hep-ph0603149/met.tex
1: \documentclass[aps,prd,eqsecnum,twocolumn,amsfonts,showpacs]{revtex4}
2: \usepackage{graphics}
3: 
4: \setlength{\unitlength}{1cm}
5: 
6: \newcommand{\beq}{\begin{equation}}
7: \newcommand{\eeq}{\end{equation}}
8: \newcommand{\beqs}{\begin{eqnarray}}
9: \newcommand{\eeqs}{\end{eqnarray}}
10: \newcommand{\lsim}{\mathrel{\raisebox{-
11: .6ex}{$\stackrel{\textstyle<}{\sim}$}}}
12: \newcommand{\gsim}{\mathrel{\raisebox{-
13: .6ex}{$\stackrel{\textstyle>}{\sim}$}}}
14: \newcommand{\pslash}{p\hspace{-0.067in}\slash}
15: \newcommand{\qslash}{q\hspace{-0.067in}\slash}
16: \newcommand{\kslash}{k\hspace{-0.067in}\slash}
17: \newcommand{\drawsquare}[2]{\hbox{%
18: \rule{#2pt}{#1pt}\hskip-#2pt%  left vertical
19: \rule{#1pt}{#2pt}\hskip-#1pt%  lower horizontal
20: \rule[#1pt]{#1pt}{#2pt}}\rule[#1pt]{#2pt}{#2pt}\hskip-#2pt%  upper horizontal
21: \rule{#2pt}{#1pt}}% right vertical
22: \newcommand{\fund}{\raisebox{-.5pt}{\drawsquare{6.5}{0.4}}}%  fund
23: \newcommand{\sym}{\raisebox{-.5pt}{\drawsquare{6.5}{0.4}}\hskip-0.4pt%
24:         \raisebox{-.5pt}{\drawsquare{6.5}{0.4}}}%  symmetric second rank
25: \newcommand{\asym}{\raisebox{-3.5pt}{\drawsquare{6.5}{0.4}}\hskip-6.9pt%
26:         \raisebox{3pt}{\drawsquare{6.5}{0.4}}}%  antisymmetric second rank
27: 
28: \begin{document}
29: 
30: \title{Extended Technicolor Models with Two ETC Groups}
31: 
32: 
33: \author{Neil D. Christensen}
34: %\thanks{neil.christensen@sunysb.edu}
35: \author{Robert Shrock}
36: %\thanks{robert.shrock@sunysb.edu}
37: 
38: \affiliation{C.N. Yang Institute for Theoretical Physics \\
39: State University of New York, Stony Brook, NY 11794}
40: 
41: \begin{abstract}
42: 
43: We construct extended technicolor (ETC) models that can produce the large
44: splitting between the masses of the $t$ and $b$ quarks without necessarily
45: excessive contributions to the $\rho$ parameter or to neutral flavor-changing
46: processes.  These models make use of two different ETC gauge groups, such that
47: left- and right-handed components of charge $Q=2/3$ quarks transform under the
48: same ETC group, while left- and right-handed components of charge $-1/3$ quarks
49: and charged leptons transform under different ETC groups.  The models thereby
50: suppress the masses $m_b$ and $m_\tau$ relative to $m_t$, and $m_s$ and $m_\mu$
51: relative to $m_c$ because the masses of the $Q=-1/3$ quarks and charged leptons
52: require mixing between the two ETC groups, while the masses of the $Q=2/3$
53: quarks do not.  A related source of the differences between these mass
54: splittings is the effect of the two hierarchies of breaking scales of the two
55: ETC groups.  We analyze a particular model of this type in some detail.
56: Although we find that this model tends to suppress the masses of the first two
57: generations of down-type quarks and charged leptons too much, it gives useful
58: insights into the properties of theories with more than one ETC group. 
59: 
60: \end{abstract}
61: 
62: \pacs{14.60.PQ, 12.60.Nz, 14.60.St}
63: 
64: %\keywords{technicolor}
65: 
66: \maketitle
67: 
68: \vspace{16mm}
69: 
70: \newpage
71: \pagestyle{plain}
72: \pagenumbering{arabic}
73: 
74: \section{Introduction}
75: \label{intro}
76: 
77: 
78: It is possible that electroweak symmetry breaks via the formation of a bilinear
79: condensate of fermions subject to a new, asymptotically free, strong gauge
80: interaction, generically called technicolor (TC) \cite{tc}.  To communicate
81: this symmetry breaking to the standard model (technisinglet) fermions, one
82: embeds technicolor in a larger, extended technicolor (ETC) theory \cite{etc}
83: (for some reviews, see \cite{etcrev}). To satisfy constraints from
84: flavor-changing neutral-current (FCNC) processes, the ETC vector bosons that
85: mediate generation-changing transitions must have large masses. To produce the
86: hierarchy in the masses of the observed three generations (families) of
87: fermions, the ETC vector boson masses have a hierarchical spectrum reflecting
88: the sequential breaking of the ETC gauge symmetry on mass scales ranging from
89: approximately $10^3$ TeV down to the TeV level. These theories are tightly
90: constrained by precision electroweak measurements \cite{pdg,lepewwg}.  Modern
91: technicolor theories are designed so that as the energy scale decreases, the
92: gauge coupling becomes large but runs very slowly (``walks''); this behavior
93: enhances masses of standard-model (SM) fermions and pseudo-Nambu-Goldstone
94: bosons \cite{wtc,chiralpt}.  Models of dynamical electroweak symmetry breaking
95: are very ambitious in their goals, which include dynamical generation of not
96: just the $W$ and $Z$ masses but also the entire spectrum of quark and lepton
97: masses.  It is thus understandable that no fully realistic models of this type
98: have been developed.  One longstanding challenge for these models has been to
99: account for the large splitting between the masses of the $t$ and $b$ quarks
100: without producing excessively large contributions to the parameter
101: $\rho=m_W^2/(m_Z^2 \cos^2\theta_W)$ or to neutral flavor-changing processes.
102: 
103: In this paper we shall formulate general classes of models that can plausibly
104: achieve this goal.  These models are based on the general idea suggested in
105: Ref. \cite{met1}, namely to obtain the splitting of $m_t$ and $m_b$ without
106: overly large contributions to $\rho$ by using two ETC groups, $G_{ETC}$ and
107: $G_{ETC}'$ and arranging that the masses of quarks of charge 2/3 involve only
108: the exchange of gauge bosons belonging to $G_{ETC}$, while the masses of quarks
109: of charge $-1/3$ arise from diagrams involving gauge bosons of both $G_{ETC}$
110: and $G_{ETC}'$, requiring mixing between these two sets of gauge bosons.  The
111: necessity of this mixing suppresses the masses of the down-type quarks relative
112: to those of up-type quarks.  From among the general classes of ETC models, we
113: construct and study in some detail one explicit ETC model embodying this idea,
114: including an analysis of the sequential breaking of the ETC gauge symmetry that
115: is necessary to produce the hierarchy in the three standard-model fermion
116: generations.  We also extend the mechanism to leptons so that the mass of the
117: charged lepton in each generation is suppressed relative to that of the up-type
118: quark.  A related factor in producing intragenerational mass splittings is the
119: fact that, for a given generation $j$, the respective breaking scales
120: $\Lambda_j$ and $\Lambda_j'$ of the $G_{ETC}$ and $G_{ETC}'$ gauge symmetries
121: may be somewhat different.  We shall focus on a model with one standard-model
122: family of technifermions (with an additional technineutrino) and also remark on
123: models in which the left- and right-handed chiral components of the
124: technifermions transform, respectively as one SU(2)$_L$ doublet and two
125: SU(2)$_L$ singlets. We also comment on neutrino masses.  Although the model is
126: rather complicated, we believe that it is useful as an explicit, moderately
127: ultraviolet-complete realization, of the strategy of using two different ETC
128: groups to account for the splitting between $m_t$ and $m_b$.
129: 
130: 
131: Before proceeding, we briefly review some past efforts to address the problem
132: of $t$-$b$ mass splitting in ETC theories.  One approach used a one-family
133: technicolor model and SU(2)$_L$-singlet, charge $-1/3$ vectorlike $b'$ quarks
134: which mix with the $b$ quark and reduce its mass relative to that of the $t$
135: quark \cite{at94}.  However, this model and similar ones with $b'$ quarks do
136: not satisfy the criteria for the diagonality of the hadronic weak neutral
137: current in terms of mass eigenstates, viz., that all quarks of a given
138: chirality have the same weak isospin $T$ and $T_3$ \cite{gw}.  Consequently,
139: such models generically can have problems with excessively large contributions
140: to hadronic flavor-changing neutral current processes.  A different approach to
141: the problem of splitting the $t$ and $b$ masses while maintaining acceptably
142: small corrections to the $\rho$ parameter used a single ETC gauge group with
143: different ETC representations for the left- and right-handed chiral components
144: of the down-type quarks and charged leptons \cite{ssvz,jt,ckm,kt}.  However, as
145: we showed in Refs. \cite{ckm,kt}, this approach also encounters problems with
146: (i) excessive suppression of down-quark and charged lepton masses, and (ii)
147: excessively large contributions to flavor-changing neutral current processes,
148: in particular, $K^0-\bar K^0$ mixing. Related discussions are contained in our
149: Refs. \cite{dml,qdml}.
150: 
151: This paper is organized as follows.  In Section II we describe the general
152: structure of one-family models with two ETC gauge groups. In Section III we
153: construct and study a specific model of this type, including the sequential
154: breakings of the ETC symmetry groups and the generation of quark and lepton
155: masses. Section IV contains remarks on some other phenomenological issues such
156: as flavor-changing neutral current processes, neutrino masses, and the 
157: minimization of the $S$ parameter. Section V contains our conclusions.
158: 
159: 
160: \section{Models with Two ETC Gauge Groups}
161: \label{generalmodels}
162: 
163: \subsection{Gauge Group}
164: 
165: We take the technicolor group to be SU($N_{TC}$) with the minimal nonabelian
166: value, $N_{TC}=2$.  There are several reasons for this choice: (i) it reduces
167: technicolor contributions to the electroweak $S$ parameter describing heavy
168: fermion loop corrections to the $Z$ boson propagator
169: \cite{pt,scalc1,scalc2,sred} (a perturbative estimate of which is proportional
170: to $N_{TC}$ for technifermions in the fundamental representation of
171: SU($N_{TC}$)); (ii) with eight or nine vectorially coupled Dirac
172: technifermions, it can plausibly have the desired walking behavior
173: \cite{wtc,chiralpt}; and (iii) it makes possible a mechanism for obtaining
174: light neutrino masses \cite{nt,lrs}. The walking behavior can occur naturally
175: as a result of an approximate infrared-stable fixed point which is larger than,
176: but close to, a critical value at which the technicolor theory would go over
177: from a confined phase with spontaneous chiral symmetry breaking to a nonabelian
178: Coulomb phase \cite{wtc,chiralpt}.
179: 
180: 
181: We take the technifermions to transform according to the fundamental
182: representation of SU(2)$_{TC}$ and focus mainly on models in which they
183: comprise one standard-model family.  This SU(2)$_{TC}$ arises dynamically from
184: the sequential breaking of the two ETC groups.  As will be shown below, the
185: last stage of this sequence, which yields the technicolor group, entails the
186: breaking of a direct product group ${\rm SU}(2)_{ETC} \times {\rm
187: SU}(2)'_{ETC}$ to its diagonal subgroup.  To embed this direct product group
188: in the two respective larger ETC groups, taken to be $G_{ETC} = {\rm
189: SU}(N_{ETC})$ and $G_{ETC}' = {\rm SU}(N_{ETC}')$, one gauges the generational
190: indices and combines them with the two sets of SU(2) group indices, leading to
191: the relation
192: %
193: \beq
194: N_{ETC}=N_{ETC}' = N_{gen.}+N_{TC} \ , 
195: \label{nrel}
196: \eeq
197: %
198: where $N_{gen.}=3$ is the number of SM fermion generations.  With $N_{TC}=2$,
199: this then yields 
200: %
201: \beq
202: G_{ETC} = {\rm SU}(5)_{ETC} \ , \quad G_{ETC}' = {\rm SU}(5)'_{ETC} \ . 
203: \label{getcgetcp}
204: \eeq
205: %
206: Thus, in this class of models the meaning of a standard-model generation is
207: somewhat different from the meaning in a conventional ETC model with only a
208: single ETC group; here, for down-type quarks and charged leptons, there are
209: really two kinds of generations for the two chiralities, corresponding to
210: different ETC gauge groups.  We shall also use additional strongly coupled
211: gauge interactions to produce the desired sequential ETC symmetry-breaking
212: pattern. These include two SU(2) hypercolor gauge interactions, each
213: corresponding to a different ETC group, denoted SU(2)$_{HC}$ and SU(2)$'_{HC}$
214: \cite{hc}.  The ETC symmetry breaking occurs in sequential stages: (i)
215: SU(5)$_{ETC}$ breaks to SU(4)$_{ETC}$ at a scale denoted $\Lambda_1$ and
216: SU(5)$'_{ETC}$ breaks to SU(4)$'_{ETC}$ at a scale $\Lambda_1'$, where the
217: subscript 1 is assigned because it is at this stage that the first-generation
218: quarks and leptons split off from the remaining four components in fundamental
219: representations of SU(5)$_{ETC}$ and SU(5)$'_{ETC}$; (ii) SU(4)$_{ETC}$ and
220: SU(4)$'_{ETC}$ break to SU(3)$_{ETC}$ and SU(3)$'_{ETC}$, respectively, at the
221: lower scales scales $\Lambda_2$ and $\Lambda_2'$ where the second-generation
222: fermions split off from the remaining three components of the above
223: representations; and (iii) SU(3)$_{ETC}$ and SU(3)$'_{ETC}$ break to
224: SU(2)$_{ETC}$ and SU(2)$'_{ETC}$, respectively, at the lower scales 
225: $\Lambda_3$ and $\Lambda_3'$, where third-generation fermions split off from
226: the remaining two components of the above representations. 
227: 
228: A basic requirement for these models is that there must be a mechanism for
229: communicating between SU(5)$_{ETC}$ and SU(5)$'_{ETC}$ in order to produce an
230: SU(2)$_{TC}$ group that leads to a technifermion condensate.  This mechanism
231: must involve all of the five indices of both the SU(5)$_{ETC}$ and
232: SU(5)$'_{ETC}$ in order to give masses to all three generations of fermions.
233: {\it A priori}, one could consider several possible ways of producing this
234: communication, thereby defining corresponding classes of models.  In the models
235: that we shall study here, this communication is achieved by the use of a
236: strongly coupled gauge interaction, which we shall call metacolor (MC), which
237: mediates between the ETC and ETC$'$ groups.  In the explicit models considered
238: here, we take the metacolor gauge group to be SU(2)$_{MC}$.  The communication
239: is effected by metacolor-induced condensates of a set of SM-singlet fermions
240: that transform as nonsinglets under SU(5)$_{ETC}$ and SU(2)$_{MC}$ with another
241: set that transform as nonsinglets under SU(5)$'_{ETC}$ and SU(2)$_{MC}$.  We
242: denote this class of ETC models as EMC for ``ETC with MC'' \cite{efc}.
243: 
244: 
245: In an EMC type of ETC model, starting with a low-energy effective field theory
246: involving descendents of the original ETC groups, which we will denote as
247: $H_{ETC} \times H_{ETC}'$, with $H_{ETC} \subset {\rm SU}(5)_{ETC}$ and
248: $H_{ETC}' \subset {\rm SU}(5)_{ETC}'$, the metacolor interaction produces
249: the breaking $H_{ETC} \times H_{ETC}' \to {\rm SU}(2)_{TC}$.  We denote the
250: scale of this breaking as $\Lambda_{MC}$, which is also the scale where the
251: metacolor gauge interaction becomes strong.  This scale should be smaller than
252: each of the lowest generational symmetry-breaking scales in the two respective
253: ETC sector, which are involved in the dynamical production of third-generation
254: fermion masses, $\Lambda_3$ and $\Lambda_3'$, since if this were not the case,
255: i.e., if there were a single ETC group operative at a scale $\ge \Lambda_3$,
256: then the basic mechanism considered here for splitting the $t$ and $b$ masses
257: would not be operative.  Since $\Lambda_3 < \Lambda_j$ for $j=1,2$ and
258: $\Lambda_3' < \Lambda_j'$, $j=1,2$, it follows that
259: %
260: \beq
261: \Lambda_{MC} < {\rm min}(\Lambda_3, \ \Lambda_3') \ . 
262: \label{lammcineq}
263: \eeq
264: %
265: In turn, this implies that in the above discussion, 
266: $H_{ETC} = {\rm SU}(2)_{ETC}$ and $H_{ETC}' = {\rm SU}(2)_{ETC}'$. 
267: 
268: Thus, the respective full gauge group of the fundamental theory is 
269: %
270: \beqs
271: & & G =  {\rm SU}(5)_{ETC} \times {\rm SU}(5)'_{ETC} \cr\cr
272: & \times &  
273: {\rm SU}(2)_{HC} \times {\rm SU}(2)'_{HC}
274:  \times {\rm SU}(2)_{MC} \times G_{SM}
275: \label{g}
276: \eeqs
277: %
278: where $G_{SM} = {\rm SU}(3)_c \times {\rm SU}(2)_L \times {\rm U}(1)_Y$ is the
279: standard-model gauge group with $N_c=3$. 
280: 
281: 
282: \subsection{Standard-Model Fermion Content}
283: 
284: We next discuss the choices of standard-model fermion representations under the
285: ETC and ETC$'$ gauge groups.  We assign the left-handed quark and techniquark
286: SU(2)$_L$ doublets to transform as a fundamental representation of
287: SU(5)$_{ETC}$ and a singlet under SU(5)$'_{ETC}$.  (Our choice of a fundamental
288: rather than conjugate fundamental representation of SU(5)$_{ETC}$ is a
289: convention.) Since we want the up-type and down-type quark masses in each of
290: the second and third generations to be unsuppressed and suppressed,
291: respectively, it follows that the right-handed components of the up- and
292: down-type quarks and techniquarks should be assigned to the (5,1) and (1,5)
293: representations of ${\rm SU}(5)_{ETC} \times {\rm SU}(5)'_{ETC}$, respectively.
294: The fermions that are nonsinglets under $G_{SM}$ are singlets under the
295: hypercolor and the metacolor group.  This then determines the
296: quark sector of our model, which is displayed below, where the numbers indicate
297: the representations under the nonabelian factor groups in $G$, and the
298: subscript gives the weak hypercharge $Y$:
299: %
300: \beqs 
301: & & Q_L: \ (5,1; \{1\};3,2)_{1/3,L} \cr\cr
302: & & u_R: \ (5,1; \{1\};3,1)_{4/3,R} \cr\cr
303: & & d_R: \ (1,5; \{1\};3,1)_{-2/3,R}
304: \label{1fam_quarks} 
305: \eeqs
306: %
307: where $\{1\}$ means singlet under ${\rm SU}(2)_{HC} \times {\rm SU}(2)'_{HC}
308: \times {\rm SU}(2)_{MC}$.  Here and in the rest of the paper we use a compact
309: notation in which, for example,
310: %
311: \beqs
312: u_R \equiv u^{aj}_R & \equiv & (u^{a1},u^{a2},u^{a3},u^{a4},u^{a5})_R \cr\cr
313:                     & \equiv & (u^a,c^a,t^a,U^{a4},U^{a5})_R
314: \label{uexample}
315: \eeqs
316: %
317: where $a$ and $j$ are color and SU(5)$_{ETC}$ indices, and so forth for the
318: other fields.  One could also consider a model in which $d_R$ transforms as
319: $(1,\bar 5;1,1,1;3,1)_{-2/3,R}$ rather than $(1,5;1,1,1;3,1)_{-2/3,R}$. This
320: would entail further reduction of the down-quark masses; we shall focus here on
321: the choice in eq. (\ref{1fam_quarks}).  (In passing, we note that the
322: assignment of $d_R$ to $(\bar 5,1;1,1,1;3,1)_{-2/3,R}$ would produce a model
323: similar to the one that we studied in Refs. \cite{ckm,kt}; while this would
324: succeed in intragenerational reduction of down-quark masses relative to
325: up-quark masses, it would yield overly large ETC contributions to
326: flavor-changing neutral current processes.)
327: 
328: In order to suppress the charged lepton mass relative to the up-quark mass in
329: each generation, we shall make the left- and right-handed components of the
330: charged leptons transform according to different ETC groups. Given our
331: restriction to fundamental and conjugate fundamental representations, 
332: we shall consider two different cases, which we label as $L1$ and $L2$:
333: %
334: \beqs
335: {\rm case} \ L1: & & L_L: \ (1,5;\{1\};1,2)_{-1,L} \cr\cr
336: & & e_R: \ (5,1;\{1\};1,1)_{-2,R} 
337: \label{1fam_leptons1} 
338: \eeqs
339: %
340: and 
341: %
342: \beqs
343: {\rm case} \ L2: & & L_L: \ (1,\bar 5;\{1\};1,2)_{-1,L} \cr\cr
344: & & e_R: \ (\bar 5,1;\{1\};1,1)_{-2,R}  \ . 
345: \label{1fam_leptons2} 
346: \eeqs
347: %
348: One can then characterize a given model as being of type $L1$ or $L2$. 
349: 
350: 
351: \subsection{Anomaly Constraints on SM-Singlet Fermion Content} 
352: 
353: A requirement in the construction of these models is the absence of any local
354: gauge anomalies and, for the SU(2) groups, also the absence of any global
355: $\pi_4$ anomalies.  For the model with lepton assignments of type L1,
356: SM-nonsinglet fermions and technifermions contribute the following terms to
357: these gauge anomalies: (i) for the cubic SU(5)$_{ETC}$ anomaly (written for
358: right-handed chiral components), $A(Q^c_R)=-2N_c=-6$, $A(u_R)=N_c=3$, and
359: $A(e_R)=1$, for a total of $-2$; (ii) for the cubic SU(5)$'_{ETC}$ anomaly
360: (again for right-handed chiral components): $A(L^c)_R=-2$ and $A(d_R)=3$, for a
361: total of 1. Here and below, we always write standard-model singlet fields as
362: right-handed.  We cancel the above cubic anomalies for SU(5)$_{ETC}$ and
363: SU(5)$'_{ETC}$ with SM-singlet, SU(5)$_{ETC}$-nonsinglet fermions $f_R$, which
364: contribute the amounts
365: %
366: \beq 
367: {\rm case} \ L1: \ 
368: \sum_{f_R} A(f_R) =  2 \ , \quad \sum_{\tilde f_R} A'(\tilde f_R) =  -1 \ . 
369: \label{arsum_1fam_L1}
370: \eeq
371: %
372: For the model with lepton assignment L2, by similar reasoning, we
373: have the constraint
374: %
375: \beq 
376: {\rm case} \ L2: \ 
377: \sum_{f_R} A(f_R)=4 \ , \quad \sum_{\tilde f_R} A'(\tilde f_R)=-5 \ .
378: \label{arsum_1fam_L2}
379: \eeq
380: %
381: One can check that these models also have zero SU(5)$^2$U(1)$_Y$ and 
382: SU(5)$'^2$U(1)$_Y$ gauge anomalies. 
383: 
384: Before proceeding, we remark on the corresponding constraints for gauge anomaly
385: cancellation in the effective field theories that result from the sequential
386: breaking of the ${\rm SU}(5)_{ETC} \times {\rm SU}(5)'_{ETC}$ symmetry to ${\rm
387: SU}(N)_{ETC} \times {\rm SU}(N')'_{ETC}$, where $N$ and $N'$ decrease through
388: the values 4 and 3.  The origin of the conditions
389: (\ref{arsum_1fam_L1})-(\ref{arsum_1fam_L2}) is the contributions of the
390: standard-model fermion multiplets to the respective ETC and ETC$'$ anomalies,
391: and since these SM-fermion ETC multiplets are fundamental representations for
392: each of the descendent ETC and ETC$'$ groups, their anomaly contributions
393: remain the same.  It follows that for each of the low-energy effective field
394: theories invariant under the various ${\rm SU}(N)_{ETC} \times {\rm
395: SU}(N')'_{ETC}$ gauge groups with $N$ and $N'$ taking on values 4 and 3, the
396: generalizations of conditions (\ref{arsum_1fam_L1})-(\ref{arsum_1fam_L2}) hold,
397: where now $A$ and $A'$ denote the respective contributions to the ${\rm
398: SU}(N)_{ETC}$ and ${\rm SU}(N')'_{ETC}$ anomalies from the massless SM-singlet
399: fermions $f_R$ and $\tilde f_R$ which are nonsinglets under these two groups.
400: (That is, in the respective sums $\sum_{f_R}$ and $\sum_{\tilde f_R}$, one has
401: removed fermions that have gained dynamical masses due to the formation of
402: condensates at higher scales.)  The SM-singlet sectors of the descendant
403: low-energy effective field theories resulting from the sequential symmetry
404: breaking of the ETC and ETC' symmetries satisfy these anomaly constraints, as
405: can be checked explicitly.
406: 
407: 
408: \section{A Model of Type $L1$}
409: \label{model1}
410: 
411: \subsection{Standard-Model Singlet Fermion Content} 
412: 
413: 
414: Here we construct and analyze a model of type $L1$. 
415: We must first choose an SM-singlet fermion sector that
416: satisfies the anomaly constraints of eq. (\ref{arsum_1fam_L1}).  Here we use
417: one relatively simple solution to these constraints for the SM-singlet
418: fermions:
419: %
420: \beqs
421: & & \psi^{ij}_R \ : \            (10,1;1,1,1;1,1)_{0,R} \cr\cr 
422: & & {\cal N}^i_R \: \             (5,1;1,1,1;1,1)_{0,R} \cr\cr 
423: & & \zeta^{ij,\alpha}_R \  :     (10,1;2,1,1;1,1)_{0,R} \cr\cr
424: & & \omega^\alpha_{p,R} \ :    2(1,1;2,1,1;1,1)_{0,R} \cr\cr
425: & & \chi_{i,\lambda,R} \ : \   (\bar 5,1;1,1,2;1,1)_{0,R} 
426: \label{f_EMMEL1}
427: \eeqs
428: %
429: and
430: %
431: \beqs
432: & & \tilde \psi_{i'j',R} \ : \ (1,\overline{10};1,1,1;1,1)_{0,R} \cr\cr
433: & & \tilde \zeta_{i'j',\alpha',R} \ : \ (1,\overline{10};1,2,1;1,1)_{0,R}\cr\cr
434: & & \tilde \omega_{\alpha',p,R} \ : \quad 2(1,1;1,2,1;1,1)_{0,R} \cr\cr
435: & & \tilde \chi^{i',\lambda}_R \ : \ (1,5;1,1,2;1,1)_{0,R} \ , 
436: \label{ftilde_EMMEL1}
437: \eeqs
438: %
439: where $1 \le i,j \le 5$ are SU(5)$_{ETC}$ indices, $1 \le i',j' \le 5$ are
440: SU(5)$'_{ETC}$ indices, $\alpha=1,2$ are SU(2)$_{HC}$ indices, $\alpha'=1,2$
441: are SU(2)$'_{HC}$ indices, $\lambda=1,2$ are SU(2)$_{MC}$ indices, and $p=1,2$
442: refer to the two copies of the fields $\omega^\alpha_{p,R}$ and $\tilde
443: \omega^{\alpha'}_{p,R}$. (It is necessary to have an even number of copies of
444: the $\omega^\alpha_{p,R}$ and $\tilde \omega^\alpha_{p,R}$ fields in order to
445: avoid global $\pi_4$ anomalies in the SU(2)$_{HC}$ and SU(2)$'_{HC}$ gauge
446: sectors.)  In the above equations, the tilde denotes SM-singlet,
447: SU(5)$_{ETC}$-singlet, SU(5)$'_{ETC}$-nonsinglet fermions, and we use the fact
448: that the representations of SU(2) are (pseudo)real.  At certain points below,
449: we shall denote technicolor indices by $t$ to distinguish them from
450: generational indices $j,k \in \{1,2,3\}$.
451: 
452: With this content of massless fermions, the SU(5)$_{ETC}$, SU(5)$'_{ETC}$,
453: SU(2)$_{HC}$, SU(2)$'_{HC}$, and SU(2)$_{MC}$ interactions are all
454: asymptotically free.  The leading coefficients of the various beta functions
455: are given by \cite{beta} 
456: %
457: \beq
458: b_0^{(SU(5)_{ETC})}=11 \ , \quad b_0^{(SU(5)'_{ETC})}=13 \ ,
459: \label{b0etc_model1}
460: \eeq
461: %
462: %
463: \beq
464: b_0^{(SU(2)_{HC})}=b_0^{(SU(2)'_{HC})}=\frac{10}{3} \ , 
465: \label{b0hc_model1}
466: \eeq
467: %
468: and
469: %
470: \beq
471: b_0^{(SU(2)_{MC})}=4 \ . 
472: \label{b0mc_model1}
473: \eeq
474: %
475: The beta functions for the other two nonabelian groups in $G$, SU(3)$_c$,
476: SU(2)$_L$, are also asymptotically free, with leading coefficients
477: $b_0^{(SU(3)_c)}=13/3$ and $b_0^{(SU(2)_L)}=2/3$.  These values of beta
478: function coefficients for standard-model gauge groups also hold for all of the
479: other 1-family models considered here, since they all have the same
480: content of SM-nonsinglet fermions. 
481: 
482: 
483: \subsection{General Structure of ETC Gauge Symmetry Breaking} 
484: 
485: The ETC and ETC$'$ gauge symmetries are chiral, so that when they become
486: strong, sequential breaking of each occurs naturally \cite{tumb}.  This
487: breaking also involves additional strongly coupled gauge interactions.  The
488: breakings of the SU(5)$_{ETC}$ to SU(2)$_{ETC}$ and of SU(5)$'_{ETC}$ to
489: SU(2)$'_{ETC}$ are driven by the condensation of SM-singlet fermions.  The
490: SM-singlet fermions that condense at a given scale acquire dynamical masses of
491: order this scale and hence decouple from the effective theory at lower
492: energies.
493: 
494: We identify plausible preferred condensation channels using a generalized
495: most-attractive-channel (GMAC) approach that takes account of one or more
496: strong gauge interactions at each breaking scale, as well as the energy cost
497: involved in producing gauge boson masses when gauge symmetries are broken. An
498: approximate measure of the attractiveness of a channel $R_1 \times R_2 \to
499: R_{cond.}$ is $\Delta C_2 = C_2(R_1)+C_2(R_2)-C_2(R_{cond.})$ \ \cite{gap},
500: where $R_j$ denotes the representation under a relevant gauge interaction and
501: $C_2(R)$ is the quadratic Casimir invariant \cite{casimir}.
502: 
503: \subsection{SU(5)$_{ETC} \to$ SU(4)$_{ETC}$ and SU(5)$'_{ETC} \to$ 
504: SU(4)$'_{ETC}$ Breaking} 
505: 
506: As the energy scale decreases from high values, the SU(5)$_{ETC}$ coupling
507: increases, as governed by the coefficient $b_0^{(SU(5)_{ETC})}$ in
508: eq. (\ref{b0etc_model1}).  We envision that at an energy scale $\Lambda_1$ of
509: order $10^3$ TeV, $\alpha_{_{ETC}}$ becomes sufficiently large to cause a
510: condensate in the channel
511: %
512: \beqs
513: & & (10,1;1,1,1;1,1)_{0,R} \times (10,1;1,1,1;1,1)_{0,R}
514: \to \cr\cr
515: & & \to (\bar 5,1;1,1,1;1,1)_0
516: \label{1010channel_model1}
517: \eeqs
518: %
519: with $\Delta C_2 = 24/5$, breaking ${\rm SU}(5)_{ETC}$ to ${\rm SU}(4)_{ETC}$.
520: This is a most attractive channel; i.e., there is no other channel with a
521: higher value of $\Delta C_2$ \cite{other1}. With no loss of generality, we take
522: the breaking direction in SU(5)$_{ETC}$ as $i=1$, thereby splitting off the
523: first-generation up quarks from the remaining components of the corresponding
524: SU(5) fields with indices $i \in \{2,3,4,5\}$.  With respect to the unbroken
525: ${\rm SU}(4)_{ETC}$, we have the decomposition $10 = 4 + 6$. (Note that $6
526: \approx \bar 6$ in SU(4).)  We denote the fundamental representation of
527: SU(4)$_{ETC}$, $(4,1;1,1,1;1,1)_{0,R}$, as $\alpha^{1i}_R \equiv \psi^{1i}_R$
528: for $2 \le i \le 5$, and the antisymmetric tensor representation
529: $(1,6;1,1,1;1,1)_{0,R}$ as $\xi^{ij}_R \equiv \psi^{ij}_R$ for $2 \le i,j \le
530: 5$. The associated condensate is
531: %
532: \beq 
533: \langle \epsilon_{1 i j k \ell} \xi^{ij \ T}_R C \xi^{k \ell}_R \rangle
534: = 8\langle 
535: \xi^{23 \ T}_R C \xi^{45}_R - 
536: \xi^{24 \ T}_R C \xi^{35}_R +
537: \xi^{25 \ T}_R C \xi^{34}_R \rangle \ . 
538: \label{xixicondensate_model1}
539: \eeq
540: %
541: where here and below, summations over repeated indices are understood.  Linear
542: combinations of the six $\xi^{ij}_R$ fields involved in this condensate pick up
543: masses of order $\Lambda_1$. 
544: 
545: Similarly, at a comparable scale $\Lambda_1' \simeq 10^3$ TeV, we envision that
546: $\tilde \alpha_{_{ETC}}$ becomes sufficiently strong to produce condensation in
547: the channel
548: %
549: \beqs
550: & & (1,\overline{10};1,1,1;1,1)_{0,R} \times 
551:     (1,\overline{10},;1,1,1;1,1)_{0,R} \to \cr\cr
552: & & \to (1,5;1,1,1;1,1)_0
553: \label{10b-10bchannel_model1}
554: \eeqs
555: %
556: breaking SU(5)$'_{ETC}$ to SU(4)$'_{ETC}$. Again, this is a most attractive
557: channel, with $\Delta C_2=24/5$.  With no loss of generality, we take the
558: breaking direction in SU(5)$'_{ETC}$ as $i'=1$; this entails the separation of
559: the first generation of down-type quarks and charged leptons from the
560: components of the corresponding SU(5)$'_{ETC}$ fields with indices lying in the
561: set $\{2,3,4,5\}$.  In analogy with our notation for SU(4)$_{ETC}$, we denote
562: the conjugate fundamental representation $(1,\bar 4;1,1,1;1,1)_{0,R}$ and
563: antisymmetric conjugate tensor representation $(1,\bar 6;1,1,1;1,1)_{0,R}$ of
564: SU(4)$'_{ETC}$ as $\tilde \alpha_{1i,R} \equiv \tilde \psi_{1i,R}$ for $2 \le i
565: \le 5$ and $\tilde \xi_{ij,R} \equiv \tilde \psi_{ij,R}$ for $2 \le i,j \le 5$.
566: The associated SU(5)$'_{ETC}$-breaking, SU(4)$'_{ETC}$-invariant condensate is
567: %
568: \beqs
569: \langle \epsilon^{1i'j'k' \ell'}
570: \tilde \xi^T_{i'j',R} C \tilde \xi_{k' \ell',R} \rangle & = & 8\langle 
571: \tilde \xi^T_{23,R} C \tilde \xi_{45,R} - 
572: \tilde \xi^T_{24,R} C \tilde \xi_{35,R} \cr\cr  
573: & + & \tilde \xi^T_{25,R} C \tilde \xi_{34,R} \rangle \ . 
574: \label{xixitildecondensate} 
575: \eeqs
576: %
577: The six $\tilde \xi_{i'j',R}$ fields involved in this condensate pick up masses
578: of order $\Lambda_1'$. (Again, the actual mass eigenstates are linear
579: combinations of these fields; henceforth, we shall often suppress this when it
580: is not important for the discussion.)  As was true in our previous studies of
581: ETC models, at lower energy scales, different patterns of ETC breaking can
582: occur, depending on the relative strengths of the ETC and HC gauge couplings.
583: We shall focus on one pattern here.
584: 
585: \subsection{ETC Symmetry Breaking at Lower Mass Scales}
586: 
587: In the energy interval just below the lower of the two scales $\Lambda_1$ and
588: $\Lambda_1'$, the effective theory is invariant under the gauge group
589: %
590: \beqs
591: & & {\rm SU}(4)_{ETC} \times {\rm SU}(4)'_{ETC} \times {\rm SU}(2)_{HC} \times
592: {\rm SU}(2)'_{HC} \times \cr\cr
593: & & \times {\rm SU}(2)_{MC} \times G_{SM} \ . 
594: \label{g44}
595: \eeqs
596: %
597: Since the SU(4)$_{ETC}$ and SU(4)$'_{ETC}$ gauge interactions are
598: asymptotically free, the corresponding couplings $\alpha_{_{ETC}}$ and $\tilde
599: \alpha_{_{ETC}}$ continue to increase as the energy scale decreases.  As the
600: energy scale descends through the value $\Lambda_2 \simeq 50$ TeV, the
601: SU(4)$_{ETC}$ and SU(2)$_{HC}$ couplings become sufficiently strong to lead
602: together to condensation in the channel
603: %
604: \beqs
605: & & (4,1;2,1,1;1,1)_{0,R} \times (6,1;2,1,1;1,1)_{0,R} \to 
606: \cr\cr
607: & & (\bar 4,1;1,1,1;1,1)_0  \ . 
608: \label{4x6channel} 
609: \eeqs
610: %
611: This condensation channel preserves SU(2)$_{HC}$ and breaks SU(4)$_{ETC}$ to
612: SU(3)$_{ETC}$.  The Casimir operators measuring the attractiveness of this
613: channel are $\Delta C_2 = 5/2$ for SU(4)$_{ETC}$ and $\Delta C_2 = 3/2$ for 
614: SU(2)$_{HC}$.  The associated condensate is
615: %
616: \beqs 
617: & & \langle \epsilon_{\alpha\beta}\epsilon_{12jk \ell}\zeta^{1j,\alpha\ T}_R 
618: C \zeta^{k \ell,\beta}_R \rangle = 2\langle
619: \epsilon_{\alpha\beta}( \zeta^{13,\alpha \ T}_R C \zeta^{45,\beta}_R \cr\cr
620: & & -
621: \zeta^{14,\alpha \ T}_R C \zeta^{35,\beta}_R +
622: \zeta^{15,\alpha \ T}_R C \zeta^{34,\beta}_R ) \rangle \ , 
623: \label{4x6zetacondensate} 
624: \eeqs
625: %
626: and the twelve $\zeta^{ij,\alpha}_R$ fields in this condensate gain
627: masses of order $\Lambda_2$.  
628: 
629: Analogously, as the energy scale decreases through the value 
630: $\Lambda_2' \simeq 15$ TeV, the SU(4)$'_{ETC}$ and
631: SU(2)$'_{HC}$ couplings become sufficiently strong to lead together to
632: condensation in the channel
633: %
634: \beqs
635: & & (1,\bar 4;1,2,1;1,1)_{0,R} \times (1,\bar 6;1,2,1;1,1)_{0,R} \to 
636: \cr\cr
637: & & \to (1,4;1,1,1;1,1)_0
638: \label{4bx6bchannel} 
639: \eeqs
640: %
641: breaking SU(4)$'_{ETC}$ to SU(3)$'_{ETC}$.  The quadratic Casimir invariants
642: for this channel are $\Delta C_2 = 5/2$ for SU(4)$'_{ETC}$ and $\Delta C_2 =
643: 3/2$ for SU(2)$'_{HC}$.  The condensate is
644: %
645: \beqs 
646: & & \langle \epsilon^{\alpha'\beta'} \, \epsilon^{12j'k' \ell'} \, 
647: \tilde \zeta^T_{1j',\alpha',R} C \tilde \zeta_{k' \ell',\beta',R} \rangle =
648: \cr\cr & = & 2\langle \epsilon_{\alpha'\beta'}( 
649: \tilde \zeta^T_{13,\alpha',R} C \tilde \zeta_{45,\beta',R} \cr\cr 
650: & & 
651: - \tilde \zeta^T_{14,\alpha',R} C \tilde \zeta_{35,\beta',R} +
652: \tilde \zeta^T_{15,\alpha',R} C \tilde \zeta_{34,\beta',R} ) \rangle \ , \cr\cr
653: & & 
654: \label{4bx6bzetatildecondensate} 
655: \eeqs
656: %
657: and the twelve $\tilde \zeta_{i'j',\alpha',R}$ fields in this condensate gain
658: masses $\sim \Lambda_2'$. 
659: 
660: 
661: The effective theory just below $\Lambda_2'$ is invariant under the gauge
662: group
663: %
664: \beqs
665: & & {\rm SU}(3)_{ETC} \times {\rm SU}(3)'_{ETC} \times {\rm SU}(2)_{HC} \times
666: {\rm SU}(2)'_{HC} \times \cr\cr
667: & & \times {\rm SU}(2)_{MC} \times G_{SM} \ . 
668: \label{g33}
669: \eeqs
670: %
671: Since the SU(3)$_{ETC}$, SU(3)$'_{ETC}$, SU(2)$_{HC}$, SU(2)$'_{HC}$, and
672: SU(2)$_{MC}$ interactions are asymptotically free, their couplings continue to
673: increase as the energy scale decreases.  At the scale $\Lambda_3$ of a few TeV,
674: the SU(3)$_{ETC}$ and SU(2)$_{HC}$ interactions trigger condensation in the
675: channel
676: %
677: \beqs
678: & & (3,1;2,1,1;1,1)_{0,R} \times (3,1;2,1,1;1,1)_{0,R} \to \cr\cr
679: & & \to (\bar 3,1;1,1,1;1,1)_0 \ , 
680: \label{33to3bchannel} 
681: \eeqs
682: %
683: where the numbers indicate the representations under the group (\ref{g33}).
684: This condensation is invariant under SU(2)$_{HC}$ and breaks SU(3)$_{ETC}$ to
685: SU(2)$_{ETC}$.  Its attractiveness is given by the Casimir invariants
686: $\Delta C_2 = 4/3$ for SU(3)$_{ETC}$ and $\Delta C_2 = 3/2$ for SU(2)$_{HC}$.
687: Without loss of generality, we may use the original SU(3)$_{ETC}$ gauge
688: symmetry to orient the condensate so that it takes the form
689: %
690: \beq 
691: \langle \epsilon_{123jk} \epsilon_{\alpha\beta} \zeta^{2j,\alpha \ T}_R
692: C \zeta^{2k,\beta}_R \rangle \ = \ 2\langle \epsilon_{\alpha\beta}
693:  \zeta^{24,\alpha}_R C  \zeta^{25,\beta}_R \rangle \ . 
694: \label{33to3bcondensate} 
695: \eeq
696: %
697: 
698: Similarly, at a scale $\Lambda_3'$, the SU(3)$'_{ETC}$ and SU(2)$'_{HC}$
699: are envisioned to lead together to a condensation in the channel
700: %
701: \beqs
702: & & (1,\bar 3;1,2,1;1,1)_{0,R} \times (1,\bar 3;1,2,1;1,1)_{0,R} \to \cr\cr
703: & & \to (1,3;1,1,1;1,1)_0 \ , 
704: \label{3b3bto3channel} 
705: \eeqs
706: %
707: breaking SU(3)$'_{ETC}$ to SU(2)$'_{ETC}$. The condensate is 
708: %
709: \beq 
710: \langle \epsilon^{123j'k'} \epsilon^{\alpha'\beta'} 
711: \tilde \zeta^T_{2j',\alpha',R}
712: C \tilde \zeta_{2k',\beta',R} \rangle \ = \ 2\langle \epsilon^{\alpha'\beta'}
713: \tilde \zeta^T_{24,\alpha',R} C \tilde \zeta_{25,\beta',R} \rangle \ . 
714: \label{3b3bto3condensate} 
715: \eeq
716: %
717: 
718: We next discuss several additional condensations driven by the SU(2)$_{HC}$ and
719: SU(2)$'_{HC}$ interactions.  In the low-energy effective field theory just
720: below min($\Lambda_3,\Lambda_3'$), the massless SM-singlet, ETC-nonsinglet
721: fermions consist of $\zeta^{ij,\alpha}_R$ and $\tilde\zeta_{i'j',\alpha',R}$
722: with $ij=12, 23$ and $i'j'=12, 23$, together with $\omega^\alpha_{p,R}$
723: and $\tilde \omega_{\alpha',p,R}$ with $p=1,2$.  In this energy interval, the
724: SU(2)$_{ETC}$, SU(2)$'_{ETC}$, and SU(2)$_{HC}$ gauge couplings continue to
725: grow.  The hypercolor interaction naturally produces (HC-singlet) condensates
726: of the various remaining HC-doublet fermions.  In each case, $\Delta C_2 =3/2$
727: for the hypercolor interaction.  Since the condensates (\ref{33to3bcondensate})
728: and (\ref{3b3bto3condensate}) were formed via a combination of the attractive
729: SU(2)$_{HC}$ and SU(2)$'_{HC}$ interaction with, respectively, the
730: SU(3)$_{ETC}$ and SU(3)$'_{ETC}$ interactions, while the present condensates
731: are formed only by the SU(2)$_{HC}$ or SU(2)$'_{HC}$ interaction, and have the
732: same value of $\Delta C_2$ for the HC and HC$'$ groups, it follows that the
733: scales at which they form, denoted $\Lambda_s$ and $\Lambda_s'$ (where $s$
734: denotes SU(2)$_{ETC}$-singlet and SU(2)$'_{ETC}$-singlet) satisfy (i)
735: $\Lambda_s \le \Lambda_3$ and $\Lambda_s' \le \Lambda_3'$.  There are twelve
736: condensates of this type,
737: %
738: \beq 
739: \langle \epsilon_{\alpha\beta} \zeta^{12,\alpha \ T}_R C
740: \zeta^{23,\beta}_R \rangle 
741: \label{z12z23condensate} 
742: \eeq
743: %
744: \beq 
745: \langle \epsilon_{\alpha\beta} \zeta^{12,\alpha \ T}_R C
746: \omega^\beta_{p,R} \rangle
747: \label{z12omegacondensate} 
748: \eeq
749: %
750: \beq 
751: \langle \epsilon_{\alpha\beta} \zeta^{23,\alpha \ T}_R C
752: \omega^\beta_{p,R} \rangle
753: \label{z23omegacondensate} 
754: \eeq
755: %
756: \beq 
757: \langle \epsilon_{\alpha\beta} \omega^{\alpha \ T}_{1,R} C
758: \omega^\beta_{2,R} \rangle 
759: \label{omega_selfcondensate} 
760: \eeq
761: %
762: \beq 
763: \langle \epsilon^{\alpha'\beta'} \tilde \zeta^T_{12,\alpha',R}
764: C \tilde \zeta_{23,\beta',R} \rangle 
765: \label{z12z23tildecondensate} 
766: \eeq
767: %
768: \beq 
769: \langle \epsilon^{\alpha'\beta'} 
770: \tilde \zeta^T_{12,\alpha',R} C
771: \tilde \omega_{\beta',p,R} \rangle
772: \label{z12omegatildecondensate} 
773: \eeq
774: %
775: \beq 
776: \langle \epsilon^{\alpha'\beta'} 
777: \tilde \zeta^T_{23,\alpha',R} C
778: \tilde \omega_{\beta',p,R} \rangle
779: \label{z23omegatildecondensate} 
780: \eeq
781: %
782: \beq 
783: \langle \epsilon^{\alpha'\beta'} 
784: \tilde \omega^T_{\alpha',1,R} C
785: \tilde \omega_{\beta',2,R} \rangle 
786: \label{omega_tildeselfcondensate} 
787: \eeq
788: %
789: where $p=1,2$. Here we shall take $\Lambda_s \simeq \Lambda_s' \simeq
790: \Lambda_3$.
791: 
792: In the energy interval just below ${\min}(\Lambda_3, \Lambda_3')$, the
793: theory is invariant under the gauge group
794: %
795: \beqs
796: & & {\rm SU}(2)_{ETC} \times {\rm SU}(2)'_{ETC} \times {\rm SU}(2)_{HC} \times
797: {\rm SU}(2)'_{HC} \times \cr\cr
798: & & \times {\rm SU}(2)_{MC} \times G_{SM} \ . 
799: \label{g22}
800: \eeqs
801: %
802: With the fermions that have gained dynamical masses at scales $\ge \Lambda_s,
803: \Lambda_s'$ integrated out, the resultant low-energy effective field theory
804: contains 14 chiral fermions transforming as doublets under SU(2)$_{ETC}$,
805: namely
806: %
807: \beq
808: Q_L^{aj}, \ \  u_R^{aj}, \ \ e_R^j, \ \ \alpha^{1j}_R, \ \ {\cal N}^j_R, \ 
809: \chi_{j,\lambda,R} 
810: \label{su2doublets}
811: \eeq
812: %
813: and eight chiral fermions transforming as chiral doublets under
814: SU(2)$'_{ETC}$, namely
815: %
816: \beq
817: L_L^{j'}, \ \ d_R^{aj'}, \ \ \tilde \alpha_{1j',R}, \ \ 
818: \tilde \chi^{j',\lambda}_R
819: \label{su2primedoublets}
820: \eeq
821: %
822: with $j=4,5$, $j'=4,5$, and $\lambda=1,2$. 
823: 
824: As the energy scale decreases through the value $\Lambda_{MC}$, the metacolor
825: interaction gets sufficiently strong to lead to the condensation of the
826: metacolor-nonsinglet fermions.  We assume that $\Lambda_{MC}$ is of order a few
827: TeV. Applying a GMAC argument, we infer that the favored condensation channel,
828: as regards metacololor, is $2 \times 2 \to 1$, with condensate 
829: %
830: \beq 
831: \langle \sum_{\lambda=1}^2 \sum_{j,k'=1}^5 \chi^T_{j,\lambda,R} C \tilde
832: \chi^{k',\lambda}_R \rangle
833: \label{mccondensate}
834: \eeq
835: % 
836: where $\lambda$ are MC indices, and $j$ and $k'$ are SU(5)$_{ETC}$ and
837: SU(5)$'_{ETC}$ indices, respectively.  Although these latter two groups,
838: SU(5)$_{ETC}$ and SU(5)$'_{ETC}$, are no longer invariance groups of the
839: effective theory at this scale $\Lambda_{MC}$, the range of the indices in the
840: condensate (\ref{mccondensate}) is still $1 \le j,k' \le 5$ since the
841: $\chi_{j,\lambda,R}$ and $\tilde \chi^{k',\lambda}_R$ are still massless
842: fermions at this scale, and this condensate can be bound solely by the
843: SU(2)$_{MC}$ interaction.  For the components $j,k' \in \{4,5\}$, a vacuum
844: alignment argument implies that the condensate (\ref{mccondensate}) is of the
845: form ${\rm const.} \times \delta_j^{k'}$ so that it breaks ${\rm SU}(2)_{ETC}
846: \times {\rm SU}(2)'_{ETC}$ to the diagonal subgroup ${\rm SU}(2)_d \equiv {\rm
847: SU}(2)_{TC}$. The components of the $\chi$ and $\tilde \chi$ fermions involved
848: in this condensate thus gain dynamical masses of order $\Lambda_{MC}$.  The
849: three normalized diagonal linear combinations of the gauge bosons of ${\rm
850: SU}(2)_{ETC} \times {\rm SU}(2)'_{ETC}$ are the massless gauge bosons of
851: SU(2)$_{TC}$ and the three orthogonal linear combinations gain masses of order
852: $\Lambda_{MC}$.  Thus, the communication between the SU(5)$_{ETC}$ and
853: SU(5)$'_{ETC}$ groups takes place at the scale $\Lambda_{MC}$.  This
854: communication, via the condensate, (\ref{mccondensate}) achieves two important
855: goals: (i) connecting fermions with SU(5)$_{ETC}$ generation indices $j=1,2,3$
856: and fermions with SU(5)$'_{ETC}$ generation $k'=1,2,3$ indices; and (ii)
857: producing the exact SU(2)$_{TC}$ group which will break electroweak
858: interactions at a lower energy scale. 
859: 
860: The effective field theory just below $\Lambda_{MC}$ is thus invariant under
861: the gauge group
862: %
863: \beq
864: {\rm SU}(2)_{TC} \times {\rm SU}(2)_{HC} \times {\rm SU}(2)'_{HC} 
865: \times {\rm SU}(2)_{MC} \times G_{SM} \ . 
866: \label{glow}
867: \eeq
868: %
869: Since the SU(2)$_{HC}$, SU(2)$'_{HC}$, and SU(2)$_{MC}$ gauge interactions
870: confine, the particles in this effective low-energy field theory are singlets
871: under all of these groups.  With the fermions having dynamically generated
872: masses $\ge \Lambda_{MC}$ integrated out, the resultant low-energy effective
873: SU(2)$_{TC}$ theory consists of the 18 chiral doublets given by
874: eqs. (\ref{su2doublets}) and (\ref{su2primedoublets}) with the $\chi$'s
875: removed, namely (with numbers denoting representations under the group of
876: eq. (\ref{glow})),
877: %
878: \beqs
879: & & Q_L^t: \ (2;\{1\};3,2)_{1/3,L} \ , \cr\cr
880: & & U_R^t: \ (2;\{1\};3,1)_{4/3,R} \ , \quad 
881:     D_R^t: \ (2;\{1\};3,1)_{-2/3,R} \ , \cr\cr
882: & & L_L^t: \ (2;\{1\};1,2)_{-1,L} \ , \quad 
883:     E_R^t: \ (2;\{1\};1,1)_{-2,R} \ , \cr\cr
884: & & \alpha^{1t}_R, \ {\cal N}^t_R, \ \tilde \alpha_{1t,R}: \ 
885: 3(2;\{1\};1,1)_{0,R} 
886: \label{1fam_fermions_tc} 
887: \eeqs
888: %
889: where $t=4,5$ refer to SU(2)$_{TC}$ indices and, following standard notation,
890: we denote technifermions with capital letters. Neglecting the SM interactions,
891: which are weak at this scale, this theory is vectorial, with nine Dirac
892: technifermions, including three electroweak-singlet technineutrinos. (Here we
893: make use of the fact that the representations of SU(2) are (pseudo)real to
894: re-express the technineutrinos in vectorial form as regards their technicolor
895: couplings.) The technicolor gauge interaction has the necessary property of
896: asymptotic freedom, and furthermore, to within the uncertainties inherent in
897: the analysis of a strong-coupling gauge theory, one may plausibly consider that
898: it could have walking behavior, so that the technicolor coupling $\alpha_{TC}$,
899: evolves slowly in the interval below $\Lambda_{MC}$ down to the technicolor
900: condensation scale.
901: 
902: 
903: As the energy scale decreases further to the value that we shall denote 
904: $\Lambda_{TC}$, the SU(2)$_{TC}$ technicolor theory
905: produces technifermion condensates that break 
906: ${\rm SU}(2)_L \times {\rm U}(1)_Y$ to U(1)$_{em}$.  These condensates are
907: %
908: \beq
909: \langle \bar U_{a,t} U^{a,t} \rangle \ , \ 
910: \langle \bar D_{a,t} D^{a,t} \rangle \ , \ 
911: \langle \bar E_t E^t \rangle \ , \ 
912: \label{FFbar}
913: \eeq
914: %
915: \beq
916: \langle \bar N_{t,L} \alpha^{1t}_R \rangle + h.c., \ 
917: \langle \bar N_{t,L} {\cal N}^t_R \rangle + h.c., \ 
918: \langle \epsilon^{123tv} \bar N_{t,L} \tilde \alpha_{1v,R} \rangle + h.c. 
919: \label{NNbar}
920: \eeq
921: %
922: where implied sums are over the color indices $a$ and TC indices $t,v$.  In a
923: one-family technicolor model such as this, from the relation $m_W^2=(g^2/4)
924: F_{TC}^2 N_D$ with $N_D=N_c+1=4$ the number of technifermion electroweak
925: doublets, one has $F_{TC} \simeq 125$ GeV and, taking $\Lambda_{TC} \simeq 2
926: F_{TC}$, this gives $\Lambda_{TC} \simeq 250$ GeV. The technicolor interaction
927: also naturally produces Majorana condensates involving only the right-handed,
928: SM-singlet technineutrinos (which do not break electroweak symmetry), namely,
929: %
930: \beqs
931: & & \langle \epsilon_{123tv} \alpha^{1t \ T}_R C {\cal N}^{v}_R \rangle + 
932: h.c., \ \langle \alpha^{1t \ T}_R C \tilde \alpha_{1t,R} \rangle + h.c., \cr\cr
933: & & \langle {\cal N}^{t \ T}_R C \tilde \alpha_{1t,R} \rangle + h.c.
934: \label{tcmajorana}
935: \eeqs
936: %
937: The exact gauge symmetry of the theory at energies below the electroweak scale,
938: $\Lambda_{TC}$, is
939: %
940: \beqs
941: & & {\rm SU}(2)_{TC} \times {\rm SU}(2)_{HC} \times {\rm SU}(2)'_{HC} 
942: \times {\rm SU}(2)_{MC} \times \cr\cr
943: & & {\rm SU}(3)_c \times {\rm U}(1)_{em} \ . 
944: \label{exactsymmetry}
945: \eeqs
946: %
947: 
948: The technifermions $F$ involved in the condensates of
949: eqs. (\ref{FFbar})-(\ref{tcmajorana}) gain dynamical masses that are
950: generically denoted $\Sigma_F$.  Since other interactions are weaker than
951: technicolor at this scale, these technifermion condensates are expected to be
952: nearly equal for different $F$'s, and consequently so are the corresponding
953: dynamical masses $\Sigma_F$ \cite{lq}. Hence, in particular,
954: %
955: \beq
956: \frac{|\Sigma_U - \Sigma_D|}{\Sigma_U + \Sigma_D} \ll 1 \ , \quad 
957: \frac{|\Sigma_N - \Sigma_E|}{\Sigma_N + \Sigma_E} \ll 1 \ . 
958: \label{sigmadiff}
959: \eeq
960: %
961: For the overall magnitude of the common $\Sigma_F$, the estimates used, e.g.,
962: in Ref. \cite{ckm}, give $\Sigma_F \simeq 2 \Lambda_{TC} \simeq 500$ GeV.
963: 
964: Therefore, the model preserves custodial symmetry very well and can naturally
965: yield acceptably small corrections to the parameter $\rho = \alpha_{em}(m_Z)T$.
966: An estimate of the contribution in the present model gives a result similar to
967: the value that we found in Ref. \cite{ckm}.  We recall the reasoning that went
968: into that estimate.  Let us denote the TC/ETC corrections to $\rho$ as $\Delta
969: \rho$.  The one-loop ($1\ell$) contribution involving technifermions yields
970: %
971: \beq
972: (\Delta \rho )_{1\ell} \simeq  \frac{N_{TC} G_F}{8\pi^2 \sqrt{2}} \biggl [ 
973: N_c f_\rho(\Sigma_U^2,\Sigma_D^2) + f_\rho(\Sigma_N^2,\Sigma_E^2) \biggr ]
974: \label{deltarho}
975: \eeq
976: %
977: where \cite{veltman}
978: \beq
979: f_\rho(x,y)=x+y-\frac{2xy}{x-y} \, \ln \Big ( \frac{x}{y} \Big ) \ .
980: \label{fxy}
981: \eeq
982: %
983: We denote $\Sigma_U-\Sigma_D=\epsilon_{_{UD}}$ and
984: $\Sigma_N-\Sigma_E=\epsilon_{_{NE}}$.  Since among standard-model gauge
985: interactions only the weak hypercharge U(1)$_Y$ distinguishes between $U$ and
986: $D$ (and, separately, $N$ and $E$), one may estimate the standard-model
987: contribution to these technifermion mass differences as $\epsilon_{_{UD}}
988: \simeq \epsilon_{_{NE}} \simeq (\alpha_Y/\pi) \Sigma_F$, where $\alpha_Y =
989: g'^2/(4\pi)$.  With $\alpha_{em}(m_Z) \simeq 1/128$ and $\sin^2\theta_W(m_Z)
990: \simeq 0.232$, one has $\alpha_Y(m_Z) = 1.0 \times 10^{-2}$, which is also
991: approximately the value of $\alpha_Y(\mu)$ at a scale $\mu = \Lambda_{TC}$.
992: This gives a standard-model contribution $\epsilon_{_{UD}}/\Sigma_F \simeq
993: \epsilon_{_{NE}}/\Sigma_F \simeq 0.3 \times 10^{-2}$.  Other contributions
994: arise from the different manner in which the ETC and ETC$'$ interactions treat
995: the $U$ and $D$ (and $N$ and $E$) technifermions (see below). Using the Taylor
996: series expansion
997: %
998: \beq
999: f_\rho((\Sigma+\epsilon)^2,\Sigma^2) = \epsilon^2 \bigg [ \frac{4}{3} -
1000:   \frac{1}{15} \frac{\epsilon^2}{\Sigma^2} + 
1001: O \Big (  \frac{\epsilon^3}{\Sigma^3} \Big )  \bigg ] \ , 
1002: \label{fexpand}
1003: \eeq
1004: %
1005: we can express this contribution as 
1006: %
1007: \beq
1008: (\Delta \rho )_{1\ell} \simeq 
1009: \frac{N_{TC} G_F}{6 \pi^2 \sqrt{2}}(N_c \epsilon_{_{UD}}^2+\epsilon_{_{NE}}^2) 
1010: \ . 
1011: \label{deltarho_expand}
1012: \eeq
1013: %
1014: For a rough estimate, setting $\epsilon_{_{UD}} \simeq \epsilon_{_{NE}}$, we
1015: obtain $(\Delta \rho )_{1\ell} \simeq 2 N_{TC}G_F
1016: \epsilon_{_{UD}}^2/(3\pi^2\sqrt{2})$.  Using $N_{TC}=2$ and
1017: $\epsilon_{_{UD}}/\Sigma_{U,D} \simeq 10^{-2}$ then yields $(\Delta \rho
1018: )_{1\ell} \simeq 3 \times 10^{-5}$, or equivalently, $(\Delta T)_{1\ell} \simeq
1019: 4 \times 10^{-3}$, which is safely small.  The higher-lying dynamics of
1020: unbroken gauge interactions, such as metacolor, leading to the technicolor
1021: theory should be subsumed in this estimate. Next, one considers contributions
1022: involving explicit ETC and ETC$'$ gauge boson exchanges. These can be roughly
1023: estimated by recalling that the momentum scale of the technicolor mass
1024: generation mechanism is set by $\Lambda_{TC}$, and the emission and
1025: reabsorption of an ETC or ETC$'$ gauge boson will lead to a denominator factor
1026: of at most $1/\Lambda_3^2$. This yields the estimate of Ref. \cite{ckm}, namely
1027: %
1028: \beq 
1029: \Delta \rho \simeq \frac{2 b \Lambda_{TC}^2}{3 \Lambda_3^2} \ ,
1030: \label{delta_rho} 
1031: \eeq
1032: %
1033: where $b \simeq O(1)$.  Numerically, this gives $(\Delta \rho) \simeq
1034: (4 \times 10^{-3})b$ or equivalently, $T \simeq 0.5b$.  (We note that this
1035: estimate and the one in Ref. \cite{ckm} are slightly smaller than the one given
1036: in Ref. \cite{met1}.)  For $b \lsim 0.5$, this is consistent with current
1037: experimental constraints \cite{pdg,lepewwg}.  From our studies, this success in
1038: splitting $m_t$ and $m_b$ while plausibly maintaining sufficiently small
1039: corrections to the $\rho$ parameter, appears to generalize beyond just this
1040: particular EMC $L1$ model to other EMC-type ETC models with two ETC gauge
1041: groups.
1042: 
1043: \subsection{ETC Gauge Bosons}
1044: 
1045: For a SM fermion $f_\chi$ transforming as a 5 of SU(5)$_{ETC}$, the basic
1046: coupling to the SU(5)$_{ETC}$ gauge bosons (which is vectorial) is
1047: %
1048: \beq
1049: {\cal L} = g_{_{ETC}} \bar f_j (T_a)^j_k(V_a)^\lambda \gamma_\lambda f^k
1050: \label{gff}
1051: \eeq
1052: %
1053: where the $T_a$, $a=1,...,24$ are the generators of the Lie algebra of
1054: SU(5)$_{ETC}$ and the $V_a$ are the corresponding ETC gauge fields. For a
1055: fermion $f$ transforming as a 5 of SU(5)$'_{ETC}$, the coupling is the
1056: same with the replacement of $g_{_{ETC}}$ by $g_{_{ETC'}}$ and $(V_a)$ by
1057: $\tilde V_a$.  For nondiagonal transitions, $j \ne k$, it is convenient to use
1058: the fields $V^j_k = \sum_a V_{a,\lambda} (T_a)^j_k$ for SU(5)$_{ETC}$, whose
1059: absorption by $f^k$ yields $f^j$, with coupling
1060: $g_{_{ETC}}/\sqrt{2}$, analogous to the $W^\pm$ in SU(2)$_L$ and similarly for
1061: SU(5)$'_{ETC}$, with the changes noted before.  We take the diagonal (Cartan)
1062: generators for both SU(5)$_{ETC}$ and SU(5)$'_{ETC}$ to be
1063: %
1064: \beqs
1065: & & T_{24} \equiv T_{d1}=(2\sqrt{10})^{-1}{\rm diag}(-4,1,1,1,1), \cr\cr
1066: & & T_{15} \equiv T_{d2}=(2\sqrt{6})^{-1}{\rm diag}(0,-3,1,1,1), \cr\cr
1067: & & T_8    \equiv T_{d3}=(2\sqrt{3})^{-1}{\rm diag}(0,0,-2,1,1), \cr\cr
1068: & & T_3 = (1/2){\rm diag}(0,0,0,-1,1) \ .  
1069: \label{tds}
1070: \eeqs
1071: %
1072: The ETC gauge bosons that couple to these diagonal generators $T_{dj}$ are
1073: denoted $V_{dj}$ for SU(5)$_{ETC}$ and $\tilde V_{dj}$ for SU(5)$'_{ETC}$.
1074: 
1075: 
1076: 
1077: When SU(5)$_{ETC}$ breaks to SU(4)$_{ETC}$, the nine ETC gauge bosons in the
1078: coset SU(5)$_{ETC}$/SU(4)$_{ETC}$, namely, $V^1_j$, $(V^1_j)^\dagger=V^j_1$,
1079: $j=2,3,4,5$, and $V_{d1}$, gain masses $M_1 \simeq \Lambda_1$. When
1080: SU(4)$_{ETC}$ breaks to SU(3)$_{ETC}$, the seven ETC gauge bosons $V^2_j$ and
1081: $(V^2_j)^\dagger=V^j_2$, $j=3,4,5$, together with $V_{d2}$, gain masses $\simeq
1082: \Lambda_2$.  Finally, when SU(3)$_{ETC}$ breaks to SU(2)$_{TC}$, the five ETC
1083: gauge bosons $V^3_j$, $(V^3_j)^\dagger=V^j_3$, $j=4,5$, together with $V_{d3}$,
1084: gain masses $\simeq \Lambda_3$.  The analogous statements hold for
1085: SU(5)$'_{ETC}$ with the replacements of $V^j_k$ by $\tilde V^{j'}_{k'}$,
1086: $V_{dj}$ by $\tilde V_{dj'}$, and $\Lambda_j$ by $\Lambda_j'$.  The SM-singlet
1087: fermions responsible for these breakings also, through quantum loops, lead to
1088: mixing among the $V$ bosons and, separately, among the $\tilde V$ bosons, so
1089: that they are not exact mass eigenstates.  The mixing is small, being
1090: suppressed by ratios of the hierarchical ETC and ETC$'$ scales.  There is also
1091: mixing of the ETC and ETC$'$ groups, which takes place at the scale
1092: $\Lambda_{MC}$ via the condensate (\ref{mccondensate}).  This is necessary for
1093: generating down-quark and lepton masses.  A graph that contributes to this
1094: mixing is shown in Fig. \ref{chigraph}.
1095: 
1096: %
1097: \begin{figure}
1098: \begin{center}
1099: \includegraphics[3in, 8in][4in, 9.5in]{chigraph.ps}
1100: \end{center}
1101: \caption{\footnotesize{A graph contributing to the ETC gauge boson mixing
1102: $\tilde V^{k'}_{t'} \leftrightarrow V^j_t$, where $j,k' \in \{1,2,3\}$ are
1103: generational indices and $t,t' \in \{4,5\}$ connect with technicolor indices in
1104: SU(2)$_{TC}$, the diagonal subgroup of ${\rm SU}(2)_{ETC} \times {\rm
1105: SU}(2)'_{ETC}$.}}
1106: \label{chigraph}
1107: \end{figure}
1108: 
1109: 
1110: \subsection{Quark and Lepton Masses}
1111: 
1112: The effective theory describing the physics at energies $E < \Lambda_{TC}$,
1113: obtained by integrating out the ETC and TC gauge bosons and all of the heavy
1114: fermions, contains the mass matrix of the up-type quarks, 
1115: %
1116: \beq
1117: {\cal L}_u = -\bar u_{j,L} M^{(u)}_{jk} u^k_R + h.c.,
1118: \label{lmup}
1119: \eeq
1120: %
1121: and the corresponding mass matrix of the down-type quarks, 
1122: %
1123: \beq
1124: {\cal L}_d = -\bar d_{j,L} M^{(d)}_{jk'} d^{k'}_R + h.c.,
1125: \label{lmdn}
1126: \eeq
1127: %
1128: with $j,k' \in \{1,2,3\}$.  For the present model, of L1 type, the mass
1129: matrix for the charged leptons is 
1130: %
1131: \beq
1132: {\cal L}_e = -\bar e_{j',L} M^{(e)}_{j'k} e^{k}_R + h.c.,
1133: \label{lme}
1134: \eeq
1135: %
1136: with $j',k \in \{1,2,3\}$. An analogous operator, with obvious interchange of
1137: primed indices, applies for a model of type L2. 
1138: 
1139: 
1140: \begin{figure}
1141: \begin{center}
1142: \includegraphics[3.5in, 8.7in][4.25in, 9.9in]{ugraph.ps}
1143: \end{center}
1144: \caption{\footnotesize{A graph generating $\bar u_{j,L} M^{(u)}_{jk} u^k_R$
1145: where $j,k \in \{1,2,3\}$ are generational indices and $t\in \{4,5\}$ are
1146: technicolor indices. The $V^j_t$ are SU(5)$_{ETC}$ vector bosons.}} 
1147: \label{ugraph}
1148: \end{figure}
1149: 
1150: The elements of the up-type quark mass matrix arise from the diagram in
1151: Fig. \ref{ugraph}. The diagonal elements of this matrix are 
1152: %
1153: \beq
1154: M^{(u)}_{jj} \simeq \frac{\kappa \eta \Lambda_{TC}^3}{\Lambda_j^2}
1155: \label{mfu}
1156: \eeq
1157: %
1158: where $\kappa \sim O(10)$ is a numerical prefactor from the integration (see,
1159: e.g., \cite{ckm}) and $\eta$ is a walking factor
1160: %
1161: \beq
1162: \eta = \exp[\int_{\Lambda_{TC}}^{\Lambda_w} (d\mu/\mu) \gamma(\alpha(\mu))]
1163: \label{eta}
1164: \eeq
1165: %
1166: where the technicolor theory has walking behavior between $\Lambda_{TC}$ and a
1167: scale denoted $\Lambda_w$.  With the anomalous dimension $\gamma \simeq 1$ as
1168: in a walking theory, this factor is then $\eta \simeq \Lambda_w/\Lambda_{TC}$.
1169: In the current class of theories, it is plausible that there could be walking
1170: up to the scale $\Lambda_{MC}$, so that $\eta \simeq
1171: \Lambda_{MC}/\Lambda_{TC}$.  For the third-generation standard-model fermions,
1172: the entries $M^{(f)}_{33}$ should give reasonable estimates of the
1173: corresponding masses of $u^3 \equiv t$, $d^3 \equiv b$, and $e^3 \equiv \tau$
1174: which are not changed significantly by off-diagonal entries.  In particular,
1175: for the top quark, assuming the above value of $\eta$ and using the value
1176: $\Lambda_3 \simeq 3$ TeV, one obtains a value of $M^{(u)}_{33}$ and hence $m_t$
1177: that is acceptably close to the experimental pole mass $m_t \simeq 175$
1178: GeV. For $m_t$, the difference between this pole mass and the running mass
1179: evaluated at 175 GeV is negligible; for the other quarks, we use the running
1180: masses evaluated at the scale $\mu_t = 175$ GeV.  In view of the substantial
1181: uncertainties in the dynamically generated standard-model fermion masses due to
1182: the strong-coupling nature of the TC and ETC theories, we consider an estimate
1183: for a fermion mass to be acceptable if it is within a factor of 2-3 of the
1184: measured mass.  If the contributions of off-diagonal element in $M^{(u)}$ are
1185: sufficiently small, then the diagonal element $M^{(f)}_{22}$ is dominant in
1186: determining the mass of the corresponding second-generation fermions $u^2
1187: \equiv c$.  Assuming that this is the case, the value of $\Lambda_2 \simeq 45$
1188: TeV, which is consistent with the renormalization group equations for the
1189: SU(4)$_{ETC}$ theory, yields an acceptable result for $m_c$, i.e., $m_c(\mu_t)
1190: \simeq 0.6$ GeV, corresponding to the pole mass $m_c = 1.3$ GeV. These values
1191: are listed in Table \ref{scales} together with other scales in the model.  The
1192: property $m_u < m_d$ for first-generation quarks, which is the opposite of the
1193: pattern $m_c \gg m_s$, $m_t \gg m_b$ of the second and third generations,
1194: requires that some off-diagonal elements of $M^{(u)}$ and $M^{(d)}$ play an
1195: important role in determining the first-generation quark masses.
1196: 
1197: 
1198: \begin{table}
1199: \caption{\footnotesize{Summary table of ETC and ETC$'$ breaking scales and
1200: other strong-coupling scales, in units of TeV, in the model.  In the third
1201: column we list a physical quantity or constraint(s) that are highly correlated
1202: with the numerical value of the given scale. In the lines for $\Lambda_s$ and
1203: $\Lambda_s'$, $f=u,d,e$ refers to up- and down-quark quarks and charged
1204: leptons.  In the line for $\Lambda_{MC}$, the notation $\{m_d\}$, \ $\{m_e\}$
1205: indicates that $\Lambda_{MC}$ affects the values of all of the down-type quarks
1206: and charged leptons.  See text for detailed discussion.}}
1207: \begin{center}
1208: \begin{tabular}{|c|c|c|} \hline\hline
1209: scale & value & comments  \\ \hline
1210: $\Lambda_1$         & $\simeq 10^3$ & FCNC constraints          \\ \hline
1211: $\Lambda_2$         & $\simeq 45$   & $m_c$ value               \\ \hline
1212: $\Lambda_3$         & $\simeq 3$    & $m_t$ value               \\ \hline
1213: $\Lambda_1'$        & $\simeq 10^3$ & FCNC constraints          \\ \hline
1214: $\Lambda_2'$        & $\simeq 15$   & ETC hierarchy             \\ \hline
1215: $\Lambda_3'$        & $\simeq 15$   & $m_b$ value               \\ \hline
1216: $\Lambda_s$         & $\simeq 3$    & $M^{(f)}$, off-diag.      \\ \hline
1217: $\Lambda_s'$        & $\simeq 3$    & $M^{(f)}$, off-diag.      \\ \hline
1218: $\Lambda_{MC}$      & $\simeq 2$    & $\{m_d\}$, \ $\{m_e\}$    \\ \hline
1219: $\Lambda_{TC}$      & $\simeq 0.25$ & $m_W$, \ $m_Z$           \\ \hline\hline
1220: \end{tabular}
1221: \end{center}
1222: \label{scales}
1223: \end{table}
1224: 
1225: 
1226: The off-diagonal entries $M^{(u)}_{jk}$, $j \ne k$, arise via diagrams
1227: involving ETC vector boson mixing of the form $V^k_t \to V^j_t$.  This mixing
1228: is indicated by the cross on the ETC vector boson propagator in
1229: Fig. \ref{ugraph}, where it is understood that since the ETC and TC couplings
1230: are strong, further gauge boson exchanges not suppressed by large propagators
1231: are implicitly included.
1232: %
1233: \beq
1234: M^{(u)}_{jk} \simeq 
1235: \frac{\kappa \eta \ ({}^j_t \Pi^{k}_t) \Lambda_{TC}^3}{\Lambda_j^2 \Lambda_k^2}
1236: \label{mfuoffdiagonal}
1237: \eeq
1238: %
1239: where ${}^j_t \Pi^k_t$ denotes the relevant nondiagonal ETC propagator
1240: insertion that produce the transition $V^k_t \to V^j_t$.  Additional virtual
1241: SU(5)$_{ETC}$ exchanges not overly suppressed by large mass scales are
1242: understood to be included, since the corresponding gauge couplings are large.
1243: 
1244: %
1245: \begin{figure}
1246: \begin{center}
1247: \includegraphics[3.5in, 8.7in][4.25in, 9.9in]{dgraph.ps}
1248: \end{center}
1249: \caption{\footnotesize{A graph generating the mass matrix for charge $q=-1/3$
1250: quarks, $\bar d_{j,L} M^{(d)}_{jk'} d^{k'}_R$, where $j,k' \in \{1,2,3\}$ are
1251: generational indices and $t,t' \in \{4,5\}$ connect with technicolor indices in
1252: SU(2)$_{TC}$, the diagonal subgroup of ${\rm SU}(2)_{ETC} \times {\rm
1253: SU}(2)'_{ETC}$.  The $V^j_t$ and $\tilde V^{k'}_{t'}$ are SU(5)$_{ETC}$ and
1254: SU(5)$'_{ETC}$ vector bosons; since the corresponding gauge couplings are
1255: strong, this graph is understood to represent also additional exchanges of
1256: these gauge bosons, insofar as they are not suppressed by large mass scales.}}
1257: \label{dgraph}
1258: \end{figure}
1259: %
1260: 
1261: The down-quark and charged lepton masses are suppressed, since all elements of
1262: the mass matrices $M^{(d)}_{jk'}$ and $M^{(e)}_{j'k}$ require the $V-\tilde V$
1263: mixing.  The elements of the matrix $M^{(d)}$ arise from the graph in Fig.
1264: \ref{dgraph}.  We have
1265: %
1266: \beq
1267: M^{(d)}_{jk'} \simeq 
1268: \frac{\kappa \eta \ ({}^j_t \Pi^{k'}_{t'}) \Lambda_{TC}^3}
1269:  {\Lambda_j^2 \Lambda_k'^2}
1270: \label{mfd}
1271: \eeq
1272: %
1273: where ${}^j_t \Pi^{k'}_{t'}$ denotes the relevant nondiagonal ETC vector bosons
1274: propagator insertions and exchanges that produce the transition $\tilde
1275: V^{k'}_{t'} \to V^j_t$. Again, additional SU(5)$_{ETC}$ and SU(5)$'_{ETC}$
1276: exchanges not overly suppressed by large mass scales are understood to be
1277: included, since the corresponding gauge couplings are large.  Since the
1278: metacolor condensate (\ref{mccondensate}) occurs at the scale $\Lambda_{MC}$,
1279: it follows that the dynamical masses for the metacolor fermions are soft for
1280: higher momentum scales.  Performing the loop integral in Fig. \ref{chigraph},
1281: one thus finds that ${}^j_t \Pi^{j'}_{t'} \propto \Lambda_{MC}^2$ for $j=j'$
1282: (with smaller values for ${}^j_t \Pi^{k'}_{t'}$ for $j \ne k'$). 
1283: The largest eigenvalues of $M^{(u)}$ and $M^{(d)}$, i.e., the masses of $m_t$
1284: and $m_b$, are determined by the respective elements $M^{(u)}_{33}$ and
1285: $M^{(d)}_{33}$.  From eqs. (\ref{mfu}) and (\ref{mfd}), one then has the
1286: relation
1287: %
1288: \beq
1289: \frac{m_b}{m_t} \simeq \left ( \frac{\Lambda_{MC}}{\Lambda_3'} \right )^2
1290: \label{mbtratio}
1291: \eeq
1292: %
1293: since the factor of $1/\Lambda_3^2$ cancels in this ratio. 
1294: Using the pole mass $m_b \simeq 4.3$ GeV or equivalently the running mass
1295: $m_b(\mu_t) \simeq 3.0$ GeV, with $m_t \simeq 175$ GeV as above,
1296: eq. (\ref{mbtratio}) yields the ratio $\Lambda_3'/\Lambda_{MC} \simeq
1297: 7.6$.  With the illustrative value $\Lambda_{MC} \simeq 2$ TeV, this can be
1298: achieved with $\Lambda_3' \simeq 15$ TeV.  Because of the strong-coupling
1299: nature of the physics involved here, these are only rough estimates, but they
1300: demonstrate that the model can achieve the requisite $t-b$ mass splitting. 
1301: 
1302: However, the model encounters difficulty in trying to account for the masses of
1303: the standard-model fermions of the lower two generations.  To see this, we
1304: first consider the situation in the approximation where one neglects the
1305: off-diagonal terms in the mass matrices. Then the model would yield the
1306: generalization of eq. (\ref{mbtratio}), viz.,
1307: %
1308: \beq
1309: \frac{m_{d_j}}{m_{u_j}}=\bigg ( \frac{\Lambda_{MC}}{\Lambda_j'} \bigg )^2
1310: \label{mdjmujratio}
1311: \eeq
1312: %
1313: for generation $j$.  Taking ratios for $j=2$ and $j=3$, this would imply the
1314: double ratio relation
1315: %
1316: \beq
1317: \frac{(m_s/m_c)}{(m_b/m_t)}=\left ( \frac{\Lambda_3'}{\Lambda_2'} \right )^2 
1318: \label{doubleratio23}
1319: \eeq
1320: %
1321: (where all of the masses are evaluated at a common scale, taken here to be
1322: $\mu_t$).  But because $m_s/m_c > m_b/m_t$ (the values being roughly 1/10 and
1323: 1/60, respectively), this would require that $\Lambda_3' > \Lambda_2'$, which
1324: is impossible, since the sequential breaking ${\rm SU}(4)'_{ETC} \to {\rm
1325: SU}(3)'_{ETC}$ guarantees that $\tilde \Lambda_2 \ge \tilde \Lambda_3$.  This
1326: problem would be mitigated as much as possible if $\Lambda_2' \simeq
1327: \Lambda_3'$, i.e. the sequential breaking of SU(4)$'_{ETC}$ to SU(3)$'_{ETC}$
1328: and thence to SU(2)$'_{ETC}$ occurs at comparable scales (see Table
1329: \ref{scales}). 
1330: 
1331: There is also a problem with the ratios of down-quark masses of different
1332: generations.  With the same simplification of neglecting off-diagonal elements
1333: in the mass matrix $M^{(d)}$, one has
1334: %
1335: \beq
1336: \frac{m_{d_j}}{m_{d_k}} = \bigg ( \frac{\Lambda_k \Lambda_k'}
1337: {\Lambda_j \Lambda_j'} \bigg )^2 \ . 
1338: \label{djdkratio}
1339: \eeq
1340: %
1341: This yields ratios for $m_s/m_b$ and $m_d/m_s$ that are too small to fit
1342: experimental values.  One is thus motivated to consider the full down-quark
1343: mass matrix in order to assess whether this improves the predictions for the
1344: ratios (\ref{djdkratio}).  We define the ratios
1345: %
1346: \beq
1347: r_{jk}  \equiv \left ( \frac{\Lambda_j}{\Lambda_k} \right )^2 \ , \quad
1348: r'_{jk} \equiv \left ( \frac{\Lambda_j'}{\Lambda_k'} \right )^2 \ .  
1349: \label{rjk}
1350: \eeq
1351: % 
1352: We can then write eq. (\ref{mfd}) equivalently as 
1353: %
1354: \beq
1355: M^{(d)} \simeq  m_b \left ( \begin{array}{ccc}
1356:     r_{31}r'_{31}  &  r_{31}r'_{32}   &  r_{31}   \\
1357:     r_{32}r'_{31}  &  r_{32}r'_{32}   &  r_{32}   \\
1358:     r'_{31}        &  r'_{32}         &  1  \end{array} \right ) \ . 
1359: \label{matrixratio}
1360: \eeq
1361: %
1362: Diagonalizing this matrix, one finds that the presence of the off-diagonal
1363: elements does not change the mass eigenvalues sufficiently from the 
1364: diagonal-matrix case, so that the masses for the first two generations,
1365: $m_d$ and $m_s$, are still too small.  If these off-diagonal elements were
1366: larger than our estimates above, this problem might be ameliorated somewhat. 
1367: 
1368: 
1369: %
1370: \begin{figure}
1371: \begin{center}
1372: \includegraphics[3.5in, 8.7in][4.25in, 9.9in]{egraph.ps}
1373: \end{center}
1374: \caption{\footnotesize{A graph generating the mass matrix for charged leptons,
1375: $\bar e_{j',L} M^{(e)}_{j'k} e^{k}_R$, where $j',k \in \{1,2,3\}$ are
1376: generational indices and $t,t' \in \{4,5\}$ connect with technicolor indices in
1377: SU(2)$_{TC}$, the diagonal subgroup of ${\rm SU}(2)_{ETC} \times {\rm
1378: SU}(2)'_{ETC}$.  Notation is as in Fig. \ref{dgraph}. }}
1379: \label{egraph}
1380: \end{figure}
1381: %
1382: 
1383: 
1384: In this model of type L1, the elements of the charged lepton mass matrix in
1385: eq. (\ref{lme}) are generated via the graph shown in Fig. \ref{egraph}, leading
1386: to the result 
1387: %
1388: \beq
1389: M^{(e)}_{j'k} \simeq 
1390: \frac{\kappa \eta \, 
1391: ({}^{j'}_t \Pi^k_t) \Lambda_{TC}^3}{\Lambda_j'^2 \Lambda_k^2} \ . 
1392: \label{mfe}
1393: \eeq
1394: %
1395: The model succeeds in producing the desired $t-\tau$ mass splitting, but, as
1396: was the case with $M^{(d)}$ and for the same reasons, the values of $m_e$ and
1397: $m_\mu$ are too small.  For comparison, we note that a similar problem with
1398: overly small $m_{d,s}$ and $m_{e,\mu}$ masses was encountered in the model that
1399: we studied earlier in Refs. \cite{ckm,kt} which used relatively conjugate
1400: representations for the left- and right-handed chiral components of the
1401: down-quarks and charged leptons to obtain intragenerational mass splittings.
1402: 
1403: 
1404: In addition to the suppression of the $Q=-1/3$ quark mass relative to the
1405: $Q=2/3$ quark mass in the upper two generations, a viable model should also
1406: incorporate a mechanism for suppressing the charged lepton mass relative to the
1407: $Q=2/3$ and $Q=-1/3$ quark masses in each generation.  In a theory such as the
1408: present one with walking, the technicolor coupling varies slowly with energy
1409: over an extended interval and is only slightly less than the critical value for
1410: condensate formation.  Therefore, small perturbations that would normally be of
1411: negligible importance can become significant.  In particular, the attractive
1412: QCD interaction can naturally expedite techniquark condensation, relative to
1413: technilepton condensation (in a one-family model), so that the former occurs at
1414: a higher energy scale than the latter, giving rise to the inequality $\Sigma_Q
1415: > \Sigma_L$ for the resultant dynamical techniquark and technilepton masses,
1416: and consequently to an increase of the quark masses, relative to the charged
1417: lepton masses, in each generation \cite{qcdcor,met1}.  However, this, by
1418: itself, does not split the masses of the charge 2/3 quarks from those of the
1419: charge $-1/3$ quarks.  For that purpose, one could invoke the U(1)$_Y$
1420: hypercharge gauge interactions, which are weaker.
1421: 
1422: \section{Some Further Phenomenological Topics}
1423: 
1424: \subsection{Flavor-Changing Neutral Current Processes}
1425: 
1426: The present models can satisfy constraints from flavor-changing neutral current
1427: processes.  The basic observation here is the generalization of the point made
1428: in Ref. \cite{ckm} to the case of two ETC groups: because both of the ETC gauge
1429: interactions act in a vectorial, rather than chiral, manner on the
1430: standard-model quarks, processes such as $\bar K^0 \leftrightarrow K^0$, $\bar
1431: D^0 \leftrightarrow D^0$, and $\bar B^0_q \to \bar B^0_q$ with $q=d,s$ are
1432: suppressed.  For example, to take the transition among these that imposes the
1433: most severe constraint, namely the one with neutral kaons, in the $\bar K^0 \to
1434: K^0$ transition, an initial $s_L \bar d_L$ produces a $V^2_1$ ETC gauge boson,
1435: but this cannot directly yield a $d_L \bar s_L$ in the final-state $K^0$; the
1436: latter is produced by a $V^1_2$.  This requires ETC gauge boson mixing of the
1437: form $V^2_1 \to V^1_2$.  In the present models, as in the simpler vectorial
1438: model with a single ETC group studied in Ref. \cite{ckm,kt}, this mixing
1439: introduces a suppression by a factor $\simeq (\Lambda_3/\Lambda_1)^2$,
1440: resulting in an ETC contribution to $\bar K^0 - K^0$ mixing and hence to
1441: $\Delta m_K$ of order $\Lambda_3^2\Lambda_{QCD}^3/\Lambda_1^4$, where $\Delta
1442: m_K = m_{K_L}-m_{K_S} = (0.7 \times 10^{-14})m_{K^0}$, and we ignore factors of
1443: $1/(4\pi^2)$ in view of the strong-coupling nature of the calculation.
1444: Requiring that the ETC contribution be small compared with the experimentally
1445: measured value of $\Delta m_K$ implies that the ratio $\Lambda_3
1446: \Lambda_{QCD}/\Lambda_1^2 \ll 10^{-7}$.  This constraint is satisfied, for
1447: example, by the illustrative values of $\Lambda_1 \simeq 10^3$ TeV and
1448: $\Lambda_3 = 3$ TeV used here, which give $\Lambda_3 \Lambda_{QCD}/\Lambda_1^2
1449: \simeq 0.6 \times 10^{-9}$.  Similarly, in the same $\bar K^0 \to K^0$
1450: transition, an initial $s_R \bar d_R$ produces a $\tilde V^2_1$ ETC gauge
1451: boson, but this cannot directly yield a $d_R \bar s_R$ in the final-state
1452: $K^0$; the latter is produced by a $\tilde V^1_2$.  This requires ETC gauge
1453: boson mixing of the form $\tilde V^2_1 \to \tilde V^1_2$, which causes a
1454: suppression by a factor $\simeq (\Lambda_3'/\Lambda_1')^2$.  By the same
1455: reasoning as above, we require that the ratio $\Lambda_3'
1456: \Lambda_{QCD}/(\Lambda_1')^2 \ll 10^{-7}$.  This constraint is satisfied for
1457: the values of these parameters used here, which give a value of $\simeq 3
1458: \times 10^{-9}$ for this ratio.  Analogous remarks apply for the contributions
1459: via $s_L \bar d_L \to V^2_1 \to \tilde V^1_2 \to d_R \bar s_R$ and $s_R \bar
1460: d_R \to \tilde V^2_1 \to V^1_2 \to d_L \bar s_L$. 
1461: 
1462: 
1463: The same type of suppression occurs for other neutral pseudoscalar meson
1464: mixings such as $D \leftrightarrow \bar D$, $B_d \leftrightarrow \bar B_d$, and
1465: $B_s \leftrightarrow \bar B_s$.  The upper limit on the decay $K^+ \to \pi^+
1466: \mu^+ e^-$ and hence on the elementary process $\bar s \to \bar d \mu^+ e^-$,
1467: which is mediated by $V^1_2$ and $\tilde V^1_2$, is also satisfied with the
1468: values of $\Lambda_1$ and $\Lambda_1'$ that we use.  Similar consistency checks
1469: can be carried out for other processes and quantities due to dimension-5 and
1470: dimension-6 operators, as we have analyzed these in earlier works
1471: \cite{ckm}-\cite{qdml}.
1472: 
1473: 
1474: \subsection{Neutrino Masses} 
1475: 
1476: In Ref. \cite{nt} a mechanism was presented for producing light neutrino masses
1477: in an ETC theory.  This was studied further in Refs. \cite{lrs,ckm}.  Here we
1478: remark on how this mechanism can be implemented in the current models
1479: containing two ETC gauge groups.  For this purpose, we recall that even in an
1480: ETC theory with only one ETC group, below the highest scale, $\Lambda_1$, of
1481: ETC breaking, there are actually two plausible patterns of breaking.  These
1482: were labelled $G_a$ and $G_b$ in Ref. \cite{nt}, and, in modified form,
1483: sequences S1 and S2 in Refs. \cite{ckm,kt}.  In the discussion above we have
1484: concentrated on the two-ETC group generalization of sequence S1.  For
1485: considerations of neutrino masses, sequence S2 is also of interest.  One would
1486: thus be led to consider a generalization of this sequence to the case of two
1487: ETC groups relevant to the present models.
1488: 
1489: 
1490: \subsection{Remarks on a Model with a Minimal Technifermion Sector}
1491: 
1492: Two continuing concerns that one has with ETC models that contain a full
1493: standard-model family of technifermions are the substantial contribution to the
1494: electroweak parameter $S$ and the presence of many pseudo-Nambu-Goldstone
1495: bosons (PNGB's), whose masses must be elevated (e.g., by walking enhancement)
1496: to lie above experimental lower bounds.  These concerns motivate consideration
1497: of ETC models which use the minimum set of electroweak-nonsinglet
1498: technifermions, comprising (for a given TC index) a single left-handed
1499: SU(2)$_L$ doublet and the corresponding two right-handed fields.  We briefly
1500: comment here on how one might construct a model of this type with two ETC
1501: groups. 
1502: 
1503: We denote the technicolor contribution to $S$ as $(\Delta S)^{(TC)}$.  In
1504: general, the $S$ parameter \cite{pt} measures heavy-particle contributions to
1505: the $Z$ self-energy via the term $4 s_W^2 c_W^2
1506: \alpha^{-1}_{em}(m_Z)[\Pi^{(NP)}_{ZZ}(m_Z^2) - \Pi^{(NP)}_{ZZ}(0)]/m_Z^2$,
1507: where $s_W^2 = 1-c_W^2 = \sin^2\theta_W$, evaluated at $m_Z$ (see \cite{pdg}
1508: for details).  Since technifermions are strongly interacting on the scale $m_Z$
1509: used in the definition of $S$, one cannot reliably apply perturbation theory to
1510: calculate $(\Delta S)^{(TC)}$ \cite{scalc1,scalc2}; at most, it provides a
1511: rough guide, namely, $(\Delta S)^{(TC)}_{pert.}  = N_D/(6\pi)$, where $N_D$
1512: denotes the total number of new technifermion SU(2)$_L$ doublets (counting
1513: technicolors).  As is well known, for a one-family technicolor with
1514: technifermions transforming according to the fundamental representation of
1515: SU($N_{TC}$), $N_D=N_{TC}(N_c+1)=8$, so that $(\Delta
1516: S)^{(TC)}_{pert.}=4/(3\pi) \simeq 0.4$, which is larger than the experimentally
1517: preferred region.
1518: 
1519: Now consider a TC/ETC model with a minimal electroweak-nonsinglet technifermion
1520: sector.  For general $N_{TC}$, we again take the technifermions to
1521: transform according to the fundamental representation of SU($N_{TC}$).  The
1522: transformation properties of the SM-nonsinglet technifermions under 
1523: ${\rm SU}(N_{TC}) \times G_{SM}$ are given by
1524: %
1525: \beqs
1526: & & F^{p,t}_L = {F^{1,t}_L \choose F^{2,t}_L} \ : \ 
1527: (N_{TC},1,2)_{0,L}, 
1528: \cr\cr
1529: & & F^{t \ (\pm 1/2)}_R \ : \ \ (N_{TC},1,1)_{\pm 1,R},
1530: \label{1d}
1531: \eeqs
1532: %
1533: where $p=1,2$ and $t$ are the SU(2)$_L$ and SU($N_{TC}$) indices, the
1534: superscripts in parentheses indicate electric charge, and the subscripts are
1535: hypercharge and chirality.  (The charges on $F^{p,t}_L$ are obvious and hence
1536: are suppressed in the notation.)  For this model, $(\Delta
1537: S)^{(TC)}_{pert.}=N_{TC}/(6\pi)$.  Hence, for $N_{TC}=2$, $(\Delta
1538: S)^{(TC)}_{pert.}=1/(3\pi) \simeq 0.1$, which is sufficiently small to agree
1539: with experimental constraints. Although one-doublet technicolor models, as
1540: such, do not have walking, one can add SM-singlet, TC-nonsinglet fermions so as
1541: to produce walking \cite{ts}.  Another advantage of a one-doublet TC model is
1542: that all of the three Nambu-Goldstone bosons (NGB's) that arise due to the
1543: formation of technicondensates are absorbed to make the $W^\pm$ and $Z$ massive
1544: so that there are no problems with unwanted PNGB's.
1545: 
1546: In this model, as the energy scale descends to $\Lambda_{TC}$, the TC
1547: interaction naturally leads to the formation of the technifermion condensates
1548: %
1549: \beq
1550: \langle \bar F_{1,t,L} F^{(1/2),t}_R \rangle \ , \quad
1551: \langle \bar F_{2,t,L} F^{(-1/2),t}_R \rangle
1552: \label{ffbarcond}
1553: \eeq
1554: %
1555: breaking electroweak symmetry in the desired manner.  This channel has $\Delta
1556: C_2=(N_{TC}^2-1)/N_{TC}$, i.e., 3/2 for $N_{TC}=2$.  We note that for
1557: $N_{TC}=2$ an equally attractive channel would involve the condensate
1558: %
1559: \beqs
1560: & & \langle \epsilon_{pq} \epsilon_{st} F^{ps \ T}_L C F^{qt}_L \rangle \cr\cr
1561: & = & \langle \epsilon_{st} ( F^{1 s \ T}_L C F^{2 t}_L - 
1562: F^{2 s \ T}_L C F^{1 t}_L )\rangle
1563: \label{ffcond}
1564: \eeqs
1565: %
1566: where $p,q$ and $s,t$ are SU(2)$_L$ and SU(2)$_{TC}$ indices, respectively.
1567: One would not want (\ref{ffcond}) to be the only technifermion condensate to
1568: occur, since it is invariant under the entire group $G$ in eq. (\ref{g}) and
1569: thus, in particular, does not break the electroweak gauge symmetry.  It would
1570: be worthwhile in future work to investigate how this type of TC model with a
1571: minimal electroweak-nonsinglet technifermion sector could be embedded in a
1572: larger theory with two ETC gauge groups. 
1573: 
1574: 
1575: \section{Conclusions} 
1576: 
1577: In this paper we have formulated a class of extended technicolor models that
1578: can plausibly produce the observed mass splitting $m_t \gg m_b$ without
1579: excessive contributions to the $\rho$ parameter or to neutral flavor-changing
1580: processes.  These models use two different chiral ETC groups \cite{met1} such
1581: that left- and right-handed components of charge 2/3 quarks transform under the
1582: same ETC group, while left- and right-handed components of charge $-1/3$ quarks
1583: and charged leptons transform under different ETC groups.  The models suppress
1584: $m_b$ and $m_\tau$ relative to $m_t$, and $m_s$ and $m_\mu$ relative to $m_c$
1585: because the masses of the $Q=-1/3$ quarks and charged leptons require mixing
1586: between the two ETC groups, while the masses of the $Q=2/3$ quarks do not.  We
1587: have constructed and analyzed in detail one explicit model of this type.
1588: Clearly, since the relative sizes of the quark masses in the first generation
1589: are opposite to the order for the higher two generations, i.e., $m_d > m_u$,
1590: the strategy used in these classes of models would not, by itself, be expected
1591: to account for this.  However, it is also difficult to account for the absolute
1592: sizes of $m_s$ and $m_d$, and of $m_\mu$ and $m_e$; the suppression mechanism
1593: makes these smaller than the respective observed values.  A similar problem was
1594: encountered in a model using relatively conjugate representations of a single
1595: ETC gauge group for left- and right-handed chiral components to obtain
1596: intragenerational mass splittings in Refs. \cite{ckm,kt}.  Although the model
1597: is rather complicated, we believe that it is useful as explicit, moderately
1598: ultraviolet-complete realization of the strategy of using two different ETC
1599: groups to account for the splitting between $m_t$ and $m_b$. 
1600: 
1601: 
1602: \acknowledgments
1603: 
1604: We thank Prof. T. Appelquist for valuable discussions and Y. Bai for a valuable
1605: comment.  This research was partially supported by the grant NSF-PHY-00-98527.
1606: 
1607: \begin{thebibliography}{99}
1608: 
1609: \bibitem{tc}
1610: S. Weinberg, Phys. Rev. D {\bf 19}, 1277 (1979);
1611: L. Susskind, {\it ibid.} D {\bf 20}, 2619 (1979); 
1612: S. Weinberg, Phys. Rev. D {\bf 13}, 974 (1976). 
1613: 
1614: \bibitem{etc}
1615: S. Dimopoulos, L. Susskind, Nucl. Phys. {\bf B155}, 23, (1979);
1616: E. Eichten, K. Lane, Phys.  Lett. B {\bf 90}, 125 (1980).
1617: 
1618: \bibitem{etcrev}
1619: K. Lane, hep-ph/0202255; C. Hill and E. Simmons, Phys. Rep. {\bf 381}, 235 
1620: (2003); R. S. Chivukula, M. Narain, and J. Womersley, in Ref. \cite{pdg}.
1621: 
1622: \bibitem{pdg}
1623: http://pdg.lbl.gov.
1624: 
1625: \bibitem{lepewwg}
1626: http://lepewwg.web.cern.ch/LEPEWWG/plots
1627: 
1628: \bibitem{wtc}
1629: B. Holdom, Phys. Lett. B {\bf 150}, 301 (1985); K. Yamawaki,
1630: M. Bando, and K. Matumoto, Phys. Rev. Lett. {\bf 56}, 1335 (1986);
1631: T. Appelquist, D. Karabali, and L. C. R. Wijewardhana, Phys. Rev. Lett. {\bf
1632: 57}, 957 (1986); T. Appelquist and L.C.R. Wijewardhana, Phys. Rev. D
1633: {\bf 35}, 774 (1987); Phys. Rev. D {\bf 36}, 568 (1987).
1634: 
1635: \bibitem{chiralpt}
1636: T. Appelquist, J. Terning, and L. C. R. Wijewardhana, Phys. Rev. Lett.
1637: {\bf 77}, 1214 (1996); Phys. Rev. T. Appelquist, A. Ratnaweera, J. Terning, 
1638: and L. C. R. Wijewardhana, Phys. Rev. D {\bf 58}, 105017 (1998). 
1639: 
1640: \bibitem{met1}
1641: T. Appelquist, N. Evans, and S. Selipsky, Phys. Lett. B {\bf 374}, 145 (1996).
1642: 
1643: \bibitem{at94}
1644: T. Appelquist, J. Terning, Phys. Rev. D {\bf 50}, 2116 (1994).
1645: 
1646: \bibitem{gw}
1647: S. Glashow and S. Weinberg, Phys. Rev. D {\bf 15}, 1958 (1977). 
1648: 
1649: \bibitem{ssvz}
1650: P. Sikivie, L. Susskind, M. Voloshin and V. Zakharov,
1651: Nucl. Phys. B {\bf 173}, 189 (1980).
1652: 
1653: \bibitem{jt}
1654: J. Terning, Phys. Lett. B {\bf 344}, 279 (1995).
1655: 
1656: \bibitem{ckm}
1657: T. Appelquist, M. Piai and R. Shrock,
1658: Phys.\ Rev.\ D {\bf 69}, 015002 (2004)
1659: 
1660: \bibitem{kt}
1661: T. Appelquist, N. Christensen, M. Piai and R. Shrock,
1662: Phys. Rev. D {\bf 70}, 093010 (2004).
1663: 
1664: \bibitem{dml}
1665: T. Appelquist, M. Piai and R. Shrock,
1666: Phys.\ Lett.\ B {\bf 593}, 175 (2004)
1667: 
1668: \bibitem{qdml}
1669: T. Appelquist, M. Piai and R. Shrock,
1670: Phys.\ Lett.\ B {\bf 595}, 442 (2004)
1671: 
1672: \bibitem{pt}
1673: M. Peskin and T. Takeuchi, Phys. Rev. D {\bf 46}, 381 (1992).
1674: 
1675: \bibitem{scalc1} 
1676: M. Golden and L. Randall, Nucl. Phys. {\bf B} 361, 3 (1991);
1677: R. Johnson, B.-L. Young, and D. McKay, Phys. Rev. D {\bf 43}, R17 (1991);
1678: R. Cahn and M. Suzuki, Phys. Rev. D {\bf 44}, 3641 (1991).
1679: 
1680: \bibitem{scalc2} 
1681: T. Appelquist and G. Triantaphyllou, Phys. Lett. B {\bf 278}, 345
1682: (1992); R. Sundrum and S. Hsu, Nucl. Phys. B {\bf 391}, 127 (1993);
1683: T. Appelquist and J. Terning, Phys. Lett. {\bf B315}, 139 (1993);
1684: T. Appelquist, J. Terning, L.C.R. Wijewardhana, Phys. Rev. Lett. {\bf 79}, 2767
1685: (1997).  T. Appelquist and F. Sannino, Phys. Rev. D {\bf 59}, 067702 (1999);
1686: S. Ignjatovic, L. C. R. Wijewardhana, and T. Takeuchi, {\it ibid.} D {\bf 61},
1687: 056006 (2000); M. Harada, M. Kurachi, and K. Yamawaki, in {\it Proc. Dynamical
1688: Symmetry Breaking Workshop} (Dec. 2004), p. 125; M. Harada, M. Kurachi, and
1689: K. Yamawaki, hep-ph/0509193 (we thank Dr. Kurachi for useful discussions on
1690: this work); N. Christensen and R. Shrock, Phys. Rev. Lett. {\bf 94}, 241801
1691: (2005).
1692: 
1693: \bibitem{sred} 
1694: Some recent works on models to minimize TC contributions to $S$
1695: include Refs. \cite{sann,ts}.  
1696: 
1697: \bibitem{sann}
1698: D. Hong, S. Hsu, and F. Sannino, Phys. Lett. B {\bf 597}, 89 (2004).
1699: 
1700: \bibitem{ts}
1701: N. Christensen and R. Shrock, Phys. Lett. B {\bf 632}, 92 (2006).
1702: 
1703: \bibitem{nt}
1704: T. Appelquist and R. Shrock, Phys. Lett. {\bf B548}, 204 (2002).
1705: 
1706: \bibitem{lrs}
1707: T. Appelquist and R. Shrock, Phys. Rev. Lett. {\bf 90}, 201801 (2003).
1708: 
1709: \bibitem{hc} 
1710: The reason that it is necessary to use two different hypercolor groups is
1711: that if one were to try to use just one, it would not be asymptotically free,
1712: and hence the HC interaction would not increase as the energy scale decreases.
1713: Consequently, this HC group would not fulfill its job of helping to produce
1714: various symmetry-breaking bilinear fermion condensates.
1715: 
1716: \bibitem{efc} An alternative is to avoid a metacolor group and instead have the
1717: communication between the ETC and ETC$'$ groups carried out by one or more
1718: fermions that transform as nonsinglet(s) under both the ETC and ETC$'$ groups;
1719: this will be discussed elsewhere with T. Appelquist, Y. Bai, and M. Piai.
1720: 
1721: \bibitem{beta}
1722: Our notation for the beta function of a gauge group $G_j$ is indicated below,
1723: where $\alpha_j =g_j^2/(4\pi)$, 
1724: $$
1725: \beta_j = \frac{d \alpha_j}{dt} = - \frac{\alpha_j^2}{2\pi}\left ( b_0^{(j)} +
1726: \frac{b_1^{(j)}}{4\pi}\alpha_j + O(\alpha_j^2) \right )
1727: $$
1728: and $t=\ln \mu$ with $\mu$ being the momentum scale. 
1729: 
1730: \bibitem{tumb}
1731: S. Raby, S. Dimopoulos, and L. Susskind, Nucl. Phys. {\bf B} 169, 373 (1980).
1732: 
1733: \bibitem{gap} 
1734: In the approximation of a single-gauge-boson exchange, the
1735: critical coupling for chiral fermions transforming according to the
1736: representations $R_1$ and $R_2$ of a gauge group $G$ to form a condensate
1737: transforming as $R_{cond.}$ in an SU($N$) gauge theory is given by the
1738: condition $\alpha \ge \alpha_{cr}$, where $\alpha_{cr} \simeq 2\pi/(3\Delta
1739: C_2)$, $\Delta C_2 = [C_2(R_1)+C_2(R_2)-C_2(R_c)]$, and $C_2(R)$ is the 
1740: quadratic Casimir invariant \cite{casimir}. 
1741: 
1742: \bibitem{casimir} 
1743: Our notation for the Casimir invariants $C_2(R)$ and $T(R)$
1744: of the representation $R$ is given by ${\cal D}_R(T_a)^i_j{\cal D}_R(T_a)^j_k =
1745: C_2(R)\delta^i_k$ and $({\cal D}_{\cal R}(T_a))^i_j \ ({\cal D}_{\cal
1746: R}(T_b))^j_i= T({\cal R})\delta_{ab}$, where $\{a,b\}$ and $\{i,j,k\}$ denote
1747: group and representation indices and sums over repeated indices are
1748: understood.
1749: 
1750: \bibitem{other1} For completeness, we note that there are other condensation
1751: channels with the same value of $\Delta C_2 = 24/5$, such as
1752: $(5,1;1,1,1;3,2)_{1/3,L} \times (\bar 5,1;1,1,1;\bar 3,1)_{-4/3,L} \to
1753: (1,1;1,1,1;1,2)_{-1}$ with condensate $\langle \bar Q_{a,j,L} u^{a,j}_R
1754: \rangle$ and $(5,1;1,1,1;1,1)_{0,R} \times (\bar 5,1;1,1,2;1,1)_{0,R} \to
1755: (1,1;1,1,2;1,1)_0$ with condensate $\langle {\cal N}^{j \ T}_R C
1756: \chi_{j,\lambda,R} \rangle$.  These are both undesired; the former would break
1757: SU(2)$_L$ at too high a scale, and the latter would break SU(2)$_{MC}$.  Our
1758: scenario, in which the condensation (\ref{xixicondensate_model1}) occurs, and
1759: the above two condensations do not occur, at this scale, is consistent with
1760: GMAC arguments, within the uncertainties concerning the properties of strongly
1761: coupled field theories. 
1762: 
1763: \bibitem{lq} 
1764: A possible mechanism for suppressing a common dynamical technilepton mass
1765: $\Sigma_N \simeq \Sigma_E$ relative to a common techniquark mass $\Sigma_U
1766: \simeq \Sigma_D$ is briefly discussed later in the text.  Since this preserves
1767: custodial symmetry, it does not directly contribute to the TC correction to
1768: $\rho$.
1769: 
1770: \bibitem{veltman}
1771: M. Veltman, Nucl. Phys. B {\bf 123}, 89 (1977); Acta Phys. Pol. B {\bf 8}, 475
1772: (1977).  Note that $f_\rho$ has the properties $f_\rho(x,y) =f_\rho(y,x)$,
1773: $f_\rho(x,x)=0$, and, for the physical case $x,y \ge 0$, $f_\rho(x,y) \ge 0$.
1774: 
1775: \bibitem{qcdcor}
1776: B. Holdom, Phys. Rev. Lett. {\bf 60}, 1233 (1988); 
1777: T. Appelquist and O. Shapira, Phys. Lett. B {\bf 249}, 83 (1990). 
1778: 
1779: \end{thebibliography}
1780: 
1781: \end{document}
1782: 
1783: