1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: % %
3: % ON DISTINGUISHING RADIONS FROM HIGGS BOSONS %
4: % %
5: % Prasanta Kumar Das, Santosh Kumar Rai and Sreerup Raychaudhuri %
6: % %
7: % Date of last modification: 31-12-2004 %
8: % %
9: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
10: \documentclass[12pt]{article}
11: \usepackage{graphicx}
12: \topmargin -0.05in
13: \textheight 9.25in
14: \textwidth 6.50in
15: \oddsidemargin -0.30in
16: \parindent 0.25in
17: \def\baselinestretch{1.35}
18: %\setcounter{page}{0}
19: %------------------------------------------------------------------------%
20: \begin{document}
21: %------------------------------------------------------------------------%
22: \def\ie{{\it i.e.}}
23: \def\eg{{\it e.g.}}
24: \def\qv{{\it q.v.}}
25: \def\ib{{\it ibid.}}
26: \def\bc{\begin{center}}
27: \def\ec{\end{center}}
28: \def\bq{\begin{quotation}}
29: \def\eq{\end{quotation}}
30: \def\bit{\begin{itemize}}
31: \def\eit{\end{itemize}}
32: \def\benum{\begin{enumerate}}
33: \def\eenum{\end{enumerate}}
34: \def\be{\begin{equation}}
35: \def\ee{\end{equation}}
36: \def\bearr{\begin{eqnarray}}
37: \def\eearr{\end{eqnarray}}
38: \def\bdm{\begin{displaymath}}
39: \def\edm{\end{displaymath}}
40: \def\bfl{\begin{flushleft}}
41: \def\efl{\end{flushleft}}
42: \def\bfr{\begin{flushright}}
43: \def\efr{\end{flushright}}
44: \def\del{\partial}
45: \def\goes{\longrightarrow}
46: \def\epem{e^+e^-}
47: \def\MR{M_\Phi}
48: \def\VR{\Lambda_\Phi}
49: %------------------------------------------------------------------------%
50: % Title Page %
51: %------------------------------------------------------------------------%
52: %\thispagestyle{empty}
53: %\bfr
54: %\today \\
55: %hep-ph/0410244
56: %\efr
57: %\vskip 20pt
58: \bc
59: {\Large\bf
60: On Distinguishing Radions from Higgs Bosons} \\
61: \vskip 10pt
62: {\sl Prasanta Kumar Das $^a$, Santosh Kumar Rai} $^b$ and
63: {\sl Sreerup Raychaudhuri $^b$} \\
64: \bigskip
65: $^a$~{\rm Department of Physics, Chung Yuan Christian University, \\
66: 22 Pu-Jen, Pu-chung Li, Chung-Li (32023), Taiwan, R.O.C.}\\
67: {\rm Electronic address:} {\sf pdas@phys.cycu.edu.tw} \\
68: \bigskip
69: $^b$~{\rm Department of Physics, Indian Institute of Technology,
70: Kanpur 208016, India.} \\
71: {\rm Electronic addresses:} {\sf sreerup@iitk.ac.in, skrai@iitk.ac.in} \\
72: \vskip 20pt
73: {\large\bf ABSTRACT}
74: \ec
75: \bq \noindent {\footnotesize
76: Radion couplings are almost identical to Higgs boson couplings, making it very
77: difficult to distinguish the two states when the radion mass and vacuum
78: expectation value are similar to those of the Higgs boson. The only real
79: difference lies in the fact that the coupling of radions to off-shell fermions is
80: proportional to the momentum rather than the mass of the fermion. This extra
81: contribution gets cancelled in all tree-level processes and shows up only in
82: loop-induced processes like $\Phi \to \gamma\gamma$ and $\Phi \to gg$. We perform
83: a careful calculation of the branching ratios and establish that they can prove
84: crucial in clearly distinguishing a radion from a Higgs boson. This claim is made
85: concrete by evaluating the exclusive cross-sections in a radiative process
86: involving elementary scalars.}
87: \eq
88: %\vfill
89: \vskip 10pt
90: %------------------------------------------------------------------------%
91: % Body of the Paper %
92: %------------------------------------------------------------------------%
93: %\newpage
94:
95: \noindent
96: In recent years the two-brane model of Randall and Sundrum\cite{RS1} has
97: attracted a great deal of attention because it provides an elegant
98: solution to the thirty-odd year-old hierarchy problem of high energy
99: physics. The most attractive feature of the 1+4 dimensional
100: Randall-Sundrum (RS) model is that it explains the large hierarchy
101: between the electroweak scale (0.1 -- 1 TeV) and the Planck scale
102: ($10^{16}$ TeV) in terms of an exponential damping of the gravitational
103: field across a small compact fifth dimension, without recourse to
104: unnaturally large numbers\footnote{However, it is only fair to mention
105: that this is achieved at the expense of a delicate fine-tuning of the
106: five-dimensional cosmological constant with the energy densities on two
107: branes at opposite ends of the fifth dimension.}. Since the hierarchy of
108: scales is generated by an exponential damping across the fifth dimension,
109: the size of this dimension requires to be just at the right value to
110: ensure that the hierarchy is indeed a factor of $\frac{M_{Pl}}{M_{ew}}
111: \sim 10^{16}$. Since the fifth dimension has the topology of a
112: once-folded circle, with two $D_3$-branes at the fixed points, this
113: amounts to fixing the distance $R_c$ between the two branes very
114: precisely. This, in turn, would require a mechanism to protect the radius
115: against large quantum corrections. The absence of such a mechanism is a
116: major flaw in the original braneworld model of Arkani-Hamed, Dimopoulos
117: and Dvali\cite{ADD}. The question was left unresolved even in the
118: original work\cite{RS1} of RS, but an elegant model to explain this was
119: given shortly afterwards by Goldberger and Wise\cite{GW2}. They used the
120: simple device of generating a force between the two branes which would
121: ensure equilibrium when the distance between them is precisely the radius
122: $R_c$ required to generate the required hierarchy. Because of the folded
123: structure of the fifth dimension, it is only necessary to generate an
124: attractive force between the branes --- since each brane is,
125: topologically speaking, on {\sl both} sides of the other, the two pulls
126: will balance at the equilibrium point. The attractive force is modelled
127: by postulating the time-honoured device of a scalar field which lives in
128: all five dimensions (bulk) and has quartic self-interactions, in the
129: bulk, as well as in projection on the branes. It is necessary only to
130: tune the vacuum expectation values (vev's) of the scalar field on the two
131: branes to get an attractive force as required. In fact, it can be easily
132: shown that the potential has an extremely steep minimum at the argument
133: $R_c$, indicating that the hierarchy is fixed very accurately for small
134: oscillations of the bulk size about this minimum.
135:
136: \bigskip\noindent
137:
138: An important consequence of the (original) RS model is that there exists
139: on the TeV-brane (which represents the observable Universe), a scalar
140: field $\Phi$, which is very much like a dilaton field and has been dubbed
141: the {\sl radion}. The RS metric, with radial fluctuations, is written in
142: the form
143: \be
144: ds^2 = e^{-2KT(x)\varphi} ~g_{\mu\nu}(x) dx^\mu dx^\nu - [T(x)]^2 ~d\varphi^2
145: \ee
146: where $T(x)$ is a modulus field representing dilatation of the bulk, $K$
147: is the bulk curvature and $\varphi$ is an angular coordinate describing
148: the fifth dimension. We can now show\cite{GW1} that the five-dimensional
149: Einstein-Hilbert action reduces, in four dimensions, to a theory with
150: Kaluza-Klein gravity and a scalar term
151: \be
152: S_\Phi = \int d^4x \sqrt{-|g|} ~~\frac{1}{2} \del_\mu \Phi(x) \del^\mu
153: \Phi(x)
154: \ee
155: where the (massless, free) scalar field
156: \be
157: \Phi(x) = \sqrt{24M_5^3/K}~e^{-\pi K T(x)}
158: \ee
159: is the radion field. The Goldberger-Wise stabilisation mechanism, with a
160: bulk scalar field $B(x,\varphi)$ then creates an effective scalar
161: $\Phi^4$-potential for the radion field $\Phi(x)$ with a minimum at
162: $\langle T(x)\rangle = R_c$, where $KR_c \simeq 11.7$, the value required
163: to generate the electroweak hierarchy. This potential includes a radion
164: mass term $m_\Phi^2 \Phi^2$ where $m_\Phi$ is determined by the mass
165: $m_B$ and couplings of the bulk scalar field. Though $m_B$ is unknown,
166: arguing\cite{GW2} that $m_B$ should be of the order of the Planck scale
167: --- which is the only fundamental scale in the RS model --- leads to the
168: result that $m_\Phi$ should be of the order of the electroweak
169: symmetry-breaking scale, i.e. $m_\Phi \sim 100$~GeV.
170:
171: \bigskip\noindent
172: Phenomenology of the radion field\cite{DHR1} starts with the coupling of
173: the radion to ordinary matter, consisting of the Standard Model (SM)
174: fields. This interaction which arises from the usual gravitational
175: coupling to matter, is the same as the coupling of a dilaton field, viz.,
176: \be
177: {\cal L}(x) = -\frac{1}{\VR} ~\Phi(x) ~\eta^{\mu\nu}T_{\mu\nu}(x)
178: \label{radcoup}
179: \ee
180: where $\Lambda_\Phi$ is the radion vev, corresponding to the minimum of
181: the radion potential on the TeV brane and $T_{\mu\nu}$ is the
182: energy-momentum tensor composed of SM fields. The radion Feynman rules
183: can, therefore, be read off from (for example), the expressions given in
184: Ref.\cite{HLZ} by simply substituting the radion for the dilaton and
185: $\VR$ for $\bar M_P$. It turns out that the couplings are rather similar
186: to those of Higgs fields to other SM particles, though, of course, the
187: radion couplings originate from the
188: couplings of the bulk scalar\footnote{It may be noted that the vev,
189: $\VR$ is a free parameter arising from the couplings of the bulk scalar
190: and has no
191: relation to the parameter $\Lambda_\pi$ which is strongly constrained by
192: bounds on the graviton mass \cite{lambdapi}.}
193: while Higgs boson couplings arise from the Standard Model sector. An
194: important --- and for this work, crucial --- difference arises in the
195: fact that there are {\it momentum-dependent} terms in the radion coupling
196: to matter, which are not present in the Higgs boson coupling. To see
197: this, we write out in full the energy-momentum tensor for a scalar field
198: $S(x)$, a fermion field $\psi(x)$ and a vector gauge field $V_\mu(x)$.
199: \bearr
200: \eta^{\mu\nu}~T_{\mu\nu} & = &
201: -2 \left[(D^\mu S)^\dagger (D_\mu S) - 2 M_S^2 ~S^\dagger S \right]
202: \nonumber \\ & &
203: -3i \bar{\psi} \not{\!\!D} \psi + 4 m_\psi~\bar{\psi}\psi + \frac{3i}{2}
204: \del^\mu \left[ \bar{\psi}\gamma_\mu\psi\right]
205: \nonumber \\ & &
206: - M_V^2 ~V_\mu(x) V^\mu(x)
207: \label{tensor}
208: \eearr
209: where $D^\mu = \del^\mu + igT_a V^\mu_a$ and $T_a$ is the gauge group
210: generator in the appropriate representation. The explicit form of this
211: for the Standard Model fields is given in Ref.\cite{KCH}. This
212: interaction tells us that the vertex for $\Phi(p) \to \bar{\psi}(k_1) +
213: \psi(k_2)$ is given (in momentum space) by
214: \be
215: {\cal L}_{\Phi \psi\bar \psi} = \frac{3}{2\VR} \bar u(k_1) \left(
216: \not{\!k_1} + \not{\!k_2} - \frac{8}{3}m_\psi \right) u(k_2) ~\Phi(p) \ .
217: \label{offshell}
218: \ee
219: If the fermions are on-shell, we can use the Dirac equation to write the above
220: equation as
221: \be
222: {\cal L}_{\Phi \psi\bar \psi} =
223: -\frac{m_\psi}{\VR} \bar u(k_1) u(k_2) \Phi(p) \ ,
224: \label{onshell}
225: \ee
226: which is a Yukawa coupling very reminiscent of the Higgs boson.
227: Obviously, if the fermions are {\it off-shell}, the coupling is
228: different. An immediate consequence of the above is that radions, unlike
229: Higgs bosons can have significant couplings to light fermions, such as
230: electrons and $u,d$ quarks, if any of the fermions is off-shell and has
231: large energy-momentum values.
232:
233: \bigskip\noindent
234: Once produced, a massive radion will clearly decay. Being a constituent
235: field of the metric tensor the radion couples, as described in Eqns.
236: (\ref{radcoup}) and (\ref{tensor}), to {\sl all} pairs of SM particles.
237: Naturally, only those decays which are kinematically allowed will occur.
238: For fermionic decay modes $\Phi \to f\bar{f}$ (on-shell), the partial
239: decay widths will be suppressed by the factor $m_f^2/\Lambda_\Phi^2$,
240: since the fermionic states will be on-shell and the radionic coupling
241: will be Higgs boson-like. Thus, we need to consider only the following
242: decay channels:
243: \bearr
244: \Phi & \goes & \gamma\gamma,~gg \nonumber \\
245: & \goes & \tau^+\tau^- ,~c\bar{c},~b\bar{b} \nonumber \\
246: & \goes & W^+W^-,~ Z^0Z^0,~H^0H^0 \nonumber \\
247: & \goes & t\bar{t}
248: \nonumber
249: \eearr
250: Detailed formulae for these are given in Ref.\cite{KCH}. However, in this
251: last-mentioned work, the formulae for the decay widths $\Phi \goes \gamma
252: \gamma,~g g$ have been calculated assuming the coupling in
253: Eqn.(\ref{onshell}) rather than the one in Eqn.(\ref{offshell}), and are,
254: therefore, somewhat inaccurate. This is because loop-fermions will
255: obviously be off-shell and there will be consequent modifications to the
256: decay amplitude itself. It is worth noting that though the lighter
257: fermions will now couple to the scalar field through their momenta rather
258: than their masses, there is a helicity flip involved in the
259: scalar-vector-vector one-loop diagrams, which results in an amplitude
260: proportional to the fermion masses. As a result, it is only the top quark
261: loop which makes any significant contribution --- which is also the case
262: with Higgs bosons. The differences arise, then, solely from the extra
263: off-shell terms in the $\Phi t\bar t$ coupling.
264:
265: \bigskip\noindent
266: In this work, we have calculated the one-loop-mediated decay widths
267: afresh, using the off-shell coupling of Eqn.(\ref{offshell}). The
268: relevant Feynman diagrams are listed in Figure~1. Of these, the triangle
269: diagrams marked (A) and (B) are similar to those responsible for the
270: process $H^0 \to \gamma\gamma$. The ones marked (C) and (D) arise from
271: non-renormalisable couplings of the radion (dilaton) to a photon and a
272: fermion pair, which arise at the lowest order in an effective theory of
273: gravity coupling to matter. It is worth mentioning at this point that the
274: presence of these vertices is responsible, in tree-level processes like,
275: for example, $e^+ e^- \to Z \Phi$ or $e^+ e^- \to \ell^+\ell^-\Phi$, for
276: precisely cancelling out the momentum-dependent part of the
277: $\ell^+\ell^-\Phi$ coupling and rendering the cross-section totally
278: Higgs-like. However, the situation is different inside a loop diagram, as
279: we shall presently argue.
280:
281: % ------------------------------------------------------------------
282: \begin{center}
283: \includegraphics[height=3in]{Fig1.eps}
284: \end{center}
285: \noindent {\bf Figure 1}.
286: {\footnotesize\it Feynman diagrams with a top quark loop contributing to the process
287: $\Phi \goes \gamma\gamma$ at the one-loop level.}
288: % ------------------------------------------------------------------
289:
290:
291: \bigskip\noindent
292: Following the usual procedure\cite{ABJ} for calculating the amplitude for a
293: process like $\Phi(p) \goes \gamma(k_1)\gamma(k_2)$, we write the amplitude as
294: \be
295: {\cal M}(\Phi\to\gamma\gamma) =
296: \left[ A(p^2)k_1^\nu k_2^\mu + B(p^2) \eta^{\mu\nu} \right]
297: \varepsilon^*_\mu(k_1)\varepsilon^*_\nu(k_2)
298: \ee
299: which is consistent with Lorentz-symmetry and the transverse nature of
300: the photon. Imposition of gauge symmetry at once leads to the Ward
301: identity
302: \be
303: B(p^2) = - A(p^2) k_1.k_2
304: \ee
305: which relates the (naively divergent) form factor $B(p^2)$ to the finite form
306: factor $A(p^2)$, and hence acts as a regulator for the process. The amplitude
307: then becomes
308: \be
309: {\cal M}(\Phi\to\gamma\gamma) =
310: A(p^2) \left( k_1^\nu k_2^\mu - k_1.k_2 ~\eta^{\mu\nu} \right)
311: \varepsilon^*_\mu(k_1)\varepsilon^*_\nu(k_2)
312: \ee
313: which means that it is only necessary to calculate the finite form factor
314: $A(p^2) = A(M_\Phi^2)$ in order to get the decay width. Since $A(p^2)$
315: can be calculated by evaluating the coefficients of $k_1^\nu k_2^\mu$
316: alone, it can now be seen, by writing down the Feynman amplitudes for the
317: diagrams marked (A)--(D) in Figure~1, that the contributions to $A(p^2)$
318: arise from those marked (A) and (B), but not from those marked (C) and
319: (D). Thus, the exact cancellation of momentum-dependent terms, which
320: renders the effective radion coupling Higgs-like in the tree-level, does
321: {\sl not} go through at the one-loop level. This also ensures that the
322: diagrams (A) and (B) in Figure~1 have residual momentum-dependent effects
323: and justifies corrections to the partial decay width of Ref.\cite{KCH}
324: --- which is the thrust of our present work.
325:
326: \bigskip\noindent
327: For the two-photon decay mode, then, our final results are
328: \be
329: \Gamma(\Phi \to \gamma\gamma) = \frac{1}{64\pi} \frac{M_\Phi^3}{\Lambda_\Phi^2}
330: \left(\frac{ \alpha}{\pi}\right)^2 \left|I_\gamma\right|^2
331: \qquad
332: {\rm where}
333: \qquad I_\gamma = b_{QED} + I_W + \sum_{f} N_c~Q_f^2~I_f
334: \ee
335: In the above, $N_c$ is the number of colours of the fermion $f$ and $Q_f$
336: is the fermionic charge. The QED beta function (appearing because of the
337: trace anomaly) is given by\cite{UMA1,UMA2}
338: \bearr
339: b_{QED} & = & \frac{20}{9} ~{\rm for}~~M_\Phi \leq 2M_W \nonumber \\
340: & = & \frac{31}{18} ~{\rm for}~~2M_W < M_\Phi \leq 2m_t\nonumber \\
341: & = & \frac{12}{6} ~{\rm for}~~M_\Phi > 2m_t
342: \eearr
343: and the loop integral functions $I_W$ and $I_f$ are given by
344: \bearr
345: I_W & = & -1 - \frac{3}{2}\lambda_W + \frac{3}{2}\lambda_W
346: (1 - \frac{1}{2}\lambda_W)~{\rm F}(\lambda_W) \nonumber \\
347: I_f & = & - 8 \lambda_f - \lambda_f ( 4 \lambda_f - 1)~{\rm F}(\lambda_f)
348: \label{function}
349: \eearr
350: where $\lambda_i = \left({2m_i}/{M_\Phi}\right)^2$, with $i$ running over
351: all the particles involved, and ${\rm F}(\lambda)$ is given by
352: \bearr
353: {\rm F}(\lambda)
354: & = & -2\left[\sin^{-1} \frac{1}{\sqrt{\lambda}}\right]^2 ~~{\rm for}~~
355: \lambda \geq 1 \nonumber \\
356: & = & -\frac{\pi^2}{2} + \frac{1}{2}\log^2
357: \frac{1 + \sqrt{1 - \lambda}} {1 - \sqrt{1 - \lambda}}
358: - i\pi\log \frac{1 + \sqrt{1 - \lambda}} {1 - \sqrt{1 - \lambda}}
359: ~~{\rm for}~~ \lambda < 1 \nonumber
360: \eearr
361: This partial decay width differs from the analogous $H^0 \to
362: \gamma\gamma$ decay width\cite{BP} in two major particulars, viz.,
363: \bit
364: \item The presence of the trace anomaly, i.e. the $b_{QED}$ term.
365: \item A different function $I_f$ in Eqn.~(\ref{function}) from that given in, for
366: example, Ref.\cite{BP}.
367: \eit
368:
369: \bigskip\noindent
370: We now go on to calculate the very-similar process $\Phi \goes gg$, which
371: yields a partial decay width
372: \be
373: \Gamma(\Phi\to gg) = \frac{1}{64 \pi} \frac{M_\Phi^3}{\Lambda_\Phi^2}
374: \left(\frac{ \alpha_s}{\pi}\right)^2 \left|I_g\right|^2
375: \qquad
376: {\rm where}
377: \qquad I_g = b_{QCD} + \sum_{f} \sqrt{2}~I_f
378: \ee
379: and, obviously, there is no $I_W$ contribution. The QCD beta function is given
380: by the usual formula
381: \bearr
382: b_{QCD} = 11 - \frac{2}{3}N_f
383: & = & \frac{23}{3} ~{\rm for}~~M_\Phi \leq 2M_t \nonumber \\
384: & = & \frac{21}{3} ~{\rm for}~~M_\Phi > 2M_t
385: \eearr
386: and for $\alpha_s$ we take the usual running value governed by $b_{QCD}$.
387:
388: % ------------------------------------------------------------------
389: \begin{center}
390: \includegraphics[height=3in]{Fig2.eps}
391: \end{center}
392: \vskip -15pt
393: \noindent {\bf Figure 2}.
394: {\footnotesize\it Branching ratios for ($a$) a radion and ($b$) a Higgs boson
395: (of the Standard Model) as a function of the mass. Kinks at kinematic
396: thresholds
397: are mostly due to numerical instabilities. Note the enormous difference in the
398: two-gluon decay mode for the two cases.}
399: % ------------------------------------------------------------------
400:
401: \bigskip\noindent
402: Branching ratios of the radion to different decay channels may now be
403: calculated by combining the above formulae with those given in, for
404: example, Ref.\cite{KCH}, and varying the radion mass $M_\Phi$. Obviously
405: there will be no dependence on the radion vev, since all the partial
406: decay widths contain the same factor $\Lambda_\Phi^{-2}$. We have
407: exhibited our results in Figure~2($a$), which show the principal
408: branching ratios of the radion, assuming a Higgs boson mass of 150~GeV.
409: For comparison, Figure~2($b$) shows a similar set of branching ratios
410: (except the HH decay mode) for a Standard Model Higgs boson with masses
411: run over the same range. The dot-dashed lines marked $gg^{\cite{KCH}}$
412: and $\gamma\gamma^{\cite{KCH}}$ correspond to earlier results presented
413: in Ref.~\cite{KCH} where the momentum dependence of the
414: radion-fermion-antifermion coupling had not been taken into account. It
415: may be noted that the difference is quite significant, and indicates that
416: a cancellation takes place between the finite part of the loop diagram
417: and the trace anomaly, which is more pronounced for a heavy radion.
418:
419: \bigskip\noindent
420: As the figure shows, the decay patterns of the two scalar particles in
421: question exhibit a great deal of similarity but have some significant
422: differences also. Once above the $WW$ and $ZZ$ thresholds, both decay
423: primarily to weak bosons, with a small percentage of top-anti top decays
424: when the corresponding threshold is crossed. Both show significant
425: branching ratios for the $WW^*$ and $ZZ^*$ modes in the scalar mass range
426: between $M_W$ to $2M_W$ and $M_Z$ to $2M_Z$. At small masses, again, both
427: show large branching ratios for the $b\bar b$ decay mode, as well as some
428: for the $c\bar c$ and $\tau^+\tau^-$ channels. However, there the
429: similarity ends. The radion $\Phi$ has a $\Phi \to HH$ decay channel,
430: which is obviously forbidden for the Higgs boson.
431: Of greater interest is the loop-mediated decay $\Phi \to
432: \gamma\gamma$, which is at the level of a few per mil when the radion is
433: light, but is much smaller for a Higgs boson of corresponding mass. (Of
434: course, such light Higgs bosons have not been found at LEP, so this decay
435: mode is not really of much use in distinguishing radions from Higgs
436: bosons.) However, the real {\sl pi\`ece de resistance} is the $\Phi \to
437: gg$ channel, which has a branching ratio around two orders of magnitude
438: larger than that of the usual $H \to gg$ process. This branching also
439: dominates when only the trace anomaly contribution is taken and people
440: have presented ways of distinguishing radions and Higgs in this
441: light\cite{huitu}. Apart from enhancing the branching ratio for a radion
442: decaying to two hadronic jets significantly above the similar decay of
443: the Higgs boson, it reduces, (for a light radion, the branching ratio to
444: a $b\bar b$ pair quite significantly.) In fact, we find that for light
445: radions, the {\sl dominant} decay mode is to gluon pairs, while for a
446: light Higgs boson, the dominant decay mode is to $b \bar b$. Since the
447: last decay can be pinned down with a fair degree of efficiency by
448: $b$-tagging methods, we obtain another means of distinguishing between
449: radions and Higgs bosons.
450:
451: \bigskip\noindent
452: As an example of the efficacy of these ideas, we now consider a 1~TeV linear
453: $e^+e^-$ collider and calculate the production of a radion in association
454: with a $Z^0$ boson through a process of the form
455: \bearr
456: e^+ e^- & \goes & Z^0 + \Phi \nonumber \\
457: & \hookrightarrow & \ell^+ \ell^- + \Phi \nonumber
458: \eearr
459: which is then compared with the usual Higgs-strahlung
460: process\cite{HSTAHL} obtained by replacing $\Phi$ by $H^0$ in the above
461: process and considering the case when the masses and vev's are equal (or
462: comparable). In the above $\ell = e, \mu, \tau$, and we have folded in
463: the relevant detection efficiencies (i.e. 90\% for $\ell = e, \mu$ and
464: 80\% for $\ell = \tau$). Of course, there is a fundamental difference in
465: the two cases because the electroweak vev $v_{ew} \simeq 246$~GeV is
466: known, while the radion vev $\Lambda_\Phi$ is an unknown parameter. This
467: feature is also taken care of in our analysis.
468:
469: \bigskip\noindent
470: The discussion in the preceding paragraph makes it clear that it is
471: interesting to focus on two kinds of final states, viz.
472: \benum
473: \item $e^+ e^- \goes \ell^+ \ell^- +$ two jets, which arises when the scalar
474: particle decays to a pair of light quarks or gluons\footnote{We exclude
475: $\tau^\pm$ decays because these produce narrow jets which can be identified
476: as $\tau^\pm$ with 80-90\% efficiency.}, which then undergo fragmentation to
477: form a pair of hadronic jets.
478: Clearly, for a Higgs boson, this final state will receive contributions
479: mainly from the decays $H^0 \to b\bar b$ and $H^0 \to c \bar c$, with a
480: minuscule contribution due to $H^0 \to gg$. However, the radion decay
481: will have a much larger contribution from the $gg$ mode, and hence the
482: overall branching ratio to jets should be somewhat higher.
483:
484: \item $e^+ e^- \goes \ell^+ \ell^- + b \bar b$, which simply means that
485: the final state in the above contains two tagged $b$-jets. The decay
486: width for $\Phi \to b \bar b$ is roughly the same as that for $H^0 \to b
487: \bar b$ when the masses and couplings are the same. However, the presence
488: of the two-gluon decay mode makes the branching ratio for the $b \bar b$
489: mode fall quite a bit as compared to that for the Higgs boson when the
490: radion is light. Of course, the $b \bar b$ cross-section will have to be
491: convoluted with efficiency factors, which we take\cite{BTAG} to be
492: $\eta_b = 0.45$ for each tagged $b$-quark.
493: \eenum
494: For both kinds of final states, the cross-section will be proportional
495: to, respectively, $v_{ew}^{-2}$ and $\Lambda_\Phi^{-2}$, and a direct
496: comparison between the Higgs boson and the radion is meaningful only if
497: these match, i.e. $\Lambda_\Phi = v_{ew} \simeq 246$~GeV --- a
498: possibility which, though not ruled out, may not, in general, be
499: realised. However, if we consider the {\sl ratio} of the two processes,
500: viz.
501: $\frac{\sigma(e^+ e^- \goes \ell^+ \ell^- + {\rm two~jets})}
502: {\sigma(e^+ e^- \goes \ell^+ \ell^- + b \bar b)}$
503: the dependence on the vev cancels out and the differences between the two
504: cases are, therefore, more robust. In fact, the underlying scalar
505: production process being the same, this ratio is more-or-less equal to
506: the ratio of the branching fractions $\frac{B(H/\Phi \to {\rm
507: two~jets)}}{B(H/\Phi \to b \bar b)}$, the only difference being due to
508: efficiency factors.
509:
510: % ------------------------------------------------------------------
511: \begin{center}
512: \includegraphics[height=2in]{Fig3.eps}
513: \end{center}
514: \noindent {\bf Figure 3}.
515: {\footnotesize\it Cross sections (in fb)
516: for radiative scalar production in association
517: with $Z^0$ bosons, with scalar decay into ($a$) two jets and ($b$) two tagged
518: $b$-jets, as a function of the scalar mass. The solid (red) line denotes the
519: prediction from a radion and the dashed (black) line that from a Higgs boson.
520: We set $\Lambda_\Phi = v_{ew} \simeq 246$~GeV, so that radion and Higgs
521: production cross-sections match. The ratio of the two cross-sections is shown in
522: ($c$).}
523: % ------------------------------------------------------------------
524:
525: \bigskip\noindent
526: In Figure~3 ($a$) and ($b$), we illustrate our results for the two processes
527: discussed above, namely, \\
528: \hspace*{1.5in} ($a$) $e^+e^- \goes \ell^+\ell^-$ + two jets, and \\
529: \hspace*{1.5in} ($b$) $e^+e^- \goes \ell^+\ell^-$ + two $b$-jets. \\
530: The solid (red) line denotes the radion-mediated cross-section, while the
531: dashed (black) line denotes the Higgs-mediated one. In generating the
532: above curves, we have imposed a few kinematic acceptance cuts on the
533: final state particles, viz.,
534: \benum
535: \item The final-state leptons should have transverse momentum
536: $p_T^{(\ell)} > 20$~GeV.
537: \item The final-state leptons should have pseudo-rapidity
538: $\eta^{(\ell)} < 2.0$.
539: \item The final-state jets should be clearly distinguishable by having their
540: thrust axes separated by $\Delta R_{JJ} = \sqrt{\Delta \eta_{JJ}^2 +
541: \Delta \phi_{JJ}^2} > 0.4$, which is the usual criterion adopted at,
542: for example, the LEP and Tevatron colliders.\footnote{In a parton-level
543: calculation, this is simply implemented by calculating $\Delta R$ for the
544: parent partons, without using a fragmentation algorithm.}
545: \item The final-state jets should have transverse momentum
546: $p_T^{(J)} > 10$~GeV.
547: \item The final-state jets should have pseudo-rapidity $\eta^{(J)} < 2.0$.
548: \eenum
549: The $b$-tagging efficiency has been taken to be 45\%, which is consistent
550: with the LEP value and is probably a conservative estimate than
551: otherwise. It should be noted that the graphs show the excess
552: cross-section after removing the non-Higgs part of the Standard Model
553: contributions (such as $e^+e^- \goes ZZ^*$, etc.). These lead to a large
554: SM four-fermion background, which is, however, easily reducible by
555: selecting only events corresponding to peaks in the $\ell^+\ell^-$ and
556: dijet ($b \bar b$) invariant masses. We have not exhibited the background
557: analysis in this work because we wish to focus on the {\it distinction}
558: between the two types of scalar resonances, rather than the mere
559: detection of a scalar resonance (for which several discussions are
560: already available in the literature).
561:
562: \bigskip\noindent
563: A glance at the cross-sections exhibited in Figure 3 ($a$) and ($b$) will
564: make it clear that for scalar masses well below the $ZZ$-threshold, the
565: cross-sections for $\ell^+\ell^-$ plus two jets final state are almost
566: identical in the two cases, but there is considerable difference if we
567: tag $b$-jets for scalar masses up to at least 150 GeV. Above the
568: $ZZ$-threshold there is again significant difference in the total
569: cross-section, which may be attributed to the enhanced decays of the
570: radion to gluon jets. Noting that the plots are semi-logarithmic in
571: nature, the deviations between the two cases are quite large.
572: Interestingly, the two processes complement each other in the sense that
573: each shows a deviation in the mass region where the other process does
574: not. This shows up very clearly in Figure 3~($c$), where the ratio of the
575: two cross-sections is plotted and there is a very large deviation between
576: the two cases all through the mass range shown. We thus have a simple and
577: robust method of distinction between production of the two kinds of
578: scalar particle: simply measure the cross sections for $e^+e^- \goes
579: \ell^+\ell^-$ + two jets and for $e^+e^- \goes \ell^+\ell^-$ + two
580: $b$-jets and compute the ratio. A large ratio ($>10$) indicates a radion,
581: while a smaller ratio ($5 - 10$) indicates a Higgs particle. Of course,
582: if the radion vev $\Lambda_\Phi$ is very large, the radion effectively
583: decouples from the Standard Model fields and, though Figure 3 ($c$)
584: remains unchanged, the radion production cross-sections shown in
585: Figure~3($a,b$) dwindle accordingly, so that at some stage they become
586: impossible to measure. This case -- though not improbable -- is not the
587: point of interest to us in this work.
588:
589: \bigskip\noindent
590: To summarise, then, this work consists of two parts. In the first part,
591: we have correctly computed the decay width of a radion into two photons
592: or into two gluons using the non-Higgs-like coupling of the radion to
593: off-shell fermions (specifically to off-shell top quarks, in this case).
594: This leads to modest changes in the two-photon branching ratios. However,
595: we predict large changes in the $b\bar b$ branching ratio for light
596: radions and to the overall dijet branching ratio for heavy radions, both
597: occurring because of a greatly-enhanced two-gluon decay mode. Using these
598: results, we compute a `radion-strahlung' process at a 1~TeV linear
599: $e^+e^-$ collider and show that the processes $e^+e^- \goes \ell^+\ell^-$
600: + two jets, with and without $b$-tagging, may be used to distinguish
601: signals from the on-shell production of a Higgs boson from those arising
602: from a genuine radion production event. The ratio of the two cases is a
603: robust method to distinguish between the two cases, even if the radion
604: vev $\Lambda_\Phi$ is not well-determined.
605:
606: \bigskip\noindent
607: Before ending, it needs to be mentioned that we have not taken into
608: account the possibility\cite{MIX} of {\sl mixing} between the radion and
609: the Higgs boson. This is permissible, and, even if precluded at the
610: tree-level, will be generated by quantum corrections, because the radion
611: and the Higgs boson carry the same set of gauge quantum numbers. In this
612: case, however, it is hardly meaningful to talk of the Higgs boson and the
613: radion separately --- there will just be two scalar states of disparate
614: masses, and with couplings scaled suitably by a mixing angle $\xi$. Our
615: work is, therefore, relevant principally in the limit $\xi \to 0$.
616: However, it is worth mentioning that even if $\xi$ is finite, the
617: calculation of the decay widths of the new scalar states to a photon pair
618: or a gluon pair, will be along the lines indicated in the present work,
619: and hence, our efforts will not have been entirely in vain.
620:
621: \bq\noindent\small
622: The authors acknowledge many useful discussions with the late Uma Mahanta, and
623: wish to express a deep sense of loss at his passing. This work is, therefore,
624: dedicated to his memory.
625: \eq
626: \normalsize
627: %------------------------------------------------------------------------%
628:
629: %------------------------------------------------------------------------%
630: % Bibliography %
631: %------------------------------------------------------------------------%
632: %\newpage
633: \begin{thebibliography}{99}
634: \bibitem{RS1}
635: L.~Randall and R.~Sundrum,~ {\it Phys. Rev. Lett.}~{\bf 83},~3370~(1999).
636: \bibitem{ADD}
637: N.~Arkani-Hamed, S.~Dimopoulos and G.R.~Dvali,~{\it Phys. Lett.} {\bf B429},
638: ~263~(1998);
639: I.~Antoniadis, N.~Arkani-Hamed, S.~Dimopoulos and G.R.~Dvali,~{\it Phys. Lett.}
640: {\bf B436},~257~(1998);
641: N.~Arkani-Hamed, S.~Dimopoulos and G.R.~Dvali,~{\it Phys. Rev.} {\bf D59},
642: ~086004~(1999).
643: \bibitem{GW2}
644: W.D.~Goldberger, M.B.~Wise,~{\it Phys. Rev. Lett.}~{\bf 83},~4922~(1999).
645: \bibitem{GW1}
646: W.~D.~Goldberger, M.B.~Wise,~{\it Phys. Rev.} {\bf D60},~107505~(1999).
647: \bibitem{DHR1}
648: H.~Davoudiasl, J.L.~Hewett and T.G.~Rizzo,~{\it Phys. Rev. Lett.}~{\bf 84},
649: ~2080~(2000).
650: \bibitem{HLZ}
651: T.~Han, J.D.~Lykken and R.J.~Zhang,~{\it Phys. Rev.} {\bf D59},~105006~(1999).
652: \bibitem{lambdapi}
653: H.~Davoudiasl, J.L.~Hewett and T.G.~Rizzo,~{\it Phys. Rev.}~{\bf D63},
654: ~075004~(2001).
655: \bibitem{KCH}
656: Kingman Cheung,~{\it Phys. Rev.} {\bf D63},~056007~(2001).
657: \bibitem{ABJ}
658: See, for example, {\it Current Algebra and Anomalies}, by S.~Treiman and
659: R.~Jackiw, Princeton University Press (1986).
660: \bibitem{UMA1}
661: U.~Mahanta and S.~Rakshit,~{\it Phys. Lett.} {\bf B480},~176~(2000).
662: \bibitem{UMA2}
663: U.~Mahanta and A.~Datta,~{\it Phys. Lett.} {\bf B483},~196~(2000).
664: \bibitem{BP}
665: V.~Barger and R.J.N.~Phillips, {\it Collider Physics}, Addison-Wesley
666: Publishing Company~(1987).
667: \bibitem{huitu}
668: A.~Datta and K.~Huitu,~{\it Phys. Lett.} {\bf B578},~376~(2004).
669: \bibitem{HSTAHL}
670: J.R.~Ellis {\it et al},~{\it Nucl. Phys.} {\bf B106},~292~(1976);
671: B.L.~Ioffe and V.A.~Khoze,~{\it Sov. J. Part. Nucl.} {\bf 9},~50~(1978);
672: B.W.~Lee {\it et al},~{\it Phys. Rev. Lett.} {\bf 38},~883~(1977).
673: \bibitem{BTAG} See, for example, K.~Desch, {\sl Particle Searches at a Linear
674: Collider}, ICHEP2000 (Osaka), 2000.
675: \bibitem{MIX}
676: G.F.~Giudice, R.~Rattazzi and J.D.~Wells,~{\it Nucl. Phys.}
677: {\bf B595},~250~(2001);
678: C.~Csaki, M.~Graesser and G.~Kribs,~{Phys. Rev.} {\bf D63},~065002~(2001).
679: \end{thebibliography}
680: \vfill
681: %------------------------------------------------------------------------%
682: \end{document}
683: %------------------------------------------------------------------------%
684: