hep-ph0004133/na.tex
1: 
2: \documentstyle[a4,12pt]{article}
3: 
4: \input{epsf}
5: 
6: \begin{document}
7: \titlepage 
8: \begin{flushright}
9: CERN-TH/2000-111\\\
10: SPhT 00/46\\
11: hep-ph/0004133
12: \end{flushright}
13: \vskip 7pt
14: \begin{center}
15: {\bf \Large \bf
16: Models with Inverse Sfermion Mass Hierarchy and Decoupling of
17: the SUSY FCNC Effects}
18: \end{center}
19: 
20: \vskip 1.2cm
21: \begin{center}
22: { \bf Ph. Brax\footnote{email: philippe.brax@cern.ch}\footnote{On leave from
23:  CEA-SACLAY, Service de Physique Th\'eorique, 
24: F-91191 Gif-sur-Yvette, France}}
25: \vskip 1pt
26: \end{center}
27: \centerline {Theory Division, CERN, CH-1211, Geneva, 23, Switzerland}
28:  
29: \vskip 1cm
30: \begin{center}
31: { \bf C. A. Savoy\footnote{email:savoy@spht.saclay.cea.fr}}
32: \end{center}
33: \vskip 1pt
34: \centerline{ CEA-SACLAY, Service de Physique Th\'eorique, 
35: F-91191 Gif-sur-Yvette, France}
36: \vskip 1 cm
37: \leftline{Key Words: Beyond Standard Model, Supergravity Models}
38: \vskip 2cm
39: 
40: \abstract
41: { We study the decoupling of the first two
42: squark and slepton families in order to lower the flavour changing
43: neutral current effects. Models with inverse sfermion mass hierarchy
44: based upon gauged $U(1)$ flavour symmetries provide a natural framework
45: where  decoupling can be implemented. Decoupling requires a large gap
46: between the Fermi scale and the supersymmetry breaking scale.
47: Maintaining the electroweak symmetry breaking at the Fermi scale
48: requires some fine-tuning that we investigate by solving the two-loop
49: renormalization group equations. We show that the two-loop effects are
50: governed by the anomaly compensated by the Green-Schwarz mechanism and
51:  can be determined from the quark and lepton masses. The
52: electroweak breaking constraints lead to a small $\mu$ scenario where
53: the LSP is Higgsino-like. }
54: 
55: \newpage
56: 
57: \section{Introduction}
58: Many sfermion mass patterns have been suggested in order to
59:  alleviate the effects of the flavour non-diagonal contributions
60: in the supersymmetric Flavour Changing Neutral Current (FCNC) problems.
61: Three main mechanisms have been proposed:\.  (i) {\it degeneracy} in
62: the scalar masses;\. (ii) {\it alignment} between fermion and sfermion
63: mass matrices;\. (iii) {\it decoupling} of all virtual supersymmetric
64: effects by large scalar masses.  Degeneracy  seems  natural  in the
65: context of gravity mediated sypersymmetry breaking although it
66: could also be pointed out that a generic flavour dependence of the
67: K\"{a}hler potential tends to spoil  this degeneracy in the presence of
68: a spontaneously broken flavour symmetry, if unprotected by this
69: symmetry.  In the framework of gauged flavour symmetries the induced
70: $D$-type soft masses are especially dangerous in that respect.
71: Actually, gauge mediation of supersymmetry breaking at a relatively low
72: scale would lead to negligible FCNC effects if the gravitino is light
73: enough but it apparently faces inherent problems to induce the correct
74: electroweak symmetry breaking.
75: 
76: The implementation of the alignment prescription via a rather ad hoc
77: choice of flavour symmetries has been advocated \cite{nls} to be  the
78: only natural solution, although some of the arguments used have to be
79: revised in the framework of broken supergravity theories. The examples
80: of such symmetries are sufficiently contrived that one feels compelled
81: to look for another possibility. On the other hand, decoupling has
82: always served as a possible remedy  for the inconsistencies of any
83: model beyond the standard theory.  However, third generation scalars
84: together with the Higgs and gaugino sectors which control the
85: electroweak symmetry breaking should be kept light enough to avoid too
86: much fine-tuning.  Since FCNC effects are more stringent within the
87: first two generations of quarks and leptons, it has been envisaged
88: that the first two generations could be considerably heavier than the
89: other supersymmetric particles\cite{dg} - actually, an efficient
90: suppression of FCNC effects by decoupling alone would require very
91: heavy masses.  As a matter of fact, the first and second families of
92: sfermion masses do enter the MSSM expression for the $Z$ boson mass
93: when one takes into account the  two loop effects . This has been used
94: to put limits on the mass gap between generations of scalars, mostly by
95: requiring the absence of an excessive fine-tuning.  Other bounds have
96: been obtained from the positivity of the stop masses\cite{mura}.  We
97: will comment on these generic bounds later on.
98: 
99: Our main interest is to cast the decoupling approach within
100: spontaneously broken supersymmetry scenarios.  In this context a very
101: large gap between families seems to be difficult to reach in the
102: supergravity mediated framework due to the form of the scalar masses,
103: assumed to be flavour diagonal, for simplicity,
104: \begin{equation} 
105: m_i^2= m_{3/2}^2 +\frac{\left< F_{a}\right> \left<
106: \bar F_{\bar b}\right> }{M_{P}^2}\partial_a\partial_{\bar b} \ln
107: (\partial _i \partial_{\bar i} K) +X_i \left< D_X\right>\ ,\label{smass}
108: \end{equation} 
109: where $K$ is the K\"{a}hler potential, $X_i$ is the gauge charge of the
110: state whilst $\left< F_a\right>$ and $\left< D_X\right>$ are the
111: auxiliary fields responsible for supersymmetry breaking, which
112: (vectorially) add up to $\sqrt{3} m_{3/2} M_P$.  This formula displays
113: the dependence of the sfermion masses on :({\sl i}) flavour independent
114: terms such as $m_{3/2}^2$;({\sl ii}) $\left< F_a\right>$ along
115: directions where there is a dependence of the K\"{a}hler metric in the
116: matter field families; ({\sl iii}) $\left< D_X\right>$ along flavour
117: symmetries.  One easily realizes from (\ref{smass}) that scalars in
118: different generations cannot be split by more than one order of
119: magnitude, unless ({\sl i}) is compensated by ({\sl ii}), like in
120: no-scale supergravity.
121: 
122: In the presence of supersymmetry breaking and flavour symmetry breaking
123: there are two sources of ``induced'' supersymmetry breaking capable to
124: yield flavour dependent scalar mass splittings. Let us consider a
125: simple situation\cite{FN} in order to illustrate this point: an Abelian
126: flavour symmetry $U(1)_X$ broken by the value of a Frogatt-Nielsen
127: field $\phi$ with $m_{3/2}\le \left<\phi\right> \le M_P$. Assuming a
128: vanishing cosmological constant, from the supergravity Lagrangians one
129: obtains\cite{dps,dd} the induced supersymmetry breaking $\left<
130: F_{\phi}\right> \sim m_{3/2}\left<\phi\right>$ along the $\phi$
131: direction and $\left< D_X\right>\sim m_{\phi}^2$ where $m_{3/2}^2$ is
132: the total supersymmetry breaking and $m_{\phi}^2$ is the $\phi$ soft
133: mass, of $O(m_{3/2}^2).$ In this case, the $D$-type splitting is always
134: relevant, the $F$-type one being proportional to $\left<\phi\right> ^2
135: / M_P^2 .$
136: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
137: \subsection{\sl iMSSM}
138: %
139: Despite the strong motivations for models with very heavy scalars the
140: only concrete ones discussed so far are those coined "inverse hierarchy
141: models" (iMSSM). They are based on the assumption that an anomalous
142: $U(1)_X$ gauge symmetry is present  in the flavour sector of the
143: theory. The anomaly is fixed by the Green-Schwarz mechanism which then
144: determines the scale of the flavour symmetry breaking
145: $\left<\phi\right>.$ The induced $D$-term produces a mass splitting,
146: $m_i^2-m_j^2 = (X_i- X_j)m_{\phi}^2 .$ In models where the anomalous
147: $U(1)_X$ is responsible for the fermion mass hierarchy the charge
148: difference are roughly related to the fermions masses\cite{dps}
149: \begin{equation}
150: m_i^2-m_j^2 \propto \ln \left(\frac{m^F_j}{m^F_i}\right)\ .\label{hier}
151: \end{equation}
152: This leads to an inverse hierarchy in the sfermion masses compared to
153: the fermionic hierarchy (quarks and leptons).
154: 
155: This fact was first pointed out in the framework of general broken
156: supergravity coupled to Abelian flavour gauge symmetry\cite{dps} and,
157: subsequently, in models with dynamical supersymmetry breaking\cite{dd}.
158: It has been noticed that, because the top Yukawa coupling is of $O(1)$,
159: there is the relation $(X_{t_L}+X_{t_R}+X_{H_2})=0$, for the charges of
160: the top-Higgs sector, implying that the soft mass combination,
161: $(m_{U_3}^2+m_{Q_3}^2+m_{H_2}^2)$ receives no contribution from the
162: $D-$terms. Since this is the combination appearing in the one-loop
163: correction to the boson mass $M^2_Z$, the radiative gauge symmetry
164: breaking is automatically protected, at one-loop, against large
165: $D-$terms that are due to {\sl any} gauge symmetry broken at scales
166: below $M_P$.
167: 
168: It goes without saying that the FCNC effects are particularly dangerous in
169: the iMSSM framework. It has been suggested that this problem can be
170: evaded by a suitable combination of degeneracy, alignment and, last but
171: not least, decoupling of the first two generations\cite{dps,nel}. The
172: latter has prompted us to evaluate the two-loop corrections. This is
173: done in Section 2.
174: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  
175: \subsection{\sl A limit on two-loop effects} 
176: %
177: In Section 3, we show that for inverse
178: hierarchy models based on $D$-term splitting the dominant contribution
179: to the two loop terms in the renormalization evolution of the scalar
180: masses is proportional to the anomalies of the $U(1)_X$  group. For the
181: most relevant case of  gauged $U(1)_X$ symmetries the anomalies must be
182: cancelled by the Green-Schwarz mechanism. We will concentrate on
183: theories such as the weakly coupled heterotic string with only one
184: anomalous $U(1)_X$ and one dilaton-axion to implement the Green-Schwarz
185: mechanism.
186: 
187: An interesting aspect of this result is that despite the large variety
188: of $X$ charges and choice of symmetry to explain the fermion hierarchy
189: the $U(1)_X$ anomaly can be fixed in a rather model independent way.
190: Indeed by using the previously obtained relations between the anomaly
191: $\cal{A}$ and the fermion masses one gets\cite{ir}
192: \begin{equation}
193: m_u m_c m_t (m_e m_{\mu} m_{\tau})^3 (m_d m_s m_b )^{-2}\approx
194: \left( \frac{g^{2}_{X} \cal{A} }{32\pi^2 }\right) ^{\mbox{${\cal A}/2$}}
195: \sin ^3\beta \cos^3\beta (174\mbox{GeV})^6 \ ,\label{mass1}
196: \end{equation}
197: where the quark and lepton masses are taken at the scale $g_X\sqrt{\cal
198: A}M_P/4\pi$, $\tan \beta$ is the ratio between the two Higgs vacuum
199: expectation values (vev's), $g_X$ is the $U(1)_X$ gauge coupling and
200: ${\cal A}$ is the $U(1)_X$ anomaly with respect to the standard model
201: gauge groups.  
202: 
203: The evaluation of (\ref{mass1}) yields  ${\cal A}\approx 25\pm 3 .$
204: Remarquably enough this leads to a prediction of the Cabibbo angle
205: $\theta_C\approx 0.2$ in these models. 
206: 
207: This establishes a quite model independent estimate of the two loop
208: effects in inverse hierarchy models. It turns out to be much smaller
209: than values considered in previous discussions inspired by these
210: models\cite{dg,mura}.
211: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
212: \subsection{\sl Maximal hierarchy limit}
213: %
214: As already stressed in the iMSSM context, the supersymmetric flavour
215: problem requires very heavy sfermions in the first two generations,
216: hence a large  supersymmetry breaking scale, $m^2_{3/2}>>M_Z^2$.  The
217: fine-tuning problem becomes crucial.  In the cMSSM, where the
218: coefficient of the universal soft scalar masses $m_0^2$ in the
219: expression for $M_Z^2$ is strongly suppressed -- a Nature fine-tuning
220: of the top mass $m_t$ -- for moderate and large values of $\tan \beta$,
221: the necessary fine-tuning is mostly between $M_{1/2}^2$ and $\mu^2$.
222: This requires relatively large values for $\mu,$ yielding gaugino-like
223: LSP.  Several studies in the literature\cite{poko} have already
224: discussed how the predictions are modified by departing from the cMSSM
225: scalar degeneracy. The main point is that, in this case, the scalar
226: masses can also participate in cancelling the $M_{1/2}^2$ term in the
227: expression for $M_Z^2$, allowing for relatively low values of $\mu.$
228: For these small $\mu$ solutions, the spectrum wiil consist of scalars
229: heavier than gauginos, which are also heavier than the higgsinos.  The
230: advantage of this class of models is that the fine-tuning now occurs
231: mostly in the ratio $M_{1/2}^2/ m^2_{3/2}$, which is more obviously
232: related to the supersymmetry breaking mechanism, while the origin of
233: the $\mu$ parameter, which could now be of $O(M_Z),$ remains more
234: mysterious.
235: 
236: In Section 4, we investigate the possibility of large values of
237: $m^2_{3/2}$ together with small $\mu$ values. The two-loop correction,
238: controlled by the anomaly, contributes in some cases.  In
239: this regime, the iMSSM does reveal an inversed mass spectrum as compared
240: with the cMSSM. We discuss approximate constraints in the neighbourhood
241: of the 'infinitely' fine-tuned solution for large $\tan\beta,$ but this
242: rough approximation turns out to be quite appropriate from our numerical
243: analysis.
244: 
245: The iMSSM mass patterns are displayed in Section 5.  We summarize our
246: conclusions in the last section.  The case of more than one $U(1)$
247: flavour symmetries is sketched in the Appendix.
248: 
249: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
250: \section{Two Loop Renormalization Effects}
251: 
252: Let us first treat the two loop renormalization group equations of the
253: scalar masses in the following approximation,
254: \begin{equation}
255: \frac{dm_i^2}{dt}= -8\pi^2C_A(i)\left[ M_A^2 -
256: \frac{g_A^2}{16\pi^2}\mbox{tr}\left(\frac{C_Am^2}{d_A}\right)\right] +
257: 2g_1^2Y_i \left[ S + \frac{g_A^2}{4\pi^2}\mbox{tr}(YC_Am^2)\right] 
258: \label{rge}
259: \end{equation}
260: where the indices $A=1,2,3$ correspond to the gauge group factors
261: $U(1), SU(2), SU(3)$ respectively, $g_A$ is the corresponding gauge
262: coupling, $d_A$ the algebra dimensions, $M_A$ is the gaugino mass and
263: $C_A(i)$ is the Casimir eigenvalue for the fermion/sfermion labelled by
264: $i$. Finally, 
265: \begin{equation}
266: S= \mbox{tr} (Ym^2)
267: \end{equation}
268: introduces a $Y-$dependent term in the scalar masses.
269: 
270: The approximation (\ref{rge}) is appropriate for the class of models
271: discussed herein and corresponds to neglecting the two-loop corrections
272: proportional to $M_A^2$ which are suppressed by a factor of
273: $g_A^2C_A(i)/16\pi^2$ with respect to the one loop terms. As displayed,
274: (\ref{rge}) is not valid for the stops. Including the top Yukawa
275: couplings to extend the calculation to the third family scalar masses
276: and consistently neglecting the terms in $M_A^2$ which are higher
277: order in $g_A^2$, yields the solution
278: \begin{eqnarray}
279: m_i^2&=&m_i^2(0)-\frac{a(i)}{12}\rho \left( 3\bar{m}^2 +8M_{1/2}^2+
280: (1-\rho)(A_{Q_3}(0) +2M_{1/2})^2-\delta \right)\nonumber \\
281: &+& \vert t\vert C^2_A(i)g_A^2\left (8\left (g_A^2+
282: \frac{1}{2}\right)M_{1/2}^2 -\delta\right) + 
283: \frac{Y_i}{22}\left[ S(t)-S(0)\right] \label{massren}
284: \end{eqnarray}
285: at the Fermi scale where the coefficients $a(i)$ are
286: $a(U_3)=4,\ a(Q_3)=2,\ a(H_2)=6$ and zero otherwise, $3\bar{m} ^2 =
287: (m_{U3}^2+m_{Q3}^2+m_{H2}^2),$ $A_{Q_3}$ is the soft coupling associated to
288: the top Yukawa coupling, and finally,
289: \begin{equation}
290: \rho=\frac{h_t}{(h_t)}_{F.P.} \approx \frac{0.72}{\sin^2\beta} \ .
291: \end{equation}  
292: In  approximating the solutions for $m_i^2$ we are anticipating and
293: taking advantage of the fact that the two-loop term
294: \begin{equation}
295: \delta=\frac{1}{4\pi^2}\mbox{tr}\left(\frac{C_Am^2}{d_A}\right)\vert_{t=0}
296: \label{delta}
297: \end{equation}
298: is almost independent of the index $A$ in the classes of models
299: discussed here, that we now turn to discuss.
300: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
301: \subsection{\sl In the cMSSM}
302: %
303: The most important effect of the radiative corrections on the
304: $SU(2)\times U(1)$ breaking appears in the Higgs parameter 
305: \begin{eqnarray}
306: m_{H_2}^2&\approx &m_{H_2}^2(0) + 0.52(M_{1/2}^2 -0.15\delta) 
307: -0.014 S_0 \nonumber \\
308: & -&\frac{0.36}{\sin^2\beta}\left(\bar{m}^2+8M_{1/2}^2
309: +\left( 1-\frac{0.72}{\sin^2\beta}\right)(A_0+2M_{1/2})^2
310: -\delta\right) \ .
311: \end{eqnarray}
312: Therefore the two loop effects due to possible heavy scalars in the
313: first two generations become relevant for $\delta \sim O({\rm
314: few}M_{1/2}^2).$ For instance assuming a degeneracy amongst the heavy
315: scalars of mass $m$ in the first two generations this corresponds to $m
316: \sim 5 M_{1/2}.$
317: 
318: It is well known that in the cMSSM with boundary conditions
319: $m_i^2(t=0)=m_0^2$ the coefficient of $m_0^2$ in $M_Z^2$ is small for
320: large values of $\tan \beta$. Indeed the dependance of $m_{H_2}^2$ on
321: $m_0^2$ in the $\tan\beta >> 1$ limit is
322: \begin{equation}
323: m_{H_2}^2\approx -0.1 m_0^2 +0.3 \delta -2.76 M^2_{1/2} \ .
324: \end{equation}
325: Taking the traces over all the MSSM scalars one gets
326: $\mbox{tr}(C_3/8)=6,\ \mbox{tr}(C_2/3)=7,\ \mbox{tr}C_1=6.6,$ 
327: so that
328: \begin{equation}
329: \delta\approx \frac{m_0^2}{2\pi} \label{deluni} \ .
330: \end{equation}
331: The possibility of obtaining  a relatively large value of $m_0^2$
332: without worsening the fine-tuning in $M_Z^2$ remains at the two-loop
333: level.
334: 
335: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
336: \section{Anomalous U(1) Models} 
337: 
338: As already emphasized in the introduction the natural realization of
339: the inverse hierarchy for squarks and sleptons occurs in models where
340: the fermion mass hierarchy is explained by a Frogatt-Nielsen mechanism
341: with an anomalous $U(1)_X$ local flavour symmetry. Recently it has been
342: advocated that in type IIB orientifold models of string theory one
343: could expect different anomalous  $U(1)_X$ with their anomalies being
344: cancelled by a corresponding number of moduli fields \cite{aqi}.  We
345: shall ignore this possibility and remain within the more traditional
346: heterotic-like picture with only one anomalous $U(1)_X$ \cite{dws}. Of
347: course one could postulate the existence of other anomaly-free $U(1)$
348: flavour gauge symmetries in order to explain the fermion mass
349: hierarchy. As discussed in the Appendix the results are essentially
350: similar in the multi-$U(1)$ models.
351: 
352: We refer to the comprehensive literature on this subject for the
353: details and quote the main results only. The anomaly cancellation is
354: provided by the Green-Schwarz mechanism. A necessary condition for the
355: compensation of the anomalies with respect to the standard model gauge
356: symmetries as well as the gravitational anomaly is the equality
357: \begin{equation}
358: {\cal A}_1= {\cal A}_2={\cal A}_3=\frac{\mbox{tr}(X)}{24}\ ,\label{Acond}
359: \end{equation}
360: where ${\cal A}_A= 2\mbox{tr}(XC_A)/d_A .$ The Kac-Moody levels $k_A$
361: have all been taken to be one to simplify the discussion. The
362: coefficiient of the Fayet-Iliopoulos term required by the $U(1)_X$
363: gauge invariance -- which is nothing but the contribution of the
364: dilaton to $D_X$ -- is given by
365: \begin{equation}
366: \xi^2= \frac{k_X g_X^2 {\cal A}}{32\pi ^2} \ .\label{xi} 
367: \end{equation}
368: Introducing a Froggat-Nielsen field $\phi$ which is standard
369: model gauge singlet with a charge $X=-1$ (by a suitable
370: normalization of the $U(1)_X$ charges), the $U(1)_X$ gauge
371: symmetry is broken at the minimum of the scalar potential where
372: \begin{eqnarray}
373: \vert \phi \vert ^2&=&\xi^2 M_P^2 \ , \nonumber \\
374: \left<F_{\phi}\right> &\sim &m_{3/2}\xi M_P \ ,\nonumber \\
375: \left<D_X\right> &=&m_{\phi}^2 \ ,\label{breaking}
376: \end{eqnarray} 
377: where $ m_{\phi}^2$ is the soft mass given by the supersymmetry
378: breaking mechanism.  The $\left< F_{\phi}\right>$ and $\left<
379: D_{X}\right>$ vev's are  the induced supersymmetry breaking terms. The
380: former contributes to the scalar masses only as $\xi^2m_{3/2}^2$ and is
381: neglected here\footnote{Though, as discussed in \cite{dps}, this
382: contribution is relevant in the discussion of FCNC effects.}. The latter
383: gives an important D-term contribution to the scalar masses so that at
384: the scale $\xi M_P$,
385: \begin{equation}
386: m_i^2=\hat m_i^2 +X_i \left< D_X\right> \ ,\label{XD}
387: \end{equation}
388: where $\hat m_i^2$ is the contribution from the supersymmetry breaking
389: independent of the $D_X$ breaking and the charges $X_i$ are model
390: dependent.
391: %%%%%%%%%%%%%%%%%%%%%
392: \subsection{\sl Two-loop Correction and Anomaly}
393: %%%%%%%%%%%%%%%
394: In the inverse hierarchy models the two loop contributions of the
395: heavy scalars to the renormalization group equations turns out to
396: be quite independent of the choice of charges. Indeed, one gets
397: from the definition (\ref{delta}) and the masses (\ref{XD}),
398: \begin{equation}
399: 4\pi^2\delta =\sum_i \frac{C_A(i)}{d_A}(\hat m_i^2 +X_i<D_X>) \ ,
400: \end{equation}
401: which yields
402: \begin{equation}
403: \delta=\hat \delta +\frac{1}{8\pi^2}{\cal A}\left< D_X\right>\ . 
404: \label{delA}
405: \end{equation}
406: In obtaining this result we have used the equality of the anomalies
407: (\ref{Acond}). Hence, the main two-loop contribution, coming from the
408: $D-$terms in the scalar masses, is proportional to the anomaly
409: ${\cal A}$. Of course, the two-loop corrections coming from $D-$terms
410: corresponding to non-anomalous $U(1)'$s cancel.   
411: 
412: Interestingly enough, the anomaly ${\cal A}$ can be calculated from its
413: relation to the fermion masses. Indeed, even if the $X-$charges that
414: control the fermion masses are model dependent, it is possible to
415: display a combination of masses that only depends on the charges
416: through ${\cal A,}$ as we now turn to discuss.
417: %%%%%%%%%%%%%%%%%%%%%%
418: \subsection{\sl Calculation of the Anomaly}
419: %%%%%%%%%%%%%%%%%
420: In the anomalous $U(1)_X$ approach to the fermion hierarchy, with the
421: abelian flavour symmetry breaking given by the small parameter $\xi$
422: as discussed above, the quark and lepton Yukawa couplings to the
423: Higgses are given by
424: \begin{equation}
425: Y_{f_i} \sim  \xi ^{f_{Li} + f_{Ri} + h} \ ,\label{Yuky}  
426: \end{equation}
427: where the fermion name ({\sl e.g.}, $q_{i}$) also denotes its
428: $X-$charge (resp., $X(u_{Li})=X(d_{Li})),$ and $h$ is the $X-$charge of
429: the appropriate Higgs field. In particular, one obtains, from the
430: values of the third generation Yukawa couplings, the relations:
431: \begin{eqnarray}
432: q_3 + u_3 + h_2 &\approx & 0 \ ,\nonumber \\ 
433: q_3 + d_3 + h_1 &\approx & l_3 + e_3 + h_1 \approx \frac{4 - 
434: \ln{\tan \beta}}{\ln{\xi}} \ ,\label{charges} 
435: \end{eqnarray}
436: and $\xi$ is roughly the Cabibbo angle, $\theta_{C}\approx \xi.$
437: The charges of the other family fermions are more or less fixed by
438: their Yukawa couplings and the Kobayashi-Maskawa mixings.
439: 
440: Because of the condition (\ref{Acond}) on the anomalies, ${\cal A}$
441: is related to the fermion masses as follows,
442: \begin{eqnarray}
443: m_um_cm_t\left( m_em_{\mu}m_{\tau}\right) ^3(m_dm_sm_b)^{-2}\approx
444: \xi ^{\mbox{${\cal A}$}}\sin ^3\beta \cos^3\beta (174\mbox{GeV})^6 \ ,
445: \label{calA}
446: \end{eqnarray}
447: where $\xi^2=\frac{g_X^2{\cal A}}{32\pi^2}$, while for the Higgs
448: charges one obtains,
449: \begin{equation}
450: (h_1 + h_2) \ln{\xi} \approx 
451: \ln \left( \frac{m_d m_s m_b}{m_e m_{\mu}m_{\tau}} \right) \ .
452: \label{higgs}
453: \end{equation}
454: 
455: The quark and lepton masses are defined at the scale $\xi M_P .$ 
456: Putting the experimental masses in (\ref{calA}) yields 
457: \begin{eqnarray}
458: {\cal A}\ln(\frac{32\pi^2}{g_X^2{\cal A}})
459: \approx (90 \pm 3\ln{\sin{2\beta}} ) \ ,\label{Aeq}
460: \end{eqnarray}
461: giving ${\cal A}\approx 25\pm 3$ and $\xi \approx 0.2,$ with the GUT
462: value $g_X^2 = .5.$ This is in reasonable agreement with the relation
463: $\theta_C\approx \xi$. From (\ref{higgs}) one gets $(h_1 + h_2) \approx
464: 0 ,$ a result that we shall use later. In the Appendix we discuss the
465: model dependence of these relations.
466: 
467: Therefore, as a typical result, the relevant two-loop contribution 
468: to the low-energy scalar masses (\ref{massren}) is given by  a
469: relatively low value, 
470: \begin{equation}
471: \delta \approx \frac {m_{\phi}^2}{\pi}+ \hat{\delta} \label{delgen}
472: \end{equation}
473: where we have used (\ref{breaking}). 
474: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
475: \subsection{\sl i+cMSSM }
476: %
477: Let us first evaluate the impact of this two loop correction in a
478: simple model (i+cMSSM) with universality assumed for the primordial
479: supersymmetry breaking, namely, a contribution $m_0^2$ to all scalar
480: soft masses, and with an anomalous $U(1)_X$ flavour symmetry as
481: discussed above. In this case, $\left< D_X\right> = m_0^2,$ and the
482: scalar masses at the flavour symmetry breaking scale are
483: \begin{equation}
484: m^2_i=m_0^2 \left( 1 + X_i \right).
485: \end{equation}
486: From (\ref{charges}), the parameter $\bar{m}^2$ of the one-loop
487: correction in (\ref{massren}) is equal to $m_0^2,$ and, from
488: (\ref{delgen}) and (\ref{deluni}), the two-loop contribution depends on
489: \begin{equation}
490: \delta\approx {m_0^2}\left( \frac{1}{2\pi} + \frac{1}{\pi} \right)
491: \approx 0.5 m_0^2 \label{cdel}
492: \end{equation}
493: The corresponding contribution to $m_{H_2}^2$ is
494: \begin{equation}
495: \Delta m_{H_2}^2 \approx \frac{0.2 m_0^2}{\sin^2\beta}
496: \end{equation}
497: which is about three times larger than the corresponding parameter in
498: the cMSSM. As we shall discuss in the next section, contrarily to what
499: happens in the cMSSM, this alone could be enough to allow for small
500: $\mu$ models, in the large $\tan{\beta}$ limit, even if $X(H_2 ) = 0.$
501: 
502: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
503: \subsection{\sl iMSSM}
504: %
505: We now turn to discuss a more elaborate model\cite{dps} exhibiting the
506: inverse hierarchy for the scalar masses, where some flavour dependence
507: is incorporated into the K\"{a}hler potential besides the anomalous
508: $U(1)_X$ gauged flavour symmetry and the Green-Schwarz mechanism.  The
509: key assumption, which allows  the model to be  predictive, is that the
510: {\sl fermion mass hierarchies} are fixed solely by the abelian flavour
511: symmetries, not by the moduli dependence.  Amazingly, this assumption
512: turns out to imply that the (mass)$^2$ differences between sfermions
513: with the same SM quantum numbers are proportional to the gravitino
514: (mass)$^2,$ $m_{3/2}^2.$ Even without specifying the primordial
515: supersymmetry breaking along the dilaton and moduli directions in the
516: iMSSM, the sfermion masses can be parameterized in a simple and
517: suggestive way, which we now turn to summarize.
518: 
519: Let us denote by $\Phi_i$ the matter superfields and their scalars
520: where $\Phi=Q,U,D,L,E$ refers to the standard model fields and
521: $i=1,2,3$ refers to the family index. We denote by $\phi_i$ the $X$
522: charges. At the scale $\xi M_P$ the soft terms satisfy the relations
523: \begin{eqnarray}
524: m_{\Phi_i}^2-m_{\Phi_j}^2&=&(\phi_i-\phi_j)m_{3/2}^2, \nonumber \\  
525: m_{U_3}^2+m_{Q_3}^2+m_{H_2}^2&=&M_{1/2}^2,\nonumber \\
526: m_{D_3}^2+m_{Q_3}^2+m_{H_1}^2&=&M_{1/2}^2+(d_3+q_3+h_1)m_{3/2}^2,\nonumber\\
527: m_{E_3}^2+m_{L_3}^2+m_{H_1}^2&=&M_{1/2}^2+(e_3+l_3+h_1)m_{3/2}^2,\nonumber\\
528: m_{H_2}^2 + m_{H_1}^2&=&(2+h_2+h_1)m_{3/2}^2,\nonumber \\ 
529: A_{U_i}&=&(u_i+q_i+h_2 )m_{3/2}-M_{1/2}\nonumber \\
530: A_{D_i}&=&(d_i+q_i+h_1)m_{3/2}-M_{1/2}\nonumber \\
531: A_{L_i}&=&(e_i+l_i+h_1)m_{3/2}-M_{1/2}\nonumber \\ 
532: B&=&\left( 2+(h_2+h_1)\theta(h_2+h_1)\right) m_{3/2} \label{dpsmodel}
533: \end{eqnarray}
534: Actually, the terms proportional to the $U(1)_X$ charges come out as a
535: particular combination of the $\left< D_X\right>$ induced breaking and
536: the supersymmetry breaking in the moduli sector, which give rise to
537: this general form for the mass splitting between families. We have only
538: considered the soft terms which are diagonal in the family indices,
539: although the pattern of the off-diagonal terms give constraints on the
540: $U(1)_X$ charges from the FCNC bound. Other relations for the soft
541: terms will be spelt out later.
542: 
543: The fermion hierarchy requires a relatively strong family ordering of
544: the charges $\phi_1>\phi_2 >\phi_3$.  We concentrate on this situation
545: and even more on the case where $M_{1/2}<m_{3/2}$. However we do not
546: have to impose any particular choice for the charges $\phi_i$, in many
547: of the physical issues discussed below since the two loop corrections
548: that are relevant to the inverse hierarchy scenario are controlled by
549: the anomaly. This is fixed by the fermion masses to
550: \begin{equation}
551: \delta =\frac{{\cal A}m_{3/2}^2}{8\pi^2}\approx \frac{m_{3/2}^2}{\pi}\ .
552: \label{deli}
553: \end{equation}
554: With the relations in (\ref{dpsmodel}) one has for instance
555: \begin{equation}
556: m_{H2}^2=m_{H2}^2(t=0) -\frac{0.36}{\sin^2\beta}\left(\left( 10-
557: \frac{0.72}{\sin^2\beta}\right) M_{1/2}^2 -\delta\right) 
558: +0.52( M_{1/2}-0.15\delta)
559: \end{equation}
560: Therefore, the two-loop corrections are basically negligible in
561: the iMSSM. Nevertheless, the $\delta$ term is of some importance in
562: the discussion of the next section.  
563: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%5
564: \section{Higgsino-like LSP and Electro-Weak Symmetry Breaking}
565: 
566: It is well-known that by departing from the universality assumptions of
567: the cMSSM many of its striking predictions are dramatically affected.
568: In particular the nature of the LSP can change. This is mainly a matter
569: of competition between  the $\mu^2$ and $M_{1/2}^2$ parameters that
570: appear with different signs in the supersymmetric expression for
571: $M^2_Z$. The cMSSM coefficient of the universal soft scalar masses
572: $m_0^2$ is strongly suppressed  for rather large values of $\tan
573: \beta$. The necessary fine-tuning between $M_{1/2}^2$ and $\mu^2$ then
574: favours a lighter gaugino than the Higgsino.  In this section we show
575: that in the iMSSM discussed in the previous sections, the Higgsino
576: turns out to be a natural option for the LSP. Let us  sketch 
577: the situation within an analytic approximation to the supersymmetric
578: $SU(2)\times U(1)$ breaking. In terms of $t=\tan \beta, $ the minimum
579: equations read
580: \begin{eqnarray}
581: m_{H_1}^2-m_{H_2}^2-B\mu t(1-\frac{1}{t^2})=M_Z^2
582: \frac{1-t^2}{1+t^2}\nonumber \\
583: m_{H_1}^2+m_{H2}^2 + 2\mu^2-B\mu t(1+\frac{1}{t^2}) =0
584: \label{minimum} 
585: \end{eqnarray}
586: at the classical level. The radiative corrections are important
587: but the main contributions can be included by redefining
588: \begin{eqnarray}
589: \hat{m_1}^2 &=& m_{H_1}^2 -\frac{M_Z^2}{2}\frac{1-t^2}{1+t^2}
590: +3\frac{h_t^2}{16\pi^2}\mu^2 \ln \left(\frac{m_{\tilde t_1}
591: {m_{\tilde t_2}}}{m_t^2}\right) \nonumber \\
592: \hat{m_2}^2 &=& m_{H_2}^2 +\frac{M_Z^2}{2}\frac{1-t^2}{1+t^2}
593: +3\frac{h_t^2}{16\pi^2}(m_{\tilde t_1}^2+m_{\tilde
594: t_2}^2)\ln \left(\frac{m_{\tilde t_1}{m_{\tilde
595: t_2}}}{m_t^2}\right) \ . \label{newvariable}
596: \end{eqnarray}
597: From the minimum equations  we deduce that
598: \begin{equation}
599: \mu=\frac{Bt}{2}(1\pm \sqrt{1-\frac{4\hat{m}_{1}^2}{B^2 t^2}}) \ ,
600: \end{equation}
601: so that the small $\mu$ solution leading to a Higgsino-like LSP
602: corresponds to the minus sign in the previous equation. For large
603: enough values of $t$ (to be discussed later) such that $0<
604: 4\hat{m}_{1}^2 << B^2t^2$ one gets the following relations
605: \begin{eqnarray}
606: \mu&\approx &\frac{\hat{m}_{1}^2}{Bt}(1+\frac{\hat{m}_{1}^2}{B^2t^2})
607: \nonumber \\
608: \hat{m}_{2}^2&\approx &\frac{\hat{m}_{1}^2}{t^2}
609: (1-\frac{\hat{m}_{1}^2}{B^2}) \label{master}
610: \end{eqnarray}
611: Namely the small $\mu$ regime corresponds to $\mu^2 \propto t^{-2}$ and
612: $m_{H_2}^2\propto t^{-2}$. This suggests to expand the solutions in
613: powers of $t^{-2}$. This is quite unphysical but mathematically sound.
614: Let us start with an approximation to the low energy masses
615: parameterized as follows
616: \begin{eqnarray}
617: m_{H_1}^2&=&m_0^2(1-\sigma_H)+\frac{1}{2}(M_{1/2}^2 -0.15\delta)
618: \nonumber \\
619: m_{H_2}^2&=&m_{H1}^2 +2\sigma_H m_0^2 
620: - 0.36\left( 1+\frac{1}{t^2}\right) \left(\bar m^2
621: +8M_{1/2}^2+\Delta-\delta\right) \label{m_H}
622: \end{eqnarray}
623: where
624: \begin{eqnarray}
625: \Delta = \left( 0.28-\frac{0.72}{t^2} \right)\left(A_{U_3}
626: +2M_{1/2}\right)^2
627: \end{eqnarray}
628: is model dependent. In (\ref{m_H}), $m_0$ is the universal scalar
629: mass in the cMSSM and $m_0 = m_{3/2}$ in the iMSSM, as discussed before,
630: and we have assumed $m_{H_1}^2+m_{H_2}^2=2m_0^2$
631: as consistent with $h_1+h_2=0$.
632: 
633: As a first approximation we determine $M_{1/2}^2 / m_0^2$ by taking the
634: large fine-tuning limit, $M_Z^2<<m_0^2.$  We also neglect the radiative
635: corrections and we keep only the relevant powers of $\tan \beta.$ Then,
636: one can solve (\ref{master}) for $M_{1/2}$ in each of the models
637: discussed in the previous sections.  For the sake of illustration, we
638: take the values predicted by the iMSSM, $A_{U_3} = -M_{1/2}$ and $B =
639: 2m_0,$ but the latter only enters into the term $\propto t^{-2} .$
640: \vskip .2 cm
641: %
642: \leftline{{\bf a) cMSSM}}
643: %
644: \vskip .2 cm
645: \noindent
646: In this model, $\sigma_H=0,\ \bar m^2=m_0^2,\ \delta\approx
647: m_0^2/(2\pi) .$ One gets
648: \begin{equation}
649: \frac{M_{1/2}^2}{m_0^2}\approx
650: -0.02 - \frac{0.7}{t^2}
651: \end{equation}
652: which excludes the small $\mu$ solution in the limit $m_0^2
653: >>M_Z^2$ as well-known.
654: \vskip .2 cm
655: %
656: \leftline{{\bf b) i+cMSSM}}
657: %
658: \vskip .2 cm
659: \noindent In this case, the universality is only broken by the $U(1)_X$
660: $D-$terms, so that $\bar m^2=m_0^2 $ from (\ref{charges}),
661: $\delta\approx 3m_0^2/(2\pi)$ from (\ref{cdel}), $\sigma _H = h_2,$
662: yielding, 
663: \begin{equation}
664: \frac{M_{1/2}}{m_0^2}\approx  0.04 + 0.4 \sigma_H 
665: -\frac{0.7+0.3\sigma_H}{t^2}
666: \end{equation}
667: The existence of this solution, especially for $h_2=0$, is due to the
668: larger two-loop contributions. The gauginos are much lighter than the
669: sfermions for these small $\mu$ solutions. In this example, a
670: cancellation must occur between the one-loop term in $M_{1/2}^2$ and
671: the two-loop term in $m_0^2$ for large values of the soft masses.
672: \vskip .2 cm
673: %
674: \leftline{{\bf c) iMSSM}}
675: %
676: \vskip .2 cm
677: It is characterized by $\bar m^2=M_{1/2}^2,$ from (\ref{dpsmodel}),
678: and $\delta \approx m_{3/2}^2/\pi ,$ from (\ref{deli}). This leads to
679: \begin{equation}
680: \frac{M_{1/2}^2}{m_{3/2}^2}\approx 0.36(1+\sigma_H) + .05
681: -\frac{0.7+0.3\sigma_H}{t^2} \label{rapp}
682: \end{equation}
683: Roughly, the parameter $\sigma _H $ can take values in the range [-1,0].
684: {\sl E.g.,} if we take the value $\sigma _H = -0.25$  and $t=2$, we get
685: $0.17$ for the ratio (\ref{rapp}), which is close to the values
686: obtained in a scanning of the parameter space. Notice that the two-loop
687: (anomaly) term contributes by about one-third to this result. The ratio
688: in (\ref{rapp}) means a real fine-tuning, and in our numerical analysis
689: (after reintroduction of the radiative corrections and $M_Z$ in the
690: expressions) the deviations from this `infinite fine-tuning' limit are
691: rather small. In order to allow for a big hierarchy in the sfermion
692: masses, one has to take rather small values of $M_{1/2}/m_{3/2} ,$ by
693: increasing $|\sigma_H | .$ This ratio is related to the Goldstino angle
694: $\sin\theta_G= M_{1/2}/\sqrt{3}m_{3/2}$. In a sense this is a better
695: variable to be tuned than $\mu/m_{3/2}$ since it is simply related to
696: the nature of the supersymmetry breaking. Still it has to be fine-tuned
697: to match a quantity which, in the iMSSM, depends on the parameter
698: $\sigma_H$, related to the properties of the Higgs fields under the
699: $U(1)_X$ and modular symmetries.
700: 
701: Notice that the condition for a Higgsino-like LSP, $\mu^2<M_{1/2}^2/6
702: ,$ is fulfilled with $\tan \beta > 3,$ for $\sigma _H > -.75 .$
703: Otherwise, the LSP can be gaugino-like, while the lightest chargino
704: remains Higgsino-like.  Therefore, the small $\mu$ solution of the
705: iMSSM is generically characterized by {\sl (i)} large values of
706: $m_{3/2},$ {\sl i.e.,} very heavy sfermions of the two first
707: generations; {\sl (ii)} smaller values of $M_{1/2},$ {\sl i.e.,}
708: moderate gaugino masses; and {\sl (iii)} $\mu$ as low as $O(M_Z),$ {\sl
709: i.e.,} a Higgsino as the LSP.
710: 
711: Radiative corrections have been neglected in this discussion, but the
712: main effect of their inclusion is to increase the value of $\tan \beta$
713: for a given set of parameters.
714: 
715: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
716: \section{The iMass Spectrum}
717: 
718: In this section we shall discuss the typical mass spectrum that one can
719: derive from inverse hierarchy models. We will also comment briefly on
720: the predictions for the FCNC effects, a main issue in these models
721: because of the large mass splitting between the families.
722: 
723: The mass spectrum of the iMSSM version\cite{dps}  depends on three
724: parameters $\sigma_H$, $\sigma_L$ and $\sigma_Q$ which measure the
725: departure from scalar mass universality between the two Higgs
726: doublets,  the leptons and the quarks of the third family,
727: respectively.  These parameters define the solutions of
728: (\ref{dpsmodel}) and so include the dependence on the corresponding
729: $X-$charges and, in this model, on their flavour dependent K\"{a}hler
730: geometry. They are family independent. Then, the family dependent mass
731: terms, accordingly to (\ref{dpsmodel}), depend only on the $X-$charge
732: differences, {\it e.g.,} $q_{1}-q_{3}.$  Such charges can be chosen to
733: get a good agreement with fermionic mass patterns and mixing angles.
734: The choice of charges plays also a role in the $S$ term in the masses
735: (at the Fermi scale). The correction to the masses due to this term is
736: \begin{equation}
737: \delta m_i^2 =\frac{Y_i}{22}(S-S_0)  \ ,\label{Sterm}
738: \end{equation}
739: where $S_0$ comprises a term like $\hbox{tr} (XY) m_{3/2}^2 .$ From the
740: renormalization group equations one finds that the evolution of $S$ is
741: given in first approximation by
742: \begin{equation}
743: S-S_0=(\frac{g_1^2}{g_0^2}-1)S_0 \approx -0.6 S_0 \ .
744: \end{equation}
745: The effect of this contribution has been often overestimated in the 
746: literature. In any instance, the term (\ref{Sterm}) can be consistently 
747: included in the definition of $\sigma_H$, $\sigma_L$ and $\sigma_Q ,$
748: without loss of  generality. This is understood in what follows.
749: 
750: The masses of the third family sleptons are then given by $(\kappa=
751: M_Z^2\cos 2\beta )$
752: \begin{eqnarray}
753: m^2_{\tilde\tau_L}&=&(1+\sigma_l)m_{3/2}^2 +0.5(M_{1/2}^2-0.15\delta)+
754: 0.4\kappa \nonumber \\
755: m^2_{\tilde\nu_L}&=&(1+\sigma_l)m_{3/2}^2 +0.5(M_{1/2}^2-0.15\delta)+
756: -0.5\kappa \nonumber \\
757: m^2_{\tilde\tau_R}&=& 1.16 M_{1/2}^2 -(\sigma_l-\sigma_H)m_{3/2}^2 
758: -0.03\delta +0.23\kappa \label{sleptons}
759: \end{eqnarray}
760: and the third family squark masses are given by
761: \begin{eqnarray}
762: m^2_{\tilde b_L}&=& (1+\sigma_q)m_{3/2}^2 +6.9M_{1/2}^2 -0.5\delta
763: -\frac{\rho}{6}((10-\rho)M_{1/2}^2-\delta)+0.42 \kappa\nonumber\\
764: m^2_{\tilde b_R}&=&-(\sigma_q-\sigma_H)m_{3/2}^2 +7.4M_{1/2}^2
765: -0.43\delta +0.75 \kappa\nonumber\\
766: m^2_{\tilde t_L}&=& (1+\sigma_q)m_{3/2}^2 +6.9M_{1/2}^2 -0.5\delta
767: -\frac{\rho}{6}((10-\rho)M_{1/2}^2-\delta)-0.35 \kappa\label{squarks}\\
768: m^2_{\tilde t_R}&=&-(2+\sigma_q+\sigma_H)m_{3/2}^2 +7.4M_{1/2}^2
769: -0.44\delta -\frac{\rho}{3}((10-\rho)M_{1/2}^2-\delta)-0.15 \kappa
770: \nonumber
771: \end{eqnarray}
772: 
773: Since we are  interested in the case $ M_{1/2} << m_{3/2}, $ the
774: allowed range for the parameters $\sigma_H$, $\sigma_L$ and $\sigma_Q
775: ,$ is strongly constrained, and $\sigma_H$ has also  to be
776: consistent with the electroweak break conditions (\ref{master}).
777: 
778: Let us now present some generic features of the spectrum. The
779: differences in the charges between the first two generations and the
780: third one are model dependent to some extent. However, one can minimize
781: these uncertainties by considering the combinations that are more
782: directly related to the fermion masses, as given by (\ref{Yuky}).  For
783: the first family sfermions, as compared to the third family ones, one
784: finds,
785: \begin{eqnarray}
786: m^2_{\tilde{e}_L}+m^2_{\tilde{e}_R} -m^2_{\tilde\tau_L}-m^2_{\tilde\tau_R}
787: &\approx& \frac{\ln (m_e / m_{\tau} )}{\ln \xi}m_{3/2}^2 \, \nonumber \\
788: m^2_{\tilde{d}_L}+m^2_{\tilde{d}_R} -m^2_{\tilde{b}_L}-m^2_{\tilde{b}_R}
789: &\approx& \frac{\ln (m_d / m_{b} )}{\ln \xi}m_{3/2}^2 \, \nonumber \\
790: m^2_{\tilde{u}_L}+m^2_{\tilde{u}_R} -m^2_{\tilde{t}_L}-m^2_{\tilde{t}_R}
791: &\approx& \frac{\ln (m_u / m_{t} )}{\ln \xi}m_{3/2}^2 \ . \label{sfirst}
792: \end{eqnarray}
793: These are high energy relations that remain valid at low energies as
794: far as the masses of the third generation are taken from
795: (\ref{sleptons}) and (\ref{squarks}) without the terms proportional to
796: $\rho .$ Analogous expressions hold for the second generation of
797: sfermions.  Of course, parity conjugated sfermions are usually splitted
798: by a large amount with respect to the above averages. If we take, as an
799: example, $\sigma_H \approx \sigma_L\approx \sigma_Q \approx -1 ,$ which
800: leads to relatively light charginos (without further motivation for
801: this particular choice), all the third family sfermions are as light as
802: the gauginos, the first and second families are much heavier. The
803: two-loop contributions are important in this case where the fine-tuning
804: between $ M_{1/2}$ and $m_{3/2}$ is large.
805: 
806: \begin{figure}
807: \epsfxsize=12.cm
808: $$\epsfbox{rap.ps}$$
809: \caption{The mass spectrum in units of $m_{3/2}$. We have displayed the 
810: lightest chargino mass and the left-right average masses for the $U$ squarks, 
811: the $D$ squarks and the sleptons as a function of $\sigma_H$
812:  appropriately shifted. }
813: \end{figure}
814: 
815: In the numerical analysis, the radiative corrections are included, and
816: the parameter space is scanned around the maximal fine-tuning values.
817: As expected we find Higgsino-like LSP's degenerate with  the lightest
818: chargino. The chargino masses can be as low as $100$ GeV.  We have
819: explicitly cut the spectrum by (arbitrarily) imposing that the MSSM
820: Higgs mass is greater than $100$ GeV. We do find Higgses within the
821: $100-109$ GeV slot corresponding to a value of $\tan \beta$ which does
822: not exceed $18$. Among the squarks the left sbottoms are the lightest.
823: We present in fig. 1 the mass spectrum as a function of $ \sigma_H$. We
824: have rescaled the masses and display them in units of $m_{3/2}$.  As
825: expected the hierarchy between families is not destroyed by the
826: evolution down to the Fermi scale. The values of $m_{3/2}$ chosen in
827: the figure are below $2$  TeV.  Higher values of $m_{3/2}$ would not
828: modify the picture, only the fine-tuning would be more severe.
829: 
830: Let us come back to one of the issues which prompted our study: the
831: FCNC effects and the decoupling of the first two families. The mass
832: insertions that are usually used to evaluate the FCNC
833: contributions\cite{mass} are roughly given in terms of the $X$ charges
834: of the particles by
835: \begin{equation}
836: \delta_{ij}\sim 2\frac{\vert X_i -X_j \vert }{X_i + X_j}
837: \xi^{\vert X_{i}-X_{j}\vert }
838: \end{equation}
839: The strong constraints on the mass insertions  with $i=1$ and $j=2$,
840: suggests\cite{dps} a choice of some degeneracy and some alignment in
841: the diagonal soft masses by choosing $d_1=d_2$ and $e_1=e_2$.
842: However, this is not enough and we still need large values of
843: the supersymmetry breaking parameter, $m_{3/2} \sim 2{\rm TeV}$ for a
844: sufficient FCNC decoupling.  Indeed, as noticed before, the flavour
845: dependence of the soft terms coming from the $\left< F\right>$
846: supersymmetry breaking have also to be taken into account. They are
847: reduced by at least a factor $\xi ^2,$ as follows from
848: (\ref{breaking}), and more model dependent, but still dangerous enough
849: for the $K-\bar{K}$ system. Fortunately we do get such high values of
850: $m_{3/2}$ in our numerical scanning without further effort. Yet, the
851: contribution to $\epsilon _K$ comes out close to the phenomenological
852: bounds in this model, in spite of the combined use of all three
853: anti-FCNC mechanisms, degeneracy and alignment from the equality of
854: some charges, together with decoupling through a relatively large
855: supersymmetry breaking scale in the scalar sector. This is the price to
856: pay for the inverse sfermion mass hierarchy. 
857: 
858: It is worth noticing that the small $ M_{1/2}/m_{3/2}$ and $ A/m_{3/2}$
859: ratios that characterize these models are what is needed\cite{abe} to avoid 
860: charge and colour breaking vacua without need for further cosmological
861: assumptions. In the large $\mu$ version of the model, one can reduce the
862: hierarchy and increase the degeneracy by increasing the $ M_{1/2}/m_{3/2}$
863: ratio. Besides the fact that it would bring back the issue of a fine-tuning
864: of the $\mu / M_{1/2}$ ratio, this would be strongly constrained by the
865: wrong vacuum problems.
866: 
867: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
868: \section{Summary and Concluding Remarks}
869: We have studied the decoupling of the first two squark families in
870: order to lower the FCNC effects. This has been done using the gauged
871: $U(1)$ flavour symmetries which had already been utilised to explain the
872: fermion masses  and the mixing angles. Within this framework we have
873: focused on more model-independent results. In particular as soon as one
874: tries to induce large mass hierarchies one faces the fine tuning
875: problem in the electroweak sector. Indeed the Fermi scale has to be
876: maintained although the supersymmetry breaking scale is pushed up
877: beyond the TeV limit.  This forces to study carefully the diverse
878: compensations in the $M_Z^2$ equation. As a result one has to resort to
879: a two-loop analysis of the Higgs sector. Fortunately we have shown that
880: the two-loop effects are solely governed by the Green-Schwarz anomaly
881: which is determined from the fermion masses. This allows a thorough
882: study of the minimum equations, and the possibility of a scenario
883: where  the fine-tuning appears in the $(M_{1/2},m_{3/2})$ sector with a
884: small value for $\mu$.  This differs from the usual cMSSM where $\mu$
885: is large. This leads to a Higgsino-like LSP and a characteristic mass
886: spectrum. In particular we find light charginos. On the contrary, the
887: sfermions of the first and second families should be of order a few TeV,
888: a nice experimental signature, indeed.
889: 
890: Of course, the inverse hierarchy models based on abelian flavour
891: symmetries are especially affected by the FCNC problems. The
892: supersymmetry breaking scale required to get an efficient decoupling
893: would be very high. Therefore, it is not clear whether they are a good
894: choice to escape the flavour changing effects in spite of the natural
895: prediction of heavy sfermions in the first two families. On the other hand they also
896: possess some other nice features: a natural small scale from the
897: Fayet-Iliopoulos term, the presence of anomalous abelian symmetries in
898: superstring solutions, the simplicity of the fitting to the puzzling
899: fermion hierarchy and, last but not least, the relation between the
900: fermion and sfermion spectrum. A compromise could be obtained with
901: additional flavour symmetries, for instance non-abelian ones, to reduce
902: the splitting between the sfermion in the first two families. More
903: speculatively, one could hope that the more recent developments in
904: string theory -- see for instance \cite{aqi, alda} and references
905: therein -- would provide new insights into the old quarrel of
906: supersymmetry with flavour. 
907:    
908: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
909: \newpage
910: \section{Appendix}
911: In the case of more than one abelian flavour charges, $X_i,$ $(i=1,....,
912: n),$ we introduce an equal number of scalars, $\Phi ^i,$ so that all
913: the $U(1)$'s are broken. For the consistency of the model, we make the
914: following assumptions:
915: 
916: \noindent A) Only one $U(1)$ symmetry is anomalous, which we call
917: $X_1,$ and only the corresponding $D-$term has a Fayet-Iliopoulos term
918: with coefficient $\xi .$ This is related to the anomaly $\cal{A}$
919: through the Green-Schwarz mechanism by (\ref{xi}). It is mandatory that
920: the other abelian charges fulfil the analogous of (\ref{Acond}) and
921: that they do not introduce any other anomaly.
922: 
923: \noindent B) Let us denote by $-\phi _{ij}$ the charge $X_i$ of the
924: scalar $\Phi ^j$. They are chosen so that there is no term in the
925: superpotential with the $\Phi '$s alone. These $U(1)$ charges are
926: normalized so that all the corresponding coupling constants are equal.
927: 
928: The relevant soft-terms are the masses $m_i$ of the scalars $\Phi ^i.$
929: Let $\phi ^{-1}_{\  ij}$ be the inverse of the charge matrix $\phi$
930: defined above, which has an inverse because of our assumption B). The
931: equivalent of (\ref{breaking}) is now,
932: \begin{eqnarray}
933: \vert \phi_i\vert ^2&=& \phi ^{-1}_{\  i1}\xi^2 M_P^2, \nonumber \\
934: \left<D_{X_i}\right> &=&m_{j}^2 \phi ^{-1}_{\  ji}\label{breakingnew}
935: \end{eqnarray} 
936: Then, (\ref{calA}) is modified by the replacement,
937: \begin{eqnarray}
938: \xi ^{\mbox{${\cal A}$}}\ \ \longrightarrow \ \  \prod \left( \phi ^{-1}_{\  i1}\xi^2
939: \right)^{\mbox{${\cal{A}}_i /2$}} , \label{replace}
940: \end{eqnarray} 
941: where ${\cal A}_i = \phi ^{-1}_{\  i1}{\cal A}.$ Therefore, in the
942: pluri-$U(1)$ case, the resulting value for the anomaly can be slightly
943: different  from the value in section $3$.
944: 
945: Finally, the contributions to the sfermion masses from the
946: $\left<D_{X_i}\right>$ terms are $ m_{j}^2 \phi ^{-1}_{\  ji}X_i (a)$,
947: where $X_i (a)$ is the corresponding fermion charge. The contribution
948: to the two-loop scalar masses becomes,
949: \begin{eqnarray}
950: \delta=\hat \delta +\frac{1}{8\pi^2}m_{j}^2 \phi ^{-1}_{\  j1}{\cal A}
951: \label{deltamulti}
952: \end{eqnarray}
953: where $\delta$ is defined in (\ref{delta}). This allows for some
954: variation with respect to the values discussed in section $3$, but the
955: two-loop contributions generically remain as small.
956: 
957: 
958:  
959: \begin{thebibliography}{99}
960: 
961: \bibitem{nls}
962: M. Leurer, Y. Nir, N. Seiberg,  {\it Nucl. Phys.} {\bf B398} (1993) 319;
963: {\bf B420} (1994) 468;
964: Y. Nir and N. Seiberg, {\it Phys.\ Lett. }{\bf B309 }(1993) 337.
965: 
966: \bibitem{dg}
967: S. Dimopoulos; G.F. Giudice, {\it Phys. Lett.} {\bf 357B} (1995) 573;
968: A. G. Cohen, D. B. Kaplan, A. E. Nelson, {\it Phys. Lett.}{\bf 388B}
969: 588 (1996);
970: A. Pomarol, D. Tommasini, {\it Nucl.\ Phys. }{\bf B466 }(1987) 3.
971: 
972: \bibitem{mura} 
973: N. Arkani-Hamed, H. Murayama, {\it Phys. Rev.}{\bf D56 }(1997) 6733;
974: K. Agache, M. Graesser, {\it Phys. Rev.}{\bf D59 }(1999) 015007.
975: 
976: \bibitem{FN}  
977: C. D. Froggatt and H. B. Nielsen, {\it Nucl.\ Phys. }%
978: {\bf B147 }(1979) 277;
979: J. Bijnens and C. Wetterich, {\it Nucl.\ Phys. }{\bf B283 }(1987) 237.
980: P. Ramond, R.G. Roberts and G. G. Ross, {\it Nucl.\
981: Phys.\ }{\bf B406 }(1993) 19.
982: 
983: \bibitem{ir}  L. E. Ib\'{a}\~nez and G. G. Ross, 
984: {\it Phys. Lett. }{\bf B332 }(1994) 100;
985: P. Bin\'{e}truy and P. Ramond, {\it Phys.\ Lett.\ }{\bf %
986: B350 }(1995) 49;
987: V. Jain and R. Shrock, {\it Phys.\ Lett. }{\bf B352 }%
988: (1995) 83;
989: E. Dudas, S. Pokorski and C. A. Savoy, hep-ph/9504292, 
990: Phys. Lett. {\bf B356} (1995) 45;
991: Y. Nir, hep-ph/9504312, Phys. Lett. {\bf B354} (1995) 107.
992: 
993: \bibitem{dps}
994: E. Dudas, S. Pokorski and C. A. Savoy, 
995: {\it Phys.\ Lett. }{\bf B369 } (1996) 255;
996: E. Dudas, C. Grojean, S. Pokorski and C. A. Savoy, 
997: {\it Nucl. Phys.} {\bf B481} (1996) 85.
998: 
999: \bibitem{dd}
1000: P. Bin\'{e}truy and E. Dudas, {\it Phys. Lett. }{\bf B389} (1996) 503;
1001: G. Dvali, A. Pomarol, {\it Phys. Rev. Lett. }{\bf 77} (1996) 3728.
1002: 
1003: \bibitem{nel}
1004: S. Ambrosiano, A. E. Nelson, {\it Phys. Lett. }{\bf B411} (1997) 283;
1005: A. E. Nelson, D. Wright, {\it Phys. Rev.}{\bf D56 }(1997) 1598
1006: 
1007: \bibitem{jack}
1008: I. Jack, D. R. T. Jones, S. P. Martin, M. T. Vaughn and Y. Yamada
1009: {\it Phys. Rev.} {\bf D50} (1994) 5481.
1010: \bibitem{poko}
1011: M. Olechowski, S. Pokorski, {\it Phys.\ Lett.\ }{\bf B344 }(1995) 201; 
1012: N. Polonski, A. Pomarol, {\it Phys. Rev. Lett. }{\bf 73} (1994) 2292;
1013: D. Matalliotakis, H. P. Nilles, {\it Nucl.\ Phys.\ }{\bf B435 }(1995) 115.
1014: P. H. Chankowski, J. Ellis, S. Pokorski, 
1015: {\it Phys.\ Lett.\ }{\bf B423 }(1998), 327;
1016: G. L. Kane, S. F. King, 
1017: 
1018: \bibitem{aqi} 
1019:  L. E. Ib\'a\~{n}ez, C. Munoz, S. Rigolin, {\it Nucl.\ Phys.\ }{\bf B553}
1020: (1999) 43; 
1021: L.E. Ib\'a\~{n}ez, R. Rabadan , A.M. Uranga, {\it Nucl.\ Phys.\ }{\bf B542}
1022: (1999) 112. 
1023: 
1024: \bibitem{dws}  
1025: M. Dine, N. Seiberg and E. Witten, {\it Nucl.\ Phys.\ }%
1026: {\bf B289 }(1987) 585;
1027: J. Atick, L. Dixon and A. Sen, {\it Nucl.\ Phys. }{\bf B292 }(1987) 109;
1028: M. Dine, I. Ichinose and N. Seiberg, {\it Nucl.\ Phys. }{\bf B293 }(1987)
1029: 253.
1030:  
1031: \bibitem{mass} 
1032: F. Gabbiani, E. Gabrielli, A. Masiero, L. Silvestrini,
1033: {\it Nucl. Phys.} {\bf B447} (1996) 321.
1034: 
1035: \bibitem{abe}
1036: J.A. Casas, A. Lleyda, C. Munoz, {\it Phys. Lett. }{\bf B380} (1996) 59.
1037: S.A. Abel, C.A. Savoy, {\it Nucl. Phys.} {\bf B532} (1998) 3.
1038: S.A. Abel, C.A. Savoy, {\it Phys. Lett. }{\bf B444} (1998) 119.
1039: 
1040: \bibitem{alda} 
1041: G. Aldazabal, L.E. Ibanez, F. Quevedo, {\it JHEP} {\bf 0001:031} (2000).
1042: 
1043: \end{thebibliography}
1044: 
1045: \end{document} 
1046: 
1047: