hep-ph0607231/s.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{Behavior of the $S$ Parameter in the Crossover Region Between Walking 
45: and QCD-Like Regimes of an SU($N$) Gauge Theory} 
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 consider a vectorial, confining SU($N$) gauge theory with a variable number,
61: $N_f$, of massless fermions transforming according to the fundamental
62: representation.  Using the Schwinger-Dyson and Bethe-Salpeter equations, we
63: calculate the $S$ parameter in terms of the current-current correlation
64: functions. We focus on values of $N_f$ such that the theory is in the crossover
65: region between the regimes of walking behavior and QCD-like (non-walking)
66: behavior.  Our calculations indicate that the contribution to $S$ from a given
67: fermion decreases as one moves from the QCD-like to the walking regimes.  The
68: implications of this result for technicolor theories are discussed.
69: 
70: \end{abstract}
71: 
72: \pacs{PACS No.}%{14.60.PQ, 12.60.Nz, 14.60.St}
73: 
74: %\keywords{} 
75: 
76: \maketitle
77: 
78: \vspace{16mm}
79: 
80: \newpage
81: \pagestyle{plain}
82: \pagenumbering{arabic}
83: 
84: 
85: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
86: \section{Introduction}
87: \label{sec:Introduction}
88: 
89: 
90: The properties of a vectorial gauge theory as a function of the fermion content
91: are of fundamental importance.  Here we consider such a theory (in $(3+1)$
92: dimensions at zero temperature and chemical potential) with gauge group SU($N$)
93: and $N_f$ massless fermions transforming according to the fundamental
94: representation of this group.  For $N = N_c=3$ and $N_f=2$, this is an
95: approximation to actual quantum chromodynamics (QCD) with just the $u$ and $d$
96: quarks, since their current-quark masses are much smaller than the scale
97: $\Lambda_{QCD} \simeq 400$ MeV.  We restrict here to the range $N_f < (11/2)N$
98: for which the theory is asymptotically free.  An analysis using the two-loop
99: beta function and Schwinger-Dyson equation (reviewed below) leads to the
100: inference that for $N_f$ in this range, the theory includes two phases: (i) for
101: $0 \le N_f \le N_{f,cr}$ a phase with confinement and spontaneous chiral
102: symmetry breaking (S$\chi$SB), and (ii) for $N_{f,cr} \le N_f \le (11/2)N$ a
103: non-Abelian Coulomb phase with no confinement or spontaneous chiral symmetry
104: breaking. We shall refer to $N_{f,cr}$, the critical value of $N_f$, as the
105: boundary of the non-Abelian Coulomb (conformal) phase \cite{bz}.
106: 
107: For $N_f$ slightly less than $N_{f,cr}$, the theory exhibits an approximate
108: infrared (IR) fixed point. Let the SU($N$) running gauge coupling be denoted as
109: $g(\mu)$, where $\mu$ denotes the energy or momentum scale, and let
110: $\alpha(\mu) =\bar g(\mu)^2/(4\pi)$. As $\mu$ decreases from large values,
111: $\alpha(\mu)$ grows to be O(1) at a scale $\Lambda$, but increases only rather
112: slowly as $\mu$ decreases below $\Lambda$, so that there is an extended
113: interval in energy below $\Lambda$ where $\alpha$ is large, but slowly running
114: (``walking'').  In addition to its intrinsic field-theoretic interest, this
115: walking behavior is an essential ingredient of modern technicolor models of
116: dynamical electroweak symmetry breaking \cite{tc}, providing the requisite
117: enhancement of Standard-Model (SM) fermion masses \cite{wtc1}-\cite{chipt3}.
118: As $N_f$ approaches $N_{f,cr}$ from below, quantities with dimensions of mass
119: vanish continuously; i.e., the chiral phase transition separating phases (i)
120: and (ii) is continuous.
121:  
122: 
123: In this paper, we shall use solutions of the Schwinger-Dyson (SD) and
124: Bethe-Salpeter (BS) equations to compute a derivative of the difference of the
125: vector and axial-vector current-current correlation functions,
126: $\Pi'_{VV}(0)-\Pi'_{AA}(0)$.  Up to a multiplicative factor, this is the
127: coefficient $\bar L_{10}$ of one of the terms in the effective chiral
128: Lagrangian for the theory \cite{gl85,chipt}.  Moreover, in the context in which
129: one considers the SU($N$) theory as a technicolor (TC) model, with $N=N_{TC}$,
130: the above quantity is proportional to the correction to the $Z$ propagator due
131: to virtual electroweak-nonsinglet technicolor particles, often denoted as $S$
132: \cite{pt,ab}, \cite{pdg,lepewwg}. We focus on the crossover region between the
133: walking regime that occurs for $N_f \lsim N_{f,cr}$ and the QCD-like
134: (non-walking) regime that occurs for smaller $N_f$.
135: 
136: There are several motivations for this work. The quantity $S$, or equivalently
137: $\bar L_{10}$, is an intrinsic property of the SU($N$) theory, and it is of
138: interest to understand how this quantity depends on $N_f$. Further, our
139: calculations have important implications for technicolor models of dynamical
140: electroweak symmetry breaking \cite{wtc1}-\cite{chipt3}. For this application,
141: as noted, we identify the SU($N$) group with the technicolor gauge group.
142: Precision electroweak data \cite{pdg,lepewwg} determine allowed regions of
143: values of $S$ and the other $Z$ and $W$ propagator corrections denoted $T$ and
144: $U$, and yield a stringent constraint on the contributions from new particles
145: in technicolor models.  In order to assess the viability of these models, it is
146: necessary to have a reliable calculation of the contribution to $S$ from
147: technicolor particles.  A perturbative calculation of $S$ is not reliable since
148: the technifermions are strongly interacting on the relevant scale $\sim m_Z$.
149: Although it is possible to carry out a nonperturbative calculation for
150: technicolor theories that behave like scaled-up QCD (by using QCD data as input
151: for the relevant spectral functions), such theories are excluded since they
152: cannot produce sufficiently large standard-model fermion masses.  It is more
153: difficult to carry out a nonperturbative estimate of $S$ for technicolor models
154: that have walking behavior; such estimates have been presented in
155: Refs. \cite{ats}-\cite{iwts}.  These suggest that the contribution to $S$ per
156: technifermion electroweak doublet is reduced in the walking region. 
157: Meson masses and $f_P$ (the generalization of $f_\pi$) were
158: calculated in this walking regime in Ref. \cite{mmw}. (See also Ref. \cite{mme}
159: for the analogous calculations in QCD.)  A motivation for our
160: calculations in this paper is to gain further insight into the behavior of $S$
161: by studying its behavior in the crossover region between the walking and
162: QCD-like regimes. One of the reasons for concentrating on this crossover region
163: is that the theory still has an approximate infrared fixed point here, and
164: hence one can study the dependence of $S$ on $N_f$ without having to
165: introduce a model-dependent cutoff on the growth of the SU($N$) gauge coupling
166: in the infrared that was required in calculations of $S$ via the SD and BS
167: equations for small $N_f$ values, such as $N_f=3$ in QCD \cite{hys,pms}.  As
168: our previous calculations of meson masses and $f_P$ in this crossover region
169: showed, \cite{sg}, although it is a restricted interval in $N_f$ or
170: equivalently, the value of the infrared fixed point, it is sufficient large to
171: observe a significant change between walking and QCD-like behavior. 
172: 
173: This paper is organized as follows. Section II is devoted to a review some
174: background material concerning the beta function, approximate infrared fixed
175: point and walking behavior, and technicolor models. Section III contains
176: definitions of the current-current correlation functions and related spectral
177: functions together with the expression for $S$ in terms of these correlation
178: functions and an equivalent integral of spectral functions.  In this section we
179: also give the definition of $\bar L_{10}$ as a coefficient of a certain
180: operator in the low-energy effective chiral Lagrangian for the theory.  Section
181: IV explains how the current correlators are obtained from Bethe-Salpeter
182: amplitudes while Section V discusses the method of solution of the
183: Bethe-Salpeter equation.  Our results are presented in Section VI and their
184: implications for technicolor theories in Section VII.  An appendix discusses
185: some analytic approximations.
186: 
187: 
188: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
189: \section{Some Preliminaries} 
190: \label{sec:large_Nf}
191: 
192: In this section we review some background for our calculations.  For the theory
193: under consideration, with an SU($N$) gauge group and $N_f$ massless fermions in
194: the fundamental representation, the renormalization group (RG) equation for the
195: running coupling $\alpha(\mu)$ is 
196: %
197: \beq
198: \beta = \mu \frac{d \alpha(\mu)}{d \mu} = - \frac{\alpha(\mu)^2}{2\pi}
199: \left ( b_0 + \frac{b_1}{4\pi}\alpha + O(\alpha^2) \right ) \ ,
200: \label{eq:RGE_for_alpha}
201: \eeq
202: %
203: where $\mu$ is the momentum scale. The two terms listed are scheme-independent.
204: The next two higher-order terms have also been calculated but are
205: scheme-dependent; their inclusion does not significantly affect our results.
206: For the relevant case of an asymptotically free theory, $b_0 > 0$ so that an
207: infrared fixed point exists if and only if $b_1 < 0$.  This coefficient $b_1$
208: is positive for $0 \le N_f \le N_{f,IR}$, where $N_{f,IR}=(34N^3)/(13N^2-3)$,
209: and negative for larger $N_f$.  For $N=3$, $N_{f,IR} \simeq 8.1$
210: \cite{integer}.  The value of $\alpha$ at this IR fixed point, denoted
211: $\alpha_*$, is given by $\alpha_* = -4\pi b_0/b_1$.  Substituting the known
212: values of these terms \cite{b0,b1}, one has
213: %
214: \beq
215: \alpha_* = \frac{-4\pi(11N -2N_f)}{34N^2-13N N_f+3N^{-1}N_f} \ .
216: \label{alpha_irfp}
217: \eeq
218: %
219: Solving Eq. (\ref{alpha_irfp}) for $N_f$ in terms of $\alpha_*$ yields
220: %
221: \beq
222: N_f=\frac{2N^2[17N(\alpha_*/\pi)+22]}{(13N^2-3)(\alpha_*/\pi)+8N} \ .
223: \label{nfsol}
224: \eeq
225: %
226: As is evident from Eqs. (\ref{alpha_irfp}) and (\ref{nfsol}), $\alpha_*$ is a
227: monotonically decreasing function of $N_f$ and equivalently $N_f$ is a
228: monotonically decreasing function of $\alpha_*$, for $N_{f,IR} \le N_f \le
229: (11/2)N$. 
230: 
231: To study the dependence of $S$ on $N_f$, what we actually vary is the value of
232: the approximate IR fixed point $\alpha_*$, which depends parametrically on
233: $N_f$.  For definiteness, we shall take $N=3$; however, as will be seen,
234: $N$ only enters indirectly, via the dependence of the value of the infrared
235: fixed point $\alpha_*$ (Eq. (\ref{alfcrit}) below) on $N_c$.  Hence, our
236: findings may also be applied in a straightforward way, with appropriate changes
237: in the value of $\alpha_*$, to an SU($N$) gauge theory with a different value
238: of $N$.
239: 
240: 
241: In the one-gluon exchange approximation, the Schwinger-Dyson gap equation for
242: the inverse propagator of a fermion transforming according to the
243: representation $R$ of SU($N$) has a nonzero solution for the dynamically
244: generated fermion mass, which is an order parameter for spontaneous chiral
245: symmetry breaking, if $\alpha \ge \alpha_{cr}$, where $\alpha_{cr}$ is given by
246: %
247: \beq
248: \frac{3 \alpha_{cr} C_2(R)}{\pi} = 1,
249: \label{alfcritcondition}
250: \eeq
251: %
252: and $C_2(R)$ denotes the quadratic Casimir invariant for the representation $R$
253: \cite{casimir}.  Using
254: %
255: \beq
256: C_2(fund.) \equiv C_{2f}=\frac{N^2-1}{2N}
257: \label{c2f}
258: \eeq
259: for the fundamental representation yields
260: %
261: \beq
262: \alpha_{cr} = \frac{2\pi N}{3(N^2-1)} \ .
263: \label{alfcrit}
264: \eeq
265: %
266: For the case $N=3$ that we use for definiteness here, Eq. (\ref{alfcrit})
267: gives $\alpha_{cr} = \pi/4 \simeq 0.79$.  To estimate $N_{f,cr}$, one solves
268: the equation $\alpha_* = \alpha_{cr}$, yielding the result \cite{chipt3}
269: %
270: \beq
271: N_{f,cr} = \frac{2N(50N^2-33)}{5(5N^2-3)} \ .
272: \label{nfcr}
273: \eeq
274: %
275: For the values $N=3$ and $N=2$ this gives $N_{f,cr} \simeq 11.9$ and $N_{f,cr}
276: \simeq 7.9$, respectively.  These estimates are only rough, in view of the
277: strongly coupled nature of the physics. Effects of higher-order gluon 
278: exchanges have been studied in Ref. \cite{alm}.  In principle, lattice gauge
279: simulations should provide a way to determine $N_{f,cr}$, but the groups that
280: have studied this have not reached a consensus \cite{lgt}.
281: 
282: 
283: If $\alpha_*$ is less than the critical value, $\alpha_{cr}$, for a bilinear
284: fermion condensate to form, the above IR fixed point is exact, with the
285: coupling $\alpha$ approaching $\alpha_*$ from below as the energy scale $E$
286: decreases from large values to zero.  Let us denote the fermions as $f_i^a$
287: with $a=1,...,N$ and $i=1,...,N_f$.  If $\alpha_* > \alpha_{cr}$, as $E$
288: descends from large values, the coupling $\alpha$ eventually exceeds the above
289: critical value, the fermion condensates $\langle \sum_{a=1}^{N} \bar f_{a,i}
290: f^a_i \rangle$ (with no sum on $i$) form, and are equal for each flavor
291: $i=1,...,N_f$ (with electroweak interactions negligibly small relative to the
292: SU($N$) interaction).  Accordingly, the global ${\rm SU}(N_f)_L \times {\rm
293: SU}(N_f)_R \times {\rm U}(1)_V$ symmetry (where U(1)$_V$ is fermion number) is
294: broken to generalized isospin times fermion number, ${\rm SU}(N_f)_V \times
295: U(1)_V$.
296: 
297: Associated with this, the fermions pick up dynamical masses $\Sigma$ and are
298: integrated out as the energy scale decreases below $\Sigma$.  Hence, in this
299: case, the IR fixed point is only approximate since in the effective field
300: theory for energies below $\Sigma$, the form of the beta function is that for
301: the pure gauge theory, $N_f=0$.  The case where $\alpha_*$ is close to, and
302: slightly larger than, $\alpha_{cr}$, yields walking behavior.  In the strong
303: walking regime, the dynamical fermion mass, and also hadron masses are
304: exponentially smaller than the scale $\Lambda$ at which the coupling first
305: becomes O(1) as the energy scale decreases from large values.  Although our SD
306: and BS equations are semi-perturbative, the analysis is self-consistent in the
307: sense that our $\alpha_{cr}$ really is the value at which, in our
308: approximation, one passes from the confinement phase to the non-Abelian Coulomb
309: phase, and our values of $\alpha$ do span the interval over which there is a
310: crossover from walking to QCD-like (i.e., non-walking) behavior.
311: 
312: As is evident from the above results, decreasing $N_f$ below $N_{f,cr}$ has the
313: effect of increasing $\alpha_*$ and thus moving the theory deeper in the phase
314: with confinement and spontaneous chiral symmetry breaking, away from the
315: boundary with the non-Abelian Coulomb phase.  This is the key parametric
316: dependence that we shall use for our study.  In Refs. \cite{mmw,pmsw} the range
317: of $\alpha_*$ used for the calculation of meson masses was chosen to be $0.89
318: \le \alpha_* \le 1.0$, an interval where there is pronounced walking behavior.
319: For the case $N=3$ considered in Ref. \cite{pmsw} and here, given the
320: above-mentioned value, $\alpha_{cr}=\pi/4$, it follows that this lower limit,
321: $\alpha_*=0.89$, is about 12 \% greater than this critical coupling.  The
322: reason for this choice of lower limit on $\alpha_*$ was that the calculation of
323: $S$ involves very strong cancellations as $\alpha_* - \alpha_{cr} \to 0^+$,
324: rendering it progressively more and more difficult to obtain accurate numerical
325: results in this extreme walking limit.  For our present study of $S$ we
326: consider an interval extending to larger couplings, from $\alpha_*=1.0$ to
327: $\alpha_*=1.8$.  Our upper limit is chosen in order for the ladder
328: approximation used in our solutions of the Schwinger-Dyson and Bethe-Salpeter
329: equations to have reasonable reliability.  From Eq. (\ref{nfsol}) it follows
330: that $\alpha_*=0.89$ corresponds to $N_f=11.65$, about 2 \% less than
331: $N_{f,cr}$.  For a coupling as large as $\alpha_* = 1.8$, the semi-perturbative
332: methods used to derive eqs. (\ref{alpha_irfp}) and (\ref{nfsol}) are subject to
333: large corrections from higher-order perturbative, and from nonperturbative,
334: contributions; recognizing this, the above upper limit of $\alpha_*$
335: corresponds to $N_f \simeq 10.3$, a roughly 13 \% reduction from
336: $N_{f,cr}=11.9$.  Although this shift appears to be by only a modest amount
337: when expressed in terms of $N_f$, in terms of $\alpha_*$ it is a factor of
338: two, and our calculations in Ref. \cite{sg} showed a dramatic change in the
339: values of meson mass ratios and $f_P/\Lambda$ in this range, with these values
340: changing from their walking limits toward QCD-like values.  Hence we anticipate
341: that this range can be sufficient to study the shift in the value of $S$, and
342: our results confirm this.
343: 
344: If $N_f > N_{f,cr}$, i.e., $\alpha_* < \alpha_{cr}$ so that this IR fixed point
345: of the two-loop RG equation is exact, then, denoting $b \equiv b_0/(2\pi)$, the
346: solution to this equation can be explicitly
347: written~\cite{Gardi,ExplicitSolution} in the entire energy region as
348: %
349: \beq
350: \alpha(\mu)=\alpha_\ast \left[\ W(e^{-1}(\mu/\Lambda)^{b \alpha_*})+1 \ 
351: \right]^{-1},
352: \label{Lambert}
353: \eeq
354: %
355: where $W(x) = F^{-1}(x)$, with $F(x) = x e^x$, is the Lambert $W$
356: function, and $\Lambda$ is a RG-invariant
357: scale defined by~\cite{chipt2}
358: %
359: \beq
360:  \Lambda \ \equiv\  \mu \ \exp \left[ -\frac{1}{b} \left \{ 
361: \frac{1}{\alpha_\ast}
362: \ln\left( \frac{\alpha_\ast - \alpha(\mu)}{\alpha(\mu)} \right)
363:         + \frac{1}{\alpha(\mu)} \right \} \right] .
364: \label{eq:Lambda}
365: \eeq 
366: %
367: Now since we are studying the confined phase with $N_f < N_{f,cr}$, ($\alpha_*
368: > \alpha_{cr}$) with spontaneous chiral symmetry breaking, $\alpha_*$ is only
369: an approximate, rather than exact, IR fixed point.  Hence, the solution
370: (\ref{Lambert}) is only applicable in an approximate manner to our case; for
371: momenta much less than the dynamical fermion mass $\Sigma$, the fermions
372: decouple, and in this very low-momentum region, with the fermions integrated
373: out, the resultant $\alpha$ would increase above the value $\alpha_*$ at the
374: approximate IR fixed point.  However, since $\Sigma \ll \Lambda$ in a walking
375: or near-walking theory, it follows that this lowest range of momenta makes a
376: small contribution to the relevant integrals to be evaluated in our
377: calculations.  Hence, over most of the integration range for these integrals
378: where the coupling $\alpha$ is large, it is approximately constant and equal to
379: its fixed-point value, $\alpha_*$ (see Fig. 2 of Ref. \cite{mmw}).  This means
380: that one can use, as a reasonable approximation, the expression 
381: %
382: \beq
383:   \alpha(\mu) = \alpha_\ast \, \theta(\Lambda-\mu),
384: \label{run_approx}
385: \eeq
386: %
387: where $\theta$ is the step function.  (This is the same approximation used in
388: Refs. \cite{sg,mmw}.)  Thus, in the walking region and the adjacent
389: crossover region, the calculations have the advantage that one can avoid having
390: to introduce an artificial cutoff on the growth of $\alpha$ in the infrared, in
391: contrast to the situation for smaller $N_f$, where the walking behavior
392: disappears and this cutoff is necessary. 
393: 
394: Since an important application of our results is to technicolor models, we
395: briefly mention some relevant features of these models.  As noted, the
396: technicolor gauge theory has a gauge group SU($N_{TC}$) and an asysmptotically
397: free coupling that gets large at the TeV scale \cite{tc}.  It contains a set of
398: massless, vectorially coupled technifermions.  The left-handed components of
399: these fermions transform as doublets under SU(2)$_L$.  The spontaneous chiral
400: symmetry breaking and formation of a bilinear technifermion condensate breaks
401: the electroweak symmetry from ${\rm SU}(2)_L \times {\rm U}(1)_Y$ to
402: U(1)$_{em}$, producing masses for the $W$ and $Z$ given to leading order by
403: $m_W^2 = m_Z^2 \cos^2 \theta_W = (g^2/4)N_{D,TF} F_{TC}^2$, where $F_{TC}$ is
404: the technicolor analogue of $f_\pi$. In order to give masses to the
405: Standard-Model fermions (which are technisinglets), it is necessary to embed
406: technicolor in a larger, extended technicolor (ETC) theory \cite{etc,tcm}, with
407: interactions that transform technifermions to the Standard-Model fermions and
408: vice versa.  To satisfy constraints from flavor-changing neutral-current (FCNC)
409: processes, the ETC vector bosons that mediate generation-changing transitions
410: must have large masses, ranging from a few TeV to $10^3$ TeV.  For our present
411: study, concerned with $S$, we concentrate on the technicolor theory at the
412: scale of a few hundred GeV, with the ETC gauge bosons integrated out.
413: 
414: We focus here on models in which the technifermions transform according to the
415: fundamental representation of the SU($N_{TC}$) gauge group.  Two simple
416: examples are the so-called one-doublet and one-family technicolor models.  A
417: one-doublet TC model has $N_f=2$ technifermions, denoted $U$ and $D$, whose
418: chiral components transform according to
419: %
420: \beqs
421: & & F_L = {U_L \choose D_L} \ : \ (N_{TC},1,2)_{0,L}, \cr\cr
422: & & U_R \ : \ \ (N_{TC},1,1)_{1,R}, \cr\cr
423: & & D_R \ : \ \ (N_{TC},1,1)_{-1,R} \ . 
424: \label{1d}
425: \eeqs
426: %
427: where the numbers in parentheses refer to the dimensions of the representations
428: of ${\rm SU}(N_{TC}) \times {\rm SU}(3)_c \times {\rm SU}(2)_L$ and the
429: subscripts refer to the weak U(1)$_Y$ hypercharges.  The value $N_{TC}=2$ has
430: been preferred in recent TC/ETC model-building \cite{at94,as} for several
431: reasons, including the fact that it (i) minimizes technicolor contributions to
432: the $S$ parameter, (ii) can naturally produce a walking theory in a one-family
433: model (see below), and (iii) makes possible a mechanism to explain
434: light neutrino masses \cite{as}.  The one-doublet TC model has one
435: SU(2)$_L$ doublet of technifermions for each technicolor index, which we
436: express as $N_{D,TF}=1$, and hence a total number of SU(2)$_L$ doublets
437: of technifermions equal to $N_{D,tot}=N_{D,TF}N_{TC}=N_{TC}$.  The TC sector
438: with just these $N_f=2$ technifermions would not exhibit walking behavior, but
439: one can add SM-singlet technifermions to produce a theory that does have such
440: behavior \cite{ts}.  In a one-family TC model the technifermions transform as
441: %
442: \beqs
443: & & Q_L: \ (N,3,2)_{1/3,L} \cr\cr
444: & & u_R: \ (N,3,1)_{4/3,R} \cr\cr
445: & & d_R: \ (N,3,1)_{-2/3,R} \cr\cr
446: & & L_L: \ (N,1,2)_{-1,L} \cr\cr
447: & & N_R: \ (N,1,1)_{0,R} \cr\cr
448: & & E_R: \ (N,1,1)_{-2,R}
449: \label{1fam_quarks}
450: \eeqs
451: %
452: Hence, this type of technicolor models contains $N_f=2(N_c+1)=8$
453: technifermions. As is evident from Eq.  (\ref{nfcr}), with $N=N_{TC}=2$, the
454: value $N_f=8$ is close to the value $N_{f,cr}$ and hence, to within the
455: accuracy of the two-loop beta function analysis, this technicolor model can
456: naturally exhibit walking behavior.  Reverting to general $N=N_{TC}$ for our
457: discussion, this one-family technicolor model thus has $N_{D,TF}=(N_c+1)=4$
458: \ SU(2)$_L$ doublets for each technicolor index, and hence a total of 
459: $N_{D,tot.}=4N_{TC}$ SU(2)$_L$ doublets of technifermions. 
460: 
461: Because of the spontaneous chiral symmetry breaking in the technicolor theory,
462: the technifermions pick up dynamical masses $\Sigma_{TC}$ proportional to
463: $F_{TC} N_{TC}^{-1/2}$, where we have included the $N_{TC}$-dependent factor
464: that would be present in the large-$N_{TC}$ limit, since $f_P$ and
465: $\Sigma_{TC}$ scale, respectively, like $N_{TC}^{1/2}$ and $N_{TC}^0$ in this
466: limit.  For the one-doublet and one-family TC models, $F_{TC} \simeq 250$ and
467: 125 GeV, respectively.  In QCD, the constituent quark mass $\Sigma \simeq 3.5
468: f_\pi$, and one expects a roughly similar ratio in TC theories (see Fig. 3 in
469: our previous work \cite{sg}).  Since the SM gauge couplings are small at the
470: technicolor scale, different technifermions are expected to have roughly
471: degenerate dynamical masses, and the contributions of the techniquark and
472: technilepton doublets to one-loop corrections to the $Z$ propagator are
473: approximately equal.
474: 
475: 
476: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
477: 
478: \section{Expression for $S$ in Terms of Current-Current Correlation 
479: Functions}
480: \label{sec:Spectral}
481: 
482: 
483: As a measure of corrections to the $Z$ propagator arising from heavy particles
484: in theories beyond the Standard Model, $S$ was originally defined as 
485: \cite{pt}
486: %
487: \beq
488: S = \frac{4 s_W^2 c_W^2}{\alpha_{em}(m_Z)} \left. \frac{d \Pi^{(NP)}_{ZZ}(q^2)}
489: {dq^2} \right\vert_{q^2 = 0} 
490: \label{s}
491: \eeq
492: %
493: where $s_W^2 = 1-c_W^2 = \sin^2\theta_W$, evaluated at $m_Z$ and the
494: superscript $NP$ refers to the fact that the definition includes new physics
495: beyond the Standard Model.  In the case of technicolor, the technifermions 
496: are taken to have zero masses; because of the spontaneous chiral symmetry 
497: breaking in the TC theory, they pick up dynamical masses $\Sigma_{TC}$ of order
498: the technicolor scale.  More recent analyses of precision electroweak data
499: define $S$ slightly differently, replacing the derivative at $q^2=0$ by a
500: discrete difference (in the $\overline{MS}$ scheme) \cite{pdg} 
501: %
502: \beq
503: S_{PDG} = \frac{4 s_W^2 c_W^2}{\alpha_{em}(m_Z)} \Bigg [ 
504: \frac{\Pi^{(NP)}_{ZZ}(m_Z^2) - \Pi^{(NP)}_{ZZ}(0)}{m_Z^2} \Bigg ] \ . 
505: \label{spdg}
506: \eeq
507: %
508: The difference between these definitions is small if the heavy physics scale
509: $\Sigma_{TC}$ satisfies $(2\Sigma_{TC}/m_Z)^2 \gg 1$, as is the case in the
510: TC models considered here. 
511: 
512: 
513: For our purposes it will be convenient to use the original definition, Eq.
514: (\ref{s}).  The implications of our results for technicolor theories would be
515: essentially the same if we used the expression (\ref{spdg}).  With either
516: definition, since a one-loop heavy fermion correction to the $Z$ propagator has
517: a prefactor $(g^2+g'^2)/(16\pi^2)$, where $g$ and $g'$ are the respective
518: SU(2)$_L$ and U(1)$_Y$ gauge couplings, and since $(g^2+g'^2)/(16\pi^2)=
519: \alpha_{em}/(4\pi s_W^2 c_W^2)$, the prefactor $4 s_W^2 c_W^2/\alpha_{em}(m_Z)$
520: in the definition of $S$ cancels out the leading dependence on the electroweak
521: gauge couplings (evaluated at the scale $m_Z$), yielding a quantity that
522: depends on the intrinsic properties of the strongly coupled SU($N$) gauge
523: theory.
524: 
525: 
526: Now, suppressing the SU($N_{TC}$) gauge index, we write the fermions as a
527: vector, $\psi = (\psi_i,...,\psi_{N_f})$. We then define vector and the
528: axial-vector currents as
529: %
530: \beqs
531:    V_\mu^a(x) &=& \bar\psi(x) T^a \gamma_\mu \psi(x) \cr\cr
532:    A_\mu^a(x) &=& \bar\psi(x) T^a \gamma_\mu \gamma_5\psi(x) \, ,
533:   \label{def:currents}
534: \label{vacurrents}
535: \eeqs
536: %
537: where the $N_f \times N_f$ matrices $T^a$ ($a=1,..., N_f^2-1$) are the
538: generators of $SU(N_f)$ with the standard normalization ${\rm Tr}(T^a T^b) =
539: \frac{1}{2}\delta^{ab}$.  In terms of these currents, the two-point
540: current-current correlation functions $\Pi_{VV}$ and $\Pi_{AA}$ are defined via
541: the equations
542: %
543: \beqs 
544: && i\int d^4x\ e^{i q \cdot x}\ \langle 0|T(J_\mu^a(x) J_\nu^b(0))|0 \rangle \
545: \cr\cr
546: &=& \delta^{ab}\left(\frac{q_\mu q_\nu}{q^2}-g_{\mu\nu}\right)\Pi_{JJ}(q^2) \ ,
547: \label{pijjdef}
548: \eeqs
549: %
550: where $J_\mu^a(x) = V_\mu^a(x), A_\mu^a(x)$. With the above normalization of
551: $T^a$, $\Pi(q^2)$ measures the contributions to the time-ordered product in
552: Eq. (\ref{pijjdef}) per fermion.  Given that, to a good approximation,
553: different technifermion doublets contribute equally to $S$, it is natural to
554: define a reduced quantity, $\hat S$, that represents the contribution to $S$
555: from each such pair, viz.,
556: %
557: \beq
558: \hat S = \frac{S}{N_D} \ . 
559: \label{shat}
560: \eeq
561: %
562: 
563: Then, in terms of the current-current correlation functions defined above,
564: $S$, as defined in Eq. (\ref{s}), is given by 
565: %
566: \beq
567: \hat S = 4 \pi \left. \frac{d}{d q^2} \left[ \Pi_{VV}(q^2) - \Pi_{AA}(q^2)
568:    \right] \right\vert_{q^2 = 0}\, ,
569:  \label{scor}
570: \eeq
571: %
572: It is convenient to define the compact notation 
573: %
574: \beq
575: \Pi_{V-A}(q^2) \equiv \Pi_{VV}(q^2) - \Pi_{AA}(q^2) \ . 
576: \label{piva}
577: \eeq
578: %
579: 
580: As is evident from Eq. (\ref{scor}), one may also consider $S$ in a different
581: context, namely that of an abstract vectorial SU($N$) gauge theory with $N_f$
582: massless fermions transforming according to the fundamental representation of
583: this group, and with all other interactions much weaker in strength than the
584: SU($N$) gauge interaction.  In this case, in contrast to technicolor models,
585: where $N_f$ is even (since the technifermions have left-handed components
586: forming SU(2)$_L$ doublets), $N_f$ can be even or odd (being restricted to be
587: less than $N_{f,cr}$ so that the theory is in the confinement phase with
588: spontaneous chiral symmetry breaking).  Here, one could naturally define the
589: contribution to $S$ from each fermion individually, namely, $S/N_f$.  However,
590: since our main application will be to technicolor, we shall continue, as in
591: Ref. \cite{pmsw}, to present our results in terms of $\hat S$.
592: 
593: Because of the asymptotic freedom of the SU($N$) theory, for Euclidean $q^2$
594: much larger than $\Lambda^2$, dimensional considerations imply that,
595: asymptotically, $\Pi_{V-A}(q^2) \simeq \langle \bar\psi\psi \rangle^2/q^4$ up
596: to logs arising from anomalous dimensions.  Combining this property with the 
597: analytic properties of $\Pi_{V-A}(q^2)$, one can write the following 
598: dispersion relation, with $t \equiv q^2$:
599: %
600: \beq
601: \Pi_{V-A}(t) = \frac{1}{\pi} \int_0^\infty ds \, 
602: \frac{ Im(\Pi_{V-A}(s))}{s-t-i\epsilon}
603: \label{disprel}
604: \eeq
605: %
606: The dimensionless spectral function $\rho_J(s)$ is defined as 
607: %
608: \beq
609: \rho_J(s) \equiv \frac{Im(\Pi_{JJ}(s))}{\pi s}
610: \label{rho}
611: \eeq
612: %
613: This spectral function encodes information about the hadronic states produced
614: by the current $J$.  In terms of these spectral functions, one then has
615: %
616: \beq
617: \hat S = 4 \pi \int_0^\infty  \frac{ds}{s} \bigg [ \rho_V(s)-\rho_A(s) \bigg ] 
618: \label{w0}
619: \eeq
620: %
621: An early discussion of the integral on the right-hand side of Eq. (\ref{w0})
622: and its connection to $\pi^+ \to \ell^+ \nu_\ell \gamma$ decay appeared in
623: \cite{dmo}.  It is of interest to comment on the two equivalent expressions,
624: Eqs. (\ref{scor}) and (\ref{w0}) for $\hat S$.  The first of these,
625: Eq. (\ref{scor}), expresses $\hat S$ in terms of the slope of $\Pi_{V-A}(q^2)$
626: at $q^2=0$.  In contrast, the second, Eq. (\ref{w0}) expresses it as an
627: integral of $1/s$ times the difference of the physical spectral functions for
628: the vector and axial vector currents in the timelike region, which depends on
629: the spectrum of hadronic states and the strengths of their couplings to these
630: currents.  Thus, naively, one might think that these two expressions depend on
631: rather different properties of the SU($N$) theory.  The fact that they are
632: actually equivalent is a consequence of the analytic properties of
633: $\Pi_{JJ}(q^2)$ which are used, via the Cauchy theorem, to derive the resultant
634: dispersion relation.  Thus, although it is evaluated at a single point, the
635: derivative $\Pi_{V-A}'(0)$ ``knows'' about the full hadronic spectrum,
636: including the presence or absence of walking.
637: 
638: In passing, we note that the spectral functions directly in terms of the 
639: products of currents, as 
640: %
641: \beqs
642: & & i\int d^4x \ e^{i q \cdot x} \ \langle 0|T(V_\mu^a(x)V_\nu^b(0))|0\rangle  
643: \cr\cr
644: &=& -\delta^{ab} \int_0^\infty \, ds \, 
645: \frac{\rho_V(s)(sg_{\mu\nu}-q_\mu q_\nu)}{s-q^2-i\epsilon} 
646: \label{rhov}
647: \eeqs
648: %
649: and
650: %
651: \beqs
652: && i \int d^4x \ e^{i q \cdot x} \ \langle 0|T(A_\mu^a(x)A_\nu^b(0))|0\rangle  
653: \cr\cr 
654: &=& -\delta^{ab}\bigg [ \int_0^\infty \, ds \,
655: \frac{\rho_A(s)(sg_{\mu\nu}-q_\mu q_\nu)}{s-q^2-i\epsilon} + 
656: \frac{q_\mu q_\nu}{q^2} f_P^2 \ \bigg ] \ , \cr\cr
657: & & 
658: \label{rhoa}
659: \eeqs
660: %
661: where one explicitly separates out the contribution from the ($J=0$ part 
662: of) $\rho_A(s)$ at $s=0$ due to the Nambu-Goldstone bosons (NGB's). In
663: Eq. (\ref{rhoa}), the quantity $f_P$ (where $P$ stands for ``pseudoscalar'') is
664: the $N_f$-flavor generalization of $f_\pi$ defined by the transition matrix 
665: element
666: %
667: \beq
668:    \langle 0 \vert A_\mu^a (0) \vert \pi^b(q) \rangle 
669:    = i q_\mu f_P \delta^{ab}\, ,
670: \label{fp}
671: \eeq 
672: %
673: with $a,b=1,2,\ldots,N_f^2-1$.  In actual QCD, the chiral symmetry is
674: explicitly broken by the $u$ and $d$ current-quark masses (and also by
675: electroweak interactions), so that the pions decay, and, in particular, the
676: dominant weak decay of the $\pi^+$, $\pi^+ \to \mu^+ \nu_\mu$ has a rate
677: proportional to $f_\pi^2$. Thus, $f_P$ might be called the generalized
678: Nambu-Goldstone decay constant, but we will avoid this term, since in our basic
679: SU($N$) theory with other interactions turned off, these Nambu-Goldstone bosons
680: are exactly massless and do not decay.  In the chiral limit of QCD, with
681: $m_u=m_d=0$, it has been estimated that $(f_\pi)_{ch.lim.}/f_\pi \simeq 0.935$
682: \cite{gl85}, so that, with the physical value $f_\pi=92.4 \pm 0.3$, one infers
683: that $(f_\pi)_{ch.lim.} \simeq 86$ MeV (with a theoretical uncertainty of
684: several per cent from the chiral extrapolation).  This slight decrease will not
685: be important for our work.
686: 
687: The constant $f_P$ may be calculated by first solving the Schwinger-Dyson
688: equation for the momentum-dependent dynamical fermion mass $\Sigma(p)$ and then
689: substituting this into the Pagels-Stokar relation \cite{psrel},
690: %
691: \beq
692:  f_P^2 \, = \, \frac{N_c}{4\pi^2} \int_0^\infty y \, dy \,
693:        \frac{\Sigma^2(y) \, - \,  \frac{y}{4} \,
694: \frac{d}{dy} \left[ \Sigma^2(y) \right] }
695:        { [ y \ +\  \Sigma^2(y) \ ]^2}  \ .
696: \label{psrel}
697: \eeq
698: %
699: Calculations using this method~\cite{hys,mmw,sg} have shown (for $N=N_c=3$)
700: that as $N_f$ increases from the value $N_f=2$ toward $N_{f,cr}$, the
701: generalized quantity $f_P$ decreases, as is expected, since $f_P$ is an order
702: parameter for spontaneous chiral symmetry breaking.  Furthermore, in the strong
703: walking limit $N_f \nearrow N_{f,cr}$, i.e., $\alpha_* \searrow \alpha_{cr}$,
704: it has been found that $f_P$ vanishes like \cite{my,miranskybook}
705: %
706: \beq
707: f_P = c_f \Lambda \, \exp
708: \bigg [ -\pi \Big ( \frac{\alpha_*}{\alpha_{cr}} - 1 \Big )^{-1/2} \bigg ] \ ,
709: \label{sigsol}
710: \eeq
711: %
712: where $c_f$ is a constant.  In Ref. \cite{sg} we calculated $f_P$ in the
713: crossover region between this extreme walking limit and smaller values of
714: $N_f$, corresponding to larger values of $\alpha_*$, closer to QCD with $N_f=2$
715: or 3 and found a dramatic growth in $f_P/\Lambda$, approaching values nearer to
716: QCD, as expected.
717: 
718: 
719: Because of the asymptotic freedom and consequent asymptotic decay of
720: $\Pi_{V-A}(q^2)$ for large Euclidean $q^2$ and the fact that the fermions are
721: massless in the underlying SU($N$) theory, the following integral relations
722: (first and second Weinberg sum rules) hold \cite{wsumrule}-\cite{dg}:
723: %
724: \beq
725: \int_0^\infty \, ds \, \bigg [\rho_V(s)-\rho_A(s) \bigg ] = 
726: \Pi_{V-A}(0) = f_P^2 
727: \label{w1}
728: \eeq
729: %
730: and
731: %
732: \beq
733: \int_0^\infty \, ds \, s \,  \bigg [\rho_V(s) - \rho_A(s) \bigg ] = 0 \ . 
734: \label{w2}
735: \eeq
736: %
737: Clearly, Eqs. (\ref{w0}), (\ref{w1}), and (\ref{w2}) are of similar form,
738: viz., integrals of $\rho_V(s)-\rho_A(s)$ weighted, respectively by $s^{-1}$, 
739: 1, and $s$.  Because of these different factors, the contributions to
740: the integral (\ref{w0}) for $\hat S$ are weighted toward smaller 
741: $s$ values, while the contributions to the integral (\ref{w2}) are weighted
742: toward larger $s$ values, as compared with those for the integral in Eq. 
743: (\ref{w1}). 
744: 
745: Finally, in the context of an abstract SU($N$) gauge theory (with other
746: interactions turned off), it is worthwhile to mention the connection between
747: $S$ and the coefficient denoted $\bar L_{10}$.  To explain this connection, we
748: recall the definition of $\bar L_{10}$. Provided that $N_f \ge 2$, the theory
749: contains $N_f^2-1$ massless Nambu-Goldstone bosons.  (For this discussion, we
750: turn off electroweak interactions completely; for our application to
751: technicolor, we do not turn them off, and hence three of the would-be
752: Nambu-Goldstone bosons become the longitudinal modes of the $W^\pm$ and
753: $Z$. Furthermore, in the TC/ETC context, other would-be NGB's gain masses from
754: color and ETC interactions that explicitly break the ${\rm SU}(N_f)_L \times 
755: {\rm SU}(N_f)_R$ global chiral symmetry.) 
756: A useful description of the low-energy physics for energies $E \ll 4 \pi f_P$
757: is then provided by a chiral Lagrangian (see. e.g., \cite{gl85} and references
758: to earlier work therein).  This is a function of the chiral fields
759: %
760: \beq
761: U = \exp \left ( \frac{2 i \sum_{a=1}^{N_f^2-1} \pi^a T^a}{f_P} \right ) \ . 
762: \label{uchiral}
763: \eeq
764: %
765: For example, for QCD with just the two light quarks $u$ and $d$, $U$ would have
766: the form $U=e^{i \pi \cdot \tau/f_\pi}$.  Now let us denote the elements of the
767: global symmetry groups ${\rm SU}(N_f)_L$ and ${\rm SU}(N_f)_R$ as $U_L$ and
768: $U_R$.  Then under a transformation by an element of the chiral symmetry group
769: ${\rm SU}(N_f)_L \times {\rm SU}(N_f)_R$, $U \to U_L U U_R^\dagger$.  One can
770: also formally introduce external gauge fields (contracted with the respective
771: generators of ${\rm SU}(N_f)_L$ and ${\rm SU}(N_f)_R$), $L_\mu$ and $R_\mu$.
772: These have the transformation properties $L_\mu \to U_L L_\mu U_L^\dagger - i
773: (\partial_\mu U_L) U_L^\dagger$ and $R_\mu \to U_R R_\mu U_R^\dagger - i
774: (\partial_\mu U_R) U_R^\dagger$.  In terms of these, one constructs the
775: covariant derivative $D_\mu U = \partial_\mu U - i L_\mu U + i U R_\mu$.  One
776: also defines external field strengths $(F_L)_{\mu\nu}$ and $(F_R)_{\mu\nu}$ in
777: the usual manner.  The lowest-order term in this effective Lagrangian is then
778: %
779: \beq
780:  \frac{f_P^2}{4} \, {\rm Tr} \left [ (D_\mu U)^\dagger (D^\mu U) \right ] \ . 
781: \label{lowestterm}
782: \eeq
783: %
784: In terms of these quantities, $L^{(r)}_{10}$ is the coefficient of the term 
785: %
786: \beq
787: O_{10} = {\rm Tr}(U^\dagger (F_L)_{\mu\nu} U (F_R)^{\mu\nu}) , 
788: \label{o10}
789: \eeq
790: %
791: where the superscript $(r)$ refers to the renormalized quantity. From this, one
792: obtains a quantity which is constructed to be independent of the
793: renormalization scale $\mu$ \cite{gl85,chipt} 
794: %
795: \beq
796: \bar L_{10} = L_{10}^{(r)}(\mu) + \frac{1}{192\pi^2} 
797: \Bigg [ \ln \left ( \frac{m_\pi^2}{\mu^2} \right ) + 1 \Bigg ]
798: \label{l10bar}
799: \eeq
800: %
801: The relation with $S$ is then 
802: %
803: \beq
804: S = -16 \pi \bar L_{10} \ . 
805: \label{slrel}
806: \eeq
807: %
808: Note that although $\bar L_{10}$ requires $N_f \ge 2$ to be well-defined, it
809: does not require $N_f$ to be even.  For the specific case of QCD, fits to
810: experimental data, including, in particular, the radiative decay $\pi^+ \to e^+
811: \nu_e \gamma$, yield the value \cite{gl85,hyrev}
812: %
813: \beq
814: S = 0.33 \pm 0.04 \ . 
815: \label{svalue}
816: \eeq
817: %
818: 
819: To the extent that this is dominated by the two light quarks $u$ and $d$, one
820: has $N_f=2$, and hence $N_D=1$, so that the measured value of $S$ for QCD is
821: also equal to $\hat S$.  The fact that the light-quark vector and axial-vector
822: mesons $\rho$ and $a_1$ largely saturate the expression for $S$,
823: Eq. (\ref{w0}), is consistent with this conclusion.  An approximate calculation
824: of $\hat S$ has been performed using the ladder approximation to the
825: Schwinger-Dyson and (inhomogeneous) Bethe-Salpeter equations for QCD ($N=3$)
826: with $N_f=2$ quarks of negligible mass \cite{hys}.  Studies have also been done
827: for the case where one neglects the strange quark mass $m_s$, i.e., $N=3$,
828: $N_f=3$ \cite{hys,pms}.  Since for either of these values of $N_f$ the beta
829: function of the QCD theory does not exhibit an infrared fixed point, it is
830: necessary to cut off the growth of the strong coupling.  For typical cutoffs,
831: it was found that the SD-IBS calculations tended to yield slightly too large a
832: value of $\hat S = S$, namely $\hat S \simeq 0.45 - 0.5$ \cite{hys,pms}, rather
833: than a value in the $1\sigma$ experimental range $0.29 \lsim S \lsim 0.37$.
834: This suggests that in the SD-IBS approach, used for a vectorial confining
835: SU($N$) gauge theory with small values of $N_f$ such that the theory has no
836: perturbative IR fixed point, with a typical IR cutoff on the coupling, tends to
837: overestimate $S$ by about a factor of 1.4.  Since the calculation of $S$ in QCD
838: is a problem in strongly coupled, nonperturbative physics, and the
839: calculational method that was used is only approximate (neglecting, for
840: example, instanton contributions), one should probably not be surprised that it
841: does not precisely reproduce the measured value of $S$.
842: 
843: 
844: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
845: \section{Current-Current Correlation Functions in Terms of Bethe-Salpeter 
846: Amplitudes}
847: \label{sec:correlator-BS}
848:  
849: In this section, we explain how the current-current correlation functions are
850: obtained from the Bethe-Salpeter amplitudes, which will be calculated via the
851: inhomogeneous Bethe-Salpeter (IBS) equation \cite{bs}-\cite{marisroberts03},
852: \cite{miranskybook}.  These Bethe-Salpeter amplitudes $\chi^{(J)}$, where $J=V$
853: or $A$, are essentially form factors, whose behavior in the timelike region
854: describes the coupling of the given current to physical hadronic bound states
855: that can be produced by this current, with an analytic continuation into the
856: spacelike region.  Here we will only need these amplitudes for the spacelike
857: region $q^2 < 0$ and at the point $q^2=0$. The amplitudes may be defined in
858: terms of the three-point vertex function as
859: %
860: \beqs 
861: & & \delta_j^k \left( T^a \right)_{f}^{f'} \int \frac{d^4 p}{(2 \pi)^4} 
862: \, e^{- i p \cdot r} \, \chi^{(J)}_{\alpha \beta}(p;q,\epsilon) = \cr\cr
863: & & = \epsilon^\mu \int d^4 x \, e^{i q \cdot x} 
864: \langle 0 \vert T(\psi^k_{\alpha f}(r/2) \ \bar\psi_{jf'\beta}(-r/2)\
865: J_\mu^a(x)) | 0 \rangle, \cr\cr
866: & & 
867: \label{eq:three-point}
868: \eeqs
869: %
870: where $(f,f')$, $(j,k)$ and $(\alpha, \beta)$ are, respectively, the flavor,
871: gauge, and spinor indices.  Closing the fermion legs of the above
872: three-point vertex function and taking the limit $r \rightarrow 0$, one can
873: express the current correlator in terms of the Bethe-Salpeter amplitude as
874: %
875: \beqs
876: & &  \Pi_{JJ}(q^2) = \cr\cr
877: & & \frac{1}{3} \Bigg ( \frac{N}{2} \Bigg ) \sum_{\epsilon} 
878:  \int \frac{d^4 p}{i (2\pi)^4} 
879: {\rm Tr} \left[ \left(\epsilon \cdot G^{(J)}\right) 
880:    \chi^{(J)}(p;q,\epsilon) \right] \cr\cr
881: & & 
882:  \label{eq:Pi_JJ}
883: \eeqs
884: %
885: where \cite{gamcon}
886: %
887: \beq
888: G_\mu^{(V)} = \gamma_\mu,\ \ \ \ G_\mu^{(A)} = \gamma_\mu \gamma_5,
889: \eeq
890: %
891: and an average has been taken over the polarizations, so that $\Pi_{JJ}(q^2)$
892: does not depend on the polarization $\epsilon$.
893:  
894:  
895: We expand the Bethe-Salpeter amplitude $\chi_{\alpha
896: \beta}^{(J)}(p;q,\epsilon)$ in terms of a complete bispinor basis with basis
897: elements $\Gamma^{(J)}_i$ and the invariant amplitudes $\chi^{(J)}_i$ as
898: %
899: \beq
900:    \left[ \chi^{(J)}(p;q,\epsilon) \right]_{\alpha \beta} 
901:    =\ \sum_{i=1}^{8} \left[ \Gamma_i^{(J)}(p;\hat{q},\epsilon)
902:    \right]_{\alpha \beta} \chi^{(J)}_{i} (p;q) ,
903:  \label{eq:chi-expanded}
904: \eeq
905: %
906: where we define
907: %
908: \beq
909: Q^2 \equiv -q^2
910: \label{Qsq}
911: \eeq
912: %
913: so that $Q^2 > 0$ in the spacelike region, and set 
914: %
915: \beq
916: \hat{q}_\mu \equiv \frac{q_\mu}{\sqrt{Q^2}}
917: \label{qhat}
918: \eeq
919: %
920: The bispinor basis elements can be chosen in such a manner
921: that they have the same spin, parity and charge conjugation as the
922: corresponding current $J_\mu^a(x)$.  For the vector vertex we adopt the
923: following bispinor basis elements:
924: %
925: \beqs
926:  &&  \Gamma^{(V)}_1 = \fsl{\epsilon},\ \ 
927:      \Gamma^{(V)}_2 = \frac{1}{2} [\fsl{\epsilon},\fsl{p}] 
928:                             (p \cdot \hat q), \ \ 
929:      \Gamma^{(V)}_3 = \frac{1}{2} [\fsl{\epsilon},\fsl{\hat q}], \cr\cr
930:  &&  \Gamma^{(V)}_4 = \frac{1}{3!}[\fsl{\epsilon},\fsl{p},\fsl{\hat q}],\ \ 
931:      \Gamma^{(V)}_5 = (\epsilon \cdot p),\ \ 
932:      \Gamma^{(V)}_6 = \fsl{p} (\epsilon \cdot p), \cr\cr
933:  &&  \Gamma^{(V)}_7 = \fsl{\hat q}(p \cdot \hat q) 
934:             (\epsilon \cdot p),\ \ 
935:      \Gamma^{(V)}_8 = \frac{1}{2} [\fsl{p},\fsl{\hat q}]
936:             (\epsilon \cdot p),
937: \label{vbispinors}
938: \eeqs
939: %
940: where $[a,b,c] \equiv a[b,c] + b[c,a] + c[a,b]$.  For the axial-vector vertex
941: we use the bispinor basis elements
942: %
943: \beqs
944:  && \hspace{-4mm}
945:      \Gamma^{(A)}_1 = \fsl{\epsilon}\ \gamma_5 ,\  
946:      \Gamma^{(A)}_2 = \frac{1}{2} [\fsl{\epsilon},\fsl{p}] 
947:                             \gamma_5  ,\  
948:      \Gamma^{(A)}_3 = \frac{1}{2} [\fsl{\epsilon},\fsl{\hat q}]\ 
949:               (p \cdot \hat q)
950:                  \ \gamma_5 , \cr\cr
951:  && \hspace{-4mm}
952:       \Gamma^{(A)}_4 = \frac{1}{3!}[\fsl{\epsilon},\fsl{p},
953:               \fsl{\hat q}] \gamma_5,\ \ 
954:      \Gamma^{(A)}_5 = (\epsilon \cdot p)\ (p \cdot \hat q)\ 
955:         \gamma_5, \cr\cr
956:  && \hspace{-4mm}
957:      \Gamma^{(A)}_6 = \fsl{p} (\epsilon \cdot p)\ \gamma_5,\ \ 
958:      \Gamma^{(A)}_7 = \fsl{\hat q}\ (\epsilon \cdot p)\ 
959:              (p \cdot \hat q)
960:                  \ \gamma_5, \cr\cr
961:  && \hspace{-4mm}
962:      \Gamma^{(A)}_8 = \frac{1}{2} [\fsl{p},\fsl{\hat q}]
963:              (\epsilon \cdot p)\
964:       (p \cdot \hat q)\ \gamma_5.
965: \label{avbspinors}
966: \eeqs
967: %
968: Given the charge conjugation properties of the vector and axial-vector
969: currents and the above choice of the bispinor basis elements, it 
970: follows that invariant amplitudes $\chi_i^{(J)}$ are even functions of $p\cdot
971: \hat{q}$.
972:  
973: 
974: In the present analysis it is convenient to choose the Lorentz reference frame
975: so that only the timelike component of $q^\mu$ is nonzero.  Since we are
976: working in the spacelike region (which avoids physical mass singularities), we
977: thus use a Wick rotation with 
978: %
979: \beq
980:    q^\mu = (i Q , 0 , 0 , 0 ) .
981: \label{qparam}
982: \eeq
983: %
984: Similarly, for the relative momentum $p^\mu$ we perform a Wick rotation and
985: parametrize it in terms of the real variables $u$ and $w$ (with
986: dimensions of mass) as
987: %
988: \beq
989:    p \cdot q = - Q\  u \ ,\ \  p^2 = - u^2 - w^2 .
990: \label{pparam}
991: \eeq
992: %
993: Hence, the invariant amplitudes $\chi^{(J)}_i$ are functions of $u$ and $w$:
994: %
995: \beq
996:    \chi_i^{(J)} = \chi_i^{(J)}(u,w;Q) .
997: \label{chidep}
998: \eeq
999: %
1000: {}Owing to the charge-conjugation properties of the Bethe-Salpeter amplitude
1001: $\chi^{(J)}$ and the bispinor basis elements defined above, the invariant
1002: amplitudes $\chi_i^{(J)}(u,w)$ satisfies the relation
1003: %
1004: \beq
1005:    \chi_i^{(J)}(u,w;Q) = \chi_i^{(J)}(-u,w;Q) \ .
1006:  \label{eq:even-property}
1007: \eeq
1008: %
1009: Using this property of the invariant amplitudes, we rewrite 
1010: Eq.~(\ref{eq:Pi_JJ}) as 
1011: %
1012: \beqs
1013:   && \hspace{-1.5cm}
1014:    \Pi_{VV}(Q^2) \ =\  \frac{N}{\pi^3} \int_0^\infty du 
1015:    \int_0^\infty dw\ w^2 \cr\cr
1016:   &&\left[ - \chi^{(V)}_1(u,w;Q) + \frac{w^2}{3} \chi^{(V)}_6(u,w;Q) \ \right],
1017:  \label{eq:PiVV}
1018:  \eeqs
1019:  \beqs
1020:   && \hspace{-1.5cm}
1021:    \Pi_{AA}(Q^2) \ =\  \frac{N}{\pi^3} \int_0^\infty du 
1022:    \int_0^\infty dw\ w^2 \cr\cr
1023: &&\left[\ \ \chi^{(A)}_1(u,w;Q)  - \frac{w^2}{3} \chi^{(A)}_6(u,w;Q) \ \right].
1024: \label{eq:PiAA}
1025: \eeqs
1026: %
1027: Here, we have used the expanded form of the Bethe-Salpeter amplitude given in 
1028: Eq.~(\ref{eq:chi-expanded}) and carried out the three-dimensional 
1029: angular integration.
1030:  
1031:  From Eqs.~(\ref{eq:PiVV}) and (\ref{eq:PiAA}), it follows that 
1032: %
1033: \beqs
1034:  \Pi_{V-A}(Q^2) &=&
1035:  \frac{1}{3}\left (\frac{N}{2}\right )\sum_{\epsilon} \int \frac{d^4 p}{i (2
1036:  \pi)^4}  \cr\cr
1037:  &&{\rm Tr} \bigg [\fsl{\epsilon} \chi^{(J)}(p;q,\epsilon) -
1038:  \fsl{\epsilon} \gamma_5 \chi^{(A)}(p;q,\epsilon) \bigg ] \cr\cr
1039:  &=&  \frac{N}{\pi^3} \int_0^\infty du \int_0^\infty dw\ w^2 \cr\cr 
1040:  &&\Bigg[ - \bigg( \chi^{(V)}_1(u,w;Q) + \chi^{(A)}_1(u,w;Q) \bigg ) \cr\cr
1041:  && + \frac{w^2}{3} \bigg ( \chi^{(V)}_6(u,w;Q) + \chi^{(A)}_6(u,w;Q) \bigg )
1042:  \ \Bigg]. \cr\cr
1043: & & 
1044: \label{eq:PiV-A}
1045: \eeqs 
1046: % 
1047: Although both $\Pi_{VV}(Q^2)$ and $\Pi_{AA}(Q^2)$ are individually
1048: logarithmically divergent, the underlying ${\rm SU}(N_f)_L \times {\rm
1049: SU}(N_f)_R$ chiral symmetry guarantees that these divergences cancel in the
1050: difference $\Pi_{VV} - \Pi_{AA}$, which is therefore finite.
1051: 
1052: 
1053: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1054: \section{Inhomogeneous Bethe-Salpeter equation}
1055: \label{sec:IBS}
1056:  
1057: In this section we discuss the full (inhomogeneous) Bethe-Salpeter equation,
1058: which we will use to calculate current-current correlation functions and, from
1059: these, $S$ (and, as a check, also $f_P$).  The IBS equation is a
1060: self-consistent description of the coupling of a current $J^a_\mu$ to
1061: fermion-antifermion bound states.  This coupling is represented by the
1062: Bethe-Salpeter amplitude $\chi^{(J)}$.  The four-momenta assigned to the
1063: fermion and antifermion are $p_\psi = p + (q/2)$ and $p_{\bar \psi}=-p+(q/2)$
1064: so that the total momentum of the bound state is $q$, and the relative momentum
1065: of the $\psi$ and $\bar \psi$ is $2p$. Since we are dealing with $J=1$ bound
1066: states, the bound-state amplitude also depends on the polarization vector
1067: $\epsilon$, which satisfies $\epsilon \cdot q = 0$ and $\epsilon \cdot \epsilon
1068: = -1$.  A grapical indication of the inhomogeneous Bethe-Salpeter equation
1069: structure is given in Fig.~\ref{fig:IBSeq}. 
1070: %
1071: \begin{figure}[t]
1072:    \begin{center}
1073:      \includegraphics[height=1.8cm]{IBSeq.eps}
1074:    \end{center}
1075:  \caption{A graphical expression of the IBS equation in the 
1076:  (improved) ladder approximation.}
1077:  \label{fig:IBSeq}
1078:  \end{figure}
1079: %
1080: The IBS equation is 
1081: %
1082: \beq
1083:    T(p;q)\ \chi^{(J)}(p;q,\epsilon) \ \ =\ \ 
1084:    \epsilon \cdot G^{(J)}\ 
1085:    +\ K(p;k) \ast \chi^{(J)}(k;q,\epsilon) .
1086:  \label{eq:IBSeq}
1087: \eeq
1088: %
1089: Here the kinetic part $T$ is
1090: %
1091: \beq
1092:   T(p;q) =   - S_F^{-1}(p + q/2) \otimes  S_F^{-1}(p - q/2) \ ,
1093:  \label{Tdef}
1094: \eeq
1095: %
1096: where $S_F(p)=1/(A(p)\fsl{p} - \Sigma(p))$. We follow the standard procedure of
1097: using the Landau gauge in calculations with the Schwinger-Dyson and
1098: Bethe-Salpeter equations since in this gauge the fermion wave function
1099: renormalization factor $A(p)=1$.  The physical results are, of course,
1100: gauge-invariant. The Bethe-Salpeter kernel $K$ in the improved ladder
1101: approximation is expressed as
1102: %
1103: \beqs
1104: \hspace{-1cm}
1105:  K(p;k) &=&  C_{2f} \, \frac{\bar{g}^2(p,k)}{(p-k)^{2}} \times \cr\cr 
1106:   &\times&  \left(-g_{\mu\nu} + \frac{(p-k)_\mu (p-k)_\nu}{(p-k)^2}
1107:           \right) \cdot \gamma^\mu \otimes \gamma^\nu ,
1108: \label{Kdef}
1109: \eeqs
1110: %
1111: where $C_{2f}$ was given in Eq. (\ref{c2f}). 
1112: In the above expressions we use the tensor product notation
1113: %
1114: \beq
1115:    (A \otimes B) \,\chi  =  A\, \chi\, B \ ,
1116: \label{aob}
1117: \eeq
1118: %
1119: and the inner product notation
1120: %
1121: \beq
1122:   K(p;k) \ast \chi^{(J)}(k;q,\epsilon) =   
1123:    \int \frac{d^4 k}{i(2\pi)^4}\  K(p,k)\  \chi(k;q) \ .
1124: \label{kchi}
1125: \eeq
1126: %
1127: where summations over Dirac indices are understood.  (In contrast, for our
1128: previous calculations \cite{mmw,sg} of meson masses, we only needed to use the
1129: homogeneous Bethe-Salpeter equation.)
1130: 
1131: 
1132: As in Refs. \cite{Kugo:1992pr,mmw,pms,pmsw,sg}, we make the ansatz
1133: for the running coupling, after Euclidean rotation,
1134: %
1135: \beq
1136: \alpha(p_E,q_E) = \alpha(p_E^2+q_E^2) \ ,
1137: \label{gsqform}
1138: \eeq
1139: %
1140: where the subscript denotes Euclidean.  Since $\alpha$ would naturally depend
1141: on the gluon momentum squared, $(p-q)^2 = p^2+q^2-2p \cdot q$, the functional
1142: form (\ref{gsqform}) amounts to dropping the scalar product term, $-2p \cdot
1143: q$.  As discussed in Ref. \cite{sg}, this is a particularly reasonable
1144: approximation in the case of a walking gauge theory because most of the
1145: contribution to the integral comes from a region of Euclidean
1146: momenta where $\alpha$ is nearly constant.  Hence, the shift upward or downward
1147: due to the $-2p \cdot q$ term in the argument of $\alpha$ has very little
1148: effect on the value of this coupling for the range of momenta that make the
1149: most important contribution to the integral. The approximation (\ref{gsqform})
1150: enables one to carry out the angular integration analytically. 
1151: %
1152: 
1153: The momentum-dependent dynamical mass $\Sigma(p)$ for the fermion is obtained
1154: from the Schwinger-Dyson equation,
1155: %
1156: \beq
1157:    \Sigma(p) = - K(p,k) \ast S_F(p).
1158: \label{sdeq}
1159: \eeq
1160: %
1161: We use the same kernel $K(p,k)$ here as in the IBS equation in order to respect
1162: the ${\rm SU}(N_f)_L \times {\rm SU}(N_f)_R$ chiral symmetry~\cite{
1163: Kugo:1992pr,Bando-Harada-Kugo}.  The numerical method that is used for solving
1164: the SD equation and the IBS equation involves approximating the respective
1165: integrals by discrete sums and is the same as in Ref.~\cite{pms,pmsw}.
1166: The reader is referred to these papers for more details on this method.
1167: 
1168: 
1169: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1170: \section{Results and Discussion}
1171: \label{sec:results}
1172: 
1173: In this section we present the results of our calculations of $\Pi_{V-A}(Q^2)$
1174: and $\hat S$ using the Schwinger-Dyson and (inhomogeneous) Bethe-Salpeter
1175: equations.  It is appropriate to include an obvious cautionary remark that
1176: these calculations involve strong couplings $\alpha$ of order unity, and
1177: therefore there could be significant corrections to the (improved) ladder
1178: approximation used in our solutions of the Schwinger-Dyson and Bethe-Salpeter
1179: equations.  Accordingly, as one check on the reliability of our methods, we
1180: have also carried out a comparison of $f_P$ obtained from the SD and IBS
1181: equations via eq. (\ref{w1}) with $f_P$ obtained from the SD equation via the
1182: Pagels-Stokar relation.  
1183: 
1184: In Fig.\ref{fig:PiV-A} we plot our calculated values of $\Pi_{V-A}(Q^2)/f_P^2$
1185: for $\alpha_* = 1.8, 1.6, 1.4, 1.2$ and 1.0, as functions of the dimensionless
1186: quantity $Q^2/f_P^2$.
1187: %
1188:  \begin{figure}[t]
1189:    \begin{center}
1190:      \includegraphics[height=6.2cm]{PiV-A_over_fpi2.eps}
1191:    \end{center}
1192:  \caption{Plot of $\Pi_{V-A}(Q^2)/f_P^2$ as a function of $Q^2$ in the
1193: spacelike interval $0 \le Q^2/f_P^2 \le 1.2$, for $\alpha_* = 1.0, 1.2, 1.4,
1194: 1.6, 1.8$.  As indicated, the horizontal axis refers to the quantity
1195: $Q^2/f_P^2$.  For comparison, $\Pi_{V-A}(Q^2)/f_P^2$ for QCD with $N_f=3$ 
1196: massless quarks is also plotted.  See text for further details.}
1197:  \label{fig:PiV-A}
1198:  \end{figure}
1199: %
1200: %
1201: The slope of each curve at $Q^2=0$ is equal to $-{\hat S}/(4\pi)$ for the given
1202: value of $\alpha_*$.  We first observe that $\Pi_{V-A}(Q^2)$ is almost linear
1203: at $Q^2=0$, with a small positive second derivative.  This is the justification
1204: for our earlier statement that the implications of our findings for technicolor
1205: would be essentially the same whether we used the definition of $\hat S$ in
1206: terms of a derivative at $Q^2=0$ or a finite difference, analogously to
1207: Eq. (\ref{spdg}). Secondly, from Fig.  \ref{fig:PiV-A} it is clear that the
1208: magnitude of the slope, and hence $\hat S$, decreases as $\alpha_*$ decreases
1209: toward the chiral phase transition point $\alpha_{cr}$.
1210: 
1211: In Fig.~\ref{fig:S-hat_n}, we plot the values of ${\hat S}$, derived from the
1212: slope of $\Pi_{V-A}(Q^2)$ at $Q^2=0$, for several values of $\alpha_*$.  As
1213: indicated by the subscript $n$, the values are normalized by the value of $\hat
1214: S$ at $\alpha_*=1.8$.
1215: %
1216: %
1217: \begin{figure}[t]
1218:    \begin{center}
1219:      \includegraphics[height=6.4cm]{S-hat_n.eps}
1220:    \end{center}
1221:  \caption{Plot of ${\hat S}_n$ for several values of $\alpha_*$ in the 
1222: range $0.9 \le \alpha_\ast \le 1.8$. As indicated by the subscript $n$, the 
1223: values are normalized by the value of $\hat S$ at $\alpha_*=1.8$.}
1224:  \label{fig:S-hat_n}
1225: \end{figure}
1226: %
1227: %
1228: This figure shows that $\hat S$ decreases by about 40 \% as $\alpha_*$ is
1229: reduced from 1.8 to 0.9, or equivalently (c.f. Eq. (\ref{nfsol})) as $N_f$ is
1230: increased from 10.3 to 11.6.  As our calculation of meson masses in
1231: Ref. \cite{sg} showed, this is a crossover region, in which the theory is
1232: changing from QCD-like, non-walking behavior at smaller $N_f$ to the walking
1233: regime at larger $N_f$ approaching $N_{f,cr}$.  Reinserting the factor of
1234: $N_D=N_f/2$ to get $S$ itself, we obtain a decrease by about 30 \% in $S$,
1235: since $N_D$ only increases by about 10 \% over this range.  Thus, our
1236: calculation shows that for this range of values, $\hat S$ decreases
1237: significantly as one moves from the QCD-like to the walking regimes.  This
1238: finding is an important result of our present study. 
1239: 
1240: 
1241: As a check on our calculational methods, we compare $f_P$ calculated in two
1242: different ways: via the Pagels-Stokar relation, Eq. (\ref{psrel}), denoted
1243: $(f_P)_{PS}$, and by the first Weinberg sum rule (W1) or equivalent relation
1244: $\Pi_{V-A}(0)=f_P^2$ in Eq. (\ref{w1}), denoted $(f_P)_{W1}$.  In view of the
1245: fact that the Pagels-Stokar itself is approximate and that our solution of the
1246: SD and IBS equations involves the ladder approximation and the neglect of
1247: completely nonperturbative contributions such as those due to instantons, we do
1248: not expect exact agreement between these two different methods of calculation.
1249: In Fig. \ref{fig:fpi_ratio} we present a plot of the ratio
1250: $(f_P)_{PS}/(f_P)_{W1}$ as a function of $\alpha_*$.  The closeness of this
1251: ratio to unity gives one measure of the accuracy and reliability of our
1252: calculations.
1253: %
1254: %
1255: \begin{figure}[t]
1256:    \begin{center}
1257:      \includegraphics[height=6.4cm]{fpi_ratio.eps}
1258:    \end{center}
1259:  \caption{Plot of the ratio $(f_P)_{PS}/(f_P)_{W1}$, where $(f_P)_{PS}$ and 
1260: $(f_P)_{W1}$ denote $f_P$ calculated via Eq. (\ref{psrel}) and via 
1261: Eq. (\ref{w1}), respectively, as a function of $\alpha_*$.}
1262:  \label{fig:fpi_ratio}
1263: \end{figure}
1264: %
1265: %
1266: We see that the ratio is within about 20 \% of unity and is essentially
1267: independent of $\alpha_*$ in the range considered, with the Pagels-Stokar
1268: method yielding a slightly smaller value than the expression in terms of the
1269: current-current correlation functions.  This gives us further confidence in the
1270: results of our SD-IBS calculation of $\hat S$.
1271: 
1272: 
1273: Our finding that $\hat S$ and $S$ are reduced as one moves from the QCD-like
1274: regime toward the walking regime of an SU($N$) gauge theory is in agreement
1275: with the approximate analytic results of Refs. \cite{ats,acd,as_s,iwts}, and 
1276: it complements those works, being based on a numerical solution of the
1277: Schwinger-Dyson and Bethe-Salpeter equations. In the sub-interval 
1278: $0.9 \le \alpha_* \le 1.0$ closer to the walking limit our results coincide 
1279: with those in Ref. \cite{pmsw}.  Our use of a larger interval
1280: has the advantage that we are able to observe a larger reduction in $\hat S$ as
1281: $\alpha_*$ decreases than was done in Ref.  \cite{pmsw}.  We have not attempted
1282: here to examine the extreme walking limit $(\alpha_* - \alpha_{cr})/\alpha_{cr}
1283: \to 0^+$.  As discussed in Ref. \cite{pmsw}, it becomes increasingly difficult
1284: to obtain an accurate numerical solution for $\hat S$ in this limit because of
1285: very strong cancellations.  
1286: 
1287: 
1288: In addition to this decreasing trend of $\hat S$, one may also discuss the
1289: absolute magnitude of $\hat S$.  Our calculation yields $\hat S = 0.47$ at
1290: $\alpha_*=1.8$. If $\hat S$ continues to be a monotonically function of $N_f$
1291: (and hence, in the range of interest here, also a monotonic function of
1292: $\alpha_*$), then an extrapolation of our calculation to the QCD case with
1293: $N_f=2$ or, with the strange quark mass neglected, $N_f=3$, would predict a
1294: value of $\hat S \gsim 0.5$.  This is similar to the value that was obtained in
1295: earlier studies using the SD and IBS equations in a different manner than here,
1296: where it was necessary to introduce an cutoff on the growth of the strong
1297: coupling in the infrared \cite{hys,pms}.  In Ref. \cite{pms} it was shown that
1298: if one used a cutoff that led to a very large value of the coupling, one could
1299: get a result for $S$ in agreement with the experimental value (\ref{svalue}),
1300: but the reliability of the calculational method in the presence of such a large
1301: coupling was not clear.  We shall adopt the optimisitc viewpoint here of giving
1302: greater weight to the change in $\hat S$ as a function of $N_f$ than to the
1303: absolute value of $\hat S$ itself.  Equivalently, one could envision applying
1304: an overall correction factor of about 2/3 to the absolute magnitude of $\hat S$
1305: so that the value for small $N_f$ matches that in QCD.  Physically, this
1306: factor would be regarded as correcting for the strong-coupling effects not
1307: included in the SD-IBS analysis.  
1308: 
1309: 
1310: Although one cannot use perturbation theory reliably to calculate $S$ in a
1311: strongly coupled gauge theory, the perturbative result is often used in
1312: discussions of constraints on new physics, and hence it is worthwhile to see
1313: how our results compare with the perturbative computation. A one-loop
1314: perturbative calculation with degenerate fermions having effective masses
1315: satisfying $(2\Sigma/m_Z)^2 \gg 1$ yields the well-known result
1316: $S_{pert.}=N_{D,tot.}/(6 \pi)$ where here $N_{D,tot.}=N_D N$, i.e,
1317: %
1318: \beq
1319: \hat S_{pert.} = \frac{N}{6\pi} \ . 
1320: \label{s_hat_pert}
1321: \eeq
1322: %
1323: In QCD with $N_f=2$ and $N=N_c=3$, this perturbative calculation would predict
1324: $S_{QCD,pert.} \simeq 1/(2\pi) \simeq 0.16$.  The experimental value in
1325: Eq. (\ref{svalue}) is approximately twice this; $S_{QCD} \simeq 2
1326: S_{QCD,pert.}$.  The reductions that we have found in $\hat S$ and $S$ for the
1327: range of $\alpha_*$ investigated suggest that in a walking theory, much or all
1328: of the above factor of 2 might be removed, and the true value of $S$ might well
1329: be comparable to, or, indeed, perhaps less than, the perturbative estimate.
1330: 
1331: For reference, in the one-doublet and one-family technicolor models, the
1332: perturbative expressions for the techniparticle contributions to $S$ are
1333: $S_{TC,pert.}=N_{TC}/(6\pi)$ and $S_{TC,pert.}=2N_{TC}/(3\pi)$, respectively.
1334: With $N_{TC}=2$, these take the values $S_{QCD,pert.} \simeq 0.1$ and 0.4,
1335: respectively. Fits to precision electroweak data yield allowed regions in $S$
1336: and two other parameters describing modifications of the $Z$ and $W$
1337: propagators by new physics beyond the Standard Model, namely the parameter $T$
1338: measuring violations of custodial SU(2) from this new physics and a third
1339: parameter, $U$, of somewhat less importance here.  Since the Standard Model
1340: expression for $S$ includes a term $(1/(6\pi))\ln (m_H/m_{H,ref.})$, the
1341: resultant allowed regions depend on the choice of the reference value of the SM
1342: Higgs mass, $m_{H,ref.}$.  The comparison of these with a technicolor theory is
1343: complicated by the fact that technicolor has no fundamental Higgs field;
1344: sometimes one formally uses $m_{H,ref.} \sim 1$ TeV for a rough estimate, since
1345: the SM with $m_H \sim 1$ TeV has strong longitudinal vector boson scattering,
1346: as does technicolor. However, this may involve some double-counting when one
1347: also includes contributions to $S$ from technifermions, whose interactions and
1348: bound states (e.g., techni-vector mesons) are responsible for the strong $W^+_L
1349: W^-_L$ and $Z_LZ_L$ scattering in a technicolor framework.  The current fit
1350: \cite{pdg,lepewwg} disfavors values of $S \gsim 0.2$.  Our findings in this
1351: paper suggest, in agreement with the previous works noted above using different
1352: methods, that the constraint on walking technicolor models could be less severe
1353: than would be inferred from the perturbative formula for the technifermion
1354: contribution to $S$.
1355: 
1356: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1357: \section{Summary}
1358: \label{sec:Summary}
1359: 
1360: In summary, using numerical solutions of the Schwinger-Dyson and
1361: (inhomogeneous) Bethe-Salpeter equation, we have calculated the $S$ parameter
1362: as a function of the approximate infrared fixed point, $\alpha_*$, or
1363: equivalently, the number of massless fermions, $N_f$, in a vectorial, confining
1364: SU($N$) gauge theory.  We have focused on the crossover region between the
1365: walking and QCD-like (non-walking) regimes.  Our results show that $\hat S$ and
1366: also $S$ decrease significantly as $\alpha_*$ decreases in this range.  This
1367: trend agrees with earlier indications of a decrease in $S$ in walking gauge
1368: theories.  We have discussed the implications for technicolor models.
1369: 
1370: \bigskip
1371: 
1372: M.K. thanks Profs. M. Harada and K. Yamawaki for the collaborations on the
1373: related Refs. \cite{mmw,pmsw}. R.S. thanks T. Appelquist for useful comments.  
1374: This research was partially supported by the grant NSF-PHY-03-54776. 
1375: 
1376: 
1377: \section{Appendix}
1378: \label{appendix}
1379: 
1380: In this appendix we present some results on vector meson dominance (VMD) and
1381: their relevance to the crossover region that we are studying numerically.  In
1382: QCD itself, VMD has served as a useful approximate model.  The matrix elements
1383: for the production of a vector or axial-vector meson from the vacuum are 
1384: %
1385: \beq
1386: \langle \rho^a | V^b_\mu | 0 \rangle = i \delta^{ab} f_\rho \epsilon_\mu
1387: \label{rhomatrixelement}
1388: \eeq
1389: %
1390: %
1391: \beq
1392: \langle a_1^a | A^b_\mu | 0 \rangle = i \delta^{ab} f_{a_1} \epsilon_\mu
1393: \label{a1matrixelement}
1394: \eeq
1395: %
1396: where $\epsilon_\mu$ are the respective polarization four-vectors. In QCD,
1397: $f_\rho \simeq 154$ MeV, so that the first Weinberg sum rule yields 
1398: $f_{a_1} =\sqrt{f_\rho^2-f_\pi^2} = 123$ MeV. A pole-dominated form for
1399: the current-current correlation functions in the (simplistic) narrow-resonance
1400: approximation is
1401: %
1402: \beq
1403: \Pi_{V-A}(q^2) = \frac{f_V^2 m_V^2}{m_V^2-q^2-i\epsilon} 
1404:                - \frac{f_A^2 m_A^2}{m_A^2-q^2-i\epsilon} \ , 
1405: \label{piform}
1406: \eeq
1407: %
1408: where, for QCD, the ground-state vector and axial-vector mesons are the
1409: $\rho(776)$ and $a_1(1230)$, i.e., $V = \rho$ and $A = a_1$.  
1410: Using the principal value relation
1411: %
1412: \beq
1413: \frac{1}{s-s_0 \mp i\epsilon} = P \bigg ( \frac{1}{s-s_0} \bigg ) 
1414: \pm i \pi \delta(s-s_0) \ , 
1415: \label{pvrel}
1416: \eeq
1417: %
1418: the pole-dominated form (\ref{piform}) gives, for the spectral functions, 
1419: %
1420: \beq
1421: \rho_V(s) = f_V^2 \delta(s-m_V^2)
1422: \label{rho_v}
1423: \eeq
1424: %
1425: and
1426: %
1427: \beq
1428: \rho_A(s) = f_A^2 \delta(s-m_A^2)
1429: \label{rho_a}
1430: \eeq
1431: %
1432: Although these are useful simplifications, more realistic analyses of the
1433: spectral function sum rules take into account the finite widths of the $\rho$
1434: and $a_1$ resonances, and also nonresonant contributions \cite{dg}. 
1435: 
1436: Expanding Eq. (\ref{piform}) for large Euclidean $q^2$, one obtains
1437: %
1438: \beq
1439: \Pi_{V-A}(q^2) = -\frac{(f_V^2 m_V^2 - f_A^2 m_A^2)}{q^2} + 
1440: O \Bigg ( \frac{1}{q^4} \Bigg ) \ . 
1441: \label{piexpand}
1442: \eeq
1443: %
1444: The asymptotic freedom of the SU($N$) gauge theory implies that for $|q^2| \gg
1445: \Lambda^2$, $\Pi_{V-A}(q^2) \propto 1/q^4$, so this yields  the relation
1446: %
1447: \beq
1448: f_V^2 m_V^2 - f_A^2 m_A^2 = 0 \ , 
1449: \label{w2vmd}
1450: \eeq
1451: %
1452: which is the second Weinberg sum rule, evaluated in the VMD approximation. 
1453: Substituting the forms for the spectral functions into Eq. (\ref{w0}) yields 
1454: %
1455: \beq
1456: S_{VMD} = -16\pi \bar L_{10} = 4\pi \Bigg ( \frac{f_V^2}{m_V^2}
1457:                                            -\frac{f_A^2}{m_A^2} \Bigg ) 
1458: \label{svmd}
1459: \eeq
1460: %
1461: Similar substitutions into the first and second Weinberg sum rules yield the
1462: relations 
1463: %
1464: \beq
1465: f_V^2-f_A^2=f_P^2 
1466: \label{w1vmd}
1467: \eeq
1468: %
1469: and Eq. (\ref{w2vmd}).  The narrow-width VMD relations are approximate, since
1470: they neglect the sizeable widths of the $\rho$ and $a_1$, but the Weinberg
1471: relations follow from the asymptotic freedom of QCD when one sets $m_u=m_d=0$.
1472: 
1473: As $N_f$ increases and there is a crossover in the SU($N$) theory from QCD-like
1474: behavior to walking behavior, there are resultant changes in how these integral
1475: relations are satisfied.  For both the QCD-like and walking regimes, the
1476: underlying asymptotic freedom of the theory implies that the current-current
1477: correlation functions $\Pi_{V-A}(q^2)$ have a $1/q^4$ asymptotic falloff (up to
1478: logs) for Euclidean $q^2 \gg \Lambda^2$.  However, in a walking regime, the
1479: scale of chiral symmetry breaking, given by $f_P \sim \Sigma$, is much less
1480: than $\Lambda$, and in the interval $\Sigma^2 \ll q^2 \ll \Lambda^2$,
1481: $\Pi_{V-A}(q^2)$ have a less rapid falloff, $\sim 1/q^{4-2\gamma})$, where
1482: $\gamma$ is the anomalous dimension of the bilinear fermion operator and is
1483: $\gamma \simeq 1$ for strong walking, whence a $\Pi_{JJ}(q^2) \sim 1/q^2$
1484: falloff in this interval. Physically, the coupling $\alpha$ is strong, of order
1485: O(1) but slowly varying for an extended interval of energies between $\Sigma$
1486: and $\Lambda$.  Thus it is plausible that in a walking theory, rather than just
1487: the lowest-lying vector and axial-vector resonances making important
1488: contributions to the various spectral function integrals, there could be
1489: significant contributions from a number of higher-lying states with the same
1490: quantum numbers \cite{lanerev}.  In this case, still retaining the VDM
1491: approximation, one would generalize the expressions above to be sums over these
1492: higher-lying states.  For example, the evaluation of Eqs. (\ref{w0}),
1493: (\ref{w1}), and (\ref{w2}) in the VMD approximation would read
1494: %
1495: \beq
1496: \hat S_{VMD} =  4\pi \sum_i \Bigg ( \frac{f_{V_i} ^2}{m_{V_i}^2}
1497:                                    -\frac{f_{A_i}^2}{m_{A_i}^2} \Bigg ) 
1498: \label{svmdwalk}
1499: \eeq
1500: %
1501: %
1502: \beq
1503: \sum_i(f_{V_i}^2-f_{A_i}^2)=f_P^2 
1504: \label{w1vmdwalk}
1505: \eeq
1506: %
1507: and
1508: %
1509: \beq
1510: \sum_i (f_{V_i}^2 m_{V_i}^2 - f_{A_i}^2 m_{A_i}^2) = 0 \ . 
1511: \label{w2vmdwalk}
1512: \eeq
1513: %
1514: More generally, however, it is not clear that the VMD model would apply in this
1515: walking regime.  One analytic approximation is to use VMD for a partial
1516: evaluation representing contributions from $\sqrt{s} \simeq \Sigma$ and
1517: supplement this with a term due to a fermion loop \cite{as_s}.  In the region
1518: that we have concentrated on in this paper, in which the theory is crossing
1519: over between strong walking behavior and QCD-like behavior, one expects that
1520: the saturation of the spectral function integrals has a form that is
1521: intermediate between walking and QCD-like.  In our actual calculation of $\hat
1522: S$ via Eq. (\ref{scor}), we only need to calculate $\Pi_{V-A}'(0)$, which we do
1523: by means of numerical solutions of the relevant SD and IBS equations, so we do
1524: not have to deal with questions of how the spectral functions behave in a
1525: walking gauge theory. In future work it would be worthwhile to use the
1526: connection between these two different ways of calculating $S$ to undersand the
1527: hadronic spectrum better in such a walking gauge theory.
1528: 
1529: 
1530: \begin{thebibliography}{99}
1531: 
1532: \bibitem{bz}
1533: An early paper on the phase structure of vectorial gauge theories is 
1534: T.~Banks and A.~Zaks,
1535: %``On The Phase Structure Of Vector - Like Gauge Theories With
1536: %Massless Fermions,'' 
1537: Nucl.\ Phys.\ B {\bf 196} (1982), 189.
1538: %%CITATION = NUPHA,B196,189;%%
1539: 
1540: \bibitem{tc}
1541: S. Weinberg, Phys. Rev. D {\bf 19}, 1277 (1979);
1542: L. Susskind, {\it ibid.} D {\bf 20}, 2619 (1979); see also 
1543: S. Weinberg, Phys. Rev. D {\bf 13}, 974 (1976).
1544: 
1545: \bibitem{wtc1}
1546: B. Holdom, Phys. Lett. B {\bf 150}, 301 (1985).
1547: 
1548: \bibitem{wtc2}
1549: K. Yamawaki, M. Bando, and K. Matumoto, Phys. Rev. Lett. {\bf
1550: 56}, 1335 (1986).
1551: 
1552: \bibitem{chipt1}
1553: T. Appelquist, D. Karabali, and L. C. R. Wijewardhana, Phys. Rev. Lett. {\bf
1554: 57}, 957 (1986); T. Appelquist and L. C. R. Wijewardhana, Phys. Rev. D
1555: {\bf 35}, 774 (1987); Phys. Rev. D {\bf 36}, 568 (1987).
1556: 
1557: \bibitem{chipt2}
1558: T. Appelquist, J. Terning, and L. C. R. Wijewardhana,
1559: Phys. Rev. Lett.  {\bf 77}, 1214 (1996).
1560: 
1561: \bibitem{my}
1562: V. Miransky and K. Yamawaki, Phys. Rev. D {\bf 55}, 5051 (1997); {\it ibid.}
1563: {\bf 56}, E 3768 (1997). See also V. Miransky and P. Fomin,
1564: Sov. J. Part. Nucl. {\bf 16}, 203 (1985).
1565: 
1566: \bibitem{chipt3}
1567: T. Appelquist, A. Ratnaweera, J. Terning, and
1568: L. C. R. Wijewardhana, Phys. Rev. D {\bf 58}, 105017 (1998).
1569: 
1570: \bibitem{gl85}
1571: J. Gasser and H. Leutwyler, Nucl. Phys. B {\bf 250}, 465 (1985); 
1572: {\it ibid.} B {\bf 250}, 517 (1985). See also the papers in
1573: Refs. \cite{chipt}-\cite{hyrev}. 
1574: 
1575: \bibitem{chipt}
1576: J. Gasser and H. Leutwyler, Phys. Rept. C {\bf 87}, 77 (1982);
1577: J. Gasser, H. Leutwyler, A. Pich, and E. de Rafael, Nucl. Phys. {\bf 321}, 311
1578: (1989); G. Colangelo, J. Gasser, and H. Leutwyler, Nucl. Phys. B {\bf 603}, 
1579: 125 (2001). 
1580: 
1581: \bibitem{resonances}
1582: % The role of resonances in chiral perturbation theory
1583: G. Ecker, J. Gasser, A. Pich, and E. de Rafael, Nucl. Phys. B {\bf 321}, 311
1584: (1989); G. Ecker, J. Gasser, H. Leutwyler, A. Pich, and E. de Rafael,
1585: Phys. Lett. B {\bf 223}, 425 (1989); M. Knecht and E. de Rafael, 
1586: Phys. Lett. B {\bf 424}, 335 (1998). 
1587: 
1588: \bibitem{meissner}
1589: U. Meissner, Rept. Prog. Phys. {\bf 56}, 903 (1993). 
1590: 
1591: \bibitem{hyrev}
1592: M. Harada and K. Yamawaki, Phys. Rept. {\bf 381}, 1 (2004).
1593: 
1594: \bibitem{pt}
1595: M.~E.~Peskin and T.~Takeuchi,
1596: %``A New Constraint On A Strongly Interacting Higgs Sector,''
1597: Phys.\ Rev.\ Lett.\  {\bf 65}, 964 (1990);
1598: %%CITATION = PRLTA,65,964;%%
1599: %M.~E.~Peskin and T.~Takeuchi,
1600: %``Estimation of oblique electroweak corrections,''
1601: Phys.\ Rev.\ D {\bf 46}, 381 (1992). 
1602: %%CITATION = PHRVA,D46,381;%%
1603: 
1604: \bibitem{ab}
1605: G. Altarelli and R. Barbieri, Phys. Lett. B {\bf 253}, 161 (1991);
1606: G. Altarelli, R. Barbieri, F. Caravaglios,
1607: Int. J. Mod. Phys. A {\bf 13}, 1031 (1998).
1608: 
1609: \bibitem{pdg}
1610: See http://pdg.lbl.gov. 
1611: 
1612: \bibitem{lepewwg}
1613: http://lepewwg.web.cern.ch/LEPEWWG/plots.
1614: 
1615: \bibitem{scalc1}
1616: B. Holdom and J. Terning, Phys. Lett. B {\bf 247}, 88 (1990); 
1617: M. Golden and L. Randall, Nucl. Phys. {\bf B} {\bf 361}, 3 (1991);
1618: R. Johnson, B.-L. Young, and D. McKay, Phys. Rev. D {\bf 43} (1991) R17;
1619: R. Cahn and M. Suzuki, Phys. Rev. D {\bf 44}, 3641 (1991).
1620: 
1621: \bibitem{ats}
1622: T. Appelquist and G. Triantaphyllou, Phys. Lett. B {\bf 278}, 345 (1992).
1623: 
1624: \bibitem{acd}
1625: R. Sundrum and S. Hsu, Nucl. Phys. B {\bf 391}, 127 (1993).
1626: 
1627: \bibitem{hys}
1628: M.~Harada and Y.~Yoshida,
1629: %``QCD S parameter from inhomogeneous Bethe-Salpeter equation,''
1630: Phys.\ Rev.\ D {\bf 50}, 6902 (1994). 
1631: %[arXiv:hep-ph/9406402].
1632: %%CITATION = HEP-PH 9406402;%%
1633: 
1634: \bibitem{as_s}
1635: T. Appelquist and F. Sannino, Phys. Rev. D {\bf 59}, 067702 (1999).
1636: 
1637: \bibitem{iwts}
1638: S. Ignjatovic, L. C. R. Wijewardhana, and T. Takeuchi, Phys. Rev. 
1639: D {\bf 61}, 056006 (2000).
1640: 
1641: \bibitem{pmsw}
1642: M. Harada, M. Kurachi, and K. Yamawaki, Prog. Theor. Phys. {\bf 115}, 765
1643: (2006); see also {\it  Proc. of 2004 International Workshop on Dynamical 
1644: Symmetry Breaking, 2004}, eds. M. Harada and K. Yamawaki, 
1645: (Nagoya Univ., 2005), p. 125.
1646: 
1647: \bibitem{sg}
1648: M.~Kurachi and R.~Shrock,
1649: %``Study of the Change from Walking to Non-Walking Behavior in a Vectorial
1650: %Gauge Theory as a Function of Nf,'' 
1651: hep-ph/0605290.
1652: %%CITATION = HEP-PH 0605290;%%
1653: 
1654: \bibitem{mmw} 
1655: M.~Harada, M.~Kurachi and K.~Yamawaki,
1656: %``Meson masses in large N(f) QCD from Bethe-Salpeter equation,''
1657: Phys.\ Rev.\ D {\bf 68}, 076001 (2003).
1658: %[arXiv:hep-ph/0305018].
1659: %%CITATION = HEP-PH 0305018;%% 
1660: 
1661: \bibitem{mme}
1662: K.~I.~Aoki, M.~Bando, T.~Kugo, M.~G.~Mitchard and H.~Nakatani,
1663: Prog. Theor. Phys. {bf 84}, 683 (1990); T.~Kugo, {\it in Proc. of 1991 Nagoya
1664: Spring School on Dynamical Symmetry Breaking}, ed. K. Yamawaki (World
1665: Scientific Pub. Co., Singapore, 1992), p. 35; K.~I.~Aoki, T.~Kugo and
1666: M.~G.~Mitchard, Phys. Lett. B {\bf 266}, 467 (1991); Phys. Lett. B {bf 286}, 
1667: 355 (1992).
1668: 
1669: \bibitem{pms}
1670: M.~Harada, M.~Kurachi and K.~Yamawaki,
1671: %``pi+ - pi0 mass difference from the Bethe-Salpeter equation,''
1672: Phys.\ Rev.\ D {\bf 70}, 033009 (2004).
1673: %[arXiv:hep-ph/0403120]
1674: %%CITATION = HEP-PH 0403120;%%
1675: 
1676: \bibitem{integer}
1677: %
1678: Here and below, when we mention non-integral values of $N_f$, it is implicitly
1679: understood that physical values of $N_f$ are, of course, non-negative integers,
1680: and the non-integral values are defined via an analytic continuation away from
1681: these physical values.
1682: 
1683: \bibitem{b0}
1684: D. Gross and F. Wilczek, Phys. Rev. Lett. {\bf 30}, 1343 (1973);
1685: H. D. Politzer, Phys. Rev. lett. {\bf 30}, 1346 (1973). 
1686: 
1687: \bibitem{b1}
1688: W. Caswell, Phys. Rev. Lett. {\bf 33}, 244 (1974); D. R. T. Jones, Nucl. Phys.
1689: B {\bf 75}, 531 (1974).
1690: 
1691: \bibitem{casimir} The Casimir invariant $C_2(R)$ of the representation $R$ is
1692: defined by ${\cal D}_R(T_a)^i_j{\cal D}_R(T_a)^j_k = C_2(R)\delta^i_k$, where
1693: $\{a,b\}$ and $\{i,j,k\}$ denote group and representation indices and sums over
1694: repeated indices are understood.
1695: 
1696: \bibitem{alm}
1697: T. Appelquist, K. Lane, and U. Mahanta, Phys. Rev. Lett. {\bf
1698: 61}, 1553 (1988); T. Appelquist, U. Mahanta, D. Nash, and L.C.R.  Wijewardhana,
1699: Phys. Rev. D {\bf 43}, 646 (1991);
1700: U. Mahanta, Phys. Rev. D {\bf 45}, 1405 (1992).
1701: 
1702: \bibitem{lgt} 
1703: Y.  Iwasaki et al., Phys. Rev. Lett. {\bf 69}, 21 (1992);
1704: Phys. Rev. D {\bf 69}, 014507 (2004); 
1705: P. Damgaard, U. Heller, A. Krasnitz, and P. Olesen, Phys. Lett. B {\bf 400},
1706: 169 (1997); R.  Mawhinney, Nucl. Phys. B (Proc. Suppl.) {\bf 63A-C}, 212 
1707: (1998); R.  Mawhinney, Nucl. Phys. B (Proc. Suppl.) {\bf 83}, 57 (2000).
1708: 
1709: \bibitem{Gardi}
1710: E.~Gardi and M.~Karliner,
1711: %``Relations between observables and the infrared fixed-point in
1712: %{QCD},'' Nucl.\ Phys.\ B {\bf 529}, 383 (1998); 
1713: %[arXiv:hep-ph/9802218]
1714: %%CITATION = HEP-PH 9802218;%%
1715: E.~Gardi, G.~Grunberg and M.~Karliner,
1716: %``Can the {QCD} running coupling have a causal analyticity
1717: %structure?,'' J. High Energy Phys. {\bf 07}, 007 (1998). 
1718: %[arXiv:hep-ph/9806462]
1719: %%CITATION = HEP-PH 9806462;%%
1720: 
1721: \bibitem{ExplicitSolution}
1722: R.~M.~Corless, G.~H.~Gonnet, D.~G.~E.~Hare, D.~J.~Jeffrey and D.~E.~Knuth,
1723: Adv.\ Comput.\ Math.\  {\bf 5}, 329 (1996).
1724: 
1725: \bibitem{etc}
1726: S. Dimopoulos, L. Susskind, Nucl. Phys. {\bf B155}, 23, (1979);
1727: E. Eichten, K. Lane, Phys.  Lett. B {\bf 90}, 125 (1980);
1728: E.~Farhi and L.~Susskind, Phys. Rep. {\bf 74}, 277 (1981). 
1729: 
1730: \bibitem{tcm} 
1731: Some recent reviews of TC/ETC models are given in
1732: \cite{lanerev,etcrev} and some recent studies include \cite{at94,as,sann,lm}.
1733: We do not consider here models that hypothesize a strong (``topcolor'') 
1734: interaction that produces a $\langle \bar t t \rangle$ condensate; these are
1735: reviewed in Refs. \cite{lanerev,etcrev}. 
1736: 
1737: \bibitem{lanerev}
1738: K. Lane, hep-ph/0202255.
1739: 
1740: \bibitem{etcrev}
1741: C. Hill and E. Simmons, Phys. Rep. {\bf 381}, 235 (2003); 
1742: R. S. Chivukula, M. Narain, and J. Womersley, in Ref. \cite{pdg}.
1743: 
1744: \bibitem{at94}
1745: T. Appelquist and J. Terning, Phys. Rev. D {\bf 50}, 2116 (1994).
1746: 
1747: \bibitem{as}
1748: T. Appelquist and R. Shrock, Phys. Lett. B {\bf 548}, 204 (2002); 
1749: Phys. Rev. Lett. {\bf 90}, 201801 (2003); T. Appelquist, M. Piai, and
1750: R. Shrock, Phys. Rev. D {\bf 69}, 015002 (2004); {\it ibid.}  
1751: D {\bf 70}, 093010 (2004);  Phys. Lett. B {\bf 593}, 175 (2004); 
1752: {\it ibid.} B {\bf 595}, 442 (2004). 
1753: 
1754: \bibitem{sann}
1755: D. Hong, S. Hsu, and F. Sannino, Phys. Lett. B {\bf 597}, 89 (2004); 
1756: D. Dietrich, F. Sannino, K. Tuominen, Phys. Rev. D {\bf 72}, 055001 (2005). 
1757: 
1758: \bibitem{lm}
1759: K. Lane and A. Martin, Phys. Rev. D {\bf 71}, 076007 (2005); 
1760: Phys. Lett. B {\bf 635}, 118 (2006). 
1761: 
1762: \bibitem{ts}
1763: N. D. Christensen and R. Shrock, Phys. Lett. B {\bf 632}, 92 (2006); see also
1764: N. D. Christensen and R. Shrock, Phys. Rev. Lett. {\bf 94}, 241801 (2005).
1765: 
1766: \bibitem{dmo}
1767: T.~Das, V.~S.~Mathur and S.~Okubo,
1768: %``Low-Energy Theorem In The Radiative Decays Of Charged Pions,''
1769: Phys.\ Rev.\ Lett.\  {\bf 19}, 859 (1967).
1770: %%CITATION = PRLTA,19,859;%%
1771: 
1772: \bibitem{psrel}
1773: H. Pagels and S. Stokar, Phys. Rev. D {\bf 20}, 2947 (1979).
1774: 
1775: \bibitem{miranskybook}
1776: V.~A.~Miransky, {\it Dynamical symmetry breaking in quantum field theories}, 
1777: (Singapore, World Scientific, 1993).
1778: 
1779: \bibitem{wsumrule}
1780: S.~Weinberg,
1781: %``Precise Relations Between The Spectra Of Vector And Axial Vector
1782: %Mesons,'' 
1783: Phys.\ Rev.\ Lett.\  {\bf 18} (1967), 507.
1784: %%CITATION = PRLTA,18,507;%%
1785: 
1786: \bibitem{sumrule2} 
1787: C. Bernard, A. Duncan, J. LoSecco, and S. Weinberg,
1788: Phys. Rev. D {\bf 12}, 792 (1975); (1976); E. Floratos, S. Narison, and E. de
1789: Rafael, Nucl. Phys. B {\bf 155}, 115 (1979); 
1790: S.~Weinberg, {\it The Quantum Theory of Fields} (Cambridge Univ. Press,
1791: Cambridge, 1996), v. 2, p. 266. 
1792: 
1793: \bibitem{svz}
1794: M. Shifman, A. Vainshtein, and V. Zakharov, Nucl. Phys. B {\bf 147}, 385
1795: (1979). 
1796: 
1797: \bibitem{dg}
1798: A numerical analysis of these sum rules in QCD in terms of data on the spectral
1799: functions for is given in J. Donoghue and E. Golowich, Phys. Rev. D {\bf 49}, 
1800: 1513 (1994). 
1801: 
1802: \bibitem{bs}
1803: E. Salpeter and H. Bethe, Phys. Rev. {\bf 84}, 1232 (1951). 
1804: 
1805: \bibitem{nakanishi}
1806: %``A General Survey Of The Theory Of The Bethe-Salpeter Equation,''
1807: For an early review, see N. Nakanishi, Prog. Theor. Phys. Suppl.  {\bf 43}, 1
1808: (1969).
1809: %%CITATION = PTPSA,43,1;%%
1810: 
1811: \bibitem{lane74}
1812: K. Lane, Phys. Rev. D {\bf 10}, 2605 (1974).
1813: 
1814: \bibitem{mn74}
1815: T. Maskawa and H. Nakajima, Prog. Theor. Phys. {\bf 52}, 1326 (1974).
1816: % Schwinger-Dyson Equation For Massless Vector Theory And Absence Of Fermion
1817: % Pole.
1818: 
1819: \bibitem{kugo}
1820: R. Fukuda and T. Kugo, Nucl. Phys. B {\bf 117}, 250 (1974);
1821: T. Kugo, Phys. Lett. B {\bf 76}, 625 (1978).
1822: 
1823: \bibitem{higashijima}
1824: K.~Higashijima,
1825: %``Dynamical Chiral Symmetry Breaking,''
1826: Phys.\ Rev.\ D {\bf 29}, 1228 (1984).
1827: %%CITATION = PHRVA,D29,1228;%%
1828: 
1829: \bibitem{Aoki:1990aq}
1830: K.~I.~Aoki, M.~Bando, T.~Kugo and M.~G.~Mitchard,
1831: %``Asymptotics Of The Pseudoscalar Bethe-Salpeter Amplitude,''
1832: Prog.\ Theor.\ Phys.\  {\bf 85}, 355 (1991).
1833: %%CITATION = PTPKA,85,355;%%
1834: 
1835: \bibitem{Kugo:1992pr}
1836: T.~Kugo and M.~G.~Mitchard,
1837: %``The Chiral Ward-Takahashi identity in the ladder approximation,''
1838: Phys.\ Lett.\ B {\bf 282}, 162 (1992). 
1839: %%CITATION = PHLTA,B282,162;%%
1840: 
1841: \bibitem{Bando-Harada-Kugo}
1842: M.~Bando, M.~Harada and T.~Kugo,
1843: %``External gauge invariance and anomaly in BS vertices and bound
1844: %states,''
1845: Prog.\ Theor.\ Phys.\  {\bf 91}, 927 (1994). 
1846: %[arXiv:hep-ph/9312343]
1847: %%CITATION = HEP-PH 9312343;%%
1848: 
1849: \bibitem{avsrev}
1850: R. Alkofer and L. von Smekal, Phys. Rept. {\bf 353}, 281 (2001).
1851: 
1852: \bibitem{marisroberts03}
1853: P. Maris and C. D. Roberts, Int. J. Mod. Phys. E {\bf 12}, 297 (2003).
1854: 
1855: \bibitem{gamcon}
1856: Our gamma matrix and metric conventions are those of Bjorken and Drell. 
1857: 
1858: \end{thebibliography}
1859: 
1860: \end{document}
1861: 
1862: 
1863: 
1864: