1: %\documentclass[draft]{ws-procs9x6}
2: \documentclass{ws-procs9x6}
3:
4: \newcommand{\beq}{\begin{equation}}
5: \newcommand{\eeq}{\end{equation}}
6: \newcommand{\beqs}{\begin{eqnarray}}
7: \newcommand{\eeqs}{\end{eqnarray}}
8: \newcommand{\lsim}{\mathrel{\raisebox{-
9: .6ex}{$\stackrel{\textstyle<}{\sim}$}}}
10: \newcommand{\gsim}{\mathrel{\raisebox{-
11: .6ex}{$\stackrel{\textstyle>}{\sim}$}}}
12: \newcommand{\pslash}{p\hspace{-0.067in}\slash}
13: \newcommand{\qslash}{q\hspace{-0.067in}\slash}
14: \newcommand{\kslash}{k\hspace{-0.067in}\slash}
15: \newcommand{\drawsquare}[2]{\hbox{%
16: \rule{#2pt}{#1pt}\hskip-#2pt% left vertical
17: \rule{#1pt}{#2pt}\hskip-#1pt% lower horizontal
18: \rule[#1pt]{#1pt}{#2pt}}\rule[#1pt]{#2pt}{#2pt}\hskip-#2pt% upper horizontal
19: \rule{#2pt}{#1pt}}% right vertical
20: \newcommand{\fund}{\raisebox{-.5pt}{\drawsquare{6.5}{0.4}}}% fund
21: \newcommand{\sym}{\raisebox{-.5pt}{\drawsquare{6.5}{0.4}}\hskip-0.4pt%
22: \raisebox{-.5pt}{\drawsquare{6.5}{0.4}}}% symmetric second rank
23: \newcommand{\asym}{\raisebox{-3.5pt}{\drawsquare{6.5}{0.4}}\hskip-6.9pt%
24: \raisebox{3pt}{\drawsquare{6.5}{0.4}}}% antisymmetric second rank
25:
26:
27: \begin{document}
28:
29: \title{Some Recent Results on Models of Dynamical Electroweak Symmetry
30: Breaking}
31:
32: \author{R. Shrock$^*$}
33:
34: \address{C. N. Yang Institute for Theoretical Physics\\
35: State University of New York\\
36: Stony Brook, New York 11794, USA\\
37: $^*$email: robert.shrock@sunysb.edu}
38:
39:
40: \begin{abstract}
41: We review some recent results on models of dynamical electroweak symmetry
42: breaking involving extended technicolor.
43:
44: \end{abstract}
45:
46:
47: \bodymatter
48:
49:
50: \section{Introduction}
51:
52: The origin of electroweak symmetry breaking (EWSB) is an outstanding unsolved
53: question in particle physics. In the Standard Model with gauge group $G_{SM} =
54: {\rm SU}(3)_c \times {\rm SU}(2)_L \times {\rm U}(1)_Y$, this symmetry breaking
55: is produced by postulating a Lorentz scalar Higgs field $\phi = {\phi^+ \choose
56: \phi^0}$ with weak isospin $I=1/2$ and weak hypercharge $Y=1$ and assuming that
57: its potential, $V=\mu^2 \phi^\dagger \phi + \lambda (\phi^\dagger \phi)^2$, has
58: $\mu^2 < 0$, thereby leading to a nonzero vacuum expection value for $\phi$.
59: However, this mechanism is unsatisfying for several reasons: (i) the EWSB is
60: put in by hand and no explanation is provided as to why $\mu^2$ is negative
61: when, {\it a priori}, it could be positive; (ii) $\mu^2$ and hence $m_H^2 =
62: -2\mu^2 = 2\lambda v^2$ are unstable to large radiative corrections from much
63: higher energy scales (the gauge hierarchy problem), so that extreme fine-tuning
64: is needed to keep the Higgs mass of order the electroweak scale; (iii) the SM
65: accomodates, but does not explain, fermion masses and mixing, via Yukawa
66: couplings of the generic form (suppressing the matrix structure) $m_f \simeq
67: y_f v/\sqrt{2}$; the $y_f$ values and generational hierarchy are put in by hand
68: with some $y_f$'s ranging down to $10^{-5}$ with no explanation.
69:
70: These facts have motivated an alternative approach based on dynamical
71: electroweak symmetry breaking driven by a strongly coupled vectorial gauge
72: interaction, associated with an exact gauge symmetry, called technicolor (TC)
73: \cite{tc}. The EWSB is produced by the formation of bilinear condensates of
74: technifermions. To communicate this symmetry breaking to the SM fermions
75: (which are technisinglet), one embeds technicolor in a larger, extended
76: technicolor (ETC) theory \cite{etc,etcrev}. In this talk we will review some
77: of our recent results in this area \cite{nt}-\cite{sg} obtained in
78: collaboration with T. Appelquist, M. Piai, N. Christensen, and M. Kurachi. (The
79: work with Kurachi is also discussed in his talk \cite{mk}.)
80:
81: As further motivation, one may recall that in both of two major previous cases
82: where fundamental scalar fields were used to model spontaneous symmetry
83: breaking, the actual underlying physics did not involve fundamental scalar
84: fields but instead a bilinear fermion condensate. The first of these was
85: superconductivity, where the phenomenological Landau-Ginzburg model made use of
86: a complex scalar field $\phi = \rho e^{i\theta}$ with a free energy functional
87: of the form $V= c_2|\phi|^2 + c_4|\phi|^4$, where $c_2 \propto (T - T_c)$ as $T
88: \to T_c$ (with $T_c$ being the critical temperature). Hence, for $T < T_c$,
89: $c_2 < 0$ and $\langle \phi \rangle \ne 0$. However, the true origin of
90: superconductivity is the dynamical formation of a condensate of Cooper pairs,
91: $\langle e e \rangle$. A second example is spontaneous chiral symmetry
92: breaking in hadronic physics. In the Gell-Mann L\'evy $\sigma$ model for this
93: phenomenon, the chiral symmetry breaking is a result of the coefficient of the
94: quadratic term in the potential being arbitrarily chosen to be negative,
95: producing a nonzero vev of the $\sigma$ field, quite analogous to the Higgs
96: mechanism. However, the real origin of spontaneous chiral symmetry breaking in
97: quantum chromodynamics (QCD) is the dynamical formation of a bilinear quark
98: condensate $\langle \bar q q \rangle$. Perhaps these previous examples might
99: serve as a guide for thinking about the physics underlying EWSB.
100:
101: Actually, one already knows of a source of dynamical electroweak symmetry
102: breaking, namely QCD. Consider, for simplicity, QCD with $N_f=2$ quarks, $u$,
103: $d$, taken to be massless. The quark condensate $\langle \bar q q \rangle =
104: \langle \bar q_L q_R \rangle + \langle \bar q_R q_L \rangle$, transforms as
105: $I_w = 1/2$, $|Y|=1$. By itself this theory would produce nonzero $W$ and $Z$
106: masses $m_W^2 \simeq (g^2/4) f_\pi^2$, \ $m_Z^2 \simeq (1/4)(g^2+g'^2)f_\pi^2$.
107: With $f_\pi = 92$ MeV, this yields $m_W \simeq 30$ MeV, $m_Z \simeq 34$
108: MeV. These $W$ and $Z$ masses satisfy the tree-level relation $\rho=1$, where
109: $\rho = m_W^2/(m_Z^2 \cos^2\theta_W)$. While the scale here is too small by $
110: \sim 10^3$ to explain the observed $W$ and $Z$ masses, it suggests how to
111: construct a model with dynamical EWSB.
112:
113: We shall consider a technicolor theory with gauge group ${\rm SU}(N_{TC})$
114: and a set of technifermions, generically denoted $F$, with zero Lagrangian
115: masses, transforming according to the fundamental representation of the
116: group. The scale of confinement and chiral symmetry breaking in this theory is
117: denoted $\Lambda_{TC}$ and is of order the electroweak scale. One assigns the
118: SU(2)$_L$ representations of the technifermions so that their left and right
119: components form SU(2)$_L$ doublets and singlets, respectively. A minimal
120: choice is the ``one-doublet'' model, with ${F_u^\tau \choose F_d^\tau}_L$ and
121: $F_{uR}^\tau$, $F_{dR}^\tau$, where $\tau$ is a TC gauge index and $Y=0$
122: ($Y=\pm 1$) for the SU(2)$_L$ doublet (singlets). Since the technicolor theory
123: is asymptotically free, it follows that, as the energy scale decreases,
124: $\alpha_{TC}$ increases, eventually producing condensates $\langle \bar F_u F_u
125: \rangle$ and $\langle \bar F_d F_d \rangle$ transforming as $I_w=1/2$, $|Y|=1$,
126: breaking EW sym. at $\Lambda_{TC}$. It follows that, to leading order, $m_W^2
127: = g^2f_{TC}^2 N_D/4$ and $m_Z^2 = (g^2 + g'^2) f_{TC}^2 N_D/4$, where $f_{TC}
128: \lsim \Lambda_{TC}$ is the TC analogue of $f_\pi \lsim \Lambda_{QCD}$ and $N_D$
129: is the number of SU(2)$_L$ technidoublets ($N_D=1$ in this case). These masses
130: satisfy the tree-level relation $\rho=1$ \cite{ssvz}. For this minimal
131: example, $f_{TC} \simeq 250$ GeV. Another class of TC models uses one SM
132: family of technifermions \cite{onefamily},
133: %
134: \beq
135: {U^{a \tau} \choose D^{a \tau}}_L \ , \quad \quad U^{a \tau}_R, \ \
136: D^{a \tau}_R
137: \eeq
138: %
139: \beq
140: {N^{\tau} \choose E^{\tau}}_L \ , \quad \quad N^{\tau}_R, \ \ E^{\tau}_R
141: \eeq
142: %
143: (where $a$ and $\tau$ are color and technicolor indices, respectively)
144: with the usual $Y$ assignments. Again, there is technifermion condensate
145: formation, with approximately equal condensates $\langle \bar F F \rangle$ for
146: $F=U^a, \ D^a, \ N, \ E$, generating dynamical technifermion masses
147: $\Sigma_{TC} \sim \Lambda_{TC}$. In this class of models $N_D=N_c+1=4$, so
148: $f_{TC} \simeq 125$ GeV.
149:
150:
151: Technicolor has the potential to solve/explain various problematic and/or
152: mysterious features of the Standard Model: (i) given the asymptotic freedom of
153: the TC theory, the condensate formation and hence EWSB are automatic and do not
154: require any {\it ad hoc} parameter choice like $\mu^2 < 0$ in the SM; (ii)
155: because TC has no fundamental scalar field, there is no hierarchy problem;
156: (iii) because $\langle \bar F F \rangle = \langle \bar F_L F_R \rangle +
157: \langle \bar F_R F_L \rangle$, technicolor explains why the chiral part of
158: $G_{SM}$ is broken and the residual exact gauge symmetry, ${\rm SU}(3)_c \times
159: {\rm U}(1)_{em}$, is vectorial. The fact that this latter symmetry is
160: vectorial is, of course, crucial in allowing nonzero mass terms for fermions
161: that are nonsinglets under color or charge.
162:
163:
164: In order to give masses to quarks and leptons, one must communicate the EWSB
165: in the technicolor sector to these SM fermions, and hence, as mentioned above,
166: one must embed technicolor in the larger ETC theory, with ETC gauge bosons
167: $V^i_\tau$ transforming (technisinglet) SM fermions into technifermions and
168: vice versa. To satisfy constraints on flavor-changing neutral current (FCNC)
169: processes, the ETC gauge bosons must have large masses. These masses can arise
170: from self-breaking (tumbling) of the ETC chiral gauge symmetry. The ETC theory
171: is arranged to be asymptotically free, so as the energy decreases from a high
172: scale, the ETC coupling $\alpha_{_{ETC}}$ grows and eventually becomes large
173: enough to form condensates which sequentially break the ETC symmetry group down
174: at scales $\Lambda_j$, $j=1,2,3$ to a residual exact TC subgroup, where
175: %
176: \beq
177: \Lambda_1 \simeq 10^3 \ {\rm TeV}, \quad\quad
178: \Lambda_2 \simeq 50-100 \ {\rm TeV}, \quad\quad
179: \Lambda_3 \simeq {\rm few \ TeV},
180: \label{lambdavalues}
181: \eeq
182: %
183: We will mainly focus on SU($N_{ETC}$) ETC models with one-family TC. These
184: gauge the Standard Model fermion generation index and combine it with the TC
185: index $\tau$, so
186: %
187: \beq
188: N_{ETC}=N_{gen.} + N_{TC} \ = \ 3 + N_{TC} \ .
189: \label{nrel}
190: \eeq
191: %
192:
193: To be viable, modern TC models are designed to have a coupling $g_{_{TC}}$ that
194: gets large, but runs slowly (``walks'') over an extended interval of energy
195: (WTC) \cite{wtc1}-\cite{chiv}. For sufficiently many technifermions, the TC
196: beta function has a second zero (approximate infrared fixed point of the
197: renormalization group) at a certain $\alpha_{_{TC}}=\alpha_{_{IR}} \ne 0$. As
198: the number of technifermions, $N_f$, increases, $\alpha_{_{IR}}$ decreases. In
199: WTC, one arranges so that $\alpha_{_{IR}}$ is slightly greater than the
200: critical value, $\alpha_{cr}$, for $\langle \bar F F \rangle$ formation. As
201: $N_f \nearrow N_{f,cr}$, $\alpha_{_{IR}} \searrow \alpha_{cr}$. Combining the
202: calculation of $\alpha_{_{IR}}$ from the two-loop beta function and an estimate
203: of $\alpha_{cr}$ from the Schwinger-Dyson equation, one finds $N_{f,cr} =
204: 2N_{TC}(50N_{TC}^2-33)/[5(5N_{TC}^2-3)]$ \cite{chipt2}. Hence, for $N_{TC}=2$,
205: one has $N_{f,cr} \simeq 8$. Thus, if $N_{TC}=2$, a one-family TC theory, with
206: its $N_f=N_w(N_c+1)=8$ technifermions plausibly exhibits walking behavior.
207:
208:
209: In a walking TC theory, as energy scale decreases, $\alpha_{_{TC}}$ grows, but
210: its rate of increase, $|\beta|$, decreases toward zero as $\alpha_{_{TC}}
211: \nearrow \alpha_{_{IR}}$, where $\beta=0$. Eventually, $\alpha_{_{TC}}$
212: exceeds $\alpha_{cr}$, $\langle \bar F F \rangle$ forms, and the technifermions
213: gain dynamical masses $\Sigma_{TC} \simeq \Lambda_{TC}$. WTC has several
214: advantages: (i) SM fermion masses are enhanced by the factor $\eta = \exp \big
215: [ \int_{\Lambda_{TC}}^{\Lambda_w} \, \mu^{-1} d\mu \
216: \gamma(\alpha_{_{TC}}(\mu)) \big ]$; (ii) hence, one can increase ETC scales
217: $\Lambda_i$ for a fixed $m_{f_i}$, reducing FCNC effects; (iii) $\eta$ also
218: enhances masses of pseudo-Nambu Goldstone bosons (PNGB's); and (iv) the walking
219: can reduce the value of the $S$ parameter \cite{pt}, given by
220: %
221: \beq
222: \frac{\alpha_{em}S}{\sin^2(2\theta_W)} = \frac{\Pi_{ZZ}(m_Z^2)-\Pi_{ZZ}(0)}
223: {m_Z^2} \ .
224: \label{s}
225: \eeq
226: %
227: The value $N_{TC}=2$ is also motivated by the fact that it makes possible a
228: mechanism to explain light neutrinos in (E)TC \cite{nt,lrs}. Substituting this
229: value $N_{TC}=2$ into eq. (\ref{nrel}) yields $N_{ETC}=5$.
230:
231:
232:
233: \section{ETC Models}
234:
235:
236: TC/ETC models are very ambitious, since a successful model would explain not
237: only EWSB but also the spectrum of Standard Model fermion masses. Thus it is
238: perhaps not surprising that no fully realistic model of this type has been
239: constructed yet. These models are subject to several strong constraints from
240: neutral flavor-changing current processes and precision electroweak data. We
241: have studied the properties of several types of ETC models in our work. One of
242: these has a gauge group $G = {\rm SU}(5)_{ETC} \times {\rm SU}(2)_{HC} \times
243: G_{SM}$. In addition to ETC, this has another gauge interaction, hypercolor
244: (HC), which helps to produce the desired ETC symmetry-breaking pattern. The SM
245: fermions and corresponding technifermions transform according to the
246: representations
247: %
248: \beq
249: Q_L: \ (5,1,3,2)_{1/3,L} \ , \quad \ u_R: \ (5,1,3,1)_{4/3,R} \ ,
250: \quad \ d_R: \ (5,1,3,1)_{-2/3,R} \ ,
251: \label{qreps}
252: \eeq
253: %
254: %
255: \beq
256: L_L: \ (5,1,1,2)_{-1,L}, \quad\quad e_R: \ (5,1,1,1)_{-2,R} \ ,
257: \eeq
258: %
259: where the subscripts denote $Y$. For example, writing out the components of
260: $e_R$, one has $(e^1, e^2, e^3, e^4, e^5)_R \ \equiv \ (e, \mu, \tau, E^4,
261: E^5)_R$, where the last two entries are the charged technileptons. In
262: addition, the model includes the SM-singlet ETC-nonsinglet fermions
263: %
264: \beq
265: \psi_{ij,R}: \ (\overline{10},1,1,1)_{0,R} \ , \quad\quad
266: \zeta^{ij,\alpha}_R: \ (10,2,1,1)_{0,R} \ , \quad\quad
267: \omega^{\alpha}_{p,R} : \ 2(1,2,1,1)_{0,R}
268: \eeq
269: %
270: where here $1 \le i,j \le 5$ are SU(5)$_{ETC}$ indices, $\alpha=1,2$ is an
271: SU(2)$_{HC}$ index, and $p=1,2$ is a copy number.
272: In this model the SM fermions and corresponding technifermions have vectorial
273: couplings to ETC gauge bosons, but the SM-singlet sector makes the full ETC
274: theory a chiral gauge theory. Hence, there are no fermion mass terms in
275: the lagrangian. This theory has no gauge or global anomalies.
276:
277: Since the ETC theory is asymptotically free, its gauge coupling grows as the
278: energy scale decreases. The ETC breaking occurs because of the formation of
279: ETC-noninvariant bilinear condensates of ETC-nonsinglet, SM-singlet fermions.
280: To analyze the stages of symmetry breaking, we identify plausible preferred
281: condensation channels using a generalized most-attractive-channel (MAC)
282: approach. We envision that as the energy decreases from high values down to $E
283: \sim \Lambda_1 \sim 10^3$ TeV, the coupling $\alpha_{_{ETC}}$ becomes
284: sufficiently large to produce condensation in the attractive channel
285: $(\overline{10},1,1,1)_{0,R} \times (\overline{10},1,1,1)_{0,R} \to
286: (5,1,1,1)_0$, breaking ${\rm SU}(5)_{ETC} \to {\rm SU}(4)_{ETC}$. With respect
287: to the unbroken ${\rm SU}(4)_{ETC}$, we have $(\overline{10},1,1,1)_{0,R} =
288: (\bar 4,1,1,1)_{0,R} + (\bar 6,1,1,1)_{0,R}$; we denote the $(\bar
289: 4,1,1,1)_{0,R}$ as $\alpha_{1i,R} \equiv \psi_{1i,R}$ for $2 \le i \le 5$ and
290: the $(\bar 6,1,1,1)_{0,R}$ as $\xi_{ij,R} \equiv \psi_{ij,R}$ for $2 \le i,j
291: \le 5$. The associated condensate is then
292: %
293: \beq
294: \langle \epsilon^{1 i j k \ell} \xi^T_{ij,R} C \xi_{k \ell,R} \rangle =
295: 8 \langle \xi^T_{23,R} C \xi_{45,R} - \xi^T_{24,R} C \xi_{35,R} +
296: \xi^T_{25,R} C \xi_{34,R} \rangle \ .
297: \eeq
298: %
299: The six fields $\xi_{ij,R}$, $2 \le i,j \le 5$, involved in this condensate
300: gain dynamical masses $\simeq \Lambda_1$.
301:
302: At energy scales below $\Lambda_1$, depending on relative strengths of gauge
303: couplings, different symmetry-breaking sequences can occur. Again, these arise
304: via the dynamical formation of condensates involving SM-singlet ETC-nonsinglet
305: fermions. For example, one sequence, S1, leads to the breaking
306: ${\rm SU}(4)_{ETC} \to {\rm SU}(3)_{ETC}$ at $\Lambda_2 \simeq 50-100$ TeV and
307: finally ${\rm SU}(3)_{ETC} \to {\rm SU}(2)_{TC}$ at $\Lambda_3 \simeq$ few TeV,
308: with SU(2)$_{HC}$ unbroken. Another sequence, S2, involves a breaking
309: ${\rm SU}(4)_{ETC} \to {\rm SU}(2)_{ETC}$ at $\Lambda_{23} \simeq 50$ TeV and
310: ${\rm SU}(2)_{HC} \to {\rm U}(1)_{HC}$ at a scale $\Lambda_{BHC} \lsim
311: \Lambda_1$. In all cases, SU(2)$_{TC}$ is an exact gauge symmetry.
312: At the lowest scale, $\Lambda_{TC}$, the technifermion condensates form,
313: breaking electroweak symmetry and giving masses to the $W$ and $Z$.
314:
315:
316: Certain fermion condensates contribute to nondiagonal ETC gauge boson
317: propagator corrections and hence mixing. For example, in sequence $S_1$, the
318: mixing $V^\tau_1 \to V^\tau_3$, $\tau =4,5$ is induced by the graph in Fig.
319: \ref{ckmfig10},
320: %
321: \begin{figure}[hbtp]
322: \begin{center}
323: \includegraphics[3in, 8in][4in, 9.5in]{ckmfig10.ps}
324: \end{center}
325: \vspace{-1mm}
326: \caption{Graph for $V^\tau_1 \to V^\tau_3$ with $\tau=4,5$, for sequence S1.}
327: \label{ckmfig10}
328: \end{figure}
329: %
330: with the result
331: %
332: \beq
333: {}^\tau_3 \Pi^\tau_1(0) \simeq \frac{\alpha_{_{ETC}}}{4\pi} \, \int \,
334: (k^2 dk^2) \frac{k^4 \, \Sigma_3(k)^2}{[k^2 + \Sigma_3(k)^2]^4} \ ,
335: \eeq
336: %
337: where $\Sigma_3 \simeq \Lambda_3$. This yields
338: ${}^\tau_3 \Pi^\tau_1(0) \simeq {\rm const.} \times \Lambda_3^2$.
339: In sequence $S_2$, the mixing
340: $V^\tau_2 \to V^\tau_3$, $\tau =4,5$ is induced by the graph in Fig.
341: \ref{ckmfig14},
342: %
343: \begin{figure}[hbtp]
344: \begin{center}
345: \includegraphics[3in, 8in][4in, 9.5in]{ckmfig14.ps}
346: \end{center}
347: \vspace{-1mm}
348: \caption{Graph for $V^\tau_2 \to V^\tau_3$ with $\tau=4,5$, for sequence
349: S2.}
350: \label{ckmfig14}
351: \end{figure}
352: %
353: giving ${}^\tau_3 \Pi^\tau_2(0) \simeq {\rm const.} \times \Lambda_{23}^2$.
354: We find that the feature that nondiagonal ETC gauge boson propagator
355: corrections ${}^\tau_i \Pi^\tau_j(0)$ are proportional to the square of the
356: lowest ETC scale (or smaller) is generic in this type of ETC model, reflecting
357: a type of approximate generational symmetry. Other ETC gauge boson mixings are
358: similarly suppressed.
359:
360:
361: \section{Fermion Masses and Mixing}
362:
363: Figure \ref{mfij} shows a one-loop graph contributing to the mass matrix
364: element $M^{(f)}_{ij}$ for a SM fermion $f$ (up- or down-type quark or charged
365: lepton) appearing in the operator $\bar f_{i,L} M^{(f)}_{ij} f_{j,R} + h.c.$.
366: In this figure we distinguish the first three ETC indices, which refer to SM
367: fermion generations, and the indices 4,5 which are TC, by denoting the latter
368: as $\tau=4,5$.
369: %
370: \begin{figure}[hbtp]
371: \begin{center}
372: \includegraphics[3in, 8in][4in, 9.5in]{ffndiag.ps}
373: \end{center}
374: \vspace{-1mm}
375: \caption{Graph contributing to the fermion mass matrix element
376: $M^{(f)}_{ij}$.}
377: \label{mfij}
378: \end{figure}
379: %
380: An estimate for diagonal entries is (with no sum on $i$)
381: %
382: \beq
383: M^{(f)}_{ii} \simeq \frac{\kappa \, (N_{TC}/2) \, \eta \,
384: \Lambda_{TC}^3} {\Lambda_i^2} \ ,
385: \eeq
386: %
387: where $\kappa \simeq O(10)$ is a numerical factor from the integral and in WTC,
388: $\eta \simeq \Lambda_3/\Lambda_{TC}$. This is only a rough estimate, since
389: ETC coupling is strong, so higher-order diagrams are also important.
390: The sequential breaking of the ETC symmetry at the highest scale $\Lambda_1$,
391: the intermediate scale $\Lambda_2$, and the lowest scale $\Lambda_3$, thus
392: produces the generational hierarchy in the fermion masses.
393:
394:
395: Insertions of the nondiagonal ETC propagator corrections (indicated by the
396: cross on the gauge boson line in Fig. \ref{mfij}) give rise to
397: off-diagonal elements of the $M^{(f)}$. These have the form
398: %
399: \beq
400: M^{(f)}_{ij} \simeq \frac{\kappa \, \eta \, \Lambda_{TC}^3 \, {}^j_\tau
401: \Pi^i_\tau}{\Lambda_i^2 \Lambda_j^2} \ .
402: \eeq
403: %
404: Although the resultant mixing is not fully realistic, this model shows how not
405: just diagonal but also off-diagonal elements of SM fermion mass matrices
406: $M^{(f)}$ and hence CKM mixing, could arise dynamically \cite{ckm}. Further
407: corrections to $M^{(f)}$ arise from SM gauge interactions, in particular,
408: SU(3)$_c$ and U(1)$_Y$, and from direct diagonal ETC gauge boson exchanges, in
409: particular, $V_{d3}$ (having mass $\sim \Lambda_3$ and corresponding to the
410: SU(5)$_{ETC}$ generator $T_{d3}=(2\sqrt{3})^{-1}{\rm diag}(0,0,-2,1,1)$).
411:
412: \section{Mechanism for Light Neutrinos}
413:
414: An old puzzle in TC/ETC theories was how to explain light neutrino masses. A
415: solution to this was given in \cite{nt} and analyzed further in \cite{lrs,ckm}.
416: The $\alpha_{1j,R}$ with $j=2,3$ are right-handed electroweak-singlet neutrinos
417: and get induced Dirac neutrino mass terms connecting with $(n^1,n^2,n^3)_L =
418: (\nu_e,\nu_\mu,\nu_\tau)_L$. These Dirac masses $\bar n_{i,L} M_D
419: \alpha_{1j,R}+ h.c.$ cannot be generated by the usual one-loop ETC graphs that
420: produce diagonal quark and charged lepton masses and are thus suppressed. The
421: $\alpha_{1j,R}$ also have induced Majorana mass terms $\alpha_{1i,R}^T C r_{ij}
422: \alpha_{1j,R}+h.c.$ For example, with sequence S2, denoting the relevant Dirac
423: mass terms as $b_{ij}$, one finds $b_{ij} \simeq \kappa \Lambda_{TC}^4/
424: \Lambda_{23}^3 \simeq O(0.1)$ MeV for $2 \le i,j \le 3$; further, $r_{23}
425: \simeq \kappa \Lambda_{BHC}^3 \Lambda_{23}^3/\Lambda_1^5$. Numerically,
426: $|r_{23}| \simeq O(10^2)$ GeV. The resultant electroweak-nonsinglet neutrinos
427: are, to very good approximation, linear combinations of three mass eigenstates,
428: with $\nu_3$ mass
429: %
430: \beq
431: m(\nu_3) \simeq \frac{(b_{23}+b_{22})^2}{r_{23}} \simeq
432: \frac{\kappa \Lambda_{TC}^8 \Lambda_1^5}{\Lambda_{23}^9 \Lambda_{BHC}^3} \ .
433: \label{mnu3}
434: \eeq
435: %
436: With the above-mentioned numerical values and
437: $\Lambda_{BHC} \lsim \Lambda_1$, we find $m(\nu_3) \simeq 0.05$ eV,
438: consistent with experimental results on neutrino oscillations.
439: Similarly,
440: %
441: \beq
442: \frac{m(\nu_2)}{m(\nu_3)} =
443: \bigg ( \frac{b_{23}-b_{22}}{b_{23}+b_{22}} \bigg )^2 \ ,
444: \label{mnu2}
445: \eeq
446: %
447: which is again consistent with experimental data. Since $|r_{23}| >>
448: |b_{ij}|$, this is a seesaw, but quite different from the SUSY GUT seesaw; the
449: Majorana masses $r_{ij}$ that underly the seesaw are not GUT-scale and are
450: actually much smaller than the ETC scales $\Lambda_i$.
451:
452:
453: \section{Constraints from Neutral Flavor-Changing Current Processes}
454:
455:
456: An early concern was that ETC interactions would lead to excessively large
457: flavor-changing neutral current processes. A particularly severe constraint
458: arises from $K^0 - \bar K^0$ mixing. Early treatments wrote the effective
459: Lagrangian for this as
460: %
461: \beq
462: {\cal L}_{eff} \simeq \frac{1}{\Lambda_{ETC}^2} [\bar s \gamma_\mu d]^2 \ ,
463: \eeq
464: %
465: where $\Lambda_{ETC}$ was a generic ETC scale. To suppress FCNC effects
466: adequately, it was thought that $\Lambda_{ETC}$ had to be so high that there
467: would be excessive suppression of SM fermion masses. However, we have shown,
468: using our reasonably UV-complete theories, that in the present type of ETC
469: theory this old view was too pessimistic; ETC contributions to FCNC processes
470: are smaller than had been inferred with the above naive ${\cal L}_{eff}$,
471: because of residual approximate generational symmetries \cite{ckm}-\cite{kt}.
472:
473:
474: The coupling of the ETC gauge bosons to the SM fermion mass eigenstates is
475: given by
476: %
477: \beq
478: {\cal L}_{int} = g_{_{ETC}} \sum_{f,j,k}
479: \bar f_j \gamma_\lambda ({\cal V}^\lambda)^j_k f^k
480: \equiv g_{_{ETC}} \sum_{f,j,k}
481: \bar f_{m,j} \gamma_\lambda (A^\lambda)^j_k f^k_m \ ,
482: \eeq
483: %
484: where ${\cal V}$ is a matrix containing the ETC gauge fields, $A^\lambda \equiv
485: U^{(f)} {\cal V}^\lambda U^{(f) \ -1}$, and the $U^{(f)}$ are the unitary
486: transformations that diagonalize $M^{(f)}$ via $U^{(f)} M^{(f)} U^{(f) \ -1} =
487: M^{(f)}_{diag.}$. We parametrize each $U^{(f)}$ with a PDG-type parametrization
488: in terms of $\theta_{12}^{(f)}$, $\theta_{13}^{(f)}$, $\theta_{23}^{(f)}$, and
489: $\delta^{(f)}$. The approximate generational symmetry in the ETC sector can
490: naturally yield relatively small CKM mixing with $\theta^{(f)}_{jk}$ depending
491: on ratios of smaller to larger ETC scales.
492:
493: We have analyzed ETC contributions to a number of processes, including the
494: neutral meson mixings $K^0 - \bar K^0$, $D^0 - \bar D^0$, $B^0_d - \bar B_d$,
495: $B^0_s - \bar B_s$, and decays such as $\mu^+ \to e^+ e^+ e^-$, involving
496: four-fermion operators \cite{ckm,kt}. As an example, consider $K^0 - \bar K^0$
497: mixing and the resultant $K_L - K_S$ mass difference $\Delta m_{K_L K_S}$. The
498: SM contribution is consistent with the experimental value, $\Delta m_{K_L
499: K_S}/m_K = 0.70 \times 10^{-14}$. A key to the suppression is the fact that, in
500: terms of ETC eigenstates, an $s \bar d$ in a $\bar K^0$ produces a $V^2_1$ ETC
501: gauge boson, but this cannot directly yield a $d \bar s$ in the final-state
502: $K^0$; the latter is produced by a $V^1_2$. Hence, this requires either the
503: ETC gauge boson mixing $V^2_1 \to V^1_2$ or the mixing of ETC quark eigenstates
504: to produce mass eigenstates. The contribution from $V^2_1 \to V^1_2$ yields a
505: coefficient
506: %
507: \beq
508: c \ \sim \ \frac{1}{\Lambda_1^2} \ {}^1_2 \Pi^2_1 \ \frac{1}{\Lambda_1^2}
509: \ \sim \ \frac{\Lambda_3^2}{\Lambda_1^2} \frac{1}{\Lambda_1^2} \ \ll \
510: \frac{1}{\Lambda_1^2} \ .
511: \label{ccalc}
512: \eeq
513: %
514: With above values, $\Lambda_1 \sim 10^3$ TeV, $\Lambda_3 \sim 3$ TeV, the
515: suppression factor is $(\Lambda_3/\Lambda_1)^2 \simeq 10^{-5}$. So, rather
516: than the naive result $\Delta m_{K_L K_S}/m_K \sim
517: \Lambda_{QCD}^2/\Lambda_1^2$, this yields the much smaller result $\Delta
518: m_{K_L K_S}/m_K \sim \Lambda_3^2 \, \Lambda_{QCD}^2/\Lambda_1^4 \sim 10^{-18}$.
519: Hence, the dominant ETC contributions arise from the mixing of ETC eigenstates
520: of quarks to form mass eigenstates. First, $s \bar d$ can couple to $V_{d2}$
521: (having mass $\simeq \Lambda_2$ and corresponding to the SU(5)$_{ETC}$
522: generator $T_{d2}=(2\sqrt{6})^{-1}{\rm diag}(0,-3,1,1,1)$), with coefficient
523: $\theta^{(d)}_{12}$. The resultant diagram with exchange of this gauge boson
524: $V_{d2}$ contributes $\sim (\theta^{(d)}_{12})^2/\Lambda_2^2$ to the amplitude.
525: Requiring this to be small relative to SM contribution yields the constraint
526: $|\theta_{12}^{(d)}| \lsim 10^{-2}$. Second, $s \bar d$ can couple to $V_{d3}$,
527: with coefficient $\theta^{(d)}_{13}\theta^{(d)}_{23}$. The resultant diagram
528: contributes $\sim (\theta^{(d)}_{13}\theta^{(d)}_{23})^2/ \Lambda_3^2$ to the
529: amplitude. This yields the constraint $|\theta_{13}^{(d)}\theta_{23}^{(d)}|
530: \lsim 0.4 \times 10^{-3}$. A comprehensive analysis of constraints from
531: processes involving dimension-six four-fermion operators was given in
532: \cite{ckm,kt}. A related analysis is \cite{lm}.
533:
534: We have also studied ETC contributions to (dimension-5) diagonal and transition
535: lepton and quark dipole moments, and constraints from limits on $\mu^+ \to e^+
536: \gamma$, $\tau^+ \to \ell^+ \gamma$, the measured muon $(g-2)$ and $b \to s
537: \gamma$ decay, and limits on lepton, neutron, atomic electric dipole moments
538: \cite{dml}. Again, we found that this type of ETC model can be consistent
539: with existing data.
540:
541: \section{Precision Electroweak Constraints}
542:
543: One may ask what the momentum dependence of a SM fermion mass is in this type
544: of theory. This question was answered in \cite{sml};
545: the running mass $m_{f_j}(p)$ exhibits the power-law decay
546: %
547: \beq
548: m_{f_j}(p) \propto \frac{\Lambda_j^2}{p^2}
549: \label{mfp}
550: \eeq
551: %
552: for Euclidean momenta $p \gg \Lambda_j$, where $f_j$ is a fermion of generation
553: $j$. (Here we neglect logarithmic factors, which are subdominant relative to
554: this power-law falloff.) Thus, $m_t(p)$ and $m_b(p)$ decay like
555: $\Lambda_3^2/p^2$ for $p \gg \Lambda_3$, while $m_u(p)$ and $m_d(p)$ are hard
556: up to the much higher scale $\Lambda_1$. We have investigated whether
557: precision electroweak data are consistent with these power-law decays of SM
558: fermion masses and have found that they are. The largest effects occur for
559: $m_t$. Consider, e.g., the $t$-quark contribution to $\rho$. The conventional
560: (hard, one-loop) result for this is $(\Delta \rho )_{t,hard} \simeq 3 G_f
561: m_t^2/(8 \pi^2 \sqrt{2})$. The power-law decay of $m_t$ above $\Lambda_3$
562: changes this to $(\Delta \rho)_{t} = (\Delta \rho)_{t,hard}[ 1 - a_\rho
563: (m_t^2/\Lambda_3^2)]$, where $a_\rho$ is positive, $\sim O(1)$. The softness
564: of $m_t$ thus slightly reduces the violation of custodial symmetry. We find a
565: similarly small change in the $(t,b)$ contribution to $S$ relative to the
566: conventional hard-mass result.
567:
568: As noted, the $S$ parameter places a stringent constraint on TC/ETC theories.
569: A naive perturbative calculation gives, for the TC contribution to $S$, the
570: result $S_{pert.} \simeq N_{TC} \, N_D/(6\pi)$, where $N_D$ denotes
571: the number of technifermion EW doublets (given that the dynamical masses of the
572: technifermions in each EW doublet are nearly degenerate, as should be true to
573: satisfy the $\rho$-parameter constraint). However, this calculation assumes
574: that the technifermions are weakly interacting, whereas actually, they are
575: strongly interacting, at the relevant scale, $\sim m_Z$; hence, this
576: perturbative formula cannot be expected to be reliable. This is
577: important, since otherwise, even with the minimal value, $N_{TC}=2$, it would
578: yield an excessively large value of $S_{pert.} = 4/(3\pi) \simeq 0.4$
579: for a one-family TC model (and the marginally acceptable value of
580: $S_{pert.} = 1/(3\pi) \simeq 0.1$ for a TC model with one EW technifermion
581: doublet. Experimentally, $S \lsim 0.1$. Several studies have found that
582: walking can reduce $S$ \cite{scalc}-\cite{ads},\cite{s}. In
583: particular, Refs. \cite{hky} and \cite{s} have used solutions of
584: Schwinger-Dyson and Bethe-Salpeter equations to evaluate $S$ \cite{mk}.
585:
586:
587: One way to reduce $S$ would be to reconsider TC models with only one
588: technifermion electroweak doublet. But with this technifermion content alone,
589: these models would not have the walking behavior that is needed not just to
590: reduce $S$ but also to generate adequate fermion masses. One can maintain the
591: necessary walking behavior in a TC model with one EW doublet of technifermions
592: by adding a requisite set of SM-singlet technifermions \cite{ts}. The
593: SU(2)$_{TC}$ theory should have $N_f \simeq 8$ Dirac technifermions for
594: walking. The electroweak-nonsinglet technifermions comprise two, so we need
595: six more. One adds six SM-singlet, Dirac fermions (i.e., 12 chiral components)
596: that transform as 2's under SU(2)$_{TC}$. These do not contribute to $S$ or
597: $T$. Note that in this theory, $[G_{ETC}, G_{SM}] \ne 0$, so the ETC gauge
598: bosons carry color and charge. Embedding TC in ETC is thus more complicated
599: than for one-family TC models. Another approach is to use technifermions in
600: higher-dimensional representations \cite{lanerep,fs}, \cite{ts}. In general,
601: the $S$ constraint remains a concern for TC/ETC theories.
602:
603: \section{Splitting of $t$ and $b$ Masses}
604:
605: Another challenge for TC/ETC models is to account for the splitting of the $t$
606: and $b$ masses without excessive contributions to $\rho-1$ (violation of
607: custodial SU(2) symmetry). One cannot do this by having $\Sigma_{TC,U}$
608: significantly larger than $\Sigma_{TC,D}$ since this would violate the
609: custodial symmetry too much. One could consider trying to achieve this
610: splitting using a class of ETC models in which left and right components of
611: up-type quarks and techniquarks transform the same way under SU(5)$_{ETC}$, but
612: the left and right components of down-type quarks and techniquarks transform
613: according to relatively conjugate representations. However, we showed that
614: such models have serious problems with excessively large FCNC's \cite{kt}. For
615: example, consider a model in which $Q_L$ and $u_R$ are 5's of SU(5)$_{ETC}$ and
616: $d_R$ is a $\bar 5$. So, e.g., the $\bar K^0 - K^0$ transition can proceed
617: directly via $s_L \bar d_L \to V^2_1 \to d_R \bar s_R$ without ETC gauge boson
618: mixing. This gives too large a value for $\Delta m_{K_L K_S}$.
619:
620:
621: A different idea would be to use two ETC gauge groups, say ${\rm
622: SU}(5)_{ETC} \times {\rm SU}(5)_{ETC}'$ such that the left- and right-handed
623: components of charge $Q=2/3$ quarks transform under the same ETC group, while
624: left- and right-handed components of charge $-1/3$ quarks and charged leptons
625: transform under different ETC groups \cite{aes,met}.
626: These models thereby suppress the masses $m_b$ and $m_\tau$ relative to $m_t$,
627: etc. because generating $m_b$ requires mixing between the two ETC groups, which
628: is suppressed, while $m_t$ does not. However, they tend to produce too much
629: suppression of the masses of first- and second-generation down-type quarks and
630: charged leptons \cite{met}. Thus, a satisfactory explanation of $t-b$
631: splitting appears to remain a challenge for ETC models.
632:
633: \section{Dynamical Breaking of Higher Gauge Symmetries}
634:
635: To what extent can one embed a TC/ETC in a theory having higher gauge symmetry,
636: using dynamical symmetry breaking to break this higher symmetry? Such higher
637: unification would be desirable in order to explain features not explained by
638: the standard model, including charge quantization, prediction of relative sizes
639: of gauge couplings, and unification of quarks and leptons. In \cite{lrs}
640: we constructed asymptotically free gauge theories exhibiting dynamical
641: breaking of the left-right, strong-electroweak gauge group
642: %
643: \beq
644: G_{LR} = {\rm SU}(3)_c \times {\rm SU}(2)_L \times {\rm SU}(2)_R \times
645: {\rm U}(1)_{B-L}
646: \label{glr}
647: \eeq
648: %
649: (where $B$ and $L$ denote baryon and lepton number) and
650: its extension to
651: %
652: \beq
653: G_{422}={\rm SU}(4)_{PS} \times {\rm SU}(2)_L \times {\rm SU}(2)_R \ ,
654: \label{g422}
655: \eeq
656: %
657: where ${\rm SU}(4)_{PS} \supset {\rm SU}(3)_c \times {\rm U}(1)_{B-L}$ is the
658: Pati-Salam group. These models technicolor for electroweak breaking, and
659: extended technicolor for the breaking of $G_{LR}$ and $G_{422}$ and the
660: generation of fermion masses, including a seesaw mechanism for neutrino masses.
661: These models explain why $G_{LR}$ and $G_{422}$ break to ${\rm SU}(3)_c \times
662: {\rm SU}(2)_L \times {\rm U}(1)_Y$, and why this takes place at a scale ($\sim
663: 10^3$ TeV) which is large compared to the electroweak scale. In particular,
664: the model with the gauge group $G_{422}$ achieves charge quantization in the
665: context of dynamical breaking of all symmetries.
666:
667: We have also investigated various more ambitious unification schemes involving
668: TC/ETC \cite{tg}. In particular, we have studied the possibility of unifying a
669: one-family technicolor group with the SM gauge group in a simple group, but
670: find that it appears difficult to obtain the requisite symmetry breaking
671: dynamically.
672:
673:
674: \section{Other Phenomenology}
675:
676: There are several other relevant topics for discussion. One has to do with the
677: spectrum of these theories. As a confining gauge theory, technicolor produces
678: a spectrum of TC-singlet techni-hadrons composed of technifermions and
679: technigluons. This spectrum exhibits differences in a walking theory, as
680: compared with a QCD-like theory. For example, while $m_{a_1}/m_\rho = 1.6$ in
681: QCD, the analogous ratio $m_{(a_1)_{TC}}/m_{\rho_{TC}}$ is close to unity in
682: the walking limit \cite{mm}. This is natural, since a theory with walking
683: behavior has $N_f$ near to, although slightly less than, the critical value
684: $N_{f,cr}$ at which the theory would go over from the confined phase with
685: spontaneous chiral symmetry breaking to a chirally symmetric phase. It is of
686: interest to investigate how these and other hadron masses change as one
687: decreases $N_f$ (hence increases $\alpha_{IR}$) to move from the walking regime
688: near $N_{f,cr}$ toward the QCD-like regime at smaller $N_f$; this was done in
689: \cite{sg}. Searches for the production and decays of these technihadrons at the
690: CERN Large Hadron Collider will be of great importance in testing technicolor
691: theories. For example, by analogy with $\rho \to \pi \pi$ in QCD, since the
692: technipions are absorbed to become the longitudinal components of the $W$ and
693: $Z$, one would have $\rho_{TC}^0 \to W_L^+ W_L^-$, $\rho_{TC}^+ \to W_L^+
694: Z_L^0$, etc. A related issue concerns pseudo-Nambu Goldstone bosons. While
695: walking raises the masses of many of these PNGB's, they can present a
696: phenomenological challenge for the model. It will be natural to search further
697: for these at the LHC.
698:
699: \section{Summary}
700:
701: The question of the origin of electroweak symmetry is still not answered, and
702: dynamical EWSB via technicolor and extended technicolor remains an
703: interesting possibility. This approach is strongly constrained by
704: flavor-changing neutral current data and precision electroweak measurements.
705: We have reviewed here some recent progress on TC/ETC models. Clearly, there
706: are a number of challenges for such TC/ETC models. Soon, experiments at the
707: LHC will show whether these ideas are realized in nature.
708:
709:
710: \section*{Acknowledgments}
711: I would like to thank my collaborators on works discussed here, T. Appelquist
712: and M. Piai and, in subsequent papers, N. Christensen and M. Kurachi. I also
713: thank M. Harada, M. Tanabashi, and K. Yamawaki, for organizing this SCGT06
714: conference. Our research was partially supported by the grant
715: NSF-PHY-03-54776.
716:
717: \begin{thebibliography}{40}
718:
719: \bibitem{tc}
720: S. Weinberg, Phys. Rev. D {\bf 19}, 1277 (1979);
721: L. Susskind, Phys. Rev. D {\bf 20}, 2619 (1979); see also
722: S. Weinberg, Phys. Rev. D {\bf 13}, 974 (1976).
723:
724: \bibitem{etc}
725: S. Dimopoulos and L. Susskind, Nucl. Phys. B {\bf 155}, 237 (1979);
726: E. Eichten and K. Lane, Phys. Lett. B {\bf 90}, 125 (1980).
727:
728: \bibitem{etcrev}
729: Some recent reviews are K. Lane, hep-ph/0202255; C. Hill and E. Simmons,
730: Phys. Rep. {\bf 381}, 235 (2003);
731: R. S. Chivukula, M. Narain, J. Womersley, in http://pdg.lbl.gov
732:
733: \bibitem{nt}
734: T. Appelquist and R. Shrock, Phys. Lett. {\bf B548}, 204 (2002).
735:
736: \bibitem{lrs}
737: T. Appelquist and R. Shrock, Phys. Rev. Lett. {\bf 90}, 201801 (2003).
738:
739: \bibitem{ckm}
740: T. Appelquist, M. Piai, and R. Shrock, Phys. Rev. D {\bf 69}, 015002 (2004).
741:
742: \bibitem{dml}
743: T. Appelquist, M. Piai, and R. Shrock, Phys. Lett. B {\bf 593}, 175 (2004);
744: {\it ibid.}, {\bf 595}, 442 (2004).
745:
746: \bibitem{kt}
747: T. Appelquist, N. Christensen, M. Piai, and R. Shrock, Phys. Rev. D {\bf 70},
748: 093010 (2004).
749:
750: \bibitem{sml}
751: N. Christensen and R. Shrock, Phys. Rev. Lett. {\bf 94}, 241801 (2005).
752:
753: \bibitem{tg}
754: N. Christensen and R. Shrock, Phys. Rev. D {\bf 72}, 035013 (2005).
755:
756: \bibitem{ts}
757: N. Christensen and R. Shrock, Phys. Lett. B {\bf 632}, 92 (2006).
758:
759: \bibitem{met}
760: N. Christensen and R. Shrock, Phys. Rev. D {\bf 74}, 015004 (2006).
761:
762: \bibitem{s}
763: M. Kurachi and R. Shrock, Phys. Rev. D {\bf 74}, 056003 (2006).
764:
765: \bibitem{sg}
766: M. Kurachi and R. Shrock, JHEP {\bf 12}, 034 (2006).
767:
768: \bibitem{mk}
769: M. Kurachi, in these SCGT06 proceedings.
770:
771: \bibitem{ssvz}
772: P. Sikivie, L. Susskind, M. Voloshin, and V. Zakharov, Nucl. Phys. B
773: {\bf 173}, 189 (1980).
774:
775: \bibitem{onefamily}
776: See, e.g., E. Farhi and L. Susskind, Phys. Rept. {\bf74}, 277 (1981).
777:
778: \bibitem{wtc1}
779: B. Holdom, Phys. Lett. B {\bf 150}, 301 (1985).
780:
781: \bibitem{wtc2}
782: K. Yamawaki, M. Bando, and K. Matumoto, Phys. Rev. Lett. {\bf
783: 56}, 1335 (1986).
784:
785: \bibitem{chipt1}
786: T. Appelquist, D. Karabali, and L. C. R. Wijewardhana, Phys. Rev. Lett. {\bf
787: 57}, 957 (1986); T. Appelquist and L. C. R. Wijewardhana, Phys. Rev. D
788: {\bf 35}, 774 (1987); Phys. Rev. D {\bf 36}, 568 (1987).
789:
790: \bibitem{at94}
791: T. Appelquist and J. Terning, Phys. Rev. D {\bf 50}, 2116 (1994).
792:
793: \bibitem{chipt2}
794: T. Appelquist, J. Terning, and L. C. R. Wijewardhana,
795: Phys. Rev. Lett. {\bf 77}, 1214 (1996);
796: T. Appelquist, A. Ratnaweera, J. Terning, and
797: L. C. R. Wijewardhana, Phys. Rev. D {\bf 58}, 105017 (1998).
798:
799: \bibitem{my}
800: V. Miransky and K. Yamawaki, Phys. Rev. D {\bf 55}, 5051 (1997).
801:
802: \bibitem{chiv}
803: R. S. Chivukula, Phys. Rev. D {\bf 55}, 5238 (1997).
804:
805: \bibitem{pt}
806: M. Peskin and T. Takeuchi, Phys. Rev. Lett. {\bf 65}, 964 (1990);
807: Phys. Rev. D {\bf 46}, 381 (1992).
808:
809: \bibitem{lm}
810: K. Lane and A. Martin, Phys. Rev. D {\bf 71}, 015011 (2005).
811:
812: \bibitem{scalc}
813: R. Cahn and M. Suzuki, Phys. Rev. D {\bf 44}, 3641 (1991);
814: T. Appelquist and G. Triantaphyllou, Phys. Lett. B {\bf 278}, 345 (1992);
815: R. Sundrum and S. Hsu, Nucl. Phys. B {\bf 391}, 127 (1993);
816: T. Appelquist and F. Sannino, Phys. Rev. D {\bf 59}, 067702 (1999);
817: S. Ignjatovic, L. C. R. Wijewardhana, and T. Takeuchi, Phys. Rev.
818: D {\bf 61}, 056006 (2000).
819:
820: \bibitem{hky}
821: M. Harada, M. Kurachi, and K. Yamawaki, Prog. Theor. Phys. {\bf 115}, 765
822: (2006).
823:
824: \bibitem{ads}
825: D. K. Hong and H.-U. Yee, Phys. Rev. D {\bf 74}, 015011 (2006);
826: J. Hirn and V. Sanz, Phys. Rev. Lett. {\bf 97}, 121803 (2006);
827: M. Piai, hep-ph/0608241.
828:
829: \bibitem{lanerep}
830: E. Eichten and K. Lane, Phys. Lett. B {\bf 222}, 274 (1988).
831:
832: \bibitem{fs}
833: D. Hong, S. Hsu, and F. Sannino, Phys. Lett. B {\bf 597}, 89 (2004);
834: F. Sannino and K. Tuominen, Phys. Rev. D {\bf 71}, 051901 (2005).
835:
836: \bibitem{aes}
837: T. Appelquist, N. Evans, and S. Selipsky, Phys. Lett. B {\bf 374}, 145 (1996).
838:
839: \bibitem{mm}
840: % meson masses in large N_f QCD
841: M. Harada, M. Kurachi, and K. Yamawaki, Phys. Rev. D {\bf 68}, 076001 (2003).
842:
843: \end{thebibliography}
844:
845: \end{document}
846: