1: \documentstyle[preprint,aps,prc,floats,psfig,tighten,epsfig,axodraw,epic]{revtex}
2: \draft
3: \renewcommand{\thefootnote}{\arabic{footnote}}
4: \def\gev{\rm GeV}
5: \def\fbi{\rm fb^{-1}}
6: \def\epem{e^+e^-}
7: \def\mh{m_h^{}}
8: \def\lsim{\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}}}
9: \def\gsim{\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}}}
10: \def\Re{{\cal R}\!e}
11: \def\Im{{\cal I}\!m}
12: %
13: \let\jnfont=\rm
14: \def\NPB#1,{{\jnfont Nucl.\ Phys.\ B }{\bf #1},}
15: \def\PLB#1,{{\jnfont Phys.\ Lett.\ B }{\bf #1},}
16: \def\PRD#1,{{\jnfont Phys.\ Rev.\ D }{\bf #1},}
17: \def\PRL#1,{{\jnfont Phys.\ Rev.\ Lett.\ }{\bf #1},}
18:
19: \begin{document}
20:
21: \draft
22: \renewcommand{\thefootnote}{\arabic{footnote}}
23:
24: \preprint{
25: \vbox{\hbox{\bf MADPH-01-1230}
26: \hbox{\bf WM-01-107}
27: \hbox{\bf hep-ph/0106277}
28: \hbox{June, 2001}}}
29:
30: \title{\bf Search for $t \to ch$ at $\epem$ Linear Colliders}
31: \author{Tao Han\footnote{than@pheno.physics.wisc.edu}
32: and Jing Jiang\footnote{jiang@pheno.physics.wisc.edu}}
33: \address{Department of Physics, University of Wisconsin,\\
34: 1150 University Avenue, Madison, WI 53706, USA}
35: \author{Marc Sher\footnote{sher@physics.wm.edu}}
36: \address{Nuclear and Particle Theory Group, Physics Department, \\
37: College of William and Mary, Williamsburg, VA 23187, USA}
38: \maketitle
39:
40: %\vskip 0.5cm
41:
42: \begin{abstract}
43: We study the rare top-quark decay $t\rightarrow ch$,
44: where $h$ is a generic Higgs boson, at a linear collider.
45: If kinematically accessible,
46: all models contain this decay at some level due to
47: quark flavor mixing. Some models, such as Model III of the
48: two-Higgs doublet model, have a tree-level
49: top-charm-Higgs coupling, and the branching ratio is close to $0.5\%$.
50: Others, such as the MSSM, have a coupling induced at one-loop,
51: and can have a branching ratio in the range of
52: $10^{-5}-5\times 10^{-4}$.
53: We find that a linear collider of $\sqrt{s} = 500$ GeV and
54: a luminosity of $500$ fb$^{-1}$ will begin to be sensitive to
55: this range of the coupling.
56:
57: \end{abstract}
58: \pacs{PACS numbers: 12.60.Fr, 14.80.Cp, 14.65.Ha, 12.15.Mm}
59: \newpage
60: %\baselineskip 24pt
61:
62: \section{Introduction}
63:
64: Two of the most important unanswered puzzles of elementary
65: particle physics are the nature of electroweak symmetry breaking and the
66: origin of flavor. Both are implemented in the
67: Standard Model of particle physics in an {\it ad hoc} manner.
68: Often, extensions of the Standard Model (SM)
69: treat these two problems separately;
70: extended Higgs models do not address the
71: flavor problem and flavor symmetry models do not substantively
72: address the nature of electroweak symmetry breaking.
73: Yet, there are no dynamical models satisfactorily addressing
74: the two puzzles coherently.
75: Since the top quark is so heavy, with a mass on the
76: order of the electroweak symmetry breaking scale, it is
77: tempting to consider the possible special role of the top
78: quark in the electroweak symmetry breaking sector and in
79: the quark flavor sector, and its interactions could shed
80: light on both problems.
81: Flavor-changing-neutral-current (FCNC) decays of the top
82: quark \cite{eilam} could be very sensitive to various
83: extensions of the Standard Model that address both of these puzzles.
84:
85: Most studies of flavor-changing-neutral-current top-quark decays
86: have examined the decay of the top quark into a charmed quark
87: and a gauge boson \cite{tcv} (either a gluon, photon or a $Z$).
88: However, the flavor-changing-neutral-current decay most likely
89: to shed light on the nature of electroweak symmetry breaking
90: is the decay of the top quark into a charmed quark
91: and a Higgs boson \cite{eilam,mele} (or the decay of the Higgs
92: boson into a top and a charmed quark if kinematically preferred).
93: It is this interaction that we study in this paper.
94:
95: All models will contain a top-charm-Higgs (TCH) vertex at some
96: level\footnote{Whether a TCH vertex exists does, in part, depend on
97: nomenclature. Strictly speaking,
98: one should consider the ``Higgs" to be the other member of the
99: isodoublet containing the longitudinal components of the weak vector
100: bosons; after all,
101: it is this field that is associated with spontaneous symmetry
102: breaking. With
103: that definition, the Yukawa couplings are necessarily flavor-diagonal.
104: Here, we use
105: the word ``Higgs" to indicate a scalar field whose mass eigenstates
106: contain some part
107: of the field associated with spontaneous symmetry breaking.}.
108: Even in the Standard
109: Model, a one-loop diagram gives a contribution, but it is GIM-suppressed
110: and extremely small \cite{mele}.
111: In extensions of the Standard Model, however,
112: such a vertex
113: can occur either at tree level or at a non-GIM-suppressed one-loop level.
114:
115: One of the simplest extensions of the Standard Model has a tree-level TCH
116: vertex. This is the general two-Higgs doublet model, referred to as Model
117: III, in
118: which no discrete symmetries are imposed to avoid flavor-changing neutral
119: currents. In this case, the quark Yukawa coupling matrices consist of
120: two parts,
121: $y=y_1+y_2$, where $y_i$ is the coupling to the i'th Higgs doublet. Since
122: diagonalizing $y$ will not, in general, diagonalize $y_1$ and $y_2$, there
123: will be
124: tree-level flavor-changing neutral currents. Cheng and
125: Sher \cite{chengsher} noted
126: that if one wishes to avoid fine-tuning, then the observed structure of
127: the mass
128: and mixing hierarchy suggests that the FCNC vertex $\xi_{ij}\ \bar q_iq_j h$
129: is given by the geometric mean of the Yukawa couplings of
130: the quark fields\footnote{We note that some authors may have
131: used a different normalization for $\lambda_{ij}$ by a factor of
132: $\sqrt{2}$ larger than ours. Since the ansatz is order of magnitude
133: only, either definition is acceptable. Similarly, we have not
134: worried about the running mass effects at different renormalization
135: scales, and we simply take the pole masses for $m_t$ and $m_c$ for
136: our analysis.}
137: \begin{equation}
138: \xi_{ij} = \lambda_{ij}{\sqrt{m_im_j}\over{v}}~,
139: \label{coup}
140: \end{equation}
141: where $v=246$ GeV is the weak scale,
142: and the $\lambda_{ij}$ are naturally all of ${\cal O}(1)$.
143: This ansatz has no conflict with current phenomenological
144: observation, and yet predicts rich physics in the heavy-flavor
145: generations. Under this ansatz, the
146: top-charm-scalar coupling $\xi_{ct}$ is the biggest in the model,
147: and is approximately $0.05$.
148:
149: Model III is fairly general. A number of more specific models have a
150: large TCH coupling as well.
151: Burdman \cite{burdman} studied a topcolor model \cite{hill},
152: which gave a large TCH coupling of the same order as Model III
153: (although his ``Higgs"
154: boson had a mass larger than that of the top quark). A topflavor
155: model \cite{heyuan} analyzed the flavor-changing coupling of the
156: top and
157: charm to a top-pion, finding a TCH coupling that is substantially larger
158: than that
159: of Model III. Recently, a bosonic topcolor model \cite{aranda} was
160: proposed that
161: automatically has the TCH coupling as the {\it largest} fermionic coupling
162: of the
163: Higgs boson. An extensive analysis of models with additional singlet
164: quarks ($Q$) by Higuchi and Yamamoto \cite{higuchi} gives a
165: flavor-changing coupling which is proportional to the product of the $cQ$
166: and
167: $tQ$ mixings; this can be larger than the Model III coupling.
168: In general, in
169: models which treat the top quark differently than other quarks
170: motivated by its weak-scale mass, one might expect TCH coupling of
171: Model III size or larger.
172:
173: There have been numerous studies of Model III with the above ansatz,
174: examining $K,\ B,\ \mu$ and $\tau$ decays, as well as $K^0$,
175: $B^0$ and $D^0$ mixing \cite{sheryuan}. Here, however,
176: we will be concerned with the large TCH coupling only.
177:
178: The rate for the decay $t\rightarrow c h$ was first calculated by
179: Hou \cite{hou92}. For a top-quark mass of $175$ GeV, the
180: branching ratio of $t\rightarrow c h$ to $t\rightarrow b W^+$ is
181:
182: \begin{eqnarray}
183: \label{BR}
184: {\Gamma(t\rightarrow c h)\over \Gamma(t\rightarrow b W^+)}
185: &\approx&\lambda_{ct}^2\ \frac{m_c}{m_t}\ \frac{(1-{m_h^2 \over m_t^2})^2}
186: {(1-\frac{m_W^2}{m_t^2})\,(1-\frac{m_W^2}{m_t^2}-2\,
187: \frac{m_W^4}{m_t^4})} \\ \nonumber
188: &\approx& 0.009\
189: \lambda_{ct}^2\ \left(1-{m_h^2\over m_t^2}\right)^2.
190: \end{eqnarray}
191: This implies that the branching fraction of $t\to ch$ can be typically
192: of $10^{-3}$ or higher for $\lambda_{ct}\sim {\cal O}(1)$.
193: If the Higgs is heavier than the top, one can have
194: $h\rightarrow \bar{t}c+\bar{c}t$, as first discussed
195: in Ref.~\cite{hou92}. This leads to the process $e^+e^-\rightarrow
196: h^0A^0\rightarrow
197: \bar{t}\bar{t}cc+\bar{c}\bar{c}tt$ at a linear collider \cite{houlin},
198: to like-sign top-pair
199: production at the LHC \cite{houlinmayuan}, and to the linear collider process
200: $e^+e^-\rightarrow \bar{\nu}_e\nu_e(\bar{t}c+\bar{c}t)$ \cite{barshalom}.
201: These processes all assume a Higgs boson heavier than the top, however,
202: and recent
203: electroweak precision data \cite{electroweak}, as well as hints from
204: LEP \cite{lep},
205: prefer the scenario of a rather light Higgs boson,
206: making $t\rightarrow c h$
207: kinematically accessible.
208: As one can see from Eq.~(\ref{BR}), the branching ratio for
209: $\lambda_{ct}=1$ and a $115$ GeV Higgs mass is $3 \times 10^{-3}$.
210:
211: If one looks at one-loop processes, there are many more
212: possibilities. Most extensions of the Standard Model
213: will avoid the GIM suppression that makes the
214: coupling so small in the Standard Model \cite{mele}.
215: The most popular extension is the MSSM.
216: In this model, there {\it must} be a Higgs boson lighter
217: than the top quark, and so the decay will occur. An extensive analysis of
218: $t\rightarrow c H$, where
219: $H$ refers to any one of the three neutral scalars in the MSSM, was
220: carried out in Ref.~\cite{yangli}. They show that the dominant
221: contribution comes about from loops with gluinos and squarks, and the
222: branching ratio
223: ranges from
224: $10^{-5}$ to $5 \times 10^{-4}$ over the MSSM parameter space. In an
225: R-parity
226: violating SUSY model, a one-loop contribution can also give \cite{ehyz} a
227: branching ratio as large as $10^{-5}$.
228:
229: We see that models with a tree-level TCH coupling have branching ratios of
230: the order
231: of $10^{-3}-10^{-2}$, and models with a one-loop coupling can have
232: branching ratios
233: of the order of $10^{-5}-10^{-4}$.
234: What level is experimentally detectable at high energy colliders?
235: Recently, discovery
236: limits at the LHC were calculated \cite{aguilar} and are approximately
237: $5\times 10^{-5}$. While the results are quite encouraging, detailed
238: Monte Carlo simulation would be needed to draw a definitive conclusion
239: due to the complicated background issue at hadron colliders.
240: In a linear collider, backgrounds are much more manageable, although
241: signal cross sections are lower.
242: In this paper, we assume that a light Higgs boson will be
243: observed and examine in detail the discovery potential for
244: its FCNC coupling at a linear collider.
245:
246: %%%%%%%
247: %figure 1
248: %%%%%%%
249:
250: \begin{figure}[tb]
251: \begin{center}
252: \begin{picture}(240,100)(0,0)
253: \ArrowLine(10,20)(40,50)
254: \Text(10,12)[]{$e^-$}
255: \ArrowLine(40,50)(10,80)
256: \Text(10,88)[]{$e^+$}
257: \Photon(40,50)(80,50){4}{4}
258: \Text(60,62)[]{$\gamma,Z$}
259: \ArrowLine(110,20)(80,50)
260: \Text(110,12)[]{$\bar{t}$}
261: \ArrowLine(80,50)(110,80)
262: \Text(110,88)[]{$c$}
263: \dashline{5}(95,65)(110,50)
264: \Text(110,42)[]{$h$}
265: %
266: \ArrowLine(130,20)(160,50)
267: \Text(130,12)[]{$e^-$}
268: \ArrowLine(160,50)(130,80)
269: \Text(130,88)[]{$e^+$}
270: \Photon(160,50)(200,50){4}{4}
271: \Text(180,62)[]{$\gamma,Z$}
272: \ArrowLine(230,20)(200,50)
273: \Text(230,12)[]{$\bar{t}$}
274: \ArrowLine(200,50)(230,80)
275: \Text(230,88)[]{$c$}
276: \dashline{5}(215,35)(230,50)
277: \Text(230,58)[]{$h$}
278: \end{picture}
279: \end{center}
280: \caption{Tree level Feymann diagrams for $\epem \to \bar{t}ch$ process.
281: Diagrams for $\epem \to t\bar{c}h$ process are similar.}
282: \label{fey}
283: \end{figure}
284:
285: \section{Top-charm-Higgs coupling at a linear Collider}
286:
287: \subsection{Production Cross Section}
288:
289: Using the coupling constant given in Eq.~(\ref{coup}),
290: we wish to explore the limits
291: obtainable on $\lambda_{ct}$ at a linear collider.
292: We consider the process
293: \begin{equation}
294: \label{tbch}
295: \epem \to \bar{t}ch,\ t\bar{c}h~,
296: \end{equation}
297: where the corresponding Feynman diagrams are shown in Fig.~\ref{fey}.
298: The dominant contribution to process~(\ref{tbch})
299: is from $\epem \to t\bar{t}$ followed by $t\, (\bar t)$ decay into
300: $ch\, (\bar ch)$, namely
301: $\sigma(\epem\to t\bar{t})\times [Br(t \to ch)
302: +Br(\bar{t} \to \bar{c}h)]$.
303: The decay branching ratio is calculated and shown in
304: Fig.~\ref{total}(a), where we see that for $m_h = 120$ GeV and
305: $\lambda_{ct} = 1$, $t \to ch$
306: has a branching ratio of $2.8\times10^{-3}$.
307: At tree level, top-pair production
308: has a total cross section of $580$ fb at the center-of-mass
309: energy $\sqrt s=500$ GeV. Thus each channel will
310: have a cross section of $1.6$ fb.
311: The total cross section of $\epem \to \bar{t}ch$ and $t\bar{c}h$
312: is presented in Fig.~\ref{total}(b) as a function of $\sqrt{s}$,
313: where and henceforth $m_h = 120$ GeV is chosen for illustration.
314: The total cross section scales as $\lambda_{ct}^2$ and is of
315: order of 1 fb for $\lambda_{ct} = 1$.
316:
317: %%%%%%%
318: %figure 2
319: %%%%%%%
320:
321: \begin{figure}[tb]
322: \centerline{\psfig{file=roots.ps,width=5.0in}}
323: \bigskip
324: \caption{(a) Branching ratio of
325: $t \to ch$ decay as a function of $m_h$ and (b) Cross section for
326: $t\bar c h+\bar tch$ production as a function of
327: center-of-mass energy in $\epem$ collisions for $m_h = 120$ GeV.}
328: \label{total}
329: \end{figure}
330:
331: \subsection{Signal And Background}
332:
333: Since we are motivated to consider a light Higgs boson
334: with a mass around 120 GeV,
335: we concentrate on the process in which the Higgs boson
336: decays to $b\bar{b}$ and
337: $t$ decays into $bW$. The $W$ can then undergo either
338: hadronic (2 jets) or leptonic ($\ell^\pm\nu$) decays.
339: The $h \to b\bar{b}$ branching ratio for $m_h = 120$ GeV
340: is about $85\%$.
341: The signal thus consists of the following channels
342: in the final state
343: %
344: \begin{eqnarray}
345: \label{hadr}
346: \epem &\to& b\bar{b}, \, b + 3\,jets \\
347: {\rm or}\quad &\to& b\bar{b}, \, b\, \ell^\pm \nu + 1\,jets.
348: \label{lept}
349: \end{eqnarray}
350: %
351:
352: The primary background to the above signal is from the SM top-pair
353: production $\epem \to t\bar{t} \to \bar{b}W^+ bW^-$, with one $W$
354: decaying hadronically and the other decaying either hadronically or
355: leptonically, depending on which signal of Eqs.~(\ref{hadr})
356: and (\ref{lept}) we are looking at.
357: In order to identify the signal from the background, we
358: require that 3 $b$'s be tagged. The efficiency for a single $b$
359: tagging is taken to be $65\%$ \cite{jackson}. This still does not
360: eliminate the background: For the case that both $W$'s decay
361: hadronically, one out of the four non-$b$ jets from $W$ decay
362: to light quarks may be misidentified
363: as a $b$-jet. We assume the misidentification to be
364: $1\%$ for each jet. Thus for four jets the total misidentification
365: probability is $4\%$. Similarly, for the case that one of the two
366: $W$'s decays leptonically, one of the two non-$b$
367: jets from $W$ decay to light quarks is misidentified as a
368: $b$-jet. The SM background induced by one $W$ decay into $\bar{c}b$
369: is negligible due to the small size of the CKM mixing matrix
370: element $V_{cb}\sim 0.04$. The entries of the first row in
371: Table~\ref{rates} show the signal and background
372: cross sections at $\sqrt{s} = 500$ GeV before applying any
373: kinematical cuts but including decay branching fractions and
374: $b$-tagging efficiencies. We find that the background rate is
375: still overwhelming.
376: %%%%%%%
377: % Table 1
378: %%%%%%%
379: \begin{table}[tb]
380: \begin{center}{
381: \begin{tabular}{@{\hspace*{0.5cm}}ccccc@{\hspace*{0.5cm}}}
382: & signal&bg(hadronic)&bg(leptonic)&bg(total) \\ \hline
383: no cuts & 0.75 & 6.56 & 3.23 & 9.79 \\
384: basic cuts & 0.60 & 0.28 & 0.26 & 0.54 \\
385: addi. cuts & 0.52 & 0.13 & 0.12 & 0.25 \\
386: \end{tabular}
387: \caption[]{Cross sections in fb at $\sqrt s=500$ GeV
388: for signal ($m_h=120$ GeV and $\lambda_{ct} = 1$)
389: and hadronic, leptonic and total backgrounds
390: before and after cuts.}
391: \label{rates}}
392: \end{center}
393: \end{table}
394: %\vskip -0.5cm
395:
396: %%%%%%%%%
397: % figure 3
398: %%%%%%%%%
399:
400: \begin{figure}[tb]
401: \centerline{
402: \psfig{file=crucial.ps,width=5in}}
403: \caption{Normalized distribution for the signal (solid), hadronic
404: background (dashes) and leptonic background (dots) with respect to the
405: reconstructed masses (a) $M_h^{rec}$, (b) $M_t^{rec}$,
406: (c) $M_{\bar{t}}^{rec}$ and (d) $M_W^{rec}$,
407: with $\sqrt s=500$ GeV and $m_h=120$ GeV.}
408: \label{crucial}
409: \end{figure}
410:
411: We next try to make use of the distinctive signal kinematics by
412: reconstructing $h$, $t$, $\bar{t}$ and $W$ masses from the
413: final state momenta. To simulate the detector response, we smear
414: the jet energies according to a Gaussian spread
415: \begin{equation}
416: {\Delta E\over E}={45\%\over \sqrt E} \oplus 2\%~,
417: \end{equation}
418: where $\oplus$ denotes a sum in quadrature.
419: For the signal, we first consider the case
420: where $W$ decays into two jets. Of the three tagged $b$'s,
421: we choose two $b$'s whose invariant mass is closest to $m_h$ and identify
422: them as the $b$'s coming from $h$ decay ($M^{rec}_h$).
423: Then from the three light jets we pick the jet that,
424: together with $b\bar{b}$ selected above, produces an invariant mass
425: ($M_t^{rec}$) that is closest to $m_t$.
426: This jet is now identified as ${c}$ from
427: $t\to {c} b\bar b$ decay. The remaining two jets can be used to
428: construct the $W$ mass $M^{rec}_W$ and with the third $b$,
429: we have $M^{rec}_{\bar{t}}$.
430: If $W$ decays into $l\nu$, it is simpler. We simply need to choose
431: two of the three $b$-jets to construct $M^{rec}_h$. The only non-$b$
432: jet is identified as ${c}$. We get $M^{rec}_W$ from the momenta of
433: $\ell^\pm$ and missing energy assigned to $\nu$,
434: after adding the momentum of the third $b$, we have $M^{rec}_{\bar{t}}$.
435: For the background in which $W$ decays hadronically, we randomly pick one
436: from the 4 non-$b$ jets and assume it to be misidentified as a $b$-jet.
437: Then we follow the same construction as for the signals,
438: we will get $M_h^{rec}$, $M^{rec}_t$, $M^{rec}_{\bar{t}}$ and
439: $M^{rec}_W$. For the background when $W$ decays
440: leptonically, we randomly choose one of the two non-$b$ jets to be
441: misidentified as $b$ and again follow the analysis for the signals to
442: get the four reconstructed invariant masses. The normalized differential
443: distribution with respect to these reconstructed masses are shown in
444: Fig.~\ref{crucial} for signals and backgrounds. The characteristic
445: difference between the signal and background is evident from
446: these mass distributions. The first crucial reconstructed mass is
447: $M^{rec}_h$. A peak structure in this variable provides
448: the confidence of the signal
449: observation and gives the discrimination power between the signal
450: and background. The wide width is due to the jet-energy smearing
451: discussed earlier.
452: The other distinctive mass variable is $M^{rec}_t=M_{bbc}$,
453: which reconstructs $m_t$ for the signal and no structure for the
454: background due to the incorrect jet combination. Similarly,
455: $M^{rec}_{\bar t}\approx m_t$ for the signal but it spreads out
456: for the background, as seen in Fig.~\ref{crucial}(c).
457: Those distributions
458: motivate us to device more judicial kinematical cuts.
459: Our basic kinematical cuts are listed in Table~\ref{cuts}
460: based on these four reconstructed masses.
461:
462: %%%%%%%%
463: % table 2
464: %%%%%%%%
465: \begin{table}[tb]
466: \begin{center}
467: \begin{minipage}{4in}
468: \begin{center}{
469: \begin{tabular}{@{\hspace*{1.5cm}}cc@{\hspace*{1.5cm}}}
470: %\begin{tabular}{cc}
471: basic cuts & additional cuts \\ \hline
472: $110 < M_h^{rec} < 130$ & $20 < E_c^{lab} < 130$ \\
473: $160 < M_t^{rec},\ M_{\bar{t}}^{rec} < 190$ & $40 < E_c^{rest}<60$\\
474: $65 < M_W^{rec} < 95$ & $E_b^{min} < 120,\ 80 < E_b^{max}$ \\
475: \end{tabular}
476: \caption{The basic and additional kinematical cuts applied
477: to our signal-to-background optimization (all values in GeV).}
478: \label{cuts}}
479: \end{center}
480: \end{minipage}
481: \end{center}
482: \end{table}
483:
484: %\vskip -0.5cm
485:
486: Additional cuts can be applied to further
487: increase the signal-to-background ratio. We first notice that
488: the charm-jet energy from the top decay is monotonic in the top-rest
489: frame as a result of a two-body decay
490: \begin{equation}
491: E_c^{rest}={m_t\over 2}(1-{m_h^2\over m_t^2}),
492: \label{ec}
493: \end{equation}
494: which is about 50 GeV for $m_h=120$ GeV. In contrast,
495: the faked charm jet from $W$ decay will
496: have an energy more spread out.
497: The normalized distributions are shown in Fig.~\ref{addi}.
498: Figure \ref{addi}(a) is
499: the $E_c$ distribution in the $e^+e^--$lab frame which presents
500: a nearly clear range due to the Lorentz boost of the top quark
501: motion, where the sharp end-point in $E_c$ is sensitive
502: to $m_h$, that may provide an independent kinematical
503: determination for $m_h$.
504: Figure \ref{addi}(b) shows the $E_c$ distribution
505: in the top-quark rest frame based on the $t\to b\bar bc$
506: reconstruction. We see that $E_c$ is monotonic near 50 GeV
507: modulus to the jet-energy resolution. This provides a good
508: discriminator for the signal and background. In fact,
509: the reconstructed Higgs boson energy in the top-quark
510: rest frame should be also monotonic $E_h^{rest}={m_t\over 2}
511: (1+m_h^2/m_t^2)$, but it is correlated with $E_c^{rest}$.
512: We also look at the two $b$-jets coming from $h$
513: decay and examine the harder and softer
514: ones of the two in energy, separately, as illustrated
515: in Figs.~\ref{addi}(c) and \ref{addi}(d).
516: The optimal cuts are given in Table~\ref{cuts}.
517: After applying the basic and additional cuts
518: the signal-to-background ratio improves significantly,
519: as summarized in the last two rows of Table~\ref{rates}.
520:
521: %%%%%%%
522: %Figure 4
523: %%%%%%%
524:
525: \begin{figure}[tb]
526: \centerline{\psfig{file=addi.ps,width=5in}}
527: \caption{Normalized distributions for signal (solid), hadronic
528: background (dashes) and leptonic backgrounds (dots) with respect to
529: (a) $E_c^{lab}$, (b) $E_c^{rest}$, (c) $E_b^{min}$
530: and (d) $E_b^{max}$, with $\sqrt s=500$ GeV and $m_h=120$ GeV.}
531: \label{addi}
532: \end{figure}
533:
534: \subsection{Sensitivity to the $tch$ coupling}
535: Given the efficient signal identification and substantial background
536: suppression achieved in the previous section,
537: we now estimate the sensitivity to the FCNC $tch$
538: coupling from this reaction using Gaussian statistics, which is applicable
539: for large signal event samples. We define the statistical significance by
540: %
541: \begin{equation}
542: \sigma = {N_S\over \sqrt{N_S+N_B}},
543: \end{equation}
544: %
545: with $N_S$ and $N_B$ being the number of signal and background
546: events. We note that $\sigma=3$ (called 3$\sigma$) approximately
547: corresponds to a $99\%$ confidence level.
548: Figure~\ref{sensi} presents the 3$\sigma$ (2$\sigma$) sensitivity to the FCNC
549: couplings by the solid (dashed) curve as a function of integrated luminosity
550: for $\sqrt s=500$ GeV and $m_h=120$ GeV. Recall that at
551: $\lambda_{ct} =1$, the $t \to ch$ branching ratio is
552: about $2.8\times10^{-3}$. Such an order of magnitude can be
553: anticipated in models with tree-level FCNC couplings, and
554: can be easily observed at a LC with an integrated luminosity
555: of less than 40 fb$^{-1}$. As $\lambda_{ct}$ varies,
556: the branching ratio scales like $2.8\times10^{-3} \lambda_{ct}^2$.
557: For 1-loop induced FCNC decays such as in SUSY models, the
558: branching ratios can be about $10^{-5} - 5\times 10^{-4}$,
559: corresponding to $\lambda_{ct}$ of $0.06-0.4$. A linear collider
560: with $500\ {\rm fb}^{-1}$ integrated luminosity
561: will begin to be sensitive to this range of the coupling
562: of $\lambda_{ct}\approx 0.4\ (0.3)$ at a 3$\sigma$ (2$\sigma$)
563: level, but a higher luminosity will be needed to extend the
564: coverage of the parameter space to a level about 0.2.
565:
566: %%%%%%%
567: % Figure 5
568: %%%%%%%
569:
570: \begin{figure}[tb]
571: \centerline{\psfig{file=sensi.ps,width=3in}}
572: \caption{3$\sigma$ sensitivity (solid) to the FCNC $tch$ couplings at
573: $\sqrt s=500$ GeV for $m_h=120$ GeV as a function of integrated
574: luminosity. The dashed curve is for $2\sigma$ sensitivity.}
575: \label{sensi}
576: \end{figure}
577:
578: \section{Conclusions}
579:
580: Models with a top-charm-Higgs coupling come in two categories: Those with
581: a tree-level coupling and those with a coupling induced at one-loop. In the
582: former case, such as the Model III two-Higgs doublet model, one expects
583: $\lambda_{ct}$ to be of ${\cal O}(1)$.
584: If the Higgs is somewhat lighter than the top quark,
585: then the $t\rightarrow ch$ decay will be detected within the first
586: $40\ {\rm fb}^{-1}$ of running of a linear collider,
587: as indicated in Fig.~\ref{sensi}.
588: In the latter case, such as supersymmetric models (in which we {\it know}
589: that one of the Higgs bosons will be sufficiently light),
590: the branching ratios can be in the range of
591: $10^{-5}$ to a few times $10^{-4}$.
592: This corresponds to a $\lambda_{ct}$ of $0.06-0.4$. A linear collider
593: with $500\ {\rm fb}^{-1}$ integrated luminosity
594: will begin to be sensitive to this range of
595: the coupling at a 3$\sigma$ level,
596: and higher luminosity will be needed to substantially extend the
597: coverage of the parameter space.
598:
599: \vskip 0.2cm
600: {\it Acknowledgments}:
601: The work of T.H. and J.J. was supported in part by
602: a DOE grant No.~DE-FG02-95ER40896 and in part by
603: the Wisconsin Alumni Research Foundation. The work of M.S. was supported
604: in part by an NSF grant No.~PHY-9900657.
605:
606: \begin{thebibliography}{99}
607:
608: \bibitem{eilam} G. Eilam, J.L. Hewett and A. Soni, Phys. Rev. D{\bf 44},
609: 1473 (1991),
610: erratum: Phys. Rev. D{\bf 59}, 039901 (1999).
611:
612: \bibitem{tcv}
613: T.~Han, R.~D.~Peccei, and X.~Zhang, \NPB454, 527 (1995);
614: %%CITATION = NUPHA,B454,527;%%
615: T.~Han, K.~Whisnant, B.-L.~Young, and X.~Zhang, \PRD55, 7241 (1997);
616: \PLB385, 311 (1996);
617: %%CITATION = PHRVA,D55,7241;%%
618: %%CITATION = PHLTA,B385,311;%%
619: M.~Hosch, K.~Whisnant, and B.-L.~Young, \PRD56, 5725 (1997);
620: %%CITATION = PHRVA,D56,5725;%%
621: E. Malkawi and T. Tait, \PRD54, 5758 (1996);
622: %%CITATION = PHRVA,D54,5758;%%
623: T. Tait and C.P. Yuan, \PRD55, 7300 (1997);
624: %%CITATION = PHRVA,D55,7300;%%
625: T. Han, M. Hosch, K. Whisnant, B.-L. Young and X. Zhang, \PRD58, 073008 (1998);
626: %%CITATION = PHRVA,D58,073008;%%
627: F. del Aguila and J.A. Aguilar-Saavedra, \PLB462, 310 (1999);
628: %%CITATION = PHLTA,B462,310;%%
629: F. del Aguila and J.A. Aguilar-Saavedra, \NPB576, 56 (2000).
630: %%CITATION = NUPHA,B576,56;%%
631:
632: \bibitem{mele} B. Mele, S. Petrarca and A. Soddu, Phys. Lett. B{\bf 435},
633: 401
634: (1998).
635:
636: \bibitem{chengsher}T.P. Cheng and M. Sher, Phys. Rev. D{\bf 35}, 3484
637: (1987).
638:
639: \bibitem{burdman}G. Burdman, Phys. Rev. Lett. {\bf 83}, 2888 (1999).
640:
641: \bibitem{hill} C.T. Hill, Phys. Lett. B {\bf 345}, 483(1995).
642:
643: \bibitem{heyuan} H.-J. He and C.-P. Yuan, Phys. Rev. Lett. {\bf 83}, 28
644: (1999); H.-J. He, T. Tait and C.-P. Yuan, Phys. Rev. D{\bf
645: 62}, 011702
646: (2000).
647:
648: \bibitem{aranda}A. Aranda and C. Carone, Phys. Lett. B{\bf 488}, 351
649: (2000).
650:
651: \bibitem{higuchi} K. Higuchi and K. Yamamoto, Phys. Rev. D{\bf 62}, 073005
652: (2000).
653:
654: \bibitem{sheryuan}M. Sher and Y. Yuan, Phys. Rev. D{\bf 44}, 1461 (1991);
655: D. Chang, W.S. Hou and W.Y. Keung, Phys. Rev. D{\bf 48}, 217 (1993);
656: D. Atwood, L. Reina and A. Soni, Phys. Rev. Lett. {\bf 75}, 3800
657: (1995);
658: Y. Grossman, Z. Ligeti and E. Nardi, Nucl. Phys. B{\bf 465}, 369 (1996);
659: D. Atwood, L. Reina and A. Soni, Phys. Rev. D{\bf 54}, 3296 (1996);
660: S.Y. Choi, C.S. Kim, Y.J. Kwon and S.H. Lee, Phys. Rev D{\bf 57}, 7023
661: (1998);
662: G. Lopez Castro, R. Martinez and J.H. Munoz, Phys. Rev. D{\bf 58} 033003
663: (1998);
664: D. Bowser-Chao, K. Cheung and W.Y. Keung, Phys. Rev. D{\bf 59}, 115006
665: (1999);
666: K. Huitu, C.-D. Lu, P. Singer and D.-X. Zhang, Phys. Lett. B{\bf 445}
667: 394 (1999);
668: T.M. Aliev and M. Savci, Phys. Rev. D{\bf 60} 014005 (1999), J.
669: Phys. G{\bf 26}, 997 (2000);
670: S. Bergmann and Y. Nir, JHEP {\bf 9909}, 31 (1999);
671: J. Diaz-Cruz and J.J. Toscano, Phys. Rev. D{\bf 62}, 116005 (2000);
672: Z. Xiao, C.S. Li and K.T. Chao, Phys. Rev. D{\bf 62}, 094008
673: (2000); Phys. Rev. D{\bf 63}, 074005 (2001);
674: S. Fajfer and P. Singer, Phys. Rev. D{\bf 62}, 117702 (2000).
675:
676: \bibitem{hou92}W.-S. Hou, Phys. Lett. B{\bf 296}, 179 (1992).
677:
678: \bibitem{houlin}W.-S. Hou and G.-L. Lin, Phys. Lett. B{\bf 379}, 261
679: (1996).
680:
681: \bibitem{houlinmayuan}W.-S. Hou, G.-L. Lin, C.-Y. Ma and C.-P. Yuan, Phys.
682: Lett.
683: B{\bf 409}, 344 (1997).
684:
685: \bibitem{barshalom}S. Bar-Shalom, G. Eilam, A. Soni and J. Wudka, Phys.
686: Rev. Lett. {\bf 79}, 1217 (1997);
687: W.-S. Hou, G.-L. Lin and C.-Y. Ma, Phys. Rev.
688: D{\bf 56}, 7434 (1997).
689:
690: \bibitem{electroweak} J. Erler and P. Langacker,
691: Eur. Phys. J. {\bf C15}, 95 (2000).
692:
693: \bibitem{lep} A.N. Okpara, in the proceedings of
694: {\it 36th Rencontres de Moriond on QCD and Hadronic Interactions,
695: Les Arcs, France, 17-24 Mar 2001}, hep-ph/0105151.
696:
697: \bibitem{yangli} J. M. Yang and C.S. Li, Phys. Rev. D{\bf 49}, 3412
698: (1994);
699: J. Guasch and J. Sola, Nucl. Phys. B{\bf 562}, 3 (1999);
700: S. Bejar, J. Guasch and J. Sola, hep-ph/0101294.
701:
702: \bibitem{ehyz} G. Eilam, A. Gemintern,
703: T. Han, J.M. Yang and X. Zhang, hep-ph/0102037.
704:
705: \bibitem{jackson}
706: C. Damerell and D. Jackson,
707: in {\it Proceedings of the 1996
708: DPF/DPB Summer Study on New Directions for High Energy Physics}.
709: \bibitem{aguilar} J.A. Aguilar-Saavedra and G.C. Branco, Phys. Lett. B{\bf
710: 495}, 347
711: (2000).
712: \end{thebibliography}
713:
714: \end{document}
715:
716:
717:
718:
719:
720:
721:
722:
723: