hep-ph0702264/new.tex
1: \documentclass[showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
2: \usepackage{epsfig}
3: \usepackage{graphicx}% Include figure files
4: \usepackage{dcolumn}% Align table columns on decimal point
5: \usepackage{bm}% bold math
6: 
7: \let\jnfont=\rm
8: \def\NPB#1,{{\jnfont Nucl.\ Phys.\ B }{\bf #1},}
9: \def\PLB#1,{{\jnfont Phys.\ Lett.\ B }{\bf #1},}
10: \def\EPJC#1,{{\jnfont Eur.\ Phys.\ Jour.\ C }{\bf #1},}
11: \def\PRD#1,{{\jnfont Phys.\ Rev.\ D }{\bf #1},}
12: \def\PRL#1,{{\jnfont Phys.\ Rev.\ Lett.\ }{\bf #1},}
13: \def\MPLA#1,{{\jnfont Mod.\ Phys.\ Lett.\ A }{\bf #1},}
14: \def\JPG#1,{{\jnfont J.\ Phys.\ G}{\bf #1},}
15: \def\CTP#1,{{\jnfont Commun.\ Theor.\ Phys.\ }{\bf #1},}
16: \def\ZPC#1,{{\jnfont Z.\ Phys.\ C }{\bf #1},}
17: \def\JHEP#1,{{\jnfont JHEP \ }{\bf #1},}
18: \def\q_slash{\not{\hbox{\kern-2.1pt $q$}}}
19: \def\p_slash{\not{\hbox{\kern-4.0pt $p$}}}
20: \def\k_slash{\not{\hbox{\kern-2.1pt $k$}}}
21: \def\no{\normalsize}
22: \def\un{\underline}
23: \begin{document}
24: 
25: \preprint{\parbox{1.2in}{\noindent CUMQ/HEP~144 }}
26: 
27: \title{\ \\[10mm]
28: SUSY-induced FCNC top-quark processes at the Large Hadron Collider
29: }
30: 
31: \author{ J. J. Cao$^1$, G. Eilam$^1$, M. Frank$^2$, K. Hikasa$^3$,
32:         G. L. Liu$^4$, I. Turan$^2$, J. M. Yang$^5$ \\~~
33: }
34: \affiliation{
35: $^1$ Physics Department, Technion, 32000 Haifa, Israel\\
36: $^2$ Department of Physics, Concordia University, Montreal, H4B 1R6, Canada\\
37: $^3$ Department of Physics, Tohoku University, Sendai 980-8578, Japan\\
38: $^4$ Service de Physique Theorique CP225, Universite Libre de
39:           Bruxelles, 1050 Brussels, Belgium \\
40: $^5$ Institute of Theoretical Physics, Academia Sinica, Beijing 100080, China
41: }
42: 
43: \begin{abstract}
44: We systematically calculate various flavor-changing neutral-current top-quark processes
45: induced by supersymmetry at the Large Hadron Collider, which include five decay modes and six
46: production channels. To reveal the characteristics  of these
47: processes, we first  compare the dependence of
48: the rates for these channels on the relevant supersymmetric parameters,
49: then we scan the whole parameter space to find their maximal
50: rates, including all the direct and indirect current
51: experimental constraints on the scharm-stop flavor mixings. We
52: find that, under all these constraints, only a few channels,
53:  through $c g \to t $ at parton-level and $t \to c
54: h $, may be observable at the Large Hadron Collider.
55: \end{abstract}
56: 
57: \pacs{14.80.Ly, 11.30.Hv}
58: \maketitle
59: 
60: \section{\bf Introduction}
61: 
62: The study of various flavor-changing neutral-current (FCNC)
63: processes has been  shown to be  very useful.  In particular, as
64: the heaviest fermion in the standard model (SM), the top quark may play a special
65: role in FCNC phenomenology. In  the SM the FCNC interactions of
66: the top quark are extremely suppressed by the GIM mechanism, so
67: that no FCNC top quark rates can reach an observable level at
68: current or future colliders \cite{tcvh-sm,tcgg-sm,eetc-sm}. Thus,
69: the observation of any FCNC top quark process would be a robust
70: evidence for new physics beyond the SM.  Due  to its heaviness,
71: the top quark is very sensitive to new physics. Indeed, several models beyond the
72: SM often predict much larger FCNC top quark
73: interactions \cite{Larios:2006pb}. Such FCNC interactions can induce various
74: top quark production and decay channels, which can be explored in future
75: collider experiments
76: \cite{Aguilar-Saavedra:2004wm,Aguilar-saavedra:linear} and serve
77: as a good probe for new physics.
78: 
79: So far, much effort has been spent on the exploration of the FCNC
80: top quark interactions. On the experimental side, the Tevatron CDF
81: and D0 collaborations have reported interesting bounds on the FCNC
82: top quark decays from Run 1 experiment and will tighten the bounds
83: from the ongoing Run 2 experiments \cite{cdfd0}. On the
84: theoretical side, various FCNC top quark decays and top-charm
85: associated productions at high energy colliders were extensively
86: studied in the SM \cite{tcvh-sm,tcgg-sm,eetc-sm}, the Minimal
87: Supersymmetric Standard Model (MSSM)
88: \cite{tcv-mssm,tch-mssm,eetc-mssm,pptc-mssm,Gad-pptc-mssm} and
89: other new physics models \cite{tc-TC2,tcv-TC2,other}. These
90: studies showed that the SM predictions for such
91: processes are far below the detectable level. However, some new
92: physics can enhance them by several orders of magnitude, which makes
93: them potentially accessible at future colliders.
94: 
95: The Large Hadron Collider (LHC) at CERN will be a powerful
96: machine for studying the top quark properties such as its FCNC
97: interactions since it will produce top quarks copiously. Analysis
98: \cite{Aguilar-Saavedra:2004wm} showed that some FCNC top quark
99: rare decays with branching ratios as low as ${\cal O}(10^{-5})$
100: could be accessible at the LHC. Due to its high energy and high
101: luminosity, the LHC will be the main utility for exploring
102: FCNC top quark production channels \cite{Aguilar-Saavedra:2004wm}.
103: 
104: The minimal supersymmetric standard model (MSSM)  is a leading candidate for new
105: physics beyond the SM and its consequences will be extensively explored at the LHC.
106: In this paper, we throughly
107: investigate various top quark FCNC processes in the
108: framework of the MSSM.  A characteristic feature of the model is that, in addition to the
109: FCNC interactions generated at loop level by the CKM mixing matrix
110: in the SM, it predicts FCNC interactions from soft SUSY breaking
111: terms \cite{susyflavor}. These additional FCNC interactions depend
112: on the squark flavor mixings, and for the case of top quark, they
113: are sensitive to the potentially large mixings between charm
114: squarks (scharms) and top squarks (stops).
115: 
116: In this paper, we will examine six production channels, which proceed through the parton-level:
117: \begin{eqnarray}
118: && g g \to t \bar{c}, \label{pro-1} \\
119: && c g \to t, \\
120: && c g \to t g,  \\
121: && c g \to t Z, \\
122: && c g \to t \gamma,\\
123: && c g \to t h,
124: \label{processes}
125: \end{eqnarray}
126: and five FCNC top quark decay modes:
127: \begin{eqnarray}
128: t & \to & c g,  \label{decy-1}\\
129: t & \to & c g g,\\
130: t & \to & c Z, \\
131: t & \to & c \gamma,\\
132: t & \to & c h,
133: \label{fcncdecays}
134: \end{eqnarray}
135: 
136: Among these, the decays $t \to c g, cgg, c Z, c \gamma, c h$
137: and the production $ g g \to t \bar{c}$ have already been studied
138: in the MSSM  before \cite{tcv-mssm,pptc-mssm,Gad-pptc-mssm}, while the others
139: have not been studied so far. Here we
140: perform a comprehensive study of all these processes in the MSSM
141: for the following purposes:
142: \begin{itemize}
143: \item[(1)] Since in the MSSM all these processes are induced
144: mainly by scharm-stop mixings and each of them involves the same
145: set of SUSY parameters, they are correlated. Although
146: some of them have been studied in the literature, they were treated
147: individually in different papers. Even though a combined analysis has been
148: done in \cite{Aguilar-Saavedra:2004wm} within the framework of effective Lagrangian where the
149: coefficients of all FCNC interactions are independent, such analysis is
150: missing in an explicit model.
151: Therefore, a comprehensive and comparative study of all these processes in an explicit model is necessary.
152: \item[(2)] Through a comparative study of all these
153: channels, we could determine the relative size of their rates. This is
154: useful since the LHC experiments can in principle measure
155: each of them and the pattern of relative rates can be tested. Only
156: by considering all these processes  together, we might know which
157: one has the largest rate and will hopefully  be discovered at
158: the LHC, if the MSSM is the correct framework.
159: \item[(3)] By
160: scrutinizing the dependence of the rates of these transitions on the
161: relevant SUSY parameters, one can determine the most sensitive parameters
162: of the model and then discuss how the future LHC measurements could
163: possibly bound them.
164: 
165: \item[(4)] Performing the scan over the whole parameter
166: space, subject to all the direct and indirect current experimental
167: constraints on the scharm-stop flavor mixings \cite{cao}, the maximal
168: rate for each process can be determined and, in this way, we can
169: pinpoint which ones are hopefully observable at the LHC. Of
170: course, this does not mean that one should give up searching
171: for those low-rate processes at the LHC. As stated above, the LHC
172: measurements can readily place bounds on the sensitive SUSY
173: parameters and such bounds are complementary to the current
174: experimental constraints, most of which are indirect constraints.
175: \end{itemize}
176: This paper is organized as follows. In Sec. II we discuss the
177: possible sources of flavor violation in the MSSM and give the FCNC
178: interaction Lagrangian relevant to our calculations. In Sec. III
179: we introduce a method to calculate various top quark FCNC
180: processes. This method, as will be shown, can greatly simplify our
181: calculations. The predictions of the rates are given in Sec. IV,
182: with emphasis on  illustrating the
183: characteristics  of their dependence on the relevant SUSY
184: parameters. In Sec. V we consider various experimental constraints
185: on the sources of flavor violation and scan the parameter space to
186: find the maximal rates at the LHC.  We  draw our conclusion in
187: Sec. VI.  Finally, we give the expressions for the loop results
188: in the Appendix.
189: 
190: 
191: \section{\bf FCNC interactions in the MSSM}
192: 
193: There are two sources of flavor
194: violation in the MSSM \cite{susyflavor}. The first one arises from the
195: flavor mixings of up-quarks and down-quarks, which are
196: described by the CKM matrix (inherited from the SM).
197: The second one results from the misalignment between the rotations
198: that diagonalize the quark and squark sectors due to the presence of
199: soft SUSY breaking terms. This source can induce large top quark
200: FCNC processes and is the focus of investigation in this paper.
201: 
202: In the super-CKM basis with states ($\tilde u_L$, $\tilde c_L$,
203: $\tilde t_L$, $\tilde u_R$, $\tilde c_R$, $\tilde t_R$) for
204: up-squarks and ($\tilde d_L$, $\tilde s_L$, $\tilde b_L$, $\tilde
205: d_R$, $\tilde s_R$, $\tilde b_R$) for down-squarks, the $6\times
206: 6$ squark mass matrix ${\cal M}^2_{\tilde q}$ ($\tilde q=\tilde u,
207: \tilde d$) takes the form \cite{susyflavor}
208: \begin{eqnarray}
209: {\cal M}^2_{\tilde q}=\left( \begin{array}{ll}
210:   (M^2_{\tilde q})_{LL}+ m_q^2 + \cos 2\beta M_Z^2 ( T_3^q - Q_q
211: s_W^2) \hat{\mbox{\large 1}} &\;\;\;(M^2_{\tilde q})_{LR}- m_q \mu (\tan \beta)^{- 2 T_3^q}\\
212:   (M^2_{\tilde{q}})_{LR}^\dag - m_q \mu (\tan \beta)^{- 2 T_3^q}
213:              &\;\;\;(M^2_{\tilde{q}})_{RR}+ m_q^2 + \cos
214: 2\beta M_Z^2 Q_q s_W^2 \hat{\mbox{\large 1}}  \end{array} \right)
215: , \label{sq-matrix}
216: \end{eqnarray}
217: where the soft mass parameters $(M^2_{\tilde q})_{LL}$, $(M^2_{\tilde
218: q})_{LR}$ and $(M^2_{\tilde q})_{RR}$ are $3 \times 3$ matrices in
219: flavor space, $\hat{\mbox{\large 1}}$ stands for the unit matrix,
220: $m_q$ is the diagonal quark mass matrix, $T_3^q=1/2$ for
221: up-squarks and $T_3^q=-1/2$ for down-squarks, and $\tan \beta
222: =v_2/v_1$ is the ratio of the vacuum expectation values of the
223: Higgs fields. In general, the soft mass parameters are flavor
224: non-diagonal. Since the low energy experimental data, such as
225: $K^0-\bar K^0$, $D^0-\bar D^0$ and $B^0_d-\bar B^0_d$ mixings,
226: require the flavor mixings involving the first generation
227: squarks to be negligibly small \cite{susyflavor}, we only consider
228: the flavor mixings of the second and third generations and
229: parametrize the soft mass parameters as
230: \begin{eqnarray}
231: (M^2_{\tilde{u}})_{LL} &= & \left ( \begin{array}{ccc}
232:     M_{Q_1}^2 &  0                           &  0 \\
233:     0         &  M_{Q_2}^2                   & \delta_{LL} M_{Q_2} M_{Q_3} \\
234:     0         &  \delta_{LL} M_{Q_2} M_{Q_3} & M_{Q_3}^2 \end{array}  \right ), \nonumber \\
235:  (M^2_{\tilde{u}})_{LR} &=&  \left (
236: \begin{array}{ccc}
237: 0     &  0   &  0 \\
238: 0     &  0   & \delta_{LR} M_{Q_2} M_{U_3}\\
239: 0     &  \delta_{RL} M_{U_2} M_{Q_3}  & m_t A_t \end{array}  \right ), \nonumber \\
240:  (M^2_{\tilde{u}})_{RR} &= & (M^2_{\tilde{u}})_{LL}|_{M_{Q_i}^2 \to M_{U_i}^2,~ \delta_{LL} \to
241:  \delta_{RR}}, \label{up squark}
242: \end{eqnarray}
243: for up-type squarks. Similarly, for down-squarks  we have
244: \begin{eqnarray}
245: (M^2_{\tilde{d}})_{LR}& = & \left ( \begin{array}{ccc}
246: 0          &  0                  &  0 \\
247: 0                                &  0   & \delta_{LR}^d M_{Q_2} M_{D_3}\\
248: 0 &  \delta_{RL}^d M_{D_2} M_{Q_3}  & m_b A_b \end{array}  \right
249: ),  \nonumber  \\
250: (M^2_{\tilde{d}})_{RR} &= & (M^2_{\tilde{u}})_{LL}|_{M_{Q_i}^2 \to
251: M_{D_i}^2,~ \delta_{LL} \to \delta_{RR}^d}. \label{delta'-LR}
252: \end{eqnarray}
253: Due to $SU_L(2)$ gauge invariance, $(M^2_{\tilde d })_{LL}$ is
254: given by
255: \begin{eqnarray}
256: (M^2_{\tilde{d}})_{LL} = V_{CKM}^\dag (M_{\tilde{u}}^2)_{LL}
257: V_{CKM}.  \label{SU2}
258: \end{eqnarray}
259: Note that the mixing parameters, $\delta^d$ in the down sector defined in
260: Eq.~(\ref{delta'-LR}) are independent of $\delta$ in the up sector defined in
261: Eq.~(\ref{up squark}), and in general, $\delta_{LR} \neq
262: \delta_{RL}$. For the diagonal elements of left-right mixings in
263: Eq.~(\ref{up squark}) and Eq.~(\ref{delta'-LR}), we only kept the
264: terms of third-family squarks, since we adopted the popular
265: assumption that they are proportional to the corresponding quark
266: masses.
267: 
268: It is clear that the mixing parameters in the squark mass matrices
269: affect both the squark mass and its interactions. For example, in
270: the presence of flavor mixings, squark-quark interactions are given by
271: \begin{eqnarray}
272: V(\bar{q} X \tilde{q}_{\alpha}^{\prime}) \; = \;
273:  \Gamma_{q}^{i \alpha} \;
274:  V(\bar{q} X \tilde{q}^{\prime}_i)~,  \label{interaction}
275: \end{eqnarray}
276: where $V(\bar{q} X \tilde{q}_{\alpha}^{\prime})$ denotes the
277: interaction in squark mass-eigenstates, $V(\bar{q} X
278: \tilde{q}^{\prime}_i)$ is that in the interaction basis, $X$ may
279: be gluino, neutralino or chargino,
280: and $\Gamma_q$ is the unitary matrix which diagonalizes the squark
281: mass matrix. For the convenience in the following discussions,
282: we give the interaction Lagrangian for up-type quarks
283: \cite{susyflavor,mssmfeynmanrule}:
284: \begin{eqnarray}
285: {\cal{L}}_{u \tilde{u} \tilde g}&=& \sum_{i=1}^{3}\sqrt{2} g_s \,
286: T^a_{st} \left[ \bar u^{s}_i \,(\Gamma_U)^{i \alpha} P_L\, \tilde
287: g^a \,\tilde u^{t}_\alpha - \bar u^{s}_i \,(\Gamma_U)^{(i+3)
288: \alpha}\,P_R \,\tilde g^a \,\tilde u^{t}_\alpha + \text{h.c.} \right]\, , \label{interaction1}  \\
289: {\cal{L}}_{u\tilde{u}\tilde{\chi}^{0}}&=&\sum_{n=1}^{4}\sum_{i=1}^{3}
290: \frac{g}{\sqrt{2}} \left\{ \bar{u}_{i}\,N_{n1}^{*}\,\frac{4}{3}
291: \tan \theta _{W} \,P_L\,\tilde{\chi}_{n}^{0}\,(\Gamma_U)^{(i+3)
292: \alpha}\,\tilde{u}_\alpha -\bar{u}_{i}\,N_{n4}^{*}\,
293: \frac{(m_u)_{ij}}{M_W \sin \beta}
294: \,P_L\,\tilde{\chi}_{n}^{0}\,(\Gamma_U)^{j \alpha
295: }\,\tilde{u}_\alpha
296: \right.   \nonumber \\
297: &&- \left.\bar{u}_{i}\, \left( N_{n2}+  \frac{1}{3}N_{n1}\tan
298: \theta _{W}\right)\,P_R \,\tilde{\chi}_{n}^{0}\,(\Gamma_U)^{i
299: \alpha}\,\tilde{u}_\alpha -\bar{u}_{i}\,N_{n4}\,
300: \frac{(m_u)_{ij}}{M_W \sin \beta}
301: \,P_R\,\tilde{\chi}_{n}^{0}\,(\Gamma_U)^{(j+3) \alpha}
302:    \tilde{u}_\alpha \right\} \,,  \\
303: {\cal{L}}_{u\tilde{d}\tilde{\chi}^{+}} &= &\sum_{\sigma=1}^{2}\,
304: \sum_{i,j=1}^{3} g \left\{ \bar{u} _ {i}\,[V_{\sigma
305: 2}^{*}\,(\frac{m_u}{\sqrt{2} M_W \sin \beta} V_{CKM})_{ij}]
306: \,P_L\,\tilde{\chi} _{\sigma}^{+}\,(\Gamma_D)^{j \alpha
307: }\,\tilde{d}_{\alpha}-\bar{u}_{i} [ U_{\sigma 1} (V_{CKM})_{ij}]
308: P_R\,
309: \tilde{\chi}_{\sigma}^{+}\,(\Gamma_D)^{j \alpha} \,\tilde{d}_\alpha \right.   \nonumber \\
310: & &  \left. +\,\bar{u}_{i}\,[U_{\sigma 2}\,(V_{CKM}
311: \frac{m_d}{\sqrt{2} M_W \cos \beta})_{ij}]
312: \,P_R\,\tilde{\chi}_{\sigma}^{+}\,(\Gamma_D)^{(j+3) \alpha
313: }\,\tilde{d}_\alpha \right\} +\text{h.c.} \,, \label{interactions}
314: \end{eqnarray}
315: where $T^{a}$ are the $SU(3)_{c}$ generators, $i=1,2,3$ is the
316: generation index, $\alpha =1, \ldots, 6$ is the squark
317: flavor index, $s$ and $t$ are color indices, $N$ is the $4\times 4$
318: rotation matrix defined by $N^{*}M_{\tilde
319: \chi^0}N^{-1}=\mathrm{diag}(m_{\tilde {\chi}_{1}^{0}},\,m_{\tilde
320: {\chi}_2^0}, \,m_{\tilde {\chi}_3^0}, \,m_{\tilde {\chi}_4^0})$,
321:  the index $\sigma$ refers to chargino mass eigenstates, and $V$
322: and $U$ are the usual chargino rotation matrices defined by
323: $U^{*}M_{\tilde {\chi} ^{+}}V^{-1}=\mathrm{diag} (m_{\tilde {\chi}
324: _{1}^{+}},m_{\tilde {\chi} _{2}^{+}})$. From the above interactions
325: one can see that  FCNC neutralino and
326: gluino interactions only arise from up-type squark
327: mixings, while the  FCNC chargino interactions are induced from both the off-diagonal elements in the CKM matrix, and
328: from the flavor mixings in down-type squark mass matrix.
329: 
330: Although each of the above interactions contributes to the top quark
331: FCNC transitions by  gaugino mediated loops, the contributions could be of quite
332: different magnitude.
333: Since both the neutralino and gluino contributions depend on the
334: same parameters in the up-type squark mass matrix, their different
335: coupling strength indicate that the neutralino contribution is
336: much smaller than the gluino contribution, except for a very massive gluino and light neutralino
337: scenario.
338: Noting that B-physics requires small
339:  $\delta^d \leq {\cal O}(0.1)$ \cite{bsr,Ball,B-summary},
340: the FCNC interactions induced by charginos are in general not
341: large. Recently, these three types of contributions to $ g g \to t
342: \bar{c}$ at the LHC were simultaneously calculated
343: in \cite{Gad-pptc-mssm}, and it was shown that both the neutralino
344: and the chargino contribution are several orders  of
345: magnitude smaller than the gluino contribution for most SUSY
346: parameter space. Since we are
347: mostly interested in the parameter regions with large predictions for FCNC processes,
348: in this paper we consider only the
349: gluino-mediated contributions.
350: 
351: Even if only the gluino-mediated loops are considered in calculating the
352: top quark FCNC interactions, the model has still a large parameter set.
353: Beside the gluino mass, there are nine
354: soft mass parameters in the scharm-stop mass matrix, which
355: complicates our analysis.
356: In order to simplify our calculations, we neglect
357: the charm quark mass. Then the  amplitude squared for any top
358: quark FCNC mode/channel considered in this paper can be decomposed as
359: \begin{eqnarray}
360: |M|^2 = |M|^2_L + |M|^2_R,   \label{decompose}
361: \end{eqnarray}
362: where $|M|^2_L$ ($|M|^2_R$) is the amplitude squared with a left-handed (right-handed)
363: charm quark, as either an external state or internal state in Feynman diagrams.
364: Furthermore, by
365: using the mass insertion method\cite{mass insertion}, one can easily
366: find that $|M|^2_L$ ($|M|^2_R)$ vanishes if there is no
367: left-handed (right-handed) scharm mixings with stop, and that these amplitudes have
368: a weaker dependence on right-handed (left-handed) scharm
369: mixings than on the left-handed (right-handed) scharm mixings.
370: These features, verified numerically by our calculations,
371: motivate us to consider the case with only left-handed scharm mixings in top quark
372: FCNC processes. In this case, the relevant soft mass
373: parameters are reduced to seven, since we set $\delta_{RL}\delta_{RR} =0 $ and $M_{U_2}^2 $ is irrelevant to our
374: calculation (see Eq.~(\ref{up squark})). Throughout this paper, we
375: always consider this case, but we note that for those transitions
376: not involving $W, Z$ bosons, the results for $|M|^2_L $ can be
377: applied to $|M|^2_R $ with the substitutions
378: $R \leftrightarrow L$ and $M_{U_i} \leftrightarrow M_{Q_i}$.
379: 
380: Another reason for considering only left-handed scharm mixings with stops
381: is that these are well motivated in popular flavor-blind
382: SUSY breaking scenarios, such as the mSUGRA model \cite{sugra} and
383: gauge-mediated SUSY-breaking models \cite{gmsb}. In these models,
384: the sfermion-mass matrices are flavor diagonal at the
385: SUSY-breaking scale, but the Yukawa couplings can induce flavor
386: mixings when evolving the matrices down to the electroweak scale.
387: Estimates of these radiatively induced off-diagonal squark-mass
388: terms indicate the magnitude for left-handed flavor mixings are
389: proportional to bottom quark mass, while those for the right-handed
390: scharm are proportional to charm quark mass \cite{hikasa}.
391: Therefore, in phenomenological studies of scharm-stop mixings, one
392: usually assumes the existence of left-handed scharm mixings.
393: 
394: Finally, it should be pointed out that although we make use of the
395: parametrization in Eqs.(12-14), which is widely used in the literature
396: for the calculations by mass insertion approximation,
397: our calculations are the full computation in the mass eigentsate basis
398: of squarks (that is we first diagonalize the squark mass matrices
399: and then perform the loop calculations in the mass eigentsate basis).
400: Such a treatment, unlike the mass insertion method \cite{mass insertion}
401: which makes sense only for $\delta's < 1$,  can allow for $\delta's > 1$
402: (as will be shown in Sec. V, in some cases $\delta's > 1$ can be permitted by
403: all experimental constraints because we use non-universal squark mass
404: parameters, that is $M_Q$, $M_U$ and $M_D$ are not degenerate).
405: Note that although all our numerical results are obatined from
406: such full computation in the mass eigenstate basis,
407: we will utilize the mass insertion method when we try to
408: qualitatively explain the behaviors of some results.
409: 
410: \section{\bf The Effective Vertex Method }
411: 
412: We  introduce a method which can  greatly simplify our
413: calculations since it avoids repetition of the evaluation of a same
414: loop-corrected vertex in different places, or in
415: different processes. All results in this paper were obtained by
416: this method, and some of them were cross-checked by other tools
417: such as {\tt FormCalc} \cite{Hahn}.
418: 
419: The key point of our method is the so-called ``effective vertex".
420: To illustrate this method we consider $ g g
421: \to t \bar{c}$ as an example. The Feynman diagrams for this process
422: are shown in Fig.~\ref{general}. The SUSY-QCD contributions
423: to the $c-t$ transition and the vertex $t\bar c g$,
424: as well as the box diagrams, are given in Fig.~\ref{SQCD diagram}.
425: Noting that the amplitude for Fig.~1(a) can be
426: split into two terms,  one containing a charm quark propagator,
427: and the other containing a top quark propagator
428: \begin{eqnarray}
429: M_a  \propto \frac{i} {\q_slash - m_t} i \Sigma(q)
430: \frac{i}{\q_slash - m_c} = \frac{i (\q_slash + m_t )}{m_c^2 -
431: m_t^2} \ i \Sigma(q) \ \frac{i}{\q_slash - m_c} +
432:  \frac{i}{\q_slash- m_t} \ i \Sigma (q) \  \frac{i (\q_slash + m_c)}{m_t^2 -
433:  m_c^2},   \label{technique}
434: \end{eqnarray}
435: we collect the first term together with Fig.~1(e, f),
436: and combine the second term together with
437: Fig.~1(g, h). After this arrangement, we can
438: define a momentum dependent effective $\bar{t}cg$ interaction as
439: \begin{eqnarray}
440: \Gamma^{eff}_{\mu} (p_t, p_c) &= & \Gamma_\mu^{\bar{t}cg}
441: (p_t,p_c) + i \Sigma (p_t) \  \frac{i (\p_slash_t + m_c)}{m_t^2 -
442: m_c^2} \Gamma_\mu^{\bar{q}qg} + \Gamma_\mu^{\bar{q}qg} \frac{i
443: (\p_slash_c + m_t )}{m_c^2 - m_t^2} \ i \Sigma(p_c), \label{eff}
444: \end{eqnarray}
445: where $\Gamma_\mu^{\bar{t}cg} $ is the penguin diagram
446: contribution to the effective interaction and
447: $\Gamma_\mu^{\bar{q}qg}$ is the usual QCD vertex. Then
448: the calculation of Fig.~\ref{general}(a-h) is
449: equivalent to the calculation of the ``tree" level transition
450: depicted in Fig.~\ref{effective}(a-c), which obviously has a
451: simpler structure. By following this method, our calculations
452: can be greatly simplified, since the effective  $\bar{t}cg$ interaction
453: appears in many processes, and the effective
454: $\bar{t}cg$ interaction is the same for all channels considered in the paper.
455: \vspace*{-.5cm}
456: %% fig.1 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
457: \begin{figure}[hbt]
458: \begin{center}
459: \epsfig{file=fig1.ps,width=10cm, height=8.5cm}
460: \vspace*{-0.7cm}
461: \caption{Feynman diagrams for $g g \to t \bar{c}$. Additional
462: diagrams with the two gluons interchanged are not shown.}
463: \label{general}
464: \end{center}
465: \end{figure}
466: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
467: %%%% fig.2 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
468: \begin{figure}[htb]
469: \begin{center}
470: \epsfig{file=fig2.ps,width=12cm, height=5.5cm }
471: \vspace*{-0.7cm}
472: \caption{SUSY-QCD contribution
473: to $c-t$ transition, $\bar{t} c g$ interaction and box diagrams
474: for $ g g \to t \bar{c}$.} \label{SQCD diagram}
475: \end{center}
476: \end{figure}
477: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
478: %%%% fig.3 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
479: \begin{figure}[hbt]
480: \begin{center}
481: \epsfig{file=fig3.ps,width=12cm, height=5.5cm}
482: \vspace*{-0.3cm}
483: \caption{Effective diagrams for the process $g g \to t \bar{c}$. Here
484: $\Gamma^{eff}_{\mu}$ is the effective $\bar{t}c g$ interaction
485: defined in Eq.~(\ref{eff}). } \label{effective}
486: \end{center}
487: \end{figure}
488: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
489: 
490: Of course, the effective vertex $\bar{t}cg$ is model-dependent
491: since its components $\Gamma_\mu^{\bar{t}cg}$ and $\Sigma$ are
492: model-dependent. We obtained their expressions in SUSY QCD
493: analytically. We retain the tensor loop functions rather than
494: expanding them in terms of scalar loop functions as usual
495: \cite{Hooft}. This method makes our results quite compact and also
496: simplifies our Fortran codes which will be discussed below.
497: 
498:  The effective vertex $\bar{t}c g$ in
499: Eq.~(\ref{eff}) is a 4-component Lorentz vector and also
500: a $4 \times 4$ matrix in Dirac spinor space.  In its realization in
501: Fortran coding, we use a dimension-three array $V(i,j,k)$
502: with $i$ (=1,2,3,4) labeling the Lorentz index and  $j,k$ (=1,2,3,4)
503: labeling the spinor indices. We also use arrays to encode
504: other quantities such as Lorentz vectors, Dirac spinors, Lorentz tensors
505: and Dirac $\gamma$ matrices. The steps to calculate the effective
506: interaction in Fortran code can be then summarized as follows:
507: \begin{itemize}
508: \item  Input the matrices $P_{L,R}$, $\gamma^\mu P_{L,R}$ and
509: $\sigma^{\mu \nu}$.  For any other matrices
510: encountered in the calculation, we use $\gamma$ algebra to
511: generate its elements.
512: 
513: \item Use the mass splitting method\cite{Barger} to generate events
514: with momentums for the initial and final particles.
515: 
516: \item For a generated event with fixed  momenta,
517:       the components of any tensor loop function can be calculated numerically and stored in arrays.
518: 
519: \item  Generate the $\gamma $ matrices $\gamma^{\mu_1} \cdots
520:        \gamma^{\mu_n}$ and contract its Lorentz indices with those of
521:        tensor loop functions and those of quark momenta to obtain the
522:        effective vertex.
523: \end{itemize}
524: To calculate the amplitude of $g g \to t \bar{c}$, we also
525: need to calculate box diagrams. Such calculations are usually tedious
526: if the four-point  tensor loop functions are expanded in terms of
527: scalar loop functions.  Since we choose to retain the
528: tensor loop functions and contract the indices
529: numerically, our results are quite compact, as shown for
530: $ g g \to t \bar{c}$ in the Appendix.
531: The general expression for a box diagram is the sum of fermion chain
532: of the form $\left ( \bar{u} \gamma^{\mu_1} \cdots \gamma^{\mu_n} u
533: \right ) \times D_{\mu_i \cdots \mu_j} \times p_{\mu_k} \cdots $
534: and its numerical value is calculated by the following steps
535: \begin{itemize}
536: \item Input the wave functions for fermions and define the
537: multiplication of $\gamma$ matrix with the wave function.
538: 
539: \item Generate the tensor, say $ \bar{u} \cdots \gamma^{\mu_{n-1}}
540: \gamma^{\mu_{n}} u$, and contract its indices with those of loop
541: functions $D_{\mu_i \cdots \mu_j}$ and those of vectors involved,
542: to get the value of each term in the amplitude.
543: \end{itemize}
544: 
545: With the method introduced above, we can also easily calculate
546: other FCNC interactions. Let us take the calculation of $ c g \to t Z $ as
547: an example. Its Feynman diagrams can be obtained from
548: Figs.~\ref{general}-\ref{SQCD diagram} by removing those
549: involving triple-gluon interaction and gluon-gluino-gluino
550: interaction, and then replacing any gluon with Z boson.  Using the
551: technique from Eq.~(\ref{technique}) to introduce the effective
552: $\bar{t}c g$ interaction and the effective $ \bar{t}c Z$
553: interaction, one can again get simplified diagrams similar to
554: Fig.~\ref{effective}(b-e).
555: 
556:  Once  $ g g \to t \bar{c}$ is
557: calculated, evaluation of the others $ t \to c g$, $ c g \to t g$ and $ t \to c
558: g g$  becomes rather easy.  The decay  $t
559: \to c g$ is now a tree level interaction induced by the interaction
560: $\bar{t} c g$. The amplitudes (or their conjugates) for $ c g \to t
561: g $ and $ t \to c g g$ are related to that of $ g g \to t
562: \bar{c}$, and can be easily obtained by making some simple replacements
563: which can be easily realized in our code.
564: 
565: In the Appendix, we list the explicit forms of all the
566: penguin-induced FCNC interactions discussed in this paper. These
567: interactions are needed to obtain the effective top FCNC interactions.
568: 
569: \section{\bf Numerical results and dependence on SUSY parameters}
570: 
571: We know, from the discussion in Section II,
572: that the
573: calculations of the SUSY-QCD contributions induced by the flavor
574: mixings between left-handed scharm and stop, depend on the parameters
575: $M_{Q_{2,3}}$, $M_{U_3}$, $X_t = A_t - \mu \cot \beta$, $
576: m_{\tilde{g}}$, $\delta_{LL}$ and $\delta_{LR}$. In
577: this section, we investigate the dependence of the numerical
578: results on these parameters.  We first show the results for
579: top quark rare decays. These decay modes occur only via one
580: effective interaction and thus their dependence on the parameters
581: is relatively simple. We perform a comparative study and plot the
582: results for these decays together, to illustrate their dependence
583: on a given set of parameters. After analyzing the features of the
584: top quark rare decays, we extend the study to the FCNC top quark
585: productions. They usually involve two effective
586: interactions and several box diagrams, and thus their dependence
587: on SUSY parameters is more complex.
588: 
589: The SM parameters used in our calculations  are \cite{pdg}
590: \begin{eqnarray}
591: && m_t = 172.7 {~\rm GeV}, ~~m_b= 4.8 {~\rm GeV}, ~~m_Z = 91.19 {~\rm GeV}, \nonumber \\
592: && \sin \theta_W = 0.2228, ~~\alpha_s (m_t) = 0.1095, ~~\alpha =1/128. \label{sm-para}
593: \end{eqnarray}
594: After the assumptions discussed in Sec. III, about 10 SUSY
595: parameters are still involved.  We will show below the dependence
596: on SUSY parameters of the top FCNC processes.  When one of the
597: parameters is varied, the others will be fixed to their
598: ``central'' values, taken as
599: %
600: \begin{eqnarray}
601:  \label{susy-para1}
602: M_{\rm SUSY}= M_{Q_{3}} = M_{U_3} = M_{Q_{2}}= 500 {~\rm GeV},
603: \quad X_t = 1000 {~\rm GeV}, \quad m_{\tilde{g}} = 250 {~\rm GeV},
604: \quad \tan \beta = 5.
605: \end{eqnarray}
606: %
607: The values of $\delta_{LL}$ and $\delta_{LR}$ will be shown in the
608: figures. With the exception of the last plot in this section (Fig.
609: 12), we adopt the so-called $m_h^{\rm max}$ scenario \cite{mhmax}
610: which is widely discussed in Higgs physics, and which assumes that
611: all the soft mass parameters are degenerate
612: %
613: \begin{eqnarray}
614: M_{\rm SUSY} = M_{Q_i} = M_{U_i} = M_{D_i},
615: \end{eqnarray}
616: %
617: and that all the trilinear couplings are also degenerate,
618: $A_{u_i}=A_{d_i}$, with $X_t/M_{\rm SUSY}=2$.
619: In investigating the
620: processes $ t \to c h $ and $ c g \to t h $, we used the
621: loop-corrected lightest Higgs boson mass and the effective Higgs
622: mixing angle\cite{hollik,cao}. These two quantities involve two
623: additional parameters $\mu $ and $m_A$, which are fixed as
624: \begin{eqnarray} \label{susy-para2}
625: \mu = m_A = 500 {~\rm GeV}.
626: \end{eqnarray}
627: 
628: In our calculations we use CTEQ6L \cite{cteq} to generate the
629: parton distributions with renormalization scale $\mu_R $ and
630: factorization scale $\mu_F$, chosen to be $\mu_R = \mu_F = m_t$.
631: To make our predictions more realistic, we applied some kinematic
632: cuts. For example, for the three body decay $t \to c g g$  we
633: require that the energy of each decay product be larger than $15$
634: GeV and the separation of any two final states be more than $15^o$
635: in the top quark rest frame. For the top quark production
636: channels, we require that the transverse momentum of each produced
637: particle be larger than $15$ GeV and their pseudo rapidity be less
638: than 2.5 in the laboratory frame. Moreover, for $c g \to t$
639: followed by $t\to b W$, we do not require the top quark exactly on
640: mass shell and instead we require the invariant mass of bottom
641: quark and W boson in a region of $m_t - 3 \Gamma_t \leq M_{b W}
642: \leq m_t + 3 \Gamma_t $ ($\Gamma_t$ is the top quark width). This
643: requirement was once used in \cite{Hosch} to investigate the
644: observability of this channel at hadron colliders in the effective
645: Lagrangian framework.
646: 
647: Furthermore,  we vary the flavor mixings, $\delta_{LL}$ and
648: $\delta_{LR}$, over a wide range, with the only requirement that
649: they satisfy current collider searches for sparticles and Higgs
650: bosons\cite{pdg}:
651: \begin{eqnarray}
652: m_{\tilde{q}} \geq 96 {\rm ~GeV}, \quad m_{\tilde{g}} \geq 195  {\rm ~GeV}, \quad
653: m_h \geq 85  {\rm ~GeV}.  \label{bound5}
654: \end{eqnarray}
655: In principle, some low energy data, such as $b-s$ transition and
656: $\delta \rho$, can also constrain these mixings \cite{cao}. But
657: those so-called indirect constraints are usually quite
658: complicated. To simplify the discussion in this section we do not
659: impose these indirect constraints, but we address such a question
660: in the next section. We checked that the conclusions obtained in
661: this section  are valid in the region favored by these indirect
662: constraints \footnote{Another advantage of allowing the parameters
663: to vary within a large range is that, for most of the processes
664: considered in this paper, the dependence of their rates on
665: $\delta_{RL}$ and $\delta_{RR}$ is similar to that on
666: $\delta_{LR}$ and $\delta_{LL}$. But the indirect constraints on
667: them are quite different: while the indirect constraints on
668: $\delta_{LL}$ and $\delta_{LR}$ may be quite stringent, the limits
669: on $\delta_{RL}$ and $\delta_{RR}$ are rather weak \cite{cao}.
670: Therefore, the allowed range of $\delta_{RL}$ and $\delta_{RR}$ is
671: much larger than $\delta_{LL}$ and $\delta_{LR}$.}.
672: 
673: %%%%%%%%fig.4 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
674: \begin{figure}[htb]
675: \begin{center}
676: \epsfig{file=fig4.ps,width=14cm,height=10cm}
677: \caption{The branching ratios of FCNC top quark decays. Unspecified parameters
678: are given in Eqs.~(\ref{sm-para}-\ref{susy-para2}). The
679: values of $\delta_{LL}$ and $\delta_{LR}$ are arbitrarily chosen,
680: and a smaller value will lower the rates, but not change the
681: tendencies of these curves.} \label{decay1}
682: \end{center}
683: \end{figure}
684: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
685: 
686: %%%% fig.5 (old 8-9  \label{decay5} \label{decay6} )
687: \begin{figure}[htb]
688: \begin{center}
689: \epsfig{file=fig5.ps,width=14cm,height=6cm } \caption{Same as
690: Fig.~\ref{decay1}, but as functions of $X_t (= A_t - \mu \cot \beta)$.}
691: \label{decay5}
692: \end{center}
693: \end{figure}
694: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
695: 
696: %%%% fig.6 (old 10-13  \label{decay7}\label{decay8}\label{decay9}\label{decay10})
697: \begin{figure}[htb]
698: \begin{center}
699: \epsfig{file=fig6.ps,width=14cm,height=12cm } \caption{Same as
700: Fig.~\ref{decay1}, but as functions of the mixing parameters.}
701: \label{decay7}
702: \end{center}
703: \end{figure}
704: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
705: 
706: In Figs.~\ref{decay1}-\ref{decay7}, we present the branching ratios
707: of various FCNC top quark decays, defined  with respect to the width $\Gamma_t(t \to b W)$
708: ($\simeq 1.45$ GeV). We plot the six branching ratios for the decays
709: in Eq.~(\ref{decy-1})-(\ref{fcncdecays}) as functions of
710: $m_{\tilde{g}}$, $M_{SUSY}$ and $X_t$ in the upper two diagrams of
711: Fig.~\ref{decay1}, the lower two diagrams of Fig.~\ref{decay1} and
712: Fig.~\ref{decay5}, respectively. We also show the dependence of the
713: branching ratios on the squark mixing parameter
714: $\delta_{LL}$ ($\delta_{LR}$) by
715: fixing the value of $\delta_{LR}$ ($\delta_{LL}$) in
716: Fig.~\ref{decay7}. These figures show some common features of
717: all the gauge boson decay modes. The first one is that, as the
718: sparticles become heavy, the branching
719: ratios drops monotonously (see Fig.~\ref{decay1}),
720:  a reflection of the decoupling property of
721: the MSSM. The second is that, as shown in Fig.~\ref{decay5}, the branching
722: ratios increase with the increase of $X_t$. This is because $X_t$
723: affects the squark mass splittings and could alleviate the cancellation between different loop
724: contributions. The third feature is that the branching
725: ratios increase rapidly
726: with the flavor mixing $\delta_{LL}$ and $\delta_{LR}$, because the flavor
727: mixings  can not only enhance the
728: coupling strength of squark flavor changing interactions, but also
729: enlarge the squark mass splittings. This
730:  is illustrated in the first two diagrams of
731: Fig.~\ref{decay7}. These last two effects combined
732:  make the branching ratios very sensitive to the flavor mixing
733: parameters.
734: 
735: By comparing the case $\delta_{LL} \neq 0 $ with the case
736: $\delta_{LR} \neq 0$ throughout Figs.~\ref{decay1}-\ref{decay7},
737: one finds that  $\delta_{LR}$ induces larger rates with weaker
738: dependence on sparticle masses. Furthermore, when both $\delta_{LL}$
739: and $\delta_{LR}$ are non-zero, as can be inferred from the last
740: two diagrams of Fig.~\ref{decay7}, the $\delta_{LL}$ and the
741: $\delta_{LR}$ dependences interfere destructively. To understand such
742: behaviors, we resort to the mass insertion method,  as it can give more intuitive results \cite{mass
743: insertion}. We take the decay $t \to c g$ as an example.
744: By gauge invariance, the general expression for
745: the effective $\bar{t}c g$ interaction takes the form
746: \begin{eqnarray}
747: \Gamma_\mu = F_1 (k^2) \bar{t}  T^a (k^2 \gamma_\mu - k_\mu
748: \k_slash ) P_L  c g^a -  m_t F_2 (k^2) \bar{t} T^a i \sigma_{\mu
749: \nu} k^\nu P_L c g^a \label{effective expression}
750: \end{eqnarray}
751: where $F_{1,2}(k^2)$ are form factors arising from loop
752: calculations. For the decay $ t \to c g$,  $F_1$ does not contribute,
753:  since the gluon momentum $k^\mu $ satisfies $k^2=0$
754: and $ k \cdot \epsilon = 0$. So only the dipole moment term is
755: relevant to our discussion.  Noting that the dipole changes both the flavor and the
756: chirality of the fermions, we may infer
757: the form of $m_t F_2$ in the mass insertion approximation. If only
758: $\delta_{LL}$ is considered for flavor changing, $m_t F_2 $ must
759: be
760: \begin{eqnarray}
761: m_t F_2 = \frac{m_t \delta_{LL}}{M_{SUSY}^2} A +
762: \frac{m_{\tilde{g}} \delta_{LL} m_t X_t}{M_{SUSY}^4} B,
763: \label{form1}
764: \end{eqnarray}
765: where $M_{SUSY} =max(m_{\tilde{g}}, M_{\tilde{q}})$, and $A, B$
766: are dimensionless constants coming from loop functions with
767: $1/M_{SUSY}^2$ factored out. The first term corresponds to the top
768: chirality
769: flipping contribution,  i.e., obtained by using the relation
770: $\bar{u}_t\!\p_slash_t = m_t \bar{u}_t$, while the second term corresponds to
771: the  $\tilde{t}_L-\tilde{t}_R$ mixing contribution, and  thus associated with
772: the gluino mass. The situation is quite different for $\delta_{LR}
773: \neq 0$, which alone can be responsible for both flavor
774: changing and chirality flipping. In this case, $m_t F_2$ should be
775: \begin{eqnarray}
776: m_t F_2 = \frac{m_{\tilde{g}} \delta_{LR}}{M_{SUSY}^2}C.
777: \label{form2}
778: \end{eqnarray}
779: where $C$, like $A$ and $B$, is a dimensionless constant coming
780: from loop functions with $1/M_{SUSY}^2$ factored out. Comparing
781: Eq.~(\ref{form1}) with Eq.~(\ref{form2}), we find that the
782: latter is larger if $m_{\tilde{g}} \gg m_t$. Assuming that $m_{\tilde{g}} \simeq M_{\tilde{q}}$, Eq.~(\ref{form1})
783: scales like $1/M_{SUSY}^2$ while Eq.~(\ref{form2}) scales like
784: $1/M_{SUSY}$. This explains the fact  that $\delta_{LR}$
785: induces larger rates with weaker dependence on sparticle masses.
786: Moreover, a detailed calculation shows that $m_t F_2$ in
787: Eq.~(\ref{form1}) is of opposite sign to that in Eq.~(\ref{form2}), which
788:  means that the $\delta_{LL}$ contribution tends to cancel the
789: $\delta_{LR}$ contribution.
790: 
791: The decay $ t \to c h$ has similar features to the decay into a
792: gauge boson except for a rather complicated dependence on
793: $M_{SUSY}$ and $X_t$, as shown in Figs.~\ref{decay1}-\ref{decay5}.
794: For the effective $\bar{t} c h$ interaction, with both the top and
795: the charm quarks on-shell, the general expression is
796: \begin{eqnarray}
797: \Gamma^{eff}_{\bar{t}ch} = D \bar{t} P_L c h,
798: \end{eqnarray}
799: where $D$ is a form factor emanating from the loop calculation.
800: This interaction,
801: like the effective $\bar{t}c g$ interaction, also involves both
802: flavor change and chirality flip. So in
803: some aspects, the behavior for $ t \to c h$ should be similar to $
804: t \to c g$. But its dependence on $X_t$ is more complicated than
805: other decay modes. The reason may be that $X_t$ not only affects
806: the masses and mixings of squarks, but also affects the Higgs
807: boson mass and its mixing angles, and, further, it enters the
808: interaction of $\tilde{q}^\ast \tilde{q} h$.
809: 
810: An interesting feature, shown in Figs.~\ref{decay1}-\ref{decay7} is
811: that the branching ratio of $ t\to c g$ is always smaller than
812: the higher order decay $ t \to c g g$. This was
813: observed in the SM \cite{tcgg-sm} and the MSSM
814: \cite{Gad-pptc-mssm}, and it indicates that the QCD corrections to
815: $t \to c g$ may be important. Two reasons may account for this behavior.
816: One is that the QCD factor in the amplitude-square for $t \to c g
817: g$ is much larger than that for $ t \to c g$. The other reason is that,
818: unlike the case $t \to c g$, $F_1$ in Eq.~(\ref{effective
819: expression}) also contributes to $t \to c g g$ and this
820: contribution is important. From
821: Figs.~\ref{decay1}-\ref{decay7}, one finds that, in most cases
822: \begin{eqnarray}
823: Br( t \to c g g ) > Br( t \to c g ) > Br ( t \to c Z ) > Br(t \to
824: c  \gamma),
825: \end{eqnarray}
826: and in some cases $Br( t \to c h) $ may be the largest.
827: 
828: %%%% fig.7 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
829: \begin{figure}[htb]
830: \begin{center}
831: \epsfig{file=fig7.ps,width=14cm,height=11cm }
832: \caption{The cross sections of the FCNC top quark productions at the LHC.
833: Each curve labeled by the final states corresponds to a certain production channel.}
834: \label{product1}
835: \end{center}
836: \end{figure}
837: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
838: %%%% fig.8 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
839: \begin{figure}[htb]
840: \begin{center}
841: \epsfig{file=fig8.ps,width=14cm,height=6cm }
842: \caption{Same as
843: Fig.~\ref{product1}, but as a function of  $X_t (= A_t - \mu \cot \beta)$.} \label{product5}
844: \end{center}
845: \end{figure}
846: 
847: %%%% fig.9 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
848: \begin{figure}[htb]
849: \begin{center}
850: \epsfig{file=fig9.ps, width=14cm,height=11cm }
851: \caption{Same as Fig.~\ref{product1}, but as a function of the mixing parameters.}
852: \label{product7}
853: \end{center}
854: \end{figure}
855: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
856: 
857: After giving the branching ratios for the FCNC top decays comparatively,
858: we now consider the FCNC top quark production channels. At the LHC the
859: production proceeds through the six parton-level processes shown in
860: Eqs.~(\ref{pro-1}-\ref{processes}), among which only $ g g \to t
861: \bar{c}$ has been extensively studied in the
862: MSSM\cite{pptc-mssm,Gad-pptc-mssm}. In calculating the
863: cross section for each channel, we also include the rate for its
864: charge conjugate channel. In the following discussions
865: we use the parton process to label its contribution to the hadronic
866: one.
867: Thus, the cross sections mentioned below refer to
868: the hadronic ones.
869: 
870: The dependence of the cross sections on SUSY parameters is plotted
871: in Figs.~\ref{product1}-\ref{product7}. These figures  exhibit the
872: same features as the FCNC top quark decays in
873: Fig.~\ref{decay1}-\ref{decay7}. Thus we do not repeat drawing
874: attention to the same features, and we only point out three
875: remarkable points of these diagrams. The first is that the rate of
876: $ c g \to t$ is generally larger than that for $ g g \to t
877: \bar{c}$. This is possible because the charm quark in the parton
878: distributions mainly comes from the splitting of a gluon
879: \cite{cteq} and thus $c g \to t $ can be seen as $ g g \to t
880: \bar{c}$ with the final charm quark going along the beam pipe.
881: Analyzing the signal versus background
882: \cite{Hosch,pptc-background}, it seems that $ c g \to t $ provides
883: a better opportunity for the observation of top production. The
884: second point is that the rate of $c g \to t g $ is comparable to
885: $g g \to t \bar{c}$ in most cases. This coincides with the results
886: in \cite{pptc-Han} where both modes were studied in the effective
887: Lagrangian approach. Since the two processes have the same signals
888: at colliders, namely single top plus one light-quark jet, one
889: should combine these two  when searching for the FCNC top
890: production events. The third point is that in most cases the cross
891: section for $c g \to t h$ is one order of magnitude smaller than
892: that for $g g \to t \bar{c}$. The reason is that although $c g \to
893: t h$ can proceed either through $\bar{t} c g$ interaction or  $
894: \bar{t} c h$ interaction, the two interactions interfere
895: destructively and thus the combined contributions are suppressed.
896: 
897: As shown above, in most of parameter space the
898: rates  are expected to be quite small at the LHC.  In
899: order to study the observability of such rare processes,
900: it is necessary to scan the whole
901: parameter space to figure out
902: the maximum value that each process can reach.
903: We call the `favorable region' the part of parameter
904: space which gives the maximum value for the rate of
905: an individual channel. Of course, such favorable region is
906: process-dependent. In what follows we take $g g \to t \bar{c}$ as
907: an example to carry out such a parameter scan explicitly. In
908: the next section, the maximal rates
909: for all processes will be tabulated, after discussing
910: the effects of the indirect constraints
911: on the flavor mixing parameters.
912: 
913: In Fig.~\ref{m1} we plot the maximal cross section
914:   for $ g g \to t \bar{c}$ as a function of $X_t$ with
915: non-zero $\delta_{LL}$ (left panel) and $\delta_{LR}$ (right panel).
916: The maximum value is obtained by fixing $M_{SUSY}$ and $X_t$ but varying
917: $m_{\tilde{g}}$ and $\delta_{LL}$ or $\delta_{LR}$.
918: In searching for such maximal values,
919: we required $m_{\tilde{q}} \geq 100$ GeV  and $m_{\tilde{g}} \geq 200$ GeV.
920: 
921: For non-zero $\delta_{LL}$, we find that the
922: parameter points for the maximal cross sections correspond to the
923: lightest squark mass and the gluino mass fixed at their
924: allowed lower bound, i.e., 100 GeV and 200 GeV, respectively. This
925: can be easily understood from the decoupling property of the MSSM.
926: At such optimum points, if $M_{SUSY}$ is fixed, $X_t$ value is
927: related to $\delta_{LL}$. Since both $X_t$ and
928: $\delta_{LL}$ are the off-diagonal elements in squark mass matrix,
929: a large $X_t$ will correspond to a small $\delta_{LL}$ and vice
930: versa. Then from Eq.~(\ref{form1}), one can
931: see that neither large nor small
932: $X_t$ can predict the largest rate for $ g g \to t \bar{c}$. This
933: explains the behavior of each curve in the left panel of
934: Fig.~\ref{m1}.
935: 
936: In Fig. ~\ref{m2}, we plot the maximal cross section as a function
937: of mixing parameters. For non-zero
938: $\delta_{LL}$ mixings (left panel), we see that the maximal value of each
939: curve lies at a moderate $\delta_{LL}$, which agrees with the
940: above analysis. Among the three curves for different $M_{SUSY}$,
941: the maximal value is obtained for $M_{SUSY} = 800 $ GeV. The reason is that
942: $M_{SUSY} = 800 $ GeV implies heavier masses for the other two squarks
943: and alleviates the cancellation between different diagram
944: contributions.
945: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
946: %%%% fig.10
947: \begin{figure}[htb]
948: \begin{center}
949: \epsfig{file=fig10.ps,width=15cm,height=6.5cm} \vspace*{-0.3cm}
950: \caption{The maximal cross section of $t\bar c$ production at the
951: LHC, proceeding through the parton-level channel $g g \to t
952: \bar{c}$, as a function of $X_t$ by varying $m_{\tilde{g}}$ and
953: $\delta_{LL}$ or $\delta_{LR}$.} \label{m1}
954: \end{center}
955: \end{figure}
956: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
957: %%%% fig.11
958: \begin{figure}[htb]
959: \begin{center}
960: \epsfig{file=fig11.ps,width=15cm,height=6.5cm } \vspace*{-0.3cm}
961: \caption{Same as Fig.~\ref{m1}, but as a function of mixing
962: parameters. } \label{m2}
963: \end{center}
964: \end{figure}
965: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
966: %%%% fig.12
967: \begin{figure}[htb]
968: \begin{center}
969: \epsfig{file=fig12.ps,width=10cm,height=10cm }
970: \caption{Same as Fig.~\ref{m1}, but as a function of $M_{SUSY}$ under
971: different cases specified in the text.}
972: \label{m5}
973: \end{center}
974: \end{figure}
975: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
976: 
977: For the case of non-zero $\delta_{LR}$, maximal cross sections
978: also appear  at the lower bounds of the lightest squark mass and
979: the gluino mass. This property together with the effective
980: $\bar{t} c g$ coupling in Eq.~(\ref{form2}), imply that the
981: maximal value for each curve should lie at a large $\delta_{LR}$,
982: or by the correlation, at small $X_t$. (Both $X_t$ and
983: $\delta_{LR}$ appear as the off-diagonal elements in squark mass
984: matrix. Thus, for a given $M_{SUSY}$, their large values enlarge
985: the mass splitting between squark mass eigenstates and lead to a
986: small mass for the lightest squark. Due to the lower bound on the
987: lightest squark mass, $\delta_{LR}$ and $X_t$ cannot be both
988: large.) This is in agreement with the behaviors of the curves for
989: non-zero $\delta_{LR}$ shown in the right panels of
990: Figs.~\ref{m1}-\ref{m2}.  For the right panel of Fig.~\ref{m1}, we
991: checked that for $M_{SUSY} = 800 $ GeV, varying $X_t$ from 0 to
992: $500$ GeV only resulted in a change of $0.006$ for $\delta_{LR}$.
993: This is the reason that the maximal cross section is insensitive
994: to $X_t$ for $X_t \leq 500$ GeV.
995: 
996: In Fig.~\ref{m5} we show the dependence of the maximal cross
997: section on $M_{SUSY}$ for non-zero $\delta_{LL}$ or $\delta_{LR}$ values.
998: To get the maximal cross sections, we fix the lightest squark mass
999: as $100$ GeV and vary the value of $\delta_{LL}$ or $\delta_{LR}$.
1000: We considered three cases,
1001: \begin{itemize}
1002: \item Case I: $M_{Q_2} = M_{Q_3} = M_{U_3} = M_{SUSY}$, \item Case
1003: II:  $M_{Q_3} = M_{U_3} = M_{SUSY} $, $M_{Q_2} = 1.2 M_{SUSY}$,
1004: \item Case III: $M_{Q_3} = M_{SUSY}, M_{U_3} = 0.8 M_{SUSY},
1005: M_{Q_2} = 1.2 M_{SUSY}$.
1006: \end{itemize}
1007: These cases are motivated by mSUGRA model
1008: \cite{sugra} where the three squark masses are generated from the same soft
1009: breaking mass parameter $m_0$ at supersymmetry breaking scale, but are split
1010: due to quark Yukawa couplings when they evolve down to electroweak
1011: scale \cite{spectrum}. From this figure we see that the maximal
1012: cross section values increases with $M_{SUSY}$, and for non-zero
1013: $\delta_{LL}$ ($\delta_{LR}$), case I (II) gives the largest
1014: prediction for the cross sections. These results indicate that in
1015: searching for the
1016: maximum values of the cross section, we should treat $M_{Q_2}$,
1017: $M_{Q_3}$ and $M_{U_3}$ as free parameters and explore all the
1018: possibilities for  these masses.
1019: 
1020: Note that in the above discussions we have only considered the
1021: case for non-zero $\delta_{LL}$ or $\delta_{LR}$. But as pointed
1022: below Eq.~(\ref{decompose}), most of our conclusions should  be
1023: valid for the case non-zero $\delta_{RL}$ or $\delta_{RR} \neq 0$
1024: with an interchange of $L$ and $R$.
1025: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1026: 
1027: \section{\bf Low-energy constraints on FCNC top quark interactions}
1028: \subsection{Constraints on scharm-stop flavor mixings}
1029: We know, from the discussions in the preceding section, that the
1030: rates of the FCNC top quark processes depend strongly on the
1031: flavor mixing parameters $\delta_{LL}$ and $\delta_{LR}$, which
1032: are treated as free. Of course, such a treatment is
1033: informative and useful to the LHC experiments since it allows to
1034: directly place limits on the mixing parameters once the
1035: measurements are made at the LHC.
1036: 
1037: However, it is worth noting that these mixing parameters may be subject to
1038: various direct and indirect experimental constraints. Firstly,
1039: since the mixing terms appear as the non-diagonal elements of
1040: squark mass matrices, they can affect the squark mass spectrum,
1041: especially by enlarging the mass splitting between squarks. Therefore they
1042: should be constrained by the squark mass bounds from the direct
1043: experimental searches. At the same time, since the squark loops
1044: affect the precision electroweak quantities such as $M_W$ and the
1045: effective weak mixing angle
1046: $\sin^2\theta_{eff}$\cite{hollik,mw-mssm}, such mixings could also be
1047: constrained by precision electroweak measurements. As
1048: shown in\cite{hollik,cao}, to a good approximation, the
1049: supersymmetric corrections to the electroweak quantities contribute
1050: through the $\delta\rho$ parameter and thus sensitive to the mass
1051: splitting of squarks. Secondly, the processes governed by $b \to s$
1052: transition like $B_s-\bar B_s$ mixings \cite{Ball} and $b \to s
1053: \gamma$\cite{bsr} can provide rich information about the $\tilde
1054: s-\tilde b$ mixings. Through the SU(2) relation between the up-squark
1055: and down-squark mass matrices (see Eq.~(\ref{SU2})) and also
1056: through the electroweak quantities (since all squarks contribute
1057: to electroweak quantities via loops), the information can be
1058: reflected in the up-squark sector and hence constrain the scharm-stop
1059: mixings. Thirdly, we note that the chiral flipping mixings of
1060: scharm-stop come from the trilinear $H_2 \tilde{c}_L^\ast
1061: \tilde{t}_R$ interactions\cite{susyflavor}. Such interactions can
1062: lower the lightest Higgs boson mass $m_h$ via squark loops and
1063: thus should be subject to the current experimental bound on $m_h$.
1064: 
1065: In \cite{cao} these constraints were examined and it was shown that,
1066:  although they usually depend on additional unknown parameters
1067: in the down-type squark mass matrix, a combined analysis can still
1068: severely restrict the mixings $\delta_{LL}$ and $\delta_{LR}$
1069: in most cases. Here we extend the analysis of  \cite{cao}
1070: by providing more examples about these constraints and then
1071: perform a detailed investigation of the maximum rates for various
1072: top quark FCNC processes both with and without these constraints.
1073: 
1074: The constraints considered in our paper are $b \to s\gamma$,
1075: $B_s-\bar{B}_s$ mixing, $\delta M_W $ and $\delta \sin^2
1076: \theta_{eff}$ and the lightest Higgs boson mass. In the following,
1077: we first recapitulate these constraints and then apply them
1078: to scharm-stop mixings.
1079: \begin{itemize}
1080: \item[(1)] {\it $ b \to s \gamma$:} In the MSSM, there are four kinds of
1081: loops contributing to $ b \to s \gamma$ mediated respectively by
1082: the charged Higgs bosons, charginos, neutralinos and gluinos, each
1083: of which may be sizable \cite{bsr}. For a light charged Higgs
1084: mass, the contribution from the charged Higgs is quite large and
1085: always has the same sign as the SM contribution, and thus
1086: enhance the branching ratio to very high values \cite{charged higgs}. The
1087: current $b \to s \gamma$ data require either a  sufficiently heavy
1088: charged Higgs boson, or its contribution should  be canceled by other parts
1089: of SUSY effects. For the other three kinds of contributions, depending
1090: on SUSY parameters, they may interfere constructively or
1091: destructively with the SM effects and thus their relative sizes are not
1092: fixed \cite{bsr-character}. The effect of these properties of SUSY
1093:  on $ b \to s \gamma$ makes it necessary to consider all the
1094: contributions simultaneously in discussing $ b \to s \gamma $
1095: constraints on the squark flavor mixing parameters.
1096: 
1097: Current measurement of the branching ratio for $ b \to s \gamma$
1098: is rather precise, with $3 \sigma$ bounds given by \cite{heavy
1099: flavor}
1100: \begin{eqnarray}
1101: 2.53 \times 10^{-4} < Br ( b \to s \gamma ) < 4.34 \times 10^{-4}.
1102: \label{bound1}
1103: \end{eqnarray}
1104: With the SM prediction $Br^{NLO}( b \to s \gamma ) = (3.53 \pm 0.30)
1105: \times 10^{-4}$ \cite{bsr-sm} and a favored negative $C_7$ by $ b
1106: \to s l^+ l^- $ \cite{bsll}, where $C_7$ denotes the Wilson
1107: coefficient for the electromagnetic dipole operator ${\cal O}_7$,
1108: this decay can severely restrict the SUSY parameters. Our numerical
1109: results indicate that it is very sensitive to $\delta^d_{LR}$ and
1110: $\delta^d_{RL}$ for most cases, and for large $\tan \beta$, it is
1111: sensitive to $\delta_{LL}$ as well. The same results also
1112: indicate that $b \to s \gamma$ depends weakly on $\delta_{LR}$ which
1113: affects the decay via chargino-mediated loops,  but under all circumstances $ b
1114: \to s \gamma$ is not sensitive to $\delta_{RL}$ and $\delta_{RR}$.
1115: 
1116: \item[(2)] {\it $ B_s-\bar{B}_s$ mixing:} In the MSSM, although the loops
1117: mediated by charged Higgs boson, chargino and neutralino
1118: contribute to $B_s-\bar{B}_s$ mixing, their effects are generally
1119: much smaller than the SM contribution\cite{Ball}. So when
1120: discussing the constraint of $B_s-\bar{B}_s$ mixing on squark flavor
1121: mixing parameters, we only consider gluino contributions.
1122: Recently, the D0 collaboration gave the first two-side bound
1123: on the mass splitting between $B_s$ and $\bar{B}_s$ \cite{Abazov:2006dm}
1124: \begin{eqnarray}
1125: 17 ~{\rm ps}^{-1} < \Delta M_s < 21  ~{\rm ps}^{-1}   \quad (90\%~  {\rm C.L.}) \label{B_s}
1126: \end{eqnarray}
1127: This result is in agreement with the SM prediction, which is
1128: estimated as $21.3 \pm 2.6 ~{\rm ps}^{-1}$ by the UTfit group
1129: \cite{Bona:2005vz} and $20.9^{+4.5}_{-4.2} ~{\rm ps}^{-1}$ by the
1130: CKMfitter group \cite{Charles:2004jd}. After considering various
1131: uncertainties, the bounds in Eq.~(\ref{B_s}) can be re-expressed
1132: as\cite{Endo:2006dm}
1133: \begin{eqnarray}
1134: 0.55 < | 1 + M_{12}^{SUSY}/M_{12}^{SM} | < 1.37,   \label{bound2}
1135: \end{eqnarray}
1136: where $M_{12}$ is the transition matrix element for
1137: $B_s-\bar{B}_s$ transition. As pointed out in\cite{Ball},
1138: $B_s-\bar{B}_s$ mixing is very sensitive to the combinations
1139: $\delta_{LL} \delta_{RR}^d $ and $\delta_{LR}^d \delta_{RL}^d$
1140: and thus can put rather stringent constraints on any of $\delta^d$s.
1141: 
1142: \item[(3)] {\it $\delta M_W $ and $\delta \sin^2 \theta_{eff}$:} In the MSSM,
1143: the corrections to $M_W$ and $\sin^2 \theta_{eff} $ involve the
1144: calculation of the gauge boson self energy, and among all kinds of
1145: contribution to the self energy, those from squark loops are most
1146: important\cite{rho}.  As a good approximation, $\delta M_W$ and
1147: $\delta \sin^2 \theta_{eff}$ are related to $\delta \rho$
1148: by\cite{hollik}
1149: \begin{eqnarray}
1150: \delta M_W \simeq \frac{M_W}{2} \frac{c_W^2}{c_W^2 -s_W^2} \delta \rho, \nonumber \\
1151: \delta \sin^2 \theta_{eff} \simeq -\frac{c_W^2 s_W^2}{c_W^2
1152: -s_W^2} \delta \rho ,  \label{relation}
1153: \end{eqnarray}
1154: where
1155: \begin{eqnarray}
1156:  \delta \rho \equiv \frac{\Sigma_Z(0)}{M_Z^2} - \frac{\Sigma_W(0)}{M_W^2}.
1157: \end{eqnarray}
1158: Since the couplings are stronger for left-handed squarks
1159: than for right-handed squarks, $\delta \rho $ is sensitive to the
1160: mass splittings between left-handed up-squarks and down-squarks \cite{rho}.
1161: As far as $\delta_{LL}$ is concerned, due to
1162: the $SU(2)$ relation in Eq.~(\ref{SU2}), it changes up-squark and
1163: down-squark mass spectra simultaneously and thus its effects on
1164: $\delta \rho $ are generally small even for large $\delta_{LL}$.
1165: For $\delta_{LR}$ and $\delta_{RL}$, they are independent of
1166: $\delta_{LR}^d $ and $ \delta_{RL}^d$, which are very small as
1167: required by $b-s$ transition \cite{bsr,Ball,B-summary}. Thus
1168: a large $\delta_{LR}$ or $\delta_{RL}$ can induce a sizable mismatch
1169: between up-squark and down-squark mass spectra. As a result, large
1170: $\delta_{LR}$ or $\delta_{RL}$ can significantly change $\delta \rho$
1171: \cite{cao}.
1172: 
1173: With the recent analysis of the LEP data, the uncertainties in
1174: measuring $M_W$ and $\sin^2 \theta_{eff}$ were significantly
1175: lowered to read \cite{lep}
1176: \begin{eqnarray}
1177: \delta M_W < 34 {\rm ~MeV}, \quad \delta \sin^2 \theta_{eff} < 15 \times
1178: 10^{-5}. \label{bound3}
1179: \end{eqnarray}
1180: These uncertainties imply that the new physics influence on
1181: $\delta \rho $ should be lower than $5.5 \times 10^{-4}$.
1182: 
1183: \item[(4)] {\it Higgs boson mass $m_h$:} In the MSSM the
1184: loop-corrected lightest
1185: Higgs boson mass $m_h$ is defined as the pole of the corrected
1186: propagator matrix, which can be obtained by solving the equation
1187: \cite{Dabelstein}
1188: \begin{eqnarray}
1189:  & & \left[p^2 - m_{h, tree}^2 + \hat{\Sigma}_{hh}(p^2) \right]
1190:  \left[p^2 - m_{H, tree}^2 + \hat{\Sigma}_{HH}(p^2) \right]
1191: -\left[\hat{\Sigma}_{hH}(p^2)\right]^2 = 0 , \label{mass-equation}
1192: \end{eqnarray}
1193: where $m_{h, tree}$ and $m_{H, tree}$ are the tree-level masses of
1194: the neutral Higgs bosons $h$ and $H$, and  $\hat\Sigma_i (p^2)$ ($i= h h$, $h H$, $H H$)
1195: are the renormalized Higgs boson self energies. Among all SUSY
1196: contributions to the Higgs boson self energies, those from top and
1197: stop loops are by far dominant because of the large top quark Yukawa
1198: couplings\cite{h-original}. In the presence of the flavor mixings
1199: in the up-squark mass matrix, stops will mix with other squarks,
1200: and in this case, the dominant contribution comes from the up-squark
1201: sector \cite{hollik,feynhiggs}.
1202: 
1203: Our results indicate that the Higgs boson mass is more sensitive
1204: to $\delta_{LR}$ than to $\delta_{LL}$.  The reason is that
1205: $\delta_{LL}$ affects the Higgs boson mass only by changing the
1206: squark interaction through the unitary matrix $\Gamma$ in
1207: Eq.~(\ref{interaction}) while $\delta_{LR}$ can
1208: also appear directly in the coupling of trilinear $H_2 \tilde{c}_L
1209: \tilde{t}_R$ interaction, since the $\delta_{LR} M_{Q_2} M_{U_3}$ term
1210: in up-squark mass matrix comes from this trilinear
1211: interaction, and this  can reduce the lightest Higgs
1212: boson mass.
1213: Current collider searches for a MSSM Higgs boson have given the
1214: lower bounds on $m_h$ in five benchmark scenarios in Higgs
1215: physics\cite{benchmark}. In our calculation, we use the limit
1216: \begin{eqnarray}
1217: m_h >\left \{
1218: \begin{array}{ll} 92.8  {~\rm GeV} & \quad {~\rm for} \ m_h^{max}\ {~\rm scenario} \\
1219:                   & \\
1220: 85   {~\rm GeV}  & \quad   {~\rm for ~other ~scenarios}  \end{array} \right . \label{bound4}
1221: \end{eqnarray}
1222: \end{itemize}
1223: As the first example of these constraints, we consider the
1224: $m_h^{max}$ scenario with the parameters given in
1225: Eqs.~(\ref{susy-para1},\ref{susy-para2}). To determine these
1226: constraints, we also need to specify the values of the parameters
1227: in the down-squark mass matrix and the gaugino mass $M_2$. In
1228: Fig.~\ref{constr1} we show the allowed region in
1229: $\delta_{LL}-\delta_{LR}$ plane for the parameters
1230: \begin{eqnarray}
1231: &&M_{S}= M_{Q_{i}} = M_{U_i} = M_{D_{i}}= 500 {~\rm GeV}, ~~X_t 1000 {~\rm GeV}, ~~M_{\tilde{g}} = 250 {~\rm GeV}, ~~\tan \beta
1232: =5,  \nonumber \\
1233: &&\mu = M_A = M_2 = 500 {~\rm GeV},~~ \delta_{RL}=\delta_{RR}0,~~\delta^d_{LR}=\delta^d_{RL}=\delta^d_{RR}=0. \label{susy-para}
1234: \end{eqnarray}
1235: 
1236: %%%% fig.13
1237: \begin{figure}[htb]
1238: \begin{center}
1239: \epsfig{file=fig13.ps,width=9.cm,height=7.cm } \vspace*{-0.3cm}
1240: \caption{Experimental constraints on $\delta_{LL}$ versus
1241: $\delta_{LR}$ with the parameters given in Eq.~(\ref{susy-para}).
1242: The region under or left to each curve corresponds to the allowed
1243: region. The dashed-line enclosed area is allowed by $b \to s
1244: \gamma$. The colored area is the overlap region satisfying all the
1245: constraints.} \label{constr1}
1246: \end{center}
1247: \end{figure}
1248: 
1249: %%%% fig.14
1250: \begin{figure}[htb]
1251: \begin{center}
1252: \vspace*{-0.3cm} \epsfig{file=fig14.ps,width=9.0cm,height=7.cm }
1253: \caption{Same as Fig.~\ref{constr1}, but for $M_A = 200 $ GeV.}
1254: \label{constr2}
1255: \end{center}
1256: \end{figure}
1257: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1258: 
1259: %%%% fig.15
1260: \begin{figure}[htb]
1261: \begin{center}
1262: \epsfig{file=fig15.ps,width=9.0cm,height=7cm}
1263: \vspace*{-0.3cm}
1264: \caption{Same as Fig.~\ref{constr1}, but for $ M_S= 1 $ TeV.}
1265: \label{constr3}
1266: \end{center}
1267: \end{figure}
1268: 
1269: %%%% fig.16
1270: \begin{figure}[htb]
1271: \begin{center}
1272: \epsfig{file=fig16.ps,width=9.0cm,height=7cm}
1273: \vspace*{-0.3cm}
1274: \caption{Same as Fig.~\ref{constr1}, but for $M_{\tilde{g}} = 500 $ GeV.}
1275: \label{constr4}
1276: \end{center}
1277: \end{figure}
1278: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1279: 
1280: %%%% fig.17
1281: \begin{figure}[htb]
1282: \begin{center}
1283: \epsfig{file=fig17.ps,width=9.0cm,height=7cm}
1284: \vspace*{-0.3cm}
1285: \caption{Same as Fig.~\ref{constr1}, but for $\delta^d_{LR}=0.005$.}
1286: \label{constr5}
1287: \end{center}
1288: \end{figure}
1289: 
1290: %%%% fig.18
1291: \begin{figure}[htb]
1292: \begin{center}
1293: \epsfig{file=fig18.ps,width=9.0cm,height=7cm}
1294: \vspace*{-0.3cm}
1295: \caption{Same as Fig.~\ref{constr1}, but for the parameters in
1296: Eq.~(\ref{max-para}).
1297: In this case the Higgs boson mass cannot impose any constraints.}
1298: \label{constr6}
1299: \end{center}
1300: \end{figure}
1301: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1302: 
1303: Each curve in this figure corresponds to an experimental bound
1304: from Eqs.~(\ref{bound5}), (\ref{bound1}), (\ref{bound2}),
1305: (\ref{bound3}) and (\ref{bound4}); while the colored area is the
1306: overlap region satisfying all the constraints. The distinctive
1307: character of this figure is that $ b \to s \gamma $ requires a
1308: non-zero $\delta_{LL}$. The reason is that with the fixed
1309: parameters in Eq.~(\ref{susy-para}), especially with
1310: $\delta^d_{LR}= 0 $, the charged Higgs contribution enhances the
1311: SM contribution and, as a result, a none-zero gluino contribution
1312: is needed to cancel such effects. From Fig.~\ref{constr1} we see
1313: that the allowed region is mainly determined by $ b \to s \gamma$
1314: and $ B_s-\bar{B}_s$ mixing. To show the dependence of such
1315: allowed region on SUSY parameters, we vary the values of $M_A$,
1316: $M_S$, $M_{\tilde{g}}$ and $\delta_{LR}^d$ one at a time, and get
1317: the allowed region (colored area) in Figs.~\ref{constr2},
1318: \ref{constr3}, \ref{constr4} and \ref{constr5}, respectively.
1319: Explicitly, Fig.~\ref{constr2} corresponds to the parameters in
1320: Eq.~(\ref{susy-para}) but with $M_A = 200 $ GeV,
1321: Fig.~\ref{constr3} corresponds to the parameters in
1322: Eq.~(\ref{susy-para}) but with $M_S = 1000 $ GeV ,  and
1323: Fig.~\ref{constr4} and Fig.~\ref{constr5} are drawn in a similar
1324: way. The shift of the allowed region can be well understood by the
1325: properties of each constraint. Take Fig.~\ref{constr2} as an
1326: example. As $M_A$ becomes smaller, the charged Higgs contribution
1327: to $b \to s \gamma$ further enhance the SM contribution and
1328: consequently a larger $\delta_{LL}$ is needed to cancel such
1329: effect and satisfy the bound in Eq.~(\ref{bound1}). Since the
1330: constraint from $B_s-\bar{B}_s$ mixing is not changed, the overlap
1331: region diminishes gradually and finally vanishes for $M_A \simeq
1332: 200 $ GeV.
1333: 
1334: As the second example, we consider the following parameters as an
1335: input
1336: \begin{eqnarray}
1337: &&M_{Q_2}= 400  {\rm ~GeV}, \quad  M_{Q_3} = 1000  {\rm ~GeV},
1338: \quad M_{U_3} = 120  {\rm ~GeV}, \quad M_{\tilde{g}} = 196  {\rm
1339: ~GeV}, \quad M_A = 160  {\rm ~GeV},
1340: \nonumber \\
1341: &&X_t = 33  {\rm ~GeV}, \quad \mu = -330  {\rm ~GeV}, \quad
1342: M_2 = 860  {\rm ~GeV}, \quad M_{D_i} = M_{Q_1}= M_{D_1} = 500  {\rm ~GeV}, \nonumber \\
1343: && \delta_{LR}^d= 0.0026, \quad \delta_{RL}=\delta_{LR}= 0, \quad
1344: \delta_{RL}^d=\delta_{RR}^d=0.  \label{max-para}
1345: \end{eqnarray}
1346: This set corresponds to a point in the parameter space where $c g
1347: \to t$ is maximized for non-zero $\delta_{LR}$ (see following
1348: discussion about Table 1 and Table 2) and the results are depicted
1349: in Fig.~\ref{constr6}.
1350:  Since the squark masses are not universal, $\delta_{LR}>1$ can still satisfy all the
1351: constraints and thus is allowed. The reason for this is that all
1352: the constraints actually limit the size of the product $\delta_{LR} M_{Q_2}
1353: M_{U_3}$. For the case discussed here, $M_{Q_2}$ and $M_{U_3}$ are
1354: not large and thus a large $\delta_{LR}$ is allowed.
1355: 
1356: A common property of the above several figures is that
1357: $B_s-\bar{B}_s$ mixing and $b \to s \gamma$ require a small
1358: $\delta_{LL}$ value. This is a general feature, which accounts for
1359: the significant suppression of the maximal predictions for various
1360: top quark FCNC interactions after considering all the constraints
1361: (see the results in case I($\delta_{LR}=0$) of Table 1) . We also
1362: considered the constraints on $\delta_{RL}$ and $\delta_{RR}$,
1363: which, as we pointed out earlier in this section, do not
1364: significantly affect $b \to s \gamma$ and $B_s-\bar{B}_s$ mixing,
1365: and hence are constrained only by $\delta \rho$ and the Higgs
1366: boson mass. We found that by comparing with the constraints on
1367: $\delta_{LR}$ and $\delta_{LL}$, the constraints on these two
1368: mixing parameters, especially on $\delta_{RR}$, are rather weak.
1369: Let us consider parameters in Eq.~(\ref{susy-para}) as an example.
1370: For $\delta_{LL}=\delta_{LR}=0$, our results indicate that
1371: $\delta_{RR}$ and $\delta_{RL}$ should be less than $0.76$ and
1372: $0.46$, respectively. Such constraints come from $\delta M_W $ and
1373: $\delta \sin^2 \theta_{eff}$, and are sensitive to $X_t$. For $X_t
1374: =0 $,
1375:   the bounds will be relaxed to $0.98$ for
1376: $\delta_{RR}$ and $0.75$ for $\delta_{RL}$.
1377: 
1378: \subsection{Maximal predictions in MSSM for FCNC top quark processes}
1379: 
1380: With the constraints discussed above, we perform a scan over
1381: the SUSY parameter space to search for the maximal predictions of the
1382: MSSM on various top quark FCNC processes. We consider two cases: (I) only
1383: $\delta_{LL}\neq 0$ and (II) only $\delta_{LR} \neq 0$. For
1384: each case, we require the parameters to vary in the following
1385: ranges \footnote{There is a typo in Eq.~(15) of Ref.~\cite{cao}:
1386: $\delta_{LL}$ and $\delta_{LR}$ were required to vary between 0 and 2
1387: rather than between 0 and 1. We checked that allowing $\delta$ parameters
1388: to be larger than 2 does not affect the results in Table I, but more samples
1389: are needed to get the results in Table I. In our scan, we required
1390: $\delta_{RL}^d=\delta_{RR}^d=0$ and $\delta_{LL}, \delta_{LR} > 0$.
1391: Relaxing these requirements does not change our results but lowers
1392: the efficiency to search for the maximal predictions in a certain
1393: number of samples. In our scan, we also found that our results are
1394: not sensitive to the values of $M_{D_i}$, $M_{U_1}$, $M_{Q_1}$ and
1395: $A_b$.}
1396: \begin{eqnarray}
1397: && 2 < \tan\beta < 60, \quad  \quad \quad \quad \quad \quad  0< M_{Q_i,U_i, D_i}<1 {\rm ~TeV}, \nonumber \\
1398: && 94 {\rm ~GeV} < m_A <1 {\rm ~TeV}, \quad \quad 195 {\rm ~GeV}
1399: <m_{\tilde g} < 1 {\rm ~TeV},
1400:                                                                 \nonumber \\
1401: && 0< \delta_{LL}\ or \ \delta_{LR} < 2, \quad \quad \quad \quad -1 {\rm ~TeV}  < \mu, M_2 < 1 {\rm ~TeV}, \nonumber \\
1402: && 0< \delta_{LR}^b < 0.1, \quad \quad  \quad \quad \quad \quad \; -2 {\rm ~TeV} < A_{t,b} < 2 {\rm
1403: ~TeV} .
1404: \end{eqnarray}
1405: To manifest the effect of the combined constraints on the maximal rates,
1406:  we present two types of predictions: one by only
1407: requiring the squark, chargino and neutralino masses satisfy their
1408: current lower bounds
1409: \begin{eqnarray}
1410: m_{\tilde{u}}> 96  {\rm ~GeV}, \quad m_{\tilde{d}} > 89  {\rm ~GeV}, \quad
1411: m_{\tilde{\chi}^0} > 46  {\rm ~GeV}, \quad m_{\tilde{\chi}^+} > 94  {\rm ~GeV};
1412: \label{mass constraints}
1413: \end{eqnarray}
1414: and the other by imposing all constraints in Eqs.~(\ref{mass
1415: constraints}, \ref{bound1}, \ref{bound2}, \ref{bound3},
1416: \ref{bound4}). With five million samples for each process in
1417: either case, we obtain the maximal predictions and they are given in Table I.  From
1418: the table one can see that the combined constraints can significantly
1419: decrease the MSSM predictions for top-quark FCNC channels at the
1420: LHC, especially for the case I, and among these FCNC processes, $c g
1421: \to t h $ is the most affected one after imposing these constraints. The reason is that,
1422: as we pointed out before, there is a cancellation between the contribution
1423: from the effective interactions $\bar{t} c g$ and $\bar{t} c h$,
1424: and the constraints only allow a region with strong cancellation.
1425: 
1426: \vspace*{0.2cm}
1427: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1428: \noindent {\small Table 1: Maximal predictions for top-quark FCNC
1429: processes
1430:      induced by stop-scharm mixings via gluino-squark loops in the MSSM.
1431:      For the production channels we show the hadronic cross sections at the LHC
1432:      obtained from the given parton-level channels
1433:      and the corresponding charge-conjugate channels.
1434:      For the decays we show the branching ratios. LHC sensitivities listed in the last column are
1435:      for 100fb$^{-1}$ integrated luminosity.}
1436: \vspace*{0.1cm}
1437: 
1438: \begin{center}
1439: \begin{tabular}{|c|c|c|c|c|c|} \hline
1440: & \multicolumn{2}{c|}{$\delta_{LL}\neq
1441: 0$}&\multicolumn{2}{c|}{$\delta_{LR}\neq 0$} & LHC sensitivity \\
1442: \cline{2-5}
1443:         & constraints& constraints& constraints & constraints & at $3 \sigma$ level \\
1444:         &  masses&  all    & masses &   all   &   \\ \hline
1445:  $t \to ch$ & $1.2 \times 10^{-3} $ & $ 2.0 \times
1446: 10^{-5}$ & $2.5\times 10^{-2}$ & $ 6.0 \times 10^{-5}$ & $5.8 \times 10^{-5}$ \cite{tch-Aguilar} \\
1447: \hline $t \to c g $ & $5.0 \times 10^{-5}$ & $5.0 \times 10^{-6} $
1448: & $1.3 \times 10^{-4}$ & $3.2 \times 10^{-5}$ & $  - $   \\
1449: \hline $t \to c g g$ & $6.1 \times 10^{-5}$ & $7.1\times 10^{-6}$ & $1.5 \times 10^{-4}$ & $3.5 \times 10^{-5}$ & $-$ \\
1450: \hline $t \to c Z$ & $5.0 \times 10^{-6}$ & $5.7 \times 10^{-7}$
1451: &$1.2 \times 10^{-5}$ &
1452:   $1.8 \times 10^{-6}$                                  &  $3.6 \times 10^{-5}$ \cite{tcz-Han,tcz-Atlas}  \\ \hline
1453: $t \to c \gamma $ & $9.0 \times 10^{-7}$ & $1.5 \times 10^{-7}$ &
1454: $1.3 \times 10^{-6}$ & $5.2 \times 10^{-7}$ & $1.2 \times 10^{-5}$ \cite{tcr-Han,tcr-Beneke} \\
1455: \hline \hline
1456: $cg \to t$ & 1450 fb & 225 fb& 3850 fb & 950 fb         & 800 fb\cite{Hosch} \\
1457: \hline $gg \to t\bar{c}$ & 1400 fb & 240 fb & 2650 fb & 700 fb
1458:  & 1500 fb \cite{pptc-Han,pptc-background}
1459: \\ \hline $ c g \to tg$ & 800 fb & 85 fb & 1750 fb & 520 fb
1460: & 1500 fb \cite{pptc-Han,pptc-background}\\ \hline $cg \to t
1461: \gamma$ & 4 fb & 0.4 fb& 8 fb & 1.8 fb & 5 fb \cite{zt-Aguilar}
1462: \\ \hline $cg \to tZ$ & 11 fb & 1.5 fb & 17 fb & 5.7 fb & 35 fb \cite{zt-Aguilar}  \\
1463: \hline $c g \to th$ & 550 fb & 18 fb & 12000 fb & 24 fb & 200 fb \cite{tch-Aguilar}  \\
1464: \hline
1465: \end{tabular}
1466: \end{center}
1467: 
1468: \vspace*{0.2cm}
1469: \begin{center}
1470: \noindent {\small Table 2:  SUSY parameters leading to the maximal
1471: predictions for $\delta_{LR}\neq 0$ in Table 1.} \vspace*{0.2cm}
1472: 
1473: \vspace*{0.1cm}
1474: \begin{tabular}{|c|c|c|c|c|c|c|} \hline
1475: process & $M_{Q_2} (GeV) $  &$M_{Q_3} (GeV) $ & $M_{U_3} (GeV) $ &
1476: $X_t (GeV) $ & $m_{\tilde{g}} (GeV) $ &  $\delta_{LR}$ \\ \hline
1477: \hline $t \to ch$ & 1000 & 900 & 225 & 500 & 195 & \ \ 1.1\ \  \\
1478: \hline $t \to c g $ & 310 & 985 & 90 & 100  & 195 & \ \ 1.72\ \  \\
1479: \hline $t \to c g g$ & 310 & 980 & 90 & 35  & 195 & \ \ 1.75\ \  \\
1480: \hline $t \to c Z$ & 500 & 900 & 165 & 600 & 198 & \ \ 1.1 \ \
1481: \\ \hline
1482: $t \to c \gamma $ & 290 & 1000 & 90 & 20 & 196 & \ \ 1.72\ \  \\
1483: \hline \hline
1484: $cg \to t$ & 400 & 990 & 120 & 33 & 196 & 1.5 \\
1485: \hline $gg \to t\bar{c}$ & 310 & 990 & 85 & 80 & 196 & \ \ 1.8 \ \
1486: \\ \hline $ c g \to tg$ & 490 & 900 & 125 & 0 & 197 & \ \ 1.45 \ \  \\
1487:  \hline $cg \to t \gamma$ & 280 & 1000 & 85 & 25 & 197 & \ \ 1.78
1488:  \ \
1489:  \\ \hline $cg \to tZ$ & 370  & 920 & 80 & 115 & 196 & \ \ 1.86 \ \ \\
1490: \hline $c g \to th$ & 280 & 1000 & 85 & 23 & 197 & \ \ 1.77\ \ \\
1491: \hline
1492: \end{tabular}
1493: \end{center}
1494: 
1495: As one can see from Table 1, predictions for the case with
1496: non-zero $\delta_{LR}$ are larger than for the one with non-zero
1497: $\delta_{LL}$. In Table 2 we list the SUSY parameters leading to the
1498: maximal predictions in Table 1 for the case II ($\delta_{LR}\neq
1499: 0$).  It is seen from Table 2 that the `favorable' parameters for
1500: the maximal rates are process dependent. Since these parameters
1501: could be first tested by seeking top quark FCNC signals at LHC, in
1502: Table 3 we present the predictions for all processes with two sets
1503: of parameters, called `Point 1' and `Point 2', where $t\to c h$ and
1504: $c g \to t $ are required to be maximized, respectively. It is seen
1505: that `Point 1'  favors $cg\to t$ as the production channel but
1506: `Point 2' favors $t\to cgg$ among the decay modes.
1507: 
1508: Note that in Table 1 we only showed the cases of $\delta_{LL}\neq 0$
1509: and $\delta_{LR} \neq 0$. For $\delta_{RL}\neq 0$, we found that the
1510: maximal rate of $t \to cg$ is $1.3 \times 10^{-4}$ if only the
1511: squark mass constraints are included and $6 \times 10^{-5}$ with all
1512: the constraints. For $\delta_{RR} \neq 0$, the maximal rate of $t
1513: \to cg$ is $5.0 \times 10^{-5}$ with only the squark mass
1514: constraints and $4.85 \times 10^{-5}$ with all the constraints.
1515: These results can be easily understood since, as discussed in
1516: \cite{cao} and in this section, the constraints on $\delta_{RL}$ and
1517: $\delta_{RR}$ are weaker than those for $\delta_{LR}$ and
1518: $\delta_{RR}$, respectively.
1519: 
1520: \vspace*{0.2cm} \noindent {\small Table 3: SUSY predictions for the
1521: rates of top quark FCNC processes at two points of parameter space:
1522: `Point 1' maximizes $t \to c h$ and `Point 2' maximizes $c g \to t $
1523: channel. }
1524: 
1525: \vspace*{0.1cm}
1526: \begin{center}
1527: \begin{tabular}{|c|c|c|} \hline
1528: process & Point 1 & Point 2  \\ \hline $t \to c g$ & $2.0 \times
1529: 10^{-5} $ & $ 3.1 \times 10^{-5}$ \\ \hline $t \to c g g $ & $2.9
1530: \times 10^{-5} $ & $ 3.5 \times 10^{-5}$ \\ \hline $t \to c Z$ &
1531: $1.4 \times 10^{-6} $ & $ 1.2 \times 10^{-6}$  \\ \hline $t \to c
1532: \gamma $ & $2.6 \times 10^{-7}$& $ 5.2 \times 10^{-7}$ \\ \hline
1533: $t \to c h$ & $6.0 \times 10^{-5} $ & $ 7.0 \times 10^{-9}$ \\
1534: \hline $c g \to t$ & 660 fb & 950 fb \\ \hline $g  g \to t \bar{c}$
1535: & 450 fb & 690 fb \\ \hline $c g \to t g $ & 280 fb & 415 fb \\
1536: \hline $c g \to t Z$ & 2.3 fb & 4.9 fb \\ \hline $c g \to t \gamma $
1537: & 1.2 fb & 1.8 fb \\ \hline $c g \to t h $ & 24 fb & 8.5 fb \\
1538: \hline
1539: \end{tabular}
1540: \end{center}
1541: 
1542: \subsection{Top FCNC observability at the LHC}
1543: Now we consider the observability of these top quark FCNC
1544: interactions at the LHC. This issue has been intensively investigated
1545: in the effective Lagrangian  approach  for $t \to c
1546: h$ \cite{tch-Aguilar}, $t \to c Z$ \cite{tcz-Han,tcz-Atlas}, $t
1547: \to c \gamma$ \cite{tcr-Han,tcr-Beneke}, $p p \to t +X$
1548: \cite{Hosch}, $ p p \to t \bar{c}+X$
1549: \cite{pptc-Han,pptc-background}, $p p \to t g +X
1550: $\cite{pptc-Han,pptc-background}, $ p p \to t
1551: Z+X$\cite{zt-Aguilar}, $p p \to t \gamma +X $\cite{zt-Aguilar} and
1552: $p p \to t h +X$\cite{tch-Aguilar}. At the
1553: LHC, most of the top quarks will be pair-produced. One of the tops is
1554: assumed to subsequently decay as $t\to bW$ (normal mode) while
1555: the other top goes to
1556: one of the  above channels via FCNC interactions (exotic mode).
1557: %%%It is also natural to assume that in the search for single top
1558: %%%production, the top quark decays normally.
1559: 
1560: Due to the
1561: large QCD backgrounds at the LHC, the search for these processes
1562: must be performed in the decay channels $W \to \ell
1563: \bar{\nu}_\ell$ ($\ell=e,\mu$) for $W$ boson, $Z\to \ell^+ \ell^-$
1564: for $Z$ boson and $ h \to b \bar{b}$ for Higgs boson. For any of
1565: these reactions, top quark reconstruction is required to extract the
1566: signal from its background. In Table 4, we list the signals and
1567: the main backgrounds. The detailed Monte Carlo
1568: simulations for the signals and backgrounds can be found in the
1569: corresponding literature listed in the last column of Table 1,
1570: where the LHC sensitivity for these processes is quoted.
1571: Although these sensitivities are based on the effective Lagrangian
1572: approach and may be not perfectly applicable to the MSSM, we can
1573: take them as a rough criteria to estimate the observability of
1574: these channels. Comparing these sensitivities with the SUSY
1575: predictions, one can see that only the maximal predictions for $c
1576: g \to t $ and $ t \to c h $ are slightly larger than the
1577: corresponding LHC sensitivities. This implies that the study of these
1578: processes may provide the first insight about top quark FCNC. From
1579: these sensitivities one can also see that, although the $ g g \to t
1580: \bar{c}$ channel is now the most extensively studied among the FCNC
1581: production channels \cite{pptc-mssm}, its observation needs a
1582: higher luminosity. This is because $ g g \to t \bar{c}$ has large
1583: irreducible backgrounds from single top productions in the SM
1584: \cite{pptc-background}. Note that in Table 1, we did not list the
1585: sensitivity of the LHC to $ t \to c g$, which has not been
1586: investigated because of a general belief that this decay is not well
1587: suited for detecting $\bar{t} c g$ interaction
1588: \cite{Aguilar-Saavedra:2004wm}. In fact, this decay was once
1589: investigated for the Tevatron \cite{tcg-Han}, and it was found
1590: that due to the large background, namely $W$ boson plus three
1591: jets, only a branching ratio as large as $5 \times 10^{-3}$ may be
1592: accessible with 10 fb$^{-1}$ integrated luminosity. This rate is
1593: about one order larger than that for $ t \to c \gamma$ at the
1594: Tevatron with the same luminosity \cite{tcr-Han}. For the decay $t
1595: \to c g g$, there is an additional jet in its signal and its
1596: observability needs to be studied by a detailed Monte Carlo
1597: simulation.
1598: 
1599: \vspace*{0.3cm} \noindent {\small Table 4: Experimental signature
1600: and main background for FCNC top rare decays and productions at
1601: the LHC. The top quarks are assumed to decay $t \to W^+ b \to
1602: \ell^+ \nu_\ell b$, and $Z$ and $h$ bosons decay in the channel $Z
1603: \to \ell^+ \ell^-$ and $h \to b \bar b$, respectively.}
1604: \vspace*{0.2cm}
1605: 
1606: \begin{center}
1607: \begin{tabular}{lccclcc}
1608: \hline Process & Signal & Background & & ~~~~Process &  Signal &
1609: Background
1610: \\ \hline
1611: $t \bar t$, $t \to c g$ & $j j \ell \nu b$ & $Wjjj$ & & ~~~~$c g \to
1612: t$ & $\ell \nu b$ & $Wj$  \\ \hline $t \bar t$, $t \to c g g$ & $j j
1613: j \ell \nu b$ & $Wjjj j$ && ~~~~$g g \to t \bar{c}$ & $\ell \nu b j$
1614: & $t j$  \\ \hline  & & & & ~~~~$c g \to t g$ & $\ell \nu b j $ & $
1615: t j $
1616: \\  \hline $t \bar t$, $t \to c Z$ &$\;\;$ $\ell^+ \ell^- j \ell \nu
1617: b$ & $ZWjj$ && ~~~~$c g \to t Z$ &$\;\;$ $\ell^+ \ell^- \ell \nu b$
1618: & $ZWj$  \\ \hline $t \bar t$, $t \to c \gamma$ & $\gamma j \ell \nu
1619: b$ & $\gamma Wjj$ & & ~~~~$c g \to t \gamma $ & $\gamma \ell \nu b$
1620: & $\gamma Wj$  \\ \hline $t \bar t$, $t \to c h$ & $b \bar b j \ell
1621: \nu b$ & $Wb \bar b jj$ & &
1622: ~~~~$c g \to t h$ & $b \bar b \ell \nu b$ & $t \bar t$  \\
1623: \hline
1624: \end{tabular}
1625: \end{center}
1626: 
1627: \section{Conclusions}
1628: In this paper, we  investigated systematically the SUSY-induced top
1629: quark FCNC processes at the LHC, which includes various decay modes
1630: and production channels. We performed a comparative study for all
1631: the decay modes and for all the production channels so that one can
1632: see clearly which decay mode or production channel can have a
1633: relatively large rate. The dependence of these channels
1634: on the relevant SUSY parameters is investigated in detail and its
1635: properties  are analyzed. We note that such a global study of the
1636: top quark FCNC processes has been done only in a model independent
1637: way \cite{Aguilar-Saavedra:2004wm}. We also analyzed the
1638: characteristics of the `favorable region' in SUSY parameter space
1639: where the FCNC processes are maximized. After getting an
1640: understanding of these processes, we examined the effects of all the
1641: direct and indirect experimental constraints on the scharm-stop
1642: flavor mixings and scanned the parameter space to find their maximal
1643: rates with these constraints imposed. We found that $ c g \to t$ and
1644: $ t \to c h$ are the most likely channels to be observable at the
1645: LHC if the MSSM is the correct scenario beyond the SM.
1646: 
1647: \section*{Acknowledgment}
1648: 
1649: This work is supported in part by a fellowship from the Lady Davis
1650: Foundation at the Technion, by the Israel Science Foundation (ISF), by NSERC
1651: of Canada under Grant No. SAP01105354,
1652: the National Natural Science
1653: Foundation of China under Grant No. 10475107 and 10505007,  the
1654: Grant-in-Aid for Scientific Research (No. 14046201) from the Japan
1655: Ministry of Education, Culture, Sports, Science and Technology,
1656: and by the IISN and the Belgian science policy office (IAP V/27).
1657: 
1658: \appendix
1659: 
1660: \section{Expressions for the Loop results}
1661: 
1662: If significant flavor mixings exist only between left-handed
1663: scharm with stops, the squark states $\tilde{c}_L$, $ \tilde{t}_L$
1664: and $\tilde{t}_R$ will mix together to induce various top quark
1665: FCNCs. In this case, other squark states only serve as spectators
1666: to the processes considered in this paper. So in the actual
1667: calculation, we only need to consider the squark mass matrix for
1668: $(\tilde{t}_L, \tilde{t}_R, \tilde{c}_L)$, which is given by
1669: \begin{eqnarray}
1670: \left ( \begin{array}{ccc}
1671:     M_{Q_3}^2 + m_t^2 + (\frac{1}{2} - \frac{2}{3} s_W^2 ) m_Z^2 \cos 2 \beta  &  m_t ( A_t -\mu \cot \beta)&  \delta_{LL} M_{Q_2} M_{Q_3} \\
1672:      m_t ( A_t -\mu \cot \beta)  &  M_{U_3}^2 + m_t^2 + \frac{2}{3} s_W^2 \cos 2 \beta  &  \delta_{LR} M_{Q_2} M_{U_3} \\
1673:     \delta_{LL} M_{Q_2} M_{Q_3}    &  \delta_{LR} M_{Q_2} M_{U_3} & M_{Q_3}^2 + (\frac{1}{2} - \frac{2}{3} s_W^2 ) m_Z^2 \cos 2 \beta  \end{array}  \right
1674:     ).
1675: \end{eqnarray}
1676: This mass matrix can be diagonalized by a unitary matrix $V$, and
1677: it enters the squark interactions as in Eq.~(\ref{interaction}).
1678: 
1679: In this Appendix, we list the expressions for $\Sigma$ and
1680: $\Gamma^{\bar{t}c g}_\mu $ in Eq.~(\ref{eff}) which are needed to
1681: get the effective $\bar{t} c g $ vertex. We also list the
1682: expressions for $\Gamma^{\bar{t}c \gamma} $, $\Gamma^{\bar{t} c Z}
1683: $ and $\Gamma^{\bar{t}c h} $ to calculate other effective
1684: vertices. Before presenting these expressions, we define the
1685: following abbreviations:
1686: \begin{eqnarray}
1687: R^a & = &  \sum_{\lambda=1}^3 V_{1 \lambda} V_{\lambda 3}^\dag
1688: B(p, m_{\tilde{g}},  m_{\lambda}), \quad R^b = R^a |_{V_{1
1689: \lambda} \to V_{2 \lambda}},
1690: \end{eqnarray}
1691: \begin{eqnarray}
1692: R^c & = &  \sum_{\lambda=1}^3 V_{1 \lambda} V_{\lambda 3}^\dag
1693: C(-p_c, p_c - p_t,  m_{\tilde{g}},  m_{\lambda},  m_{\lambda} ),
1694: \quad R^d = R^c |_{V_{1 \lambda} \to V_{2 \lambda}},
1695: \end{eqnarray}
1696: \begin{eqnarray}
1697: R^e & = &  \sum_{\lambda=1}^3 V_{1 \lambda} V_{\lambda 3}^\dag
1698: C(-p_c, p_c - p_t,  m_{\lambda},  m_{\tilde{g}},  m_{\tilde{g}}),
1699: \quad R^f = R^e |_{V_{1 \lambda} \to V_{2 \lambda}},
1700: \end{eqnarray}
1701: \begin{eqnarray}
1702: R^g & = &  \sum_{\lambda=1}^3 V_{1 \lambda} V_{\lambda 3}^\dag
1703: C(-p_b, p_b - p_t,  m_{\tilde{g}},  m_{\tilde{b}_L},  m_{\lambda}
1704: ),
1705: \end{eqnarray}
1706: \begin{eqnarray}
1707: R^h & = &  \sum_{\rho, \lambda=1}^3 V_{1 \lambda} F^Z_{\lambda
1708: \rho} V_{\rho 3}^\dag C(-p_c, p_c - p_t,  m_{\tilde{g}},
1709: m_{\rho},  m_{\lambda} ), \quad R^i = R^h |_{V_{1 \lambda} \to
1710: V_{2 \lambda}},
1711: \end{eqnarray}
1712: \begin{eqnarray}
1713: R^j & = &  \sum_{\rho, \lambda=1}^3 V_{1 \lambda} F^h_{\lambda
1714: \rho} V_{\rho 3}^\dag C(-p_c, p_c - p_t,  m_{\tilde{g}},
1715: m_{\rho},  m_{\lambda} ), \quad R^k = R^j |_{V_{1 \lambda} \to
1716: V_{2 \lambda}},
1717: \end{eqnarray}
1718: \begin{eqnarray}
1719: R^l & = &  \sum_{\lambda=1}^3 V_{1 \lambda} V_{\lambda 3}^\dag
1720: D(-p_t, p_2, p_1, m_{\lambda}, m_{\tilde{g}}, m_{\tilde{g}},
1721: m_{\tilde{g}} ), \quad R^m = R^l |_{V_{1 \lambda} \to V_{2
1722: \lambda}},
1723: \end{eqnarray}
1724: \begin{eqnarray}
1725: R^n & = &  \sum_{\lambda=1}^3 V_{1 \lambda} V_{\lambda 3}^\dag D(
1726: p_1, p_2, -p_t, m_{\lambda}, m_{\lambda}, m_{\lambda},
1727: m_{\tilde{g}} ),
1728:  \quad R^o = R^n |_{V_{1 \lambda} \to V_{2 \lambda}},
1729: \end{eqnarray}
1730: \begin{eqnarray}
1731: R^p & = &  \sum_{\lambda=1}^3 V_{1 \lambda} V_{\lambda 3}^\dag D(
1732: -p_t, p_2, - p_c, m_{\lambda}, m_{\tilde{g}}, m_{\tilde{g}},
1733: m_{\lambda} ), \quad R^q = R^p |_{V_{1 \lambda} \to V_{2
1734: \lambda}},
1735: \end{eqnarray}
1736: where $\rho$ and $\lambda$ are squark indices in mass eigenstate,
1737: $p_i$ is particle momentum, B, C and D are loop
1738: functions\cite{Hooft}, and $F^Z_{\lambda \rho}$ and $F^h_{\lambda
1739: \rho}$ are interaction coefficients for $\tilde{q}_\lambda^*
1740: \tilde{q}_\rho Z $ and $\tilde{q}_\lambda^* \tilde{q}_\rho h $
1741: interactions respectively, which are given by
1742: \begin{eqnarray}
1743: F^Z_{\lambda \rho} &= & (1 -\frac{4}{3} s_W^2 ) \delta_{\rho
1744:     \lambda} - V_{\lambda 2}^\dag V_{2 \rho}   \\
1745: F^h_{\lambda \rho} & = & -\frac{g}{m_W \sin \beta } \left ( m_t^2
1746:    \cos \alpha ( V_{\lambda 1}^\dag V_{1 \rho} + V_{\lambda 2}^\dag
1747: V_{2 \rho} ) + \frac{m_t ( A_t \cos \alpha + \mu \sin \alpha )}{2}
1748: (V_{\lambda 1}^\dag V_{2 \rho} +
1749: V_{\lambda 2}^\dag V_{1 \rho}) \right. \nonumber\\
1750: & & \left .  + \frac{\delta_{LR} M_{Q_2} M_{U_3} \cos \alpha}{2}
1751: (V_{\lambda 3}^\dag V_{2 \rho} + V_{\lambda 2}^\dag V_{3 \rho})
1752: \right ).
1753: \end{eqnarray}
1754: Then after factoring out the common factor $\alpha_s/4 \pi$, we
1755: obtain the expressions for $\Sigma$ and $\Gamma $s:
1756: \begin{eqnarray}
1757: \Sigma (p) &=& - 2 C_F \left ( \gamma^\mu R^a_\mu + m_{\tilde{g}}
1758:                  R^b_0 \right ) P_L,   \\
1759: \Gamma^{\bar{t} c g}_\mu &=& \frac{1}{3} g_s T^a \left ( 2
1760:   \gamma^\nu R^c_{\mu \nu} - 2 m_{\tilde{g}} R^d_\mu - (p_t +
1761: p_c)_\mu (\gamma^\nu R^c_\nu - m_{\tilde{g}} R^d_0 ) \right ) P_L
1762:  + 3 g_s T^a \left ( \p_slash_t \gamma_\mu \p_slash_c R^e_0 -
1763: \p_slash_t \gamma_\mu \gamma^\nu R^e_\nu  \right .  \nonumber\\
1764: && \left . - \gamma^\nu \gamma_\mu \p_slash_c R^e_\nu + \gamma^\nu
1765: \gamma_\mu \gamma^\lambda R^e_{\nu \lambda} + m_{\tilde{g}}^2
1766: \gamma_\mu R^e_0 - m_{\tilde{g}} \gamma_\mu \p_slash_c R^f_0 + 2
1767: m_{\tilde{g}} R^f_\mu - m_{\tilde{g}} \p_slash_t \gamma_\mu R^f_0
1768: \right ) P_L,    \\
1769: \Gamma^{\bar{b} c W}_\mu &=& \frac{g}{\sqrt{2}} C_F \gamma^\nu P_L
1770: (p_{b \mu} R^g_\nu - R^g_{\mu \nu} ),  \\
1771: \Gamma^{\bar{t} c Z}_\mu &=& \frac{g}{c_W} C_F \left ( (p_t + p_c
1772: )_\mu ( \gamma^\nu R^h_\nu - m_{\tilde{g}} R^i_0 ) - 2 (
1773: \gamma^\nu R^h_{\mu \nu} - m_{\tilde{g}} R^i_\mu ) \right ) P_L, \\
1774: \Gamma^{\bar{t} c \gamma}_\mu &=& - 2 e Q_u C_F \left ( 2
1775: \gamma^\nu R^c_{\mu \nu} - 2 m_{\tilde{g}} R^d_\mu - (p_t +
1776: p_c)_\mu (\gamma^\nu R^c_\nu - m_{\tilde{g}} R^d_0 ) \right ) P_L,\\
1777: \Gamma^{\bar{t} c h} &=&  -2 C_F \left ( \gamma^\mu R^j_\mu -
1778: m_{\tilde{g}} R^k_0 \right ) P_L,
1779: \end{eqnarray}
1780: where $C_F$ is the quadratic Casimir operator of the fundemantal
1781: representation of $SU(3)_C$.
1782: 
1783: For the box diagrams in Fig.~\ref{SQCD diagram}, their results are
1784: given by
1785: \begin{eqnarray}
1786: M_d &=& -2 \alpha_s^2 T^e T^a f^{d e b} f^{c b a}
1787: \varepsilon^d_\lambda (p_2) \varepsilon^c_\rho (p_1) \bar{u}_t
1788: \left ( - m_{\tilde{g}} m_t \gamma^\lambda
1789: (\p_slash_t - \p_slash_2 ) \gamma^\rho R^m_0  \right.\nonumber \\
1790: && +  m_{\tilde{g}} m_t \gamma^\lambda \gamma^\nu \gamma^\rho
1791: R^m_\nu + m_{\tilde{g}} \gamma^\nu \gamma^\lambda (\p_slash_t -
1792: \p_slash_2 ) \gamma^\rho R^m_\nu - m_{\tilde{g}} \gamma^\mu
1793: \gamma^\lambda \gamma^\nu
1794: \gamma^\rho R^m_{\mu \nu}  \nonumber \\
1795: &&  \quad \   - m_t \gamma^\lambda (\p_slash_t - \p_slash_2 )
1796: \gamma^\rho \gamma^\nu R^l_\nu + m_t \gamma^\lambda \gamma^\mu
1797: \gamma^\rho \gamma^\nu R^l_{\mu \nu} + \gamma^\mu \gamma^\lambda
1798: (\p_slash_t - \p_slash_2 )  \gamma^\rho \gamma^\nu R^l_{\mu \nu} -
1799: \gamma^\mu \gamma^\lambda \gamma^\nu \gamma^\rho \gamma^\sigma
1800: R^l_{\mu \nu \sigma} \nonumber \\
1801:  && \left . \quad \  + m_{\tilde{g}}^2 m_t \gamma^\lambda \gamma^\rho R^l_0 -
1802:  m_{\tilde{g}}^2 \gamma^\nu \gamma^\lambda \gamma^\rho R^l_\nu +
1803:  m_{\tilde{g}} m_t \gamma^\lambda \gamma^\rho \gamma^\nu R^m_\nu
1804:  - m_{\tilde{g}} \gamma^\mu \gamma^\lambda \gamma^\rho \gamma^\nu R^m_{\mu \nu} \right ) P_L
1805:  v_c,  \\
1806: M_e &=& - 8 \alpha_s^2 T^a T^d T^c T^a \varepsilon^d_\lambda (p_2)
1807: \varepsilon^c_\rho (p_1) \bar{u}_t \left (  \gamma^\nu R^n_{\nu
1808: \rho \lambda} + \gamma^\nu p_{1 \lambda} R^n_{\nu \rho} +
1809: m_{\tilde{g}} R^o_{\rho \lambda} + m_{\tilde{g}} p_{1 \lambda}
1810: R^o_\rho \right ) P_L v_c,  \\
1811: M_f &=& - 4 \alpha_s^2 T^a T^c T^b f^{d a b} \varepsilon^d_\lambda
1812: (p_2) \varepsilon^{c}_\rho (p_1)  \bar{u}_t
1813: \left ( m_t \gamma^\lambda (\p_slash_t - \p_slash_2 )R^p_\rho  \right.\nonumber \\
1814: && - m_t \gamma^\lambda \gamma^\nu R^p_{\nu \rho} - \gamma^\nu
1815: \gamma^\lambda  (\p_slash_t - \p_slash_2 ) R^p_{\nu \rho} +
1816: \gamma^\mu \gamma^\lambda \gamma^\nu R^p_{\mu \nu \rho} -
1817: m_{\tilde{g}} m_t \gamma^\lambda R^q_\rho \nonumber \\
1818:  && \left . \quad \  + m_{\tilde{g}} \gamma^\nu \gamma^\lambda
1819:  R^q_{\rho \nu} - m_{\tilde{g}} \gamma^\lambda (\p_slash_t - \p_slash_2
1820:  ) R^q_\rho + m_{\tilde{g}} \gamma^\lambda \gamma^\nu R^q_{\rho
1821:  \nu} + m_{\tilde{g}}^2 \gamma^\lambda R^p_\rho \right ) P_L v_c.
1822: \end{eqnarray}
1823: 
1824: \begingroup\raggedright\begin{thebibliography}{99}
1825: 
1826: \bibitem{tcvh-sm} For the FCNC top quark decays in the SM, see,
1827:                   G.~Eilam, J.~L.~Hewett and A.~Soni, \PRD44, 1473 (1991);
1828:                    Erratum-ibid.\ D {\bf 59}  039901 (1999).
1829:                    B.~Mele, S.~Petrarca and A.~Soddu, \PLB435, 401
1830:                   (1998);A.~Cordero-Cid, J.~M.~Hernandez, G.~Tavares-Velasco and J.~J.~Toscano,
1831:                    Phys.\ Rev.\ D {\bf 73}, 094005 (2006).
1832:                   %%CITATION = HEP-PH 0411188;%%
1833:                   %%CITATION = PHRVA,D44,1473;%%
1834:                   %%CITATION = PHLTA,B435,401;%%
1835: \bibitem{tcgg-sm} G.~Eilam, M.~Frank and I.~Turan, \PRD73, 053011 (2006).
1836:                   %%CITATION = PHRVA,D73,053011;%%
1837: \bibitem{eetc-sm} C.-H. Chang, X.-Q. Li, J.-X. Wang and M.-Z. Yang, \PLB313, 389 (1993)
1838:                   C.-S. Huang, X.-H. Wu and S.-H. Zhu, \PLB452, 14 (1999).
1839:                    %%CITATION = PHLTA,B313,389;%%
1840:                    %%CITATION = PHLTA,B452,14;%%
1841: \bibitem{Larios:2006pb} For a recent review, see F.~Larios, R.~Martinez and M.~A.~Perez,
1842:                         Int.\ J.\ Mod.\ Phys.\ A {\bf 21}, 3473 (2006).
1843:                         %%CITATION = HEP-PH 0605003;%%
1844: \bibitem{Aguilar-Saavedra:2004wm}
1845:               J.~A.~Aguilar-Saavedra,  Acta Phys.\ Polon.\ B {\bf 35}, 2695 (2004).
1846:               %%CITATION = HEP-PH 0409342;%%
1847: \bibitem{Aguilar-saavedra:linear}
1848:                  J.~A.~Aguilar-Saavedra and T.~Riemann, hep-ph/0102197;
1849:                  J.~A.~Aguilar-Saavedra, \PLB502, 115(2001).
1850:                  %%CITATION = HEP-PH 0102197;%%
1851:                  %%CITATION = HEP-PH 0012305;%%
1852: \bibitem{cdfd0}   M.~Paulini, hep-ex/9701019; J.~Incandela (CDF), FERMILAB-CONF-95/237-E(1995);
1853:                   D.~Gerdes, hep-ex/9706001; T.~J.~Lecompte (CDF), FERMILAB-CONF-96/021-E (1996);
1854:                   A.~P.~Heinson (D0), hep-ex/9605010.
1855:                    %%CITATION = HEP-EX 9701019;%%
1856:                    %%CITATION = HEP-EX 9706001;%%
1857:                    %%CITATION = HEP-EX 9605010;%%
1858: \bibitem{tcv-mssm}  For $t \to cV$ in the MSSM, see,
1859:                     C.~S.~Li, R.~J.~Oakes and J.~M.~Yang, \PRD49, 293 (1994);
1860:                     G.~Couture, C.~Hamzaoui and H.~Konig, \PRD52, 1713 (1995);
1861:                     J.~L.~Lopez, D.~V.~Nanopoulos and R.~Rangarajan, \PRD56, 3100  (1997);
1862:                     G.~M.~de Divitiis, R.~Petronzio and L.~Silvestrini, \NPB504, 45 (1997);
1863:                     J.~M.~Yang, B.-L.~Young and X.~Zhang, \PRD58, 055001 (1998);
1864:                     C.~S.~Li, L.~L.~Yang and L.~G.~Jin, Phys.\ Lett.\ B {\bf 599}, 92
1865:                     (2004);
1866:                     D. Delepine and S. Khalil, \PLB599, 62 (2004);
1867:                     M.~Frank and I.~Turan, \PRD74, 073014 (2006).
1868:                     %%CITATION = PHRVA,D49,293;%%
1869:                     %%CITATION = PHRVA,D52,1713;%%
1870:                     %%CITATION = PHRVA,D56,3100;%%
1871:                     %%CITATION = NUPHA,B504,45;%%
1872:                     %%CITATION = PHRVA,D58,055001;%%
1873:                     %%CITATION = PHLTA, B599, 92;%%
1874:                     %%CITATION = PHRVA,D74,073014;%%
1875: \bibitem{tch-mssm}  For $t \to ch$ in the MSSM, see,
1876:                     J.~M.~Yang and C.~S.~Li, \PRD49, 3412 (1994);
1877:                     J.~Guasch and J.~Sola, \NPB562, 3 (1999);
1878:                     G. Eilam, {\it et al.}, \PLB510, 227 (2001);
1879:                J. L. Diaz-Cruz, H.-J. He, C.-P. Yuan \PLB179,530 (2002).
1880:                     %%CITATION = PHRVA,D49,3412;%%
1881:                     %%CITATION = NUPHA,B562,3;%%
1882:                     %%CITATION = PHLTA,B510,227;%%
1883: \bibitem{eetc-mssm} For top-charm associated productions in the MSSM at linear colliders,
1884:                     see,
1885:                     J. M. Yang, Annals Phys. {\bf 316}, 529 (2005);
1886:                     J. Cao, Z. Xiong and J. M. Yang, \NPB651, 87 (2003);
1887:                     C.~S.~Li, X.~Zhang and S.~H.~Zhu, \PRD60, 077702 (1999).
1888:                       %%CITATION = HEP-PH 0409351;%%
1889:                       %%CITATION = NUPHA,B651,87;%%
1890:                       %%CITATION = PHRVA,D60,077702;%%
1891: \bibitem{pptc-mssm} For top-charm associated productions in the MSSM at hadron colliders, see,
1892:                     J.~J.~Liu, C.~S.~Li, L.~L.~Yang and L.~G.~Jin, \NPB705, 3 (2005);
1893:                      J. Guasch, {\it et al.}, Nucl. Phys. Proc. Suppl. 157, 152 (2006) [hep-ph/0601218].
1894:                     %%CITATION = NUPHA, B705,3;%%
1895:                     %%CITATION = HEP-PH, 0601218;%%
1896: \bibitem{Gad-pptc-mssm} G.~Eilam, M.~Frank and I.~Turan, \PRD74, 035012 (2006).
1897:                         %%CITATION = PHRVA,D74,035012;%%
1898: \bibitem{tc-TC2}   For exotic top production processes in TC2 models, see,
1899:                     H. J. He and C. P. Yuan, \PRL83, 28(1999);
1900:                     G. Burdman, \PRL83,2888(1999);
1901:                     J. Cao, Z. Xiong and J. M. Yang, \PRD67, 071701 (2003);  C.~Yue, {\it et al.}
1902:                     Phys.\ Lett.\ B {\bf 496}, 93 (2000); C.~Yue,
1903:                     {\it et al.}, \PLB525, 301(2002); J. Cao, {\it
1904:                     et al.} \PRD70, 114035 (2004);  F. Larios and F. Penunuri,
1905:                     \JPG30, 895(2004); J. Cao, {\it et al.} \EPJC41, 381
1906:                     (2005).
1907:                     %%CITATION = PRLTA,83,28;%%
1908:                     %%CITATION = PRLTA,83,2888%%
1909:                     %%CITATION = PHRVA,D67,071701;%%
1910:                     %%CITATION = PHLTA,B496,301;%%
1911:                     %%CITATION = PHRVA,D70,114035;%%
1912:                     %%CITATION = HEP-PH 0311056;%%
1913:                     %%CITATION = HEP-PH 0311166;%%
1914: \bibitem{tcv-TC2}  For FCNC top quark decays in TC2 theory, see,
1915:                    X.~L. Wang  {\it et al.}, \PRD50, 5781 (1994);
1916:                    C.~Yue, {\it et al.}, \PRD64, 095004 (2001);
1917:                    G.~Lu, F.~Yin, X.~Wang and L.~Wan, \PRD68, 015002 (2003).
1918:                   %%CITATION = PHRVA,D50,5781;%%
1919:                   %%CITATION = PHRVA,D64,095004;%%
1920:                   %%CITATION = PHRVA,D68,015002;%%
1921: \bibitem{other}  J.~L.~Diaz-Cruz, {\em et~al.} \PRD41, 891(1990);
1922:                   G.~Eilam, {\em et.} in Ref.~\cite{tcvh-sm};
1923:                  X.L.~Wang et al., \JPG20, L91 (1994); \CTP24, 359 (1995);
1924:                  D. Atwood, L.~Reina and A.~Soni, \PRD53, 1199 (1996);
1925:                  S.~Bar-Shalom, {\em et~al.}, \PRL79, 1217(1997);
1926:                  W.~S.~Hou, G.-L.~Lin and C.-Y.~Ma, \PRD56, 7434(1997);
1927:                  S.~Bejar, J.~Guasch and J.~Sola, \NPB600, 21 (2001);
1928:                  R. A. Diaz, R. Martinez, J.-A. Rodriguez, hep-ph/0103307;
1929:                   T.~Han and J.~Hewett, \PRD60, 074015 (1999);
1930:                   F. del Aguila, J. A. Aguilar-Saavedra, R. Miquel, \PRL82, 1628 (1999); E.~O.~Iltan,
1931:                   \PRD65, 075017 (2002); E.~O.~Iltan and I.~Turan \PRD67, 015004 (2003);
1932:                  J. A. Aguilar-Saavedra, B. M. Nobre, \PLB553, 251
1933:                  (2003);  R.~Gaitan, O.~G.~Miranda and L.~G.~Cabral-Rosetti, \PRD72, 034018
1934:                  (2005); A.~Cordero-Cid, G.~Tavares-Velasco and J.~J.~Toscano,
1935:                  \PRD72, 057701 (2005); M.~Frank and I.~Turan, \PRD72, 035008
1936:                  (2005); A.~Arhrib, K.~Cheung, C.~W.~Chiang and T.~C.~Yuan,
1937:                  \PRD73, 075015 (2006); H.~Zhou, W.~G.~Ma, Y.~Jiang, R.~Y.~Zhang and L.~H.~Wan,
1938:                   hep-ph/0107293; H.~Zhou, W.~G.~Ma and R.~Y.~Zhang, hep-ph/0208170.
1939:                   P.~M.~Ferreira and R.~Santos, \PRD73, 054025
1940:                   (2006); P.~M.~Ferreira and R.~Santos, \PRD74, 014006
1941:                   (2006).
1942:                   %%CITATION = HEP-PH 0604144;%%
1943:                   %%CITATION = HEP-PH 0601078;%%
1944:                   %%CITATION = HEP-PH 0111318;%%
1945:                   %%CITATION = HEP-PH 0208170;%%
1946:                   %%CITATION = HEP-PH 0107293;%%
1947:                   %%CITATION = HEP-PH 0410268;%%
1948:                   %%CITATION = HEP-PH 0506197;%%
1949:                   %%CITATION = HEP-PH 0602175;%%
1950:                   %%CITATION = HEP-PH 0507135;%%
1951:                   %%CITATION = PHRVA,D41,891;%%
1952:                   %%CITATION = PHRVA,D44,1473;%%
1953:                   %%CITATION = PHRVA,D50,5781;%%
1954:                   %%CITATION = JPHGB,20,L91;%%
1955:                   %%CITATION = CTPMD,24,359;%%
1956:                   %%CITATION = PHRVA,D53,1199;%%
1957:                   %%CITATION = PHRVA,D79,1217;%%
1958:                   %%CITATION = PHRVA,D56,7434;%%
1959:                   %%CITATION = NUPHA,B600,21;%%
1960:                   %%CITATION = HEP-PH 0103307;%%
1961:                   %%CITATION = PHRVA,D60,074015;%%
1962:                   %%CITATION = PRLTA,82,1628;%%
1963:                   %%CITATION = HEP-PH 0210360;%%
1964: \bibitem{susyflavor} S. Dimopoulos and D. Sutter, \NPB452, 496 (1996);
1965:                F. Gabbiani, {\it et al.}, \NPB477, 321 (1996);
1966:                 M.~Misiak, S.~Pokorski and J.~Rosiek,
1967:                 hep-ph/9703442.
1968:                %%CITATION = NUPHA,B452,496;%%
1969:               %%CITATION = NUPHA,B477,321;%%
1970:               %%CITATION = HEP-PH 9703442;%%
1971: \bibitem{cao} J.~Cao, G.~Eilam, K.~Hikasa and J.~M.~Yang, \PRD74, 031701
1972:               (2006).
1973:               %%CITATION = HEP-PH 0604163;%%
1974: \bibitem{mssmfeynmanrule} J.~Rosiek, hep-ph/9511250.
1975:               %%CITATION = HEP-PH 9511250;%%
1976: \bibitem{bsr} T. Besmer, C. Greub, T.Hurth, \NPB609, 359 (2001);
1977:               F. Borzumati, {\it et al.}, \PRD62, 075005(2000).
1978:               %%CITATION = PHRVA,D62,075005;%%
1979:               %%CITATION = NUPHA,B609,359;%%
1980: \bibitem{Ball} M.~Ciuchini, E.~Franco, A.~Masiero and L.~Silvestrini,
1981:                hep-ph/0307194; P.~Ball, S.~Khalil and E.~Kou, \PRD69, 115011 (2004);
1982:                M. Ciuchini and L. Silvestrini, \PRL97, 021803 (2006).
1983:                %%CITATION = HEP-PH 0307194;%%
1984:                %%CITATION = PHRVA,D69,115011;%%
1985:                %%CITATION = PRLTA,97,021803;%%
1986: \bibitem{B-summary} J.~Hewett {\it et al.}, hep-ph/0503261.
1987:                %%CITATION = HEP-PH 0503261;%%
1988: \bibitem{mass insertion} L.~J.~Hall, V.~A.~Kostelecky and S.~Raby,
1989:   %``New Flavor Violations In Supergravity Models,''
1990:     \NPB 267, 415 (1986); F.~Gabbiani and A.~Masiero,
1991:   %``Fcnc In Generalized Supersymmetric Theories,''
1992:            \NPB 322, 235 (1989); J.~S.~Hagelin, S.~Kelley and T.~Tanaka, \NPB415, 293 (1994).
1993:                 %%CITATION = NUPHA,B267,415;%%
1994:                  %%CITATION = NUPHA,B322,235;%%
1995:                 %%CITATION = NUPHA,B415,293;%%
1996: \bibitem{sugra} A.H. Chamseddine, R. Arnowitt and P. Nath, \PRL49, 970 (1982);
1997:                 R. Barbieri, S. Ferrara and C.A. Savoy, \PLB119, 343 (1982);
1998:                 L. Hall, J. Lykken and S. Weinberg, \PRD27, 2359
1999:                 (1983); H. P. Nilles, Phys. Report {\bf{110}}, 1
2000:                 (1984); M.~Drees and S.~P.~Martin, hep-ph/9504324.
2001:                 %%CITATION = HEP-PH 9504324;%%
2002:                 %%CITATION = PHRVA,D27,2359;%%
2003:                 %%CITATION = PRLTA,49,970;%%
2004:                 %%CITATION = PHLTA,B119,343;%%
2005:                 %%CITATION = PRPLC,110,1;%%
2006: \bibitem{gmsb} N. Arkani-Hamed, J. March-Russell and H. Murayama, \NPB 509, 3(1998);
2007:                H. Murayama, \PRL79, 18 (1997); K.-I. Izawa, Y. Nomura, K. Tobe and
2008:                T. Yanagida, \PRD56, 2886(1997); M.A. Luty, \PLB414, 71(1997).
2009:                 %%CITATION = PHRVA,D56,2886;%%
2010:                 %%CITATION = PRLTA,79,18;%%
2011:                 %%CITATION = PHLTA,B414,71;%%
2012:                  %%CITATION = NUPHA,B509,3;%%
2013: \bibitem{hikasa} K. Hikasa and M. Kobayashi, \PRD36, 724 (1987);
2014:                  M. J. Duncan, \NPB221, 285 (1983).
2015:                  %%CITATION = PHRVA,D36,724;%%
2016:                  %%CITATION = NUPHA,B221,285;%%
2017: \bibitem{Hahn} T.~Hahn and M.~Perez-Victoria,  Comput.\ Phys.\ Commun.\  {\bf 118}, 153 (1999);
2018:                 T.~Hahn,  Nucl.\ Phys.\ Proc.\ Suppl.\  {\bf 135}, 333 (2004).
2019:                %%CITATION = HEP-PH 0406288;%%
2020:                %%CITATION = HEP-PH 9807565;%%
2021: \bibitem{Hooft} G.~Passarino and M.~J.~G.~Veltman, \NPB160, 151 (1979);
2022:                 G.~'t Hooft and M.~J.~G.~Veltman, \NPB153, 365 (1979).
2023:                  %%CITATION = NUPHA,B160,151;%%
2024:                  %%CITATION = NUPHA,B153,365;%%
2025: \bibitem{Barger} Vernon D. Barger and R.J.N. Phillips, Collider Physics,
2026:                  Addison-Wesley (1987).
2027: \bibitem{pdg} W.~M.~Yao {\it et al.}  [Particle Data Group],
2028:               \JPG33, 1 (2006).
2029:               %%CITATION = JPHGB,G33,1;%%
2030: \bibitem{mhmax} M.~Carena, S.~Heinmeyer, C.~E.~M.~Wagner, G.~Weiglein, hep-ph/9912223.
2031:                %%CITATION = HEP-PH 9912223;%%
2032: \bibitem{hollik} S. Heinemeyer, {\it et al.}, \EPJC37, 481 (2004).
2033:                 %%CITATION = HEP-PH 0403228;%%
2034: \bibitem{cteq}  J.~Pumplin, A.~Belyaev, J.~Huston, D.~Stump and W.~K.~Tung,
2035:                JHEP {\bf 0602}, 032 (2006).
2036:                %%CITATION = HEP-PH 0512167;%%
2037: \bibitem{Hosch} M.~Hosch, K.~Whisnant and B.~L.~Young, \PRD56, 5725 (1997).
2038:                 %%CITATION = HEP-PH 9703450;%%
2039: \bibitem{pptc-background} T.~Stelzer, Z.~Sullivan and S.~Willenbrock, \PRD58, 094021 (1998).
2040:                   %%CITATION = HEP-PH 9807340;%%
2041: \bibitem{pptc-Han} T.~Han, M.~Hosch, K.~Whisnant, B.~L.~Young and X.~Zhang, \PRD58, 073008
2042:                  (1998).
2043:                   %%CITATION = HEP-PH 9806486;%%
2044: \bibitem{spectrum} K.~Inoue, A.~Kakuto, H.~Komatsu and S.~Takeshita,
2045:                   Prog.\ Theor.\ Phys.\  {\bf 67}, 1889 (1982);
2046:                   Prog.\ Theor.\ Phys.\  {\bf 68}, 927 (1982) [Erratum-ibid.\  {\bf 70}, 330
2047:                   (1983)]; Prog.\ Theor.\ Phys.\  {\bf 71}, 413 (1984).
2048:                   %%CITATION = PTPKA,71,413;%%
2049:                   %%CITATION = PTPKA,67,1889;%%
2050:                   %%CITATION = PTPKA,68,927;%%
2051: \bibitem{mw-mssm} S.~Heinemeyer, W.~Hollik, D.~Stockinger, A.~M.~Weber and G.~Weiglein,
2052:                   JHEP {\bf 0608}, 052 (2006).
2053:                  %%CITATION = HEP-PH 0604147;%%
2054: \bibitem{charged higgs} R.~Barbieri and G.~F.~Giudice, \PLB309, 86 (1993).
2055:                   %%CITATION = HEP-PH 9303270;%%
2056: \bibitem{bsr-character} E.~Gabrielli, K.~Huitu and S.~Khalil, \NPB710, 139
2057:                  (2005); Z.~J.~Xiao, F.~G.~Li and W.~J.~Zou, hep-ph/0603120.
2058:                  %%CITATION = HEP-PH 0603120;%%
2059:                  %%CITATION = HEP-PH 0407291;%%
2060: \bibitem{heavy flavor} HFAG Group, http://www.slac.stanford.edu/xorg/hfag.
2061: \bibitem{bsr-sm} P.~Gambino and M.~Misiak, \NPB611, 338(2001);
2062:     A.J.Buras, A.~Czarnecki, M.~Misiak, and J.~Urban,  \NPB631, 219 (2002);
2063:     C.~Bobeth, P.~Gambino, M.~Gorbahn,  and U.~Haisch, JHEP{\bf 04}, 071 (2004) 071;
2064:     A.~Ghinculov, T.~Hurth, G.~Isidori, and Y.P.~Yao, \NPB685, 351(2004).
2065:                  %%CITATION = NUPHA,B611,338;%%
2066:                  %%CITATION = NUPHA,B631,219;%%
2067:                  %%CITATION = NUPHA,B685,351;%%
2068: \bibitem{bsll} P.~Gambino, U.~Haisch and M.~Misiak,\PRL94, 061803(2005).
2069:                 %%CITATION = PRLTA,94,061803;%%
2070: \bibitem{Abazov:2006dm} V.~M.~Abazov {\it et al.}  [D0 Collaboration],
2071:             \PRL97, 021802 (2006).
2072:               %%CITATION = HEP-EX 0603029;%%
2073: \bibitem{Bona:2005vz}
2074:   M.~Bona {\it et al.}  [UTfit Collaboration], JHEP {\bf 0507}, 028 (2005).
2075:   %%CITATION = HEP-PH 0501199;%%
2076: \bibitem{Charles:2004jd}
2077:   J.~Charles {\it et al.}  [CKMfitter Group], \EPJC41, 1 (2005).
2078:       %%CITATION = HEP-PH 0406184;%%
2079: \bibitem{Endo:2006dm}
2080:   M.~Endo and S.~Mishima, \PLB640, 205 (2006).
2081:   %%CITATION = HEP-PH 0603251;%%
2082: \bibitem{rho} See, e.g., P.~Chankowski, {\it et~al.}, \NPB417, 101 (1994);
2083:                 D.~Garcia and J.~Sol\`a, Mod. Phys. Lett. {\bf A 9}, 211 (1994).
2084:                 %%CITATION = NUPHA,B417,101;%%
2085:                 %%CITATION = MPLAE,A9,211;%%
2086: \bibitem{lep} The LEP Collaborations, hep-ex/0509008.
2087:                %%CITATION = HEP-EX 0509008;%%
2088: \bibitem{Dabelstein} A. Dabelstein, \ZPC67, 495 (1995).
2089:                       %%CITATION = HEP-PH 9409375;%%
2090: \bibitem{h-original}  H.E. Haber, R. Hempfling, \PRL66, 1815 (1991);
2091:                   Y. Okada, M. Yamaguchi, T. Yanagida, Prog. Theor. Phys. {\bf 85}, 1 (1991);
2092:                                                                             \PLB262,54(1991);
2093:                  J. Ellis, G. Ridolfi, F. Zwirner, \PLB257, 83 (1991); \PLB262, 477 (1991);
2094:                  J.~R.~Espinosa and R.~J.~Zhang, JHEP {\bf 0003} (2000) 026.
2095:                  %%CITATION = PRLTA,66,1815;%%
2096:                  %%CITATION = PTPKA,85,1;%%
2097:                  %%CITATION = PHLTA,B262,54;%%
2098:                 %%CITATION = PHLTA,B257,83;%%
2099:                 %%CITATION = PHLTA,B262,477;%%
2100: \bibitem{feynhiggs}S.~Heinemeyer, Int.\ J.\ Mod.\ Phys.\ A {\bf 21}, 2659
2101:                  (2006);  T.~Hahn, W.~Hollik, S.~Heinemeyer and G.~Weiglein, hep-ph/0507009.
2102:                  %%CITATION = HEP-PH 0407244;%%
2103:                  %%CITATION = HEP-PH 0507009;%%
2104: \bibitem{benchmark} M.~Carena, {\it et al.}, \EPJC26, 601 (2003).
2105:                     %%CITATION = HEP-PH 0202167;%%
2106: \bibitem{tch-Aguilar} J.~A.~Aguilar-Saavedra and G.~C.~Branco, \PLB495, 347 (2000).
2107:                   %%CITATION = HEP-PH 0004190;%%
2108: \bibitem{tcz-Han} T.~Han, R.~D.~Peccei and X.~Zhang, \NPB454, 527 (1995).
2109:                   %%CITATION = HEP-PH 9506461;%%
2110: \bibitem{tcz-Atlas} L.~Chikovani and T.~Djobava, hep-ex/0205016.
2111:                    %%CITATION = HEP-EX 0205016;%%
2112: \bibitem{tcr-Beneke} M.~Beneke {\it et al.}, hep-ph/0003033.
2113:                   %%CITATION = HEP-PH 0003033;%%
2114: \bibitem{zt-Aguilar} F.~del Aguila and J.~A.~Aguilar-Saavedra, \NPB576, 56 (2000).
2115:                   %%CITATION = HEP-PH 9909222;%%
2116: \bibitem{tcg-Han} T.~Han, K.~Whisnant, B.~L.~Young and X.~Zhang, \PLB385, 311 (1996).
2117:                  %%CITATION = HEP-PH 9606231;%%
2118: \bibitem{tcr-Han} T.~Han, K.~Whisnant, B.~L.~Young and X. Zhang, \PRD55, 7241 (1997).
2119:                   %%CITATION = HEP-PH 9603247;%%
2120: 
2121: \end{thebibliography}
2122: \endgroup
2123: 
2124: \end{document}
2125: )
2126: