0812.4609/a.tex
1: %\documentclass[showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
2: %\documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
3: %\documentclass[showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
4: \documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
5: %\documentclass[preprint,aps]{revtex4}
6: %\documentclass[preprint,aps,draft]{revtex4}
7: %\documentclass[prb]{revtex4}% Physical Review B
8: 
9: \usepackage{epsfig}
10: \usepackage{graphicx}% Include figure files
11: \usepackage{dcolumn}% Align table columns on decimal point
12: \usepackage{bm}% bold math
13: 
14: \newcommand{\ord}{{\cal O}}
15: \def\beq{\begin{equation}}
16: \def\eeq#1{\label{#1}\end{equation}}
17: \def\eeqn{\end{equation}}
18: \newcommand\iden{\leavevmode\hbox{\small1\normalsize\kern-.33em1}}
19: \newcommand{\be}{\begin{equation}}
20: \newcommand{\ee}{\end{equation}}
21: \newcommand{\bea}{\begin{eqnarray}}
22: \newcommand{\eea}{\end{eqnarray}}
23: %\renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
24: %\newcommand{\lbl}[1]{\label{eq:#1}}
25: %\newcommand{ \rf}[1]{(\ref{eq:#1})}
26: \newcommand{\gev}{\mbox{GeV}}
27: \newcommand{\tev}{\mbox{TeV}}
28: 
29: \newcommand{\SM}{\mathrm{SM}}
30: \newcommand{\LHT}{\mathrm{LHT}}
31: \newcommand{\BR}{\mathrm{BR}}
32: %***********************************************************%
33: 
34: \let\jnfont=\rm
35: \def\NPB#1,{{\jnfont Nucl.\ Phys.\ B }{\bf #1},}
36: \def\PLB#1,{{\jnfont Phys.\ Lett.\ B }{\bf #1},}
37: \def\EPJC#1,{{\jnfont Eur.\ Phys.\ Jour.\ C }{\bf #1},}
38: \def\PRD#1,{{\jnfont Phys.\ Rev.\ D }{\bf #1},}
39: \def\PRL#1,{{\jnfont Phys.\ Rev.\ Lett.\ }{\bf #1},}
40: \def\MPLA#1,{{\jnfont Mod.\ Phys.\ Lett.\ A }{\bf #1},}
41: \def\JPG#1,{{\jnfont J.\ Phys.\ G }{\bf #1},}
42: \def\CTP#1,{{\jnfont Commun.\ Theor.\ Phys.\ }{\bf #1},}
43: \def\JHEP#1,{{\jnfont JHEP \ }{\bf #1},}
44: \def\NPPS#1,{{\jnfont Nucl.\ Phys.\ Proc.\ Suppl.\ }{\bf #1},}
45: \def\CPC#1,{{\jnfont Computl.\ Phys.\ Commun.\ }{\bf #1},}
46: \def\FP#1,{{\jnfont Fortsch.\ Phys.\ }{\bf #1},}
47: \def\btt#1{{tt$\backslash$#1}}
48: \def\BibTeX{\rm B{\sc ib}\TeX}
49: 
50: \begin{document}
51: 
52: \preprint{\parbox{1.2in}{\noindent }}
53: 
54: \title{\ \\[10mm]  Higgs boson decays and production via gluon fusion at LHC
55:                    in littlest Higgs models with T-parity}
56: 
57: \author{\ \\[2mm] Lei Wang, Jin Min Yang \\ ~}
58: 
59: \affiliation{Institute of Theoretical Physics, Academia Sinica,
60:              Beijing 100190, China \vspace*{1.5cm}}
61: 
62: \begin{abstract}
63: We study the Higgs boson decays and production via gluon fusion
64: at the LHC as a probe of two typical littlest Higgs models which
65: introduce a top quark partner with different (even and odd) T-parity
66: to cancel the Higgs mass quadratic divergence contributed by
67: the top quark. For each model we consider two different choices for
68: the  down-type quark Yukawa couplings.
69: We first examine the branching ratios of the Higgs boson decays
70: and then study the production via gluon fusion followed by
71: the decay into two photons or two weak gauge bosons.
72: We find that the predictions can be quite different for
73: different models or different choices of down-type quark
74: Yukawa couplings and all these predictions can sizably deviate
75: from the SM predictions. So the Higgs boson processes at the LHC
76: can be a sensitive probe for these littlest Higgs models.
77: \vspace*{1cm}
78: \end{abstract}
79: 
80: \pacs{14.80.Cp,12.60.Fr,11.30.Qc}
81: 
82: \maketitle
83: 
84: \section{Introduction}
85: To solve the fine-tuning problem of the Standard Model (SM), the
86: little Higgs \cite{ref1} is proposed as a kind of
87: electroweak symmetry breaking mechanism accomplished by a naturally
88: light Higgs sector.  The Higgs boson remains light, being protected by
89: the approximate global symmetry and free from one-loop quadratic sensitivity
90: to the cutoff scale.
91: The littlest Higgs model \cite{ref2} provides an economical approach
92: which implements the idea of the little Higgs.
93: However, due to the tree-level mixing of heavy and light mass
94: eigenstates, the electroweak precision tests can give strong
95: constraints on this model \cite{ref3}, which would require raising
96: the mass scale of the new particles to be much higher than TeV and
97: thus reintroduce the fine-tuning in the Higgs potential \cite{ref4}.
98: To tackle this problem, a discrete symmetry called T-parity is
99: proposed \cite{ref5}, which forbids those tree-level contributions
100: to the electroweak observables.
101: In the pioneer version of this model (hereafter called model-I)
102: \cite{ref5}, the T-parity is simply implemented by adding the
103: T-parity images for the original top quark interaction to make the
104: Lagrangian T-invariant. A characteristic prediction of this model is
105: a T-even top partner which cancels the Higgs mass quadratic
106: divergence contributed by the top quark.
107: 
108: An alternative implementation of T-parity has recently been
109: proposed (hereafter called model-II)  \cite{ref6},  where all new
110: particles including the heavy top partner responsible for
111: cancelling the SM one-loop quadratic divergence are odd under
112: T-parity. An obvious virtue of this model is that the spectrum of
113: the third-generation quark sector is simplified \cite{ref6}.
114: 
115: These littlest Higgs models with T-parity (LHT) mainly alter the
116: property of the Higgs boson and hence the hints of these models
117: can be unravelled from various Higgs boson processes
118: \cite{ref12,yuanlone,invisible,beforewang,xfhan}. Since different
119: models have different predictions for Higgs boson processes, it is
120: important to study the models comparatively. In this work we
121: perform such a comparative study for model-I and model-II focusing
122: on the decay branching ratios of the Higgs boson as well as the
123: production at the LHC via gluon fusion  followed by the decay into
124: two photons or two weak gauge bosons. Since both models can have
125: two different choices for the down-type quark Yukawa couplings, we
126: will consider the two choices for each model. In our analysis we
127: will show the predictions of these two models and compare with the
128: SM results.
129: 
130: This work is organized as follows. In Sec. II we recapitulate the LHT
131: models with emphasis on model-II since model-I has been intensively
132: discussed in the literature.
133: Then we perform a comparative study for model-I and
134: model-II focusing on the decay branching ratios of the Higgs boson
135: in Sec. III and the production at the LHC via gluon fusion followed by
136: the decay into two photons or two weak gauge bosons in Sec. IV.
137: Finally, we give our conclusion in Sec. V.
138: 
139: \section{The littlest Higgs model with T-parity}
140: 
141: The original littlest Higgs model \cite{ref2} is based on a
142: non-linear sigma model describing the spontaneous breaking of a
143: global $SU(5)$ down to a global $SO(5)$ at an energy
144: scale $f\sim\ord({\rm TeV})$. The vacuum expectation value of an
145: $SU(5)$ symmetric tensor $\Sigma$ is proportional to \beq \Sigma_0
146: \,=\, \left(\begin{array}{ccc}
147: 0& 0& \iden\\
148: 0& 1& 0\\
149: \iden& 0& 0\\
150: \end{array}\right),
151: \eeq{sigma0} where $\iden$ represents a unit $2\times 2$ matrix. The
152: low energy dynamics of the non-linear sigma model is described in terms of the
153: field
154: \beq
155: \Sigma(x) \,=\, e^{i \Pi/f} \Sigma_0 e^{i \Pi^T/f} \,=\, e^{2i \Pi/f} \Sigma_0
156: \eeq{sigma_def}
157: where
158: \beq \Pi(x)=\sum_{a=1}^{14} \pi^a(x) X^a,
159:  \eeq{pionmatrix}
160: with $\pi^a(x)$ being the Goldstone fields corresponding to 14
161: broken generators $X^a$ for the $SU(5)\to SO(5)$ breaking.
162: 
163: In the pioneer version of littlest Higgs model with T-parity
164: (model-I), the T-parity in the top quark sector is implemented by
165: simply adding the T-parity images of the original interaction to
166: make the Lagrangian T-invariant. Thus, it predicts a T-even
167: top partner which cancels the Higgs mass quadratic divergence
168: contributed by the top quark. Since there are detailed descriptions
169: for this model in the literature \cite{ref5,ref12}, we do not discuss
170: it in detail here. In the following we recapitulate an alternative
171: version of T-parity construction (model-II) \cite{ref6}.
172: 
173: In model-II, to implement T-parity in the fermion sector, it
174: requires the introduction of the mirror fermions. For each SM
175: lepton/quark doublet, under the $SU(2)_1\times SU(2)_2$ gauge
176: symmetry, two fermion doublets $q_1(2,1)$ and $q_2 ({1,2})$ are
177: introduced. They can be embedded into the incomplete representation
178: multiplets $\Psi_1$ and $\Psi_2$ of $SU(5)$. A right-handed $SO(5)$
179: multiplets  $\Psi_R$ transforming nonlinearly under the full $SU(5)$
180: is introduced to give mass to the extra fermions.  The field content
181: can be expressed as
182: \begin{equation}
183: \begin{array}{ccc}
184: \Psi_1=\left(\begin{array}{c} q_1 \\ 0 \\ 0 \end{array}\right)\,,
185: & \Psi_2=\left(\begin{array}{c} 0 \\ 0 \\ q_2
186: \end{array}\right) \,,&
187: \Psi_R=\left(\begin{array}{c} \psi_R \\ \chi_R \\ \tilde{\psi}_R
188: \end{array}\right) ,
189: \end{array}
190: \end{equation}
191: where $q_A=-\sigma_2(u_{L_A},d_{L_A})^T=(id_{L_A},-iu_{L_A})^T$ with
192: $A=1,2$ and $\tilde{\psi}_R=\left( i d'_{R}, -i u'_{R} \right)^{\rm
193: T}$.  The second component of $\psi_R$ is $ -iq_R$. The mirror
194: fermions can obtain $\ord(f)$ mass via
195: \begin{equation} \label{eq5}
196: {\cal L}_{\kappa}=-\kappa_{ij}  f \left(\bar{\Psi}^i_2 \xi
197: +\bar{\Psi}^i_1 \Sigma_0\xi^\dagger \right)\Psi^j_R +h.c.,
198: \end{equation}
199: where $\xi=e^{i\Pi/f}$, $\Omega \equiv {\rm diag}(1,1,-1,1,1)$, and
200: $i,j=1,2,3$ are the generation indices. For simplicity, we assume
201: flavor diagonal and hence we have a universal $\kappa$ in our study.
202: The fields transform under $SU(5)$ as
203: \beq
204:   \Psi_1 \rightarrow V^* \Psi_1\,
205:   ~ \Psi_2 \rightarrow V \Psi_2\,, ~ \Psi_R \rightarrow U\Psi_R,
206:   ~ \xi \rightarrow V\xi U^\dagger=U\xi\Sigma_0 V^{\rm T}\Sigma_0,
207:   ~ \Sigma \rightarrow V\Sigma V^{\rm T} \, ,
208: \eeq{su5}
209: where $V$ denotes the $SU(5)$ rotation, and $U$ is the unbroken
210: $SO(5)$ rotation and is a
211: non-linear representation of the $SU(5)$. Under T-parity the
212: transformation laws are defined as
213: \beq
214:  \Psi_1 \rightarrow -\Omega\Sigma_0 \Psi_2\, ,
215:  ~\Psi_R \rightarrow -\Omega\Psi_R\, ,
216:  ~\xi \rightarrow \Omega \xi^\dagger \Omega\, ,
217: \eeq
218: where the transformation of $\Psi_1$  follows \cite{0811.2891}, and Eq.
219: (\ref{eq5}) has the full $SU(5)$ global symmetry and thus we have
220: $q_1\rightarrow -q_2$ and $\Sigma \rightarrow
221: \tilde{\Sigma}=\Sigma_0 \Omega \Sigma^\dagger \Omega \Sigma_0$ under
222: T-parity. Under the above transformations, the Lagrangian is
223: T-invariant.
224: 
225: The Lagrangian in Eq. (\ref{eq5}) contains the new Higgs boson interactions
226: and the mass terms. For the first and second generations we have
227: \begin{eqnarray} \label{eq8}
228: {\cal L}_{\kappa} &\simeq&-\sqrt{2} \kappa f \left[\bar{d}_{L_-}
229: d'_{R}+\frac{1+c_\xi}{2} \bar{u}_{L_-} u'_R
230: -\frac{1-c_\xi}{2}\bar{u}_{L_-}q_R
231: +\frac{s_\xi}{\sqrt{2}}\bar{u}_{L_+} \chi_R \right] +{\rm h.c.}\, ,
232:  \end{eqnarray}
233: where we ignored the generation indices, and
234:  $c_\xi \equiv \cos\frac{v+h}{\sqrt{2}f}$ and
235: $s_\xi \equiv \sin\frac{v+h}{\sqrt{2}f}$ come from the non-linear
236: sigma model field $\xi$, with  $h$ and $v$ being the neutral Higgs
237: boson field and its vacuum expectation value, respectively. The
238: fermion $u_{L_{-}} = (u_{L_1}+ u_{L_2})/\sqrt{2}$ is T-odd, which
239: together with $u'_R$ gets a mass, and $u_{L_{+}} = (u_{L_1}-
240: u_{L_2})/\sqrt{2}$ is T-even and massless. The same definition also
241: applies to the down-type quarks. The fields $q_R$ and $\chi_R$ can
242: obtain large Dirac masses by introducing additional fields, as
243: discussed in \cite{ref5}. In our study we assume both masses are 3.5
244: TeV. From Eq. (\ref{eq8}) we can see that the first component of the
245: doublet $\tilde{\psi}_R$ does not appear and the T-odd down-type
246: quarks have no tree-level coupling with the Higgs boson.
247: 
248: For the top quark interaction sector, in order to cancel the
249: quadratic divergence of the Higgs mass induced by the top quark, it
250: requires the introduction of additional singlets $U_1$ and $U_2$.
251: One can write down their interaction Lagrangian as \cite{ref6}
252: \begin{equation} \label{eq9}
253: {\cal L}_t= -\frac{\lambda}{2\sqrt{2}}f\epsilon_{ijk} \epsilon_{xy}
254: \left[(\bar{Q}_1)_i \Sigma_{jx} \Sigma_{ky}U_{R_1}-
255: (\bar{Q}_2\Sigma_0\Omega)_i \tilde{\Sigma}_{jx}
256: \tilde{\Sigma}_{ky}U_{R_2}\right] +{\rm h.c.} \, ,
257: \end{equation}
258: where the indices $i,j,k$ run from 1 to 3 whereas $x,y=4,5$, and
259: $Q_1=(q_1,U_{1},0_2)^{\rm T}$ whereas $Q_2=(0_2,U_{2},q_2)^{\rm T}$.
260: Under T-parity these fields transform as $Q_1 \rightarrow
261: -\Omega\Sigma_0 Q_2,~ U_{R_1}\rightarrow U_{R_2}$. Therefore, the
262: T-parity eigenstates are defined as $U_{L_{-}} =
263: (U_{1}-U_{2})/\sqrt{2}$ (T-odd) and $U_{L_{+}} = (U_{1}+
264: U_{2})/\sqrt{2}$ (T-even), and the same definition also applies to
265: the right-handed singlet. Eq. (\ref{eq9}) will introduce mixing
266: between the light T-even and the heavy T-even fermions, which can be
267: removed by the additional interactions:
268: \begin{equation} \label{eq10}
269: {\cal L}'_t= -\frac{\lambda'}{2\sqrt{2}}f\epsilon_{lmn}
270: \epsilon_{rs}\left[(\bar{Q}_2)_l \Sigma'_{mr}
271: \Sigma'_{ns}U_{R_1}-(\bar{Q}_1\Omega\Sigma_0)_l \tilde{\Sigma}'_{mr}
272: \tilde{\Sigma}'_{ns}U_{R_2}\right] +{\rm h.c.} \, ,
273: \end{equation}
274: where $\Sigma'=\Omega \Sigma^\dagger \Omega$,
275: $\Sigma'\rightarrow\tilde{\Sigma}'=\Sigma_0\Sigma\Sigma_0$ under T-parity,
276: and the indices $l$, $m$, $n$ run from 3 to 5 whereas $r$, $s$=1,2.
277: Adding ${\cal L'}_t$ to ${\cal L}_t$ and taking
278: $\lambda'=\lambda$, we can get the following simple expression for
279: the top quark Yukawa coupling sector
280: \begin{eqnarray}
281: {\cal L}_t-{\cal L}'_t&\simeq& -\lambda f \left(s_\Sigma
282: \bar{u}_{L_+}U_{R_+}+\frac{1+c_\Sigma}{\sqrt{2}} \bar{U}_{L_-}
283: U_{R_-} \right)+{\rm h.c.}\, , \label{eq11}
284: \end{eqnarray}
285: where $c_\Sigma \equiv\cos\frac{\sqrt{2}(v+h)}{f}$ and
286: $s_\Sigma\equiv \sin\frac{\sqrt{2}(v+h)}{f}$ arise from the
287: non-linear sigma model field $\Sigma$. The field $U_{L_+}$  together
288: with $\chi_R$ gets a Dirac mass in Eq. (\ref{eq5}). From
289: Eq.(\ref{eq5}) we can get the Higgs boson interactions and the mass
290: terms for the third generation fermions
291: \begin{eqnarray}
292: {\cal L}_{\kappa} &\simeq&-\sqrt{2} \kappa f \left [ \bar{d}_{L_-}
293: d'_{R}+\frac{1+c_\xi}{2} \bar{u}_{L_-} u'_R
294: -\frac{1-c_\xi}{2}\bar{u}_{L_-}q_R \right. \nonumber \\
295: &&\left. -\frac{s_\xi}{\sqrt{2}}\bar{U}_{L_-}q_R
296:   -\frac{s_\xi}{\sqrt{2}}\bar{U}_{L_-}u'_R
297:   +\frac{s_\xi}{\sqrt{2}}\bar{u}_{L_+} \chi_R
298:   +c_\xi\bar{U}_{L_+}\chi_R \right] +{\rm h.c.}. \label{eq12}
299:  \end{eqnarray}
300: The Yukawa couplings of up-type quarks for the first and second
301: generations are given by a similar Lagrangian as for the top quark,
302: but without introducing any extra singlet fields:
303: \begin{equation} \label{eq13}
304: {\cal L}_u= -\frac{\lambda_u}{2\sqrt{2}}f\epsilon_{ijk}
305: \epsilon_{xy} \left[(\bar{\Psi}_1)_i \Sigma_{jx} \Sigma_{ky}-
306: (\bar{\Psi}_2\Sigma_0\Omega)_i \tilde{\Sigma}_{jx}
307: \tilde{\Sigma}_{ky}\right]u_{R}+{\rm h.c.}\, ,
308: \end{equation}
309: where $u_R\rightarrow u_R$ under T-parity. Eq.(\ref{eq13}) contains the
310: following Higgs boson interactions as well as the mass term for
311: up-type quarks of the first and second generations
312: \begin{equation} \label{eq14}
313: {\cal L}_u \simeq -\frac{\lambda_u}{\sqrt{2}} f s_\Sigma u_{L_+}u_R+{\rm h.c.} .
314: \end{equation}
315: After diagonalizing the mass matrix in
316: Eqs.(\ref{eq8},\ref{eq11},\ref{eq12},\ref{eq14}), we can get the
317: mass eigenstates and the Higgs couplings. For each SM fermion
318: doublet, there are $d_-$, $u_-$ , $q$ (T-odd) and $\chi$ (T-even).
319: Besides, the top quark has a T-odd  partner $T_-$ which cancels the
320: one loop quadratic divergence of Higgs mass induced by the top
321: quark.
322: 
323: Higgs boson has the couplings with other particles including
324: down-type quarks, leptons, SM gauge bosons, extra gauge bosons and
325: scalar particles. These couplings are same as in model-I and can be
326: found in \cite{ref12,taoh03}. Here we list the Higgs boson couplings
327: with the down-type quarks and the SM gauge bosons, normalized with
328: the corresponding couplings in the SM,
329: \begin{eqnarray}
330: \frac{g_{hd\bar{d}}}{g_{hd\bar{d}}^{\rm SM}}
331: &\approx&1-\frac{1}{4}\frac{v_{SM}^2}{f^2}+\frac{7}{32}
332: \frac{v_{SM}^4}{f^4} ~~~~{\rm for~Case~A}, \label{Higgs-downA} \nonumber\\
333: &\approx&1-\frac{5}{4}\frac{v_{SM}^2}{f^2}-\frac{17}{32}
334:   \frac{v_{SM}^4}{f^4} ~~~~{\rm for~Case~B},\nonumber\\
335: \frac{g_{hVV}}{g_{hVV}^{\rm SM}} &\approx&
336:   1-\frac{1}{4}\frac{v^2_{SM}}{f^2}-\frac{1}{32}\frac{v^4_{SM}}{f^4}, ~~(V=Z,W),
337: \label{eq15}
338: \end{eqnarray}
339: where
340: $G_{F}=1/(\sqrt{2}v_{sm}^2)$ with $v_{sm}=f\sqrt{1-\cos(\sqrt{2}v/f)}$.
341: The relation of down-type quark couplings also
342: applies to the lepton couplings.
343: 
344: \section{Higgs decay branching ratios in LHT models}
345: In both model-I and model-II, the heavy photon $A_H$ is the lightest
346: T-odd particle with a mass given by
347: \begin{equation}
348: M^2_{A_H} = \frac{{g^\prime}^2 f^2}{5}-\frac{{g^\prime}^2
349: v_{SM}^2}{4}.
350: \end{equation}
351: As discussed in \cite{ref18}, the scale $f$ in model-I may be as low
352: as 500 GeV, and the constraint in model-II is expected to be even
353: weaker \cite{ref6}.  For $f=500$ GeV, $A_H$ has a mass of about 65
354: GeV. Therefore, in addition to the SM decay channels, the new decay
355: $h\rightarrow A_H A_H$ will open for $m_h\geq 2m_{A_H}$ and the
356: partial width is given by \bea \Gamma(h \to A_{H} A_{H}) & = &
357: \frac{g_{hA_H A_H}^{2} m_h^3}{128 \pi m_{A_H}^4}
358:   \sqrt{1-\beta_{A_H}}\left(1-\beta_{A_H}+\frac{3}{4}\beta_{A_H}^2\right),
359: \eea where $\beta_{A_H}=4m_{A_H}^2/m_{h}^2$. Because $A_H$ is
360: stable, there are no off-shell decays $h\rightarrow A^*_{H}A^*_{H}$
361: or $h\rightarrow A_{H}A^*_{H}$. In the LHT models the partial widths
362: of the Higgs decays to the SM particles can be obtained as $\Gamma(h
363: \to XX)= \Gamma(h \to XX)_{SM}(g_{hXX}/g_{hXX}^{SM})^2$ ($X$ denotes
364: a SM particle), where $g_{hXX}/g_{hXX}^{SM}$ is predicted by the LHT
365: models and $\Gamma(h \to XX)_{SM}$ is calculated with the code
366: Hdecay \cite{hdecay} (the relevant higher order QCD and electroweak
367: corrections are considered in this code).
368: 
369: Note that in the LHT models the corrections to the tree-level decays
370: $h\to f \bar{f},WW,ZZ$ are mainly from the suppression of the
371: corresponding couplings. For the loop-induced decay $h \to gg$, in
372: addition to the top quark loops, the loops of new T-even and T-odd
373: quarks also come into play. For the decay $h \to Z\gamma$, the $W$
374: boson loop contribution is dominant \cite{invi19} and thus we only
375: consider the alteration of the Higgs coupling with the $W$ boson.
376: The decay channel $h \to \gamma\gamma$ is a focus of our discussion.
377: In addition to the contributions of top quark and $W$ boson, the new
378: charged heavy fermions, gauge bosons and scalar particles will
379: contribute to the decay $h \to \gamma\gamma$. Following the approach
380: in \cite{taohrr}, we calculate the partial decay width of  $h \to
381: \gamma\gamma$ at one-loop. Because the QCD radiative corrections are
382: rather small \cite{hdecay}, our result is precise enough.
383: 
384: In our calculation we take $\kappa=1$
385: and $\lambda'=\lambda$ in model-II, and $r=1$ in model-I.
386: Our calculations show that the results are not so sensitive to
387: these parameters, but very sensitive to the Higgs mass
388: $m_h$ and breaking scale $f$. We take $100 ~{\rm GeV}\leq m_h\leq 500~{\rm GeV}$
389: and $500~{\rm GeV} \leq f\leq 2 ~{\rm TeV}$.
390: 
391: %%%%fig.1 %%%%%%%%%%%%%%%%%%%%%%%
392: \begin{figure}[htb]
393: \epsfig{file=fig1.ps,width=13cm}
394: \vspace*{-0.5cm}
395: \caption{\small The Higgs decay branching ratios versus the Higgs mass
396:            in model-I and model-II.}
397: \label{mh500}
398:  \end{figure}
399: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
400: %%%%fig.2 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
401: \begin{figure}[htb]
402: \epsfig{file=fig2.ps,width=16cm}
403: \vspace*{-0.5cm}
404: \caption{\small  The Higgs decay branching ratios
405: normalized to the SM predictions in model-I and model-II.
406: For the decay channel $VV$, $V$ can be $Z$ or $W$, while for $f\bar f$,
407: $f$ denotes a down-type quark or lepton.}
408: \label{fa}
409:  \end{figure}
410: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
411: %%%%fig.3 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
412: \begin{figure}[htb]
413: \epsfig{file=fig3.ps,width=16cm}
414: \vspace*{-0.5cm}
415: \caption{\small Same as Fig.\ref{fa}, but for Case B.}
416: \label{fb}
417:  \end{figure}
418: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
419: 
420: In Fig.~\ref{mh500} we plot the Higgs decay branching ratios versus
421: the Higgs boson mass in model-I and model-II. Comparing the left
422: panels with the right panels, we see that the two models predict
423: very different branching ratios for $h\to gg$. Further, comparing
424: the upper panels with the lower panels, we see that for each model
425: the two cases give quite different branching ratios for $h\to gg$ or
426: $h\to \gamma\gamma$. In both models with $f=500$ GeV, the new decay
427: $h\to A_HA_H$ opens up for $m_h \geq 130$ GeV. This decay mode can
428: be dominant and over $70\%$ for $130 ~{\rm GeV}<m_h<150 ~{\rm GeV}$,
429: then it decreases as $m_h$ gets large and become comparable with $h
430: \to WW^*$ at $m_h \simeq 160$ GeV. The reason is that the Higgs
431: coupling with $A_H$ is of the electroweak strength and much
432: larger than the Yukawa coupling of $b$ quark. For $130 ~{\rm
433: GeV}<m_h<150 ~{\rm GeV}$, the decay width of $h\to A_HA_H$ is much
434: larger than the decay $h \to bb$ and the off-shell decay $h\to
435: W W^*$.  Here we fixed $f=500$ GeV and  did not show the
436: dependence on $f$. As $f$ gets larger, the decay $h\to A_HA_H$
437: becomes less important.
438: 
439: In Figs.~\ref{fa} and \ref{fb}, we plot the Higgs decay branching
440: ratios normalized to the SM predictions in model-I and model-II for
441: Case A and Case B, respectively. We see that for a small value of
442: $f$ the two models can predict quite different branching ratios from
443: the SM predictions. As $f$ gets large, the deviation from the SM
444: prediction for each decay mode becomes small and finally reduce to
445: the SM results when $f$ is up to 2 TeV. The deviation from the SM
446: prediction is also sensitive to the Higgs boson mass. Again, the
447: results show that the two models predict quite different branching
448: ratios for $h\to gg$ or $h\to \gamma\gamma$; while for other decay
449: modes the two models give the similar results. Besides, the
450: predictions of various branching ratios in Case A and Case B can be
451: sizably different for $m_h=120GeV$ and a small value of $f$.
452: In the following we give some explanations for the above features:
453: \begin{itemize}
454: \item[(1)] First we explain why the branching ratio of $h\to gg$
455: ($h\to \gamma\gamma$) in model-II is smaller (larger) than in
456: model-I. The main contributions of model-I and model-II to the decay
457: $h\to gg$ are from the loops of the fermions whose couplings to $h$
458: are given by \small
459: \begin{eqnarray}
460: -\frac{m_t}{v}y_t\bar{t}th-\frac{m_T}{v}y_T\bar{T}Th+
461: \sum_{i=1}^{3}
462: -\frac{m_{u_{-}^i}}{v}y_{u_{-}^i}\bar{u}_{-}^iu_{-}^ih
463: -\frac{m_{q^i}}{v}y_{q^i}\bar{q}^iq^i h
464: -\frac{m_{\chi^i}}{v}y_{\chi^i}\bar{\chi}^i\chi^i h
465: \end{eqnarray}
466: \normalsize
467: in model-I and
468: \small
469: \begin{eqnarray}
470: -\frac{m_t}{v}y'_t\bar{t}th
471: -\frac{m_{_{U_{-}}}}{v}y'_{_{U_{-}}}\bar{U}_{-}U_{-}h+
472: \sum_{i=1}^{3}
473: -\frac{m_{u_{-}^i}}{v}y'_{u_{-}^i}\bar{u}_{-}^iu_{-}^i
474: h-\frac{m_{q^i}}{v}y'_{q^i}\bar{q}^iq^i h
475: -\frac{m_{\chi^i}}{v}y'_{\chi^i}\bar{\chi}^i\chi^i h
476: \end{eqnarray}
477: \normalsize in model-II. Here all the particles are the mass
478: eigenstates (the diagonalization of the mass matrix was performed
479: numerically in our analysis). The contributions of these fermion
480: loops are not sensitive to the mass values when the fermion masses
481: are much larger than half of the Higgs boson mass. Hence, the
482: contributions of model-I and model-II  are approximately
483: proportional to $y^2_{_{I}}/v^2$ and $y^2_{_{II}}/v^2$, where
484: $y_{_{I}}$ and $y_{_{II}}$ denote the sum of $y$ and $y'$ in Eqs.
485: (18) and (19), respectively. As $y^2_{_{II}}$ is smaller than
486: $y^2_{_{I}}$ in the parameter space we chose (for example, Table 1
487: shows the values of $y_{_{I}}$ and $y_{_{II}}$ for $f=700$ GeV),
488: the decay width of $h\to gg$ in model-II is thus smaller than in
489: model-I. For the decay $h\to \gamma\gamma$, besides the fermion
490: loops, the boson loops also contribute but with an opposite sign.
491: While the fermion loop contribution in model-II is smaller than in
492: model-I, the contributions of boson loops are equal in both
493: models. The extent of cancellation between fermion and boson loops
494: in model-I is more severe than in model-II. Thus, the decay width
495: of $h \to \gamma\gamma$ in mode-II is larger than in model-I. On
496: the other hand, for the total decay width of the Higgs boson,
497: depending on the value of the Higgs boson mass, it can be
498: dominated by the decay $h\to b\bar{b}$, $h \to VV$ or $h \to A_H
499: A_H$, and each of these decays has the same width in both models.
500: Therefore, the branching ratio of $h\to gg$ ($h\to \gamma\gamma$)
501: in model-II is smaller (larger) than that in model-I.
502: %%%%%%%table 1 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
503: \begin{table}[t]
504: \caption{The values of $y$ and $y'$ in Eqs. (18) and (19) for $f=700$ GeV.}
505: \begin{center}
506: \begin{tabular}{|c|c|c|c|c|c|}
507: \hline $y_t$ & $y_T$ & $y_{u_{-}^1}+y_{u_{-}^2}+y_{u_{-}^3}$&
508: $y_{q_{-}^1}+y_{q_{-}^2}+y_{q_{-}^3}$&
509:  $y_{\chi_{-}^1}+y_{\chi_{-}^2}+y_{\chi_{-}^3}$& $y_{_{I}}$\\
510:  \hline
511: $0.947$ & $-0.036$& $-0.104$&
512:  $0.008$& $0$&0.815\\
513: \hline \hline $y'_t$ & $y'_{_{U_{-}}}$ &
514: $y'_{u_{-}^1}+y'_{u_{-}^2}+y'_{u_{-}^3}$&
515: $y'_{q_{-}^1}+y'_{q_{-}^2}+y'_{q_{-}^3}$&
516:  $y'_{\chi_{-}^1}+y'_{\chi_{-}^2}+y'_{\chi_{-}^3}$& $y_{_{II}}$\\
517:  \hline
518: $0.876$ & $-0.169$& $-0.057$&
519:  $0.001$& $-0.026$&0.625\\
520: \hline
521: \end{tabular}
522: \end{center}
523: \label{discovery_pot}
524: \end{table}
525: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
526: \item[(2)] The reason why the two models give similar results for
527: the decay modes of $c\bar c$ and $f\bar f$ ($f$ is a down-type quark
528: or lepton) is that the two models give respectively the same prediction
529: for the Higgs Yukawa couplings with down-type quarks, leptons, and almost
530: the same prediction for up-type quarks except top quark. Therefore, the
531: branching ratios of these decay channels are similar in both models.
532: \item[(3)] The $hb\bar{b}$ coupling in Case B is
533:            more suppressed than in Case A, as shown in Eq. (15).
534: For $m_h=120$ GeV, the decay $h \to b\bar{b}$ is dominant
535: and the total decay width in Case B is much smaller than
536: in Case A. Therefore, the predictions of various branching
537: ratios in Case B can be sizably different from those in Case A
538: for $m_h=120$ GeV and a small value of $f$.
539: \end{itemize}
540: 
541: \section{The rate $\sigma(gg\to h) \times BR(h\to \gamma\gamma ~{\rm or~} VV)$
542:          at LHC in LHT models}
543: 
544: In the SM the Higgs productions at the LHC are dominated by gluon
545: fusion process. The $h \to \gamma\gamma$ channel shows very good
546: sensitivity in the range $114 {\rm ~GeV}<m_h<140 {\rm ~GeV}$.
547: For $2m_W<m_h<2m_Z$,
548: the decay $h \to WW \to l\nu l\nu$ provides the most sensitive search
549: channel. For $m_H > 130$ GeV (except the
550: interval between $2m_W$ and $2m_Z$), the channel $h \to ZZ^* \to 4l$
551: provides excellent sensitivity \cite{hsearch}. Especially, for
552: the  channel $h \to \gamma\gamma$, with an integrated luminosity 100 fb$^{-1}$
553: (10 fb$^{-1}$) from both ATLAS and CMS,
554: the rate $\sigma(gg \to h)\times BR(h \to \gamma\gamma)$ can be
555: measured to $10\%$ ($30\%$) \cite{yuanlone,lone9}. Once we find a
556: light Higgs boson at the LHC, this channel can provide a test
557: for different models. In the LHT models,
558: $\sigma(gg \to h)$ is strongly correlated with
559: $\Gamma(h \to gg)$, which depend on the same effective coupling
560: $hgg$.  In our results we use $\sigma(gg \to h)$ to denote
561: the hadronic cross section of the Higgs production proceeding
562: through  $gg \to h$ at parton level.
563: We use CTEQ6L \cite{cteq} for parton distributions,
564: with the renormalization scale $\mu_R$ and factorization
565: scale $\mu_F$ chosen to be $\mu_R=\mu_F=m_h$.
566: 
567: %%%%%%%%Fig.4 %%%%%%%%%%%%%%%%%
568: \begin{figure}[htb]
569: \epsfig{file=fig4.ps,width=11.5cm}
570: \vspace*{-0.7cm}
571: \caption{The value of $\sigma(gg \to h) \times BR(h\to \gamma\gamma)$
572: normalized to the SM prediction in model-I and model-II.
573: The curves from bottom to top correspond to $f=500$ GeV,  600 GeV,
574: 700 GeV, 800 GeV, 1 TeV, 2 TeV, respectively.
575: The cross section $\sigma(gg \to h)$ denotes the hadronic cross section
576: proceeding through $gg \to h$.}
577: \label{mhr}
578: \end{figure}
579: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
580: %%%%Fig.5 %%%%%%%%%%%%%%%%%%%%%
581: \begin{figure}[htb]
582: \vspace{0.2cm} \epsfig{file=fig5.ps,width=11.5cm}
583: \vspace*{-0.7cm}
584: \caption{Same as Fig. \ref{mhr}, but for
585: $\sigma(gg \to h) \times Br(h\to VV)$ ($V=Z, W$).}
586: \label{mhv}
587: \end{figure}
588: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
589: 
590: In Figs.~\ref{mhr} and \ref{mhv} we plot respectively the rates of
591: $\sigma(gg \to h) \times BR(h\to \gamma\gamma)$ and $\sigma(gg \to
592: h) \times BR(h\to VV)$ ($V=Z, W$) normalized to the SM predictions
593: in model-I and model-II. We see that compared with the SM
594: predictions the LHT models can suppress the rates sizably for a
595: small value of $f$. As $f$ gets large, the suppression is weakened
596: and finally the results reduce to the SM predictions for a
597: sufficiently large $f$ (about 2 TeV). Further, the two models can
598: give very different predictions. For example,  for $m_h=150$ GeV
599: and $f=500$ GeV, the rates are suppressed to $10^{-1}$ ($10^{-2}$)
600: relative to the SM predictions in model-I (model-II). The
601: deviation of the predictions between the two models can be
602: understood as follows. The production cross section  of $gg \to h$
603: is much smaller in model-II than in model-I since the process $gg
604: \to h$ is strongly correlated with the decay $h \to gg$ (the decay
605: width of $h \to gg$ in  model-II  is much smaller than in model-I,
606: as shown and explained in the preceding section). Although the
607: decay branching ratio of $h\to \gamma\gamma$ is larger in
608: model-II, the suppression of the production cross section of $gg
609: \to h$ is dominant and thus the rate $\sigma(gg \to h) \times
610: BR(h\to \gamma\gamma)$ is smaller in model-II. Besides,
611: Figs.~\ref{mhr} and \ref{mhv} showed that the predictions in Case
612: A and Case B can be sizably different in the range of $100
613: {\rm~GeV}\leq m_h\leq 130 {\rm ~GeV}$ for a small value of $f$.
614: The reason is that the two cases give quite different branching
615: ratios for $h\to \gamma\gamma$ or $h\to gg$ (and thus give
616: different cross sections for $gg\to h$), as shown and explained in
617: the preceding section.
618: 
619: \section{Conclusion}
620: In two typical littlest Higgs models which introduce a top
621: quark partner with different (even and odd) T-parity to cancel the
622: Higgs mass quadratic divergence contributed by the top quark, we
623: calculated the branching ratios of the Higgs boson decays and
624: examined the production at the LHC via gluon fusion followed by the
625: decay into two photons or two weak gauge bosons. For each model we
626: considered two different choices for the down-type quark Yukawa
627: couplings. From our numerical results we obtained the following
628: observations: (i) For the Higgs decays, we found that with $130 {\rm
629: ~GeV}<m_h<150 {\rm ~GeV}$ and $f\simeq 500$ GeV, the new decay $h
630: \to A_H A_H$ can be the dominant mode. The two models can give very
631: different branching ratios from the SM predictions. Further, the
632: predictions between the two models can be quite different for
633: $BR(h\to \gamma\gamma)$ and $BR(h\to gg)$; while for other decay
634: modes both models give the similar predictions; (ii) For the rates
635: $\sigma(gg \to h) \times BR(h\to \gamma\gamma)$ and $\sigma(gg \to
636: h) \times BR(h\to VV)$ ($V=Z, W$) at the LHC, both models can give
637: severe suppression relative to the SM predictions, and the
638: predictions of the two models can also differ significantly; (iii)
639: For each model the two different choices for the  down-type quark
640: Yukawa couplings can also lead to different results. Therefore,
641: these Higgs processes at the LHC may be a sensitive probe for the
642: little Higgs theory and may even provide a way to distinguish the
643: different models or different scenarios.
644: 
645: \section*{Acknowledgement}
646: We thank C.-P. Yuan for discussions.
647: This work was supported  by the National Natural
648: Science Foundation of China (NNSFC) under Nos. 10821504,
649: 10725526 and 10635030.
650: 
651: \begin{thebibliography}{99}
652: \bibitem{ref1} N. Arkani-Hamed, A. G. Cohen, H. Georgi, \PLB513, 232 (2001);
653:                N. Arkani-Hamed, A. G. Cohen, T. Gregoire, J. G. Wacker,
654:                                           \JHEP0208, 020 (2002);
655:                N. Arkani-Hamed, {\it et al.},  \JHEP0208, 021 (2002);
656:                I. Low, W. Skiba, D. Smith, \PRD66, 072001 (2002);
657:                D. E. Kaplan, M. Schmaltz, \JHEP0310, 039 (2003).
658:                %%CITATION = PHLTA,B513,232;%%
659:                %%CITATION = PHRVA,D66,072001;%%
660: \bibitem{ref2}  N. Arkani-Hamed, A. G. Cohen, E. Katz, A. E. Nelson, \JHEP0207, 034 (2002);
661:                 S. Chang, \JHEP0312, 057 (2003);
662:                 T. Han, H. E. Logan, B. McElrath, L. T. Wang, \PRD67, 095004 (2003);
663:                 M. Schmaltz, D. Tucker-smith, Ann. Rev. Nucl. Part. Sci. {\bf55}, 229 (2005).
664:                 %%CITATION = PHRVA,D67,095004;%%
665: \bibitem{ref3}
666:   C.Csaki, J. Hubisz, G. D. Kribs, P. Meade, J. Terning, \PRD67, 115002 (2003);
667:   J. L. Hewett, F. J. Petriello, T. G. Rizzo, \JHEP0310, 062 (2003);
668:   C. Csaki, J. Hubisz, G. D. Kribs, P. Meade, J. Terning, \PRD68, 035009 (2003);
669:   M. C. Chen, S. Dawson, \PRD70, 015003 (2004);
670:   M. C. Chen et al.,  \MPLA21, 621 (2006);
671:   W. Kilian, J. Reuter, \PRD70, 015004 (2004).
672:   %%CITATION = PHRVA,D67,115002;%%
673:   %%CITATION = PHRVA,D68,035009;%%
674:   %%CITATION = PHRVA,D70,015003;%
675:   %%CITATION = PHRVA,D70,015004;%%%
676: \bibitem{ref4} G. Marandella, C. Schappacher and A. Strumia, \PRD72, 035041 (2005).
677:                %%CITATION = PHRVA,D72,035041;%%
678: \bibitem{ref5} H. C. Cheng and I. Low, \JHEP0309, 051 (2003);
679:                                        \JHEP0408, 061 (2004);
680:                I. Low, \JHEP0410, 067 (2004); J. Hubisz, P. Meade, \PRD71, 035016 (2005).
681:                %%CITATION = PHRVA,D71,035016;%%
682: \bibitem{ref6} H. C. Cheng, I. Low and L. T. Wang, \PRD74, 055001 (2006).
683:                %%CITATION = PHRVA,D74,055001;%%
684: \bibitem{ref12} C. R. Chen, K. Tobe, C. P. Yuan, \PLB640, 263 (2006).
685:                %%CITATION = PHLTA,B640,263;%%
686: \bibitem{yuanlone} K. Hsieh, C. P. Yuan, \PRD78,053006 (2008).
687:                %%CITATION = PHRVA,D78,053006;%%
688: \bibitem{beforewang}
689:      C. O. Dib, R. Rosenfeld, A. Zerwekh, \JHEP0605, 074 (2006);
690:      L. Wang {\it et al.}, \PRD76, 017702 (2007); \PRD75, 074006 (2007); \PRD77, 015020 (2008).
691:      %%CITATION = PHRVA,D76,017702;%%
692:      %%CITATION = PHRVA,D75,074006;%%
693:      %%CITATION = PHRVA,D77,015020;%%
694: \bibitem{xfhan} X.F. Han, L. Wang, J. M. Yang, \PRD78, 075017 (2008).
695:                %%CITATION = PHRVA,D78,075017;%%
696: \bibitem{invisible} R. S. Hundi, B. Mukhopadhyaya, A. Nyffeler, \PLB649, 280 (2007).
697:                     %%CITATION = PHLTA,B649,280;%%
698: \bibitem{0811.2891} F. del Aguila, J. I. Illana, M. D. Jenkins, arXiv:0811.2891.
699:                     %%CITATION = ARXIV:0811.2891;%%
700: \bibitem{taoh03}  T. Han {\it et al.},  \PRD67, 095004 (2003).
701:                %%CITATION = PHRVA,D67,095004;%%
702: \bibitem{ref18} J. Hubisz, P. Meade, A. Noble and M. Perelstein, \JHEP01, 135 (2006).
703: \bibitem{hdecay} A. Djouadi, J. Kalinowski, M. Spira, \CPC108, 56 (2006).
704: \bibitem{invi19} See for instance: M. Spira, \FP 46, 203 (1998).
705: \bibitem{taohrr} T. Han {\it et al.}, \PLB563, 191 (2003).
706:                     %%CITATION = PHLTA,B563,191;%%
707: \bibitem{hsearch} R. Goncalo, arXiv:0811.3778.
708: \bibitem{lone9} D. Zeppenfeld {\it et al.}, \PRD62, 013009 (2000).
709:                %%CITATION = PHRVA,D62,013009;%%
710: \bibitem{cteq}  J.~Pumplin {\it et al.}, JHEP {\bf 0602}, 032 (2006).
711:                 %%CITATION = HEP-PH 0512167;%
712: \end{thebibliography}
713: 
714: 
715: \end{document}
716: