1: \documentclass[12pt]{article}
2:
3: \usepackage{amsmath,amssymb,cite,axodraw,rotating,calc,graphics,epic,eepic,FeynMan,epsf}
4:
5: \voffset0cm
6: \hoffset0cm
7: \oddsidemargin0cm
8: \evensidemargin0cm
9: \topmargin0cm
10: \textwidth16.25cm
11: \textheight22.cm
12: \setlength{\parindent}{1cm}
13: \setlength{\parskip}{5pt plus 2pt minus 1pt}
14: \renewcommand{\baselinestretch}{1.2}
15:
16: \def\theequation{\arabic{section}.\arabic{equation}}
17:
18: \renewcommand{\textfraction}{0.1}
19: \renewcommand{\topfraction}{0.9}
20: \renewcommand{\bottomfraction}{0.9}
21:
22: \newcommand {\crossings} {\operatorname{crossings}}
23: \newcommand {\const} {\operatorname{const.}}
24: \newcommand {\tr} {\operatorname{tr}}
25: \newcommand {\sgn} {\operatorname{sgn}}
26: \newcommand {\ord} {\mathcal{O}}
27: \newcommand {\Lag} {\mathcal{L}}
28: \newcommand {\QCD} {\operatorname{QCD}}
29: \newcommand {\BRS} {{\scriptscriptstyle \operatorname{BRS}}}
30: \newcommand {\YM} {\operatorname{YM}}
31: \newcommand {\YMBKT} {\operatorname{5DQCD}}
32: \newcommand {\kin} {\operatorname{kin}}
33: \newcommand {\inv} {\operatorname{inv}}
34: \newcommand {\GF} {\operatorname{GF}}
35: \newcommand {\FP} {\operatorname{FP}}
36: \newcommand {\Res}[1] {\operatorname{Res} \Big|_{#1}}
37: \newcommand {\sss} {\scriptstyle}
38: \newcommand {\Zb} {\mathbb{Z}} %integer numbers
39: \newcommand {\Rb} {\mathbb{R}} %real numbers
40: \newcommand {\I} {\operatorname{I}}
41: \newcommand {\II} {\operatorname{II}}
42: \newcommand {\III} {\operatorname{III}}
43: \newcommand {\id} {\mathbf{1}} %identity
44: \newcommand {\shift} {\hspace{3mm}}
45:
46: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
47:
48: \begin{document}
49:
50: \begin{flushright}
51: WUE-ITP-2004-035\\
52: hep-ph/0411258\\
53: November 2004
54: \end{flushright}
55:
56: \bigskip
57:
58: \begin{center}
59: {\Large {\bf Quantization and High Energy Unitarity of}}\\[0.3cm]
60: {\Large {\bf 5D Orbifold Theories with Brane Kinetic Terms}}\\[1.4cm]
61: {\large Alexander M\"uck$^{\, a, b}$, Lars Nilse$^{\, c}$, Apostolos
62: Pilaftsis$^{\, c}$ and Reinhold R\"uckl$^{\,a}$}\\[0.4cm]
63: $^a${\em Institut f\"ur Theoretische Physik und Astrophysik,
64: Universit\"at W\"urzburg,\\ Am Hubland, 97074 W\"urzburg,
65: Germany}\\[0.2cm]
66: $^b${\em School of Physics, University of Edinburgh,\\
67: King's Buildings, Edinburgh EH9 3JZ, United Kingdom}\\[0.2cm]
68: $^c${\em School of Physics and Astronomy, University of Manchester,\\
69: Manchester M13 9PL, United Kingdom}
70: \end{center}
71: \vskip1.cm
72:
73: \centerline{\bf ABSTRACT}
74: \noindent
75: Five-dimensional field theories compactified on an $S^1/\mathbb{Z}_2$
76: orbifold naturally include local brane kinetic terms at the orbifold
77: fixed points at the tree as well as the quantum level. We study the
78: quantization of these theories before the Kaluza--Klein reduction and
79: derive the relevant Ward and Slavnov--Taylor identities that result
80: from the underlying gauge and Becchi--Rouet--Stora symmetries of the
81: theory. With the help of these identities, we obtain a generalization
82: of the equivalence theorem, where the known high-energy equivalence
83: relation between the longitudinal Kaluza--Klein gauge modes and their
84: respective would-be Goldstone bosons is extended to consistently
85: include the energetically suppressed terms in the high-energy
86: scatterings. Demanding perturbative unitarity, we compute upper
87: limits on the number of the Kaluza--Klein modes. We find that these
88: limits weakly depend on the size of the brane kinetic terms.
89:
90: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
91:
92: \newpage
93: \setcounter{equation}{0}
94: \section{Introduction}
95: \label{introduction}
96:
97: Field theories that realize extra compact dimensions offer new
98: perspectives~\cite{SS,Hosotani} to address problems associated with
99: the breaking of the electroweak gauge symmetry of the Standard Model
100: (SM). One interesting aspect of these theories is that the extra
101: components of the gauge fields may acquire a vacuum expectation value
102: through the so-called Hosotani mechanism~\cite{Hosotani} and so
103: play the role of the Higgs field~\cite{Quiros}. Another novel
104: possibility for electroweak symmetry breaking emerges if the gauge
105: symmetry is broken explicitly by boundary conditions, e.g.~by
106: compactifying the theory on an interval~\cite{Csaki}.
107:
108: Higher dimensional field theories, however, are not renormalizable, in
109: the sense that only a finite number of counterterms would be needed to
110: fix the ultra-violet (UV) infinities to all loop orders. Therefore,
111: higher dimensional field theories should be regarded as effective
112: theories, and as such, they can play a significant role in
113: understanding the low-energy properties of a more fundamental theory
114: that could, for example, be of stringy origin. Their consistent
115: formulation within the context of perturbation theory should still
116: respect basic field-theoretic properties, such as gauge invariance and
117: perturbative unitarity.
118:
119: The present study focuses on five-dimensional (5D) field theories
120: compactified on an $S^1/\mathbb{Z}_2$ orbifold. Orbifolding provides
121: a viable mechanism to introduce chiral fermions after the so-called
122: Kaluza--Klein (KK) reduction of 5D theories to 4 dimensions. However,
123: a consistent description of orbifold field theories requires the
124: inclusion of brane kinetic terms (BKTs) which are localized at the
125: orbifold fixed points~\cite{Georgi,Carena}. These new operators are
126: local and need be included as counterterms at the tree-level to
127: renormalize the UV infinities of local operators of the same form that
128: are generated at the quantum level.
129:
130: In this paper we extend our earlier approach to
131: quantization~\cite{MPR} to 5D orbifold theories that include BKTs. An
132: important aspect of our approach is that the higher dimensional theory
133: is quantized, e.g.~within the framework of generalized $R_\xi$ gauges,
134: {\em before} the KK reduction. Such a gauge-fixing procedure is not
135: trivial, since both the gauge-fixing and ghost sectors should contain
136: terms localized at the orbifold fixed points. After the KK
137: decomposition and integration of the extra dimension, the resulting
138: effective 4D theory coincides with the one that would have been
139: obtained if the KK gauge fields had been quantized mode by
140: mode~\cite{PS} in the conventional $R_\xi$ gauge.
141:
142: One of the advantages of our 5D quantization formalism is that both
143: the usual gauge and the so-called Becchi--Rouet--Stora (BRS)
144: transformations~\cite{BRS} take on a simple and finite form. In close
145: analogy to the ordinary case of 4D Quantum Chromodynamics (QCD), the
146: invariance of the Lagrangian under these transformations gives rise to
147: the corresponding 5D Ward and Slavnov--Taylor (ST)
148: identities~\cite{Ward,ST}. With the aid of these identities, we can
149: derive a generalization of the Equivalence Theorem~(ET)~\cite{CLT,LQT}
150: within the context of 5D orbifold theories~\cite{Dicus} that takes
151: account of the presence of BKTs and the energetically suppressed terms
152: in high-energy scatterings. Such a generalization is obtained by
153: following a line of argument very analogous to that for the ordinary
154: 4D case~\cite{CG} and has been assisted by the introduction of a new
155: formalism of functional differentiation that preserves the orbifold
156: constraints on the 5D fields.
157:
158: As any perturbative framework of field theory, higher dimensional
159: theories should also satisfy perturbative unitarity so as to be able
160: to make credible predictions for physical observables, such as cross
161: sections and decay widths. Requiring the validity of perturbative
162: unitarity for the $s$-wave amplitude, we compute upper limits on the
163: number of the KK modes. The restoration of high-energy unitarity
164: differs in our models from the one taking place in 5D orbifold
165: theories without BKTs~\cite{Dicus,Dominici}. In the presence of BKTs,
166: unitarity is achieved by delicate cancellations among the entire
167: infinite tower of the KK modes.
168:
169: We find, however, that the upper limits established from perturbative
170: unitarity have a weak dependence on the actual size of the BKTs. In
171: particular, large values of BKTs do not seem to screen the effect of
172: the bulk terms. This weak dependence of the unitarity limits on the
173: size of the BKTs may be understood from first principles as follows.
174: High-energy unitarity strongly depends on the UV structure of the
175: theory, since at high energies distances much smaller than the
176: compactification radius of the theory are probed. Therefore, only
177: terms related to the bulk dynamics appear to be relevant when studying
178: perturbative unitarity constraints. Evidently, such considerations
179: will be of immediate phenomenological relevance when constraining more
180: realistic higher-dimensional models into which the well-established
181: Standard Model (SM) can be embedded.
182:
183: The structure of the paper is as follows. In Section~2, after briefly
184: reviewing the ordinary QCD case, we study the quantization of 5D
185: orbifold field theories with BKTs before the KK reduction. In
186: addition, we derive the relevant Ward and ST identities which are a
187: consequence of the underlying gauge and BRS symmetries of the theory.
188: With the help of these identities, we derive in Section~3 a
189: generalization of the ET, which we call the 5D generalized equivalence
190: theorem (GET). The 5D GET provides a complete equivalence relation
191: between the longitudinal KK gauge modes and their respective would-be
192: Goldstone bosons that takes consistently into account the
193: energetically suppressed terms in high-energy scatterings. Section~4
194: is devoted to the computation of upper limits on the number of the KK
195: modes which are obtained by requiring the validity of perturbative
196: unitarity. We also analyze the dependence of these unitarity limits
197: on the actual size of the BKTs. Technical aspects of our study are
198: presented in the appendices. Finally, Section~5 summarizes our
199: conclusions.
200:
201: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
202:
203: \setcounter{equation}{0}
204: \section{Ward and Slavnov--Taylor Identities}\label{ward}
205:
206: We will first review the basic formalism of quantization and the
207: derivation of the Ward and ST identities for the conventional 4D QCD
208: case. This will help us to set up our conventions and obtain insight
209: into analogous considerations concerning quantization and derivation
210: of the corresponding Ward and ST identities within the context of 5D
211: orbifold field theories with~BKTs.
212:
213:
214: \subsection{4D QCD}
215: \label{wardQCD}
216:
217: The ordinary 4D QCD Lagrangian quantized in the $R_\xi$ gauge is given
218: by
219: \begin{equation}
220: \label{LagQCD}
221: \Lag_{\QCD}\ =\ -\,\frac{1}{4} F^a_{\mu \nu} F^{a \, \mu \nu}\: +\:
222: \Lag_{\GF}\: +\: \Lag_{\FP}\,,
223: \end{equation}
224: where $F^a_{\mu \nu} = \partial_{\mu} A^a_{\nu} - \partial_{\nu}
225: A^a_{\mu}+g f^{abc} A^b_{\mu} A^c_{\nu}$ is the field-strength tensor
226: of the gluon field $A^a_{\mu}$. Moreover, $\Lag_{\GF}$ and
227: $\Lag_{\FP}$ denote the parts of the Lagrangian, associated with the
228: gauge-fixing scheme and the Faddeev--Popov ghosts:
229: \begin{eqnarray}
230: \Lag_{\GF} &=& -\, \frac{1}{2 \xi} \big(F[A^a_{\mu}]\big)^2\ =\
231: -\, \frac{1}{2 \xi}\; (\partial^{\mu} A^a_{\mu})^2\;,\nonumber\\
232: \Lag_{\FP} &=& \overline{c}^a\, \frac{\delta F[A^a_{\mu}]}{\delta
233: \theta^b} \, c^b \ =\ \overline{c}^a\, \big(\delta^{ab} \partial^2\:
234: -\: g f^{abc}\, \partial^{\mu} A^c_{\mu}\,\big)\,c^b\; .
235: \end{eqnarray}
236: In the absence of $\Lag_{\GF}$ and $\Lag_{\FP}$, the remaining
237: (classical) part of the QCD Lagrangian $\Lag_{\QCD}$ is invariant
238: under the usual gauge transformations
239: \begin{align}
240: \label{4Dgaugetrans}
241: \delta A^a_{\mu}\ &= \ D^{ab}_{\mu}\, \theta^b\ =\ \big(\, \delta^{ab}
242: \partial_{\mu}\: -\: g f^{abc}\, A^c_{\mu}\,\big)\, \theta^b\; .
243: \end{align}
244: Hence, the gauge-invariant part of the tree-level effective action,
245: $\Gamma[A^a_{\mu}] = -\frac{1}{4} \int d^4x \; F^a_{\mu \nu} F^{a \,
246: \mu \nu}$, satisfies the relation $\Gamma [A^a_{\mu}] = \Gamma
247: [A^a_\mu + \delta A^a_\mu]$, from which the following master Ward
248: identity for QCD is easily derived:
249: \begin{equation}
250: \label{masterWardQCD}
251: \partial_{\mu} \frac{\delta \Gamma}{\delta A^a_{\mu}}\ -\
252: g f^{abc}\; \frac{\delta \Gamma}{\delta A^b_{\mu}} A^c_{\mu}\ =\ 0\; .
253: \end{equation}
254: There is a similarity of the Ward identity~(\ref{masterWardQCD}) with
255: the one derived with the background field method (BFM) in
256: QCD~\cite{Abbott}, which is a consequence of the gauge-invariance of
257: the effective action with respect to gauge transformations of the
258: background fields. Therefore, the results derived below from the
259: classical part of $\Lag_{\QCD}$ will be valid within the BFM framework
260: as well.
261:
262: After functionally differentiating~(\ref{masterWardQCD}) with respect
263: to gluon fields and subsequently setting all the fields to zero, we
264: obtain Ward identities that relate tree-level $n$-point correlation
265: functions to each other. With the convention that all momenta flow
266: into the vertex, these Ward identities may be graphically represented
267: as follows:
268: \begin{equation}
269: \begin{split}
270: -i k_{\mu} \BBB{a \; \mu}{b \; \nu}{c \; \rho}{\shift k}{p}{\shift q}
271: &= \ g f^{abd} \BB{d \; \nu}{c \; \rho}{\shift q}\: +\: g f^{acd} \BB{d \;
272: \rho}{b \; \nu}{\shift p}\\ &\phantom{}
273: \end{split}
274: \end{equation}
275: \begin{equation}
276: \begin{split}
277: -i k_{\mu} \BBBB{a \; \mu}{b \; \nu}{c \; \rho}{d \; \sigma}{k}{\shift
278: p}{q}{\shift r} &=\ g f^{abe} \BBB{e \; \nu}{c \; \rho}{d \;
279: \sigma}{k+p}{q}{\shift r}\: +\ g f^{ace} \BBB{e \;
280: \rho}{b \; \nu}{d \; \sigma}{k+q}{p}{\shift r}\\ &\phantom{}\\
281: &+\ g f^{ade} \BBB{e \; \sigma}{b \; \nu}{c \; \rho}{k+r}{p}{\shift q}\\
282: &\phantom{}\\ &\phantom{}
283: \end{split}
284: \end{equation}
285:
286: We now turn our attention to the complete QCD Lagrangian
287: (\ref{LagQCD}) quantized in the covariant $R_\xi$~gauges. Although the
288: gauge-fixing and ghost terms break the gauge invariance of
289: $\Lag_{\QCD}$, the complete QCD Lagrangian $\Lag_{\QCD}$ is still
290: invariant under the so-called BRS transformations~\cite{BRS}:
291: \begin{eqnarray}
292: \label{BRSQCD}
293: s A^a_{\mu} &=& \ D^{ab}_{\mu} c^b\;,\nonumber\\
294: s\, c^a &=& -\; \frac{g f^{abc}}{2} c^b c^c\;,\\
295: s\, \overline{c}^a &=& \frac{F[A^a_{\mu}]}{\xi}\ =\
296: \frac{\partial^{\mu} A^a_{\mu}}{\xi}\ .\nonumber
297: \end{eqnarray}
298: Note that the above form of BRS transformations is nilpotent after the
299: anti-ghost equation of motion is imposed. The nilpotency of the BRS
300: transformations is an important property that ensures unitarity of the
301: $S$-matrix operator in the subspace of the physical
302: fields~\cite{Pokorski,SMreviews}.
303:
304: Following the standard path-integral quantization approach, we define
305: the generating functional $Z$ of the connected Green functions through
306: the relation
307: \begin{equation}
308: e^{iZ}\, = \, \int\! DA \, Dc \, D\bar{c} \, \exp \Big[\,
309: i\int\! d^4x\, \big(\, \Lag_{\QCD} + J^{a \mu} A^a_{\mu} +
310: \overline{D}^a c^a + \overline{c}^a D^a + K^{a \mu} s A^a_{\mu} + M^a
311: sc^a\, \big)\,\Big]\ .
312: \end{equation}
313: In the above, $J^{a \mu}$, $\overline{D}^a$ and $D^a$ are the sources
314: for the gluons, ghosts and anti-ghosts, respectively. As usual, we
315: have included the additional sources $K^{a \mu}$ and $M^a$ for the BRS
316: variations $s A^a_{\mu}$ and $s c^a$ that are non-linear in the
317: fields.
318:
319: Given the invariance of the path-integral measure under BRS
320: transformations, it is not difficult to show that
321: \begin{equation}
322: \label{Zgen}
323: Z \big[\, J^{a \mu},\ \overline{D}^a,\ D^a,\ K^{a \mu},\ M^a\,\big]
324: \ =\
325: Z \big[\,J^{a \mu} - \mbox{$\frac{\omega}{\xi}$}\,\partial^\mu D^a,\
326: \overline{D}^a,\ D^a,\ K^{a \mu} + \omega J^{a \mu},\
327: M^a - \omega \overline{D}^a\,\big]\, ,
328: \end{equation}
329: where $\omega$ is a Grassmann parameter, i.e.~$\omega^2 = 0$. An
330: immediate consequence of~(\ref{Zgen}) is the master ST identity for
331: QCD
332: \begin{equation}
333: \label{4DST}
334: J^{a \mu} \frac{\delta Z}{\delta K^{a \mu}}\ -\ \overline{D}^a
335: \frac{\delta Z}{\delta M^a}\ +\ \frac{1}{\xi}\, D^a\, \partial^{\mu}
336: \frac{\delta Z}{\delta J^{a \mu}}\ =\ 0\, .
337: \end{equation}
338: As before, a number of ST identities can be derived from~(\ref{4DST}),
339: using functional differentiation techniques.
340:
341: An alternative and perhaps more practical approach to deriving ST
342: identities is to use the BRS invariance of the Green function defined
343: by means of the time-ordered operator formalism. To give an example,
344: let us consider the BRS invariance of the Green function associated
345: with the fields $\bar{c}^a(x) A^b_{\nu}(y) A^c_{\rho}(z)$, i.e.
346: \begin{equation}
347: \label{ersteZeile}
348: s\, \langle 0 | T \overline{c}^a(x) A^b_{\nu}(y) A^c_{\rho}(z) | 0
349: \rangle\ = \ 0\; .
350: \end{equation}
351: Employing the BRS transformations given by~(\ref{BRSQCD}), it is
352: straightforward to obtain the expression
353: \begin{eqnarray}
354: \label{zweiteZeile}
355: \frac{1}{\xi}\, \partial^{\mu}_{\sss x}\,
356: \langle 0 | T A^a_{\mu} (x) A^b_{\nu} (y) A^c_{\rho} (z) | 0 \rangle\:
357: -\: \partial^{\sss y}_{\nu} \langle 0 | T \overline{c}^a(x) c^b(y)
358: A^c_{\rho}(z)| 0 \rangle\
359: -\: \partial^{\sss z}_{\rho} \langle 0 | T
360: \overline{c}^a(x) A^b_{\nu} (y) c^c(z) | 0 \rangle\: &&\nonumber\\
361: +\: g f^{bde} \langle 0 | T \overline{c}^a(x) c^d(y) A^e_{\nu} (y)
362: A^c_{\rho}(z) | 0 \rangle\:
363: +\: g f^{cde} \langle 0 | T \overline{c}^a(x)
364: A^b_{\nu}(y) c^d(z) A^e_{\rho}(z) | 0 \rangle\ =\ 0\, .&&\nonumber\\
365: \end{eqnarray}
366: The last two terms on the LHS of~(\ref{zweiteZeile}) depend on
367: bilinear fields evaluated at the same space-time point $y$ or
368: $z$. These terms do not have one-particle poles for the external lines
369: attached to those points, and therefore cancel on-shell. The resulting
370: on-shell ST identity for the involved one-particle irreducible Green
371: functions may then be represented graphically as follows:
372: \begin{equation}
373: p^{\mu} \; \BBBblob{a \; \mu}{b \; \nu}{c \; \rho}{p}{q}{\shift r} -\
374: q_{\nu} \GGBblob{a}{b}{c \; \rho}{p}{q}{\shift r} -\ r_{\rho}
375: \GBGblob{a}{b \; \nu}{c}{p}{q}{\shift r} \ =\ 0 \; ,
376: \end{equation}
377:
378: \vspace{1.cm}
379: \noindent
380: with $q^2 = r^2 = 0$. It is not difficult to check the validity of
381: the above on-shell ST identity for the tree-level QCD
382: Lagrangian~(\ref{LagQCD}).
383:
384: For our purpose of studying high-energy unitarity in $2\to 2$
385: scatterings, more useful are ST identities that apply to on-shell
386: 4-point functions, e.g.~to the product of fields: $A^a_{\nu} (x)
387: A^b_{\nu} (y) A^c_{\rho} (z) A^d_{\sigma} (w)$. Again, starting from
388: the identity
389: \begin{equation}
390: \label{dritteZeile}
391: s\; \langle 0 | T \; \bar{c}^a(x) [\partial^{\nu} A^b_{\nu} (y)]
392: [\partial^{\rho} A^c_{\rho} (z)] [\partial^{\sigma} A^d_{\sigma} (w)]
393: | 0 \rangle\ =\ 0\, ,
394: \end{equation}
395: we arrive at the ST identity which may be given diagrammatically by
396: \begin{equation}
397: \label{vierteZeile}
398: k_{\mu} \, p_{\nu} \, q_{\rho} \, r_{\sigma} \;
399: \BBBBblob{\mu \; a}{\nu \; b}{\rho \; c}{\sigma \; d}{k}{\shift
400: p}{q}{\shift r}\ =\ 0\ .
401: \end{equation}
402:
403: \vspace{1.cm}
404: \noindent
405: As we will see in the next section and Section~3, it is the 5D analog
406: of the above on-shell ST identity which lies at the heart of the proof
407: of the GET for the 5D orbifold field theories.
408:
409:
410: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
411:
412: \subsection{5D Orbifold Theories with Brane Kinetic Terms}\label{wardCYMBKT}
413:
414: We will now study a 5D Yang--Mills theory, such as 5D QCD,
415: compactified on an $S^1/\mathbb{Z}_2$ orbifold. As has been observed
416: in~\cite{Georgi,Carena}, orbifold compactification introduces BKTs for
417: the 5D gluon field at the orbifold fixed points $y=0$ and $y = \pi R$.
418: These BKTs naturally emerge as counterterms at the tree level, so as
419: to cancel the UV infinities of the same form that are generated by
420: radiative corrections~\cite{Quiros}. Here and in the following, we
421: denote Lorentz indices enumerating the 5 dimensions with capital Roman
422: letters $M,N$ etc., with $M = \mu, 5$ ($\mu = 0,1,2,3$), and the fifth
423: compact dimension with $y\equiv x^5 \in (-\pi R,\pi R]$.
424:
425: To avoid excessive complication in our analytic results, we consider
426: 5D QCD compactified on an $S^1/\mathbb{Z}_2$ orbifold with one BKT at
427: $y=0$. However, our discussion can very analogously carry over to the
428: general case with two BKTs at $y=0$ and $y = \pi R$. To be specific,
429: the 5D QCD Lagrangian of interest reads
430: \begin{equation}
431: \label{LagYMBKT}
432: \Lag_{\YMBKT} (x,y)\ =\
433: -\, \frac{1}{4}\, \Big( 1\: +\: r_c\,\delta(y) \Big)\,
434: F^a_{MN} F^{a \; MN}\ +\ \Lag_{\rm 5D\GF}\ +\ \Lag_{\rm 5D\FP}\,,
435: \end{equation}
436: where $F^a_{MN} = \partial_M A^a_N - \partial_N A^a_M + g_5 f^{abc}
437: A^b_M A^c_N$ is the corresponding field-strength tensor for the 5D QCD
438: gluon $A^a_M$, and $r_c$ is taken to be a positive dimensionful
439: coupling for the BKT at $y=0$.\footnote{If $r_c$ is negative and
440: $|r_c| \leq 2\pi R$, the theory will contain undesirable negative norm
441: states~\cite{Santiago}.} In addition, the terms $\Lag_{\rm 5D\GF}$
442: and $\Lag_{\rm 5D\FP}$ describe the 5D gauge-fixing and Faddeev--Popov
443: ghost sectors to be discussed in detail below.
444:
445: In order that the 5D orbifold theory includes ordinary QCD with a
446: massless gluon, it is sufficient for the 5D gluon field $A^a_M$ to
447: obey the following constraints:
448: \begin{equation}
449: \label{cond}
450: A^a_M(x,y)\ =\ A^a_M(x, y+2 \pi R)\,,\quad
451: A^a_{\mu}(x,y)\ =\ A^a_{\mu}(x,-y)\,,\quad
452: A^a_5(x,y)\ =\ - A^a_5(x,-y)\; .
453: \end{equation}
454: Observe that $A^a_\mu$ ($A^a_5$) is even (odd) under the
455: $\mathbb{Z}_2$ reflection, $y \to - y$. Although the BKT at $y=0$
456: breaks explicitly the higher-dimensional Lorentz invariance of the
457: Lagrangian~(\ref{LagYMBKT}) down to 4 dimensions, the classical part
458: of the 5D QCD Lagrangian remains invariant under the 5D gauge
459: transformations
460: \begin{equation}
461: \label{5DgaugeTransformation}
462: \delta A^a_M\ =\ D^{ab}_M\, \theta^b \ = \
463: \big(\, \delta^{ab} \partial_M\: -\: g_5 f^{abc} A^c_M\,\big)\, \theta^b\, .
464: \end{equation}
465: For consistency, the 5D gauge parameter $\theta^a (x,y)$ has to be
466: periodic and even under $S^1/\mathbb{Z}_2$: $\theta^a (x,y) = \theta^a
467: (x, y+2 \pi R)$ and $\theta^a (x,y) = \theta^a (x,-y)$.
468:
469: Given the $S^1/\mathbb{Z}_2$ parities of the 5D gluon field $A^a_M
470: (x,y)$ and the gauge parameter $\theta^a (x,y)$, these quantities can
471: be expanded in infinite series in terms of orthonormal functions $f_n
472: (y)$ and $g_n (y)$ that are even and odd in $y$, respectively. More
473: explicitly, we have
474: \begin{eqnarray}
475: \label{AmuExpansion}
476: A^a_\mu (x,y) &=& \sum_{n=0}^{\infty} A^a_{(n) \mu}(x)\, f_n(y)\;,\nonumber\\
477: A^a_5 (x,y) &=& \sum_{n=1}^{\infty} A^a_{(n) 5}(x)\, g_n(y)\;,\\
478: \theta^a(x,y) &=& \sum_{n=0}^{\infty} \theta^a_{(n)}(x)\, f_n(y)\; .\nonumber
479: \end{eqnarray}
480: An important property of the orthonormal functions $f_n (y)$ and $g_n
481: (y)$ is that the resulting coefficients $A^a_{(n) \mu}(x)$ and
482: $A^a_{(n) 5} (x)$, the so-called KK modes, are mass eigenstates of the
483: effective theory, after integrating out the extra dimension. For
484: example, in the limit of a vanishing BKT term, $r_c \to 0$, one
485: recovers the standard Fourier series expansion, where the functions
486: $f_n (y)$ and $g_n (y)$ are proportional to $\cos (ny/R)$ and $\sin
487: (ny/R )$, respectively.
488:
489: In Appendix~\ref{masseigenmodeexpansion} we derive the analytic forms
490: of the orthonormal functions $f_n (y)$ and $g_n (y)$ for the complete
491: 5D orbifold theory with BKTs. To properly deal with the singular
492: behaviour of the BKTs, e.g.~at $y=0$, we introduce a regularization
493: method to analytically define $f_n (y)$ and $g_n (y)$ in the interval
494: $(-\epsilon,\epsilon)$, where $\epsilon$ is an infinitesimal positive
495: constant which is taken to zero at the end of the calculation. In
496: their definition interval $(-\pi R +\epsilon, \pi R +\epsilon]$, the
497: $\epsilon$-regularized orthonormal functions are given by
498: \begin{eqnarray}
499: \label{fn}
500: f_n(y) &=& \frac{N_n}{\sqrt{2^{\delta_{n,0}} \pi R}}
501: \times \begin{cases}
502: \ \cos m_n y\: -\: \frac{1}{2}\,m_n\,r_c\,\sin m_n y\,,\
503: &\textrm{for}\ -\pi R + \epsilon <
504: y \leq -\epsilon\\
505: \ \cos m_n y\,,\ &\textrm{for} \
506: -\epsilon < y < \epsilon\\
507: \ \cos m_n y\: +\: \frac{1}{2}\,m_n\,r_c\,\sin m_n y\,,
508: &\textrm{for}\quad \epsilon \leq y
509: \leq \pi R + \epsilon
510: \end{cases}\\[3mm]
511: \label{gn}
512: g_n(y) &=& \frac{N_n}{\sqrt{2^{\delta_{n,0}} \pi R}}
513: \times\begin{cases}
514: \ \sin m_n y\: +\: \frac{1}{2}\,m_n\,r_c\,\cos m_n y\,,\ &
515: \textrm{for}\ -\pi R+\epsilon <
516: y\leq -\epsilon \\
517: \ \sin m_n y\,,\ &\textrm{for}\
518: -\epsilon < y < \epsilon\\
519: \ \sin m_n y\: -\: \frac{1}{2}\,m_n\,r_c\,\cos m_n y\,,\ &
520: \textrm{for}\quad \epsilon \leq y
521: \leq \pi R +\epsilon\; ,
522: \end{cases}
523: \end{eqnarray}
524: where
525: \begin{equation}
526: \label{Nn}
527: N_n^{-2}\ =\ 1 + \tilde{r}_c + \pi^2 R^2 \tilde{r}_c^2 m_n^2\;,
528: \end{equation}
529: with $\tilde{r}_c = r_c/(2 \pi R) \geq 0$. Notice that the
530: orthonormal functions $f_n (y)$ and $g_n (y)$ are gauge-independent
531: and related to each other through
532: \begin{equation}
533: \label{partial}
534: \partial_5 f_n \ =\ -\, m_n g_n\,,\qquad \partial_5 g_n \ =\ m_n f_n\; ,
535: \end{equation}
536: where the derivatives are understood to be piecewise defined within
537: the respective sub-intervals given in~(\ref{fn}) and (\ref{gn}).
538: Finally, the physical masses $m_n$ of the KK gauge modes
539: $A^a_{(n)\,\mu}$ are obtained by solving numerically the
540: transcendental equation~\cite{Carena}:
541: \begin{equation}
542: \label{spectrum}
543: \frac{m_n r_c}{2}\ =\ -\, \tan \big(m_n \pi R\big)\; .
544: \end{equation}
545: In addition to the massless solution $m_0 = 0$, in the limits $r_c \to
546: 0$ and $r_c \to \infty$, (\ref{spectrum}) has the simple analytic
547: solutions $m_{n\ge 1} = n/R$ and $m_{n\ge 1} = (2n-1)/2R$, respectively.
548:
549: Employing Fourier-like convolution techniques developed in Appendix
550: C.1, we can calculate the Feynman rules of the effective 4D KK
551: theory. These are listed in Appendix~A. For example, applying the
552: convolution techniques to the 5D gauge
553: transformations~(\ref{5DgaugeTransformation}), we find the effective
554: gauge transformations for the KK modes
555: \begin{eqnarray}
556: \label{transBKT}
557: \delta A^a_{(n) \mu} &=& \partial_{\mu} \theta^a_{(n)}\: -\:
558: g f^{abc} \sum_{m, l=0}^{\infty} \sqrt{2}^{\;
559: -1-\delta_{n,0}-\delta_{m,0}-\delta_{l,0}} \; \theta^b_{(m)}\, A^c_{(l)
560: \mu}\, \Delta_{n,l,m}\; ,\nonumber\\
561: \delta A^a_{(n) 5} &=& -\, m_n \theta^a_{(n)}\: -\:
562: g f^{abc} \sum_{m=0, \; l=1}^{\infty}
563: \sqrt{2}^{\; -1-\delta_{m,0}} \; \theta^b_{(m)}\, A^c_{(l) 5}\,
564: \tilde{\Delta}_{n,l,m}\; ,
565: \end{eqnarray}
566: where $g = g_5/\sqrt{2\pi R}$ is a dimensionless coupling constant.
567: In the limit $r_c \to 0$, these transformations reduce to the ones
568: stated in~\cite{MPR} without BKTs, with the identifications:
569: $\Delta_{n,l,m} = \delta_{n,l,m}$ and $\tilde{\Delta}_{n,l,m} =
570: \tilde{\delta}_{n,l,m}$. The definitions of $\Delta_{n,l,m}$ and
571: $\tilde{\Delta}_{n,l,m}$ are given in Appendix C.2.
572:
573:
574: A technical drawback of~(\ref{transBKT}) is that the number of the
575: effective gauge-field transformations becomes infinite. Our task of
576: deriving Ward and ST identities will be greatly facilitated if we
577: introduce a formalism of functional differentiation on an $S^1 /
578: \Zb_2$ orbifold. To this end, we need first to define the
579: $\delta$-function relevant to our $S^1 / \Zb_2$ theory with BKTs. An
580: appropriate construction can be achieved by exploiting the
581: completeness relation
582: \begin{equation}
583: \label{deltaExp}
584: \delta (y-y'; r_c)\ =\ \sum_{n=0}^{\infty}\,\big[\, f_n (y)\, f_n (y')\: +\:
585: g_n (y)\, g_n(y')\,\big]\, .
586: \end{equation}
587: In the limit $r_c \to 0$, the $\delta$-function goes over to the known
588: form:
589: \begin{equation}
590: \delta (y;r_c=0)\ \equiv\ \delta (y) \ =\ \sum_{n=0}^{\infty} \,
591: \frac{1}{2^{\delta_{n,0}}\,\pi R}\ \cos \bigg( \frac{ny}{R} \bigg)\; .
592: \end{equation}
593: For any periodic test function $h(y)$ defined on $S^1$, it can be
594: shown that the defining property of the $\delta$-function,
595: \begin{equation}
596: \label{deltaDef}
597: \int_{-\pi R}^{\pi R} dy\, [ 1\, +\, r_c \delta (y) ]\;
598: h(y)\, \delta (y;r_c)\ =\ h(0)\, ,
599: \end{equation}
600: holds true. In fact, the proof of~(\ref{deltaDef}) is based on the
601: orthonormality of the mass eigenmode wavefunctions $f_n (y)$ and $g_n
602: (y)$.
603:
604: We are now in a position to introduce a functional differentiation
605: formalism that respects the $S^1/\Zb_2$ properties of the fields
606: $A^a_\mu (x,y)$ and $A^a_5 (x,y)$. To be specific, functional
607: differentiation on $S^1/\Zb_2$ may be defined as follows:
608: \begin{eqnarray}
609: \label{fdDef}
610: \frac{\delta A^a_{\mu}(x_1,y_1)}{\delta A^b_{\nu}(x_2,y_2)} &=&
611: \frac{1}{2}\,g_\mu^\nu\,\delta^{ab}\, \Big[\, \delta(y_1-y_2; r_c)\ +\
612: \delta(y_1+y_2; r_c)\,\Big]\, \delta^{(4)}(x_1-x_2)\; ,\nonumber\\
613: \frac{\delta A^a_5(x_1,y_1)}{\delta A^b_5(x_2,y_2)} &=&
614: \frac{1}{2}\, \delta^{ab}\, \Big[\,\delta(y_1-y_2; r_c)\ -\ \delta(y_1+y_2;
615: r_c)\,\Big]\, \delta^{(4)}(x_1-x_2)\;.
616: \end{eqnarray}
617: Notice that functional differentiation with respect to $A^b_\nu
618: (x_2,y_2)$ ($A^b_5 (x_2,y_2)$) is even (odd) under $y_2 \to -y_2$, as
619: it should be.
620:
621: Like in the ordinary QCD case, we may now derive Ward identities that
622: result from the 5D gauge invariance of the classical part of the
623: tree-level effective action: $\Gamma [ A^a_M ] = -\frac{1}{4}\,\int
624: d^4x \int_{-\pi R}^{\pi R} dy \, [1 + r_c \delta (y)]\, F^a_{MN} F^{a
625: \; MN}$. Making use of the newly introduced functional differentiation
626: formalism, we obtain the master Ward identity for 5D QCD
627: \begin{equation}
628: \label{masterWard5DYM}
629: \partial_M \, \frac{\delta \Gamma}{\delta A^a_M}\ -\ g_5 f^{abc}
630: \frac{\delta \Gamma}{\delta A^b_M} A^c_M\ =\ 0\; .
631: \end{equation}
632: In order to translate the 5D master Ward identity into the effective
633: 4D one, it proves useful to define the decompositions of the
634: functional derivatives $\delta / \delta A^a_{\mu} (x,y)$ and $\delta /
635: \delta A^a_5 (x,y)$ in terms of the orthonormal functions $f_n (y)$
636: and $g_n (y)$:
637: \begin{eqnarray}
638: \label{dd}
639: \frac{\delta}{\delta A^a_{\mu}(x,y)} &=& \sum_{n=0}^{\infty} f_n(y) \;
640: \frac{\delta}{\delta A^a_{(n) \mu}(x)}\,,\nonumber\\
641: \frac{\delta}{\delta A^a_5 (x,y)} &=&
642: \sum_{n=1}^{\infty} g_n(y) \; \frac{\delta}{\delta A^a_{(n) 5}(x)}\ .
643: \end{eqnarray}
644: Observe that the definitions in~(\ref{dd}) are consistent with the
645: functional differentiation rules stated in~(\ref{fdDef}). Substituting
646: (\ref{dd}) into (\ref{masterWard5DYM}) and integrating over $y$ yields
647: the effective master Ward identity for the compactified theory
648: \begin{equation}
649: \label{masterWardCYM}
650: \begin{split}
651: \partial_{\mu} \frac{\delta \Gamma}{\delta A^a_{(n) \mu}}\ +\ m_n
652: \frac{\delta \Gamma}{\delta A^a_{(n) 5}}\ =\ \; &g f^{abc} \,
653: \sum_{m,l=0}^{\infty} \sqrt{2}^{\; -1-\delta_{n,0} - \delta_{m,0} -
654: \delta_{l,0}}\\ &\times\, \bigg(\, \frac{\delta \Gamma}{\delta A^b_{(m)
655: \mu}} A^c_{(l) \mu} \Delta_{m, n, l}\ +\ \frac{\delta \Gamma}{\delta
656: A^b_{(m) 5}} A^c_{(l) 5} \tilde{\Delta}_{m,n,l}\, \bigg)\; .
657: \end{split}
658: \end{equation}
659:
660: By functional differentiation with respect to KK fields, we may derive
661: Ward identities among different $n$-point correlation functions, which
662: may graphically be represented as follows:
663: \begin{equation}
664: \label{wardCYM1}
665: \begin{split}
666: -i k_{\mu} \BBB{a \; \mu}{b \; \nu}{c \; \rho}{(n) \; k}{(m) \; p}{(l)
667: \; q} = &\ \ -m_n \SBB{a}{b \; \nu}{c \; \rho}{(n) \; k}{(m) \; p}{(l)
668: \; q} \\ &\phantom{} \\ &\phantom{} \\ +\ \sqrt{2}^{\; -1-\delta_{n,0}
669: - \delta_{m,0} - \delta_{l,0}} \, &\Delta_{n, m, l} \: g \big[ f^{abd}
670: \BB{d \; \nu}{c \; \rho}{(l) \; q} +\ f^{acd} \BB{d \; \rho}{b \;
671: \nu}{(m) \; p} \big]
672: \end{split}
673: \end{equation}
674:
675: \bigskip
676:
677: \begin{equation}
678: \begin{split}
679: -i k^{\mu} &\BBS{a \; \mu}{b \; \nu}{c}{(n) \; k}{(m) \; p}{(l) \; q} =\
680: -m_n \SBS{a}{b \; \nu}{c}{(n) \; k}{(m) \; p}{(l) \; q}\\ &\phantom{}
681: \\ &\phantom{}\\ &+\ \sqrt{2}^{\; -1-\delta_{n,0} - \delta_{m,0}} \, g
682: \, \big[ \Delta_{m,n,l} f^{abd} \BS{d \; \nu}{c}{(l) \; q} -\
683: \tilde{\Delta}_{m,n,l} f^{acd} \BS{b \; \nu}{d}{(m) \; p} \big]
684: \end{split}
685: \end{equation}
686:
687: \bigskip
688:
689: \begin{equation}
690: \begin{split}
691: -i k_{\mu} \BSS{a \; \mu}{b}{c}{(n) \; k}{(m) \; p}{(l) \; q} = &\
692: \sqrt{2}^{\; -1-\delta_{n,0}} \, \tilde{\Delta}_{m,n,l} \: g \, \big[
693: f^{abd} \SSLL{ d }{ c }{(l) \; q} +\ f^{acd} \SSLL{d}{b}{(m) \; p} \big]\\
694: &\phantom{}
695: \end{split}
696: \end{equation}
697:
698: \bigskip
699:
700: \begin{equation}
701: \begin{split}
702: -i k^{\mu} \BBBB{a \; \mu}{b \; \nu}{c \; \rho}{d \; \sigma}{(n) \;
703: k}{(m) \; p}{(l) \; q}{(k) \; r} \hspace{-3mm} = &\ \sum_{j=0}^{\infty}
704: g \, \sqrt{2}^{\; -1-\delta_{n,0} - \delta_{j,0}} \big[ \sqrt{2}^{\; -
705: \delta_{m,0}} \, \Delta_{m,n,j} f^{abe} \BBB{e \; \nu}{c \; \rho}{d \;
706: \sigma}{(j) \; k+p}{(l) \; q}{(k) \; r}\\ &\phantom{} \\ +
707: \sqrt{2}^{\; - \delta_{l,0}} \, \Delta_{l,n,j} &f^{ace} \BBB{e \;
708: \rho}{b \; \nu}{d \; \sigma}{(j) \; k+q}{(m) \; p}{(k) \; r}
709: \hspace{-5mm} +\ \sqrt{2}^{\; - \delta_{k,0}} \, \Delta_{k,n,j} f^{ade}
710: \BBB{e \; \sigma}{b \; \nu}{c \; \rho}{(j) \; k+r}{(m) \; p}{(l) \; q}
711: \hspace{-5mm} \big]\\ &\phantom{} \\ &\phantom{}
712: \end{split}
713: \end{equation}
714:
715: \bigskip
716:
717: \begin{equation}
718: \label{wardCYM5}
719: \begin{split}
720: -i k^{\mu} \BBSS{a \; \mu}{b \; \nu}{c}{d}{(n) \; k}{(m) \; p}{(l) \;
721: q}{(k) \; r} \hspace{-3mm} = &\ \sum_{j=0}^{\infty} g \, \sqrt{2}^{\;
722: -1-\delta_{n,0} - \delta_{j,0}} \big[ \sqrt{2}^{\; - \delta_{m,0}} \,
723: \Delta_{m,n,j} f^{abe} \BSS{e \; \nu}{c}{d}{(j) \; k+p}{(l) \; q}{(k)
724: \; r}\\ &\phantom{} \\ +\ \tilde{\Delta}_{l,n,j} &f^{ace} \SBS{e}{b \;
725: \nu}{d}{(j) \; k+q}{(m) \; p}{(k) \; r} \hspace{-5mm} +\
726: \tilde{\Delta}_{k,n,j} f^{ade} \SBS{e}{b \; \nu}{c}{(j) \; k+r}{(m) \;
727: p}{(l) \; q} \hspace{-5mm} \big]\\ &\phantom{}\\ &\phantom{}
728: \end{split}
729: \end{equation}
730: At this point, we should remark that $A^a_{(n) 5}$ satisfies Ward
731: identities very analogous to those found for the would-be Goldstone
732: bosons in spontaneously broken gauge theories~\cite{Papa}. Moreover,
733: we should stress again that the above 5 Ward
734: identities~(\ref{wardCYM1})--(\ref{wardCYM5}) will also hold in the
735: case of a BFM formulation of 5D Yang--Mills theories~\cite{Buras}.
736:
737: We will now discuss the part of the 5D Lagrangian in~(\ref{LagYMBKT}),
738: associated with the gauge-fixing scheme and the ghost sector. Within
739: the framework of the generalized $R_\xi$ gauges, all mixing terms
740: between the KK gauge fields $A^a_{(n)\,\mu}$ and their would-be
741: Goldstone bosons $A^a_{(n)\, 5}$ are absent. To incorporate this
742: property, we proceed very analogously to~\cite{MPR} and choose the 5D
743: gauge-fixing functional
744: \begin{equation}
745: \label{gfFunctional}
746: F[A^a_M]\ =\ \partial^{\mu} A^a_{\mu}\: -\: \xi\, \partial_5 A^a_5\; .
747: \end{equation}
748: However, we multiply the 5D gauge-fixing and the Faddeev--Popov ghost
749: terms with the expression $[ 1 + r_c \delta (y)]$ which includes
750: localized interactions of the same form as those present in the
751: gauge-kinetic part of the Lagrangian, i.e.
752: \begin{eqnarray}
753: \label{gfTerm}
754: \Lag_{\rm 5D\GF}(x,y) &=& -\, \big[ 1 + r_c \delta
755: (y) \big]\, \frac{1}{2 \xi}\;
756: \big(F[A^a_M]\big)^2\ =\ -\, \big[ 1 + r_c \delta
757: (y) \big]\, \frac{1}{2 \xi}\, \Big(\,
758: \partial^{\mu} A^a_{\mu}\: -\: \xi\, \partial_5
759: A^a_5\,\Big)^2\;,\nonumber\\[3mm]
760: \Lag_{\rm 5D\FP}(x,y) &=& \big[ 1 + r_c \delta
761: (y) \big]\, \bar{c}^a\; \frac{\delta F[A^a_M]}{\delta
762: \theta^b}\; c^b\\
763: &=& \big[ 1 + r_c \delta
764: (y) \big]\,\bar{c}^a\, \Big[\, \delta^{ab}\, \Big( \partial^2\, -\,
765: \xi\, \partial_5^2\Big)\: -\: g_5 f^{abc}\,
766: \Big(\,\partial^{\mu} A^c_{\mu}\, -\, \xi\, \partial_5\,
767: A^c_5\,\Big)\, \Big] c^b\; ,\nonumber
768: \end{eqnarray}
769: where $c^a (x,y)$ and $\bar{c}^a (x,y)$ are the 5D ghosts that are
770: even under $\Zb_2$.
771:
772: An important constraint in our formulation is the
773: condition~(\ref{surfcond}) of vanishing of the surface terms, i.e.~the
774: fact that total derivatives of fields or product of fields with
775: respect to $y$ do not contribute to the action. Unlike the naive
776: Fourier modes, the mass eigenmode wavefunctions $f_n (y)$ and $g_n
777: (y)$ comply with this crucial constraint and hence become the
778: appropriate basis of decomposition of the 5D fields
779: [cf.~(\ref{AmuExpansion})]. Detailed discussion is given in
780: Appendix~B, where we show explicitly that upon compactification,
781: $\Lag_{\rm 5D\GF}$ and $\Lag_{\rm 5D\FP}$ lead indeed to an effective
782: 4D theory quantized in the conventional $R_\xi$ gauge.
783:
784: It is not difficult to verify that the complete 5D QCD Lagrangian
785: (\ref{LagYMBKT}) is invariant under the 5D BRS transformations
786: \begin{eqnarray}
787: \label{BRS5D}
788: s A^a_M &=& \ D^{ab}_M\, c^b\;,\nonumber\\
789: s\, c^a &=& - \frac{g_5 f^{abc}}{2}\; c^b c^c\;,\\
790: s\, \overline{c}^a &=& \frac{F[A^a_M]}{\xi}\ .\nonumber
791: \end{eqnarray}
792: Following the Fourier-like convolution method developed in
793: Appendix~C.1, we derive the BRS transformations for the KK modes of
794: the effective theory
795: \begin{eqnarray}
796: \label{BRSeff}
797: s A^a_{(n) \mu} &=& \partial_{\mu} c^a_{(n)}\: -\: g f^{abc} \sum_{m,
798: l=0}^{\infty} \sqrt{2}^{\; -1-\delta_{n,0}-\delta_{m,0}-\delta_{l,0}}\,
799: \Delta_{n,m,l} \; A^c_{(m) \mu} c^b_{(l)}\; ,\nonumber\\
800: s A^a_{(n) 5} &=& - m_n c^a_{(n)}\: -\: g f^{abc} \sum_{m, l=0}^{\infty}
801: \sqrt{2}^{\; -1-\delta_{l,0}}\, \tilde{\Delta}_{n,m,l} \; A^c_{(m) 5}
802: c^b_{(l)}\; ,\nonumber\\
803: s c^a_{(n)} &=& -\; \frac{g f^{abc}}{2} \sum_{m, l=0}^{\infty}
804: \sqrt{2}^{\; -1-\delta_{n,0}-\delta_{m,0}-\delta_{l,0}}\, \Delta_{n,m,l}
805: \; c^b_{(m)} c^c_{(l)}\; ,\nonumber\\
806: s \overline{c}^a_{(n)} &=& \frac{1}{\xi}\, \partial^{\mu} A^a_{(n) \mu}\:
807: -\: m_n A^a_{(n) 5}\; .
808: \end{eqnarray}
809:
810: In close analogy to our discussion for the ordinary QCD case in the
811: previous section, we define the generating functional $Z$ of the
812: connected Green functions through the relation:
813: \begin{eqnarray}
814: \label{Z5D}
815: e^{iZ} &=& \int DA_M \, Dc \, D\overline{c} \ \exp\bigg\{\, i \int d^4x\,
816: \bigg[\,\int_{-\pi R}^{\pi R} dy\; \Lag_{\YMBKT}\\
817: &+&\!\!\! \int_{-\pi R}^{\pi R} dy\, [1 + r_c \delta (y)]\, \big(
818: J^{a M} A^a_M\: +\: \overline{D}^a c^a\: +\:
819: \overline{c}^a D^a\:
820: + \: K^{a M} s A^a_M\: +\: M^a sc^a\,\big)\, \bigg]\,\bigg\}\; .\nonumber
821: \end{eqnarray}
822: The response of the generating functional $Z$ under the 5D BRS
823: transformations gives rise to the following master ST identity for the
824: 5D theory:
825: \begin{equation}
826: \label{5DST}
827: J^{a M} \frac{\delta Z}{\delta K^{a M}}\ -\ \overline{D}^a
828: \frac{\delta Z}{\delta M^a}\ +\ \frac{1}{\xi}\; D^a\, \partial^\mu
829: \frac{\delta Z}{\delta J^{a \mu}}\ -\ D^a\,\partial_5\, \frac{\delta
830: Z}{\delta J^{a 5}} \ =\ 0\, .
831: \end{equation}
832: {}From~(\ref{5DST}), we obtain the effective master ST identity
833: \begin{equation}
834: \label{masterSTCYM}
835: \sum_{n=0}^{\infty}\; \bigg( J^{a \mu}_{(n)} \frac{\delta Z}{\delta K^{a
836: \mu}_{(n)}} + J^{a 5}_{(n)} \frac{\delta Z}{\delta K^{a 5}_{(n)}} -
837: \overline{D}^a_{(n)} \frac{\delta Z}{\delta M^a_{(n)}} + \frac{1}{\xi}
838: \partial^{\mu} \frac{\delta Z}{\delta J^{a \mu}_{(n)}} D^a_{(n)} -
839: m_n\; \frac{\delta Z}{\delta J^{a 5}_{(n)}} D^a_{(n)} \bigg)
840: \ =\ 0\;.
841: \end{equation}
842:
843: As in the ordinary QCD case, it proves more practical to derive ST
844: identities from the BRS invariance of Green functions. For instance,
845: we may translate the obvious identity,
846: \begin{equation}
847: \label{green1}
848: s\, \langle 0 | T \bar{c}^a_{(n)}(x) A^b_{(n) \nu} (z) A^c_{(n) \rho}
849: (w) A^d_{(n) \sigma} (u) |0 \rangle\ =\ 0\; ,
850: \end{equation}
851: into the corresponding on-shell ST identity, which is graphically
852: given by
853: \begin{align}
854: \begin{split}
855: \label{onshellST1}
856: \frac{p_1^{\mu}}{m_n} \hspace{-1mm} \BBBBblob{\mu}{\nu}{\rho}{\sigma}{(n)
857: p_1}{(n) \; p_2}{(n) \; k_1}{(n) \; k_2} &= \; i \hspace{-7mm}
858: \SBBBblob{}{\nu}{\rho}{\sigma}{}{}{}{} \hspace{-7mm} +
859: \frac{p_2^{\nu}}{m_n} \hspace{-5mm} \GGBBblob{}{}{\rho}{\sigma}{}{}{}{}\\
860: &\phantom{}\\ & \qquad \qquad + \frac{k_2^{\sigma}}{m_n} \hspace{-5mm}
861: \GBBGblob{}{\nu}{\rho}{}{}{}{}{} \hspace{-7mm} + \frac{k_1^{\rho}}{m_n}
862: \hspace{-7mm} \GBGBblob{}{\nu}{}{\sigma}{}{}{}{} \hspace{-7mm} \\
863: &\phantom{}
864: \end{split}
865: \end{align}
866:
867: \vspace{0.5cm}
868: \noindent
869: Likewise, the BRS invariance of the Green function pertinent to the
870: product of fields $\bar{c}^a_{(n)}(x) A^b_{(n) 5} (z) A^c_{(n) \rho}
871: (w) A^d_{(n) \sigma} (u)$, namely
872: \begin{equation}
873: \label{green2}
874: s\, \langle 0 | T \bar{c}^a_{(n)}(x) A^b_{(n) 5} (z) A^c_{(n) \rho} (w)
875: A^d_{(n) \sigma} (u) |0 \rangle\ =\ 0\; ,
876: \end{equation}
877: gives rise to following on-shell ST identity:
878: \begin{align}
879: \begin{split}
880: \label{onshellST2}
881: \frac{p_1^{\mu}}{m_n} \hspace{-1mm} \BSBBblob{\mu}{}{\rho}{\sigma}{(n) \;
882: p_1}{(n) \; p_2}{(n) \; k_1}{(n) \; k_2} &= \; i \hspace{-7mm}
883: \SSBBblob{}{}{\rho}{\sigma}{}{}{}{} \hspace{-7mm} -i \hspace{-4mm}
884: \GGBBblob{}{}{\rho}{\sigma}{}{}{}{}\\ &\phantom{}\\ & \qquad \qquad + \frac{k_2^{\sigma}}{m_n} \hspace{-5mm} \GSBGblob{}{}{\rho}{}{}{}{}{}
885: \hspace{-7mm} + \frac{k_1^{\rho}}{m_n} \hspace{-7mm}
886: \GSGBblob{}{}{}{\sigma}{}{}{}{} \hspace{-7mm} \\ &\phantom{}
887: \end{split}
888: \end{align}
889:
890: \vspace{0.5cm}
891: \noindent
892: As we will see in the next section, the above on-shell ST identities
893: are important to prove the GET in $2\to 2$ scatterings. In particular,
894: by virtue of these identities, we see again that the KK modes
895: $A^a_{(n) 5}$ behave very analogous to the would-be Goldstone bosons
896: of non-linearly realized spontaneously broken gauge
897: theories~\cite{Thompson,Ohl}. A detailed list of Feynman rules, including
898: would-be Goldstone-boson and ghost interactions, is given in Appendix
899: A.
900:
901:
902:
903: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
904:
905: \setcounter{equation}{0}
906: \section{The 5D Generalized Equivalence Theorem}
907:
908: In the high-energy limit of a spontaneously broken gauge theory, the
909: amplitude for emission or absorption of longitudinal massive vector
910: bosons equals, up to a phase and possibly up to a renormalization
911: scheme dependent constant, the amplitude for emission or absorption of
912: the associated unphysical would-be Goldstone modes. This high-energy
913: equivalence relation between the longitudinal gauge bosons and the
914: would-be Goldstone bosons constitutes the famous Equivalence Theorem
915: (ET)~\cite{CLT,LQT}.
916:
917: In the previous section, we have shown that the on-shell ST
918: identities, which are required for a rigorous proof of the
919: ET~\cite{CG}, are very analogous to those of a spontaneously broken
920: gauge theory, if the scalar gauge modes $A^{a}_{(n) 5}$ are identified
921: with the would-be Goldstone bosons. Such an identification has proven
922: very important in order to establish the validity of the ET in 5D
923: orbifold Yang--Mills theories without BKTs~\cite{Dicus}. Here, this is
924: extended to 5D orbifold theories {\em with} BKTs. In detail, the ET
925: reads:
926: \begin{equation}
927: \label{ET}
928: \begin{split}
929: T(A^{a_1}_{(n_1) L},\ldots A^{a_k}_{(n_k) L},S \to
930: A^{b_1}_{(m_1) L},\ldots A^{b_l}_{(m_l) L},S')\ & =\ \\
931: C \, i^k \, (-i)^l T(A^{a_1}_{(n_1) 5},\ldots A^{a_k}_{(n_k) 5},S
932: \to A^{b_1}_{(m_1) 5},\ldots & A^{b_l}_{(m_l) 5},S')
933: \ +\ \mathcal{O}(m_{n_i}/E ) \, ,
934: \end{split}
935: \end{equation}
936: where $m_{n_i}$ are the KK masses, $E$ is the centre of mass system
937: (c.m.s.)~energy of the high-energy scattering process. The parameter
938: $C$ is a constant that generally depends on the renormalization
939: scheme~\cite{Yao}. However, it is $C=1$ at the tree level and in
940: certain renormalization schemes that maintain the Ward identities of
941: the classical action~\cite{Papa,Denner}. Finally, $S, \, S'$
942: collectively denote all particles in the initial and final state which
943: have no longitudinal modes. In order for the ET~(\ref{ET}) to hold,
944: the underlying gauge structure of the theory has to ensure
945: cancellations of all terms that grow with energy in the amplitude on
946: the LHS of~(\ref{ET}). As we will see below, in higher-dimensional
947: models with BKTs, such cancellations are much more intricate, because
948: the complete infinite tower of the KK modes is involved.
949:
950:
951: The ET is a direct consequence of the so-called generalized
952: equivalence theorem (GET)~\cite{CG}. The GET goes beyond~(\ref{ET}),
953: by providing an exact relation between the relevant amplitudes which
954: includes the energetically subleading terms. To state the GET, we
955: first note that the longitudinal polarization vector
956: $\epsilon^{\mu}_L$ can be written as
957: \begin{align}
958: \label{longVec}
959: \epsilon^{\mu}_L(k)\ =\ k^{\mu}/m_n + a^{\mu}(k) \, ,
960: \end{align}
961: where $a^{\mu}(k) = \ord(m_n/E)$ is the remainder of the polarization
962: vector that vanishes in the high-energy limit. As an example, we may
963: consider the GET applied to the production of two longitudinal KK
964: gauge modes:
965: \begin{eqnarray}
966: \label{GETequ}
967: T(A^a_{(n) T} A^b_{(n) T} \rightarrow A^c_{(n) L} A^d_{(n) L}) &=& \nonumber\\
968: &&\hspace{-2cm}
969: -\, T(A^a_{(n) T} A^b_{(n) T} \rightarrow A^c_{(n) 5} A^d_{(n) 5})\:
970: -\: i T(A^a_{(n) T} A^b_{(n) T} \rightarrow A^c_{(n) 5} a^d_{(n)})\nonumber\\
971: &&\hspace{-2cm}
972: -i T(A^a_{(n) T} A^b_{(n) T} \rightarrow a^c_{(n)} A^d_{(n) 5})\: +\:
973: T(A^a_{(n) T} A^b_{(n) T} \rightarrow a^c_{(n)} a^d_{(n)}) \; .\qquad
974: \end{eqnarray}
975: In~(\ref{GETequ}), $a^a_{(n)}$ indicates an amplitude, where the
976: longitudinal polarization vector of the corresponding gauge boson has
977: been replaced by its remainder vector $a^{\mu}$. Hence, the
978: corresponding amplitudes are suppressed in the high-energy limit and
979: the ET~(\ref{ET}) is recovered.
980:
981: It is straightforward to apply the GET to processes with an arbitrary
982: number of longitudinal gauge modes. Specifically, the RHS
983: of~(\ref{GETequ}) may be found by performing the following four
984: operations:
985:
986: \begin{itemize}
987:
988: \item[(i)] Write down all amplitudes that result from replacing any number of longitudinal vector bosons by the respective would-be Goldstone bosons.
989:
990: \item[(ii)] Replace in each amplitude the remaining longitudinal
991: polarization vectors by $a^a_{(n)}$.
992:
993: \item[(iii)] Multiply each of the resulting amplitudes with the
994: factors $i^k (-i)^l$, where $k$ is the number of the would-be
995: Goldstone bosons in the initial state and $l$ the corresponding one in
996: the final state.
997:
998: \item[(iv)] Sum the complete set of the so-generated amplitudes.
999:
1000: \end{itemize}
1001: Observe that our GET relation in~(\ref{GETequ}) is consistent with the
1002: above rules.
1003:
1004: Let us present an explicit proof of the GET for the simple example
1005: of~(\ref{GETequ}), based on the ST identities discussed in the last
1006: section. Using the decomposition, $\epsilon_1 =- k_1/m_n + a_1$, with
1007: $k_1$ being the incoming momentum, the ST identity~(\ref{onshellST1})
1008: and the transversality condition $\epsilon \cdot k=0$, we find
1009: diagrammatically
1010: \begin{align}
1011: \label{proof2}
1012: &\BBBBblob{\epsilon_a}{\epsilon_b}{\epsilon_1}{\epsilon_2}{(n)p_1}{(n)p_2}
1013: {(n)k_1}{(n)k_2}\quad = \quad -i \hspace{-4mm}
1014: \BBSBblob{\epsilon_a}{\epsilon_b}{}{\epsilon_2}{}{}{}{}\:
1015: + \:
1016: \BBBBblob{\epsilon_a}{\epsilon_b}{a_1}{\epsilon_2}{}{}{}{}\\
1017: &\phantom{}\nonumber \\ &\phantom{}\nonumber \,
1018: \end{align}
1019: where the polarization vectors multiplying the amplitudes have been
1020: explicitly displayed at the corresponding external leg.
1021:
1022: Our next step consists in making use of $\epsilon_2=- k_2/m_n + a_2$,
1023: the ST identity~(\ref{onshellST2}), and the fact that $k_1 \cdot a_1 =
1024: m_n$. Substituting all the above into the RHS of (\ref{proof2}), we
1025: obtain
1026: \begin{align}
1027: \begin{split}
1028: \label{proof2general}
1029: \BBBBblob{\epsilon_a}{\epsilon_b}{\epsilon_1}{\epsilon_2}{(n)p_1}
1030: {(n)p_2}{(n)k_1}{(n)k_2}\ = \ - \hspace{-4mm}
1031: &\BBSSblob{\epsilon_a}{\epsilon_b}{}{}{}{}{}{} \hspace{-4mm} +\
1032: \BBGGblob{\epsilon_a}{\epsilon_b}{}{}{}{}{}{}\hspace{-4mm} +\ i\hspace{-4mm}
1033: \BBSBblob{\epsilon_a}{\epsilon_b}{}{a_2}{}{}{}{}\\ &\phantom{}\\
1034: &\phantom{}\\ &\quad -i \hspace{-4mm}
1035: \BBBSblob{\epsilon_a}{\epsilon_b}{a_1}{}{}{}{}{}\hspace{-4mm} -\
1036: \BBGGblob{\epsilon_a}{\epsilon_b}{}{}{}{}{}{} \hspace{-4mm} +\
1037: \BBBBblob{\epsilon_a}{\epsilon_b}{a_1}{a_2}{}{}{}{}\\
1038: &\phantom{}\\ &\phantom{}
1039: \end{split}
1040: \end{align}
1041: It is obvious that ghost contributions cancel, and hence
1042: (\ref{proof2general}) is nothing than (\ref{GETequ}) in diagrammatic
1043: form. Following the same reasoning, one can prove the GET for any
1044: other process. At the tree level, a proof based on the Ward
1045: identities (\ref{wardCYM1})--(\ref{wardCYM5}) is possible but tedious.
1046:
1047: Having established the GET and the resulting ET, we now show how the
1048: ET is realized in a specific tree-level calculation. Consider the
1049: elastic scattering of two longitudinally polarized gauge bosons,
1050: $A^a_{(n)L} A^b_{(n)L} \rightarrow A^c_{(n)L} A^d_{(n)L}$. In terms of
1051: Feynman diagrams, the tree-level amplitude is given by the infinite set
1052: of diagrams
1053: \begin{equation}
1054: \label{gaugeBKT}
1055: \begin{split}
1056: \BBBBblob{\epsilon_a}{\epsilon_b}{\epsilon_c}{\epsilon_d}{(n)p_1}{(n)p_2}
1057: {(n)k_1}{(n)k_2} &=
1058: \BBBB{\epsilon_a}{\epsilon_b}{\epsilon_c}{\epsilon_d}{(n)p_1}{(n)p_2}
1059: {(n)k_1}{(n)k_2} +\ \sum_{j=0}^{\infty}
1060: \sChannelB{\epsilon_a}{\epsilon_b}{\epsilon_c}{\epsilon_d}{(n)p_1}{(n)p_2}
1061: {(n)k_1}{(n)k_2}{\,
1062: \, \, \, (j)} \hspace{-1mm} +\ \textrm{crossings}\\ &\phantom{}\\ &=\
1063: iT_4+\sum_{j=0}^{\infty}\,
1064: \Big[\,iT^s_{(j)}+iT^t_{(j)}+iT^u_{(j)}\,\Big]\ =\
1065: iT_4+iT^s+iT^t+iT^u \, ,
1066: \end{split}
1067: \end{equation}
1068: where $s$, $t$ and $u$ refer to the pertinent channels. Using the
1069: Feynman rules of Appendix~A and the relation
1070: \begin{equation}
1071: \label{epsmu}
1072: \epsilon_a^{\mu}\ =\ \frac{1}{2 m_n \beta} \left[ (1+\beta^2)p_1^{\mu}
1073: - (1-\beta^2) p_2^{\mu} \right] \, ,
1074: \end{equation}
1075: with $\beta = \sqrt{1-4 m_n^2/s}$, it is straightforward to calculate
1076: the individual diagrams. More explicitly, we find
1077: \begin{eqnarray}
1078: \label{diagramamps}
1079: iT_4 &=& \Delta_{n,n,n,n} \, \frac{i g^2}{8 m_n^4} \bigg[\, f^{abe}
1080: f^{cde} s(t-u)\: +\: f^{ace} f^{bde} t\bigg(s-\frac{u}{\beta^4}\bigg)\: +\:
1081: f^{ade} f^{bce} u\bigg(s-\frac{t}{\beta^4}\bigg)\, \bigg] \, , \nonumber\\
1082: iT^s_{(j)} &=& 2^{-\delta_{j,0}} \Delta^2_{n,n,j} \, i g^2 f^{abe}
1083: f^{cde} \frac{s(u-t)}{8 m_n^4}\, \bigg(\, 1\: +\: \frac{2m_n^2}{s} \bigg)^2 \,
1084: \frac{s}{s-m_j^2} \, , \\
1085: iT^t_{(j)} &=& 2^{-\delta_{j,0}}
1086: \Delta^2_{n,n,j} \, i g^2 f^{ace} f^{bde}\, \bigg[\, \frac{u-s}{2t}\,
1087: \bigg(\, 1\: +\: \frac{t}{2 m_n^2 \beta^2} \bigg)^2\: +\:
1088: \frac{t-2u}{m_n^2 \beta^2}\, \bigg]\; \frac{t}{t-m_j^2} \, , \nonumber\\
1089: iT^u_{(j)} & = & 2^{-\delta_{j,0}}
1090: \Delta^2_{n,n,j} \, i g^2 f^{ade} f^{bce}\, \bigg[\, \frac{t-s}{2u}
1091: \bigg(\, 1\: +\: \frac{u}{2 m_n^2 \beta^2} \bigg)^2 \:
1092: +\: \frac{u-2t}{m_n^2 \beta^2}\, \bigg]\; \frac{u}{u-m_j^2} \, ,\nonumber
1093: \end{eqnarray}
1094: where the Mandelstam variables are given by
1095: \begin{equation}
1096: s = (p_1+p_2)^2 \, , \quad
1097: t = (p_1+k_1)^2 =-(1 - c_\theta ) \frac{\beta^2 s}{2} \, , \quad
1098: u = (p_1+k_2)^2 =-(1 + c_\theta) \frac{\beta^2 s}{2} \, , \quad
1099: \end{equation}
1100: and we have used the notation $c_\theta = \cos \theta$.
1101:
1102: Our task of verifying the ET simplifies significantly if the infinite
1103: sums are expanded in terms of $s$. To be precise, using
1104: \begin{equation}
1105: \label{sFactor}
1106: \frac{s}{s-m_j^2}\ =\ 1\: +\: \frac{m_j^2}{s}\: +\: \frac{m_j^4}{s^2}\: +\:
1107: \frac{m_j^4}{s^2} \frac{m_j^2}{s-m_j^2}\ ,
1108: \end{equation}
1109: the infinite sum $T^{s}$ can be split into four different sums: $T^{s}
1110: = T^{s}_{0} + T^{s}_{2} + T^{s}_{4} + T^{s}_{6}$, where the subscript
1111: indicates the power of $m_j$ in the numerator of (\ref{sFactor}).
1112: Note that each sum is convergent. As can be shown by an explicit
1113: calculation (see end of this section for a similar case), $T^{s}_{6}$
1114: includes terms of order $m_n/\sqrt{s}$, which are energetically
1115: suppressed and do not contribute to the leading part of the amplitude.
1116: Likewise, we can make analogous expansions for the $t$- and
1117: $u$-exchange graphs:
1118: \begin{equation}
1119: \begin{split}
1120: \frac{t}{t-m_j^2}\ &=\ 1 - \left( \frac{2}{1-c_\theta} \, \frac{1}{s} +
1121: \frac{8 m_n^2}{1-c_\theta} \, \frac{1}{s^2} \right) m_j^2 + \left(
1122: \frac{4}{(1-c_\theta)^2} \, \frac{1}{s^2} \right) \, m_j^4\ +\
1123: \ord( m^6_j s^{-3}) \, , \\
1124: \frac{u}{u-m_j^2}\ &=\ 1 -
1125: \left(\frac{2}{1+c_\theta} \, \frac{1}{s} + \frac{8m_n^2}{1+c_\theta}
1126: \, \frac{1}{s^2} \right) m_j^2 + \left( \frac{4}{(1+c_\theta)^2} \,
1127: \frac{1}{s^2} \right) \, m_j^4\ +\ \ord(m^6_j s^{-3}) \, .
1128: \end{split}
1129: \end{equation}
1130: The different infinite sums in $T^{s,t,u}_{0}$, $T^{s,t,u}_{2}$, and
1131: $T^{s,t,u}_{4}$ can be calculated by means of (\ref{Snnnn1}) and
1132: (\ref{Snnnn2}), which have been derived by employing complex analysis
1133: techniques developed in Appendix~B of~\cite{Paes}. In this way, we
1134: obtain
1135: \begin{equation}
1136: \label{T4}
1137: \begin{split}
1138: i T_4\ =\ i g^2\,\Delta_{n,n,n,n} \; \bigg[\, &f^{abe} f^{cde}\, \bigg(
1139: \frac{c_\theta}{8 m_n^4} s^2\ -\ \frac{c_\theta}{2 m_n^2} s\, \bigg)\\
1140: &+\, f^{ace} f^{bde}\, \bigg(\,
1141: \frac{(c_\theta+3)(c_\theta-1)}{32m_n^4} s^2\ +\
1142: \frac{1-c_\theta}{4m_n^2} s\, \bigg)\\ &+\, f^{ade} f^{bce}
1143: \bigg(\,\frac{(c_\theta-3)(c_\theta+1)}{32m_n^4} s^2\ +\
1144: \frac{c_\theta+1}{4m_n^2} s \bigg)\, \bigg]\; ,
1145: \end{split}
1146: \end{equation}
1147: \begin{eqnarray}
1148: \label{Ts}
1149: iT^s &=& ig^2 \; f^{abe} f^{cde} \; \bigg[\, \Delta_{n,n,n,n}\,
1150: \bigg( -\, \frac{c_{\theta}}{8 m_n^4} s^2\: -\:
1151: \frac{c_{\theta}}{6 m_n^2} s\: +\:
1152: \frac{5 c_{\theta}}{4}\, \bigg)\nonumber\\
1153: &&\hspace{3cm} -\ \frac{c_{\theta}}{2} X_n\, \bigg]\ +\
1154: \ord(m_n/\sqrt{s})\; ,\\[4mm]
1155: \label{Tt}
1156: iT^t &=& ig^2 \; f^{ace} f^{dbe} \; \bigg[\, \Delta_{n,n,n,n}\, \bigg(\,
1157: \frac{(c_{\theta}+3)(1-c_{\theta})}{32 m_n^4} s^2\: +\: \frac{11
1158: c_{\theta}-3}{12 m_n^2} s\: +\: \frac{8
1159: c_{\theta}^2-5c_{\theta}+9}{6(1-c_{\theta})}\, \bigg)\nonumber\\
1160: &&\hspace{3cm} +\; \frac{3+c_{\theta}}{2(1-c_{\theta})}\, X_n\bigg]\
1161: +\ \ord(m_n/\sqrt{s})\; ,\\[4mm]
1162: \label{Tu}
1163: iT^u &=& ig^2 \; f^{ade} f^{bce} \; \bigg[\, \Delta_{n,n,n,n}\, \bigg(\,
1164: \frac{(3-c_{\theta})(1+c_{\theta})}{32 m_n^4} s^2\: -\: \frac{11
1165: c_{\theta}+3}{12 m_n^2} s\:
1166: +\: \frac{8 c_{\theta}^2+5c_{\theta}+9}{6(1+c_{\theta})}\, \bigg)\nonumber\\
1167: &&\hspace{3cm} +\; \frac{3-c_{\theta}}{2(1+c_{\theta})}\, X_n\, \bigg]\
1168: +\ \ord(m_n/\sqrt{s})\, ,
1169: \end{eqnarray}
1170: where
1171: \begin{equation}
1172: \label{Xn}
1173: X_n\ =\ 8\, N_n^4\,\pi^2 R^2 \tilde{r}_c^3 m_n^2 \; .
1174: \end{equation}
1175: Collecting all contributions (\ref{T4})--(\ref{Tu}), we find that the
1176: $s^2$-contributions in the $s$-, $t$- and $u$-channel graphs cancel
1177: against terms in $T_4$. Terms linear in $s$ are identical for each
1178: colour factor. Thus, they vanish due to the Jacobi identity. The
1179: final result is then given by
1180: \begin{align}
1181: \label{LOgaugeBKT}
1182: iT_4&+iT^s+iT^t+iT^u\\
1183: &=\ i g^2\,(\Delta_{n,n,n,n} + X_n)\, \bigg[\, f^{ace} f^{dbe}
1184: \frac{c^2_\theta + 3}{2 (c_\theta - 1)}\:
1185: +\: f^{ade} f^{bce} \frac{c^2_\theta + 3}{2(c_\theta + 1)}\, \bigg]\ +\
1186: \ord ( m_n/\sqrt{s})\; . \nonumber
1187: \end{align}
1188: In (\ref{LOgaugeBKT}), we only exhibit the leading contribution
1189: $\ord(1)$, which will be essential for checking the validity of the
1190: ET.
1191:
1192: Let us now consider the RHS of the ET~(\ref{ET}). The scalar
1193: scattering amplitude is given by
1194: \begin{equation}
1195: \begin{split}
1196: \label{scalarBKT}
1197: \SSSSblob{}{}{}{}{(n)p_1}{(n)p_2}{(n)k_1}{(n)k_2} \hspace{-2mm}\ &=\
1198: \sum_{j=0}^{\infty}
1199: \sChannelS{}{}{}{}{(n)p_1}{(n)p_2}{(n)k_1}{(n)k_2}{\, \, \, \, (j)}
1200: \hspace{-2mm}\ +\quad \textrm{crossings}\\ &\phantom{}\\ & =\
1201: \sum_{j=0}^{\infty}\, \Big[\, iT^s_{(j)}+iT^t_{(j)}+iT^u_{(j)}\,
1202: \Big]\ =\ iT^s+iT^t+iT^u\, .
1203: \end{split}
1204: \end{equation}
1205: Using the Feynman rules of Appendix~A, we obtain for the individual
1206: $s$-, $t$- and $u$-channel graphs
1207: \begin{equation}
1208: \begin{split}
1209: \label{Scalar}
1210: iT^s_{(j)}\ &=\ i g^2\: 2^{-\delta_{j,0}} \tilde{\Delta}^2_{n,j,n} \,
1211: f^{abe} f^{cde}\; \frac{u-t}{2(s-m_j^2)} \ , \\
1212: iT^t_{(j)}\ &=\ i g^2\: 2^{-\delta_{j,0}} \tilde{\Delta}^2_{n,j,n} \,
1213: f^{ace} f^{dbe}\; \frac{s-u}{2(t-m_j^2)} \ , \\
1214: iT^u_{(j)}\ &=\ i g^2\:
1215: 2^{-\delta_{j,0}} \tilde{\Delta}^2_{n,j,n} \, f^{ade} f^{bce}\;
1216: \frac{t-s}{2(u-m_j^2)} \ .
1217: \end{split}
1218: \end{equation}
1219: Considering the analytic form of the couplings
1220: $\tilde{\Delta}_{n,j,n}$ in (\ref{Scalar}), it turns out that only a
1221: single expansion in terms of $1/s$, i.e.
1222: \begin{equation}
1223: \label{sumsj}
1224: \frac{s}{s-m_j^2}\ =\ 1\: +\: \frac{m_j^2}{s-m_j^2} \, ,
1225: \end{equation}
1226: will be sufficient to evaluate the leading part of the amplitude.
1227: Using (\ref{Snnnn2}), we can now sum the leading part of
1228: (\ref{Scalar}). Up to energetically subleading terms $\ord(
1229: m_n/\sqrt{s})$, the amplitude (\ref{scalarBKT}) equals
1230: (\ref{LOgaugeBKT}). This completes our check of the ET (\ref{ET}).
1231: For vanishing $r_c$, i.e.~$r_c \to 0$, we find $\Delta_{n,n,n,n} \to
1232: 3$ and $X_n \to 0$, and so recover the result in~\cite{Dicus} for the
1233: validity of the ET in 5D orbifold theories without BKTs.
1234:
1235: We conclude this section by showing that the infinite sum involving
1236: the second term in~(\ref{sumsj}) is indeed subleading of order
1237: $m_n/\sqrt{s}$. We modify the second equation of (\ref{Snnnn2}) by
1238: taking the additional factor $m_j^2/(s-m_j^2)$ into account. With
1239: this modification, an explicit calculation gives
1240: \begin{equation}
1241: \label{surfaceSum}
1242: \begin{split}
1243: \sum_{j=0}^{\infty} 2^{-\delta_{j,0}}\, \tilde{\Delta}^2_{n,j,n}\;
1244: \frac{m_j^2}{s-m_j^2}\ =\ \, &\tilde{\Delta}_{n,n,n,n}\, \frac{4
1245: m_n^2}{s-4 m_n^2}\\ &\hspace{-2cm}+\; 32 \pi^2 R^2 m_n^2 \tilde{r}_c^3
1246: N_n^4 \, \Big(\, 1 - \pi^2 R^2 m_n^2 \tilde{r}_c^2\, \Big)\, \bigg[\,
1247: \frac{m_n^2}{s-4 m_n^2} + \frac{m_n^4}{(s-4 m_n^2)^2}\,\bigg]\\
1248: &\hspace{-2cm}+\, \frac{8 \pi^4 R^4 m_n^4 \tilde{r}_c^5 N_n^4}{s+
1249: \frac{\sqrt{s}}{\pi R \tilde{r}_c} \tan \pi R \sqrt{s}} \,
1250: \frac{s^3-4m_n^2 s^2 + 4 m_n^4 s}{(s-4 m_n^2)^2}\ -\ 8 \pi^4 R^4 m_n^4
1251: \tilde{r}_c^5 N_n^4 \; .
1252: \end{split}
1253: \end{equation}
1254: The first two terms are obviously $\ord(m^2_n/s)$. The third term has
1255: got poles at the spectrum and we therefore take the high-energy limit
1256: in a discrete manner: $\sqrt{s_n} = (n-\frac{1}{4})/R$, with $n \to
1257: \infty$. The third term approaches the negative of the fourth term.
1258: Consequently, all $\ord(1)$ terms cancel and (\ref{surfaceSum}) does
1259: contribute only subleading terms of order $m_n/\sqrt{s}$.
1260:
1261: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1262:
1263: \setcounter{equation}{0}
1264: \section{High Energy Unitarity Bounds}
1265: \label{bounds}
1266: \setcounter{equation}{0}
1267:
1268: In higher-dimensional field theories, the coupling constants are
1269: dimensionful parameters. For example, for the case of 5D $N_c$-colour
1270: QCD, the coupling $g_5$ has the energy dimensions $E^{-1/2}$. On
1271: grounds of naive dimensional analysis, the $s$-wave amplitude $a_0$ of
1272: a $2 \to 2$ scattering involving 5D gluons behaves as
1273: \begin{equation}
1274: \label{a0E5}
1275: a_0 (E_5) \ \sim \ N_c\, g^2_5\, E_5\; ,
1276: \end{equation}
1277: where $E_5$ is the c.m.s.~energy in 5D. This means that at energy
1278: scales $E_5 \stackrel{>}{{}_\sim} 1/(g^2_5\,N_c)$, the transition
1279: amplitude $a_0$ exceeds 1, thereby invalidating perturbation theory.
1280: In addition, the fact that the asymptotic high-energy behaviour of
1281: $a_0$ does not approach a constant, but grows linearly with the
1282: energy, is in gross violation with the Froissart--Martin
1283: bound~\cite{Froissart}.
1284:
1285: In a theory compactified to 4D, the 3- and 4-gluon couplings are
1286: dimensionless, while the KK spectrum is infinite. In this case, the
1287: aforementioned energy upper limit due to unitarity violation is
1288: translated into a corresponding upper limit on the number of the KK
1289: modes. To be specific, one gets the very approximate upper
1290: bound~\cite{Dicus}
1291: \begin{equation}
1292: \frac{N_0}{R} \ \stackrel{<}{{}_\sim} \ \frac{1}{g^2_5\,N_c }\ .
1293: \end{equation}
1294: Obviously, this upper bound depends non-trivially on the
1295: compactification radius $R$, the 5D gauge-coupling constant $g_5$ and
1296: the number of colours $N_c$.
1297:
1298: We will now compute explicitly the perturbative unitarity limits on
1299: our 5D orbifold theory with one BKT. For this purpose, let us first
1300: describe our approach to perturbative unitarity. As usual, we start
1301: by stating the optical theorem for a general $i \to f$ transition
1302: \begin{equation}
1303: \label{OT}
1304: 2\, \operatorname{Im} T_{f i}\ =\ \sum_j \; T_{f j}\, T^*_{j i}\; ,
1305: \end{equation}
1306: where the sum over $j$ is understood to include phase-space correction
1307: factors for each of the individual states in $j$.
1308: Decomposing the transition amplitude $T_{fi}$ in terms of Legendre
1309: polynomials~\cite{Barton},
1310: \begin{equation}
1311: \label{pwExp}
1312: T_{fi}(s,c_\theta)\ =\ 16 \pi\, \sum_{l=0}^{\infty}\, (2l+1)\,
1313: P_l(c_\theta)\; a_l (s)\; ,
1314: \end{equation}
1315: with $a_l (s) = 1/(32\pi)\, \int_{-1}^{+1}\, dc_\theta\,
1316: T_{fi} P_l(c_{\theta})$, and substituting the resulting expression~(\ref{pwExp})
1317: into~(\ref{OT}), we obtain a unitarity relation for the $s$-wave
1318: ($l=0$) amplitudes
1319: \begin{equation}
1320: \label{a0OT}
1321: {\rm Im}\, [a_0]_{f i}\ =\ \sum_j \; \sigma_j \; [a_0]_{f j} \;
1322: [a_0]^*_{j i} \; .
1323: \end{equation}
1324: Here, we denote with $\sigma_j = \lambda(s, m_1^2,m_2^2)/s$ the
1325: phase-space factors for 2-particle states, with $\lambda (x,y,z) =
1326: [ x^2 + y^2 + z^2 - 2(x y + y z + z x) ]^{1/2}$.
1327:
1328: Our approach to perturbative unitarity consists in absorbing the
1329: phase-space factors $\sigma_j$ into the definition of the transition
1330: amplitudes $[a_0]_{i j}$, i.e.
1331: \begin{equation}
1332: \label{a0tilde}
1333: [\tilde{a}_0]_{ij}\ =\ \sqrt{\sigma_i \sigma_j} \; [a_0]_{ij}\; .
1334: \end{equation}
1335: This enables us to cast~(\ref{a0OT}) in the simple matrix form
1336: \begin{equation}
1337: \label{OTtilde}
1338: {\rm Im}\, \tilde{a}_0\ =\ \tilde{a}_0\, \tilde{a}_0^*\; .
1339: \end{equation}
1340: Since our unitarity limits will be deduced from $2 \to 2$ scatterings
1341: that involve longitudinal KK gauge modes, $\tilde{a}_0$ will turn out
1342: to be a square symmetric matrix. Therefore, $\tilde{a}_0$ can be
1343: diagonalized by means of a {\em real} orthogonal matrix $R$ as
1344: \begin{equation}
1345: \label{ROT}
1346: R^T\, \tilde{a}_0\, R \ =\ \hat{a}_0\;,
1347: \end{equation}
1348: where $\hat{a}_0$ is a diagonal transition amplitude matrix. The
1349: unitarity relation~(\ref{OTtilde}) implies that not only
1350: $\tilde{a}_0$, but also its imaginary part ${\rm Im}\, \tilde{a}_0$
1351: can be diagonalized by the same orthogonal matrix $R$. Consequently,
1352: the following unitarity relation may be established:
1353: \begin{equation}
1354: \label{OTrelation}
1355: {\rm Im}\, \hat{a}_0\ =\ \big({\rm Re}\,\hat{a}_0\big)^2\:
1356: +\: \big({\rm Im}\,\hat{a}_0\big)^2\;.
1357: \end{equation}
1358: Equation~(\ref{OTrelation}) directly implies the inequality
1359: \begin{equation}
1360: \label{PUL}
1361: {\rm Im}\, \hat{a}_0\ \leq \ {\bf 1}\; ,
1362: \end{equation}
1363: which is equivalent, by means of~(\ref{OTtilde}), to the requirement
1364: that the largest eigenvalue of $\hat{a}_0$ or equivalently of
1365: $\tilde{a}_0$ (in absolute value terms) should smaller than 1. Notice
1366: that our approach to perturbative unitarity differs from the one
1367: applied originally for the SM case~\cite{LQT}, in the fact that
1368: phase-space effects were not considered in the latter. However, such
1369: effects are non-negligible for the heavier KK modes, and should
1370: therefore be included consistently.
1371:
1372: Since the 5D transition amplitudes diverge with increasing energy
1373: [cf.~(\ref{a0E5})], the unitarity constraint~(\ref{PUL}) will
1374: explicitly depend on the c.m.s.~energy $\sqrt{s}$. For this reason,
1375: perturbative unitarity limits will constrain not only the maximum
1376: allowed number of KK modes, e.g.~$N_{\rm max}$, but also the
1377: c.m.s.~energy, i.e.~$\sqrt{s}_{\rm max} = 4m^2_{N_{\rm max}}$, above
1378: which unitarity is violated. To avoid complications arising mainly
1379: from IR singularities in the KK gluon scatterings, we
1380: follow~\cite{Dicus} and perform a coupled channel analysis (CCA),
1381: based on the inelastic colour-singlet processes $A^a_{(n)5} A^a_{(n)5}
1382: \to A^b_{(m)5} A^b_{(m)5}$, with $n\neq m$ and $n,m \leq N_{\rm max}$
1383: and $s = 4 m^2_{N_{\rm max}}$. Strictly speaking, our CCA relies on
1384: the physical processes $A^a_{(n)L} A^a_{(n)L} \to A^b_{(m)L}
1385: A^b_{(m)L}$, but we have made use of the ET to replace the
1386: longitudinal KK gauge bosons $A^a_{(n)L}$ with their respective
1387: would-be Goldstone bosons $A^a_{(n)5}$. Such a consideration is a
1388: good approximation for the lighter KK modes, with masses much smaller
1389: than the c.m.s~energy $\sqrt{s}$. For the heavier KK modes, with
1390: masses $m_{(n)} \sim \sqrt{s}$, our simplification may be justified by
1391: the fact that scatterings of those heavy states will be relatively
1392: suppressed by phase-space factors that occur in the transition
1393: amplitudes $[\tilde{a}_0]_{nm}$ defined
1394: in~(\ref{a0tilde}). Nevertheless, our derived upper bounds should be
1395: regarded to be slightly conservative, in the sense that we have not
1396: taken into account $2\to 2$ scatterings where all asymptotic KK gluon
1397: states have different masses. However, initial estimates convinced us
1398: that these contributions to the CCA are really
1399: negligible.\footnote{These estimates will be further consolidated by
1400: our analysis of unitarity bounds from the decay widths of the KK
1401: states, which is given at the end of the section.}
1402:
1403:
1404: The phase-space-corrected $s$-wave amplitudes $[\tilde{a}_0]_{nm}$ for
1405: the $2\to 2$ inelastic scatterings, $A^a_{(n)5} A^a_{(n)5} \to
1406: A^b_{(m)5} A^b_{(m)5}$, are given by
1407: \begin{equation}
1408: \label{a0jSum}
1409: [\tilde{a}_0]_{nm}\ =\ \frac{S_{n,m}}{s}
1410: \sum_{j=0}^{\infty}\; [a^{(j)}_0]_{nm}\; ,
1411: \end{equation}
1412: where $S_{n,m} = \sqrt{(s-4m_n^2)(s-4m_m^2)}$ are phase-space
1413: correction factors and $[a^{(j)}_0]_{nm}$ are the individual $s$-wave
1414: amplitudes mediated by the exchange of the $j$ KK gauge mode $A^c_{(j)
1415: \mu}$:
1416: \begin{eqnarray}
1417: [a_0^{(j)}]_{nm} &=& -\, \frac{g^2 N_c}{32 \pi}\; \tilde{\Delta}_{n,j,m}^2 \;
1418: 2^{-\delta_{j,0}}\nonumber\\
1419: &&\times\; \bigg[\, 1\: +\: \frac{2 (s - m_n^2 - m_m^2) + m_j^2}{S_{n,m}}\;
1420: \ln\bigg| \frac{s + 2(m_j^2 - m_n^2 - m_m^2) - S_{n,m}}{s +
1421: 2 (m_j^2 - m_n^2 - m_m^2) + S_{n,m}} \bigg|\; \bigg]\; .\qquad
1422: \end{eqnarray}
1423: It is interesting to observe that in the high-energy limit $s \gg (m_n
1424: + m_m)^2$, the $s$-wave amplitudes $[a_0^{(j)}]_{nm}$ diverge
1425: logarithmically,
1426: \begin{equation}
1427: [a^{(j)}_0]_{nm} \ = \ -\, \frac{g^2 N_c}{32 \pi}\,
1428: \tilde{\Delta}_{n,j,m}^2 \; 2^{-\delta_{j,0}}\; \bigg[\, 1\: -\: 2 \ln
1429: \bigg(\frac{s}{m_j^2}\bigg)\, \bigg]\ +\ {\cal O}\bigg(\frac{(m_n +
1430: m_m)^2}{s}\bigg) \; .
1431: \end{equation}
1432: In the limit $r_c \to 0$, the terms $j=n+m$ and $j=|n-m|$ dominate the
1433: sum (\ref{a0jSum}) and the result~\cite{Dicus} of a 5D
1434: orbifold theory without BKTs is recovered
1435: \begin{equation}
1436: [a_0]_{nm}\ =\ -\,
1437: \frac{g^2 N_c}{32 \pi}\, \bigg[\, 1\: -\:
1438: 2 \ln \bigg(\frac{s}{|n^2/R^2-m^2/R^2|}\bigg)\, \bigg]\; .
1439: \end{equation}
1440: Because of the presence of the BKT, our CCA involves the infinite
1441: sum~(\ref{a0jSum}), thus making it technically more challenging than
1442: the corresponding analysis without BKTs. We perform this infinite sum
1443: and find the maximum eigenvalue $\lambda_{\rm max}$ of $\tilde{a}_0$
1444: numerically.
1445:
1446: \begin{figure}[t]
1447: \begin{center}
1448: \includegraphics[width=0.48\textwidth]{bound1.eps}
1449: \includegraphics[width=0.48\textwidth,]{bound2.eps}
1450: \caption{\em Contour plots in the plane $(\tilde{r}_c = r_c/(2\pi R),
1451: g^2N_c)$, for fixed values of $N_{\rm max}$. The area above the
1452: contour lines is excluded by perturbative unitarity.}\label{contour}
1453: \end{center}
1454: \end{figure}
1455:
1456:
1457: In Figure~\ref{contour}, we display contour plots in the plane
1458: $(\tilde{r}_c, g^2 N_c)$, for different fixed values of the maximum
1459: KK-mode number $N_{\rm max}$, which is obtained by requiring the
1460: perturbative unitarity constraint: $|\lambda_{\rm max}| \leq 1$. The
1461: area that lies above the contour lines is excluded by perturbative
1462: unitarity. For the case without BKTs, $\tilde{r}_c = r_c/(2\pi R) =
1463: 0$, our upper limits are larger by a about of factor 2 than those
1464: presented in~\cite{Dicus}. The origin of this difference may be
1465: traced to the effect of the phase-space factors $\sigma_j$, which have
1466: not been included in the analysis of~\cite{Dicus}. In the presence of
1467: BKTs, we find that the unitarity bounds show only a weak dependence on
1468: the size $\tilde{r}_c$ of the BKTs. In particular, the maximum
1469: allowed value of $g^2 N_c$ thanks to perturbative unitarity increases
1470: only by about $5\%$. This behaviour may be understood by the fact
1471: that high-energy unitarity probes distances much smaller than the
1472: compactification radius $R$ where the effect of the BKTs becomes less
1473: significant.
1474:
1475: \begin{figure}[t]
1476: \begin{center}
1477: \includegraphics[width=0.70\textwidth]{width1.eps}\\ \vspace{3mm}
1478: \includegraphics[width=0.70\textwidth]{width2.eps}
1479: \caption{\em Numerical values for $\Gamma_n/2 m_n$ as functions of
1480: $\tilde{r}_c$ and $n$}\label{decayWidths}
1481: \end{center}
1482: \end{figure}
1483:
1484: An interesting consequence of the BKT is that the KK gauge modes are
1485: no longer stable. One would have naively expected that the larger the
1486: size $\tilde{r}_c$ of the BKT is, the larger the decay width of the $n$th KK
1487: gauge boson becomes, thereby leading to a potential violation of
1488: perturbative unitarity. However, we will show that this is {\em not}
1489: the case, in agreement with our results from the $2\to 2$ scatterings.
1490: Perturbative unitarity in the decays of the KK gauge bosons gives rise
1491: to the constraint,
1492: \begin{equation}
1493: \label{widthCond}
1494: \frac{1}{2}\,\Gamma_n\ \leq\ m_n\ ,
1495: \end{equation}
1496: where $\Gamma_n$ is the total decay width of the $n$th KK gauge boson
1497: $A^a_{(n)\mu}$. At the lowest order of perturbation theory, only
1498: two-body decays of $A^a_{(n)\mu}$ into lighter KK gauge bosons will
1499: contribute. The total decay width $\Gamma_n$ is then the sum of the
1500: partial widths and may conveniently be expressed as
1501: \begin{equation}
1502: \label{widthsSum}
1503: \Gamma_n\ =\ \frac{1}{16 \pi m_n^3}\ \sum_{k,l=0 \; (k+l \leq n)}
1504: \lambda (m^2_n, m^2_k, m^2_l)\ |T_{(n,k,l)}|^2\ ,
1505: \end{equation}
1506: where $T_{(n,k,l)}$ is the transition amplitude for decay
1507: $A^a_{(n)\mu} \to A^b_{(k)\mu} A^c_{(l)\mu}$, with $n \ge k
1508: +l$. Squaring $T_{(n,k,l)}$ and averaging over initial states, we find
1509: \begin{eqnarray}
1510: \label{pwidth}
1511: |T_{(n,k,l)}|^2 & =& \frac{1}{3} g^2 N_c \pi^2 R^2 \tilde{r}_c^3 N_n N_k
1512: N_l\; \Delta_{n,k,l}\nonumber\\
1513: &&\times\, \Big[\, m_n^4 + m_k^4 + m_l^4 +
1514: 10\big(\, m_n^2 m_k^2 + m_k^2 m_l^2 +m_l^2 m_n^2\,\big)\, \Big]\, .
1515: \end{eqnarray}
1516: Here, we should note that not all kinematically allowed decay channels
1517: of $A^a_{(n)\mu}$ contribute to $\Gamma_n$. For example, the
1518: transition amplitudes $|T_{(n,k,0)}|$, with $0 \leq k < n$, vanish
1519: identically.
1520:
1521: In Fig.~\ref{decayWidths}, we present numerical estimates of the
1522: quantity $\Gamma_n/(2m_n)$, as functions of the BKT coupling
1523: $\tilde{r}_c$ and the KK number $n$ of the decaying
1524: gauge boson $A^a_{(n)\mu}$. After its rapid increase, the value of
1525: $\Gamma_n/(2m_n)$ saturates to values much smaller than 1 at large
1526: $\tilde{r}_c$ and/or large $n$. Consequently, we find as before that
1527: BKTs do not lead by themselves to a violation of perturbative
1528: unitarity in the decays of the KK gauge bosons.
1529:
1530:
1531:
1532: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1533:
1534: \setcounter{equation}{0}
1535: \section{Conclusions}
1536: \label{conclusion}
1537: \setcounter{equation}{0}
1538:
1539: We have studied the quantization and high-energy unitarity of 5D field
1540: theories compactified on an $S^1/\mathbb{Z}_2$ orbifold that include
1541: localized gauge-kinetic terms at the orbifold fixed points. These
1542: localized interactions, the so-called BKTs, emerge naturally in these
1543: theories as counterterms at tree-level to absorb the UV infinities that
1544: are generated at the quantum level from operators of the same form.
1545:
1546: Extending our approach to quantization in~\cite{MPR}, we have
1547: developed a functional differentiation formalism to quantize the
1548: theory within the framework of generalized $R_\xi$ gauges, before the
1549: KK reduction. With the aid of this formalism, we were able to derive
1550: the master Ward and ST identities which is a reflection of the
1551: underlying gauge and BRS symmetries of the theory. The Ward and ST
1552: identities relate the divergence of a given Green function to other
1553: Green functions of the theory. By virtue of such identities, we have
1554: stated a generalized form of the ET, which we call the 5D GET. The 5D
1555: GET extends the equivalence relation between the longitudinal KK gauge
1556: modes and their respective would-be Goldstone bosons to consistently
1557: include the energetically suppressed terms in high-energy scatterings.
1558:
1559: An important property of any perturbative predictive framework of
1560: field theory is perturbative unitarity. Requiring perturbative
1561: unitarity, we have deduced upper limits on the number of the KK modes
1562: as functions of the size of the gauge coupling and the colour of the
1563: 5D Yang--Mills theory. Our approach to perturbative unitarity takes
1564: into consideration phase space corrections due to heavier KK modes and
1565: so improves upon earlier studies on the same topic~\cite{Dicus}.
1566: Because of this, the derived upper limits have been found to be weaker
1567: by a factor $\sim 2$ than those obtained in these studies. We have
1568: investigated the impact of the BKTs on these unitarity limits. We
1569: have shown that the unitarity limits weakly depend on the size of the
1570: BKTs. Such a behaviour may be attributed to the fact that at high
1571: energies distances smaller than the compactification radius are
1572: probed. As a consequence, high-energy unitarity strongly depends on
1573: the bulk dynamics of the theory, and therefore high-energy unitarity
1574: limits show a weak dependence on the size of the BKTs.
1575:
1576: Even though large BKTs may weakly affect the unitarity limits on the
1577: theory, they can, however, make the higher KK modes decouple from the
1578: dynamics of the lowest lying KK states~\cite{Carena}, e.g.~that of the
1579: gluons. Their presence may help to considerably alleviate the severe
1580: phenomenological constraints that apply to those theories from a full
1581: fledged analysis of the electroweak data~\cite{MPR2}. It is therefore
1582: of great phenomenological interest to analyze in detail the prospects
1583: of probing such BKT-dominated theories at the LHC and at future
1584: $e^+e^-$ linear colliders.
1585:
1586:
1587: \subsection*{Acknowledgements}
1588: We wish to thank Alex Pomarol for discussions, as well as
1589: Francisco del Aguila, Jose Santiago and Manuel Perez-Victoria for
1590: pointing out Ref.~\cite{Santiago} to us. The work of AP is
1591: supported in part by the PPARC research grant PPA/G/O/2002/00471. The
1592: work of AM and RR is supported by the Bundesministerium f\"ur Bildung
1593: und Forschung (BMBF, Bonn, Germany) under the contract number
1594: 05HT4WWA2.
1595:
1596:
1597:
1598:
1599: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1600: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1601:
1602: \newpage
1603: \def\theequation{\Alph{section}.\arabic{equation}}
1604: \begin{appendix}
1605:
1606: \setcounter{equation}{0}
1607: \section{Feynman Rules}\label{rules}
1608:
1609: The two-point functions for the KK mass eigenstates, deduced from the
1610: classical part of the 5D Lagrangian~(\ref{LagYMBKT}), i.e.~without the
1611: terms $\Lag_{\rm 5D\GF}$ and $\Lag_{\rm 5D\FP}$, are given by:
1612: \begin{align*}
1613: \BB{a \; \mu}{b \; \nu}{(n) \; p}
1614: &\parbox{130mm}{
1615: \begin{equation}
1616: \label{twopoint}
1617: \Gamma^{a b}_{\mu \nu} (p_{(n)})\ =\ i \delta^{ab} \big[
1618: -g_{\mu \nu}\, \big( p^2 - m_n^2\big)\: +\:
1619: p_{\mu} p_{\nu}\, \big]\; ,
1620: \end{equation}}\\
1621: \BS{a \; \mu}{b}{(n) \; p}
1622: &\parbox{130mm}{
1623: \begin{equation}
1624: \label{mixing2}
1625: \Gamma^{a b}_{\mu 5} (p_{(n)})\ =\ \delta^{ab} m_n p_{\mu}\; ,
1626: \end{equation}}\\
1627: \SSLL{a}{b}{(n) \; p}
1628: &\parbox{130mm}{
1629: \begin{equation}
1630: \Gamma^{a b}_{5 5} (p_{(n)})\ =\ i \delta^{ab} p^2\; .
1631: \end{equation}}\\
1632: \end{align*}
1633: The two-point function~(\ref{mixing2}) refers to the non-vanishing
1634: mixing between vector and scalar modes that occurs in the absence of
1635: the gauge-fixing term $\Lag_{\rm 5D\GF}$ discussed in
1636: Section~\ref{wardCYMBKT}. In the presence of $\Lag_{\rm 5D\GF}$,
1637: however, the two-point functions take on the form:
1638: \begin{align*}
1639: \BB{a \; \mu}{b \; \nu}{(n) \; p}
1640: &\parbox{130mm}{
1641: \begin{equation}
1642: \label{twopointGF}
1643: \Gamma^{ab}_{\mu \nu} (p_{(n)},q_{(n)})\ =\ i \delta^{ab} \big[
1644: -g_{\mu \nu}\, ( p^2 - m_n^2)\: +\:
1645: \big( 1 - \mbox{$\frac{1}{\xi}$}\big)\, p_{\mu} p_{\nu}\,\big]\; ,
1646: \end{equation}}\\
1647: \SSLL{a \; \mu}{b}{(n) \; p}
1648: &\parbox{130mm}{
1649: \begin{equation}
1650: \Gamma^{ab}_{55} (p_{(n)},q_{(n)})\ =\ i \delta^{ab} \big( p^2\: -\: \xi
1651: m_n^2 \big)\; ,
1652: \end{equation}}\\
1653: \GG{a}{b}{(n) \; p}
1654: &\parbox{130mm}{
1655: \begin{equation}
1656: \Gamma^{ab}_{\bar{c} c} (p_{(n)},q_{(n)})\ =\ -i \delta^{ab} \big( p^2\: -\:
1657: \xi m_n^2\, \big)\; .
1658: \end{equation}}
1659: \end{align*}
1660: Correspondingly, the propagators for the vector, scalar and ghost KK
1661: modes read:
1662: \begin{align}
1663: \label{propagator}
1664: D^{ab}_{\mu \nu} (p_{(n)})\ &=\ \frac{i
1665: \delta^{ab}}{p^2-m_n^2+i\epsilon}\ \bigg[ -\,g_{\mu \nu}\: +\: (1- \xi)\,
1666: \frac{p_{\mu} p_{\nu}}{p^2-\xi m_n^2}\, \bigg]\ ,\\
1667: D^{ab}_{55} (p_{(n)})\ &=\
1668: \frac{i \delta^{ab}}{p^2- \xi m_n^2 +i \epsilon}\ ,\\
1669: D^{ab}_{\overline{c} c} (p_{(n)})\ &=\ \frac{-i \delta^{ab}}{p^2- \xi
1670: m_n^2 +i \epsilon}\ .
1671: \end{align}
1672: Using the definitions introduced in Appendix~\ref{multiplication} for
1673: the effective gauge-coupling coefficients $\Delta_{n,m,l}$,
1674: $\tilde{\Delta}_{n,m,l}$ etc., the Feynman rules for the interactions
1675: of the KK mass eigenstates are given by
1676: \begin{align*}
1677: \BBBeast{a \; \mu}{b \; \nu}{c \; \rho}{(n) \; k}{(m) \; p}{(l) \; q}
1678: &\parbox{130mm}{
1679: \begin{equation}
1680: \begin{split}
1681: \label{triple}
1682: \Gamma^{abc}_{\mu \nu \rho} &(k_{(n)},p_{(m)},q_{(l)})\ =\ \\ &g f^{abc}
1683: \sqrt{2}^{\; -1-\delta_{n,0}-\delta_{m,0}-\delta_{l,0}} \,
1684: \Delta_{n,m,l}\\ &\times \big[g_{\mu \nu}(k-p)_{\rho}+g_{\rho
1685: \mu}(q-k)_{\nu}+g_{\nu \rho}(p-q)_{\mu} \big]\; ,
1686: \end{split}
1687: \end{equation}}\\ &\phantom{}\\
1688: \BBSeast{b \; \mu}{c \; \nu}{a}{(m) \; p}{(l) \; q}{(n) \; k}
1689: &\parbox{130mm}{
1690: \begin{equation}
1691: \begin{split}
1692: \Gamma^{abc}_{5 \mu \nu} &(k_{(n)},p_{(m)},q_{(l)})\ =\ i g f^{abc} \,
1693: g_{\mu \nu}\\ & \times \big[ m_l \sqrt{2}^{\; -1-\delta_{m,0}}
1694: \tilde{\Delta}_{n,m,l} - m_m \sqrt{2}^{\; -1-\delta_{l,0}}
1695: \tilde{\Delta}_{n,l,m} \big]\; ,
1696: \end{split}
1697: \end{equation}}\\ &\phantom{}\\
1698: \SSBeast{b}{c}{a \; \mu}{(m) \; p}{(l) \; q}{(n) \; k}
1699: &\parbox{130mm}{
1700: \begin{equation}
1701: \Gamma^{abc}_{\mu 5 5} (k_{(n)},p_{(m)},q_{(l)})\ =\ g f^{abc}
1702: \sqrt{2}^{\; -1-\delta_{n,0}} \tilde{\Delta}_{l,n,m} (q-p)_{\mu}\; ,
1703: \end{equation}}\\
1704: \end{align*}
1705: \begin{align*}
1706: \BBBB{a \; \mu}{b \; \nu}{c \; \rho}{d \; \sigma}{(n) \; k}{(m) \;
1707: p}{(l) \; q}{(k) \; r} &\parbox{130mm}{
1708: \begin{equation}
1709: \begin{split}
1710: \Gamma^{abcd}_{\mu \nu \rho \sigma} (&k_{(n)},p_{(m)},q_{(l)},r_{(k)})
1711: \ =\ \\ i g^2 &\Delta_{n, m, l, k} \sqrt{2}^{\;
1712: -2-\delta_{n,0}-\delta_{m,0}-\delta_{l,0}-\delta_{k,0}}\\ \times
1713: \big[ &f^{abe} f^{cde} (g_{\mu \sigma} g_{\nu \rho} - g_{\mu \rho}
1714: g_{\nu \sigma}) +\\ &f^{ace} f^{bde} (g_{\mu \sigma} g_{\nu \rho} -
1715: g_{\mu \nu} g_{\rho \sigma})+\\ &f^{ade} f^{bce} (g_{\mu \rho} g_{\nu
1716: \sigma} - g_{\mu \nu} g_{\rho \sigma}) \big]\; ,
1717: \end{split}\\
1718: \end{equation}}\\
1719: \BBSS{a \; \mu}{b \; \nu}{c}{d}{(n) \; k}{(m) \; p}{(l) \; q}{(k) \; r}
1720: &\parbox{130mm}{
1721: \begin{equation}
1722: \begin{split}
1723: \label{quartic}
1724: \Gamma^{abcd}_{\mu \nu 5 5} &(k_{(n)},p_{(m)},q_{(l)},r_{(k)})\ =\ i g^2
1725: \sqrt{2}^{\; -2 -\delta_{n,0}-\delta_{m,0}}\\ &\times
1726: \tilde{\Delta}_{n,m,l,k} \, g_{\mu \nu} \big[ f^{ace} f^{bde} +
1727: f^{ade} f^{bce} \big]\; ,
1728: \end{split}
1729: \end{equation}}
1730: \end{align*}
1731: \begin{align*}
1732: \GGBeast{a}{b}{c \; \mu}{(n) \; k}{(m) \; p}{(l) \; q}
1733: &\parbox{130mm}{
1734: \begin{equation}
1735: \begin{split}
1736: \Gamma^{abc}_{\overline{c} c \mu} &(k_{(n)}, p_{(m)}, q_{(l)})\ =\ \\ &-g
1737: f^{abc} \sqrt{2}^{\; -1-\delta_{n,0}-\delta_{m,0}-\delta_{l,0}} \,
1738: \Delta_{n,m,l} \, k^{\mu}\; ,
1739: \end{split}
1740: \end{equation}}\\ &\phantom{}\\
1741: \GGSeast{a}{b}{c}{(n) \; k}{(m) \; p}{(l) \; q}
1742: &\parbox{130mm}{
1743: \begin{equation}
1744: \begin{split}
1745: \label{ghostScalar}
1746: \Gamma^{abc}_{\overline{c} c 5} &(k_{(n)}, p_{(m)}, q_{(l)})\ =\ \\ &i g
1747: f^{abc} \xi m_l \sqrt{2}^{\; -1-\delta_{n,0}-\delta_{m,0}} \,
1748: \Delta_{n,m,l}\; .
1749: \end{split}
1750: \end{equation}}
1751: \end{align*}
1752: \vspace{1cm}
1753:
1754: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1755:
1756: \setcounter{equation}{0}
1757: \section{Mass Eigenmode Wavefunctions}
1758: \label{masseigenmodeexpansion}
1759:
1760: Our aim here is to determine the analytic forms of the orthonormal
1761: wavefunctions $f_n (y)$ and $g_n (y)$, which are used
1762: in~(\ref{AmuExpansion}) to express the 5D fields $A^a_\mu (x,y)$ and
1763: $A^a_5 (x,y)$ in terms of the KK mass eigenmodes $A^a_{(n)\mu} (x)$
1764: and $A^a_{(n)5} (x)$. In addition, we will derive the transcendental
1765: equation~(\ref{spectrum}) that determines the masses for the KK gauge
1766: fields.
1767:
1768: We start our discussion with the observation that the gauge-invariant
1769: part of the 5D QCD Lagrangian~(\ref{LagYMBKT}) is proportional to $[ 1
1770: + r_c \delta (y) ]$. Then, our approach to quantization in the $R_\xi$
1771: gauge can be consistently formulated if this $\delta(y)$-dependent
1772: factor is promoted to an overall integral weight of the full quantized
1773: action that includes the gauge-fixing and Faddeev--Popov ghost terms
1774: given in~(\ref{gfTerm}). In such an approach, it is important to
1775: require that total derivatives of fields or product of fields with
1776: respect to $y$ do not contribute to the action, but vanish
1777: identically. This condition on the surface terms can be achieved by
1778: requiring that
1779: \begin{equation}
1780: \label{surfcond}
1781: \int_{-\pi R+\epsilon}^{\pi R +\epsilon} dy\, \big[ 1 + r_c \delta
1782: (y) \big]\; \partial_5 f_n (y)\ =\ 0\,,\qquad
1783: \int_{-\pi R+\epsilon}^{\pi R + \epsilon} dy\, \big[ 1 + r_c \delta
1784: (y) \big]\; \partial_5 g_n (y)\ =\ 0\, .
1785: \end{equation}
1786: Bear in mind that our compact space $y$ is defined in the interval
1787: $(-\pi R +\epsilon,\ \pi R + \epsilon ]$, where $\epsilon$ is an
1788: infinitesimal positive constant which is taken to zero at the end of
1789: the calculation. Using the convolution methods presented in
1790: Appendix~C, it is not difficult to show that (\ref{surfcond}) is
1791: sufficient to ensure the vanishing of any total derivative when
1792: integrated along with the $\delta (y)$-dependent weight $[ 1 + r_c
1793: \delta (y) ]$. As a consequence of this, the undesirable mixing term
1794: $(\partial_\mu A_5)\,(\partial_5 A^\mu )$ in the gauge-kinetic part of
1795: the 5D Lagrangian cancels against a corresponding term $(\partial_5
1796: A_5)\, (\partial_\mu A^\mu)$ that occurs in the gauge-fixing
1797: Lagrangian~(\ref{gfTerm}).
1798:
1799: In addition to~(\ref{surfcond}), the wavefunctions $f_n (y)$ and $g_n
1800: (y)$ have to satisfy the orthonormality conditions:
1801: \begin{equation}
1802: \label{ortho}
1803: \begin{split}
1804: \int_{-\pi R+\epsilon}^{\pi R+\epsilon} dy \;
1805: \big[\, 1\: +\: r_c \delta(y)\,\big] &\,
1806: f_k(y) f_l(y)\ =\ \delta_{k,l}\; ,\\[3mm]
1807: \int_{-\pi R+\epsilon}^{\pi R+\epsilon} dy \;
1808: \big[\, 1\: +\: r_c \delta(y)\,\big] &\,
1809: g_k(y) g_l(y)\ =\ \delta_{k,l}\; .
1810: \end{split}
1811: \end{equation}
1812: The analytic form of $f_n (y)$ and $g_n (y)$ can be fully specified by
1813: requiring that the kinetic part of the effective 4D Lagrangian for the
1814: KK modes $A^{a}_{(n)\mu}$ and $A^a_{(n) 5}$ takes on the expected form
1815: in the conventional $R_\xi$ gauge:
1816: \begin{eqnarray}
1817: \label{4DRxi}
1818: {\cal L}^{\rm eff}_{\rm kin} & = & -\, \frac{1}{4}\;
1819: \big( \partial_\mu A^{a}_{(n)\,\nu} - \partial_\nu
1820: A^{a}_{(n)\,\mu}\big)\, \big( \partial^\mu A^{a\,\nu}_{(n)} -
1821: \partial^\nu A^{a\,\mu}_{(n)}\big)\: -\: \frac{1}{2\xi}\,
1822: \big(\partial_\mu A^{a\,\mu}_{(n)}\big)^2\: +\: \frac{1}{2}\, m^2_n
1823: A^{a}_{(n)\,\mu}\,A^{a\,\mu}_{(n)} \nonumber\\
1824: &&+\, \frac{1}{2}\, \big(\partial_\mu A^{a}_{(n)\,5}\big)\,
1825: \big(\partial^\mu A^{a}_{(n)\,5}\big)\: -\: \frac{\xi}{2}\, m^2_n\,
1826: \big(A^{a}_{(n)\,5}\big)^2\ .
1827: \end{eqnarray}
1828: The above form of the Lagrangian~(\ref{4DRxi}) is obtained if $f_n
1829: (y)$ and $g_n (y)$ satisfy the simple wave equations
1830: \begin{eqnarray}
1831: \label{fnWaveEquation}
1832: \partial^2_5\, f_n(y)\ +\ m_n^2\, f_n(y) &=& 0\; ,\\
1833: \label{gnWaveEquation}
1834: \partial^2_5\, g_n(y)\ +\ m_n^2\, g_n(y) &=& 0\; .
1835: \end{eqnarray}
1836: To find the solutions to the above wavefunction equations, we should
1837: consistently implement the constraints~(\ref{surfcond})
1838: and~(\ref{ortho}). Notice that our approach is different from the one
1839: presented in the existing literature~\cite{Santiago}.
1840:
1841: Let us now describe how to find the analytic form of the wavefunctions
1842: $f_n (y)$. Because of the presence of the function $\delta (y)$
1843: through the constraints, our starting point is the piecewise ansatz
1844: \begin{equation}
1845: \label{fnsol}
1846: f_n(y)\ =\
1847: \begin{cases}
1848: \ A_{\I} \sin m_n y\ +\ B_{\I} \cos m_n y\,, \qquad &\textrm{for}
1849: \quad -\pi R +\epsilon\, <\, y\, \leq\, -\,\epsilon\\
1850: \ A_{\II} \sin m_n y\ +\ B_{\II} \cos m_n y\,, &\textrm{for}
1851: \qquad \quad \epsilon\, \leq\, y\, \leq\, \pi R+\epsilon \; .
1852: \end{cases}
1853: \end{equation}
1854: Obviously, our regularized ansatz~(\ref{fnsol}) is a solution
1855: to~(\ref{fnWaveEquation}), which excludes an infinitesimal interval
1856: $(-\epsilon, \epsilon)$ in the neighbourhood of the singular point
1857: $y=0$. The difference between the wavefunctions $f_n (y)$ and $g_n
1858: (y)$ is that the $f_n (y)$ is an even function of $y$, whereas $g_n
1859: (y)$ is an odd one. Moreover, we must assume that both $f_n (y)$ and
1860: $g_n (y)$ are periodic functions of $y$, i.e.~$f_n (y ) = f_n (y +
1861: 2\pi R)$ and $g_n (y) = g_n (y + 2\pi R)$. Within their definition
1862: interval $(-\pi R+\epsilon, \pi R +\epsilon]$, the periodic condition
1863: for the points $y = \pm \pi R$ amounts to the constraints:
1864: \begin{equation}
1865: \label{periodic}
1866: f_n (-\pi R + \epsilon)\ =\ f_n (\pi R +\epsilon)\,,\qquad
1867: g_n (-\pi R + \epsilon)\ =\ g_n (\pi R +\epsilon)\,.
1868: \end{equation}
1869: Although the above periodicity constraint on $f_n (y)$ may be trivial
1870: because $f_n (y)$ is an even function, the one applied to $y$-odd
1871: functions, such as $\partial_5 f_n (y)$ and $g_n (y)$, can provide
1872: useful information (see discussion below).
1873:
1874: The unknown coefficients $A_{\rm I,II}$ and $B_{\rm I,II}$, as well as
1875: the KK masses $m_n$, may be determined as follows. {}From the
1876: requirement that $f_n (y)$ is an even function, i.e.~$f_n (y) = f_n
1877: (-y)$, we get the constraint
1878: \begin{equation}
1879: \label{AB}
1880: A_{\I}\ =\ - A_{\II}\ =\ A_n\,,\qquad B_{\I}\ =\ B_{\II}\ =\ B_n\; .
1881: \end{equation}
1882: Imposing the first condition of~(\ref{surfcond}) on the $y$-odd
1883: expression $\partial_5 f_n (y)$ leads to the constraint:
1884: \begin{equation}
1885: \label{formal1}
1886: \partial_5 f_n (y=0)\ =\ 0\; .
1887: \end{equation}
1888: To implement~(\ref{formal1}) in agreement with the wave
1889: equation~(\ref{fnWaveEquation}) for $f_n (y)$, we analytically define
1890: $f_n(y)$ in the interval $-\epsilon < y < \epsilon$ as follows:
1891: \begin{equation}
1892: \label{interfn}
1893: f_n (y) \ =\ R_n\, \cos m_n y\; .
1894: \end{equation}
1895: The arbitrary constant $R_n$ may be determined by the matching
1896: condition
1897: \begin{equation}
1898: \label{matching1}
1899: R_n\ =\ \lim_{\epsilon\to 0}\;
1900: \bigg(\, \lim_{y \to \epsilon^+} f_n (y)\,\bigg)\ =\
1901: \lim_{\epsilon\to 0}\; \bigg(\, \lim_{y \to -\epsilon^-} f_n (y)\,\bigg)\; .
1902: \end{equation}
1903: Notice that the above double limits are well-defined, even though $f_n
1904: (y)$ has finite jumps at the critical points $y = \pm \epsilon$. The
1905: derivative $\partial_5 f_n (y)$ of the $\epsilon$-regularized function
1906: $f_n (y)$ is also analytically well-behaved at the singular point $y =
1907: 0$, where $\delta (y)$ acts, and complies with the
1908: constraint~(\ref{formal1}). In Fig.~\ref{interpol} we display the
1909: $y$-profile of the $\epsilon$-regularized function $f_{n=2} (y)$, and
1910: show its shape in the limit $\epsilon \to 0$. Here, we should stress
1911: again that our regularization method consists in taking the limit
1912: $\epsilon \to 0$ at the very end of our calculation.
1913:
1914: \begin{figure}[h!]
1915: \begin{center}
1916: \parbox{120mm}{
1917: \begin{center}
1918: \includegraphics[width=0.6\textwidth]{fn.eps}\\
1919: \includegraphics[width=0.6\textwidth]{gn.eps}
1920: \vspace{-1cm}
1921: \end{center}}
1922: \end{center}
1923: \caption{\em The $y$-profile of the $\epsilon$-regularized mass
1924: eigenmode wavefunctions $f_{n=2} (y)$ and $g_{n=2} (y)$. The dotted
1925: lines indicate the $y$-dependence of the wavefunctions in the limit
1926: $\epsilon \to 0$, in the neighbourhood of the singular point $y=0$
1927: [cf.\ (\ref{fn}) and (\ref{gn})].}
1928: \label{interpol}
1929: \end{figure}
1930:
1931: As we will see below, $\partial_5 f_n (y)$ is an odd function that
1932: shares the same properties as $g_n (y)$. We may understand this fact
1933: by observing that if the even function $f_n (y)$ is a solution to the
1934: wave equation~(\ref{fnWaveEquation}), its derivative $\partial_5 f_n
1935: (y)$, which is an odd function, is a solution
1936: to~(\ref{fnWaveEquation}) as well. Thus, assuming {\it \`a
1937: posteriori} that, up to a constant, the odd function $\partial_5 f_n
1938: (y)$ equals $g_n (y)$, we may impose the second condition
1939: of~(\ref{surfcond}) in the form
1940: \begin{equation}
1941: \label{2ndcond}
1942: \int_{-\pi R+\epsilon}^{\pi R+\epsilon} dy \;
1943: \big[\, 1\: +\: r_c \delta(y)\,\big]\; \partial_5\,
1944: \big(\partial_5 f_n (y)\big)\ =\ 0\; .
1945: \end{equation}
1946: Using the wave equation~(\ref{fnWaveEquation}) for the function $f_n$,
1947: we get the new constraint:
1948: \begin{equation}
1949: \label{3rdcond}
1950: \partial_5 f_n (\pi R +\epsilon)\: -\: \partial_5 f_n (-\pi R +\epsilon)\:
1951: -\: \partial_5 f_n (\epsilon )\: +\: \partial_5 f_n (-\epsilon )\ =\
1952: m_n^2\, r_c\, f_n (0)\; .
1953: \end{equation}
1954: In deriving~(\ref{3rdcond}), we have neglected terms
1955: $\ord{(\epsilon)}$ that occur in the region $-\epsilon < y <
1956: \epsilon$.
1957:
1958: Our ultimate constraint arises from the fact that in addition to $f_n
1959: (y)$ which is periodic by construction, $\partial_5 f_n (y)$ has also
1960: to be periodic on the entire $y$ interval, i.e.~$\partial_5 f_n (y ) =
1961: \partial_5 f_n (y + 2\pi R)$. Applying the periodicity restriction for
1962: the points $y = \mp\pi R +\epsilon$ that lie within the definition
1963: interval, we get the non-trivial relation
1964: \begin{equation}
1965: \label{periodic2}
1966: \partial_5\, f_n (-\pi R + \epsilon) \ =\ \partial_5\, f_n (\pi R +
1967: \epsilon )\; .
1968: \end{equation}
1969: The constraints~(\ref{3rdcond}) and (\ref{periodic2}) lead to a
1970: determination of the coefficients $A_n$ and $B_n$ and the KK masses
1971: $m_n$:
1972: \begin{equation}
1973: \label{AnBn}
1974: A_n\ =\ \frac{m_n r_c}{2}\ B_n\; ,\qquad \sin m_n \pi R\: +\:
1975: \frac{m_n r_c}{2}\, \cos m_n \pi R\ =\ 0\; .
1976: \end{equation}
1977: The second equation in~(\ref{AnBn}) is equivalent to the
1978: transcendental equation~(\ref{spectrum}). The remaining freedom is an
1979: overall normalization constant $N_n$ given in~(\ref{fn}) and can be
1980: determined from the orthonormality condition~(\ref{ortho}).
1981:
1982: To determine the analytic form of $g_n (y)$, we proceed very
1983: analogously. We start with a similar ansatz
1984: \begin{equation}
1985: \label{gnsol}
1986: g_n(y)\ =\
1987: \begin{cases}
1988: \ C_{\I} \sin m_n y\: +\: D_{\I} \cos m_n y\;, \qquad &\textrm{for}
1989: \quad -\pi R +\epsilon\, <\, y \, \leq\, -\epsilon\\
1990: \ C_{\II} \sin m_n y\: +\: D_{\II} \cos m_n y\;, &\textrm{for}
1991: \qquad \quad \epsilon\, \leq\, y\, \leq\, \pi R +\epsilon\; .
1992: \end{cases}
1993: \end{equation}
1994: {}From the fact that $g_n (y)$ is an odd function, we get
1995: \begin{equation}
1996: \label{CD}
1997: C_{\I}\ =\ C_{\II}\ =\ C_n\,,\qquad D_{\I}\ =\ - D_{\II}\ =\ D_n\; .
1998: \end{equation}
1999: In view of the constraints~(\ref{CD}), $g_n (y)$ exhibits a
2000: discontinuity in the limit $y=\pm \epsilon \to 0$. Exactly as we did
2001: above for $f_n (y)$, we analytically regularize $g_n (y)$ in the
2002: interval $(-\epsilon, \epsilon)$ as
2003: \begin{equation}
2004: \label{intergn}
2005: g_n (y)\ =\ Q_n\, \sin m_n y\; .
2006: \end{equation}
2007: Observe that (\ref{intergn}) is also a solution to the wave
2008: equation~(\ref{gnWaveEquation}) for the function $g_n (y)$. Since
2009: $\partial_5 g_n (y)$ determined from~(\ref{gnsol}) has well-defined
2010: limits when $y=\pm \epsilon \to 0$, we apply the matching condition
2011: \begin{equation}
2012: \label{matching2}
2013: m_n\,Q_n\ =\ \lim_{\epsilon\to 0}\;
2014: \bigg(\, \lim_{y \to \epsilon^+}\, \partial_5 g_n (y)\,\bigg)\ =\
2015: \lim_{\epsilon\to 0}\; \bigg(\, \lim_{y \to -\epsilon^-}\,
2016: \partial_5 g_n (y)\,\bigg)\;
2017: \end{equation}
2018: to determine the constant $Q_n$. Given~(\ref{intergn}), the second
2019: equation of the condition~(\ref{surfcond}) implies that
2020: \begin{equation}
2021: \label{4thcond}
2022: g_n (\pi R +\epsilon)\: -\: g_n (-\pi R +\epsilon)\:
2023: -\: g_n (\epsilon )\: +\: g_n (-\epsilon )\ =\
2024: -\,r_c\, \partial_5 g_n (y=0)\; .
2025: \end{equation}
2026: In writing down~(\ref{4thcond}), we also made use of the fact that
2027: terms $\ord{(\epsilon )}$ arising from the integration in the interval
2028: $(-\epsilon, \epsilon)$ have been neglected. For illustration, we
2029: show in Fig.~\ref{interpol} the $y$-dependence of the
2030: $\epsilon$-regularized wavefunction $g_{n=2} (y)$, along with its
2031: $y$-profile in the limit $\epsilon \to 0$.
2032:
2033:
2034: Finally, taking into account the periodic condition~(\ref{periodic})
2035: for the functions $g_n (y)$, we obtain the relation
2036: \begin{equation}
2037: D_n\ =\ -\; \frac{ m_n r_c}{2}\ C_n\; ,
2038: \end{equation}
2039: and the second equation in~(\ref{AnBn}). As before, the overall
2040: normalization constant can be determined from the orthonormality
2041: condition~(\ref{ortho}). Finally, implementing the matching
2042: conditions (\ref{matching1}) and (\ref{matching2}), we find that $R_n
2043: = Q_n$.
2044:
2045: After having derived the analytic forms of $f_n (y)$ and $g_n (y)$
2046: given in (\ref{fn}) and (\ref{gn}), respectively, it is not difficult
2047: to see that $f_n (y)$ and $g_n (y)$ satisfy the relations
2048: in~(\ref{partial}), for the entire piecewise defined region $(-\pi R
2049: +\epsilon, \pi R + \epsilon ]$ that also includes the singular point
2050: $y=0$. Finally, we should remark that for a 5D orbifold theory
2051: without BKTs, the functions $f_n (y)$ and $g_n (y)$ go to the standard
2052: Fourier expansion modes in limit $r_c \to 0$:
2053: \begin{equation}
2054: \label{lim}
2055: \lim_{r_c \to 0}\, f_n(y)\ =\ \frac{1}{\sqrt{ 2^{\delta_{n,0}} \pi R}} \;
2056: \cos \frac{ny}{R}\ , \qquad \lim_{r_c \to 0}\, g_n(y)\ =\
2057: \frac{1}{\sqrt{\pi R}} \; \sin \frac{ny}{R}\ .
2058: \end{equation}
2059:
2060: For completeness, we present in the remainder of this appendix
2061: analytic results concerning the case of a 5D orbifold theory with two
2062: BKTs at the orbifold fixed points $y = 0$ and $y = \pi R$. We follow
2063: a very analogous approach to the one given above for the one BKT case.
2064: We start again with the same form of regularized ansaetze
2065: (\ref{fnsol}) and (\ref{gnsol}) for $f_n (y)$ and $g_n (y)$, which are
2066: piecewise defined in the interval $(-\pi R + \epsilon, -\epsilon]\
2067: \cup\ [\epsilon , \pi R -\epsilon]$. Notice that the $y$-intervals,
2068: $(-\epsilon,\epsilon)$ and $(\pi R -\epsilon, \pi R + \epsilon ]$, are
2069: excluded, because of the singular overall integral measure: $\big[\, 1
2070: + r_c\,\delta (y) + r_c \delta (y - \pi R)\,\big]$. However, most of
2071: the discussion given above goes through, with some obvious
2072: modifications related to the presence of the second $\delta$-function,
2073: $\delta (y - \pi R)$. In fact, the only crucial difference is that
2074: the constraints~(\ref{3rdcond}) and~(\ref{4thcond}) now become
2075: \begin{eqnarray}
2076: \label{2BKTcond1}
2077: \partial_5 f_n (\pi R -\epsilon)\: -\: \partial_5 f_n (-\pi R +\epsilon)\:
2078: -\: \partial_5 f_n (\epsilon )\: +\: \partial_5 f_n (-\epsilon )& =&
2079: m_n^2\, r_c\, \big[\, f_n (0)\: +\: f_n (\pi R)\,\big]\; ,\nonumber\\
2080: && \\[3mm]
2081: \label{2BKTcond2}
2082: g_n (\pi R -\epsilon)\: -\: g_n (-\pi R +\epsilon)\:
2083: -\: g_n (\epsilon )\: +\: g_n (-\epsilon ) &=&
2084: -\,r_c\, \big[\, \partial_5 g_n (0)\: +\: \partial_5 g_n (\pi
2085: R)\,\big]\; .\nonumber\\
2086: &&
2087: \end{eqnarray}
2088: Taking into consideration the above non-trivial modifications, we find
2089: that the mass eigenmode wavefunctions $f_n(y)$ and $g_n (y)$ retain
2090: the functional form given in (\ref{fn}) and (\ref{gn}), in the limit
2091: $\epsilon \to 0$, for the one BKT case. However, the normalization
2092: factor $N_n$ modifies in the presence of the second BKT at $y = \pi R$.
2093: The analytic form of $N_n$ may be determined by
2094: \begin{equation}
2095: \label{Nn2BKT}
2096: \begin{split}
2097: N_n^{-2}\ =& \ 1\: +\: 4\, \tilde{r}_c\:
2098: +\: \frac{1+2 \tilde{r}_c}{(1-\tilde{r}_c \pi R m_n)^2}\:
2099: +\: \frac{1+2 \tilde{r}_c}{(1+\tilde{r}_c \pi R m_n)^2}\\
2100: &-\ \frac{1+3\tilde{r}_c}{1-\tilde{r}_c \pi R m_n}\:
2101: -\: \frac{1+3\tilde{r}_c}{1+\tilde{r}_c \pi R m_n}\ .
2102: \end{split}
2103: \end{equation}
2104: Moreover, the transcendental equation determining the mass spectrum of
2105: the effective 4D theory may be cast into the useful factorizable form:
2106: \begin{equation}
2107: \label{spectrum2BKT}
2108: m_n\, \bigg[\, \tan \bigg(\,\frac{m_n \pi R}{2}\,\bigg)\ +\
2109: \frac{m_n r_c}{2}\, \bigg]\,
2110: \bigg[\, \cot \bigg(\,\frac{m_n \pi R}{2}\,\bigg)\ -\
2111: \frac{m_n r_c}{2}\, \bigg]\ =\ 0\; ,
2112: \end{equation}
2113: which is valid for $r_c \neq 0$ and, after some algebra, agrees
2114: with~\cite{Carena}. In Fig.~\ref{spectra}, we present typical
2115: solutions to (\ref{spectrum}) and (\ref{spectrum2BKT}) for a 5D
2116: orbifold theory with one and two BKTs, respectively.
2117:
2118: \begin{figure}[h!]
2119: \begin{center}
2120: \parbox{120mm}{
2121: \begin{center}
2122: \includegraphics[width=0.7\textwidth]{spectra.eps}
2123: \end{center}}
2124: \end{center}
2125: \caption{{\em Mass spectrum $m_n$ of the effective 4D theory with
2126: one/two BKT in the 5D Lagrangian, (\ref{spectrum}) and
2127: (\ref{spectrum2BKT}), respectively.}} \label{spectra}
2128: \end{figure}
2129:
2130:
2131: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2132:
2133:
2134: \newpage
2135: \section{Convolution Integrals for $S^1/\Zb_2$ Wavefunctions}
2136: \label{multiplication}
2137: \setcounter{equation}{0}
2138:
2139: In the process of the KK reduction, expressions that involve products
2140: of 5D fields compactified on an $S^1/\Zb_2$ orbifold, such as $A^a_\mu
2141: (x,y)$ and $A^a_5 (x,y)$, need to be integrated over the compact
2142: dimension $y$. These 5D fields have an even or odd parity under
2143: $\Zb_2$ transformations and can be expressed in terms of the KK mass
2144: eigenmodes $f_n (y)$ and $g_n (y)$, given in (\ref{fn}) and
2145: (\ref{gn}), respectively. Using convolution integral techniques, we
2146: may express the product of two $\Zb_2$-even functions, e.g.~$F(y)$ and
2147: $G(y)$, in a series of the orthonormal wavefunctions $f_n (y)$ as
2148: follows:
2149: \begin{equation}
2150: F(y)\,G(y)\ =\ \sum_{n=0}^{\infty}\, [F*G]_{(n)}\, f_n(y) \, ,
2151: \end{equation}
2152: where the coefficients $[F*G]_{(n)}$ are given by the convolution
2153: integrals
2154: \begin{equation}
2155: \label{FGn}
2156: [F*G]_{(n)}\ =\ \sum_{k,l=0}^{\infty} F_{(k)} G_{(l)}
2157: \int_{-\pi R}^{\pi R} dy \, \big[ 1 + r_c \delta(y) \big]\,
2158: f_k(y) f_l(y) f_n(y)\; .
2159: \end{equation}
2160: Likewise, the product of a $\Zb_2$-even function $F(y)$ with a
2161: $\Zb_2$-odd one $H(y)$ can be expanded in a series of the orthonormal
2162: wavefunctions $g_n (y)$:
2163: \begin{equation}
2164: F(y)\,H(y)\ =\ \sum_{n=0}^{\infty}\, [F*H]_{(n)}\, g_n(y) \, ,
2165: \end{equation}
2166: where the coefficients $[F*H]_{(n)}$ are given by the convolution
2167: integrals
2168: \begin{equation}
2169: \label{FHn}
2170: [F*G]_{(n)}\ =\ \sum_{k,l=0}^{\infty} F_{(k)} H_{(l)}
2171: \int_{-\pi R}^{\pi R} dy \, \big[ 1 + r_c \delta(y) \big]\,
2172: f_k(y) g_l(y) g_n(y)\; .
2173: \end{equation}
2174:
2175:
2176: The integrals~(\ref{FGn}) and (\ref{FHn}) appear in the calculation of
2177: the effective Feynman rules, and of the effective gauge and BRS
2178: transformations, when the 5D fields are expressed in terms of their KK
2179: modes. Specifically, defining the overlap integrals as
2180: \begin{equation}
2181: \begin{split}
2182: \int_{-\pi R}^{\pi R} dy \, \big[ 1 + r_c \delta(y) \big] \,
2183: f_k(y) \, f_l(y) \, f_n(y)
2184: \ =&\ \frac{\Delta_{k,l,n}}{2
2185: \sqrt{2^{\delta_{k,0}+\delta_{l,0}+\delta_{n,0}} \pi R}} \ , \\
2186: \int_{-\pi R}^{\pi R} dy \, \big[ 1 + r_c \delta(y) \big] \,
2187: f_k(y) \, g_l(y) \, f_n(y)
2188: \ = &\ \frac{\tilde{\Delta}_{k,l,n}}{2
2189: \sqrt{2^{\delta_{l,0}} \pi R}} \ ,
2190: \end{split}
2191: \end{equation}
2192: we may analytically calculate the coefficients $\Delta_{k,l,n}$ and
2193: $\tilde{\Delta}_{k,l,n}$ as follows:
2194: \begin{equation}
2195: \label{DeltaDownThree}
2196: \begin{split}
2197: \Delta_{k,l,n}\ & = \
2198: \Delta(m_k, m_l, m_n)\: +\: \Delta(-m_k, m_l, m_n)\:
2199: +\: \Delta(m_k, -m_l, m_n)\: +\: \Delta(m_k, m_l, -m_n) \, ,\\[1mm]
2200: \tilde{\Delta}_{k,n,l}\ &=\ -\,\Delta(m_k, m_l, m_n) \:
2201: -\: \Delta(m_k, m_l, -m_n)\: +\: \Delta(-m_k, m_l, m_n) \:
2202: +\: \Delta(m_k, -m_l, m_n) \, ,
2203: \end{split}
2204: \end{equation}
2205: where
2206: \begin{equation}
2207: \label{DeltaThree}
2208: \Delta(m_k,m_l,m_n)\ =\ N_k N_l N_n\, \left[\, \frac{\int_0^{\pi R} dy \,
2209: \cos (m_k+m_l+m_n)(y-\pi R)}{\pi R \, \cos m_k \pi R \, \cos m_l \pi R \,
2210: \cos m_n \pi R}\ +\ \tilde{r}_c\, \right] \, .
2211: \end{equation}
2212: The normalization constants $N_n$ are given in (\ref{Nn}) and
2213: $\tilde{r}_c = r_c/(2 \pi R)$. According to the above definition, the
2214: coefficient $\Delta_{k,l,n}$ is symmetric in all of its indices, while
2215: $\tilde{\Delta}_{k,l,n}$ is only symmetric under the interchange of
2216: the first and the last index. Note that the argument of the cosine in
2217: (\ref{DeltaThree}) may vanish, when some of the arguments in
2218: $\Delta(m_k,m_l,m_n)$ become negative. Taking this possibility into
2219: account, we find
2220: \begin{align}
2221: \label{Dmasskln}
2222: \Delta(m_k,m_l,m_n)\
2223: &= \
2224: \begin{cases}
2225: & \hspace{-3mm} N_k N_l N_n\, \pi^2 R^2 \tilde{r}_c^3 \; \frac{\displaystyle
2226: m_k m_l m_n}{\displaystyle m_k+m_l+m_n}\ ,
2227: \qquad \qquad
2228: \textrm{for} \quad m_k+m_l+m_n \neq 0 \\[2ex]
2229: \begin{split}
2230: & \hspace{-3mm} N_k N_l N_n \Big\{
2231: \Big[(1+\pi^2 R^2 \tilde{r}_c^2 m_k^2)(1+\pi^2 R^2
2232: \tilde{r}_c^2 m_l^2)(1+\pi^2 R^2 \tilde{r}_c^2
2233: m_n^2)\Big]^{\frac{1}{2}}\: +\:
2234: \tilde{r}_c \Big\},\\
2235: & \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \,
2236: \textrm{for} \quad m_k+m_l+m_n = 0\; .
2237: \end{split}
2238: \end{cases}
2239: \end{align}
2240: For non-vanishing $\tilde{r}_c$, the lower case is only important for
2241: the calculation of $\Delta(m_k,m_l,m_n)$, if at least one of the
2242: indices $k$, $l$, $n$ is zero and the other two are equal. When this
2243: condition is satisfied, we find
2244: \begin{eqnarray}
2245: \label{Dnn0}
2246: \Delta_{0,0,0} &=& \frac{4}{\sqrt{1+\tilde{r}_c}}\ ,
2247: \qquad \tilde{\Delta}_{n,n,0}\ =\ 0\; ,\nonumber\\
2248: \Delta_{n,n,0} &=& \tilde{\Delta}_{n,0,n}\ =\
2249: \frac{2 N_n^2}{\sqrt{1+\tilde{r}_c}}\,\Big(\, 1\: +\: \tilde{r}_c\:
2250: +\: \pi^2 R^2 \tilde{r}_c^2 m_n^2\, \Big)\; .
2251: \end{eqnarray}
2252: In any other case (up to symmetries), the coefficients
2253: $\Delta_{k,l,n}$ and $\tilde{\Delta}_{k,l,n}$ have been calculated to
2254: have the more compact analytic forms:
2255: \begin{equation}
2256: \begin{split}
2257: \Delta_{k, l, n}\ & =\ N_k N_l N_n \pi^2 R^2 \tilde{r}_c^3 \;
2258: \frac{8 m_k^2 m_l^2 m_n^2}{m_k^4+m_l^4+m_n^4-2(m_k^2 m_l^2 +
2259: m_l^2 m_n^2 + m_n^2 m_k^2)} \ , \\[2mm]
2260: \tilde{\Delta}_{k, n, l}\ & =\ N_k N_l N_n \pi^2 R^2 \tilde{r}_c^3 \;
2261: \frac{4 m_k m_l m_n^2(m_k^2+m_l^2-m_n^2)}{m_k^4+m_l^4+m_n^4-2(m_k^2
2262: m_l^2 + m_l^2 m_n^2 + m_n^2 m_k^2)} \ .
2263: \end{split}
2264: \end{equation}
2265: If two of the indices are equal, the last two formulae simplify
2266: further to
2267: \begin{eqnarray}
2268: \label{Dnnj}
2269: \Delta_{n,n,j} &=& N_n^2 N_j \pi^2 R^2 \tilde{r}_c^3 \;
2270: \frac{8 m_n^4}{m_j^2-4 m_n^2} \ , \nonumber\\
2271: \tilde{\Delta}_{n,n,j} &=& N_n^2 N_j \pi^2 R^2 \tilde{r}_c^3 \;
2272: \frac{4 m_j m_n^3}{m_j^2-4m_n^2} \ , \\
2273: \tilde{\Delta}_{n,j,n} &=& -\, N_n^2 N_j \pi^2 R^2 \tilde{r}_c^3
2274: \; \frac{4 m_n^2\, (m_j^2-2m_n^2)}{m_j^2-4 m_n^2} \ .\nonumber
2275: \end{eqnarray}
2276:
2277: Correspondingly, we may now evaluate overlap integrals that involve
2278: the product of 4 orthonormal wavefunctions,
2279: \begin{equation}
2280: \begin{split}
2281: \int_{-\pi R}^{\pi R} dy \, \big[ 1+r_c \delta(y) \big]\,
2282: f_k(y) f_l(y) f_n(y) f_m(y)\
2283: & =\ \frac{\Delta_{k,l,n,m}}{4 \pi R
2284: \sqrt{2^{\delta_{k,0}+\delta_{l,0}+\delta_{n,0}+\delta_{n,0}}}} \ ,\\[1ex]
2285: \int_{-\pi R}^{\pi R} dy \, \big[ 1+r_c \delta(y)\big]\,
2286: f_k(y) f_l(y) g_n(y) g_m(y)
2287: \ & =\ \frac{\tilde{\Delta}_{k,l,n,m}}{4 \pi R
2288: \sqrt{2^{\delta_{k,0}+\delta_{l,0}}}} \ .
2289: \end{split}
2290: \end{equation}
2291: The coefficients $\Delta_{k,l,n,m}$ and $\tilde{\Delta}_{k,l,n,m}$ are
2292: given by
2293: \begin{eqnarray}
2294: \label{Dklnm}
2295: \Delta_{k,l,n,m} &=& \Delta(m_k,m_l,m_n,m_m)\: +\:
2296: \Delta(-m_k,m_l,m_n,m_m)\: +\: \Delta(m_k,-m_l,m_n,m_m)\nonumber\\
2297: && +\,\Delta(m_k,m_l,-m_n,m_n)\: +\: \Delta(m_k,m_l,m_n,-m_m)\: +\:
2298: \Delta(-m_k,-m_l,m_n,m_m)\nonumber\\
2299: && +\, \Delta(-m_k,m_l,-m_n,m_m)\: +\:
2300: \Delta(-m_k,m_l,m_n,-m_m) \, , \\[2mm]
2301: \label{tDklnm}
2302: \tilde{\Delta}_{k,l,n,m} &=& -\,\Delta(m_k,m_l,m_n,m_m)\: -\:
2303: \Delta(-m_k,m_l,m_n,m_m)\: -\: \Delta(m_k,-m_l,m_n,m_m)\nonumber\\
2304: && -\, \Delta(-m_k,-m_l,m_n,m_m)\: +\: \Delta(m_k,m_l,-m_n,m_m)\: +\:
2305: \Delta(m_k,m_l,m_n,-m_m)\nonumber\\
2306: && +\, \Delta(-m_k,m_l,-m_n,m_m)\: +\: \Delta(-m_k,m_l,m_n,-m_m) \, ,
2307: \end{eqnarray}
2308: where $\Delta(m_k,m_l,m_n,m_m)$ is an obvious generalization
2309: of~(\ref{DeltaThree}), i.e.
2310: \begin{equation}
2311: \label{DeltaFour2}
2312: \begin{split}
2313: \Delta&(m_k,m_l,m_n,m_m)\ =\ N_k N_l N_n N_m\\[2mm]
2314: &\times\, \left\{
2315: \begin{array}{l}
2316: \pi^2 R^2 \tilde{r}_c^3 \
2317: \frac{\displaystyle m_l m_n m_m + m_k m_n m_m + m_k m_l m_m +
2318: m_k m_l m_n}{\displaystyle m_k+m_l+m_n+m_m}\ , \\[1ex]
2319: \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad
2320: \textrm{for} \quad m_k+m_l+m_n+m_m \neq 0\\[2ex]
2321: \Big\{ \Big[(1+\pi^2 R^2 \tilde{r}_c^2 m_k^2)
2322: (1+\pi^2 R^2 \tilde{r}_c^2 m_l^2)(1+\pi^2 R^2 \tilde{r}_c^2 m_n^2)
2323: (1+\pi^2 R^2 \tilde{r}_c^2 m_m^2) \Big]^{\frac{1}{2}}\: +\:
2324: \tilde{r}_c \Big\}\,,\\[1ex]
2325: \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad
2326: \textrm{for} \quad m_k+m_l+m_n+m_m = 0 \, \, .
2327: \end{array}
2328: \right.
2329: \end{split}
2330: \end{equation}
2331: It is easy to see that $\Delta_{k,l,n,m}$ is symmetric in all of its
2332: indices, while $\tilde{\Delta}_{k,l,n,m}$ is only symmetric in the first pair as well as the last two of its indices. If one of the 4
2333: indices is zero in $\Delta_{k,l,n,m}$ and $\tilde{\Delta}_{k,l,n,m}$,
2334: then (\ref{Dklnm}) and (\ref{tDklnm}) reduce to
2335: \begin{equation}
2336: \label{Dkln0}
2337: \Delta_{k,l,n,0}\ =\ \frac{2 \Delta_{k,l,n}}{\sqrt{1+ \tilde{r}_c}}\ ,\qquad
2338: \tilde{\Delta}_{0,k,l,n}\ =\ \frac{2 \tilde{\Delta}_{l,k,n}}{\sqrt{1+
2339: \tilde{r}_c}} \ .
2340: \end{equation}
2341: It is also useful to list analytic results for special combinations of
2342: indices, when the lower case in~(\ref{DeltaFour2}) becomes relevant:
2343: \begin{eqnarray}
2344: \label{Dnnnn}
2345: \Delta_{0,0,0,0}&=&\frac{8}{1+\tilde{r}_c} \ , \nonumber\\[1ex]
2346: \Delta_{n,n,m,m} &=& 2 N_n^2 N_m^2\, \Big[\, 1\: +\: \tilde{r}_c\: +\:
2347: (1- \tilde{r}_c) \tilde{r}_c^2 \pi^2 R^2 (m_n^2+m_m^2)\: +\:
2348: \tilde{r}_c^4 \pi^4 R^4 m_n^2 m_m^2\, \Big] \, ,\qquad \\[1ex]
2349: \Delta_{n, n, n, n} &=& 3 \tilde{\Delta}_{n,n,n,n}\ =\
2350: 3 N_n^4\, \Big[\, \tilde{r}_c(1-\pi^2 R^2 \tilde{r}_c^2 m_n^2)\: +\:
2351: (1+\pi^2 R^2 \tilde{r}_c^2 m_n^2)^2\, \Big] \, .\nonumber
2352: \end{eqnarray}
2353:
2354: Finally, let us comment on the limit of vanishing $r_c$ for our
2355: analytic results. Taking this limit for the expression
2356: $\Delta(m_k,m_l,m_n)$, for example, is a non-trivial task. One should
2357: take care of the fact that although one may have $m_k+m_l+m_n \neq 0$
2358: for $r_c \ne 0$, it could be that $m_k+m_l+m_n \to 0$ for $r_c \to
2359: 0$. In such a case, the denominator that appears in the first line of
2360: (\ref{Dmasskln}) approaches zero, and one has to carefully expand the
2361: BKT corrections to the KK masses in powers of $r_c$ to obtain a
2362: sensible result. In this way, we find that
2363: \begin{equation}
2364: \label{limD1}
2365: \lim_{r_c \to 0} \; \Delta(m_k,m_l,m_n)\ =\ \delta_{k+l+n,0} \; .
2366: \end{equation}
2367: Analogous considerations for $\Delta(m_k,m_l,m_n,m_m)$ lead to
2368: \begin{equation}
2369: \label{limD2}
2370: \lim_{r_c \to 0} \; \Delta(m_k,m_l,m_n,m_m)\ =\ \delta_{k+l+n+m,0}\; .
2371: \end{equation}
2372: Given that $\delta_{k+l+n,0}$ and $\delta_{k+l+n+m,0}$ are the usual
2373: Kronecker symbols, we have checked that our analytic results perfectly
2374: agree with those presented in~\cite{MPR}, within the context of 5D
2375: Yang--Mills orbifold theories without BKTs.
2376:
2377:
2378: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2379:
2380: \section{High Energy Unitarity Sum Rules}
2381: \label{summation}
2382: \setcounter{equation}{0}
2383:
2384: In proving the ET in Section~2, we encounter infinite sums that
2385: involve products of the gauge-coupling coefficients $\Delta_{k,l,n}$
2386: and $\tilde{\Delta}_{k,l,n}$ defined in Appendix~C
2387: [cf.~(\ref{DeltaDownThree})]. For instance, we have to carry out the
2388: infinite sum, $$\sum_{j=0}^{\infty}\; 2^{-\delta_{j,0}}\,
2389: \Delta^2_{n,n,j}\ ,$$ that occurs in the calculation of
2390: (\ref{gaugeBKT}). Considering~(\ref{Dnnj}), the problem reduces to
2391: finding an expression for the sum,
2392: $$\sum_{j=1}^{\infty}\ \frac{N_j^2}{(m_j^2-4m_n^2)^2}\ .$$ The
2393: calculation of such infinite sums can be carried out analytically, by
2394: extending the complex integration analysis techniques developed
2395: in~\cite{Paes}.
2396:
2397: Let us briefly describe our complex integration analysis method by
2398: considering the infinite sum $$\sum_{j=1}^{\infty}\, N_j^2\; ,$$ where
2399: the normalization factors $N_j$ are given in~(\ref{Nn}). We start
2400: with the complex analytic expression
2401: \begin{equation}
2402: \label{sameRes}
2403: \frac{1}{z\: +\: \tan \pi R z/(\tilde{r}_c \pi R)}\ \approx\ \bigg( 1+
2404: \frac{1}{\tilde{r}_c \cos^2 \pi R z} \bigg)^{-1} \; \frac{1}{z-m_j}\
2405: \equiv\ \frac{\tilde{r}_c\ N(z)}{z-m_j}\ ,
2406: \end{equation}
2407: where we have expanded it about the regions $|z-m_j|<\epsilon $, close
2408: to its poles $\pm m_j$. Here and in the following, we use the
2409: convention that $m_{-j} = -m_j$ and $m_j \ge 0$, for $j\ge 0$. In the
2410: last equality of~(\ref{sameRes}), we have defined the function $N(z)$,
2411: in a way such that it is $N (m^2_j) = N^2_j$ at the residuum
2412: of~(\ref{sameRes}).
2413:
2414: Our next step is to use Cauchy's theorem~\cite{Spiegel} to integrate
2415: the LHS of~(\ref{sameRes}) over the poles $m_j$, i.e.
2416: \begin{equation}
2417: \label{CT}
2418: \lim_{n \to \infty} \, \oint_{C_n} dz \, \bigg(
2419: z + \frac{1}{\tilde{r}_c\pi R} \tan \pi R z \bigg)^{-1}\ =\
2420: \lim_{n \to \infty} \, \oint_{C_n} dz\; \tilde{r}_c\;
2421: \sum_{j=-\infty}^{\infty}\; \frac{N (z)}{z-m_j}\ .
2422: \end{equation}
2423: The contours $C_n$ are the circles $z_n=(n+1/2)/R \; e^{i \theta}$
2424: defined in the complex plane, whose radii are taken to infinity in a
2425: discrete manner, i.e.~$n\to \infty$. This ensures that neither $m_j$,
2426: nor possible poles of the complex function $N(z)$ do lie on the
2427: contour $C_n$. The LHS of~(\ref{CT}) then becomes a simple
2428: integration over the angle $\theta$ that can be performed easily to
2429: give $2\pi i$. The RHS of~(\ref{CT}) can be calculated using Cauchy's
2430: theorem. Thus, we arrive at the equality
2431: \begin{equation}
2432: \label{here}
2433: 2 \pi i\ =\ 2 \pi i \, \sum_{j=-\infty}^{\infty} \, \tilde{r}_c N(m^2_j)\; ,
2434: \end{equation}
2435: which is equivalent to the desired result
2436: \begin{equation}
2437: \label{simpleExample}
2438: \sum_{j=1}^{\infty} N_j^2\ =\ \frac{1}{2 \tilde{r}_c(1+\tilde{r}_c)}\ .
2439: \end{equation}
2440:
2441: The above method can be extended to more complicated infinite sums
2442: that include factors, such as $\tilde{\Delta}^2_{n,j,n}$. For example,
2443: a typical infinite sum that arises in such a calculation is
2444: $$\sum_{j=1}^{\infty}\ \frac{N^2_j\,m_j^4}{(m_j^2-4m_n^2)^2}\ .$$
2445: In this case, we may exploit the analytic structure of the summing
2446: terms with respect to $m_j$, and consider the analytic continuation
2447: $m_j \to z$ in the complex plane. Then, we multiply both the LHS and
2448: the RHS of~(\ref{sameRes}) with the analytic expression
2449: $$ \frac{z^4}{(z^2-4m_n^2)^{2}}\ .$$
2450: This extra factor leaves the integral on the LHS of~(\ref{CT}) intact,
2451: since the new factor goes to 1 in the limit $|z_n| \to \infty$, for
2452: $n\to \infty$. To calculate the RHS of~(\ref{CT}), we have to
2453: properly include the contributions from the two additional double
2454: poles at $z= \pm 2 m_n$. Thus, we finally obtain
2455: \begin{equation}
2456: \label{Sexample}
2457: \sum_{j=1}^{\infty}\ \frac{N^2_j\,m_j^4}{(m_j^2-4m_n^2)^2}\ =\
2458: \frac{1}{12 \pi^4 R^4 m_n^4 \tilde{r}_c^6 N_n^4}\; \bigg(
2459: \Delta_{n,n,n,n}\: +\: \frac{3}{4}\;X_n\,\bigg)\; ,
2460: \end{equation}
2461: where $\Delta_{n,n,n,n}$ and $X_n$ are given in~(\ref{Dnnnn})
2462: and~(\ref{Xn}), respectively.
2463:
2464: With the help of the complex integration analysis method outlined
2465: above, the following list of high energy unitarity sum rules can be
2466: derived:
2467: \begin{eqnarray}
2468: \label{Snnnn1}
2469: \sum_{j=0}^{\infty} 2^{-\delta_{j,0}} \Delta_{n,n,j}^2 \!&=&\!
2470: \Delta_{n,n,n,n}\;,\qquad\qquad
2471: \sum_{j=1}^{\infty} \tilde{\Delta}_{n,n,j}^2\ =\
2472: \tilde{\Delta}_{n,n,n,n}\; ,\\
2473: \label{Snnnn2}
2474: \sum_{j=0}^{\infty} 2^{-\delta_{j,0}}
2475: \Delta_{n,n,j} \tilde{\Delta}_{n,j,n} \!&=&\! \tilde{\Delta}_{n,n,n,n}\;,
2476: \qquad
2477: \sum_{j=0}^{\infty} 2^{-\delta_{j,0}} \tilde{\Delta}_{n,j,n}^2\ =\
2478: \Delta_{n,n,n,n}\: +\: X_n\; ,\\
2479: \label{Snnmm1}
2480: \sum_{j=0}^{\infty} 2^{-\delta_{j,0}}
2481: \Delta_{n,j,m}^2 \! &=&\! \Delta_{n,n,m,m}\; ,\quad
2482: \sum_{j=0}^{\infty} 2^{-\delta_{j,0}} \tilde{\Delta}_{n,j,n}
2483: \tilde{\Delta}_{m,j,m} \ =\ \Delta_{n,n,m,m}\: +\: Y_{n,m}\; ,\\
2484: \label{Snnmm2}
2485: \sum_{j=0}^{\infty} 2^{-\delta_{j,0}} \tilde{\Delta}_{n,j,m}^2\!&=&\!
2486: \Delta_{n,n,m,m}\: +\: Y_{n,m}\; ,\qquad \\
2487: \label{Sklnm}
2488: \sum_{j=0}^{\infty} 2^{-\delta_{j,0}} \tilde{\Delta}_{k,j,l}
2489: \tilde{\Delta}_{n,j,m} \!&=&\! \Delta_{k,l,n,m}\: +\: Z_{k,l,n,m}\; ,
2490: \end{eqnarray}
2491: where
2492: \begin{eqnarray}
2493: \label{Ynm}
2494: Y_{n,m} &=& 4\,N_n^2 N_m^2\, \pi^2 R^2 \tilde{r}_c^3\, (m_n^2 + m_m^2)\;,\\
2495: \label{Zklnm}
2496: Z_{k,l,n,m} &=& 2\, N_k N_l N_n N_m\,
2497: \pi^2 R^2 \tilde{r}_c^3\, (m_k^2 + m_l^2 + m_n^2 +
2498: m_m^2) \; .
2499: \end{eqnarray}
2500: The sum rules~(\ref{Snnnn1}) and (\ref{Snnnn2}) are relevant to
2501: restore unitarity in high energy $2\to 2$ processes where all KK gauge
2502: modes in the initial and final state have the same mass $m_n$. The
2503: sum rules~(\ref{Snnmm1}) and (\ref{Snnmm2}) ensure high energy
2504: unitarity for $2\to 2$ scatterings, where the KK gauge modes in the
2505: initial and final states have the same mass $m_n$ and $m_m$,
2506: respectively, but $m_n \neq m_m$. Finally, (\ref{Sklnm}) is relevant
2507: to restore unitarity in $2\to 2$ scatterings, when all 4 asymptotic
2508: states happen to have different masses.
2509:
2510:
2511: \end{appendix}
2512:
2513: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2514: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2515: \newpage
2516:
2517: \begin{thebibliography}{99}
2518:
2519: \bibitem{SS} J. Scherk and J. H. Schwarz, Nucl.\ Phys.\ {\bf B153} (1979) 61.
2520:
2521: \bibitem{Hosotani} Y. Hosotani, Annals Phys.\ {\bf 190} (1989) 233.
2522:
2523: \bibitem{Quiros} G. von Gersdorff, N. Irges, M. Quiros, Nucl.\ Phys.\
2524: {\bf B635} (2002) 127.
2525:
2526: \bibitem{Csaki} C. Cs\'aki, C. Grojean, H. Murayama, L. Pilo and
2527: J. Terning, Phys.\ Rev.\ {\bf D69} (2004) 055006.
2528:
2529: \bibitem{Georgi} H. Georgi, A. K. Grant and G. Hailu, Phys.\ Lett.\
2530: {\bf B506} (2001) 207.
2531:
2532: \bibitem{Carena} M. Carena, T. T. M. Tait and C. E. M. Wagner, Acta
2533: Phys.\ Polon.\ {\bf 33} (2002) 2355.
2534:
2535: \bibitem{MPR} A. M\"uck, A. Pilaftsis and R. R\"uckl, Phys.\ Rev.\
2536: {\bf D65} (2002) 085037.
2537:
2538: \bibitem{PS} J. Papavassiliou and A. Santamaria, Phys.\ Rev.\ {\bf
2539: D63} (2001) 125014;\\
2540: J.~F.~Oliver, J.~Papavassiliou and A.~Santamaria,
2541: Phys.\ Rev.\ {\bf D67} (2003) 125004.
2542:
2543: \bibitem{BRS} C. Becchi, A. Rouet and R. Stora, Comm.\ Math.\ Phys.\
2544: {\bf 52} (1975) 55.
2545:
2546: \bibitem{Ward} J. C. Ward, Phys.\ Rev.\ {\bf 78} (1950) 182;\\
2547: Y. Takahashi, Nuovo Cim.\ {\bf 6} (1957) 371.
2548:
2549: \bibitem{ST} A. A. Slavnov, Theor.\ and Math.\ Phys.\
2550: {\bf 10} (1972) 99;\\ J. C. Taylor, Nucl.\ Phys.\ {\bf B33} (1971) 436.
2551:
2552: \bibitem{CLT} J. M. Cornwall, D. N. Levin and G. Tiktopoulos,
2553: Phys.\ Rev.\ {\bf D10} (1974) 1145.
2554:
2555: \bibitem{LQT} B. W. Lee, C. Quigg and H. B. Thacker, Phys.\ Rev.\ {\bf
2556: D16} (1977) 1519.
2557:
2558: \bibitem{Dicus} R. S. Chivukula, D. A. Dicus and H. J. He, Phys.\
2559: Lett.\ {\bf B525} (2002) 175;\\
2560: R.~S.~Chivukula, D.~A.~Dicus, H.~J.~He and S.~Nandi,
2561: %``Unitarity of the higher dimensional standard model,''
2562: Phys.\ Lett.\ {\bf B562} (2003) 109.
2563:
2564: \bibitem{CG} M. S. Chanowitz and M. K. Gaillard, Nucl.\ Phys.\ {\bf
2565: B261} (1985) 379;\\
2566: G.~J.~Gounaris, R.~K\"ogerler and H.~Neufeld,
2567: %``On The Relationship Between Longitudinally Polarized Vector Bosons And
2568: %Their Unphysical Scalar Partners,''
2569: Phys.\ Rev.\ {\bf D34} (1986) 3257.
2570:
2571: \bibitem{Dominici} S.~De Curtis, D.~Dominici and J.~R.~Pelaez,
2572: %``Strong tree level unitarity violations in the extra dimensional standard
2573: %model with scalars in the bulk,''
2574: Phys.\ Rev.\ {\bf D67} (2003) 076010.
2575:
2576: \bibitem{Abbott} L.~F.~Abbott,
2577: %``The Background Field Method Beyond One Loop,''
2578: Nucl.\ Phys.\ {\bf B185} (1981) 189.
2579:
2580: \bibitem{Pokorski} S. Pokorski, ``Gauge Field Theories'', Cambridge
2581: University Press 2000.
2582:
2583: \bibitem{SMreviews} K.~I.~Aoki, Z.~Hioki, M.~Konuma, R.~Kawabe and T.~Muta,
2584: %``Electroweak Theory. Framework Of On-Shell Renormalization And Study Of
2585: %Higher Order Effects,''
2586: Prog.\ Theor.\ Phys.\ Suppl.\ {\bf 73} (1982) 1;\\
2587: M.~B\"ohm, H.~Spiesberger and W.~Hollik,
2588: %``On The One Loop Renormalization Of The Electroweak Standard Model And
2589: %Its Application To Leptonic Processes,''
2590: Fortsch.\ Phys.\ {\bf 34} (1986) 687.
2591:
2592: \bibitem{Santiago} F.~del Aguila, M.~Perez-Victoria and J.~Santiago,
2593: JHEP {\bf 0302} (2003) 051, and references therein.
2594:
2595: \bibitem{Papa} J. Papavassiliou and A. Pilaftsis, Phys.\ Rev.\ {\bf D54}
2596: (1996) 5315; Phys.\ Rev.\ Lett.\ {\bf 80} (1998) 2785;
2597: Phys.\ Rev.\ {\bf D58} (1998) 053002;\\
2598: D. Binosi,
2599: %``Electroweak pinch technique to all orders,''
2600: J.\ Phys.\ {\bf G30} (2004) 1021.
2601:
2602: \bibitem{Buras} For example, see
2603: A.~J.~Buras, A.~Poschenrieder, M.~Spranger and A.~Weiler,
2604: %``The impact of universal extra dimensions on B $\to$ X/s gamma, B $\to$ X/s
2605: %gluon, B $\to$ X/s mu+ mu-, K(L) $\to$ pi0 e+ e-, and epsilon'/epsilon,''
2606: Nucl.\ Phys.\ {\bf B678} (2004) 455.
2607:
2608:
2609: \bibitem{Thompson} For example, see, R.~Delbourgo, S.~Twisk and
2610: G.~Thompson,
2611: %``Massive Yang-Mills Theory: Renormalizability Versus Unitarity,''
2612: Int.\ J.\ Mod.\ Phys.\ {\bf A3} (1988) 435.
2613:
2614: \bibitem{Ohl} T.~Ohl and C.~Schwinn,
2615: %``Unitarity, BRST symmetry and Ward identities in orbifold gauge theories,''
2616: Phys.\ Rev.\ {\bf D70} (2004) 045019.
2617:
2618: \bibitem{Yao} Y.~P.~Yao and C.~P.~Yuan,
2619: %``Modification Of The Equivalence Theorem Due To Loop Corrections,''
2620: Phys.\ Rev.\ {\bf D38} (1988) 2237;\\
2621: J.~Bagger and C.~Schmidt,
2622: %``Equivalence Theorem Redux,''
2623: Phys.\ Rev.\ D {\bf D41} (1990) 264;\\
2624: H.~J.~He, Y.~P.~Kuang and X.~Y.~Li,
2625: %``On the precise formulation of equivalence theorem,''
2626: Phys.\ Rev.\ Lett.\ {\bf 69} (1992) 2619.
2627:
2628: \bibitem{Denner} A.~Denner and S.~Dittmaier,
2629: %``Dyson summation without violating Ward identities and the Goldstone-boson
2630: %equivalence theorem,''
2631: Phys.\ Rev.\ {\bf D54} (1996) 4499.
2632:
2633: \bibitem{Paes} G. Bhattacharyya, H. V. Klapdor-Kleingrothaus, H. P\"as
2634: and A. Pilaftsis, Phys.\ Rev.\ {\bf D67} (2003) 113001.
2635:
2636: \bibitem{Froissart} A. Martin, Phys.\ Rev.\ {\bf 129} (1963) 1432;
2637: M. Froissart, Phys.\ Rev.\ {\bf 123} (1961) 1053.
2638:
2639: \bibitem{Barton} G. Barton, ``Introduction to Dispersion Techniques in
2640: Field Theory'', Benjamin 1965.
2641:
2642: \bibitem{MPR2} For a recent global-fit analysis, see,
2643: A. M\"uck, A. Pilaftsis and R. R\"uckl, Nucl.\ Phys.\
2644: {\bf B687} (2004) 55;\\
2645: R.~Barbieri, A.~Pomarol, R.~Rattazzi and A.~Strumia,
2646: %``Electroweak symmetry breaking after LEP1 and LEP2,''
2647: hep-ph/0405040.
2648:
2649: \bibitem{Spiegel} M. R. Spiegel, ``Schaum's outline series: Complex
2650: Variables'', McGraw-Hill, New York (1964).
2651:
2652:
2653:
2654: \end{thebibliography}
2655:
2656: \end{document}
2657:
2658: