(N_f)_L \times {\rm SU}(N_f)_R \times U(1)_V$ is broken
594: spontaneously to ${\rm SU}(N_f)_V \times {\rm U}(1)_V$.  We drop the explicit
595: subscript $V$ on ${\rm SU}(N_f)_V$ henceforth.  With regard to ${\rm SU}(N_f)$,
596: a $f \bar f$ meson with a given $J^{PC}$ (where $J$ denotes the spin, and $P$
597: and $C$ are the parity and charge conjugate eigenvalues) is described via the
598: Clebsch-Gordan decomposition $N_f \times \bar N_f = 1 + Adj$, where 1 and $Adj$
599: denote the singlet and adjoint representations.
600: 
601: Let the generators of the group ${\rm SU}(N_f)$ have the standard normalization
602: ${\rm Tr}(T_iT_j)=(1/2)\delta_{ij}$.  Then the hadrons transforming according
603: to the adjoint representation of ${\rm SU}(N_f)$ are comprised of (i) the set
604: of $N_f(N_f-1)$ states
605: %
606: \beq 
607: h_{\Gamma;ij} = \frac{1}{\sqrt{N_c}}\sum_{a=1}^{N_c}\bar f_a \, 
608: \Gamma \, T_{ij} \, f^a
609: \label{hij}
610: \eeq
611: %
612: where $T_{ij}$ is the $N_f \times N_f$ matrix with a 1 in the $i$'th column and
613: $j$'th row, with $1 \le i,j \le N_f$, $i \ne j$, and $\Gamma$ specifies the
614: type of particle (pseudoscalar, vector, axial-vector, scalar), and
615: (ii) the $N_f-1$ states corresponds to the generators of the Cartan subalgebra
616: of ${\rm SU}(N_f)$ given by the traceless $N_f \times N_f$ matrices
617: % 
618: \beq 
619: T_{dk} = [2k(k+1)]^{-1/2}{\rm diag}(1,1,..,1,-k,0,..,0)
620: \label{tdk} 
621: \eeq 
622: % 
623: where there are $k$ 1's and $1 \le k \le N_f-1$.  That is, $T_{d1}=(1/2){\rm
624: diag}(1,-1,...,0)$, $T_{d2}=(2\sqrt{3})^{-1}{\rm diag}(1,1,-2,0,...,0)$, etc.
625: Because of the ${\rm SU}(N_f)$ flavor symmetry, it does not matter which of
626: these $N_f^2-1$ hadrons with a given $\Gamma$ we use.  We shall refer to these
627: as the $N_f$-generalized $\rho$, $\omega$, etc.  In particular, the spectrum of
628: mesons includes a set of $N_f^2-1$ Nambu-Goldstone bosons (NGB's) with $L=S=0$
629: and $J^{PC}=0^{-+}$, transforming according to the adjoint representation of
630: ${\rm SU}(N_f)$. The corresponding $0^{-+}$ singlet with respect to ${\rm
631: SU}(N_f)$, i.e., the generalized $\eta'$, is not a Nambu-Goldstone boson
632: because the corresponding U(1)$_A$ symmetry is anomalous.  Our analysis of
633: meson masses is for the lowest-lying $f \bar f$ states.  In future work one
634: could also consider radial excitations, Regge recurrences, pure gluonic states
635: (glueballs) and the general coupled situation in which glueballs and $f \bar f$
636: mesons of the same $J^{PC}$ mix.
637: 
638: 
639: In QCD, there are several (light-quark) $\bar q q$ mesons that are of interest
640: here. For the reader's convenience, we list these in Table \ref{mesons}.  A
641: notation for the various states in the case of general $N_f$ massless quarks is
642: $S_R$, $P_R$, $V_R$, and $A_R$, standing for ``scalar, pseudoscalar, vector,
643: and axial-vector'', where the subscript $R$ denotes the representation -
644: adjoint or singlet - under the SU($N_f$) flavor symmetry group.  The
645: experimental and theoretical situation concerning the $0^{++}$ isoscalar meson
646: $f_0$ has been the subject of much discussion over the years; indeed, this
647: state may involve mixing with $q q \bar q \bar q$ mesons \cite{schechter}.
648: Because of the complications in the analysis of this state, and the expected
649: complications in a realistic analysis of its $N_f$-generalization, the
650: SU($N_f$)-singlet $0^{++}$ meson, we do not attempt to treat this in our
651: current study.  
652: 
653: 
654: \begin{table}
655: \caption{\footnotesize{Data on relevant $q \bar q$ mesons whose masses are
656: compared with Bethe-Salpeter calculations. $n {}^{2S+1} \, L_J$
657: is standard spectroscopic notation, where $n$ denotes radial quantum number.
658: The symbols adj. and sing. denote the adjoint and singlet representations of
659: the SU(2)$_V$ isospin flavor symmetry group.  Masses are given in MeV, from 
660: \cite{pdg}. The last column lists the mass divided by a typical hadronic scale,
661: $f_\pi$.}}
662: \begin{center}
663: \begin{tabular}{|c|c|c|c|c|c|c|} \hline
664: $n {}^{2S+1} \, L_J$ & $J^{PC}$ & $R_{SU(2)_V}$ & name &  $M$ & $M/f_\pi$ 
665: \\ \hline
666: $1 {}^3 \, S_1$  & $1^{--}$  & adj.  & $\rho$   & $775.8 \pm 0.5$ & 8.40 
667: \\ \hline
668: $1 {}^3 \, S_1$  & $1^{--}$  & sing. & $\omega$ & $782.6 \pm 0.1$ & 8.47
669: \\ \hline
670: $1 {}^1 \, P_1$  & $1^{+-}$  & adj.  & $b_1$    & $1229.5 \pm 3.2$ & 13.3
671: \\ \hline
672: $1 {}^1 \, P_1$  & $1^{+-}$  & sing.  & $h_1$    & $1170 \pm 20$   & $12.7 \pm
673: 0.2$ \\ \hline
674: $1 {}^3 \, P_0$  & $0^{++}$  & adj.  & $a_0$    & $984.7 \pm 1.2$ & 10.7
675: \\ \hline
676: $1 {}^3 \, P_0$  & $0^{++}$  & sing. & $f_0$    & $\sim 600^{+600}_{-200}$ &
677: $6.5^{+6.5}_{-4.3}$ \\ \hline
678: $1 {}^3 \, P_1$  & $1^{++}$  & adj.  & $a_1$    & $1230 \pm 40$ & $13.3 \pm 
679: 0.4$  \\ \hline
680: $1 {}^3 \, P_1$  & $1^{++}$  & sing. & $f_1$    & $1281.8 \pm 0.6$ & 13.9
681: \\ \hline
682: \end{tabular}
683: \end{center}
684: \label{mesons}
685: \end{table}
686: 
687: 
688: As will be seen below, in the Bethe-Salpeter equation that we use to calculate
689: the masses of the mesons, the flavor-dependent structure is simply
690: a prefactor.  Hence, the solutions of this equation have the property 
691: that, for a given radial quantum number and spectroscopic form 
692: ${}^{2S+1} \, L_J$, the SU($N_f$) flavor-singlet and flavor-adjoint mesons 
693: are degenerate: 
694: %
695: \beq
696: M(n {}^{2S+1} \, L_J; {\rm flav. \ adjoint}) = 
697: M(n {}^{2S+1} \, L_J; {\rm flav. \ singlet})
698: \label{mrel}
699: \eeq
700: %
701: In view of this, we henceforth drop the subscript $R$ and simply write $V$
702: rather than $V_{flav. adj.}$ or $V_{flav. sing.}$, etc.  Note that this is
703: different from the prediction from SU($N_f$) flavor symmetry (with degenerate
704: quarks and electroweak interactions turned off) that the members of a given
705: representation of SU($N_f$) are degenerate.  Experimentally, except for the
706: pseudoscalar mesons, the light-quark isospin-adjoint and isospin-singlet $q
707: \bar q$ mesons are nearly degenerate. The physical $\omega$ meson is very
708: nearly a singlet under isospin SU(2), so a measure of this predicted degeneracy
709: for the ground state $1^{--}$ mesons is $(M_\omega - M_\rho)/[(1/2)((M_\omega +
710: M_\rho)] = 0.87 \times 10^{-2}$, quite small.  Similarly, $(M_{f_1} -
711: M_{a_1})/[(1/2)((M_{f_1} + M_{a_1})] = 0.04 \pm 0.03$ and $(M_{h_1} -
712: M_{b_1})/[(1/2)((M_{h_1} + M_{b_1})] =-0.05 \pm 0.02$.  So for these states the
713: prediction from our Bethe-Salpeter technique for the special case $N_f=2$
714: massless quarks is in agreement with the data for light-quark mesons in QCD.
715: 
716: The situation with the $0^{-+}$ mesons is quite different.  Since the SU($N_f$)
717: flavor-singlet mesons are not NGB's, owing to the anomalous nature of the
718: U(1)$_A$ symmetry, they are split by a large mass difference from the
719: flavor-adjoint NGB's.  In this case, as noted above, the semi-perturbative
720: Bethe-Salpeter analysis does not contain the relevant physics involving
721: instantons, and hence its prediction is far from reality.  For this reason we
722: do not consider the flavor-singlet $0^{-+}$ mesons here.  As regards the
723: flavor-adjoint $0^{-+}$ mesons, since we assume massless fermions and have
724: turned off electroweak interactions, the mass $M_P$ of the flavor-adjoint
725: pseudoscalar mesons is exactly zero in our calculations.  
726: 
727: The pion decay constant $f_\pi$ provides a convenient mass scale with which to
728: normalize the hadron masses.  For comparison with our results calculated in
729: the case of general larger $N_f$, we list in Table \ref{mesons} the masses of
730: the $q \bar q$ mesons divided by $f_\pi$.  For later use we also record the
731: ratio
732: %
733: \beq
734: \frac{M_{a_1}}{M_\rho} = 1.59 \pm 0.05 \ . 
735: \label{ma1mrho_qcd}
736: \eeq
737: %
738: This is slightly larger than the prediction $M_{a_1}/M_\rho = \sqrt{2} \simeq
739: 1.414$ from a combination of vector meson dominance and spectral function sum
740: rules \cite{wein}.  Also,
741: %
742: \beq
743: \frac{M_{a_0}}{M_\rho} = 1.27 \ . 
744: \label{ma0mrho_qcd}
745: \eeq
746: %
747: 
748: 
749: An interesting result of the calculations of meson masses in the walking limit
750: in Ref. \cite{mm} was that the ratios of these masses to $f_P$ are rather
751: constant.  Specifically, it was found that in for $0.89 \le \alpha_* \le 1.0$, 
752: %
753: \beq
754: \frac{M_V}{f_P} \simeq 11 \ , 
755: \label{mvfp_ratio_wqcd}
756: \eeq
757: %
758: %
759: \beq
760: \frac{M_A}{f_P} \simeq 11.5 \ , 
761: \label{mafp_ratio_wqcd}
762: \eeq
763: %
764: and
765: %
766: \beq
767: \frac{M_S}{f_P} \simeq 4.1 \ , 
768: \label{msfp_ratio_wqcd}
769: \eeq
770: %
771: so that 
772: %
773: \beq
774: \frac{M_A}{M_V} = 1.04 
775: \label{mamvratio}
776: \eeq
777: %
778: and
779: %
780: \beq
781: \frac{M_S}{M_V} = 0.36 \ , 
782: \label{msmvratio}
783: \eeq
784: %
785: where the theoretical uncertainty is several per cent.  These ratios may be
786: compared with the values in regular QCD which, as far as the light-meson
787: spectrum is concerned, are close to the values that they would have in the
788: $N_f=2$ chiral limit (with the understanding that the pion masses would
789: actually vanish in this limit if electroweak interactions are turned off, as
790: assumed here).  For the purpose of this comparison, we do not try to use the
791: inferred chiral-limit value of $f_\pi$ \cite{fpi}, since to be consistent we
792: would have to do the same for the mesons themselves.  For the comparison
793: between the extreme walking limit (WL) and QCD, we have
794: %
795: \beq
796: \frac{(M_V/f_P)_{WL}}{(M_\rho/f_\pi)} \simeq 1.3
797: \label{mvratio_wq}
798: \eeq
799: %
800: %
801: \beq
802: \frac{(M_A/f_P)_{WL}}{(M_{a_1}/f_\pi)} \simeq 0.86 \ , 
803: \label{maratio_wq}
804: \eeq
805: %
806: and
807: %
808: \beq
809: \frac{(M_S/f_P)_{WL}}{(M_{a_0}/f_\pi)} \simeq 0.38 \ . 
810: \label{rsratio_wq}
811: \eeq
812: %
813: A major output of the present work is the elucidation of how, as $N_f$
814: decreases and $\alpha_*$ increases, the ratios of meson masses to $f_P$ begin 
815: to shift toward their QCD values.
816: 
817: \section{Calculations} 
818: 
819: Next, we describe the details of our solution of the Bethe-Salpeter equation
820: and the resulting masses of $f \bar f$ mesons.
821: 
822: \subsection{Bethe-Salpeter amplitudes}
823: 
824: We introduce the Bethe-Salpeter amplitudes $\chi$ for 
825: the scalar ($S$), pseudoscalar ($P$), vector ($V$), and 
826: axial-vector ($A$) bound states of quark and anti-quark as follows : 
827: %
828: \beqs
829:   \langle 0 \vert 
830:    T \psi_{\alpha f i}(x_+)\ \bar\psi_\beta^{f^\prime j}(x_-) 
831:   \ \vert S_a(q) \rangle  && \\
832:   = \sqrt{2} \, \delta_i^j (T_a)_f^{f^\prime} \, 
833:   e^{-iq \cdot X} \int \frac{d^4p}{(2 \pi)^4} 
834: e^{-ip \cdot r} &[\chi_{(S)}(p;q)]_{\alpha\beta}& , \nonumber 
835: \eeqs
836: %
837: %
838: \beqs
839:   \langle 0 \vert 
840:     \ T\  \psi_{\alpha f i}(x_+)\ \bar\psi_\beta^{f^\prime j}(x_-) 
841:   \ \vert P_a(q) \rangle && \\
842:   = \sqrt{2} \, \delta_i^j (T_a)_f^{f^\prime} \, 
843:   e^{-iq \cdot X} \int \frac{d^4p}{(2 \pi)^4} 
844: e^{-ip \cdot r} &[\chi_{(P)}(p;q)]_{\alpha\beta}& ,  \nonumber 
845: \eeqs
846: %
847: %
848: \beqs
849:   \langle 0 \vert 
850:     \ T\ \psi_{\alpha f i}(x_+)\ \bar\psi_\beta^{f^\prime j}(x_-) 
851:   \ \vert V_a(q,\epsilon) \rangle && \\
852:   = \sqrt{2} \, \delta_i^j (T_a)_f^{f^\prime} \, 
853:   e^{-iq \cdot X} \int \frac{d^4p}{(2 \pi)^4} 
854: e^{-ip \cdot r} &[\chi_{(V)}(p;q,\epsilon)]_{\alpha\beta}&,  \nonumber 
855: \eeqs
856: %
857: %
858: \beqs
859:   \langle 0 \vert 
860:     \ T\  \psi_{\alpha f i}(x_+)\ \bar\psi_\beta^{f^\prime j}(x_-) \ 
861:   \vert A_a(q,\epsilon) \rangle && \\
862:   = \sqrt{2} \, \delta_i^j (T_a)_f^{f^\prime} \, 
863:   e^{-iq \cdot X} \int \frac{d^4p}{(2 \pi)^4} 
864: e^{-ip \cdot r} &[\chi_{(A)}(p;q,\epsilon)]_{\alpha \beta}&, \nonumber 
865: \eeqs
866: %
867: where $x_\pm = X \pm r/2$, and ($\alpha$, $\beta$), ($f$, $f^\prime$), and
868: ($i$, $j$) denote the spinor, flavor, and color indices, respectively.
869: $\lambda_a$ represents flavor structure of the bound states.  In the case of
870: flavor-adjoint bound states, $T_a$ is the generator of $\mbox{SU}(N_f)$,
871: while in the case of flavor singlet bound states, $(T_a)_{f}^{f'}$ is the
872: identity {\bf 1}.
873: 
874: We expand the BS amplitude $\chi$ in terms of the bispinor bases 
875: $\Gamma^i$ and the invariant amplitudes $\chi^{(i)}$ as follows :
876: %
877: \beq
878: \left[ \chi_{(S,P)}(p;q) \right]_{\alpha \beta} 
879: =\ \sum_{i=1}^{4} \left[ \Gamma^i_{(S,P)}(p;q) \right]_{\alpha \beta}
880:   \chi_{(S,P)}^{(i)} (p;q) ,
881: \label{eq:chi-SP}
882: \eeq
883: %
884: %
885: \beq
886: \left[ \chi_{(V,A)}(p;q,\epsilon) \right]_{\alpha \beta} 
887: =\ \sum_{i=1}^{8} \left[ \Gamma^i_{(V,A)}(p;q,\epsilon) \right]_{\alpha \beta}
888:   \chi_{(V,A)}^{(i)} (p;q) .
889: \label{eq:chi-VA}
890: \eeq
891: %
892: The bispinor bases can be determined from the spin, parity, and charge
893: conjugation properties of the bound states.  The explicit forms of
894: $\Gamma_{(S)}^i$, $\Gamma_{(P)}^i$, $\Gamma_{(V)}^i$, and $\Gamma_{(A)}^i$ are
895: summarized in Appendix~\ref{app:bispinor-bases}.
896: 
897: We take the rest frame of the bound state as a frame of reference:
898: %
899: \beq
900:   q^\mu = ( M_B , 0 , 0 , 0 ) ,
901: \eeq
902: %
903: where $M_B$ represents the bound state mass, i.e., $M_S, \ M_P, \ M_V$ and
904: $M_A$ for scalar, pseudoscalar, vector and axial-vector meson masses,
905: respectively.  After a Wick rotation, we parametrize $p^\mu$
906: by the real variables $u$ and $x$ as
907: %
908: \beq
909:   p \cdot q = i M_B u \ ,\  p^2 = - u^2 - x^2 .
910: \eeq
911: %
912: Consequently, the invariant amplitudes $\chi^{(i)}$
913: can be expressed as functions of the variables $u$ and $x$:
914: %
915: \beq
916:   \chi^{(i)}_{(S,P,V,A)} = \chi^{(i)}_{(S,P,V,A)}(u,x) .
917: \eeq
918: %
919: {}From the charge conjugation properties for the BS amplitude $\chi$ and the
920: bispinor bases defined in Appendix~\ref{app:bispinor-bases}, the invariant
921: amplitudes $\chi^{(i)}(u,x)$ are shown to satisfy the following relation:
922: %
923: \beq
924:   \chi^{(i)}_{(S,P,V,A)}(u,x) = \chi^{(i)}_{(S,P,V,A)}(-u,x)\ .
925: \label{eq:even_chi}
926: \eeq
927: %
928: 
929: 
930: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
931: \subsection{Homogeneous Bethe-Salpeter equation}
932: 
933: The homogeneous Bethe-Salpeter (HBS) equation is the self-consistent equation
934: for the Bethe-Salpeter amplitude, and it is expressed as (see
935: Fig.~\ref{fig:HBSeq})
936: %
937: 
938: \begin{figure}
939:   \begin{center}
940:     \includegraphics[height=2.5cm]{HBSeq.eps}
941:   \end{center}
942: \caption{A graphical representation of the HBS equation 
943: in the (improved) ladder approximation}
944: \label{fig:HBSeq}
945: \end{figure}
946: 
947: 
948: %
949: \beq
950:   T \chi = K \chi \ .
951: \label{eq:HBSeq}
952: \eeq
953: %
954: The kinetic part $T$ is given by 
955: %
956: \beq
957:   T(p;q) =  S_f^{-1}(p + q/2) \otimes S_f^{-1}(p - q/2) \ ,
958: \label{def T}
959: \eeq
960: %
961: where the BS kernel $K$ in the improved ladder approximation 
962: is expressed as
963: %
964: \beq
965:   K(p;k) \ =\    C_{2f} 
966:          \frac{4\pi \alpha(p,k)}{(p-k)^{2}}
967:          \left( g_{\mu\nu} - \frac{(p-k)_\mu (p-k)_\nu}{(p-k)^2}
968:          \right) \gamma^\mu \otimes \gamma^\nu .
969: \eeq
970: %
971: In the above expressions we used the tensor product notation
972: %
973: \beq
974:   (A \otimes B) \,\chi  =  A\, \chi\, B \ ,
975: \eeq
976: %
977: and the inner product notation 
978: \beq
979:    K \chi\ (p;q) = -i \int \frac{d^4 k}{(2\pi)^4}\  K(p,k)\  \chi(k;q) \ .
980: \eeq
981: %
982: It should be noted that the fermion propagators included in $T$ in
983: eq.~(\ref{def T}) have complex-valued arguments after the Wick rotation
984: \cite{sdcomplex}.  The arguments of the mass functions appearing in the two
985: legs of the Bethe-Salpeter amplitude are expressed as
986: %
987: \beq
988:   - ( p \pm q/2 )^2 = u^2 + x^2 - \left( \frac{M_B}{2} \right)^2 
989:                       \mp i u M_B .
990: \eeq
991: %
992: In general, it is difficult to obtain mass functions for complex arguments by
993: solving the Schwinger-Dyson equation in the complex plane, especially because
994: of the analytic structure of the running coupling in the complex momentum
995: plane.  However, in the case of large $N_f$ QCD, the analyticity of the
996: two-loop running coupling constant~\cite{Gardi} makes it possible to solve for
997: the mass function in the complex plane. This leads to the following
998: approximation, in accordance with eq.~(\ref{stepfun}) \cite{mm}:
999: %
1000: \beqs
1001:   {\rm Re}\left[\alpha(X e^{i2\theta})\right]
1002: &=& \alpha_\ast\ \theta(\Lambda^2 - X),\label{eq:real}\\
1003:   {\rm Im}\left[\alpha(X e^{i2\theta})\right]
1004: &=& 0.
1005: \eeqs
1006: %
1007: where no confusion should result from the use of the symbol $\theta$ on the
1008: right-hand side of eq.~(\ref{eq:real}) as the step function. 
1009: .
1010: 
1011: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1012: \subsection{Numerical results}
1013: 
1014: We next present the results of the numerical calculations for the masses of the
1015: mesons.  We solve the homogeneous Bethe-Salpeter equation as an eigenvalue
1016: problem, namely, the Bethe-Salpeter amplitude as an eigenfunction and the mass
1017: of a bound state as an eigenvalue, denoted generically as $M_B$.  Because the
1018: eigenvalue $M_B$ appears nonlinearly in the  equation, we use so-called
1019: fictitious eigenvalue method \cite{HY} to obtain the value of $M_B$.  For
1020: details of numerical method to solve HBS equation, see Ref.~\cite{mm}.  In
1021: Fig.~\ref{m_over_lam}, we show the values of meson masses divided by $\Lambda$
1022: calculated from the Schwinger-Dyson and Bethe-Salpeter equations in the range
1023: $0.9 \le \alpha_* \le 2.5$.  In Fig.~\ref{m_over_fp} we plot the values of
1024: $M_B/f_P$ in the range of $0.9 \le \alpha_* \le 2.5$.  In Fig.~\ref{mmratio},
1025: we plot the meson mass ratios $M_A/M_V$ and $M_S/M_V$ in the range of $0.9 \le
1026: \alpha_* \le 2.5$.
1027: 
1028: 
1029: %
1030: \begin{figure}
1031:   \begin{center}
1032:     \includegraphics[height=7cm]{m_over_lam.eps}
1033:   \end{center}
1034: \caption{Values of meson masses divided by $\Lambda$ calculated from the
1035:  Schwinger-Dyson and Bethe-Salpeter equations.}
1036: \label{m_over_lam}
1037: \end{figure}
1038: %
1039: 
1040: 
1041: %
1042: \begin{figure}
1043:   \begin{center}
1044:     \includegraphics[height=7cm]{m_over_fp.eps}
1045:   \end{center}
1046: \caption{Values of meson masses divided by $F_P$ calculated from the
1047:  Schwinger-Dyson and Bethe-Salpeter equations.}
1048: \label{m_over_fp}
1049: \end{figure}
1050: %
1051: %
1052: 
1053: %
1054: \begin{figure}
1055:   \begin{center}
1056:     \includegraphics[height=7cm]{mmratio.eps}
1057:   \end{center}
1058: \caption{Ratios of meson masses calculated from the
1059:  Schwinger-Dyson and Bethe-Salpeter equations.}
1060: \label{mmratio}
1061: \end{figure}
1062: %
1063: 
1064: Our calculations yield a number of interesting results.  We summarize these
1065: for the changes in these meson masses as $\alpha_*$ increases from 0.9 to 2.5
1066: as follows.  
1067: 
1068: \begin{enumerate}
1069: 
1070: 
1071: \item 
1072: 
1073: The ratios of the meson masses divided by $\Lambda$ increase dramatically, by
1074: factors of order $10^2$, approaching values of order unity at $\alpha_* = 2.5$.
1075: This amounts to the removal of the exponential suppression of these masses
1076: which had described the walking limit at the boundary of the non-Abelian phase,
1077: as one moves away from this limit into the interior of the confined phase. 
1078: 
1079: \item 
1080: 
1081: $M_S/f_P$ increases monotonically from about 4 to 7, thereby approaching to
1082: within about 35 \% of the value 10.7 in QCD for $M_{a_0}/f_\pi$.
1083: 
1084: \item 
1085: 
1086: $M_V/f_P$ decreases from about 11 to 9, rather close to the value 8.5 for
1087: $M_\rho/f_\pi$ and $M_\omega/f_\pi$ in QCD.  As is evident from
1088: Fig. \ref{m_over_fp}, this ratio $M_V/f_P$ is roughly constant in the upper end
1089: of the interval of $\alpha_*$ values that we study.
1090: 
1091: \item 
1092: 
1093: $M_A/f_P$ behaves non-monotonically, first decreasing from roughly 11.5 to 10,
1094: but then increasing to about 11, within about 20 \% of the average of the
1095: values in QCD for the isospin-triplet and isospin-singlet axial-vector mesons,
1096: 13 for $M_{a_1}/f_\pi$ and 14 for $M_{f_1}/f_\pi$.
1097: 
1098: \item 
1099: 
1100: Thus, the ratios $M_A/M_V$ and $M_S/M_V$, which were found in Ref. \cite{mm} to
1101: have the respective values 1.04 and 0.36 in the walking limit, both increase in
1102: the interval of $\alpha_*$ that we study, reaching about 1.17 and 0.74,
1103: respectively, at $\alpha_*=2.5$.  For comparison, these ratios are
1104: approximately 1.6 and 1.3 in QCD (cf. eqs.  (\ref{ma1mrho_qcd}) and
1105: (\ref{ma0mrho_qcd})).  Although the value of the ratio $M_S/M_V$ at
1106: $\alpha_*=2.5$ is farther from its QCD value than is the case with $M_A/M_V$,
1107: it is increasing somewhat more rapidly as a function of $\alpha_*$, consistent
1108: with eventually approaching the QCD value.
1109: 
1110: \end{enumerate}
1111: 
1112: \section{Conclusions} 
1113: 
1114: In this paper we have considered a vectorial SU($N_c$) gauge theory with $N_f$
1115: massless fermions transforming according to the fundamental representation and
1116: have studied the shift in behavior from walking that occurs in the region near
1117: the boundary between the confinement phase and the non-Abelian Coulomb phase to
1118: the QCD-like behavior with a non-walking coupling.  Specifically, we have used
1119: the Schwinger-Dyson and Bethe-Salpeter equations to calculate the dynamically
1120: induced fermion mass $\Sigma$, the spontaneous chiral symmetry breaking
1121: parameter $f_P$, and the masses of the lowest-lying $q \bar q$ vector,
1122: axial-vector, and flavor-adjoint scalar mesons.  We have investigated how these
1123: change as one decreases $N_f$ below $N_{f,cr}$, or equivalently, increases
1124: $\alpha_*$ above $\alpha_{cr}$, to move away from the above-mentioned boundary
1125: into the interior of the confinement phase.  Our results show the crossover
1126: between walking and non-walking behavior in a gauge theory.
1127: 
1128: There are a number of interesting topics for future research using the methods
1129: of this paper.  It would be useful to construct a kernel for the Bethe-Salpeter
1130: equation that could include more of the relevant physics, including instantons
1131: effects.  Work is underway on this.  It would also be worthwhile to calculate
1132: the masses of radially excited mesons and mesons with internal orbital angular
1133: momenta $L \ge 2$, as well as glueballs and the mixing between glueballs and
1134: $\bar q q$ mesons.  We anticipate that the results of these calculations would
1135: exhibit the same general properties that we have found with the ground-state
1136: $\bar q q $ mesons, but it would be interesting to confirm this expectation
1137: explicitly. Another project would be to connect our study of the region in
1138: $N_f$ where there is a crossover from walking to nonwalking behavior, to the
1139: region around $N_f=2$. However, when one moves this far away from the walking
1140: regime, one loses a simplifying features of our calculation, namely the fact
1141: that we do not have to use an infrared cutoff on $\alpha$.  Given that lattice
1142: gauge theory methods provide an {\it ab initio} framework for calculating
1143: hadron masses, we hope that the lattice community will extend early efforts
1144: such as those of Ref. \cite{mawhinney} and carry out a definitive study of
1145: hadron masses in QCD with an arbitrary number of flavors.  It would be of
1146: considerable interest to compare the results of the lattice calculations with
1147: those obtained from solutions of Schwinger-Dyson and Bethe-Salpeter equations.
1148: 
1149: \acknowledgements
1150: 
1151: This research was partially supported by the grant NSF-PHY-00-98527.
1152: M.K. thanks Profs. M. Harada and K. Yamawaki for the collaboration on the
1153: related Ref. \cite{mm}, and R.S. thanks Dr. Neil Christensen for useful
1154: comments.
1155: 
1156: \bigskip
1157: \bigskip
1158: 
1159: 
1160: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1161: \appendix
1162: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1163: 
1164: \section{Bispinor bases for scalar, pseudoscalar, vector, and 
1165: axial-vector bound states}
1166: \label{app:bispinor-bases}
1167: 
1168: In this appendix we show the explicit forms of the bispinor bases for the
1169: scalar, pseudoscalar, vector, and axial-vector bound states.  Here we use the
1170: notation $\hat{q}_\mu = q_\mu / M_B $ with $M_B$ being the mass of the bound
1171: states, and $[a,b,c] \equiv a[b,c] + b[c,a] + c[a,b]$.
1172: 
1173: %%%%% scalar %%%%%
1174: Bispinor base for the scalar bound state ($J^{PC} = 0^{++}$) 
1175: is given by
1176: %
1177: \beqs
1178:   \Gamma_{(S)}^1 &=&  {\bf 1} ,\ \ 
1179:   \Gamma_{(S)}^2 =  \fsl{p} ,\ \ 
1180:   \Gamma_{(S)}^3 =  \fsl{\hat{q}} (p \cdot \hat{q}) , \nonumber\\ 
1181:   \Gamma_{(S)}^4 &=&  \frac{1}{2}\ [\fsl{p},\fsl{\hat{q}}] ,
1182: \eeqs
1183: %%%%% pseudo-scalar %%%%%
1184: and that for the pseudoscalar bound state ($J^{PC} = 0^{-+}$) 
1185: is given by 
1186: %
1187: \beqs
1188:   \Gamma_{(P)}^1 &=& \gamma_5 ,\ \ 
1189:   \Gamma_{(P)}^2 = \fsl{p} \ (p \cdot \hat{q})\ \gamma_5 ,\ \ 
1190:   \Gamma_{(P)}^3 = \fsl{\hat{q}} \ \gamma_5 , \nonumber\\
1191:   \Gamma_{(P)}^4 &=& \frac{1}{2}\ [\fsl{p},\fsl{\hat{q}}]\ 
1192:                    \gamma_5 \ .
1193: \eeqs
1194: %%%%% vector %%%%%
1195: Furthermore, for the vector bound state ($J^{PC} = 1^{--}$) we use
1196: %
1197: \beqs
1198: &&  \Gamma_{(V)}^1 = \fsl{\epsilon} ,\ \ 
1199:      \Gamma_{(V)}^2 = \frac{1}{2} [\fsl{\epsilon},\fsl{p}] 
1200:                            (p \cdot \hat q) ,\ \ 
1201:      \Gamma_{(V)}^3 = \frac{1}{2} [\fsl{\epsilon},\fsl{\hat q}] ,\nonumber\\
1202: &&     \Gamma_{(V)}^4 = \frac{1}{3!}[\fsl{\epsilon},\fsl{p},\fsl{\hat
1203:      q}] , 
1204:      \Gamma_{(V)}^5 = (\epsilon \cdot p) ,\ \ 
1205:      \Gamma_{(V)}^6 = \fsl{p} (\epsilon \cdot p) ,\nonumber\\
1206: &&   \Gamma_{(V)}^7 = \fsl{\hat q}(p \cdot \hat q) (\epsilon \cdot p) ,\ \ 
1207:      \Gamma_{(V)}^8 = \frac{1}{2} [\fsl{p},\fsl{\hat q}](\epsilon \cdot p) ,
1208: \label{V bases}
1209: \eeqs
1210: %%%%% axial-vector %%%%%
1211: and for the axial-vector bound state  
1212: ($J^{PC} = 1^{++}$) 
1213: %
1214: \beqs
1215: &&   \Gamma_{(A)}^1 = \fsl{\epsilon}\ \gamma_5 ,\ \ \ 
1216:      \Gamma_{(A)}^2 = \frac{1}{2} [\fsl{\epsilon},\fsl{p}] 
1217:                            \gamma_5  ,\ \ \ 
1218:      \Gamma_{(A)}^3 = \frac{1}{2} [\fsl{\epsilon},\fsl{\hat q}]\ 
1219:         (p \cdot \hat q) \ \gamma_5  ,\nonumber\\
1220: &&   \Gamma_{(A)}^4 = \frac{1}{3!}[\fsl{\epsilon},\fsl{p},\fsl{\hat q}]
1221:      \ \gamma_5 ,\ \ \ 
1222:      \Gamma_{(A)}^5 = (\epsilon \cdot p)\ (p \cdot \hat q)\ \gamma_5 ,\nonumber\\
1223: &&   \Gamma_{(A)}^6 = \fsl{p} (\epsilon \cdot p)\ \gamma_5  ,\ \ 
1224:      \Gamma_{(A)}^7 = \fsl{\hat q}\ (\epsilon \cdot p)\ (p \cdot \hat q)
1225:                 \ \gamma_5  ,\nonumber\\
1226: &&   \Gamma_{(A)}^8 = \frac{1}{2} [\fsl{p},\fsl{\hat q}](\epsilon \cdot p)\
1227:      (p \cdot \hat q)\ \gamma_5  .
1228: \eeqs
1229: %
1230: 
1231: 
1232: \begin{thebibliography}{99}
1233: 
1234: 
1235: \bibitem{bz}
1236: An early paper on the phase structure of vectorial gauge theories is 
1237: T. Banks and A. Zaks, Nucl. Phys. B {\bf 196}, 189 (1982). 
1238: 
1239: \bibitem{wtc1}
1240: B. Holdom, Phys. Lett. B {\bf 150}, 301 (1985).
1241: 
1242: \bibitem{wtc2} 
1243: K. Yamawaki, M. Bando, and K. Matumoto, Phys. Rev. Lett. {\bf
1244: 56}, 1335 (1986).
1245: 
1246: \bibitem{chipt1}
1247: T. Appelquist, D. Karabali, and L. C. R. Wijewardhana, Phys. Rev. Lett. {\bf
1248: 57}, 957 (1986); T. Appelquist and L. C. R. Wijewardhana, Phys. Rev. D
1249: {\bf 35}, 774 (1987); Phys. Rev. D {\bf 36}, 568 (1987).
1250: 
1251: \bibitem{chipt2}
1252: T. Appelquist, J. Terning, and L. C. R. Wijewardhana,
1253: Phys. Rev. Lett.  {\bf 77}, 1214 (1996).
1254: 
1255: \bibitem{my}
1256: V. Miransky and K. Yamawaki, Phys. Rev. D {\bf 55}, 5051 (1997); {it ibid.} 
1257: {\bf 56}, E 3768 (1997). See also V. Miransky and P. Fomin,
1258: Sov. J. Part. Nucl. {\bf 16}, 203 (1985). 
1259: 
1260: \bibitem{chipt3} 
1261: T. Appelquist, A. Ratnaweera, J. Terning, and 
1262: L. C. R. Wijewardhana, Phys. Rev. D {\bf 58}, 105017 (1998).
1263: 
1264: \bibitem{mm}
1265: % meson masses in large N_f QCD 
1266: M. Harada, M. Kurachi, and K. Yamawaki, Phys. Rev. D {\bf 68}, 076001 (2003).
1267: 
1268: \bibitem{qm}
1269: See, e.g., F. Close, {\it Introduction to Quarks and Partons} (Academic, New
1270: York, 1979). 
1271: 
1272: \bibitem{mitbag}
1273: A. Chodos, R. Jaffe, K. Johnson, C. Thorn, and V. Weisskopf, Phys. Rev. D 
1274: {\bf 9}, 3471 (1974); T. DeGrand, R. Jaffe, K. Johnson, and J. Kiskis, 
1275: Phys. Rev. D {\bf 12}, 2060 (1975). 
1276: 
1277: \bibitem{mitbag_pwave}
1278: % Excited states of confined quarks, I
1279: T. Degrand and R. Jaffe, Annals Phys. {\bf 100}, 425 (1976). 
1280: 
1281: \bibitem{lat}
1282: For recent reviews, see Lattice 2005, Dublin; 
1283: http://www.maths.tcd.ie/lat05; Lattice 2004, Fermilab, 
1284: http:/lqcd.fnal.gov/lattice04; and earlier lattice field theory symposia. 
1285: 
1286: \bibitem{bs}
1287: E. Salpeter and H. Bethe, Phys. Rev. {\bf 84}, 1232 (1951). 
1288: 
1289: \bibitem{nakanishi}
1290: %``A General Survey Of The Theory Of The Bethe-Salpeter Equation,''
1291: For an early review, see N. Nakanishi, Prog. Theor. Phys. Suppl.  {\bf 43}, 1
1292: (1969).
1293: %%CITATION = PTPSA,43,1;%%
1294: 
1295: \bibitem{mn74}
1296: T. Maskawa and H. Nakajima, Prog. Theor. Phys. {\bf 52}, 1326 (1974);
1297: % Schwinger-Dyson Equation For Massless Vector Theory And Absence Of Fermion
1298: % Pole.
1299: 
1300: \bibitem{kugo}
1301: R. Fukuda and T. Kugo, Nucl. Phys. B {\bf 117}, 250 (1974); 
1302: % Dynamical Instability Of The Vacuum In The Lagrangian Formalism Of The
1303: % Bethe-Salpeter Bound States.
1304: T. Kugo, Phys. Lett. B {\bf 76}, 625 (1978). 
1305: 
1306: \bibitem{lane74}
1307: K. Lane, Phys. Rev. D {\bf 10}, 2605 (1974). 
1308: 
1309: \bibitem{higashijima}
1310: K. Higashijima, Phys. Rev. D {\bf 29}, 1228 (1984). 
1311: 
1312: \bibitem{alm} 
1313: T. Appelquist, K. Lane, and U. Mahanta, Phys. Rev. Lett. {\bf
1314: 61}, 1553 (1988); T. Appelquist, U. Mahanta, D. Nash, and L.C.R.  Wijewardhana,
1315: Phys. Rev. D {\bf 43}, 646 (1991); 
1316: U. Mahanta, Phys. Rev. D {\bf 45}, 1405 (1992).
1317: 
1318: \bibitem{kugorev}
1319: T.~Kugo,
1320: in {\it Proc. of 1991 Nagoya Spring School on Dynamical
1321: Symmetry Breaking, Nakatsugawa, Japan, 1991}, ed. K.~Yamawaki  
1322: (World Scientific, Singapore, 1992).
1323: 
1324: \bibitem{miranskyrev}
1325: V. Miransky, {\it Dynamical Symmetry Breaking in Quantum Field Theories} 
1326: (World Scientific, Singapore, 1993). 
1327: 
1328: \bibitem{HY}
1329: % solving the bethe-salpeter equation 
1330: M. Harada and Y. Yoshida, Phys. Rev. D {\bf 53}, 1482 (1996). 
1331: 
1332: \bibitem{abkmn}
1333: % Calculating the decay constant f(pi).
1334: K.-I. Aoki, M. Bando, T. Kugo, M. Mitchard, and H. Nakatani, 
1335: Prog. Theor. Phys. {\bf 84}, 683 (1990).
1336: 
1337: \bibitem{abkm}
1338: K.-I. Aoki, M. Bando, T. Kugo, and M. Mitchard, 
1339: Prog. Theor. Phys. {\bf 85}, 355 (1991).
1340: 
1341: \bibitem{akm}
1342: % Meson properties from the ladder Bethe-Salpeter equation.
1343: K.-I. Aoki, T. Kugo, and M. Mitchard, Phys. Lett. B {\bf 266}, 467 (1991). 
1344: 
1345: \bibitem{jainmunczek93}
1346: % qqbar bound states in the bethe-salpeter formalism, incl. analytic solutions
1347: P. Jain and H. Munczek, Phys. Rev. D {\bf 48}, 5403 (1993). 
1348: 
1349: % ground-state spectrum of light-quark mesons 
1350: \bibitem{robertsmesons}
1351: C. J. Burden et al., Phys. Rev. C {\bf 55}, 2649 (1997). 
1352: 
1353: % review 
1354: \bibitem{avsrev}
1355: R. Alkofer and L. von Smekal, Phys. Rept. {\bf 353}, 281 (2001). 
1356: 
1357: % review 
1358: \bibitem{marisroberts03}
1359: P. Maris and C. D. Roberts, Int. J. Mod. Phys. E {\bf 12}, 297 (2003).
1360: 
1361: \bibitem{pimdif}
1362: % pi+ pi0 mass difference
1363: M. Harada, M. Kurachi, and K. Yamawaki, Phys. Rev. D {\bf 70}, 033009 (2004).
1364: 
1365: \bibitem{HKY}
1366: % S parameter 
1367: M. Harada, M. Kurachi, and K. Yamawaki, Prog. Theor. Phys. {\bf 115}, 765 
1368: (2006) (hep-ph/0509193). 
1369: 
1370: \bibitem{integer}
1371: %
1372: Here and below, when we mention non-integral values of $N_f$, it is implicitly
1373: understood that physical values of $N_F$ are, of course, positive integers, and
1374: the non-integral values are defined via an analytic continuation away from
1375: these physical values. 
1376: 
1377: \bibitem{b0}
1378: D. Gross and F. Wilczek, Phys. Rev. Lett. {\bf 30}, 1343 (1973); 
1379: H. D. Politzer, Phys. Rev. lett. {\bf 30}, 1346 (1973); G. 't Hooft,
1380: unpublished. 
1381: 
1382: \bibitem{b1}
1383: W. Caswell, Phys. Rev. Lett. {\bf 33}, 244 (1974); D. R. T. Jones, Nucl. Phys. 
1384: B {\bf 75}, 531 (1974). 
1385: 
1386: \bibitem{casimir} The Casimir invariant $C_2(R)$ of the representation $R$ is
1387: defined by ${\cal D}_R(T_a)^i_j{\cal D}_R(T_a)^j_k = C_2(R)\delta^i_k$, where
1388: $\{a,b\}$ and $\{i,j,k\}$ denote group and representation indices and sums over
1389: repeated indices are understood.
1390: 
1391: \bibitem{instantons}
1392: D. Caldi, Phys. Rev. Lett. {\bf 39}, 121 (1977); C. Callan, R. Dashen, and 
1393: D. Gross, Phys. Rev. D {\bf 17}, 2717 (1978); T. Appelquist and S. Selipsky, 
1394: Phys. Lett. B {\bf 400}, 364 (1997).
1395: 
1396: \bibitem{iwasaki} 
1397: Y.  Iwasaki et al., Phys. Rev. Lett. {\bf 69}, 21 (1992); 
1398: Phys. Rev. D {\bf 69}, 014507 (2004).
1399: 
1400: \bibitem{heller}
1401: P. Damgaard, U. Heller, A. Krasnitz, and P. Olesen, Phys. Lett. B {\bf 400},
1402: 169 (1997). 
1403: 
1404: \bibitem{mawhinney}
1405: %The QCD hadron spectrum and the number of dynamical quark flavors 
1406: R.  Mawhinney, Nucl. Phys. B (Proc. Suppl.) {\bf 63A-C}, 212 (1998); 
1407: %review of unquenched results at the PISA lattice conf. 
1408: R.  Mawhinney, Nucl. Phys. B (Proc. Suppl.) {\bf 83}, 57 (2000). 
1409: 
1410: \bibitem{fpi} 
1411: It has been estimated \cite{gl85} that $f_\pi/(f_\pi)_{ch.lim.} \simeq 1.06$,
1412: where $(f_\pi)_{ch.lim.}$ denotes the value of $f_\pi$ in the chiral 
1413: limit $m_u=m_d=0$.  With $f_\pi=92.4$, this gives $(f_\pi)_{ch.lim.} \simeq 
1414: 87$ MeV. 
1415: 
1416: \bibitem{gl85}
1417: J. Gasser and H. Leutwyler, Nucl. Phys. B {\bf 250}, 465 (1985). See also 
1418: Ref. \cite{chipt}. 
1419: 
1420: \bibitem{chipt} 
1421: J. Gasser and H. Leutwyler, Phys. Rept. C {\bf 87}, 77 (1982);
1422: Nucl. Phys. B {\bf 250}, 517 (1985); G. Colangelo, J. Gasser, and H. Leutwyler,
1423: Nucl. Phys. B {\bf 603}, 125 (2001); M. Harada and K. Yamawaki,
1424: Phys. Rept. {\bf 381},1 (2004); U.-G. Meissner, hep-ph/0501009.
1425: 
1426: \bibitem{pdg}
1427: For a compendium of relevant data and references, see http://pdg.lbl.gov. 
1428: 
1429: \bibitem{psrel}
1430: H. Pagels and S. Stokar, Phys. Rev. D {\bf 20}, 2947 (1979).  
1431: 
1432: \bibitem{schechter}
1433: D. Black, A. Fariborz, F. Sannino, and J. Schechter, 
1434: Phys. Rev. D {\bf 59}, 074026 (1999); 
1435: A. Fariborz, R. Jora, and J. Schechter, hep-ph/0601216. 
1436: 
1437: \bibitem{wein}
1438: S. Weinberg, Phys. Rev. Lett. {\bf 18}, 507 (1967). 
1439: 
1440: \bibitem{sdcomplex}
1441: The Schwinger-Dyson equation with variables in the complex plane was discussed
1442: by T. Kugo and Y. Yoshida, Soryushiron Kenkyu {\bf 91}, B26 (1995).  See also
1443: Ref. \cite{mm} for more a detailed discussion. 
1444: 
1445: \bibitem{Gardi}
1446: E.~Gardi and M.~Karliner, Nucl.\ Phys.\ B {\bf 529}, 383 (1998);
1447: E.~Gardi, G.~Grunberg and M.~Karliner, JHEP {\bf 9807}, 007 (1998). 
1448: 
1449: \end{thebibliography}
1450: 
1451: \end{document}
1452: 
1453: