hep-ph0201162/num.tex
1: \documentclass [a4paper]{JHEP3}
2: \title{
3: \protect\vspace{5mm}
4: Towards the realistic fermion masses with a single
5: family in extra dimensions.}
6: \author{Maxim Libanov and Emin~Nougaev\\
7: Institute for Nuclear Research of the Russian Academy of
8: Sciences,\\
9: 60th October Anniversary Prospect 7a, 117312, Moscow, Russia.\\
10: E-mail: \email{ml@ms2.inr.ac.ru, emin@ms2.inr.ac.ru}
11: }
12: \abstract{
13: In a class of multidimensional models, topology of a thick brane
14: provides three chiral fermionic families with hierarchical masses and
15: mixings in the effective four-dimensional theory, while the full model
16: contains a single vector-like generation. We carry out  numerical
17: simulations and reproduce all known Standard Model fermion masses and
18: mixings in one of these models.  }
19: \keywords{Extra Large Dimensions, Quark
20: Masses and SM Parameters,  Beyond Standard Model }
21: 
22: \begin{document}
23: 
24: \section{Introduction} \label{sec:1}
25: 
26: One of interesting features inherent in the theories with more than
27: four spacetime dimensions is a possibility to explain the mysterious
28: pattern of fermion mass hierarchies~\cite{AD,LT,LTF} (see also
29: review~\cite{Rubakov} and references therein). In the previous series of
30: works~\cite{LT,LTF,neutrino}, we have constructed a model in which a
31: single family of fermions, with vector-like couplings to the Standard
32: Model (SM) gauge groups in 6 dimensions, gives rise to three generations
33: of  chiral Standard Model  fermions in 4 dimensions. This mechanism is
34: based on  localization of fermion zero modes on a two-dimensional vortex
35: with winding number $n$. In particular, in the paper~\cite{LT} the case of
36: global vortex (i.e. group of symmetry of the vortex $U_g(1)$ is global)
37: with winding number $n=3$ was considered. In this setup three localized
38: fermionic zero modes appear due to specific Yukawa coupling to the
39: vortex scalar $\Phi $. Coupling of fermions to the SM Higgs doublet $H$
40: results in four-dimensional effective fermion masses.  Inter-generation
41: mixings occur due to explicit breaking of $U_g(1)$ symmetry.  The latter
42: point is the main drawback of this model since it does not allow to
43: gauge $U_g(1)$ symmetry. In the paper~\cite{LTF} this problem was
44: overcome.  The price for unbroken $U_g(1)$, however, is the necessity to
45: invoke higher dimension operators  in the scalar-fermion
46: interactions\footnote{The requirement of renormalizability of the theory
47: does not make much sense anyway in the six-dimensional models, since even
48: usual Yukawa scalar-fermion-fermion coupling is non-renormalizable.}.
49: More precisely, to obtain $n$ fermionic generations we consider the vortex
50: with winding number 1 but fermions  coupled to the vortex scalar
51: raised to the $n$th power. In both cases (global and gauged $U_g(1)$
52: symmetry) the hierarchical pattern of the fermion masses occurs due to
53: different profiles of the fermionic wave functions in the transverse extra
54: dimensions.  Namely, in Refs.~\cite{LT,LTF} a crude
55: dimensional analysis has been presented which shows that hierarchical mass
56: pattern has the form
57: \begin{equation}
58: 1:g^2:g^4,
59: \label{OldHierarchy}
60: \end{equation}
61: where $g$ is small Yukawa coupling, which yields the localization of
62: fermionic modes. Also, in Refs.~\cite{LT,LTF} it has been pointed out that
63: any exact prediction of the fermion masses (as well as a check of the
64: Eq.(\ref{OldHierarchy})) can be obtained only numerically.
65: 
66: In this paper we return to the models of Refs.~\cite{LT,LTF} and pay
67: special attention to a check of the Eq.(\ref{OldHierarchy}). We will see
68: that this equation literally does not hold. Namely,  in
69: the case of gauged vortex the hierarchical mass pattern takes the form
70: \[
71: 1:g:g^2.
72: \]
73: At the same time, in the case of global vortex unnatural hierarchical
74: pattern appears.  We will also evaluate fermion masses numerically.  For
75: reasonable values of parameters in the case of the gauged vortex our
76: results reproduce SM fermionic mass pattern.  Furthermore, we will obtain
77: Cabbibo-Kobayashi-Maskawa (CKM) matrix for the same set of parameters, and
78: see that six-dimensional theory reproduces the mixing matrix quite well.
79: 
80: \section{The Setup} \label{sec:2}
81: 
82: In what follows we restrict ourselves  to  the
83: model of Ref.~\cite{LTF}. In this section we give a brief description of
84: this construction.  Our notations coincide with those used in
85: Refs.~\cite{LT,LTF}. In particular, six-dimensional coordinates $x_A$
86: are labeled by capital Latin indices $A,B=0,\dots,5$. Four--dimensional
87: coordinates $x_\mu $ are labeled by Greek indices $\mu,\nu =0,\dots,3$.
88: The Minkowski metric is $g_{AB}=\mbox{diag}(+,-,\dots,-)$. We use also
89: chiral representation for six-dimensional Dirac $\Gamma $-matrices (see
90: Ref.~\cite{LT}).  Besides this we introduce polar coordinates $r$, and
91: $\theta$ in the $x_4$, $x_5$ plane.
92: 
93: The matter field content of the six-dimensional theory is summarized in
94: Table \ref{table:fields}. There are three scalar fields. One of these
95: $\Phi $, together with the $U_g(1)$ gauge field, forms a vortex, while two
96: other scalars, $X$ and $H$, develop profiles localized on the vortex.  The
97: potential term of the Lagrangian which gives rise to the non-trivial
98: profiles for the scalar fields has the following form
99: \begin{equation}
100: V_s=
101: {\lambda\over 2}\left( |\Phi|^2-v^2\right)^2+
102: {\kappa\over 2} \left( |H|^2-\mu^2\right)^2+h^2|H|^2|\Phi|^2+
103: {\rho\over 2} \left( |X|^2-v_1^2\right)^2+\eta^2|X|^2|\Phi|^2.
104: \label{ScalarPotential}
105: \end{equation}
106: 
107: \TABLE[ht]{
108: %\begin{center}
109: \begin{tabular}{|rc|c|c|c|c|c|}
110: \hline
111: \multicolumn{2}{|c|}{fields}
112: & profiles&\multicolumn{2}{|c|}{charges}&
113: \multicolumn{2}{|c|}{representations}\\
114: \cline{4-7}
115: &&&$U_g(1)$&$U_Y(1)$&$SU_W(2)$&$SU_C(3)$\\
116: \hline
117: scalar&$\Phi$&$F(r){\rm e}^{i\theta}$&+1&0&{\bf 1}&{\bf 1}\\
118: &&$F(0)=0$, $F(\infty)=v$&&&&\\
119: \hline
120: scalar&$X$&$X(r)$&+1&0&{\bf 1}&{\bf 1}\\
121: &&$X(0)=v_X$, $X(\infty)=0$&&&&\\
122: \hline
123: scalar&$H$&$H(r)$&$-1$&$+1/2$&{\bf 2}&{\bf 1}\\
124: &&$H(0)=v_H$, $H(\infty)=0$&&&&\\
125: \hline
126: fermion&$Q$&3 L zero modes&axial $+3/2$&$+1/6$&{\bf 2}&{\bf 3}\\
127: \hline
128: fermion&$U$&3 R zero modes&axial $-3/2$&$+2/3$&{\bf 1}&{\bf 3}\\
129: \hline
130: fermion&$D$&3 R zero modes&axial $-3/2$&$-1/3$&{\bf 1}&{\bf 3}\\
131: \hline
132: fermion&$L$&3 L zero modes&axial $+3/2$&$-1/2$&{\bf 2}&{\bf 1}\\
133: \hline
134: fermion&$E$&3 R zero modes&axial $-3/2$&$-1$&{\bf 1}&{\bf 1}\\
135: \hline
136: \end{tabular}
137: %\end{center}
138: \caption{Scalars and fermions with their gauge quantum numbers~\cite{LTF}
139: .
140: For convenience, we describe here also the profiles of the classical
141: scalar fields and fermionic wave functions in extra dimensions.}
142: \label{table:fields}
143: }
144: 
145: There is also one fermionic generation which consists of five
146: six-dimensional fermions $Q$, $D$, $U$, $L$ and $E$. These fermions have
147: vector-like coupling to SM gauge bosons, (a subtle issue of localization
148: of the latters we do not discuss here), and axial couplings with the
149: vortex background\footnote{For the scalar--fermion interactions, we take
150: the most general operators of the lowest order, consistent with  gauge
151: invariance. On the other hand, in the scalar self--interaction
152: (\ref{ScalarPotential}) there could be terms of the form $\Phi X^*$ or
153: $\Phi X^*|H|^2$ which we do not include. These terms, as well as terms of
154: the form $\Phi ^2X\bar Q (1-\Gamma _7)Q$ in the (\ref{VPhiF}), can be
155: forbidden, for example, by the imposition of $Z_4$ discrete symmetry:
156: $\Phi \to i\Phi $, $Q\to \exp(i\Gamma _7\frac{3\pi }{4})Q$, $X\to X$. This
157: symmetry is explisitly broken by the interaction (\ref{V2}), (\ref{V3}),
158: and, therefore the undesirable terms will be generated by the radiative
159: corrections. However, the value of these corrections is of order
160: $\frac{Y^2}{(4\pi )^3} \epsilon \Lambda ^4$ (for the operator $|H|^2\Phi
161: X^*$, for instance; $\Lambda $ is the cut off), and for our choice of the
162: parameters (see Sec. \ref{sec:5}) is of order $10^{-6}\left(\frac{\Lambda
163: }{M} \right)^4 \lambda $, where $M$ is mass scale, and $\lambda =1/M^2$.
164: So, if we choose the cut off to be of order $\Lambda \sim 10 M$, then
165: this corrections can be neglected.}
166: \begin{equation}
167: \begin{array}{c}
168: \displaystyle V_1=g_q\Phi^3\bar Q{1-\Gamma_7\over 2} Q+
169: g_u\Phi^{*3}\bar U{1-\Gamma_7\over2} U + g_d\Phi^{*3}\bar
170: D{1-\Gamma_7\over 2} D+\\
171: \displaystyle g_l\Phi^3\bar L{1-\Gamma_7\over 2} L+ g_e\Phi^{*3}\bar
172: E{1-\Gamma_7\over2} E+{\rm h.c.}
173: \end{array}
174: \label{VPhiF}
175: \end{equation}
176: \nopagebreak
177: ($\Gamma _7$ is a six-dimensional analog of four-dimensional matrix
178: $\gamma _5$).
179: 
180: Note that the scalar background $\Phi^3$, where the field $\Phi$ has
181: the winding number one, $\Phi=F(r){\rm e}^{i\theta}$, has exactly the same
182: topological properties as the background $\Phi_1$ of the vortex with
183: winding number three, $\Phi_1=F_1(r){\rm e}^{3i\theta}$. As a result, the
184: interactions Eq.~(\ref{VPhiF}) give rise to {\em three} left--handed
185: (right--handed) zero modes
186: for each of the fermions $Q$, $L$ ($U$, $D$, $E$). All of these modes are
187: localized in the core of the vortex\footnote{These modes were discussed in
188: detail in Ref.~\cite{LT}.}, and could be identified with the three
189: generations of the chiral SM fermions. The last point, however, is not
190: quite fair since these {\it zero} modes are {\it massless} from the
191: four-dimensional point of view. To give small (compared to the vortex
192: scale $\sqrt{\lambda }v$) masses to these modes we introduce the following
193: interactions of the fermions with the Higgs doublet $H$ and singlet $X$
194: \begin{equation}
195: V_2= Y_dHX\bar Q\frac{1-\Gamma_7}{2}D+ Y_u\tilde HX^*\bar
196: Q\frac{1-\Gamma_7}{2}U+ Y_lHX\bar L\frac{1-\Gamma_7}{2}E+ {\rm h.c.},
197: \label{V2}
198: \end{equation}
199: where $\tilde H_\alpha =\epsilon_{\alpha \beta }H^*_\beta
200: $, $\alpha $, $\beta $ are $SU_W(2)$ indices. Finally, there is one
201: more gauge invariant in the scalar--fermions interaction
202: \begin{equation}
203: V_3= Y_d\epsilon_dH\Phi\bar Q\frac{1-\Gamma_7}{2}D+
204: Y_u\epsilon_u\tilde H\Phi^*\bar Q\frac{1-\Gamma_7}{2}U+
205: Y_l\epsilon_lH\Phi\bar L\frac{1-\Gamma_7}{2}E+ {\rm h.c.},
206: \label{V3}
207: \end{equation}
208: (we denote Yukawa coupling constants in Eq.~(\ref{V2}) as
209: $Y_{u,d,l}$, and in Eq.~(\ref{V3}) as $Y_u\epsilon_u, \dots$, for
210: convenience). This term yields inter-generation mixings.
211: 
212: It is worth noting that to obtain small fermion masses and mixing, it is
213: not necessary to take $Y$ or $\epsilon $ small. Small masses and mixings
214: originate from small overlaps of the fermionic wave functions and the
215: scalar profiles.
216: 
217: Now let us consider this model in more detail. First of all we will
218: present stable non-trivial configuration for the scalars and
219: $U_g(1)$ gauge fields. Then we will find fermion zero modes in this
220: background.  Finally, we will calculate the fermion masses and the CKM
221: matrix.
222: 
223: 
224: 
225: \section{The vortex background}\label{sec:3}
226: 
227: As described in the previous section, in our model there are
228: three scalar fields and one $U_g(1)$ gauge field which form a non-trivial
229: vortex background. In this section we present a stable solution of the
230: static field equations which meets all the requirements.
231: 
232: The set of the static equations to the background has the
233: following form (prime denotes the derivative with respect to $r$; $e$ is
234: $U_g(1)$ gauge charge),
235: \begin{eqnarray}
236: &&F''+{1\over r}F'-{1\over r^2}F(1-A)^2
237: -\lambda v^2 F \left( F^2-1\right)-h^2FH^2-\eta^2FX^2 =0,\nonumber\\
238: &&A''-{1\over r}A'+2e^2v^2F^2(1-A)-2e^2H^2A-2e^2X^2A=0,\nonumber\\
239: &&H''+{1\over r}H'-\frac{1}{r^2}HA^2-h^2 H v^2F^2-\kappa H (H^2-\mu^2)
240: =0,\nonumber \\
241: &&X''+{1\over r}X'-\frac{1}{r^2}XA^2-\eta^2 X v^2F^2-\rho X (X^2-{v_1}^2) =0,
242: \label{qvheq}
243: \end{eqnarray}
244: with the boundary conditions
245: \begin{eqnarray}
246: &&F(r)= 0,\; H'(r)= 0,\; X'(r)=0, \;A(r)= 0,\qquad {\rm at}\ \ r=
247: 0,\nonumber\\
248: &&F(r)\to 1,\; H(r)\to 0, \;X(r)\to 0,\; A(r)\to 1,\qquad {\rm at}\ \
249: r\to\infty.
250: \label{boundarycondition}
251: \end{eqnarray}
252: Here we use the standard anzatz for Abrikosov-Nielsen-Olesen vortex
253: ($i,j=4,5$)
254: \[
255: A_{i}(x)=-\frac{1}{er^2}\epsilon _{ij}x_{j}A(r),\ \ \
256: \Phi(x)=vF(r)e^{i\theta},
257: \]
258: \begin{equation}
259: H_ \alpha (x) = \delta _{2 \alpha }H(r),\ \ \ X(x)=X(r).
260: \label{anzatz}
261: \end{equation}
262: 
263: Let us briefly discuss the choice of the boundary conditions
264: (\ref{boundarycondition}) and the anzatz (\ref{anzatz}). At infinity the
265: fields should tend to their vacuum expectation values (VEV) which are
266: determined from the potential (\ref{ScalarPotential}). In general at
267: arbitrary set of the coupling constants the potential
268: (\ref{ScalarPotential}) may admit several minima including those in which
269: $X$ and (or) $H$ are non-zero. However, there are several regions in the
270: space of the parameters where the potential (\ref{ScalarPotential}) has
271: {\it only one} minimum $\Phi =v$, $H=X=0$. In particular, one of the
272: regions is determined by the conditions $h^4>\lambda \kappa $, $\lambda
273: \rho >\eta ^4$, $\eta ^2v^2>\rho v_1^2$, and in what follows we will
274: assume that these conditions are satisfied.
275: 
276: In general, any configuration of the fields $\Phi $, $H$, and $X$ may be
277: characterized by three topological charges $Q_T=(Q_T^\Phi ,Q_T^H, Q_T^X)$.
278: The zero VEVs of $X$ and $H$ admit the existence of the topologically
279: trivial configuration of these fields (\ref{anzatz}) (in spite of the
280: interaction with winded gauge field), $Q_T=(1,0,0)$. In principle, there
281: may exist a solution of the field equation with $Q_T=(1,-1,1)$ and with
282: the same topological properties of the gauge field $A_\mu $. However, as
283: we will see, the solution with $Q_T=(1,0,0)$ is stable, and, therefore,
284: the choice of the anzatz (\ref{anzatz}) is apropriate.
285: 
286: 
287: As  mentioned in the previous section, scalar-fermion interactions
288: (\ref{V2}), (\ref{V3}) generate low energy mass terms for fermionic modes.
289: Thus, to obtain non-zero fermion masses one should have a non-trivial
290: solution for $H(r)$ and $X(r)$.  On the other hand, the boundary value
291: problem (\ref{qvheq}), (\ref{boundarycondition}) always adopts trivial
292: solutions $H=0$ or (and) $X=0$. Fortunately, it has been pointed out in
293: \cite{Witten}, that in some region of parameters these trivial  solutions
294: are unstable. Namely, the necessary conditions to have non-trivial stable
295: solutions are $h^2v^2>\kappa\mu^2$ and $\eta^2 v^2>\rho v_1^2$
296: ~\cite{Witten}.  However, it turns out that these conditions are
297: insufficient. To clarify the situation, let us study the problem of the
298: stability in more detail.  The stability of the scalar $\Phi $ and gauge
299: fields is guaranteed by topology.  To investigate the stability of the
300: static solution $H(r)$ one  linearizes equations of motion
301: (\ref{qvheq})  and obtain Schr\"{o}dinger-type problem with
302: %"
303: the potential
304: \begin{equation}
305: h^2v^2F^2(r)+\frac{1}{r^2}A^2-\kappa(\mu^2-3H^2(r)).
306: \label{shrodinger}
307: \end{equation}
308: 
309: At large $r$ the solution of  Eq.(\ref{qvheq}), $H(r)$, has
310: asymptotic $H\sim\exp(-\sqrt{h^2v^2-\kappa\mu^2}\; r)$. So, at large
311: enough $h$ the Higgs profile becomes narrow  compared to $F(r)$. This
312: means, in particular, that there is a region of $r$ where $H(r)$ becomes
313: negligible while $A(r)$ and $F (r)$ are still
314: almost equal to zero. Therefore, the main contribution to the potential
315: (\ref{shrodinger}) in this region comes from $r$-independent term $-\kappa
316: \mu ^2$ and is negative. Thus, the potential becomes not positively
317: definite, and there  may exist negative levels which  mean the
318: instability.  The numerical calculations show that this is the case.
319: 
320: \FIGURE[th]{
321: {
322: \input{picture1.tex}} \leavevmode
323: \caption{Profiles of the background at $\lambda=\kappa=\rho=1$, $v=\mu=1$,
324: $e=0.04$, $h=1.02$, $\eta=0.31$, $v_1=0.3$.
325: \label{fig1}
326: } }
327: 
328: To be specific, we have chosen the following set
329: of the parameters:  $v=\mu=1\cdot
330: M^2$, $v_1=0.3\cdot M^{2}$, $\lambda=\kappa=\rho=1\cdot M^{-2}$,
331: $h=1.02\cdot M^{-1}$, $\eta=0.31\cdot M^{-1}$, $e=0.04\cdot M^{-1}$,
332: where $M$ is a unit of mass which we will define in Section (\ref{sec:5}).
333: In numerical calculations we assume $M=1$.
334: To solve the boundary value problem (\ref{qvheq}),
335: (\ref{boundarycondition}) we have used the relaxation method \cite{Num}.
336: The solutions for profiles $F(r)$, $H(r)$, $X(r)$ and $A(r)$ in the
337: vicinity of the origin are presented in Figure \ref{fig1}.
338: The potential (\ref{shrodinger}) for these solutions is
339: positive and, therefore, the solutions are stable.
340: 
341: 
342: The vortex background in the case of global $U_g(1)$ group can be obtained
343: from the Eqs.(\ref{qvheq}), (\ref{boundarycondition}) by the setting $e=0$,
344: $A(r)=0$ ($A(\infty )=0$). The profiles for the scalar fields have the
345: same shapes as in the gauge case. We do not present the corresponding plot
346: here.
347: 
348: 
349: \section{Mass hierarchy: cases of global and gauged $U_g(1)$}
350: \label{sec:4}
351: 
352: Now we have all ingredients to find fermionic zero modes in the vortex
353: background and to obtain hierarchy of the low energy fermionic masses.
354: To do this let us first briefly remind the analysis
355: given in Refs.~\cite{LT,LTF}.
356: 
357: \subsection{Global $U_g(1)$}
358: Let us consider first the case of global $U_g(1)$ group and one
359: fermionic family, say $Q$ and $U$.
360: The part of the Lagrangian describing the interaction of
361: six-dimensional spinor  $Q$ with vortex scalar $\Phi $ is
362: \begin{equation}
363: L_Q=i\bar{Q}\Gamma^A\partial_AQ
364: -\left(g_q\Phi^3\bar Q{1-\Gamma_7\over 2} Q+{\rm h.c.}\right).
365: \label{ferint}
366: \end{equation}
367: This Lagrangian is invariant under  $U_g(1)$ rotations
368: \[
369: Q\to{\rm e}^{i{\frac{3\alpha}{2}}\Gamma_7}Q,\,\,\,\Phi \to
370: {\rm e}^{i\alpha }\Phi.
371: \]
372: In the vortex background, one can expand an arbitrary spinor in
373: eigenfunctions of the corresponding Dirac operator.
374: There is a set of massive heavy modes (with masses of order $gv^3$) and
375: zero modes.  The careful analysis given in~\cite{fermionModes,LT,LTF}
376: shows that in the vortex background with winding number one
377: there are exactly three localized zero modes  describing four-dimensional
378: left-handed spinors.  To obtain three right-handed zero modes one should
379: consider a six-dimensional spinor $U$ which has an axial charge $-3/2$
380: under $U_g(1)$ with the Lagrangian
381: \begin{equation}
382: L_U=i\bar{U}\Gamma^A\partial_AU
383: -\left(g_u\Phi^3\bar U{1+\Gamma_7\over 2} U+{\rm h.c.}\right).
384: \label{ferintU}
385: \end{equation}
386: These left-handed and right-handed zero modes have the form
387: \begin{equation}
388: \begin{array}{c}
389: \displaystyle
390: Q_{p}=
391: \left(
392: \begin{array}{c}
393: 0 \\
394: q_{p}(r){\rm e}^{i(3-p)\theta}\\
395: q_{4-p}(r){\rm e}^{i(1-p)\theta}\\
396: 0 \\
397: \end{array}
398: \right)
399: ,\,\,\,
400: \displaystyle
401: U_p=
402: \left(
403: \begin{array}{c}
404: u_{4-p}(r){\rm e}^{i(1-p)\theta} \\
405: 0\\
406: 0\\
407: u_{p}(r){\rm e}^{i(3-p)\theta} \\
408: \end{array}
409: \right) \;.
410: \end{array}
411: \label{8**}
412: \end{equation}
413: Here $p=1,2,3$  enumerates three modes (three fermionic generations). The
414: radial functions $q_{p}$ are  solutions to the following set of
415: the equations,
416: \begin{eqnarray}
417: &&\displaystyle \left\{
418: \displaystyle \begin{array}{l}
419: \displaystyle \partial _rq_{p}-\frac{(3-p)}{r}q_{p}+g_qv^3F^3q_{4-p}=0\\
420: \\
421: \displaystyle \partial _rq_{4-p}-\frac{(p-1)}{r}q_{4-p}+g_qv^3F^3q_{p}=0
422: \end{array}
423: \right.
424: \label{ML523}
425: \end{eqnarray}
426: with the normalization condition
427: \begin{equation}
428: \int\limits_{0}^{\infty
429: }\!rdr(q_p^2+q_{4-p}^2)=\frac{1}{2\pi}.
430: \label{norm}
431: \end{equation}
432: The functions $u_p(r)$ satisfy the same equations with the
433: replacement $g_q\to g_u$.
434: 
435: Since the background radial function $F(r)$ is not known in analytical
436: form, Eqs.(\ref{ML523}) can be solved only numerically. We
437: will present the numerical results hereinafter, now, however, let us
438: consider the equations in more details.
439: 
440: We are interested in obtaining low energy fermion masses
441: which originate from the integration over extra dimensions of the
442: following expression
443: \begin{equation}
444: Y_u\int\!d^2x
445: \tilde{H}X^*\bar Q\frac{1-\Gamma_7}{2}U +
446: {\rm h.c.}
447: \label{massinter}
448: \end{equation}
449: Substituting here the zero modes~(\ref{8**}), and integrating over
450: polar angle $\theta$ one finds
451: \[
452: m_{ps}^u = Y_u\int \limits_{}^{}rdrd\theta H(r)X(r)q_{p}(g_q,
453: r)u_{s}(g_u,r) {\rm e}^{i\theta (r-s)}=
454: \]
455: \begin{equation}
456: \delta _{ps}2\pi Y_u \int\limits_{0}^{\infty
457: }\!rdr H(r)X(r)q_{p}(g_q, r)u_{p}(g_u,r).
458: \label{main}
459: \end{equation}
460: It is worth noting that due to orthogonality of the zero modes and due to
461: the fact that $H$ and $X$ do not depend on $\theta $ the integral over
462: $\theta $ leads to a selection rule, $m_{ps}^u\sim \delta _{ps}$. To
463: obtain non-trivial  inter-generation mixings one should also consider the
464: term
465: \begin{equation}
466: Y_u\epsilon _u\int\!d^2x
467: \tilde{H}\Phi ^*\bar Q\frac{1-\Gamma_7}{2}U +
468: {\rm h.c.}
469: \label{massintermiss}
470: \end{equation}
471: The non-trivial $\theta $-dependence of $\Phi $ (\ref{anzatz})
472: leads to the off-diagonal elements in the mass matrix
473: \[
474: m_{ps}^u = Y_u\epsilon _u\int \limits_{}^{}rdrd\theta H(r)\Phi(r)q_{p}(g_q,
475: r)u_{s}(g_u,r) {\rm e}^{i\theta (r-s-1)}=
476: \]
477: \begin{equation}
478: \delta _{p,s+1}2\pi Y_u \epsilon _u\int\limits_{0}^{\infty
479: }\!rdr H(r)\Phi(r)q_{p}(g_q, r)u_{p-1}(g_u,r).
480: \label{main1}
481: \end{equation}
482: In the same way one finds non-zero off-diagonal elements for down-type
483: quarks $m^d_{ps}\sim\delta _{p+1,s}$. Therefore, the mass matrices have the
484: following form
485: \[
486: m^u=\left(
487: \begin{array}{ccc}
488: m_{11}&0&0\\
489: m_{21}&m_{22}&0\\
490: 0&m_{32}&m_{33}
491: \end{array}
492: \right),\ \ \ \
493: m^d=\left(
494: \begin{array}{ccc}
495: m_{11}&m_{12}&0\\
496: &m_{22}&m_{23}\\
497: 0&0&m_{33}
498: \end{array}
499: \right).
500: \]
501: 
502: 
503: To evaluate the integrals (\ref{main}) and (\ref{main1}) over $r$ we take
504: into account that the functions $H(r)$ and $X(r)$ are localized inside the
505: vortex. So, the integral (\ref{main}) is saturated near the origin, and
506: therefore only the behavior of the zero modes at $r\to 0$ is relevant. It
507: is easy to find from equations (\ref{ML523}) that $q_p\sim r^{3-p}$ at
508: $r\to 0$, however, it is not so easy to obtain the dependence on $g_q$ of
509: the zero modes. To obtain the  dependence on small Yukawa couplings $g$ of
510: the fermion masses we should take into account the normalization condition
511: (\ref{norm}) and investigate the behavior of the zero modes near the
512: origin more carefully. To do this let us consider the following simplified
513: model.
514: 
515: Let us approximate the vortex background by the functions
516: \[
517: v^3F(r)^3=\left\{
518: \begin{array}[]{ll}
519: r^3,&r<1\\
520: 1,&r\ge 1
521: \end{array}
522: \right.
523: \]
524: \begin{equation}
525: H(r)=X(r)=\left\{
526: \begin{array}[]{ll}
527: 1,&r<1\\
528: 0,&r\ge 1
529: \end{array}
530: \right.
531: \label{aproxhiggs}
532: \end{equation}
533: These functions have the correct behavior at the origin and at infinity.
534: So, this is a reasonable approximation.  The zero mode
535: equations (\ref{ML523}) with the background Eq.(\ref{aproxhiggs}) have
536: analytical solution in terms of modified Bessel functions.  However, to
537: clarify the picture,  we present here only the leading dependence
538: on $g$ of the solutions  at $r<1$ (which is only relevant for our
539: purposes), and exact solutions at $r\ge 1$:
540: \[
541: q_1(r)=
542: \left\{
543: \begin{array}[]{ll}
544: \tilde{C_1}r^2,&r<1\\
545: \frac{Cr}{\sqrt{g_q}}e^{-g_qr},&r\ge1 \end{array} \right.  \ \ \ q_2(r)=
546: \left\{
547: \begin{array}[]{ll}
548: \tilde{C_2}r,&r<1\\
549: \frac{C_mr}{\sqrt{g_q}}e^{-g_qr},&r\ge1
550: \end{array}
551: \right.
552: \]
553: \begin{equation}
554: q_3(r)= \left\{
555: \begin{array}[]{ll}
556: \tilde{C_3},&r<1\\
557: \frac{Cr}{\sqrt{g_q}}e^{-g_qr}(1+{1\over{g_qr}}),&r\ge1
558: \end{array}
559: \right.
560: \label{aproxsolution}
561: \end{equation}
562: 
563: 
564: To find the dependence on $g_q$ of the coefficients $C$ and $\tilde{C}$
565: let us assume that the normalization integral (\ref{norm}) is saturated at
566: $r\ge 1$. Then substituting the solutions (\ref{aproxsolution}) into the
567: condition~(\ref{norm}) one has $C\simeq C_m\simeq g_q^{5/2}$. Thus, from
568: the continuity of the solutions at $r=1$, we find  that
569: $\tilde{C}_{1,2}\simeq g_q^2$ and $\tilde{C}_3\simeq g_q$. This means in
570: particular that our assumption is indeed valid:  the normalization
571: integrals at $r<1$ are of order $g_q^2$ and can be neglected.
572: 
573: Now we can find the mass pattern in this simplified model. Substituting
574: the solutions (\ref{aproxsolution}) and background (\ref{aproxhiggs}) into
575: integral (\ref{main}) we have the following dependence on $g_{q,u}$ of the
576: fermion masses,
577: \begin{equation}
578: m_{11}\sim m_{22} \sim (g_q g_u)^2,\ \ \ \ m_{33}\sim (g_q g_u).
579: \label{globalmasspattern}
580: \end{equation}
581: We see that $m_{11}$ and $m_{22}$ are parametrically of the same order,
582: and therefore there is no hierarchy in the case of global $U_g(1)$. This
583: result was confirmed by the numerical evaluation of the zero modes and the
584: integral (\ref{main}).
585: 
586: \subsection{Gauged $U_g(1)$}
587: Let us concentrate now on the case of the gauged $U_g(1)$. In this case
588: the equations for the zero modes can be obtained from (\ref{ML523}) by the
589: replacement
590: \[
591: \partial _r\to \partial _r+\frac{3A(r)}{2r}.
592: \]
593: Therefore, all dependence of the zero modes on the gauge field  can be
594: absorbed into a factor
595: \begin{equation}
596: (q_p)_{\rm gauge} = \exp{\left(- \int\limits^r {3A(r')
597: \over 2r'}dr'\right)}\cdot(q_{p})_{\rm global}.
598: \label{gageoff}
599: \end{equation}
600: Approximating  the gauge field as
601: \begin{equation}
602: A(r)=
603: \left\{
604: \begin{array}[]{ll}
605: 0,&r<1\\
606: 1,&r\ge 1
607: \end{array}
608: \right.
609: \label{aproxforA}
610: \end{equation}
611: we find the following zero modes
612: \[
613: q_1(r)=
614: \left\{
615: \begin{array}[]{ll}
616: \tilde{C_1}r^2,&r<1\\
617: \frac{C}{\sqrt{g_qr}}e^{-g_qr},&r\ge1 \end{array} \right.  \ \ \ q_2(r)=
618: \left\{
619: \begin{array}[]{ll}
620: \tilde{C_2}r,&r<1\\
621: \frac{C_m}{\sqrt{g_qr}}e^{-g_qr},&r\ge1
622: \end{array}
623: \right.
624: \]
625: \begin{equation}
626: q_3(r)= \left\{
627: \begin{array}[]{ll}
628: \tilde{C_3},&r<1\\
629: \frac{C}{\sqrt{g_qr}}e^{-g_qr}(1+{1\over{g_qr}}),&r\ge1
630: \end{array}
631: \right.
632: \label{aproxsolutiongauge}
633: \end{equation}
634: 
635: The situation changes drastically compared to the global
636: $U_g(1)$ group.  The continuity at $r=1$ results in the following
637: relations between $\tilde{C_p}$ and $C$, $C_m$:
638: \begin{equation}
639: \tilde{C_1}=Cg_q^{-1/2},\qquad \tilde{C_2}=C_mg_q^{-1/2},\qquad
640: \tilde{C_3}=Cg_q^{-3/2}.
641: \label{continuitygauge}
642: \end{equation}
643: Using the normalization condition (\ref{norm}) one obtains that the
644: main contribution to the integral (\ref{norm}) for the second mode
645: comes from $r>1$, and, so, $C_m\sim g_q$.  However, the normalization
646: integral for the first and third modes is saturated now in the core $r<1$
647: by the third mode. Thus, $C\sim g_q^{3/2}$ and
648: \begin{equation}
649: \tilde{C}_p\sim g_q^{\frac{3-p}{2}}.
650: \label{cpong}
651: \end{equation}
652: With the dependence (\ref{cpong}), one finds the hierarchical
653: mass pattern in the case of gauge vortex
654: \begin{equation}
655: m^u_{pp}\sim Y_u(g_u g_q)^{\frac{3-p}{2}}
656: \label{masshier}
657: \end{equation}
658: which  was announced in the Introduction.
659: 
660: In the same way one estimates mixing terms from the Lagrangian (\ref{V3}),
661: (\ref{massintermiss})
662: \begin{equation}
663: m^u_{ps}\sim\delta _{p,s+1}Y_u\epsilon _u\sqrt{g_u}(g_u
664: g_q)^{\frac{3-p}{2}},
665: \label{massmixing}
666: \end{equation}
667: and
668: \begin{equation}
669: m^u\sim
670: Y_u\left(
671: \begin{array}{ccc}
672: g_u g_q&0&0\\
673: \epsilon _ug_u\sqrt{g_q}&\sqrt{g_ug_q}&0\\
674: 0&\epsilon _u\sqrt{g_u}&1
675: \end{array}
676: \right)
677: \label{matrixhier}
678: \end{equation}
679: 
680: 
681: \FIGURE[ht]{
682: %\begin{center}
683:    \leavevmode
684: {\input{picture2.tex}}
685: \caption{
686: The $g$-dependence of the fermion mass pattern
687: $\tilde{m}_{ps}=m_{ps}/g^{(6-p-s)/2}$ calculated in the background
688: shown at Fig.  \ref{fig1}.
689: \label{fig2}
690: }
691: % \end{center}
692: }
693: Of course, the vortex background obtained numerically
694: differs from our crude approximation. In particular, one can see
695: from  Fig. \ref{fig1} that the gauge field $A(r)$ is wider than the
696: scalar $F (r)$, and even far outside the core is almost equal to zero.
697: So, our approximation (\ref{aproxforA}) is not quite appropriate.  One may
698: expect that in the real background (Fig.  \ref{fig1}) the $g$-dependence
699: of the mass pattern  is more resembling with the dependence in the case of
700: the global $U_g(1)$ group (\ref{globalmasspattern}) when $A(r)=0$.
701: However, the numerical calculations show that the mass pattern depends
702: weakly on the width of the gauge field. We present in  Fig.~\ref{fig2}
703: five non-zero elements of the mass matrix
704: (\ref{matrixhier}) divided by the corresponding power of
705: $g$ in the case $g=g_q=g_u$. We see that $m_{ps}/g^{(6-p-s)/2}$  indeed
706: depend weakly on $g$ in the wide interesting range.
707: 
708: 
709: 
710: \section{Results for Standard Model fermions}\label{sec:5}
711: 
712: 
713: Now we are ready to find the set of the parameters of the our model ($Y$,
714: $\epsilon $, $g$) which reproduce the real SM fermionic mass pattern.
715: 
716: To do this, one has to find  eigenvalues of the mass matrix
717: $m^u_{ps}(Y_u,\epsilon_u ,g_q,g_u)$, to equate the result with known
718: fermion masses, and to obtain three equations on four unknown variable,
719: $Y_u$, $\epsilon _u$, $g_q$, $g_u$. However,  even to solve numerically
720: these equations is not an easy problem.  The point is that to find the
721: solution one has to obtain the mass matrix $m_{ps}^u$ in some range of the
722: parameters, which requires in turn the calculation of the zero modes
723: (solving the Eqs.  (\ref{ML523})) at each value of $g$.  Fortunately,
724: there is a simpler way. First of all, note that the mass matrix is
725: proportional to $Y_{u,(d,l)}$, and, therefore, any ratio of the masses does
726: not depend on $Y_{u,(d,l)}$.  Forming two ratios of the masses one obtains two
727: equations on three variables $\epsilon _u$, $g_q$, $g_u$.  The second
728: simplification is to use the established dependence of the mass matrix on
729: $g$ and $\epsilon $ (\ref{masshier}), (\ref{massmixing}). After this the
730: obtained equations can be easily solved.  In fact, in the quark sector
731: the situation is more complicated since besides the mass matrix for
732: up-quarks,  there are also the mass matrix for down quarks and the
733: CKM matrix.  Therefore, we have seven equations (four equations for the
734: mass ratios and three equations for CKM angles) on five parameters ($g_q$,
735: $g_u$, $g_d$, $\epsilon _u$, $\epsilon _d$). In our simulations we have
736: used five of seven equations to reproduce correct mass ratios and one
737: element of the CKM-matrix ($U_{33}^{CKM}$).  As a result we have
738: obtained
739: \begin{equation}
740: g_q = 0.0013\cdot M^{-5},\ \ g_u = 0.0013\cdot M^{-5},\ \ g_d=0.03\cdot
741: M^{-5}, \ \ \epsilon _u=0.9, \ \ \epsilon _d=0.1.
742: \label{g-our}
743: \end{equation}
744: Having at hand these parameters, one can find two other
745: independent elements of the CKM-matrix, but, in general, we could not
746: expect that they  will coincide with the known values.  However, we
747: have obtained the following CKM-matrix
748: \begin{equation}
749: U^{CKM}= \left(
750: \begin{array}{ccc}
751: 0.9780& 0.2084& 0.0100\\
752: 0.2086& 0.9769& 0.0453\\
753: 0.0004& 0.0464& 0.9990
754: \end{array}
755: \right),
756: \label{ckm}
757: \end{equation}
758: and see that this matrix coincides quite well with the measured CKM-matrix
759: in the Standard Model~\cite{PDG}
760: \begin{equation}
761: U_{SM}^{CKM}= \left(
762: \begin{array}{lll}
763: 0.9742 \div 0.9757& 0.219 \div 0.226& 0.002 \div
764: 0.005\\
765: 0.219 \div 0.225& 0.9734 \div 0.9749& 0.037 \div 0.043\\
766: 0.004 \div 0.014& 0.035 \div 0.043& 0.9990 \div 0.9993
767: \end{array}
768: \right).
769: \label{ckm1}
770: \end{equation}
771: 
772: 
773: 
774: In the leptonic sector the correct mass ratios are reproduced at
775: \[
776: g_l= g_e= 0.01,\ \  \epsilon_l = 0.75.
777: \]
778: 
779: To find Yukawa couplings $Y$ we should first of all understand how the
780: obtained  dimensionful parameters (that is all parameters except
781: $\epsilon $) correlate with the absolute energy scale. Let us assume that
782: the first non-zero fermion level is of order 100TeV. It means that
783: $gv^3\simeq 100$TeV. Substituting here $g=0.001$ and $v=1$ one finds our
784: units:
785: \[
786: 1\cdot M=10^5\mbox{TeV}.
787: \]
788: With this relation, one can convert all parameters measured in
789: our units to the absolute mass scale:
790: \[
791: v=\mu =10^{10}{\rm TeV}^2,\ \  v_1=3\cdot 10^{9}{\rm TeV}^2
792: \]
793: \[
794: \lambda =\kappa =\rho =10^{-10}{\rm TeV}^{-2},\ \ h=1.02\cdot 10^{-5}
795: {\rm TeV}^{-1}, \ \ \eta =3.1\cdot 10^{-6}{\rm TeV}^{-1}, \ \
796: e=4\cdot 10^{-7}{\rm TeV}^{-1}
797: \]
798: \[
799: g_q=g_u=1.3\cdot 10^{-28}{\rm TeV}^{-5},\ \ g_d=3\cdot 10^{-27}{\rm
800: TeV}^{-5},\ \ g_l=g_e=1\cdot 10^{-27}{\rm TeV}^{-5}.
801: \]
802: Now taking into account the actual values of the fermion masses one
803: finds the values of the Yukawa couplings $Y$:
804: \[
805: Y_u=3.7\cdot 10^{-17} {\rm TeV}^{-3},\ \ Y_d=2.1\cdot 10^{-20}{\rm
806: TeV}^{-3}, \ \ Y_l=1.8\cdot 10^{-21}{\rm TeV}^{-3}.
807: \]
808: 
809: It is  worth noting that we do not obtain any additional hierarchy: all
810: dimensionless combinations of the our parameters are of order unity. For
811: instance, $g_u^{1/5}/\sqrt{\lambda }=0.4$,  $Y_u^{1/3}/g_u^{1/5}=1.3$,
812: etc.
813: 
814: 
815: \section{Conclusions}
816: In a class of multidimensional models with one vector-like fermionic
817: family, the low-energy effective theory describes three chiral
818: families in four dimensions. Hierarchy of fermionic masses appears due to
819: different profiles of the fermionic wave functions in  extra
820: dimensions.  In this paper we have investigated numerically one of the
821: models of this kind. Namely, we considered the model suggested in
822: ~\cite{LT} and ~\cite{LTF}. We have shown that hierarchical mass pattern
823: indeed appears in the model ~\cite{LTF} with the gauged $U_g(1)$ group
824: which is responsible for the forming of Abrikosov--Nielsen--Olesen vortex,
825: but the dependence of the mass matrix on the parameters is slightly
826: different than it was presented in the Refs.~\cite{LT,LTF}. On the other
827: hand, in the case of global $U_g(1)$ unnatural  mass pattern appears.
828: In the gauge case we found the set of the parameters of the model which
829: reproduces well all known fermion masses and mixings, without hierarchy
830: among parameters.  Moreover, in SM quark sector we have reproduced the
831: nine known values (six masses and three mixing angles)  by the variation
832: of the seven parameters of the model: three couplings $g_{u,d,q}$, two
833: $\epsilon _{u,d}$, and two Yukawa couplings $Y_{u,d}$.
834: 
835: 
836: \acknowledgments
837: We are indebted to V.Rubakov, S.Troitsky, and J.-M. Frere for numerous
838: helpful discussions. We thank F.Bezrukov and A.Kuznetsov for the help in
839: numerical simulations.  This work is supported in part by RFFI grants
840: 99-02-18410a, 01-02-06033, by CRDF award RP1-2103, by the Council of
841: Presidential Grants and State Support of Leading Scientific Schools, grant
842: 00-15-96626, and by the programme SCOPES of the Swiss National Science
843: Foundation, project No. 7SUPJ062239, financed by Federal Department of
844: Foreign affairs.
845: 
846: 
847: \begin{thebibliography}{9}
848: 
849: \bibitem{AD}
850: N.~Arkani-Hamed and M.~Schmaltz,
851: %Phys.\ Rev.\  {\bf D61} (2000) 033005
852: \prd{61}{2000}{033005}
853: [hep-ph/9903417];\\
854: G.~Dvali and M.~Shifman,
855: %Phys.\ Lett.\  {\bf B475} (2000) 295
856: \plb{475}{2000}{295}
857: [hep-ph/0001072];\\
858: T.~Gherghetta and A.~Pomarol,
859: \npb{586}{2000}{141}
860: [hep-ph/0003129];\\
861: D.~E.~Kaplan and T.~M.~Tait,
862: \jhep{06}{2000}{020}
863: [hep-ph/0004200];\\
864: S.~J.~Huber and Q.~Shafi,
865: \plb{498}{2001}{256}
866: [hep-ph/0010195].
867: 
868: \bibitem{LT}
869: M.~V.~Libanov and S.~V.~Troitsky,
870: %Nucl.\ Phys.\ B {\bf 599} (2001) 319
871: \npb{599}{2001}{319}
872: [hep-ph/0011095].
873: 
874: \bibitem{LTF}
875: J.~M.~Frere, M.~V.~Libanov, S.~V.~Troitsky,
876: \plb{512}{2001}{169}
877: %Phys.\ Lett.\ B {\bf 512} (2001) 169
878: [hep-ph/0012306].
879: 
880: \bibitem{Rubakov}
881: V. A. Rubakov, {\it Large and infinite extra dimensions
882: }, hep-ph/0104152.
883: 
884: \bibitem{neutrino}
885: J.~M. Frere, M.~V. Libanov, S.~V. Troitsky.
886: \jhep{11}{2001}{025}
887: %JHEP {\bf 0111} (2001) 025
888: [hep-ph/0110045].
889: 
890: \bibitem{Witten}
891: E.~Witten,
892: \npb{249}{1985}{557}.
893: %Nucl.\ Phys.\  {\bf B249} (1985) 557.
894: 
895: \bibitem{Num}
896: %W.~H. Press, S. A. Teukolsky, W. T. Vetterling,  B. P. Flannery,
897: %\newblock {\it Numerical recipies in C.}
898: %\newblock Cambridge University Press, England (1992).
899: W.~H. Press {\it  et al.},
900: \newblock {\it Numerical recipies in C.}
901: \newblock Cambridge University Press, England (1992).
902: 
903: \bibitem{fermionModes}
904: R.~Jackiw and P.~Rossi,
905: \npb{190}{1981}{681};\\
906: %Nucl.\ Phys.\  {\bf B190} (1981) 681;\\
907: E.~J.~Weinberg,
908: \prd{24}{1981}{2669}.
909: %Phys.\ Rev.\  {\bf D24} (1981) 2669.
910: 
911: \bibitem{PDG} F.~J.~Gilman, K.~Kleinknecht and B.~Renk,
912: \epjc{15}{2000}{110}.
913: %Eur.\ Phys.\ J.\ C {\bf 15} (2000) 110.
914: 
915: \end{thebibliography}
916: 
917: \end{document}
918: