1: \documentclass[12pt]{article}
2:
3: \usepackage{epsfig}
4: \usepackage{pstricks}
5: \usepackage{cite}
6:
7: \def\ve{\varepsilon}
8: \def\vf{\varphi}
9: \def\m{\mu}
10: \def\beq{\begin{equation}}
11: \def\eeq{\end{equation}}
12: \def\b{\beta}
13: \def\beqa{\begin{eqnarray}}
14: \def\eeqa{\end{eqnarray}}
15: \def\D{\Delta}
16: \def\G{\Gamma}
17: \def\l{\lambda}
18: \def\L{\Lambda}
19: \def\n{\nu}
20: \def\p{\pi}
21: \def\k{\kappa}
22: \def\cl{{\cal L}}
23: \def\Bar#1{\overline{#1}} % big bar
24: \def\VEV#1{\left\langle #1\right\rangle} % < >
25: \def\dg{\dagger}
26: \def\wt#1{\widetilde{#1}}
27: \def\NO{\nonumber}
28: \def\mt{\widetilde{m}_1}
29: \def\mb{\overline{m}}
30: \def\pl#1#2#3{Phys.~Lett.~{\bf B{#1}}, #3 ({#2})}
31: \def\np#1#2#3{Nucl.~Phys.~{\bf B{#1}}, #3 ({#2})}
32: \def\prl#1#2#3{Phys.~Rev.~Lett.~{\bf #1}, #3 ({#2})}
33: \def\pr#1#2#3{Phys.~Rev.~{\bf D{#1}}, #3 ({#2})}
34: \def\cqg#1#2#3{Class.~and Quantum Grav.~{\bf {#1}}, #3 ({#2})}
35: \def\cmp#1#2#3{Commun.~Math.~Phys.~{\bf {#1}}, #3 ({#2})}
36: \def\jmp#1#2#3{J.~Math.~Phys.~{\bf {#1}}, #3 ({#2})}
37: \def\ap#1#2#3{Ann.~of Phys.~{\bf {#1}}, #3 ({#2})}
38: \def\prep#1#2#3{Phys.~Rep.~{\bf {#1}C}, #3 ({#2})}
39: \def\ptp#1#2#3{Progr.~Theor.~Phys.~{\bf {#1}}, #3 ({#2})}
40: \def\ijmp#1#2#3{Int.~J.~Mod.~Phys.~{\bf A {#1}}, #3 ({#2})}
41: \def\mpl#1#2#3{Mod.~Phys.~Lett.~{\bf A {#1}}, #3 ({#2})}
42: \def\nc#1#2#3{Nuovo Cim.~{\bf {#1}}, #3 ({#2})}
43: \def\ibid#1#2#3{{\it ibid.}~{\bf {#1}}, #3 ({#2})}
44:
45: \addtolength\textheight{115pt}
46: \addtolength\textwidth{60pt}
47: \addtolength\oddsidemargin{-37pt}
48: \setlength{\parindent}{20pt}
49: \setlength{\parskip}{4pt}
50: \frenchspacing
51: \sloppy
52: \headheight 12pt
53: \headsep 30pt
54: \footskip 24pt
55: \renewcommand{\baselinestretch}{1.2}
56: \addtolength{\topmargin}{-1.5cm}
57:
58:
59: \begin{document}
60: \title{
61: \vspace*{-2cm}
62: {\normalsize
63: \begin{minipage}{3cm}
64: DESY 05-031\\
65: February 2005
66: \end{minipage}}\hspace{\fill}\mbox{}\\[5ex]
67: %
68: {\bf\large LEPTOGENESIS AS THE ORIGIN OF MATTER}
69: }
70: %
71: \author{W. Buchm\"uller$^a$, R. D. Peccei$^b$, T. Yanagida$^c$ \\
72: \vspace{3.0\baselineskip}
73: {\normalsize\it a Deutsches Elektronen-Synchrotron DESY, 22603 Hamburg,
74: Germany}
75: \\
76: \vspace{3.0\baselineskip}
77: {\normalsize\it b Department of Physics and Astronomy, University of California
78: at Los Angeles,}\\ {\normalsize\it Los Angeles, California, 90095, USA}\\
79: \vspace{3.0\baselineskip}
80: {\normalsize\it c Department of Physics, University of Tokyo, Tokyo 113-0033,
81: Japan}\\
82: }
83:
84: \date{}
85: \maketitle
86: %
87: \thispagestyle{empty}
88: %
89: \begin{abstract}
90: \noindent
91: We explore in some detail the hypothesis that the generation of a
92: primordial lepton-antilepton asymmetry (Leptogenesis) early on in the
93: history of the Universe is the root cause for the origin of matter. After
94: explaining the theoretical conditions for producing a matter-antimatter
95: asymmetry in the Universe we detail how, through sphaleron processes, it
96: is possible to transmute a lepton asymmetry -- or, more precisely, a
97: (B-L)-asymmetry -- into a baryon asymmetry. Because Leptogenesis
98: depends in detail on properties of the neutrino spectrum, we review
99: briefly existing experimental information on neutrinos as well as the
100: seesaw mechanism, which offers a theoretical understanding of why
101: neutrinos are so light. The bulk of the review is devoted to a discussion
102: of thermal Leptogenesis and we show that for the neutrino spectrum
103: suggested by oscillation experiments one obtains the observed value for
104: the baryon to photon density ratio in the Universe, independently of any
105: initial boundary conditions. In the latter part of the review we consider
106: how well Leptogenesis fits with particle physics models of dark matter.
107: Although axionic dark matter and Leptogenesis can be very naturally
108: linked, there is a potential clash between Leptogenesis and models of
109: supersymmetric dark matter because the high temperature needed for
110: Leptogenesis leads to an overproduction of gravitinos, which alter the
111: standard predictions of Big Bang Nucleosynthesis. This problem can be
112: resolved, but it constrains the supersymmetric spectrum at low energies
113: and the nature of the lightest supersymmetric particle (LSP). Finally, as an
114: illustration of possible other
115: options for the origin of matter, we discuss the possibility that
116: Leptogenesis may occur as a result of non-thermal processes.
117: \end{abstract}
118:
119: \maketitle
120:
121: \newpage
122: \tableofcontents
123:
124: %\newpage
125:
126: \section{Introduction}
127:
128:
129: Our understanding of the Universe has deepened considerably in the last
130: 25 years, so much so that a standard cosmological model has emerged
131: \cite{SCM}. In this model, after the Big Bang, a period of inflationary
132: expansion \cite{inflation} ensued that effectively set the Universe's
133: curvature to zero. After inflation the Universe's expansion continued,
134: not in an exponential fashion but with the rate of expansion being
135: determined by which component of the Universe's energy density dominated
136: the total energy density.
137:
138:
139: In the present epoch this energy density is dominated by a, so-called,
140: dark energy component whose negative pressure causes the Universe's
141: expansion to accelerate \cite{DE}. Dark energy now accounts for
142: approximately 70\% of the total energy density, with the other 30\% of
143: the remaining energy density of the Universe's being dominated by some
144: kind of non-luminous (dark) matter. In detail, the angular distribution
145: of the temperature fluctuations of the microwave background radiation
146: measured by the
147: Wilkinson Microwave Anisotropy Probe (WMAP)
148: collaboration \cite{WMAP} determines the various
149: components of the ratio of the Universe's energy density now $\rho_o$ to
150: the critical energy density $\rho_c$,
151: $\Omega= \rho_o/\rho_c$.\footnote{The critical
152: density $\rho_c$ is the density that
153: corresponds to a closed Universe now, $ \rho_c= 3H_o^2/8\pi G_N$. Here
154: $H_o$ is the value of the Hubble parameter now. Inflation predicts that $\rho_o=\rho_c$, so that
155: $\Omega=1$. } The results are: $\Omega_{\rm{dark~energy}}= 0.73\pm 0.04$;
156: $\Omega_{\rm{matter}}= 0.27 \pm 0.04$; and $\Omega_{\rm{B}}= 0.044 \pm
157: 0.004$. Here $\Omega_{\rm{B}}$ is the contribution of baryonic matter to
158: $\Omega_{\rm{matter}}$, confirming that about 85 \% of $\Omega_{\rm{matter}}$ is indeed contributed by dark matter. The contribution
159: of neutrinos and photons is at a few per mil, or below, and is negligible.
160:
161: Although the standard cosmological model sketched above provides an
162: accurate description of the present Universe and its evolution, deep
163: questions remain to be answered. What exactly constitutes the dark
164: energy? Is it just a cosmological constant? But if that is so, why is
165: the energy scale associated with the corresponding vacuum energy density
166: [$\rho_{\rm{cc}}=E_o^4; E_o \simeq 2 \times 10^{-3}$ eV] so small?
167: Equally mysterious is the nature of the dark matter, although in this
168: case there are at least some particle physics candidates that may be the
169: source for this component of the Universe's energy density.
170:
171:
172: A further mystery is associated with the observed baryon energy density.
173: This number can be used to infer the ratio of the number density of
174: baryons to photons in the Universe, a quantity that is measured
175: independently from the primordial nucleosynthesis of light elements.
176: The WMAP results \cite{WMAP} are in agreement with the most recent
177: nucleosynthesis analysis of the primordial Deuterium abundance, but there
178: are discrepancies with both the inferred $\rm{^4He}$ and $\rm{^7Li}$
179: values \cite{nucleo}. These latter values, however, may have an
180: underestimated error \cite{OS}. Averaging the WMAP result only with that
181: coming from the primordial abundance of Deuterium gives:
182: \begin{equation}
183: \frac{n_B}{n_{\gamma}}\equiv \eta_B= 6.1 \pm 0.3 \times
184: 10^{-10}.
185: \end{equation}
186: Why does this ratio have this value?
187:
188: In this review, we will principally try to address this last question
189: which, as we shall see, is intimately related to the existence of a
190: primordial matter-antimatter asymmetry. Nevertheless,
191: we shall try, when germane, to connect our discussion with the broader
192: issues of what constitutes dark energy and dark matter.
193:
194: There is good evidence that the Universe is mostly made up of matter,
195: although it is possible that small amounts of antimatter exist
196: \cite{anti}. However, antimatter certainly does not constitute one of the
197: dominant components of the Universe's energy density. Indeed, as Cohen,
198: de Rujula, and Glashow \cite{CDG} have compellingly argued, if there
199: were to exist large areas of antimatter in the Universe they could only
200: be at a cosmic distance scale from us. Thus, along with the question of
201: why $n_B/n_{\gamma}$ has the value given in Eq. (1), there is a parallel
202: question of why the Universe is predominantly composed of baryons rather
203: than antibaryons.
204:
205: %\enlargethispage{0.5cm}
206:
207: In fact, these two questions are interrelated. If the Universe had been
208: matter-antimatter symmetric at temperatures of O(1 GeV), as the
209: Universe cools further and the inverse process $ 2\gamma \to B+ \bar{B}$
210: becomes ineffective because of the Boltzmann factor, the number density
211: of baryons and antibaryons relative to photons would have been reduced
212: dramatically as a result of the annihilation process $ B+ \bar{B} \to
213: 2\gamma $. A straightforward calculation gives, in this case,
214: \cite{ann}:
215: \begin{equation}
216: \frac{n_B}{n_{\gamma}}=\frac{n_{\bar{B}}}{n_{\gamma}} \simeq 10^{-18}.
217: \end{equation}
218: Thus, in a symmetric Universe the question is really why
219: observationally $n_B/n_{\gamma}$ is so large!
220:
221:
222:
223: It is very difficult to imagine processes at temperatures below a GeV
224: that could enhance the ratio of the number density of baryons relative
225: to that of photons much beyond the value this quantity attains when
226: baryon-antibaryon annihilation occurs.\footnote{ An exception is provided by some versions of Affleck-Dine Baryogenesis \cite{AD} where a baryon excess is produced by the decay of a scalar field very late in the history of the Universe, which reheats the Universe to temperatures of the
227: ${\cal O}(100\ {\rm MeV})$.}
228: Thus, because Eq. (2) does not
229: agree with the observed value given in Eq. (1), one is led to the
230: interesting conclusion that a primordial matter-antimatter asymmetry
231: must have existed at temperatures of ${\cal O}(1\ {\rm GeV})$ in the
232: Universe. The
233: observed value for $n_B/n_{\gamma}$ and the lack of antimatter in the
234: Universe are manifestations of this primordial asymmetry. Hence, in
235: reality, the ratio $\eta_B$ is, in effect, a measure of the number density
236: of matter minus that of antimatter relative to the photon number
237: density:
238: \begin{equation}
239: \eta_B=\frac{n_B- n_{\bar{B}}}{n_{\gamma}}= 6.1 \pm 0.3 \times
240: 10^{-10}.
241: \end{equation}
242:
243:
244: It is interesting to consider the physical origins of
245: this primordial matter-antimatter asymmetry. From the seminal work of
246: Sakharov \cite{Sakharov} one knows that, under certain conditions which
247: we will amplify later on in this article, this asymmetry can be
248: generated by physical processes. In this review we will focus on
249: Leptogenesis -- the creation of a primordial lepton-antilepton
250: asymmetry -- as the root source for the observed baryon- antibaryon
251: asymmetry of Eq. (3) \cite{FY}. In our view, Leptogenesis provides the most
252: compelling scenario for generating the observed baryon asymmetry in the
253: Universe. In particular, because Leptogenesis is closely linked with
254: parameters in the neutrino sector that can be eventually determined
255: experimentally, this scenario can be tested and can be either confirmed
256: or ruled out by data.
257:
258: \enlargethispage{0.5cm}
259:
260: The plan of this review is as follows. In Section 2 we discuss the
261: theoretical conditions necessary for producing a primordial matter-
262: antimatter asymmetry in the Universe and explain how, through a
263: mechanism first discussed by Kuzmin, Rubakov, and Shaposhnikov,
264: \cite{KRS}, it is possible to turn a lepton asymmetry into a baryon
265: asymmetry. In Section 3 we review existing experimental information on
266: the neutrino sector, as well as the seesaw mechanism \cite{seesaw} that
267: provides a theoretical framework for understanding this data. Section 4
268: discusses thermal Leptogenesis and contains the main quantitative
269: results of the review. In particular, we show in this Section that the
270: observed value for $\eta_B$ obtained from Leptogenesis significantly
271: constrains low energy neutrino properties, and vice versa. In Section 5
272: we turn to dark matter and discuss how supersymmetric candidates for
273: dark matter are significantly constrained if thermal Leptogenesis is the source
274: of the observed baryon asymmetry in the Universe. The constraint arises
275: from the overproduction of gravitinos. Section 6
276: discusses nonthermal Leptogenesis and other nonthermal processes that
277: can lead to Baryogenesis. Finally, we present our conclusion and summary
278: of results in Section 7.
279:
280:
281: \section{Theoretical Foundations}
282: \vspace{-0.2cm}
283: \subsection{Sakharov's Conditions for Baryogenesis}
284: \vspace{-0.1cm}
285: In 1967, Sakharov \cite{Sakharov} considered the consequences of the hypothesis that the observed
286: expanding Universe originated from a superdense initial state with temperature
287: of order the Planck mass, $T_i \sim M_{\rm P}$. Since he could not imagine
288: how, starting from such an initial state, one could obtain a macroscopic
289: separation of matter and antimatter, he concluded that our Universe contains
290: today only matter. That is, the Universe evolved from an initial state even under
291: charge conjugation to a state odd under charge conjugation today. He then
292: realized that in an expanding Universe a matter-antimatter asymmetry could
293: be generated dynamically, if C, CP, and baryon and lepton number were violated, and these processes were out of thermal equilibrium.
294:
295: Sakharov also described a concrete model for Baryogenesis. He proposed as the
296: origin for
297: the baryon and lepton asymmetry the CP-violating decays of maximons, hypothetical
298: neutral spin zero particles with mass of order the Planck mass. Their
299: existence leads to a departure from thermal equilibrium already at
300: temperatures $T \sim M_{\rm P}$, where a small matter-antimatter asymmetry
301: is then generated. An unavoidable consequence of this model is that protons are unstable and
302: decay. However, the proton lifetime in Sakharov's model turned out to be unobservably long,
303: $\tau_p > 10^{50}\ {\rm years}$.
304:
305: During the past four decades many models of Baryogenesis have been proposed,
306: demonstrating that the conditions Sakharov spelled out to allow Baryogenesis to take place are quite readily satisfied
307: in the Standard Model of particle physics and its extensions. There are,
308: however, significant differences among the various mechanisms suggested for producing the
309: baryon asymmetry. Grand Unified Theories (GUTs) have been of particular importance
310: for the development of realistic models of Baryogenesis \cite{yo78}. These theories
311: provide natural heavy particle candidates, whose decays can be the source of the baryon
312: asymmetry. However, in general, the simplest GUT models based on $SU(5)$ lead
313: to a creation of a (B+L)-asymmetry, with a vanishing asymmetry for B-L. As
314: will be made clear below, a (B+L)-asymmetry generated at the GUT scale
315: eventually gets erased by sphaleron processes. In Leptogenesis, heavy Majorana
316: neutrinos required by the seesaw
317: mechanism \cite{seesaw} serve to trigger Baryogenesis. Because B-L is violated, the erasure present in GUTs is avoided. In principle, Electroweak Baryogenesis \cite{electro}
318: is also an attractive possibility, as the relevant parameters could then be tested in
319: collider experiments. However, in general, the electroweak phase transition is not sufficiently out of equilibrium to generate an asymmetry of the magnitude observed in the Universe. Finally, in supersymmetric theories the baryon and lepton
320: number stored in scalar expectation values can also lead to Baryogenesis, through the, so-called,
321: Affleck-Dine mechanism \cite{AD}, which will be discussed in some detail in Section 6.
322:
323: \enlargethispage{0.7cm}
324:
325: \newpage
326:
327: \subsection{$B+L$ Violation in the Standard Model}
328:
329: Due to the chiral nature of the electroweak interactions, baryon and lepton
330: number are not conserved in the Standard Model \cite{tH}. The divergence of the
331: $B$ and $L$ currents,
332: \begin{eqnarray}
333: J^B_\mu &=& {1\over 3} \sum_{generations}\left(\overline{q_L}\gamma_\mu q_L
334: + \overline{u_R}\gamma_\mu u_R + \overline{d_R}\gamma_\mu d_R\right)\;,\\
335: J^L_\mu &=& \sum_{generations}\left(\overline{l_L}\gamma_\mu l_L
336: + \overline{e_R}\gamma_\mu e_R \right)\;,
337: \end{eqnarray}
338: is given by the triangle anomaly,
339: \begin{eqnarray}
340: \partial^\mu J^B_\mu &=& \partial^\mu J^L_\mu \nonumber\\
341: &=& {N_f\over 32 \pi^2} \left(-g^2 W^I_{\mu\nu}\widetilde{W}^{I\mu\nu}
342: + g'^2 B_{\mu\nu}\widetilde{B}^{\mu\nu}\right)\;.
343: \end{eqnarray}
344: Here $N_f$ is the number of generations, and $W^I_{\mu}$ and $B_{\mu}$ are,
345: respectively, the $SU(2)$ and $U(1)$ gauge fields with gauge couplings $g$
346: and $g'$.
347:
348: As a consequence of the anomaly, the change in baryon and lepton number is
349: related to the change in the topological charge of the gauge field,
350: \begin{eqnarray}
351: B(t_f) - B(t_i) &=& \int_{t_i}^{t_f}dt\int d^3x\partial^\mu J^B_\mu \nonumber\\
352: &=&N_f[ N_{cs}(t_f) - N_{cs}(t_i)]\;,\label{deltaB}
353: \end{eqnarray}
354: where
355: \begin{equation}
356: N_{cs}(t) = {g^3\over 96\pi^2} \int d^3x \epsilon_{ijk}\epsilon^{IJK}
357: W^{Ii}W^{Jj}W^{Kk}\;.
358: \end{equation}
359: For vacuum to vacuum transitions $W^{Ii}$ is a pure gauge configuration
360: and the Chern-Simons numbers $N_{cs}(t_i)$ and $N_{cs}(t_f)$ are integers.
361:
362: In a non-abelian gauge theory there are infinitly many degenerate ground
363: states, which differ in their value of the Chern-Simons number,
364: $\Delta N_{cs} = \pm 1, \pm2, \ldots$. The correponding points in field
365: space are separated by a potential barrier whose height is given by the
366: so-called sphaleron energy $E_{sph}$ \cite{Manton}. Because of the anomaly,
367: jumps in the
368: Chern-Simons number are associated with changes of baryon and lepton number,
369: \begin{equation}
370: \Delta B = \Delta L = N_f \Delta N_{cs}\;.
371: \end{equation}
372: Obviously, in the Standard Model the smallest jump is $\Delta B = \Delta L = \pm 3$.
373:
374: In the semiclassical approximation, the probability of tunneling between neighboring vacua is determined by instanton configurations. In the Standard Model,
375: $SU(2)$ instantons lead to an
376: effective $12$-fermion interaction
377: \begin{equation}\label{obl}
378: O_{B+L} = \prod\limits_{i=1\ldots 3}\left(q_{Li}q_{Li}q_{Li}l_{Li}\right)\;,
379: \end{equation}
380: which describes processes with $\Delta B= \Delta L=3$, such as
381: \begin{equation}
382: u^c + d^c + c^c \rightarrow d + 2 s + 2 b +t + \nu_e + \nu_\mu + \nu_\tau\;.
383: \end{equation}
384: The transition rate is determined by the instanton action and one finds \cite{tH}
385: \begin{eqnarray}
386: \Gamma &\sim& e^{-S_{\rm inst}} = e^{-{4\pi\over \alpha}} \nonumber\\
387: &=& {\cal O}\left(10^{-165}\right)\;.
388: \end{eqnarray}
389: Because this rate is extremely small, ($B+L$)-violating interactions
390: appear to be completely negligible in the Standard Model. However, this picture changes dramatically when one is in a thermal bath.
391:
392: \subsection{Sphalerons and the KRS Mechanism}
393:
394:
395: As emphasized in the seminal paper of Kuzmin, Rubakov and Shaposhnikov
396: \cite{KRS},
397: in the thermal bath provided by the expanding Universe one can make transitions between the gauge vacua not by tunneling, but through thermal fluctuations over the barrier. For temperatures larger than
398: the height of the barrier, the exponential suppresion in the rate provided by the Boltzmann factor disappears completely. Hence (B+L)-violating processes can occur at a significant rate and these processes can be in equilibrium in the expanding Universe.
399:
400: The finite-temperature transition rate in the electroweak theory is
401: determined by the sphaleron configuration \cite{Manton}, a saddle point of the field energy
402: of the gauge-Higgs system. Fluctuations around this saddle point have
403: one negative eigenvalue, which allows one to extract the transition rate.
404: The sphaleron energy is proportional to $v_F(T)$, the finite-temperature
405: expectation value of the Higgs field, and one finds
406: \begin{equation}
407: E_{sph}(T) \simeq {8\pi\over g} v_F(T)\;.
408: \end{equation}
409: Taking translational and rotational zero-modes into account, one obtains
410: for the transition rate per unit volume in the Higgs phase \cite{aml87}
411: \begin{equation}
412: {\Gamma_{B+L}\over V}=\kappa {M_W^7\over (\alpha T)^3} e^{-\beta E_{sph}(T)}\;,
413: \end{equation}
414: where $\beta = 1/T$, $M_W = g^2 v_F(T)/2$ and $\kappa$ is some constant.
415:
416: Extrapolating this semiclassical formula to the high-temperature symmetric
417: phase, where $v_F(T) = 0$, and using for $M_W$ the thermal mass, $M_W\sim g^2 T$,
418: one expects in this phase $\Gamma_{B+L}/V \sim (\alpha T)^4$.
419: However, detailed studies have shown that this naive extrapolation from the Higgs to
420: the symmetric phase is not quite correct. The relevant spatial scale for
421: non-perturbative fluctuations is the magnetic screening length
422: $\sim 1/(g^2 T)$, but the corresponding time
423: scale turns out to be $1/(g^4 T \ln{g^{-1}})$, which is larger for small
424: coupling \cite{yaffe,bodeker}. As a consequence one obtains for the sphaleron rate in the symmetric phase
425: \begin{equation}
426: \Gamma_{B+L}/V \sim \alpha^5 \ln{\alpha^{-1}} T^4.
427: \end{equation}
428:
429: It turns out that the dynamics of low-frequency gauge fields can be described by a remarkably simple
430: effective theory, derived by B\"odeker \cite{bodeker}. The color magnetic and
431: electric fields satisfy the equation of motion
432: \begin{equation}
433: \vec{D}\times\vec{B} = \sigma \vec{E} - \vec{\zeta}\;.
434: \end{equation}
435: Here $\vec{\zeta}$ is Gaussian noise, a random vector field with variance
436: \begin{equation}
437: \langle \zeta_i(x)\zeta_j(x')\rangle = 2\sigma \delta_{ij} \delta^4(x-x')\;.
438: \end{equation}
439: These equations define a stochastic three-dimensional gauge theory. The
440: parameter $\sigma$ is the `color conductivity', $\sigma = m_D^2/(3\gamma)$, where $m_D \sim gT$ is the
441: Debye screening mass and $\gamma \sim g^2T \ln(1/g)$ is the hard gauge boson damping rate. To leading-log accuracy one has
442: $1/\sigma \sim \ln{g^{-1}}$. A next-to-leading order analysis yields for the
443: sphaleron rate \cite{yaffe2}
444: \begin{equation}\label{sphrate}
445: {\Gamma_{B+L}\over V} = (10.8\pm 0.7) \left({gT\over m_D}\right)^2
446: \alpha^5 T^4 \left[\ln{\left({m_D\over \gamma}\right)} + 3.041
447: + \left({1\over \ln{(1/g)}}\right)\right]\;.
448: \end{equation}
449: The overall coefficient has been determined by a numerical lattice
450: simulation \cite{bmr00}. From Eq.~(\ref{sphrate}) one easily obtains the temperature range
451: where sphaleron processes are in thermal equilibrium:
452: \begin{equation}\label{range}
453: T_{EW} \sim 100\ {\rm GeV} < T < T_{sph} \sim 10^{12}\ {\rm GeV}\;.
454: \end{equation}
455:
456: The effective theory describing topological fluctuations of the gauge field
457: in the high-temperature phase is valid for small coupling, $g\ll 1$. Yet for
458: $T_{EW} < T < T_{sph} \sim 10^{12}\ {\rm GeV}$ one has $g={\cal O}(1)$. This implies
459: that the electric screening lenth $1/(gT)$ and the magnetic screening length
460: $1/(g^2 T)$ are not well separated and that nonperturbative corrections
461: to the sphaleron rate, Eq. (\ref{sphrate}), may be large. This will modify the
462: temperature range given in Eq. (\ref{range}), but one expects that the qualitative picture
463: of fluctuations in baryon and lepton number in the high-temperature phase
464: of the Standard Model will not be affected.
465:
466:
467: \subsection{Electroweak Baryogenesis and its Experimental Constraints}
468:
469: An important ingredient in the theory of Baryogenesis is related to the nature
470: of the electroweak transition from the high-temperature symmetric phase to the
471: low-temperature Higgs phase. Because in the Standard Model baryon number, C and
472: CP are not conserved, it is conceivable that the cosmological
473: baryon asymmetry could have been generated at the electroweak phase
474: transition \cite{KRS}, provided that this transition is of first-order,
475: because then there is also the necessary departure from thermal equilibrium.
476: This possibility has stimulated much theoretical
477: activity during the past years to determine the phase diagram of the
478: electroweak theory.
479:
480: Electroweak Baryogenesis requires that the baryon asymmetry
481: generated during the phase transition is not erased by sphaleron processes
482: afterwards.\footnote{ The produced asymmetry will be erased if, after the phase transition, (B+L)-violating processes are in equilibrium.} This leads to a condition on the jump of the Higgs vacuum
483: expectation value $v_F =\sqrt{H^{\dagger}H}$ at the critical
484: temperature \cite{sh86}:
485: \begin{equation}\label{sphbound}
486: {\Delta v_F(T_c)\over T_c} > 1\;.
487: \end{equation}
488: The strength of the electroweak transition has been studied by numerical and
489: analytical methods as function of the Higgs boson mass. For the
490: $SU(2)$ gauge-Higgs model one finds from lattice simulations as well as
491: perturbative calculations that the lower bound of Eq. (\ref{sphbound}) is
492: violated for Higgs masses above 45 GeV \cite{ja96}.\footnote{ For Higgs masses
493: below 50 GeV, the Higgs model provides a good approximation for the full Standard
494: Model.} Because the present lower bound from LEP on the Higgs mass is
495: 114 GeV \cite{PDG}, it is clear that the electroweak transition in the Standard Model is too weak for Baryogenesis. However, for special choices of
496: parameters or by adding singlet fields, in certain circumstances supersymmetric extensions of the Standard
497: Model have a sufficiently strong first-order phase transition to allow
498: Electroweak Baryogenesis to take place \cite{susyew}.
499:
500: For large Higgs masses, the nature of the electroweak transition is dominated
501: by nonperturbative effects of the $SU(2)$ gauge theory at high temperatures.
502: At a critical Higgs mass $m_H^c = {\cal O}(M_W)$, an intriguing phenomenon
503: occurs: The first-order phase transition turns into a smooth
504: crossover \cite{bp95,klrs96,bw97}, as expected on general grounds \cite{ja96}.
505: At the endpoint of a critical line of first-order transitions, which is reached
506: for $m_H=m_H^c$, the phase transition is of second order \cite{rtk98}.
507:
508: The value of the critical Higgs mass can be estimated by comparing the
509: W-boson mass $M_W$ in the Higgs phase with the magnetic mass $m_{SM}$ in the
510: symmetric phase. This yields $m_H^c \simeq 74\ \mbox{GeV}$ \cite{bp97}.
511: Numerical lattice simulations
512: have determined the precise value $m_H^c = 72.1 \pm 1.4$~GeV \cite{fo99}.
513: The analytic estimate of the critical Higgs mass can be generalized to
514: supersymmetric extensions of the Standard Model, where one finds
515: $m_h^c < 130\dots 150$~GeV \cite{ce97}, which is still compatible with
516: the present experimental lower bound.
517:
518: \subsection{The Relation Between Baryon and Lepton Asymmetries}
519:
520: In a weakly coupled plasma, one can assign a chemical potential $\m$ to
521: each of the quark, lepton and Higgs fields. In the Standard Model, with one
522: Higgs doublet $H$ and $N_f$ generations one then has $5N_f+1$ chemical
523: potentials.\footnote{In addition to the Higgs doublet, the two left-handed
524: doublets $q_i$ and $\ell_i$ and the three right-handed singlets $u_i$, $d_i$,
525: and $e_i$ of each generation each have an independent chemical potential.}
526: For a non-interacting gas of massless particles the asymmetry in the particle
527: and antiparticle number densities is given by
528: \begin{equation}\label{number}
529: n_i-\overline{n}_i={g T^3\over 6}
530: \left\{\begin{array}{rl}\b\m_i +{\cal O}\left(\left(\b\m_i\right)^3\right)\;,
531: &\mbox{fermions}\;,\\
532: 2\b\m_i+{\cal O}\left(\left(\b\m_i\right)^3\right)\;, &\mbox{bosons}\;.
533: \end{array}\right.
534: \end{equation}
535: The following analysis is based on these relations for $\b \m_i \ll 1$.
536: However, one should keep in mind that the plasma of the early Universe is
537: very different from a weakly coupled relativistic gas, owing to the presence
538: of unscreened non-abelian gauge interactions, where nonperturbative effects
539: are important in some cases.
540:
541: Quarks, leptons and Higgs bosons interact via
542: Yukawa and gauge couplings and, in addition, via the nonperturbative
543: sphaleron processes. In thermal equilibrium all these processes yield
544: constraints between the various chemical potentials \cite{ht}. The effective
545: interaction
546: of Eq. (\ref{obl}) induced by the $SU(2)$ electroweak instantons implies
547: \begin{equation}\label{sphew}
548: \sum_i\left(3\m_{qi} + \m_{li}\right) = 0\;.
549: \end{equation}
550: One also has to take the $SU(3)$ Quantum Chromodynamics (QCD)
551: instanton processes into account
552: \cite{moh}, which generate an effective interaction
553: between left-handed and right-handed quarks. The corresponding relation
554: between the chemical potentials reads
555: \begin{equation}\label{sphqcd}
556: \sum_i\left(2\m_{qi} - \m_{ui} - \m_{di}\right) = 0\;.
557: \end{equation}
558: A third condition, valid at all temperatures, arises from the
559: requirement that the total hypercharge of the plasma vanishes. From
560: Eq.~(\ref{number}) and the known hypercharges one obtains
561: \begin{equation}\label{hypsm}
562: \sum_i\left(\m_{qi} + 2 \m_{ui} - \m_{di} - \m_{li} - \m_{ei} +
563: {2\over N_f} \m_{H}\right) = 0\;.
564: \end{equation}
565:
566: The Yukawa interactions, supplemented by gauge interactions, yield relations
567: between the chemical potentials of left-handed and right-handed fermions,
568: \begin{equation}\label{myuk}
569: \m_{qi}-\m_{H}-\m_{dj} = 0\;, \quad
570: \m_{qi}+\m_{H}-\m_{uj} = 0\;, \quad
571: \m_{li}-\m_{H}-\m_{ej} = 0\;.
572: \end{equation}
573: These relations hold if the corresponding interactions are in thermal
574: equilibrium. In the temperature range
575: $100\ \mbox{GeV} < T < 10^{12}\ \mbox{GeV}$,
576: which is of interest for Baryogenesis, this is the case for gauge
577: interactions. On the other hand, Yukawa interactions are in equilibrium only
578: in a more restricted temperature range that depends on the strength of the
579: Yukawa couplings. In the following we shall ignore this complication which
580: has only a small effect on our discussion of Leptogenesis.
581:
582: Using Eq.~(\ref{number}), the baryon number density $n_B \equiv g B T^2/6$
583: and the lepton number densities $n_{L_i} \equiv L_i gT^2/6$ can be expressed in
584: terms of the chemical potentials:
585: \begin{eqnarray}
586: B &=& \sum_i \left(2\m_{qi} + \m_{ui} + \m_{di}\right)\;, \\
587: L_i &=& 2\m_{li} + \m_{ei}\;,\quad L=\sum_i L_i\;.
588: \end{eqnarray}
589:
590: Consider now the case where all Yukawa interactions are in equilibrium.
591: The asymmetries $L_i-B/N_f$ are then conserved and we have equilibrium
592: between the different generations, $\m_{li} \equiv \m_l$,
593: $\m_{qi} \equiv \m_q$, etc. Using also the sphaleron relation and the
594: hypercharge constraint, one can express all chemical potentials, and
595: therefore all asymmetries, in terms of a single chemical potential that
596: may be chosen to be $\m_l$,
597: \begin{eqnarray}\label{exam1}
598: \m_e &=& {2N_f+3\over 6N_f+3}\m_l\;, \quad
599: \m_d = -{6N_f+1\over 6N_f+3}\m_l\;, \quad
600: \m_u = {2N_f-1\over 6N_f+3}\m_l\;, \nonumber\\
601: \m_q &=& -{1\over 3} \m_l\;, \quad
602: \m_{H} = {4N_f\over 6N_f+3} \m_l\;.
603: \end{eqnarray}
604: The corresponding baryon and lepton asymmetries are
605: \begin{equation}
606: B = -{4N_f\over 3}\m_l\;, \quad L = {14N_f^2+9N_f\over 6N_f+3}\m_l\;.
607: \end{equation}
608: This yields the important connection between the $B$, $B-L$ and $L$
609: asymmetries \cite{ks}
610: \begin{equation}\label{connection}
611: B = c_s (B-L); ~~L= (c_s - 1)(B- L) \;,
612: \end{equation}
613: where $c_s = (8N_f+4)/(22N_f+13)$. The above relations hold for temperatures
614: $T\gg v_F$. In general, the ratio $B/(B-L)$ is a function of $v_F/T$
615: \cite{ls00}.
616:
617: The relations (\ref{connection}) between B-, (B-L)- and L-number
618: suggest that (B-L)-violation is needed in order to
619: generate a B-asymmetry.\footnote{In the case of Dirac neutrinos,
620: which have extremely small Yukawa couplings, one can construct Leptogenesis
621: models where an asymmetry of lepton doublets is accompanied
622: by an asymmetry of right-handed neutrinos such that the total L-number
623: is conserved and the (B-L)-asymmetry vanishes \cite{dlx}.}
624: Because the (B-L)-current has no anomaly, the value of B-L at time $t_f$,
625: where the Leptogenesis process is completed, determines the value of the
626: baryon asymmetry today,
627: \begin{equation}
628: B(t_0)\ =\ c_s\ (B-L)(t_f)\;.
629: \end{equation}
630: On the other hand, during the Leptogenesis process the strength of (B-L)-,
631: and therefore L-violating interactions can only be weak. Otherwise, because
632: of Eq. (\ref{connection}), they would wash out any baryon asymmetry.
633: As we shall
634: see in the following, the interplay between these conflicting
635: conditions leads to important constraints on the properties of neutrinos.
636:
637:
638: \section{Experimental and Theoretical Information on the Neutrino
639: Sector}
640:
641: \subsection{Results from Oscillation Experiments}
642:
643: The search for neutrino mass has a long history \cite{history}. Positive results
644: are now provided by neutrino oscillation experiments. The allowed values
645: for the mass-squared differences $\Delta m_{ij}^2$ and the mixing angles
646: $\theta_{ij}$ at the $3\sigma$ level for three generations of neutrinos are summarized below \cite{PDG}:
647: %
648: \begin{equation}
649: {\rm sin}^22\theta_{23}=0.92-1,~~ |\Delta
650: m^2_{23}|= (1.2-4.8)\times 10^{-3} {\rm eV}^2 ~~({\rm atmospheric}~ \nu)
651: \end{equation}
652: %
653: \begin{equation}
654: {\rm sin}^22\theta_{12}=0.70-0.95,~~ |\Delta m^2_{12}|=
655: (5.4-9.5)\times 10^{-5} {\rm eV}^2 ~~({\rm solar}~\nu).
656: \end{equation}
657: %
658: The CHOOZ experiment \cite{CHOOZ} gives only an upper limit on the remaining mixing angle $\theta_{13}$:
659: %
660: \begin{equation}
661: {\rm sin}^22\theta_{13}=0-0.17 ~~({\rm CHOOZ} ~\nu).
662: \end{equation}
663: %
664: The LSND experiment \cite{LSND} reports neutrino oscillations from ${\bar \nu}_\mu$
665: to ${\bar \nu}_e$. The mixing angle and mass-squared difference inferred from this experiment are ${\rm
666: sin}^22\theta =0.003-0.03$ and $|\Delta m^2|= 0.2-2~ {\rm eV}^2$. This
667: parameter region is almost excluded by negative results from a
668: comparable experiment by the KARMEN collaboration \cite{KARMEN},
669: but there still
670: remains a narrow region allowed at the $90 \%$ CL.
671:
672: It is very difficult to explain all the above data from neutrino oscillation
673: experiments within the three neutrino framework. Indeed,
674: to accommodate the LSND data one must introduce at least one sterile neutrino
675: \cite {sterile}, or make the radical assumption that CPT is not conserved
676: \cite{CPT}. In this review, for simplicity we will disregard
677: the data from the LSND experiment.
678:
679: \subsection{Information from $\beta$-decay, 2$\beta$-decay and
680: Cosmology}
681:
682: The oscillation experiments discussed in the previous subsection are
683: only sensitive to mass-squared differences. In this subsection we quote
684: results from direct searches for the absolute values of neutrino masses.
685:
686: The direct laboratory limits on the neutrino masses are summarized as
687: follows \cite{PDG}:
688: %
689: \begin{equation}
690: m_{\nu_e} < 2.5 ~{\rm eV} ~;
691: \end{equation}
692: %
693: \begin{equation}
694: m_{\nu_\mu} < 170 ~{\rm keV} ~;
695: \end{equation}
696: \begin{equation}
697: m_{\nu_\tau} < 18 ~{\rm MeV} ~.
698: \end{equation}
699: %
700: The $\nu_e$ mass measurements use the decay of tritium, $^3{\rm H}\rightarrow
701: ^3{\rm He} + {\bar \nu}_e + e^-$, which has a small $Q$ value, $Q=18.6$ keV,
702: and looks at the electron spectrum near the end point in the Kurie plot.
703: The limit on the $\nu_\mu$ mass is obtained from the two-body kinematics
704: of the pion decay, $\pi^+\rightarrow \mu^+ + \nu_\mu$. Finally, the limit on the
705: $\nu_\tau$ mass is obtained from measurements of the invariant mass distribution of
706: $3\pi$ and $5\pi$ systems in the $\tau$ decays, $\tau\rightarrow 3(5)\pi
707: + \nu_\tau$.
708:
709: Neutrinoless double $\beta$ decay experiments provide a bound
710: on an element of the Majorana mass matrix, $m_{\nu_{ee}}$ \cite{Kayser}. The best limit
711: comes from the $^{76}{\rm Ge}$ results. Because the calculation of the double
712: $\beta$ decay rate is model dependent, we quote a range for this bound:
713: %
714: \begin{equation}
715: m_{\nu_{ee}} < 0.3-0.8 ~{\rm eV}~.
716: \end{equation}
717: %
718:
719: One can derive a stringent upper limit on the sum of neutrino masses from
720: cosmology. In the early Universe neutrinos were in thermal
721: equilibrium with radiation and one can infer their number density today
722: to be $n_{i} \simeq 110~{\rm cm}^{-3}$ for each neutrino species. Although the contribution of massless neutrinos to the Universe's energy density is negligible, the contribution of
723: massive neutrinos could be important. By
724: requiring that the energy density of massive neutrinos does not exceed that of dark matter, one
725: obtains the bound:
726: %
727: \begin{equation}
728: \sum_{i}m_{i} <30h^2 ~{\rm eV},
729: \end{equation}
730: %
731: where $h$ is the Hubble constant in units of $100~{\rm km~s}^{-1}{\rm
732: Mpc}^{-1}$
733: [$h=0.71^{+0.04}_{-0.03}$ \cite{WMAP}].
734: %[$h=0.71 \begin{array}{c} +0.04\\-0.03 \end{array}$ \cite{WMAP}].
735:
736: Even if neutrinos satisfy the above limit, massive neutrinos would
737: affect the formation of cosmic structure, because the free streaming of
738: neutrinos suppresses density fluctuations at small scales.
739: The normalization of large- and small-scale fluctuations constrains the
740: contribution from neutrinos. Recent detailed analyses \cite{HR}
741: lead to the bound:
742: %
743: \begin{equation}
744: \sum_{i}m_{i} <0.65 ~{\rm eV}.
745: \end{equation}
746: %
747: It is interesting that a similar constraint, $\sum_{i} m_{i} <2.0 ~{\rm
748: eV}$, has been obtained by using the cosmic microwave background data alone
749: \cite{fukugita}.
750:
751: \subsection{The Seesaw Mechanism}
752:
753: If there are right-handed neutrinos, then neutrinos can have a Dirac mass much as quarks and charged leptons do. However, if this is the only source for their mass, it is not easy to find a natural
754: reason for the very small mass of neutrinos. With only a Dirac mass term, the smallness of
755: neutrino masses needs to be ascribed to having very tiny Yukawa coupling constants $h$ ($h \sim
756: 10^{-13}$ gives a neutrino mass $m_{\nu}\simeq 0.01$ eV). Although it is possible to imagine mechanisms that result in very small Dirac masses for
757: neutrinos,\footnote{ At the end of this subsection
758: we outline a recent approach that may explain how such extremely small Yukawa coupling constants might arise in a higher dimensional theory.}
759: the seesaw mechanism, which entails Majorana masses,
760: provides a natural explanation for the smallness of neutrino masses in
761: theories of unification at ultra-high energy scales \cite{seesaw}.
762:
763: To appreciate this point, we should first note that the Standard Model does not require neutrinos to be massive, because right-handed neutrinos
764: are not necessary for the electroweak theory. Without right-handed neutrinos,
765: these particles may acquire their mass only from what are called irrelevant
766: operators -- operators such as
767: $\ell\ell HH$ with dimension greater than four. These operators can give rise to a Majorana mass for neutrinos in theories with a cutoff. However, in the
768: limit of an infinite cutoff, neutrino masses vanish. It should be noted that if
769: one adopts the Planck scale as the Wilson cutoff for the
770: Standard Model, one finds neutrino masses to be at most $10^{-5}$ eV.
771: Thus, the observed mass $\sqrt{\Delta m^2_{\rm atm}}\simeq 0.05$ eV is
772: unable to be explained within the Standard Model.
773:
774: If one considers possible extensions of the Standard Model gauge group,
775: it is natural to consider a gauge group $G$ whose rank is at least 5,
776: since the rank of the Standard Model group,
777: $SU(3)\times SU(2)\times U(1)_{Y}$, is 4. This new group $G$ may
778: contain an extra $U(1)$ as a subgroup, in addition to the Standard Model
779: gauge group. The simplest candidate for the
780: extra $U(1)$ is a B-L gauge symmetry, because we know that the global
781: $U(1)_{B-L}$ does not have any anomaly due to the Standard Model gauge
782: interactions. However, when one gauges this B-L symmetry,
783: the theory does have
784: a self-anomaly of the B-L\\ interactions. That is, the triangle anomaly
785: of $[U(1)_{B-L}]^3$ is non-vanishing. A crucial point, however, is that this
786: anomaly is cancelled by introducing right-handed neutrinos. This also
787: cancels the mixed gravitational/B-L anomaly. Thus,
788: right-handed neutrinos are required for consistency of the theory! A famous
789: example which includes $U(1)_{B-L}$ is provided by $SO(10)$ grand unification.
790: But our argument is more general. For instance, the string brane world
791: predicts many $U(1)$'s and it is quite natural to consider some of them
792: to be anomaly free and survive as low-energy (compared with the string
793: scale) gauge symmetries.
794:
795:
796: It is usually assumed that the unification group $G$ is broken
797: down to the Standard Model group at high energies. Then, the right-handed
798: neutrinos naturally obtain large Majorana masses, because they are singlets
799: of the Standard Model and there is no unbroken symmetry to
800: protect them from acquiring a large Majorana mass. In this article we shall denote heavy
801: Majorana neutrinos as $N$.
802: Then, the masses of the neutrinos written as a matrix take a simple form:
803: %
804: \begin{equation}
805: \left(
806: \begin{array}{cc}
807: 0&m \\
808: m^T&M
809: \end{array}
810: \right)\;.
811: \end{equation}
812: %
813: Here $m$ is the Dirac mass matrix between the left-handed neutrinos $\nu$ and
814: the heavy Majorana neutrinos $N$,
815: which is of order the electroweak scale, while $M$ is the Majorana mass matrix of the heavy neutrinos $N$. For three generations, both $M$ and $m$ are $3 \times 3$ matrices.
816: Integrating out the heavy Majorana neutrinos $N$ leads to a small neutrino
817: mass via the seesaw mechanism \cite{seesaw}. For one neutrino generation one simply has that:
818: %
819: \begin{equation}
820: m_\nu \simeq \frac{m^2}{M}.
821: \end{equation}
822: %
823: We see from the above that a small neutrino mass is a reflection of the ultra-heavy mass
824: of the heavy neutrino $N$. The observed neutrino mass
825: $\sqrt{\Delta m^2_{\rm atm}}\simeq 0.05$ eV implies $M\simeq 10^{15}$ GeV,
826: which is very close to the Grand Unification scale. Thus, effectively, the small neutrino masses provide a window to new physics at
827: an ultra-high energy scale.
828:
829: One may question why the unification group $G$, or the $U(1)_{B-L}$ symmetry,
830: should be broken at such a high energy scale. Perhaps the answer to this
831: question, as we shall amplify in this review, is because, otherwise, the baryon number in the Universe would be too small
832: for us to exist! It turns out that if the
833: Universe's baryon number were to be two orders of
834: magnitude below the present observed value galaxies would not be formed
835: \cite{TR97}. One of the purposes
836: of this review article is to explain in some detail this fundamental
837: point.\\
838:
839: \noindent
840: {\bf 3.3.1 Small Neutrino Yukawa Couplings from Higher Dimensional Theories}\\
841: \noindent
842: To explain how one can generate small Dirac masses for neutrinos, consider a theory described in (4+1)-dimensional space-time,
843: while our world is on a (3+1)-dimensional hyperplane -- a so-called D3 brane.
844: The Einstein action of gravity in five-dimensional space-time is given
845: by
846: %
847: \begin{equation}
848: S= \frac{M_*^3}{16\pi}\int d^4x\int dy\sqrt{-g_5}{\cal R}_5,
849: \end{equation}
850: %
851: where $M_*$ is the gravitational scale in five-dimensional space-time,
852: and $g_5$ and ${\cal R}_5$ are the metric and the scalar curvature,
853: respectively. We assume that the fifth dimension is compactified to a
854: space of radius $L$, and consider the metric to be
855: %
856: \begin{equation}
857: d^2s = g_{\mu\nu}dx^\mu dx^\nu -dy^2,
858: \end{equation}
859: %
860: where $g_{\mu\nu}$ is the metric in four-dimensional space-time. The
861: integration over $dy$ leads to the four-dimensional action
862: %
863: \begin{equation}
864: S_4 = \frac{M_*^3L}{16\pi}\int d^4x\sqrt{-g_4}{\cal R}_4.
865: \end{equation}
866: %
867: Then, the Planck scale, $M_{\rm P}\simeq 1.2\times 10^{19}$ GeV, in
868: four-dimensional space-time is given by
869: %
870: \begin{equation}
871: M_{\rm P}^2 = M_*^3L.
872: \end{equation}
873: %
874: One can get the observed value for $M_{\rm P}$ even for $M_* = 1$ TeV
875: by taking a very large $L$.
876: The weakness of gravity in these theories is the result of having a large compactification scale in
877: the fifth dimension \cite{arkani}.
878:
879: Let us now assume that all the Standard Model particles reside on a D3 brane
880: at the boundary $y=0$ and that the right-handed neutrino lives in the
881: five-dimensional bulk \cite{arkani2}. Then the action involving the
882: right-handed neutrinos is given by
883: %
884: \begin{equation}
885: S= M_*\int d^4x\int^L_0 dy\sqrt{-g_5}{\bar N_R} i \partial\llap{/} N_R
886: + \int d^4x\int^L_0 dy\sqrt{-g_5}h{\bar N_R}\ell_LH\delta (y) +
887: {\rm h.c.}\,,
888: \end{equation}
889: %
890: Here, $\ell_L$ and $H$ are SU(2)$_L$ doublets of the left-handed leptons
891: and of the Higgs boson, respectively. One obtains the action in four
892: dimensions by integrating over $dy$, and one finds:
893: %
894: \begin{equation}
895: S_4 = M_*L\int d^4x\sqrt{-g_4}{\bar N_R} i \partial\llap{/} N_R
896: + \int d^4x\sqrt{-g_4}h{\bar N_R}\ell_LH + {\rm h.c.}\,.
897: \end{equation}
898: %
899: After renormalizing the wave function of $N_R$ so that it has a
900: canonical kinetic term, one finds that the Yukawa coupling constant is
901: suppressed by $1/\sqrt{M_*L} = M_*/M_{\rm P}$. Thus in this model one obtains an effective
902: Yukawa coupling constant $h_{\rm eff} \simeq 10^{-13}$ (corresponding to a neutrino
903: Dirac mass $m_{\nu}\simeq 0.01$ eV ) for $h=1$ and $M_* \simeq 10^3$ TeV.
904:
905: This model, however, has a serious drawback. There is no symmetry to
906: protect the right-handed neutrinos from acquiring a Majorana mass, because
907: they are singlets of the Standard Model gauge group. A solution to this
908: problem may be found by imposing a $U(1)_{B-L}$ gauge symmetry in
909: the five-dimensional bulk. As long as the B-L gauge symmetry is exact,
910: the right-handed neutrinos cannot have a Majorana mass because
911: this mass carries a non-vanishing B-L charge. If this symmetry is exact, the corresponding gauge boson is completely
912: massless. However, this may not cause any phenomenological difficulties at
913: low-energies, because the B-L gauge coupling constant must also be extremely
914: suppressed.\footnote{The B-L gauge coupling constant $\alpha_{B-L}$
915: is constrained to be $\alpha_{B-L} < 10^{-21}\alpha_{\rm em}$. This bound
916: comes from the empirical limits of the electromagnetic charges for the
917: neutron and the neutrino. That is, $Q_n = (-0.4 \pm 1.1)\times 10^{-21}$
918: \cite{chargenu} and $Q_\nu = (0.5\pm2.9)\times 10^{-21}$ \cite{chargeneu}.
919: The present model suggests $\alpha_{B-L}\simeq (\frac{M_*}{M_P})^2\times\alpha_{\rm
920: em}\simeq 10^{-25}\times\alpha_{\rm
921: em}$, which may be in an interesting region for future
922: experiments.} But, as we explained in Section 2, if the B-L gauge symmetry is
923: exact, it is very
924: difficult to account for the baryon-number asymmetry in the present
925: Universe.
926:
927: \subsection{CP-Violating Phases at Low and High Energies in the \\ Lepton
928: Sector}
929:
930: In the Standard Model, the Lagrangian for the lepton sector, augmented by including right handed neutrinos, is given by
931: %
932: \begin{eqnarray} \label{llag}
933: {\cal L} &=& {\bar \ell}_{Li} i \partial\llap{/}\ell_{Li}
934: + {\bar e}_{Ri} i \partial\llap{/}e_{Ri}
935: + {\bar N}_{Ri} i \partial\llap{/}N_{Ri} \nonumber\\
936: && + f_{ij}{\bar e}_{Ri}\ell_{Lj}H^{\dagger}
937: +h_{ij}{\bar N}_{Ri}\ell_{Lj}H - \frac{1}{2}M_{ij}N_{Ri}N_{Rj} + {\rm h.c.}\;,
938: \end{eqnarray}
939: %
940: where $i,j = \{1-3\}$ are the family-number indices. We adopt, without losing
941: generality, a basis where the matrices $f_{ij}$ and $M_{ij}$ are diagonal.
942: The Yukawa matrix $h_{ij}$ in this basis is in general complex and thus has CP-violating phases. Because for three families, the matrix
943: $h_{ij}$ has 9 complex parameters, we have 9 possible CP-violating phases. However, three of
944: these phases can be absorbed into the wave function of $\ell_L$ and hence 6
945: CP-violating phases remain physically relevant. These are known as
946: high-energy phases, because they enter in the full theory.\footnote{In
947: particular, as we will show in the next Section, the phase contributing to the generation
948: of the lepton asymmetry in the decay of $N_1$ is a combination of these
949: high-energy phases, given by $\sum{\rm Im}[(hh^{\dagger})_{1i}]^2$.}
950:
951:
952: Let us now discuss the CP-violating phases at low energies. To do that we first
953: need to integrate out the heavy Majorana neutrinos, $N_i$. Doing so the effective
954: Lagrangian for the lepton sector reduces to:
955: %
956: \begin{equation}
957: {\cal L}_{\rm eff} = {\bar \ell}_{Li} i \partial\llap{/}\ell_{Li} +
958: {\bar e}_{Ri} i \partial\llap{/}e_{Ri}
959: + f_{ii}{\bar e}_{Ri}\ell_{Li}H^{\dagger}
960: + \frac{1}{2}\sum_{k}h^T_{ik}h_{kj}\ell_{Li}\ell_{Lj}\frac{H^2}{M_k} +
961: {\rm h.c.}\,.
962: \end{equation}
963: %
964: The last term can be rewritten as
965: %
966: \begin{equation}
967: - \frac{1}{2}m_{\nu_{ij}}\ell_{Li}\ell_{Lj}\frac{H^2}{\langle H\rangle^2}\,,
968: \end{equation}
969: %
970: so that all low-energy phases appear in the mass matrix of light neutrinos.
971: Because the neutrino mass matrix is symmetric, it has 6 complex parameters, and hence one has 6 possible CP-violating phases. However, as before,
972: 3 of these 6 phases can be absorbed into the wave function of
973: $\ell_L$. Therefore, there remain only three physical low energy CP-violating
974: phases \cite{petcov}. Because they are different in number,
975: it is unfortunately very difficult to establish a
976: direct link between the low-energy and the high-energy CP-violating
977: phases \cite{BR}.
978:
979: Furthermore, in practice, it is not possible to measure all three low-energy phases.
980: One of these three phases can be measured by neutrino oscillation experiments, while neutrinoless double $\beta$ decay, if it were to be observed, would provide information on another
981: phase. However, the remaining phase is undetermined. In other
982: words, one cannot perform a complete experiment to determine the
983: neutrino mass matrix. Nevertheless, if the neutrino mass matrix were to have an extra
984: constraint, one may be able to determine all matrix elements of $m_\nu$. This
985: constraint must be independent of the frame of the family basis.
986: One example of such a constraint is the requirement that ${\rm det}(m_{\nu})=0$. In this case, we have only
987: 7 independent physical parameters including the phases in the neutrino
988: mass matrix \cite{branco}, which can be determined in principle in future
989: experiments.\footnote{As will be shown in Section 6, Affleck-Dine Leptogenesis
990: suggests a constraint, $m_{\nu_1}\simeq 10^{-9}$ eV and hence ${\rm
991: det}(m_\nu)\simeq 0$.}
992:
993: \section{Thermal Leptogenesis}
994:
995: \subsection{Lepton Number Violation and Leptogenesis}
996:
997: As we discussed above, lepton number violation is most simply realized by adding right-handed
998: neutrinos to the Standard Model. Their existence is predicted by all
999: extensions of the Standard Model containing B-L as a local symmetry
1000: and allows for an elegant explanation of the smallness of the light neutrino
1001: masses via the seesaw mechanism \cite{seesaw}.
1002:
1003: The most general Lagrangian for couplings and masses of charged leptons and
1004: neutrinos is given in Eq. (\ref{llag}).
1005: The vacuum expectation value of the Higgs field, $\VEV{H}=v_F$,
1006: generates Dirac masses $m_e$ and $m_D$ for charged leptons and neutrinos,
1007: $m_e=f v_F$ and $m_D=hv_F$, which are assumed to be much
1008: smaller than the Majorana masses $M$. This yields the light and heavy neutrino
1009: mass eigenstates
1010: \begin{equation}
1011: \n\simeq V_{\nu}^T\n_L+\n_L^c V_{\nu}^*\quad,\qquad
1012: N\simeq N_R+N_R^c\, ,
1013: \end{equation}
1014: with masses
1015: \begin{equation}
1016: m_{\n}\simeq- V_{\nu}^Tm_D^T{1\over M}m_D V_{\nu}\,
1017: \quad,\quad m_N\simeq M\, .
1018: \label{seesaw}
1019: \end{equation}
1020: In a basis where the charged lepton mass matrix $m_e$ and the Majorana
1021: mass matrix $M$ are diagonal,
1022: $V_{\nu}$ is the mixing matrix in the leptonic charged current.
1023:
1024: The right-handed neutrinos can efficiently erase any pre-existing lepton asymmetry
1025: at temperatures $T > M$, but they can also generate a lepton asymmetry
1026: by means of their out-of-equilibrium decays at temperatures $T < M$.
1027: This asymmetry is then partially transformed into a baryon asymmetry by sphaleron processes. This is the Leptogenesis mechanism proposed by
1028: Fukugita and Yanagida \cite{FY}.
1029:
1030:
1031: The decay width of the heavy neutrino $N_i$ at tree level reads,
1032: \begin{equation}
1033: \G_{Di}=\G\left(N_i\to H + \ell_L\right)
1034: +\G\left(N_i\to H^{\dagger} + \ell_L^{\dagger}\right)
1035: ={1\over8\p}(h h^\dg)_{ii} M_i\;.
1036: \label{decay}
1037: \end{equation}
1038: Once the temperature of the universe drops below the mass $M_1$, the heavy neutrinos are not able to follow the rapid change of the equilibrium
1039: distribution. Hence, the necessary deviation from thermal equilibrium ensues as a result of having a too large number density of heavy neutrinos, compared to the
1040: equilibrium density.
1041: Eventually, however, the heavy neutrinos decay, and a lepton asymmetry is
1042: generated owing to the presence of CP-violating processes. The CP asymmetry involves the interference
1043: between the tree-level amplitude and the one-loop vertex and self-energy
1044: contributions (see Fig.~(1)).
1045: In a basis, where the right-handed neutrino mass matrix $M$
1046: is diagonal, one obtains \cite{l-asymmetry} for the CP asymmetry parameter
1047: $\ve_1$ assuming hierarchical heavy neutrino masses ($M_1 \ll M_2, M_3$):
1048: \begin{equation}\label{eps}
1049: \ve_1 \simeq {3\over16\pi}\;{1\over\left(h h^\dg\right)_{11}}
1050: \sum_{i=2,3}\mbox{Im}\left[\left(h h^\dg\right)_{i1}^2\right]
1051: {M_1\over M_i}\; .
1052: \end{equation}
1053: In the case of mass differences of order the decay widths, one obtains a
1054: significant enhancement from the self-energy contribution \cite{pil99},
1055: although the influence of the thermal bath on this effect is presently unclear.
1056:
1057: The CP asymmetry of Eq. (\ref{eps}) can be obtained in a very simple way by first
1058: integrating out the heavier neutrinos $N_2$ and $N_3$ in the leptonic Lagrangian. This yields
1059: \begin{equation}\label{yuk2}
1060: \cl_{\n}^{eff} = h_{ 1j}\Bar{N_R}_1 \ell_{Lj} H
1061: -{1\over2}M_1 \Bar{N^c_R}_1 N_{R1}
1062: +{1\over 2} \eta_{ij} \ell_{Li} H \ell_{Lj} H + \mbox{ h.c.}\;,
1063: \end{equation}
1064: with
1065: \begin{equation}
1066: \eta_{ij} = \sum_{k=2}^3 h^T_{ ik}{1\over M_k} h_{ kj}\;.
1067: \end{equation}
1068: The asymmetry $\ve_1$ is then obtained from the interference of the
1069: Born graph and the one-loop graph involving the cubic and the quartic
1070: couplings.
1071: %
1072: \begin{figure}[t]
1073: \centerline{\input{DecayH.tex}}
1074: \caption{Tree level and one-loop diagrams contributing to heavy neutrino
1075: decays whose interference leads to Leptogenesis.}
1076: \end{figure}
1077: %
1078: %\begin{figure}[t]
1079: %\centerline{\psfig{file=DecayH.ps,height=4cm,width=11cm}}
1080: %\caption{Tree level and one-loop diagrams contributing to heavy neutrino
1081: %decays whose interference leads to Leptogenesis.}
1082: %\end{figure}
1083: %
1084: This includes automatically both, vertex and self-energy
1085: corrections \cite{bf00} and yields an expression for $\ve_1$ directly
1086: in terms of the light neutrino mass matrix:
1087: \begin{equation}\label{CPas}
1088: \ve_1 \simeq -{3\over 16\p}{M_1\over (h h^\dg)_{11} v_F^2}
1089: {\rm Im}\left(h^* m_\n h^\dg\right)_{11}\;.
1090: \end{equation}
1091:
1092: The CP asymmetry then leads to a (B-L)-asymmetry \cite{FY},
1093: \begin{equation}\label{basym}
1094: Y_{B-L}\simeq -Y_L = - \frac{n_L-n_{\Bar{L}}}{ s}\ = -\k\frac{\ve_1}{g_*}\;.
1095: \end{equation}
1096: Here $s$ is the entropy and, in the present epoch $s=7.04 n_\gamma$, whereas $g_*\sim 100$ is the number of degrees of freedom in the plasma.
1097: The factor $\k<1$ in the above takes into account the effect of washout processes. As we shall discuss below, in order to
1098: determine $\k$ one has to solve the Boltzmann equations.
1099:
1100: Early studies of Leptogenesis were partly motivated by trying to find alternatives to Electroweak Baryogenesis, which did not seem to produce a big enough asymmetry. Some extensions of the Standard Model were considered and, in particular, in the
1101: simple case of hierarchical heavy neutrino masses the observed value of the
1102: baryon asymmetry is naturally obtained with B-L broken at the unification
1103: scale, $M_{GUT} \sim 10^{15}$~GeV. The corresponding light neutrino masses
1104: are then very small, $m_{1,2} < m_{3} \sim 0.1$~eV, and the typical
1105: parameters for the necessary
1106: CP asymmetry and the Baryogenesis temperature are $\ve_1 \simeq 10^{-6}$ and
1107: $T_B \sim M_1 \sim 10^{10}\ {\rm GeV}$, respectively
1108: \cite{bp96,by99}.\footnote{For early work based on $SO(10)$, see \cite{Ghe}.}
1109: Subsequently, researchers realized that
1110: such small neutrino masses are consistent with the small mass differences
1111: inferred from the solar and atmospheric neutrino oscillations.
1112: This fact has given rise to a strong interest in Leptogenesis in recent years, and
1113: a large number of interesting models have been suggested \cite{models}.
1114:
1115: \subsection{Departure from Thermal Equilibrium}
1116:
1117: Leptogenesis takes place at temperatures $T \sim M_1$. For a decay width small
1118: compared to the Hubble parameter, $\G_1(T) < H(T)$, heavy neutrinos are out
1119: of thermal equilibrium, otherwise they are in thermal equilibrium \cite{KT}.
1120: The
1121: borderline between the two regimes is given by $\G_1 = H|_{T=M_1}$, which
1122: is equivalent to the condition that the effective neutrino mass
1123: \beqa\label{effm}
1124: \mt = {(m_D m_D^\dg)_{11}\over M_1}
1125: \eeqa
1126: is equal to the `equilibrium neutrino mass'
1127: \beq\label{mequ}
1128: m_* = {16\p^{5/2}\over 3\sqrt{5}} g_*^{1/2} {v_F^2\over M_{\rm P}}
1129: \simeq 10^{-3}~\mbox{eV}\;.
1130: \eeq
1131: Here we have used the Hubble parameter $H(T)\simeq 1.66\,g_*\,T^2/M_{\rm P}$
1132: where $g_*=g_{SM}=106.75$ is the total number of degrees of freedom and
1133: $M_{\rm P}=1.22\times10^{19}\,{\rm GeV}$ is the Planck mass.
1134:
1135: It is quite remarkable that the equilibrium neutrino mass $m_*$ is close to the
1136: neutrino masses suggested by neutrino oscillations,
1137: $\sqrt{\D m^2_{\rm sol}} \simeq 8\times 10^{-3}$~eV and
1138: $\sqrt{\D m^2_{\rm atm}} \simeq 5\times 10^{-2}$~eV.
1139: This encourages one to think that it may be possible to understand the cosmological baryon
1140: asymmetry via Leptogenesis as a process close to thermal equilibrium. Ideally,
1141: $\D L=1$ and $\D L=2$ processes should be strong enough at temperatures above
1142: $M_1$ to keep the heavy neutrinos in thermal equilibrium and weak enough to
1143: allow the generation of an asymmetry at temperatures below $M_1$.
1144:
1145: In general, the generated baryon asymmetry is the result of a competition
1146: between production processes and washout processes that tend to erase any
1147: generated asymmetry. Unless the heavy Majorana neutrinos are partially
1148: degenerate, $M_{2,3}-M_1 \ll M_1$, the dominant processes are decays and
1149: inverse decays of $N_1$ and the usual off-shell $\D L=1$ and $\D L=2$
1150: scatterings \cite{lut,plu}.
1151:
1152: \begin{figure}[t]
1153: \centerline{\psfig{file=BASYM.eps,height=6.5cm,width=11cm}}
1154: \caption{The evolution of the $N_1$ abundance and the $B-L$ asymmetry
1155: for a typical choice of parameters, $M_1=10^{10}\,$GeV,
1156: $\varepsilon_1=10^{-6}$, $\widetilde{m}_1=10^{-3}\,$eV
1157: and $\overline{m}=0.05\,$eV. From \cite{bdp02}.}
1158: \label{BASYM}
1159: \end{figure}
1160:
1161: The Boltzmann equations for Leptogenesis are\footnote{We use the
1162: conventions of \cite{bdp02}. We have also summed over the three lepton
1163: flavours neglecting the dependence on the lepton Yukawa couplings
1164: \cite{bcx00}.}
1165: \beqa
1166: {dN_{N_1}\over dz} & = & -(D+S)\,(N_{N_1}-N_{N_1}^{\rm eq}) \;, \label{lg1} \\
1167: {dN_{B-L}\over dz} & = & -\ve_1\,D\,(N_{N_1}-N_{N_1}^{\rm eq})-W\,N_{B-L} \;,
1168: \label{lg2}
1169: \eeqa
1170: where $z=M_1/T$. The number density $N_{N_1}$ and the amount of $B-L$
1171: asymmetry, $N_{B-L}$, are calculated in a portion of comoving volume that
1172: contains one photon at the onset of Leptogenesis, so that the relativistic
1173: equilibrium $N_1$ number density is given by $N_{N_1}^{\rm eq}(z\ll 1)=3/4$. Alternatively, one may normalize the number density to the entropy density
1174: $s$ and consider $Y_X = n_X/s$. If entropy is conserved, both normalizations are related by a constant.
1175:
1176: There are four classes of processes that contribute to the different terms in
1177: the above equations: decays, inverse decays, $\D L=1$ scatterings and $\D L=2$
1178: processes mediated by heavy neutrinos. The first three processes all modify the
1179: $N_1$ abundance and try to push it towards its equilibrium value
1180: $N_{N_1}^{\rm eq}$. Denoting by $H$ the Hubble expansion rate, the term
1181: $D = \Gamma_D/(H\,z)$ accounts for decays and inverse decays, whereas the
1182: scattering term $S = \Gamma_S/(H\,z)$ represents the $\D L=1$ scatterings.
1183: Decays also yield the source term for the generation of the $B-L$ asymmetry,
1184: the first term in Eq.~(\ref{lg2}), whereas all other processes
1185: contribute to the total washout term $W = \Gamma_W/(H\,z)$ which competes
1186: with the decay source term. The dynamical generation of the $N_1$ abundance
1187: and the $B-L$ asymmetry is shown in Fig. (\ref{BASYM}) for typical parameters.
1188:
1189:
1190: \subsection{Decays and Inverse Decays}
1191:
1192: It is very instructive to consider first a simplified picture in which decays
1193: and inverse decays are the only processes that are effective.\footnote{This section follows
1194: closely \cite{bdp04}.} For consistency, in this approximation the real
1195: intermediate state contribution to the $2\rightarrow 2$ processes has to be
1196: included. In the kinetic equations (\ref{lg1}) and (\ref{lg2}) one then has
1197: to replace $D+S$ by $D$ and $W$ by $W_{I\!D}$, respectively, where $W_{I\!D}$
1198: is the contribution of inverse decays to the washout term.
1199: The solution for $N_{B-L}$ in this case is the sum of two terms \cite{KT},
1200: \begin{equation}\label{solution}
1201: N_{B-L}(z)=N_{B-L}^{\rm i}\,e^{-\int_{z_{\rm i}}^{z}\,dz'\,W_{I\!D}(z')}
1202: -{3\over 4}\,\ve_1\,\k(z;\mt)\;.
1203: \end{equation}
1204: Here the first term accounts for an initial asymmetry which is partly reduced
1205: by washout, and the second term describes $B-L$ production from $N_1$ decays.
1206: It is expressed in terms of the {\em efficiency factor} $\k$ \cite{bcx00}
1207: which does not depend on the CP asymmetry $\ve_1$,
1208: \begin{equation}\label{lg3}
1209: \k(z)={4\over 3} \int_{z_{\rm i}}^{z}\,dz'\,
1210: D \left(N_{N_1}-N_{N_1}^{\rm eq}\right)\,
1211: e^{-\int_{z'}^{z}\,dz''\,W_{I\!D}(z'')}\;.
1212: \end{equation}
1213: As we shall see, decays and inverse decays are sufficient to describe qualitatively many properties of the full problem.
1214:
1215: We will first study in
1216: detail the regimes of weak and strong washout. If just decays and inverse
1217: decays are taken into account, these regimes correspond, respectively, to the limits $K\ll 1$ and $K\gg 1$ of the decay parameter
1218: \begin{equation}\label{decpar}
1219: K = {\G_D(z=\infty)\over H(z=1)} = {\widetilde{m}_1\over m_*} \;,
1220: \end{equation}
1221: introduced in the context of ordinary GUT baryogenesis \cite{KT}.
1222: Based on the insight into the dynamics of the non-equilibrium process gained
1223: from these limiting cases one can then obtain analytic interpolation formulas
1224: that describe rather accurately the entire parameter range.
1225:
1226: To proceed, let us first recall some basic definitions and formulas.
1227: The decay rate is given by the formula \cite{kw80},
1228: \begin{equation}
1229: \Gamma_D(z) = \Gamma_{D1}\,\left\langle {1\over \gamma} \right\rangle \;,
1230: \end{equation}
1231: where the thermally averaged dilation factor is given by the ratio of
1232: the modified Bessel functions $K_1$ and $K_2$,
1233: \begin{equation}\label{G}
1234: \left\langle {1\over \gamma} \right\rangle =
1235: {K_1(z)\over K_2(z)}\;.
1236: \end{equation}
1237: For the decay term $D$, one then obtains
1238: \begin{equation}\label{D}
1239: D(z) = K\,z\,\left\langle {1\over \gamma} \right\rangle\;.
1240: \end{equation}
1241:
1242: The inverse decay rate is related to the decay rate by
1243: \begin{equation}
1244: \Gamma_{I\!D}(z) =\Gamma_D(z)\,{N_{N_1}^{\rm eq}(z)\over N_l^{\rm eq}} \; ,
1245: \end{equation}
1246: where $N_l^{\rm eq}$ is the equilibrium density of lepton doublets. Because
1247: the number of degrees of freedom for heavy Majorana neutrinos and lepton
1248: doublets is the same, $g_{N_1}=g_l=2$, one has
1249: \begin{equation}\label{Neq}
1250: N_{N_1}^{\rm eq}(z) = {3\over 8} z^2 K_2(z) \; , \quad
1251: N_l^{\rm eq} = {3\over 4}\;.
1252: \end{equation}
1253: This yields for the contribution of inverse decays to the washout term
1254: $W$:
1255: \begin{equation}\label{WID2}
1256: W_{I\!D}(z) = {1\over 2} D(z)\,{N_{N_1}^{\rm eq}(z)\over N_l^{\rm eq}}\;.
1257: \end{equation}
1258: All relevant quantities are given in terms of the Bessel functions
1259: $K_1$ and $K_2$, which can be approximated by simple analytical expressions.
1260:
1261: In the regime {\it far out of equilibrium}, $K \ll 1$, decays occur at
1262: very small temperatures, $z\gg 1$, and the produced ($B-L$)-asymmetry
1263: is not reduced by washout effects. In this case, using Eq. (\ref{lg1}) with $S=0$, the integral for the
1264: efficiency factor given in Eq. (\ref{lg3}) becomes simply,
1265: \begin{equation}\label{oodec}
1266: \kappa(z)\simeq{4\over 3}\left(N_{N_1}^{\rm i} - N_{N_1}(z)\right) \; .
1267: \end{equation}
1268: The final value of the efficiency factor $\k_{\rm f} = \k(\infty)$ is
1269: proportional to the initial $N_1$ abundance. If
1270: $N_{1}^{\rm i}=N_{1}^{\rm eq}=3/4$, then $\k_{\rm f}=1$.
1271: But if the initial abundance is zero, then $\k_{\rm f}=0$ as well. Therefore,
1272: in this region there is the well known problem that one has to invoke
1273: some external mechanism to produce the initial abundance of neutrinos.
1274: Moreover, an initial (B-L)-asymmetry is not washed out. Thus in the regime $K \ll 1$ the results strongly depend on
1275: the initial conditions and there is little predictivity.
1276:
1277: In order to obtain the efficiency factor in the case of {\it vanishing
1278: initial $N_1$-abundance},
1279: $N_{N_1}(z_{\rm i}) \equiv N_{N_1}^{\rm i} \simeq 0$, one has to
1280: calculate how heavy neutrinos are dynamically produced by inverse decays.
1281: This requires solving the kinetic equation Eq. (\ref{lg1}) with the initial
1282: condition $N_{N_1}^{\rm i} = 0$.
1283:
1284: Let us define a value $z_{\rm eq}$ by the condition
1285: \begin{equation}\label{condeq}
1286: N_{N_1}(z_{\rm eq})=N_{N_1}^{\rm eq}(z_{\rm eq}) \;.
1287: \end{equation}
1288: Then Eq.~(\ref{lg1}) implies that the number density reaches its maximum at
1289: $z=z_{\rm eq}$. For $z > z_{\rm eq}$ the efficiency factor is
1290: always the sum of two contributions,
1291: \begin{equation}
1292: \k_{\rm f}(z) = \k^-(z) + \k^+(z) \;.
1293: \end{equation}
1294: Here $\k^-(z)$ and $\k^+(z)$ correspond to the integration domains
1295: $[z_{\rm i},z_{\rm eq})$ and $[z_{\rm eq},z)$, respectively.
1296:
1297: Consider first the case of {\it weak washout}, $K \ll 1$, which implies
1298: $z_{\rm eq} \gg 1$. One then finds,
1299: \begin{equation}\label{nkweak}
1300: N_{N_1}(z_{\rm eq})\simeq {9\pi \over 16}\ K \; .
1301: \end{equation}
1302: It turns out that to first order in $K$, there is a cancellation between
1303: $\k^+$ and $\k^-$, yielding for the final efficiency factor
1304: \begin{equation}\label{reduction}
1305: \k_{\rm f}(K)
1306: \simeq {9\pi^2\over 64}\,K^2 \; .
1307: \end{equation}
1308: Note, that Eq.~(\ref{reduction}) does not hold for $K > 1$, because
1309: in this case
1310: $z_{\rm eq}$ becomes small, and washout effects change the result.
1311:
1312: In the case of {\it strong washout}, $K\gg 1$, we can neglect the negative
1313: contribution $\k^-$, because the asymmetry generated at high temperatures
1314: is efficiently washed out. Now the neutrino abundance tracks closely the
1315: equilibrium behavior. Because $D \propto K$, one can solve Eq.~(\ref{lg1}) systematically in powers of $1/K$, which yields
1316: \begin{equation}\label{del}
1317: D\left(N_{N_1}(z)-N_{N_1}^{\rm eq}(z)\right) =
1318: {3\over 2Kz} W_{I\!D}(z)+ {\cal O}\left({1\over K}\right)\; ,
1319: \end{equation}
1320: where we have used properties of the Bessel functions. From Eqs.
1321: (\ref{lg3}) and (\ref{del}) one obtains for the efficiency
1322: factor\footnote{Because $\k^-$ does not contribute we can take the lower limit below as $z_{\rm i}$.}
1323: \begin{equation}\label{efd}
1324: \k(z) = {2\over K} \int_{z_{\rm eq}}^{z} dz'\, {1\over z'} W_{I\!D}(z')\,
1325: e^{- \int_{z'}^{z} dz''\, W_{I\!D}(z'')}\;.
1326: \end{equation}
1327: The integral is dominated by the contribution from a region around the value
1328: $z_{\rm B}$ where the integrand has a maximum, which is determined by the
1329: condition
1330: \begin{equation}\label{eqz0}
1331: W_{I\!D}(z_{\rm B}) =
1332: \left\langle {1\over\gamma}\right\rangle^{-1}(z_{\rm B})\,
1333: -\, {3\over z_{\rm B}}\;.
1334: \end{equation}
1335: For $K\gg 1$ one has $z_{\rm B}\gg 1$, and the condition (\ref{eqz0}) becomes
1336: approximately $W_{I\!D}(z_{\rm B})\simeq 1$, with $W_{I\!D}(z) > 1$ for
1337: $z < z_{\rm B}$ and $W_{I\!D}(z) < 1$ for $z > z_{\rm B}$. This means that
1338: the asymmetry produced for $z < z_{\rm B}$ is essentially erased,
1339: whereas for $z > z_{\rm B}$, washout is negligible.
1340: Hence, the expression of Eq. (\ref{efd}) is a good approximation for
1341: the final efficiency factor.
1342:
1343: One finds that a rather accurate expression for $z_B(K)$ is given by
1344: \begin{equation}
1345: z_B(K) \simeq 1+{1\over2}\,\mbox{ln}\left(1+{\pi K^2\over1024}
1346: \left[\mbox{ln}\left({3125\pi K^2\over1024}\right)\right]^5\right)\;.
1347: \label{interpol}
1348: \end{equation}
1349: The integral of Eq. (\ref{efd}), which gives the final efficiency factor in
1350: terms of $z_B(K)$, is well approximated by
1351: \begin{equation}\label{kfan}
1352: \k_{\rm f}(K) \simeq
1353: {2\over z_{\rm B}(K)K}\left(1-e^{-{1\over 2}z_{\rm B}(K)K}\right)\;.
1354: \end{equation}
1355: Both equations can also be extrapolated into the regime of weak washout,
1356: $K \ll 1$, where one obtains $\k_{\rm f} = 1$ corresponding to thermal initial
1357: abundance, $N_{N_1}^{\rm i}=N_{N_1}^{\rm eq} = 3/4$. At $K\simeq 3$ a rapid
1358: transition takes place from strong to weak washout. Even here analytical
1359: and numerical results agree within 30\%. For the case of zero initial $N_1$
1360: abundance one obtains an interpolation formula $\k_{\rm f}(K)$ analogous
1361: to Eq. (\ref{kfan}).
1362:
1363: \begin{figure}[t]
1364: \centerline{\psfig{file=KFIN.eps,height=8cm,width=12cm}}
1365: \caption{\small Final efficiency factor when the washout term $\Delta W$
1366: is neglected. From \cite{bdp04}.}
1367: \label{KFIN}
1368: \end{figure}
1369:
1370: The above discussion of decays and inverse decays can be extended to
1371: include $\Delta L=1$ and $\Delta L=2$ scattering and washout processes.
1372: In the weak washout regime, $K\ll 1$, the main effect is that the efficiency
1373: factor of Eq. (\ref{reduction}) is enhanced to $\k_{\rm f} \propto K$.
1374: Relevant effects include scattering processes
1375: involving gauge bosons \cite{pu04,gnx04} and thermal
1376: corrections to the decay and scattering rates \cite{gnx04,crx98}.
1377: The range of different results is represented in Fig. (\ref{KFIN}) by
1378: the hatched region. An additional uncertainty in the weak washout
1379: regime is due to the dependence of the final results on the
1380: initial $N_1$ abundance and a possible initial asymmetry
1381: created before the onset of Leptogenesis.
1382:
1383: The situation is very different in the strong washout regime. Here the
1384: final efficiency factor is not sensitive to the neutrino production
1385: because a thermal neutrino distribution is always reached at high
1386: temperatures. For $\mt > m_* \simeq 10^{-3}\,{\rm eV}$, the effect of
1387: $\Delta L=1$ processes on the washout is not larger than about $50\%$,
1388: as indicated by the hatched region in Fig. (\ref{KFIN}). Within these uncertainties,
1389: the final efficiency factor is given by the simple power law:
1390: \begin{equation}\label{plaw}
1391: \k_{\rm f}=(2\pm1)\,10^{-2}\,
1392: \left({0.01\,{\rm eV}\over\mt}\right)^{1.1\pm 0.1}\;.
1393: \end{equation}
1394: Both the scale of solar neutrino oscillations, $m_{\rm sol}
1395: \equiv \sqrt{\Delta m^2_{\rm sol}}\simeq 8\times 10^{-3}\,{\rm eV}$, and the
1396: scale of atmospheric neutrino oscillations, $m_{\rm atm} \equiv
1397: \sqrt{\Delta m^2_{\rm atm}} \simeq 0.05\,{\rm eV}$, are larger
1398: than the equilibrium neutrino mass $m_*$. Hence, the range of neutrino masses,
1399: and therefore $\mt$, indicated by neutrino oscillations
1400: lies entirely in the strong washout regime where
1401: theoretical uncertainties are small and the efficiency factor is still
1402: large enough to allow for successful Leptogenesis.
1403:
1404:
1405: \subsection{Bounds on Neutrino Masses}
1406:
1407: The $\D L=2$ processes with heavy neutrino exchange generate a contribution
1408: to the washout rate that depends on the absolute neutrino mass scale
1409: $\mb^2 = m_1^2 + m_2^2 + m_3^2$,
1410: \begin{equation}
1411: \D W \, \propto \, {M_{\rm P}M_1\, \mb^2\over v_F^4} \; .
1412: \end{equation}
1413: As long as $\D W$ can be neglected, the efficiency factor is independent
1414: of $M_1$. With increasing $\mb$ however, the washout rate $\D W$ becomes
1415: important and eventually prevents successful Leptogenesis. This leads to the
1416: upper bound on the absolute neutrino mass scale. \cite{bdp02,bdp022}
1417:
1418: One can also obtain a lower bound on the heavy neutrino masses \cite{di},
1419: because
1420: the CP asymmetry $\ve_1$ satisfies an upper bound \cite{di,hmy,bdp2,hlx},
1421: which is a function of $M_1$, $\mt$ and $\mb$. Since the rates entering the
1422: Boltzmann equations depend on the same quantities, there exists for arbitrary
1423: neutrino mass matrices a maximal baryon asymmetry $\eta^{\rm max}_{B}$,
1424: \begin{eqnarray}
1425: \eta_{B} &\leq& \eta^{\rm max}_{B}(\mt,M_1,\mb) \NO\\
1426: &\simeq& 0.01\ \ve_1^{\rm max}(\mt,M_1,\mb)\ \k(\mt,M_1\mb^2)\;.
1427: \end{eqnarray}
1428: Requiring the maximal baryon asymmetry to be larger than the observed one,
1429: \begin{equation}\label{cmbcon}
1430: \eta^{\rm max}_{B}(\mt,M_1,\mb) \geq \eta^{CMB}_{B} \;,
1431: \end{equation}
1432: then yields a constraint on the neutrino mass parameters $\mt$, $M_1$
1433: and $\mb$. For each value of $\mb$ one obtains a domain in the
1434: ($\mt$-$M_1$)-plane, which
1435: is allowed by successful baryogenesis. For $\mb \geq 0.20$~eV
1436: this domain shrinks to zero. One can easily translate this
1437: bound into upper limits on the individual neutrino masses. In a similar way,
1438: one finds a lower bound on $M_1$, the smallest mass of the heavy Majorana
1439: neutrinos. The resulting upper and lower bounds are \cite{bdp2}
1440: \begin{equation}
1441: m_i < 0.1\,{\rm eV}\;, \quad M_1 > 4 \times 10^8~\mbox{GeV}\;,
1442: \end{equation}
1443: where we have assumed thermal initial $N_1$ abundance. The upper bound on
1444: the light neutrino masses holds for a normal as well as an inverted hierarchy of masses.
1445: For zero initial $N_1$ abundance one obtains the more restrictive lower
1446: bound $M_1 > 2 \times 10^9~\mbox{GeV}$. For $\mt > m_*$, the baryon asymmetry
1447: is generated at the temperature $T_B \simeq M_1/z_B < M_1$. Hence the lower
1448: bound on the reheating temperature $T_i$ is less restrictive than the lower
1449: bound on $M_1$. The results of a detailed
1450: analytical and numerical calculation are summarized in Fig. (\ref{BTR}).
1451: For the lower bound on the reheating temperature one finds
1452: $T_i > 2 \times 10^9\ {\rm GeV}$ \cite{bdp04,gnx04}.\footnote{In the
1453: supersymmetric case the CP asymmetry is enhanced but also
1454: the washout processes are stronger. These two effects partly compensate each
1455: other \cite{plu98}, leading to the slightly less stringent bound
1456: $T_i > 1.5 \times 10^9\ {\rm GeV}$ \cite{db04}.}
1457:
1458: \begin{figure}[t]
1459: \centerline{\psfig{file=BTR.eps,height=6.5cm,width=11cm}}
1460: \caption{Analytical lower bounds on $M_1$ (circles) and $T_{\rm i}$
1461: (dotted line) for $m_1 = 0$, $\eta_B^{CMB} = 6\times 10^{-10}$ and
1462: $m_{\rm atm} = 0.05\,{\rm eV}$.
1463: The analytical results for $M_1$ are compared with the numerical ones
1464: (solid lines). Upper and lower curves correspond to zero and thermal
1465: initial $N_1$ abundance, respectively.
1466: The vertical dashed lines indicate the range ($m_{\rm sol}$,$m_{\rm atm}$).
1467: The gray triangle at large $M_1$ and large $\mt$ is excluded by theoretical
1468: consistency. From \cite{bdp04}.}
1469: \label{BTR}
1470: \end{figure}
1471:
1472:
1473: What is the theoretical error on the upper bound for the light neutrino
1474: masses? In order to answer this question one needs a full quantum
1475: mechanical treatment of Leptogenesis, a
1476: challenging problem! A possible starting point is the Kadanoff-Baym
1477: equations for which a systematic expansion around the Boltzmann equations
1478: can be constructed \cite{bf00}. One then has to calculate relativistic
1479: corrections, off-shell effects, `memory effects', higher order loop
1480: corrections, etc.
1481: One important effect is the running of neutrino masses between the Fermi
1482: scale and the energy scale of Leptogenesis \cite{bcx00,akx03}. Also relevant
1483: are
1484: the $\D L=1$ scattering processes involving gauge bosons \cite{pu04,gnx04}.
1485: Conceptually interesting are thermal corrections at large
1486: temperatures, $T > M_1$, which correspond to loop corrections involving
1487: gauge bosons and the top quark \cite{gnx04}. Their effect is large if thermal
1488: masses are treated as kinematical masses in the evaluation of scattering
1489: matrix elements. At sufficiently high temperatures the process
1490: $N_1 \rightarrow H \ell_L$ is then kinematically forbidden whereas the process
1491: $H \rightarrow N_1 \ell_L^{\dagger}$ is allowed by `phase space'. On the
1492: contrary, thermal correction are small if they are only included as
1493: propagator effects \cite{crx98}. It is important to clarify this issue
1494: for the treatment of non-equilibrium processes at high temperatures.
1495:
1496: The analysis \cite{gnx04} leads to the upper bound on the light
1497: neutrino masses $m_i < 0.15$~eV. In \cite{bdp04} an upper bound
1498: of $0.12$~eV has
1499: been obtained. About $0.02~{\rm eV}$ of this difference is due to
1500: the different treatment of radiative corrections \cite{akx03}, the remaining
1501: $0.01~{\rm eV}$ reflects differences in the treatment of thermal corrections. This discrepancy has to be compared with an uncertainty of
1502: about $-0.02~{\rm eV}$ due to `spectator processes' \cite{bp}, which have
1503: not been taken into account in both analyses. Hence, within the minimal
1504: seesaw model and the present status of theoretical calculations, the upper
1505: bound on the light Majorana neutrino masses is now known rather precisely.
1506:
1507: The main result of this section is summarized in Fig. (\ref{KFIN}).
1508: For $\mt > m_*$, the efficiency factor, and therefore
1509: the baryon asymmetry $\eta_B$, is independent of the initial $N_1$ abundance.
1510: Furthermore, the final baryon asymmetry does not depend on the value of an
1511: initial baryon asymmetry generated by some other mechanism \cite{bdp2}. Hence,
1512: the value of $\eta_B$ is entirely determined by neutrino properties.
1513: In this way Leptogenesis singles out the neutrino mass range
1514: \begin{equation}
1515: 10^{-3}~{\rm eV} < m_i < 0.1~{\rm eV}\;.
1516: \end{equation}
1517: The firm predictions of thermal Leptogenesis open a window into the physics
1518: of the early universe at temperatures $T_B = {\cal O}(10^{10}~ {\rm GeV})$,
1519: and we can ask what the implications are for dark matter, cosmology and
1520: particle physics.
1521:
1522: \subsection{Triplet Models and Resonant Leptogenesis}
1523:
1524: Measurements in neutrino physics determine the parameters of the neutrino mass
1525: matrix,
1526: \begin{equation}
1527: m_\n = -m_D^T{1\over M}m_D + m_\n^{\rm triplet}\; ,
1528: \end{equation}
1529: which in general contains a contribution from $SU(2)$ triplet fields \cite{wet}
1530: in addition to the seesaw term generated by $SU(2)$ singlet heavy Majorana
1531: neutrinos. So far, we have only considered the minimal case,
1532: $m_\n^{\rm triplet}=0$. Clearly, a dominant triplet
1533: contribution would destroy the connection between Leptogenesis and low energy
1534: neutrino physics.
1535:
1536: The discovery of quasi-degenerate neutrinos with masses above the bound
1537: $0.1~{\rm eV}$ would require significant modifications of minimal Leptogenesis
1538: and/or the seesaw mechanism. In this case $SU(2)$ triplet contributions to
1539: neutrino masses could be a possible way out \cite{tri,hlx,ak}. Clearly,
1540: one then has no upper bound on the light neutrino masses anymore.
1541: Yet Leptogenesis with right-handed neutrino decays can still work
1542: yielding a slightly relaxed lower bound on the heavy neutrino masses.
1543: For instance, one may have $m_i \sim 0.35~{\rm eV}$ with
1544: $M_1 > 4 \times 10^8~{\rm GeV}$ \cite{ak}.
1545:
1546: Another way to reconcile quasi-degenerate light neutrinos with Leptogenesis
1547: makes use of the enhancement of the CP asymmetry in case of
1548: quasi-degenerate heavy neutrinos \cite{fps}. For instance, to raise the
1549: upper bound from $0.1~{\rm eV}$ to $0.4~{\rm eV}$, a degeneracy of
1550: $\D M/M$ for the heavy neutrinos in the range $0.4 - 10^{-3}$ is required,
1551: depending on assumptions about the neutrino mass matrices \cite{bdp2,hlx}.
1552: In the extreme case of `resonant Leptogenesis' \cite{pu04}, CP asymmetries
1553: $\ve = {\cal O}(1)$ are reached for degeneracies
1554: $\D M/M = {\cal O}(10^{-10})$.
1555: In this case the right-handed neutrino masses may be as small as
1556: $1~{\rm TeV}$, which may lead to observable signatures at
1557: colliders. A number of models of this type have been constructed \cite{rlm},
1558: some of which make use of the relative smallness of soft supersymmetry
1559: breaking terms \cite{rls}.
1560:
1561: \subsection[Connection with Low Energy CP Violation]{Is the CP Violation in Leptogenesis Connected with the Low
1562: Energy CP Violation in the Neutrino Sector?}
1563:
1564: As was shown in Section 3, the seesaw model has 6 CP-violating phases in
1565: the Yukawa matrix $h_{ij}$. Leptogenesis depends on one combination
1566: of these 6 phases. However, there are only 3 CP-violating phases at low
1567: energies. Hence it is impossible to determine all 6 phases in the
1568: theory, even if one were to measure all 3 low-energy phases.
1569: Futhermore, as we discussed earlier, one of these low-energy phases remains undetermined by
1570: experiments feasible at low energies.
1571:
1572: Nevertheless, the effective number of high energy
1573: CP-violating phases is reduced if one of the superheavy Majorana
1574: neutrinos $N_i$ is extremely heavy and decouples from the seesaw system.
1575: In this case, the Yukawa couplings $h_{ij}$ effectively are given by a $2\times 3$
1576: matrix that contains 6 complex parameters and hence 6 phases. Three of
1577: the 6 phases can be absorbed into the wave functions of $\ell_L$ and
1578: thus one is left with only 3 CP-violating phases at high energies. In this case, the 3 low-energy
1579: CP-violating phases that appear in the neutrino mass matrix $m_\nu$ are reduced to
1580: only two physical phases, because ${\rm det}(m_\nu)\simeq
1581: 0$. Although these 2 low-energy phases are, in principle, measurable in future experiments, this is still
1582: not enough to determine all three phases in the full
1583: theory. Therefore, even in this simplified example, one cannot establish any link between the sign of the
1584: Universe's baryon-number asymmetry with the observable CP-violating
1585: phases at low energies.
1586:
1587: In the very special case where $h_{ij}$ has two zeros, one has
1588: only one CP-violating phase. In this case the CP-violating phase in
1589: neutrino oscillations is connected with the phase in
1590: Leptogenesis or, equivalently, the sign of the baryon-number asymmetry in
1591: the Universe \cite{FGY}.\footnote{The prediction of the sign of
1592: the CP-violating phase in neutrino oscillations depends on which heavy
1593: Majorana neutrino is responsible for Leptogenesis. This problem is
1594: solved in the inflaton-decay scenario in supersymmetry (SUSY) theories,
1595: because one choice
1596: is unable to produce enough lepton-number asymmetry due to the
1597: constraint on the reheating temperature $T_R < 10^7$ GeV \cite{KKM}.}
1598: Thus, in this case, one may indeed test directly the idea of Leptogenesis.
1599: It is interesting that such a restricted model, where $h_{13}=h_{21}=0$,
1600: is still consistent with data on neutrino oscillations.
1601:
1602:
1603:
1604: \section{Dark Matter Considerations}
1605:
1606:
1607:
1608: It is certainly possible that the mechanism that generates a primordial
1609: matter-antimatter asymmetry in the Universe is not physically related to
1610: the existence of a non-luminous component of the energy density of the
1611: Universe, a component that now accounts for about 25 $\%$ of the total
1612: energy density. However, it would be very interesting if these two
1613: phenomena, so central to the history of the Universe, were connected in
1614: some deep way. It turns out, as we shall see, that if Leptogenesis is
1615: the mechanism by which a primordial matter-antimatter asymmetry in the
1616: Universe is established, it considerably impacts what the dark matter in
1617: the Universe can be.
1618:
1619: Of the three viable options for dark matter, from the point of view of
1620: particle physics, two are either linked or constrained by thermal
1621: Leptogenesis and the third has clear connections to nonthermal
1622: Leptogenesis. Before discussing these points in some detail, it is
1623: useful to briefly review the extant dark matter candidates motivated by
1624: particle physics.
1625:
1626:
1627: \subsection{PQ Symmetry and Axions}
1628:
1629: It is well known that QCD admits the presence of an additional CP violating
1630: term in its Lagrangian density \cite{tH},
1631: \begin{equation}
1632: {\cal{L}}_{\theta}= \frac{\alpha_3 }{8 \pi}\theta F^{\mu \nu}_i
1633: \tilde{F}_{i \mu \nu},
1634: \end{equation}
1635: where $\tilde{F}_{i\mu\nu}=1/2\epsilon_{\alpha\beta\mu\nu}F^{\alpha\beta}_i$.
1636: If $\theta$ is non-vanishing ${\cal{L}}_{\theta}$, which is C-even and
1637: P-odd, violates CP and T invariance. Because possible CP
1638: violating parameters of the strong interactions, like the electric
1639: dipole moment of the neutron, are very tightly bounded by experiment
1640: \cite{edm}, the parameter $\theta$ must be very small ($\theta \leq
1641: 10^{-10}$) \cite{bound}. The reason for this is a mystery, and is known as
1642: the Strong CP Problem \cite{RDPCP}.
1643:
1644: Probably the most `natural' solution suggested for the Strong CP Problem
1645: is to assume that the total Lagrangian for the strong and electroweak
1646: interactions is invariant under a global chiral $U(1)_{PQ}$ symmetry
1647: \cite{PQ}. Even though this symmetry is spontaneously broken, one can
1648: show \cite{PQ} that as a result of the $U(1)_{PQ}$ symmetry the parameter
1649: $\theta$ is driven to zero. In effect, what happens is that the
1650: CP violating Lagrangian term ${\cal{L}}_{\theta}$ is replaced by a
1651: CP conserving interaction between the CP odd pseudo-Goldstone
1652: boson\footnote{Axions are not true Goldstone bosons because the $U(1)_{PQ}$
1653: symmetry is anomalous \cite{WW}. In fact, the same effective potential
1654: for axions that serves to drive $\theta$ to zero gives axions a small
1655: mass. This mass is of order \cite{RDPCP} $m_ a^2 \sim m_q
1656: \Lambda_{QCD}^3/f_a^2$,
1657: where $f_a$ is the scale where the $U(1)_{PQ}$ symmetry breaks down
1658: spontaneously, and $m_q$ is the (light) quark mass.} associated with the
1659: spontaneous breakdown of $U(1)_{PQ}$ -- the axion -- \cite{WW} and
1660: $F\tilde{F}$:
1661: \begin{equation}
1662: {\cal{L}}_{\theta} \to \frac{\alpha_3 }{8 \pi }\frac{a}{f_a} F^{\mu \nu}_i
1663: \tilde{F}_{i \mu \nu}.
1664: \end{equation}
1665:
1666:
1667: Originally, it was supposed \cite{PQ} that the $U(1)_{PQ}$ symmetry was
1668: broken at the electroweak scale. Then $f_a \sim v_F \simeq 200$ GeV and
1669: the axion mass lies in the keV range. However, axions in this mass
1670: range, which are coupled with strength $1/f_a$, have been ruled out by
1671: experiment \cite{RDPCP}. Astrophysical considerations, however, impose
1672: very strong constraints on axions much lighter than a keV, as their
1673: emission from stars would significantly alter their properties. Only
1674: axions that are sufficiently weakly coupled (hence, with large enough
1675: $f_a$ and thus a correspondingly small mass) avoid these constraints,
1676: and one finds the bound \cite{Raffelt} $f_a \geq 10^{10}$ GeV. On the
1677: other hand, $f_a$ cannot be arbitrarily large, because zero-momentum
1678: axion oscillations in the early Universe would carry enough energy
1679: density (proportional, approximately, to $f_a$) to overclose the
1680: Universe \cite{Cbounds}. Thus, for an appropriate value for $f_a$,
1681: axions can be the dark matter in the Universe. In particular, one finds
1682: \cite{Sikivie} that $f_a \simeq 10^{12}$ GeV gives $\Omega_a \simeq 1$.
1683:
1684: \subsection{ Dark Matter Candidates from Supersymmetry}
1685:
1686:
1687: Supersymmetry, a boson-fermion symmetry, has been invoked extensively as
1688: the solution of the so-called hierarchy problem. This problem is
1689: related to the fact that without some stabilizing influence radiative
1690: corrections in the electroweak theory would naturally push the Fermi
1691: scale $v_F$ to have the value of whatever cutoff delimits the validity
1692: of the theory. Typically, this cutoff is imagined to be at the Planck
1693: scale $M_P$, and why $v_F << M_P$ is the hierarchy problem. This problem
1694: is resolved if there is some low energy (spontaneously broken)
1695: supersymmetry. Due to the fermion-boson nature of supersymmetry,
1696: radiative corrections of parameters in the electroweak theory (like $v_F$) are now
1697: only logarithmically dependent on the cutoff, not quadratically
1698: dependent. Hence, effectively, if there is some low energy supersymmetry
1699: one can contemplate having a hierarchy like $v_F<<M_P$, because radiative
1700: shifts can only change $v_F$ logarithmically.
1701:
1702:
1703:
1704: In general, supersymmetric theories possess a discrete symmetry
1705: (R-symmetry) that distinguishes particles from their supersymmetric
1706: partners. As a result, the lightest supersymmetric particle (the LSP) is
1707: stable and, in principle, could be the source of the dark matter in the
1708: Universe. Indeed, it is known \cite{KT} that the energy density
1709: of particles of mass of $O(v_F)$, whose annihilation cross section is of
1710: electroweak strength, is of the order of the critical energy density
1711: that closes the Universe. With supersymmetric partners of ordinary
1712: particles having electroweak scale masses and interactions, the LSP is
1713: therefore an ideal candidate for the dark matter in the Universe
1714: \cite{SDM}. In this review we will discuss both the cases of neutralinos (the
1715: SUSY partners of gauge and Higgs boson) and of gravitinos (the spin 3/2
1716: partner of the graviton) as LSP candidates.
1717:
1718: \subsection{Extended Structures}
1719:
1720: Scalar fields are necessary ingredients of the standard electroweak
1721: model, as well as its supersymmetric extension. It is well known that
1722: theories with scalar fields can lead to the formation of nontopological
1723: solitons. These extended structures, known as Q-balls \cite{Coleman},
1724: may be stable or unstable and arise when some scalar field carries a
1725: conserved $U(1)$ charge. For example, in supersymmetric theories
1726: sleptons and squarks carry, respectively, lepton and baryon number.
1727:
1728:
1729:
1730: In supersymmetric theories, more generally, Q-balls can develop along
1731: flat directions of the scalar potential \cite{DK}. These Q-balls can, in
1732: a number of instances, carry baryon number. If the baryon number of the
1733: Q-ball is large enough, and its mass is small enough, the baryonic
1734: Q-balls are stable. Because of their stability, one can imagine that
1735: these Q-balls could be the dark matter in the Universe.\footnote{This,
1736: however, is not easily achieved because, in general, the squarks are
1737: unstable with their baryon number eventually residing on quarks. If the
1738: squarks are light enough, stability can be achieved. However, as Kasuya et al.
1739: point out \cite{KKT}, it is difficult to explain both the baryon asymmetry
1740: and the dark matter density simultaneously. Nevertheless, there are scenarios
1741: where unstable Q-balls are the source for both baryogenesis and neutralino
1742: dark matter \cite{FY02}.} Typically \cite{DK}, if stable Q-balls exist they have both very large baryon number ($B \sim 10^{26}$)
1743: and are very massive ($M_Q \leq 10^{26}$
1744: GeV). Unfortunately, this makes their detection very difficult, because
1745: their flux is very low \cite{Ara}.
1746:
1747: \subsection{ Natural Connection of Axions with Leptogenesis}
1748:
1749:
1750: The scale of $ U(1)_{PQ}$ breaking needed for axions to be the dark
1751: matter in the Universe ($f_a \simeq 0.3 \times 10^{12}$ GeV) is close
1752: enough to the mass of the lightest right handed neutrino ($M_1 \simeq
1753: 10^{10}$ GeV) needed for Leptogenesis to seek for a common linkage. In
1754: fact, the existence of such a linkage was observed long ago by
1755: Langacker, Peccei and Yanagida \cite{LPY}. What these authors observed
1756: was that if $M_1$ were due to the VEV of a scalar field $\sigma$, one
1757: could identify this field as carrying a PQ-symmetry rather that lepton
1758: number.
1759:
1760: Let us examine this assertion in a bit more detail by looking at the
1761: Yukawa interactions of the quarks and leptons with the three Higgs
1762: fields\footnote{For a PQ symmetry to exist one needs to have two
1763: $SU(2)$ doublet Higgs fields, $\phi_1$ and $\phi_2$, rather than just the
1764: single Higgs field of the Standard Model $H$ (and its Hermitian adjoint $H^{\dagger}$).} $\phi_1$, $\phi_2$ and $\sigma$.
1765: Schematically, one has
1766: \begin{equation} \label{LPQ}
1767: {\cal{L}}_{\rm{Yukawa}}= h_{\sigma} \sigma N_RN_R + h \bar{N}_R
1768: \phi_2 \ell_L + h_u \bar{u}_R \phi_2 q_L + f_d \bar{d}_R \phi_1
1769: q_L + f \bar{e}_R \phi_1 {\ell}_L + {\rm h.c.}\, .
1770: \end{equation}
1771: One sees that Eq. (\ref{LPQ}) is invariant under a PQ-symmetry, where
1772: \begin{eqnarray}
1773: \phi_1, \phi_2&\to& e^{i \alpha} \phi_1, e^{i \alpha} \phi_2\\
1774: N_R, {\it{l}}_R, u_R, d_R&\to& e^{i \alpha} N_R, e^{i \alpha}
1775: {\it{l}}_R, e^{i \alpha} u_R, e^{i \alpha} d_R,
1776: \end{eqnarray}
1777: provided that
1778: \begin{equation}
1779: \sigma \to e^{-2i \alpha} \sigma.
1780: \end{equation}
1781:
1782: To allow $<\sigma>=f_a >>v_F$, as in all invisible axion models
1783: \cite{DFSZ,KSVZ}, requires one fine tuning. In the above case,
1784: this requires the PQ-invariant term in the scalar potential
1785: \begin{equation}
1786: V=\kappa \sigma \phi_1 \phi_2 + {\rm h.c.}
1787: \end{equation}
1788: to have the constant $\kappa \sim v_F^2/f_a$, to allow electroweak
1789: symmetry breaking to occur at a scale much below the scale of $U(1)_{PQ}$
1790: symmetry breaking ($ v_F << f_a$).
1791:
1792: The Yukawa couplings of
1793: this model guarantee that the mass of the lightest right-handed neutrino
1794: and $f_a$ are related: $M_1= 2(h_{\sigma})_{11}f_a$. Thus, if axions are the source of the dark matter energy density in the Universe and the baryon aymmetry arises from Leptogenesis, because $\Omega_{\rm{DM}} \sim f_a$ and $\Omega_{\rm{B}} \sim M_1 \sim f_a$, their ratio is independent of the scale of $U_{PQ}(1)$ breaking. Hence it is perhaps not surprising that this ratio is of order unity.
1795:
1796:
1797:
1798: \subsection{The Gravitino Problem in Supersymmetric Theories}
1799:
1800:
1801: As we discussed in Section 4, for Leptogenesis to be effective, the mass
1802: of the lightest right-handed neutrino has to be greater than
1803: $2\times 10^9$ GeV. This bound, in turn, means that thermal Leptogenesis must
1804: have occurred at temperatures above $2 \times 10^{9}$ GeV. Hence, if the
1805: Universe went through an inflationary period, as all evidence seems to
1806: suggest \cite{WMAP}, the reheating temperature after inflation $T_R$
1807: must have been greater than $2 \times 10^{9}$ GeV for Leptogenesis to be
1808: the source of the matter-antimatter asymmetry in the Universe. This high
1809: reheating temperature is problematic for supersymmetric theories because
1810: it leads to an overproduction of light states, like the gravitino, with
1811: catastrophic consequences for the evolution of the Universe after
1812: inflation. Unless these observational inconsistencies can be avoided, it
1813: appears that Leptogenesis in supersymmetric theories cannot produce the
1814: desired baryon asymmetry in the Universe.
1815:
1816:
1817: The production of gravitinos after inflation has been studied in some
1818: detail \cite{KL}. The thermal production of gravitinos
1819: produced by the strong interactions of quarks, squarks, gluons and
1820: gluinos is governed by the Boltzmann equation \cite{BBB}
1821: \begin{equation} \label{evn3}
1822: \frac{dn_{3/2}}{dt} +3Hn_{3/2}=C_{3/2}(T),
1823: \end{equation}
1824: where
1825: \begin{equation}
1826: C_{3/2}(T)=\frac{3\zeta(3)\alpha_3(T)}{\pi^2}
1827: \frac{T^6}{M_P^2}\left(1+\frac{m_{\tilde{g}}^2}{3m_{3/2}^2}\right)F(T).
1828: \end{equation}
1829: Here $F(T)$ is a thermal factor of O(10) and $m_{\tilde{g}}$ and
1830: $m_{3/2}$ are, respectively, the gluino and the gravitino masses.
1831: Integrating Eq. (\ref{evn3}) to a reheating temperature $T_R$, the resulting
1832: relic density of produced gravitinos is given by
1833: \begin{equation}
1834: \Omega_{3/2}h^2\simeq 0.44 \alpha_3(T_R) \left[1
1835: +\frac{1}{3}\left(\frac{\alpha_3(T_R)}{\alpha_3(\mu)}\right)^2
1836: \left(\frac{m_{\tilde{g}}(\mu)}{m_{3/2}}\right)^2\right]
1837: \left(\frac{T_R}{10^{10}\rm{GeV}}\right)
1838: \left(\frac{m_{3/2}}{\rm{100GeV}}\right),
1839: \end{equation}
1840: where $h$ is the scaled Hubble parameter and $\mu \sim M_Z$.
1841:
1842: If gravitinos are stable (i.e. they are the LSP), the WMAP constraint on
1843: the amount of dark matter in the Universe \cite{WMAP}
1844: \begin{equation}
1845: \Omega_{DM}h^2= 0.1126^{+ 0.0161}_{-0.0181}
1846: %\begin{array}{c}+ 0.0161\\ -0.0181\end{array}
1847: \end{equation}
1848: constrains $\Omega_{3/2}h^2$ to be below this value and, for any given
1849: reheating temperature $T_R$ and gravitino mass $m_{3/2}$, gives a bound
1850: on the gluino mass. If, on the other hand, the gravitinos are not
1851: stable, their rate of production for $T_R>2 \times 10^{9}$ GeV is so
1852: large that subsequent gravitino decays completely alter the standard Big
1853: Bang Nucleosynthesis (BBN) scenario. Thus, in either case, there are
1854: severe constraints imposed on supersymmetric dark matter, which we will
1855: discuss in detail below.
1856:
1857: If the gravitino is unstable, it has a long lifetime and decays during
1858: or after BBN for an interesting range of
1859: the gravitino mass, $m_{3/2}\simeq 100 ~{\rm GeV}-10 ~{\rm TeV}$.
1860: The gravitino decay products destroy the light elements produced by the BBN
1861: and
1862: hence the relic abundance of gravitinos is constrained from above to keep the
1863: success of the BBN \cite{Falom}. This leads to an upper bound of the reheating temperature
1864: $T_R$ after inflation, since the abundance of gravitinos is proportional to
1865: the reheating temperature. A recent detailed analysis derived a strigent
1866: upper bound $T_R< 10^{6-7}$ GeV when the gravitino decay has hadronic
1867: modes (see Fig. (\ref{KKMfig})) \cite{KKM}.
1868: This upper bound is much lower than the temperature for
1869: Leptogenesis, $T_R > 2\times 10^{9}$ GeV \cite{bdp04,gnx04}.
1870: Therefore, thermal Leptogenesis seems difficult to reconcile with low energy supersymmetry if gravitino masses lie in the range $m_{3/2}\simeq 100~{\rm
1871: GeV}-10$ TeV - a natural range for Supergravity (SUGRA) models.
1872:
1873: \begin{figure}[t]
1874: \centerline{\psfig{file=KKM.eps,height=8cm,width=11cm}}
1875: \caption{Upper bounds on the reheating temperature as function of the
1876: gravitino mass for the case where the gravitino dominantly decays into a
1877: gluon-gluino pair. From \cite{KKM}.}
1878: \label{KKMfig}
1879: \end{figure}
1880:
1881: \subsection[Solutions to the Gravitino Problem in Thermal Leptogenesis]{Solutions to the Gravitino Problem in Thermal\\ Leptogenesis}
1882:
1883: There have been several attempts to solve the gravitino
1884: problem in thermal Leptogenesis. Here we will briefly review a number of these proposed solutions.
1885:
1886: One possibility has been proposed by Pilaftsis who considers quasi-degenerate
1887: heavy Majorana neutrinos $(M_1\simeq M_2)$ \cite{enhancement}. In this model
1888: the lepton-asymmetry parameter $\ve$ is enhanced by a factor of $M_1/(M_1-M_2)$
1889: and hence the decays of both $N_1$ and $N_2$ may produce enough asymmetry even for
1890: $T_R < 10^{6-7}$ GeV. However, it is difficult to find a compelling justification for having such a degeneracy in the heavy neutrino spectrum.
1891:
1892: Another proposal was made by Bolz, Buchm\"uller and Pl\"umacher \cite{BBP}, who
1893: consider the case where the gravitino is the stable lightest SUSY particle
1894: (LSP). In this case the next lightest supersymmetric particle (NLSP)
1895: is the subject of the cosmological constraint, because
1896: its decay products may destroy the light elements created by the BBN, much like
1897: the unstable gravitino. Detailed analyses show that this scenario
1898: favors the $\tilde{\tau}$ NLSP compared to the neutralino NLSP and gravitino
1899: masses below $m_{3/2} \simeq 100\ {\rm GeV}$ \cite{FIY,RRA}.\footnote{The
1900: gravitino mass is even less cnstrained if the LSP is a scalar
1901: neutrino or the gluino. For a class of supergravity models an upper bound of
1902: $5\times 10^9$ GeV on the reheating temperature has been obtained \cite{RRA}.}
1903: In general, the
1904: gravitino production can be dominated by NLSP decays \cite{frt03} or by
1905: thermal processes \cite{bhr03}.
1906:
1907: A third proposed solution makes use of gauge-mediation model of supersymmetry breaking in which the gravitino
1908: is the stable LSP with a mass $m_{3/2} < 1$ GeV. It turns out that, if the
1909: gravitino mass is $m_{3/2} < 16$ eV \cite{viel},
1910: then there is no gravitino problem. However, in this case the gravitino cannot be the dark matter. It must be something else, perhaps the axion. For the range of gravitino masses
1911: $m_{3/2}\simeq 100 ~{\rm keV}-1~ {\rm GeV}$, there is the interesting
1912: possibility
1913: that late-time entropy production in a class of gauge mediation models can
1914: naturally make the gravitino the dominant component of dark matter \cite{FIY2}.
1915: In this scenario the reheating temperature can be as high as
1916: $T_R\simeq 10^{13}$
1917: GeV. A light axino with mass ${\cal O}(1\ {\rm keV})$ as LSP
1918: and a gravitino as NLSP would solve the gravitino problem \cite{ay00}.
1919:
1920: Finally, another possible solution arises if supersymmetry breaking effects
1921: are transmitted to the Standard Model sector through the scale anomaly,
1922: resulting in very heavy gravitino masses $m_{3/2}>100$ TeV.
1923: In this case the gravitino decays before the time of BBN and hence there is no
1924: cosmological problem. However, the gravitino decay modes contain always one LSP and
1925: hence the relic abundance of the gravitino must be constrained from above so
1926: that
1927: the density of the nonthermal LSP produced by the gravitino decays does not exceed
1928: the dark-matter density.
1929: This condition leads to $T_R< 10^{11}$ GeV \cite{GGW,IKMY}, which is consistent
1930: with thermal Leptogenesis.
1931:
1932: We stress here that each of the above `solutions' predicts distinct particle spectra at the
1933: TeV scale, which
1934: may be testable in future collider experiments at the
1935: Large Hadron Collider (LHC).
1936: If one discovers supersymmetry but all the above possibilities are
1937: excluded experimentally, this would argue strongly against thermal
1938: Leptogenesis although nonthermal Leptogenesis could still be viable.
1939: On the other hand, if some of these scenarios are confirmed experimentally, thermal Leptogenesis will become much more compelling.
1940:
1941: \subsection{Leptogenesis and Lepton Flavor Violation in SUSY Models}
1942:
1943: Lepton flavor violation is another area in which thermal Leptogenesis and supersymmetry may have some linkages. If the neutrinos have a mass, lepton flavor violation (LFV) processes such as
1944: $\mu\rightarrow e + \gamma$ decay can occur. In non supersymmetric theories, these processes are strongly
1945: suppressed by a factor of $(m_\nu /M_W)^2$ in rate
1946: and hence are unmeasurable physically. However, this is not the case in the SUSY
1947: Standard Model, if the seesaw mechanism is effective.
1948:
1949: In the SUSY Standard Model scalar quarks and leptons are assumed to have
1950: a universal SUSY-breaking soft mass, $m_0$, at the Planck scale.
1951: Otherwise one would have too large flavor-changing neutral currents (FCNC).
1952: However, even then quantum corrections resulting from Yukawa interactions of the quarks
1953: generate a violation of the universality of the soft masses for scalar
1954: quarks, which induces FCNC.
1955: In the lepton sector the Yukawa couplings of the superheavy
1956: Majorana neutrinos $N_i$ also generates non-universal masses
1957: for scalar leptons that serves as a source of LFV \cite{BM}.\footnote{The
1958: Yukawa couplings $h_{ij}$ of the Higgs to the $N_i$s and leptons induce flavor
1959: dependent soft masses for scalar leptons.
1960: At the one-loop level the induced mass is given by
1961: $ (m_{\tilde \ell})^2_{ij} \simeq -6m^2_0/(4\pi)^2h^{\dagger}_{ik}h_{kj}
1962: ln(M_P/M_k)$,
1963: where $m_0$ is the universal soft mass for scalar leptons at the Planck
1964: scale $M_P$. Thus, one may obtain information on the high-energy Yukawa coupling
1965: $h_{ij}$ and the heavy neutrino masses $M_{k}$ by measuring directly the mass matrix for the scalar
1966: leptons. However, the phases in $ h^{\dagger}h$ are different from the phases
1967: contributing to Leptogenesis \cite{di01}.} If the
1968: relevant Yukawa couplings are of $O(1)$, or equivalently $M_3\simeq 10^{15}$
1969: GeV, one predicts a branching ratio for $\mu\rightarrow e + \gamma$ decay
1970: that may be
1971: testable in future experiments. However, an accurate prediction for LFV
1972: processes is very difficult, since it hinges on unkown Yukawa coupling
1973: constants \cite{MS}. In particular, the constraint coming from Leptogenesis
1974: that $M_1\simeq 10^{10}$ GeV is not strong enough to suggest that LFV
1975: processes have potentially testable rates.
1976:
1977:
1978: \section{Nonthermal Leptogenesis}
1979:
1980: Supersymmetry is an important symmetry for the unification
1981: of all interactions and all matter, and the SUSY Standard Model is
1982: considered as a plausible scenario for producing new physics at the TeV scale. Thus, it is quite
1983: interesting to consider theories where supersymmetry is spontaneously broken in a hidden sector connected to ordinary matter by gravitational strength interactions-- the SUGRA framework. The seesaw
1984: mechanism is easily incorporated into this framework. However, as we discussed in some detail in Section 5, the gravitino problem argues against thermal Leptogenesis, particularly in SUGRA.
1985:
1986:
1987: A possible solution to this problem may be provided by nonthermal Leptogenesis
1988: \cite{AD, Shafi,KMY, MSYY,MrYa,MrYa2},
1989: where one does not have a strong constraint on the reheating temperature. We
1990: will discuss here specifically nonthermal Leptogenesis via inflaton decay
1991: \cite{Shafi,KMY}, which we consider an interesting scenario.
1992: In the next subsection, we present general arguments for this
1993: scenario and show that it suggests a lower bound on the mass of the
1994: heaviest light neutrino $m_3 > 0.01$ eV.
1995: In the subsequent subsection, we will also discuss the Affleck-Dine mechanism
1996: \cite{AD} for Leptogenesis which, specifically in supersymmetric theories,
1997: is also an interesting
1998: mechanism to generate the matter-antimatter asymmetry.
1999:
2000: \subsection{Nonthermal Leptogenesis via Inflaton Decay}
2001:
2002: Inflation early on in the history of the Universe is one of the most attractive hypothesis in modern
2003: cosmology, because it not only solves long-standing problems in cosmology,
2004: like the horizon and the flatness problems \cite{guth}, but also
2005: accounts for the origin of density fluctuations \cite{density-fluctuations}.
2006: In this subsection we discuss the hypothesis that the inflaton $\Phi$ decays
2007: dominantly into a pair of the lightest heavy Majorana neutrinos,
2008: $\Phi\rightarrow N_1+N_1$. We assume, for simplicity, that other decay modes
2009: including those into pairs of $N_2$ and $N_3$ are energetically forbidden.
2010: The produced $N_1$ neutrinos decay subsequently into $H + \ell_L$
2011: or ${H}^{\dagger} + \ell_L^{\dagger}$. If the reheating temperature $T_R$ is lower
2012: than the mass $M_1$ of the heavy neutrino $N_1$, then the out-of-equilibrium
2013: condition \cite{Sakharov} is automatically satisfied.
2014:
2015: The above two channels for $N_1$ decay have different branching ratios
2016: when CP conservation is violated. Interference between tree-level and
2017: one-loop diagrams generates a lepton-number asymmetry \cite{l-asymmetry}.
2018: Following our discussion in Section 4, the lepton asymmetry parameter
2019: $\ve$ can be written as \cite{MrYa,MrYa2,susycp}
2020: \footnote{Because of supersymmetry,
2021: the asymmetry parameter $\ve$ below is a factor of 2 larger than that
2022: given in Eq. (\ref{CPas}).}
2023: \begin{equation}
2024: \ve = -\frac {3}{8\pi}\frac{M_1}{\langle H\rangle ^2}
2025: m_{3}\delta _{\rm eff}\,,
2026: \end{equation}
2027: where the effective CP-violating phase $\delta_{\rm eff}$ is given by
2028: \begin{equation}
2029: \delta_{\rm eff} = \frac{ {\rm Im}\left[h^2_{13} + \frac{m_{2}}{m_{3}}
2030: h^2_{12} + \frac{m_{1}}{m_{3}}h^2_{11}\right]}{|h_{13}|^2 + |h_{12}|^2
2031: + |h_{11}|^2}\;.
2032: \end{equation}
2033: Numerically, one obtains for the $\ve$
2034: parameter
2035: \begin{equation}
2036: \ve \simeq -2\times 10^{-6} \left( \frac{M_1}{10^{10}{\rm GeV}}\right)
2037: \left(\frac{m_{3}}{0.05 {\rm eV}}\right)\delta_{\rm eff}\;.
2038: \end{equation}
2039:
2040:
2041: The chain decays $\Phi\rightarrow N_1 + N_1$ and $N_1\rightarrow H + \ell_L$
2042: or $H^{\dagger} + \ell_L^{\dagger}$ reheat the Universe producing not only the
2043: lepton-number asymmetry but also entropy for the thermal bath.
2044: The ratio of the lepton number to entropy density after reheating
2045: is estimated to be \cite{KMY}
2046: \begin{eqnarray}
2047: \frac{n_L}{s} &\simeq& -\frac{3}{2}\ve\frac{T_R}{m_{\Phi}} \nonumber \\
2048: &\simeq& 3\times 10^{-10} \left(\frac{T_R}{10^6{\rm GeV}}\right)
2049: \left(\frac{M_1}{m_\Phi}\right)
2050: \left(\frac{m_{3}}{0.05{\rm eV}}\right),
2051: \end{eqnarray}
2052: where $m_{\Phi}$ is the inflaton mass and we have taken $\delta_{\rm eff} =1$.
2053: This lepton-number asymmetry is converted into a baryon-number asymmetry
2054: through the sphaleron effects and one obtains \cite{ht}
2055: \begin{equation}
2056: \frac{n_B}{s} \simeq -\frac {8}{23}\frac {n_L}{s}\;.
2057: \end{equation}
2058:
2059: We should stress, here, an important merit of the inflaton-decay
2060: scenario: It does not require a reheating temperature $T_R \sim M_1$,
2061: but it requires only $m_{\Phi} > 2M_1$. On the other hand, for thermal
2062: Leptogenesis to work it is necessary that $T_R \sim M_1$,
2063: which necessitates higher reheating temperature for Leptogenesis to
2064: produce enough matter-antimatter asymmetry.
2065:
2066:
2067: If one assumes that $T_R < 10^7$ GeV to satisfy the cosmological constraint on
2068: the gravitino abundance \cite{KKM} discussed earlier and
2069: uses $m_\Phi > 2M_1$, the observed baryon number to entropy ratio \cite{WMAP}
2070: gives a constraint on the heaviest light neutrino:
2071: \begin{equation} \label{m3}
2072: m_{3} > 0.01~ {\rm eV}.
2073: \end{equation}
2074: It is very interesting
2075: that the neutrino mass suggested by atmospheric neutrino oscillation experiments,
2076: $\sqrt{\Delta m^2_{\rm atm}}\simeq 0.05$ eV, just satisfies the above
2077: constraint. However, to get this bound
2078: we assumed that the inflaton decays dominantly
2079: into a pair of $N_1$s. If this branching ratio is only 10 $\%$, the
2080: lower bound on the neutrino mass exceeds the observed neutrino mass
2081: $\sqrt{\Delta m^2_{\rm atm}}\simeq 0.05$ eV.
2082:
2083: A variety of models have been considered to restore the bound of
2084: Eq. (\ref{m3}) by imposing a symmetry. However, it is perhaps most
2085: interesting to consider that the
2086: scalar partner of the heavy Majorana neutrino $N_1$ is the inflaton
2087: itself \cite{MSYY}, and the inflaton decay into a lepton plus a Higgs boson gives an effective branching ration of 100\%.
2088: In this model, one must assume that the initial
2089: value of the scalar partner of $N_1$ is much larger than the Planck
2090: scale to cause inflation (chaotic inflation \cite{linde}).
2091: However, chaotic inflation is not easily realized in SUGRA, because
2092: the minimal supergravity potential has
2093: an exponential factor, ${\rm exp}(\phi^*\phi /M^2_G)$, that prevents
2094: any scalar field $\phi$ from having a value larger than the reduced Planck scale $M_G \simeq 2.4\times 10^{18}$ GeV. Ref. \cite{GL} uses a restricted form of
2095: the Kahler potential.
2096:
2097: \subsection{Affleck-Dine Leptogenesis}
2098:
2099: In the SUSY Standard Model, in the limit of unbroken supersymmetry, some combinations of scalar fields do not enter
2100: the potential, constituting so-called flat directions
2101: of the potential.
2102: Since the
2103: potential is almost independent of these fields, they may have large initial values in the early Universe. Such flat directions
2104: receive soft masses in the SUSY-breaking vacuum. When the expansion rate
2105: $H_{\rm exp}$ of the Universe becomes comparable to their masses, the flat directions begin to
2106: oscillate around the minimum of the potential. If the
2107: flat directions are made of scalar quarks and carry baryon number, the
2108: baryon-number asymmetry can be created
2109: during these coherent oscillations. This is the
2110: Affleck-Dine (AD) mechanism for Baryogenesis \cite{AD}.
2111:
2112: QCD corrections, however, make the potential of the AD fields milder
2113: than $|\phi|^2$. This allows non-topological soliton solutions (Q-balls)
2114: \cite{Q-ball2} to form in the early Universe, as a result of the coherent oscillations in the flat
2115: directions. Because Q-balls
2116: have long lifetimes, their decays produce a huge amount of entropy at late times. To avoid this problem one must choose parameters in the SUSY theory so that the density
2117: of the lightest SUSY particle (LSP) does not exceed the dark matter
2118: density in the present Universe \cite{Q-ball2}. Although this may not be a problem, it is much safer to consider flat directions without QCD
2119: interactions, because such directions most likely do not have Q-ball solutions.
2120:
2121: The most interesting candidate \cite{MrYa} for such a flat direction is
2122: \begin{equation}
2123: \phi_i = (2H\ell_i)^{1/2},
2124: \end{equation}
2125: where $\ell_i$ is the lepton doublet field of the $i$-$th$ family.
2126: Here, $H$ and $\ell_i$ represent the scalar components of the corresponding
2127: chiral multiplets. The Yukawa
2128: interactions of $H$ make the potential of $\phi_i$ steeper than
2129: the mass term and hence there is no instability of the coherent oscillation
2130: (i.e. there are no Q-ball solutions). Because this flat direction carries
2131: lepton number, a lepton asymmetry will be created during the coherent
2132: oscillation (AD Leptogenesis) \cite{MrYa}. Sphaleron processes then transmute,
2133: in the usual fashion, this lepton asymmetry into a baryon asymmetry.
2134:
2135: The seesaw mechanism induces a dimension-five operator in the
2136: superpotential for the theory,\footnote{For ease of notation we have dropped the subscript $i$ below.}
2137: \begin{equation}
2138: W=\frac{m_{\nu}}{2|\langle H\rangle |^2}(\ell H)^2,
2139: \end{equation}
2140: where we have used a basis in which the neutrino mass matrix is
2141: diagonal. With this superpotential we have a SUSY-invariant potential
2142: for the flat direction $\phi$ given by
2143: \begin{equation}
2144: V_{\rm SUSY}=\frac{m_{\nu}^2}{4|\langle H\rangle |^4}|\phi|^6.
2145: \end{equation}
2146: In addition to the SUSY-invariant potential we have a SUSY-breaking
2147: potential,
2148: \begin{equation}
2149: \delta V= m_{\phi}^2|\phi|^2 + \frac{m_{\rm SUSY}m_{\nu}}{8|\langle H\rangle |^2}
2150: (a_m\phi^4 + {\rm h.c.}).
2151: \end{equation}
2152: Here, $a_m$ is a complex number. We take
2153: $m_{\phi} \simeq m_{\rm SUSY} \simeq 1$ TeV and $|a_m|\sim 1$.
2154: The second term in $\delta V$ is very important, because it gives rise to
2155: the lepton-number generation.
2156:
2157: We assume that the flat direction $\phi$ acquires a negative $({\rm
2158: mass})^2$ induced by the inflaton potential and rolls down to the point
2159: balanced by the SUSY-invariant potential $V_{\rm SUSY}$ during
2160: inflation. Thus, the AD field $\phi$ has an initial value of
2161: $\sqrt{H_{\rm inf}|\langle H\rangle |^2/m_{\nu}}$, where $H_{\rm inf}$
2162: is the Hubble constant (the expansion rate) during inflation. $\phi$ decreases in
2163: amplitude gradually after inflation, and begins to oscillate around
2164: the potential minimum when the Hubble constant $H_{\rm exp}$ of the
2165: Universe becomes
2166: comparable to the SUSY-breaking mass $m_{\phi}$.
2167: At the beginning of the oscillation, the AD field
2168: has the value $ |\phi_{0}| \simeq \sqrt{m_{\phi}|\langle H\rangle
2169: |^2/m_{\nu}}$ which, as shown below, is an effective initial value for
2170: Leptogenesis.
2171:
2172: Let us consider now lepton-number generation in this scenario. The evolution of the AD
2173: field $\phi$ is described by
2174: \begin{equation}
2175: \frac{\partial^2\phi}{\partial t^2} + 3H_{\rm exp}\frac{\partial \phi}{\partial t}
2176: + \frac{\partial V}{\partial \phi^*} = 0 \;,
2177: \end{equation}
2178: where $V=V_{\rm SUSY} + \delta V$.
2179: Because the lepton number is given by
2180: \begin{equation}
2181: n_{\rm L} = i\left(\frac{\partial \phi^*}{\partial t}\phi -
2182: \phi^*\frac{\partial \phi}{\partial t}\right)\;,
2183: \end{equation}
2184: the evolution of $n_{\rm L}$ is given by
2185: \begin{equation}
2186: \frac{\partial n_{\rm L}}{\partial t} + 3H_{\rm exp}n_{\rm L}
2187: = \frac{m_{\rm SUSY}m_{\nu}}{2|\langle H\rangle |^2}
2188: {\rm Im}(a_m^*\phi^{*4})\;.
2189: \end{equation}
2190:
2191: The motion of $\phi$ in the phase direction generates the lepton number. This is predominantly created just after the AD field $\phi$
2192: starts its coherent oscillation, at a time $t_{\rm osc} \simeq 1/H_{\rm
2193: osc}\simeq 1/m_{\phi}$, because the amplitude $|\phi|$ damps as $t^{-1}$
2194: during the oscillation. Thus, we obtain for the lepton number
2195: %
2196: \begin{equation}
2197: n_{\rm L}\simeq \frac{m_{\rm SUSY}m_{\nu}}{2|\langle
2198: H\rangle |^2}\delta_{\rm eff}|a_m\phi_0^4|\times t_{\rm osc}\;,
2199: \end{equation}
2200: where $\delta_{\rm eff} = {\rm sin}(4{\rm arg}\phi +{\rm arg}a_m)$
2201: represents an effective CP-violating phase. Using $m_{\rm SUSY}\simeq
2202: m_\phi$, $ |\phi_0| \simeq \sqrt{m_{\phi}|\langle H\rangle
2203: |^2/m_{\nu}}$ and $t_{\rm osc}\simeq 1/m_\phi$, we find
2204: \begin{equation}
2205: n_{\rm L}\simeq \delta_{\rm eff}m_{\phi}^2 \frac{|\langle
2206: H\rangle |^2}{2m_{\nu}}\;.
2207: \end{equation}
2208:
2209:
2210: After the end of inflation, the inflaton begins to oscillate around
2211: the potential minimum and $n_{\rm L}/\rho_{\rm inf}$
2212: stays constant until the inflaton decays. Here $\rho_{\rm inf}$
2213: is the energy density of the inflaton. The inflaton decay reheats the
2214: Universe producing entropy $s$. Because $\rho /s = 2T_R/4$, we find for the lepton-number asymmetry the expression
2215: \begin{equation}
2216: \frac{n_{\rm L}}{s} \simeq \left( \frac{\rho_{\rm inf}}{s}\right)
2217: \left(\frac{n_{\rm L}}{\rho_{\rm inf}}
2218: \right) \simeq \delta_{\rm eff}\frac{3T_R}{4M_{ G}}
2219: \frac{|\langle H\rangle |^2}{6m_{\nu}M_{G}}\;.
2220: \end{equation}
2221: Here we have used $\rho_{\rm inf} \simeq 3m_{\phi}^2M_{G}^2$ at the
2222: begining of the AD field oscillation (when most of the lepton number
2223: is generated). This lepton-number asymmetry is converted to a
2224: baryon-number asymmetry by the KRS mechanism. In this way one obtains for the baryon-number asymmetry
2225: \begin{equation}
2226: \frac{n_{\rm B}}{s} \simeq \frac{1}{23}
2227: \frac{|\langle H\rangle |^2T_R}{m_{\nu}M_{G}^2}\;.
2228: \end{equation}
2229: The observed ratio $n_{\rm B}/s \simeq 0.9\times 10^{-10}$ implies
2230: $m_\nu \simeq 10^{-9}$ eV for $T_R \simeq 10^6$ GeV. This
2231: small mass corresponds to the mass of the lightest neutrino. We should
2232: note that for such a low reheating temperature one may neglect the
2233: effects due to thermal mass for the AD field $\phi$ \cite{FHY}.
2234:
2235: \section{Conclusions and Summary of Results}
2236:
2237: In this article we have discussed the physical mechanism responsible for the origin of matter in the Universe. Both the rather
2238: large observed value for the ratio of baryons to photons, $\eta_B$, in the present epoch and the absence of antimatter are the consequences of a primordial asymmetry between matter and antimatter generated early on in the Universe. Although a variety of mechanisms have been proposed for producing this primordial asymmetry, in this review we have focused on Leptogenesis as the origin of matter. In our view, this is the most appealing scenario for the origin of matter, for at least three reasons:
2239:
2240: 1) Explicit lepton number violation is very natural once one includes right-handed neutrinos in the Standard Model. Furthermore, the lightness of the observed neutrinos strongly suggests, through the seesaw mechanism, the presence of superheavy neutrinos, whose decays can produce a lepton-antilepton asymmetry.
2241:
2242: 2) Quantum mechanically, through the KRS mechanism, one can automatically
2243: turn a leptonic asymmetry into a baryonic asymmetry. Indeed, because of the
2244: existence of these sphaleron processes, the origin of matter is linked to
2245: phenomena in the early Universe that result in the establishment of a
2246: (B-L)-asymmetry, like Leptogenesis.
2247:
2248: 3) If neutrino masses lie in the range $ 10^{-3} \rm{eV} < m_i< 0.1 \rm{eV}$, as suggested by neutrino oscillation experiments, the leptonic asymmetry produced in thermal Leptogenesis is both independent of the abundance of heavy neutrinos and of any pre-existing asymmetry and has the right magnitude to yield the observed value for $\eta_B$.
2249:
2250: Because, in the final analysis, $ \eta_B$ is just one number, it is important
2251: to ask if the particular mechanism proposed for the origin of matter has
2252: other consequences. Thus in this review we examined in some detail how
2253: Leptogenesis fit with ideas proposed to explain the dark matter that
2254: constitutes about 25\% of the Universe's energy density.\footnote{We did not
2255: try to examine models of dark energy in the light of Leptogenesis, because our understanding of dark energy is still in its infancy.} We pointed out that axionic dark matter is perfectly compatible with Leptogenesis. Indeed, it is possible to very naturally link the scale of the heavy neutrinos with that of $U(1)_{PQ}$ breaking $f_a$, so that the ratio $\Omega_{\rm DM}/ \Omega_{B}$ is independent of these large scales. The situation, however, is more complex in the case of supersymmetric dark matter.
2256:
2257: To be effective, thermal Leptogenesis needs to occur at high temperatures, above $T= 2 \times 10^9$ GeV. This means that the Universe after inflation must have reheated to at least this temperature. However, in supersymmetric theories such a high reheating temperature is problematic as it leads to an overproduction of gravitinos. When they decay,
2258: gravitinos of such abundances completely alter the primordial abundance of light elements produced in Big Bang Nucleosynthesis.
2259: The gravitino problem, however, is not fatal as there are a number of ways to mitigate the overproduction of gravitinos. Nevertheless, if Leptogenesis is at the root of the origin of matter, the supersymmetric spectrum at low energies and the nature of the LSP are quite constrained. Thus, in a sense, Leptogenesis is also quite predictive in this context.
2260:
2261: Although much of our review, very naturally, focused on thermal Leptogenesis,
2262: we also discussed two examples where matter originated through a leptonic
2263: asymmetry produced in nonthermal processes. These models, although much more
2264: speculative, illustrate some of the possible other options for the origin of
2265: matter. Naturally, in this case some of the specific predictivity is lost.
2266:
2267: \vspace{0.5cm}
2268: \noindent
2269: {\bf Acknowledgments}\\
2270: \noindent
2271: In our work on the topics discussed in this review we have benefitted from
2272: the insight of many colleagues. W.B. and T.Y. owe special thanks to
2273: M. Bolz, A. Brandenburg, P. Di Bari, K. Hamaguchi, M. Ibe, K. Izawa,
2274: T. Moroi, M. Pl\"umacher and M. Ratz. The work of T.Y. has been supported
2275: in part by a Humboldt Research Award. RDP's work was supported in part by the
2276: Department of Energy under Contract No. FG03-91ER40662, Task C.
2277:
2278: %\newpage
2279:
2280: \begin{thebibliography}{99}
2281:
2282: \bibitem{SCM} U.~Seljak, {\it {et al.}}, Phys.~Rev.~{\bf D71}, 103515 (2005).
2283:
2284: \bibitem{inflation} A.~H.~Guth, Phys.~Rev.~{\bf D23}, 347 (1981);
2285: A.~D.~Linde, Phys.~Lett.~{\bf B129}, 177 (1982);
2286: A.~Albrecht and P.~J.~Steinhardt, Phys.~Rev.~Lett.~{\bf 48}, 1220 (1982).
2287:
2288: \bibitem{DE} J.~L.~Tonry, {\it {et al.}}, Astrophys.~J.~{\bf 594}, 1 (2003);
2289: R.~A.~Knop, {\it {et al.}}, Astrophys.~J.~{\bf 598}, 102 (2003);
2290: B.~J.~Barris, {\it {et al.}}, Astrophys.~J.~{\bf 602}, 571 (2004);
2291: A.~G.~Riess, {\it{et al.}}, Astrophys.~J.~{\bf 607}, 665 (2004);
2292: for a review see, S.~Perlmutter and B.~Schmidt, in
2293: {\bf{ Supernovas and Gamma Ray Bursts}}, ed. K.~Weller
2294: (Springer Verlag, Berlin, 2003).
2295:
2296: \bibitem{WMAP} WMAP Collaboration, C.~L.~Bennett, {\it{et al}},
2297: Astrophys.~J.~Suppl.~{\bf 148}, 1 (2003); D.~N.~Spergel, {\it {et al}},
2298: Astrophys.~J.~Suppl.~{\bf 148}, 175 (2003).
2299:
2300: \bibitem{nucleo} J.~P.~Kneller and G.~Steigman,
2301: New~J.~Phys.~{\bf 6}, 117 (2004).
2302:
2303: \bibitem{OS} K.~A.~Olive and E.~D.~Skillman,
2304: Astrophys.~J.~{\bf 617}, 29 (2004).
2305:
2306: \bibitem{anti} A.~Dolgov and J.~Silk, Phys.~Rev.~{\bf D47}, 4244 (1993);
2307: M.~Y.~Khlopov, M.~G.~Rubin and A.~S.~Sakharov, Phys.~Rev.~{\bf D62},
2308: 083505 (2000).
2309:
2310: \bibitem{CDG} A.~Cohen, A.~De~Rujula and S.~Glashow,
2311: Astrophys.~J.~{\bf 495}, 539 (1998).
2312:
2313: \bibitem{ann} See, for example, M.~S.~Turner, in
2314: {\bf Intersection Between Particle Physics and Cosmology},
2315: Jerusalem Winter School for Theoretical Physics, Vol. 1, p. 99, eds.
2316: T.~Piran and S.~Weinberg (World Scientific, Singapore, 1986).
2317:
2318: \bibitem{AD} I.~Affleck and M.~Dine, Nucl.~Phys.~{\bf B249}, 361 (1985).
2319:
2320: \bibitem{Sakharov} A.~D.~Sakharov, JETP~Lett.~{\bf 5}, 24 (1967).
2321:
2322: \bibitem{FY} M.~Fukugita and T.~Yanagida, Phys.~Lett.~{\bf B174}, 45 (1986).
2323:
2324: \bibitem{KRS} V.~A.~Kuzmin, V.~A.~Rubakov and M.~A.~Shaposhnikov,
2325: \pl{155}{1985}{36}.
2326:
2327: \bibitem{seesaw}
2328: T.~Yanagida, in {\bf Proc. of the Workshop on ''The Unified Theory and the
2329: Baryon Number in the Universe''}, Tsukuba, Japan, Feb. 13-14, 1979, p. 95,
2330: eds. O.~Sawada and S.~Sugamoto, (KEK Report KEK-79-18, 1979, Tsukuba);
2331: Progr. Theor. Phys. {\bf 64}, 1103 (1980);
2332: P.~Ramond, in {\bf Talk given at the Sanibel Symposium}, Palm Coast, Fla.,
2333: Feb. 25-Mar. 2, 1979, preprint CALT-68-709.
2334:
2335: \bibitem{yo78}
2336: M.~Yoshimura, \prl{41}{1978}{281}; {\it ibid.} {\bf 42}, 746(E) (1979);
2337: D.~Toussaint, S.~B.~Treiman, F.~Wilczek and A.~Zee, \pr{19}{1979}{1036};
2338: S.~Weinberg, \prl{42}{1979}{850};
2339: S.~Dimopoulos and L.~Susskind, \pr{18}{1978}{4500}.
2340:
2341: \bibitem{electro} V. A. Rubakov and M. S. Shaposhnikov, Phys. Usp. {\bf 39}, 461 (1996).
2342:
2343: \enlargethispage{0.5cm}
2344:
2345: \bibitem{tH} G.~t'~Hooft, Phys.~Rev.~Lett.~{\bf 37}, 8 (1976);
2346: Phys.~Rev.~{\bf D14}, 4332 (1976).
2347:
2348: \bibitem{Manton}
2349: F.~R.~Klinkhammer and N.~S.~Manton, \pr{30}{1984}{2212}.
2350:
2351: \bibitem{aml87} P.~Arnold and L.~McLerran, Phys.~Rev.~{\bf D36}, 581 (1987).
2352:
2353: \bibitem{yaffe}
2354: P.~Arnold, D.~Son and L.~Yaffe, \pr{55}{1997}{6264}; \pr{59}{1999}{105020}.
2355:
2356: \bibitem{bodeker}
2357: D.~B\"odeker, \pl{426}{1998}{351}; \np{559}{1999}{502}.
2358:
2359: \bibitem{yaffe2}
2360: P.~Arnold and L.~Yaffe, \pr{62}{2000}{125014}.
2361:
2362: \bibitem{bmr00}
2363: D.~B\"odeker, G.~D.~Moore and K.~Rummukainen, Phys.~Rev. {\bf D61}, 056003 (2000).
2364:
2365: \bibitem{sh86}
2366: M.~E.~Shaposhnikov, JETP Lett. {\bf 44}, 465 (1986).
2367:
2368: \bibitem{ja96} K.~Jansen, Nucl.~Phys. (Proc.~Supp.) {\bf B47}, 196 (1996).
2369:
2370: \bibitem{PDG} Particle Data Group, S.~Eidelman, {\it et al},
2371: Phys.~Lett.~{\bf B592}, 1 (2004).
2372:
2373: \bibitem{susyew} For recent work and references, see
2374: T.~Prokopec, K.~Kainulainen, M.~G.~Schmidt and S.~Weinstock, in
2375: {\bf Strong and Electroweak Matter 2002}, ed. M.~G.~Schmidt
2376: (World Scientific, Singapore, 2003);
2377: M.~Carena, A.~Megevand, M.~Quir\'os and C.~E.~M.~Wagner,
2378: Nucl.~Phys.~{\bf B716}, 319 (2005).
2379:
2380: %\enlargethispage{0.5cm}
2381:
2382: \bibitem{bp95}
2383: W.~Buchm\"uller and O.~Philipsen, \np{443}{1995}{47}.
2384:
2385: \bibitem{klrs96}
2386: K.~Kajantie, M.~Laine, K.~Rummukainen and M.~Shaposhnikov, \prl{77}{1996}{2887}.
2387:
2388: \bibitem{bw97}
2389: B.~Bergerhoff and C.~Wetterich, in {\bf Current Topics in Astrofundamental
2390: Physics}, p. 162, eds. N.~Sanchez and A.~Zichichi (World Scientific, Singapore, 1997).
2391:
2392: \bibitem{rtk98}
2393: K.~Rummukainen, M.~Tsypin, K.~Kajantie, M.~Laine and M.~Shaposhnikov,
2394: \np{532}{1998}{283}.
2395:
2396: \bibitem{bp97}
2397: W.~Buchm\"uller and O.~Philipsen, \pl{397}{1997}{112}.
2398:
2399: \bibitem{fo99}
2400: Z.~Fodor, Nucl.~Phys.(Proc.~Supp.) {\bf B83-84}, 121 (2000).
2401:
2402: \bibitem{ce97}
2403: B.~de~Carlos and J.~R.~Espinosa, \np{503}{1997}{24}.
2404:
2405: \bibitem{ht} J.~A.~Harvey and M.~S.~Turner, \pr{42}{1990}{3344}.
2406:
2407: \bibitem{moh}
2408: R.~N.~Mohapatra and X.~Zhang, \pr{45}{1992}{2699}.
2409:
2410: \bibitem{ks}
2411: S.~Yu.~Khlebnikov and M.~E.~Shaposhnikov, \np{308}{1988}{885}.
2412:
2413: \bibitem{ls00}
2414: M.~Laine and M.~E.~Shaposhnikov, Phys.~Rev. {\bf D61}, 117302 (2000).
2415:
2416: \bibitem{dlx} K.~Dick, M.~Lindner, M.~Ratz and D.~Wright, \prl{84}{2000}{4039}.
2417:
2418: \bibitem{history} See, for example, S.~M.~Bilenky, hep-ph/0410090,
2419: report at the ``Nobel Symposium on Neutrino physics'', Haga Slott, Enkoping,
2420: Sweden, August 19-24, 2004, to be published in Physica Scripta.
2421:
2422: \bibitem{CHOOZ} CHOOZ Collaboration, M.~Apollonio {\it et al.},
2423: Phys.~Lett.~{\bf B420}, 397 (1998); CHOOZ Collaboration, M.~Apollonio
2424: {\it et al.}, Phys.~Lett.~{\bf B466}, 415 (1999); Palo Verde Collaboration,
2425: F.~Boehm, {\it et al.}, Phys.~Rev.~{\bf D64}, 112001 (2001).
2426:
2427: \bibitem{LSND} LSND Collaboration, A.~Aguilar, {\it et al.},
2428: Phys.~Rev.~{\bf D64}, 112007 (2001).
2429:
2430: \bibitem{KARMEN} KARMEN Collaboration: B.~Armbruster, {\it et al.},
2431: Phys.~Rev.~{\bf D65}, 112001 (2002).
2432:
2433: \bibitem{sterile} For a recent discussion see, for example,
2434: M.~Cirelli, G.~Marandella, A.~Strumia and F.~Vissani,
2435: Nucl.~Phys.~{\bf B708}, 215 (2005).
2436:
2437: \bibitem{CPT} See, for example, G.~Barenboim, J.~F.~Beacom, L.~Borissov
2438: and B.~Kayser, Phys.~Lett.~{\bf B537}, 227 (2002).
2439:
2440: \bibitem{Kayser} See, for example, B.~Kayser, in
2441: {\bf Neutrino Mass}, eds. G.~Altarelli and K.~Winter (Springer Tracts in
2442: Modern Physics, Berlin, 2003), p.1.
2443:
2444: \bibitem{HR} S.~Hannestad and G.~Raffelt,
2445: J.~Cosmol.~Astropart.~Phys.~{\bf 0404}, 008 (2004).
2446:
2447: \bibitem{fukugita} K.~Ichikawa, M.~Fukugita and M.~Kawasaki,
2448: Phys.~Rev.~{\bf D71}, 043001 (2005).
2449:
2450: \bibitem{TR97} See, for example,
2451: M.~Tegmark and M.~Rees, Astrophys.~J.~{\bf499} , 526 (1998).
2452:
2453: \bibitem{arkani} N.~Arkani-Hamed, S.~Dimopoulos and G.~Dvali,
2454: Phys.~Lett.~{\bf B429}, 263 (1998).
2455:
2456: \bibitem{arkani2}
2457: R.~K.~Dienes, E.~Dudas, T.~Gherghetta, Nucl.~Phys.~{\bf B557}, 25 (1999);
2458: N.~Arkani-Hamed, S.~Dimopoulos, G.~Dvali and
2459: J.~March-Russell, Phys.~Rev.~{\bf D65}, 024032 (2002); Y.~Grossman and
2460: M.~Neubert, Phys.~Lett.~{\bf B474}, 361 (2000);
2461: T.~Gherghetta, Phys.~Rev.~Lett.~{\bf 92}, 161601 (2004).
2462:
2463: \bibitem{chargenu} J.~Baumann, R.~Gahler, J.~Kalus and W.~Mampe,
2464: Phys.~Rev.~{\bf D37}, 3107 (1988).
2465:
2466: \bibitem{chargeneu} H.~F.~Dylla and J.~G.~King, Phys.~Rev.~{\bf A7},
2467: 1224 (1973).
2468:
2469: \bibitem{petcov} S.~M.~Bilenky, J.~Hosek and S.~T.~Petcov,
2470: Phys.~Lett.~{\bf B94}, 495 (1980); J.~Schechter and J.~W.~F.~Valle,
2471: Phys.~Rev.~{\bf D22}, 2227 (1980).
2472:
2473: \bibitem{BR} G.~C.~Branco and M. N. Rebelo,
2474: New~J.~Phys.~{\bf 7}, 86 (2005).
2475:
2476: \bibitem{branco} G.~C.~Branco, R.~G.~Felipe, F.~R.~Joachim and
2477: T.~Yanagida, Phys.~Lett.~{\bf B562}, 265 (2003).
2478:
2479: \bibitem{l-asymmetry} M.~Fukugita and T.~Yanagida, in ref.\cite{FY};
2480: M.~Flanz, E.~A.~Paschos and U.~Sarkar, Phys.~Lett.~{\bf B345}, 248 (1995);
2481: L.~Covi, E.~Roulet and F.~Vissani, Phys.~Lett.~{\bf B384}, 169 (1996);
2482: W.~Buchm\"uller and M.~Pl\"umacher, Phys.~Lett.~{\bf B431}, 354 (1998).
2483:
2484: \bibitem{pil99}
2485: For a discussion and references, see
2486: A.~Pilaftsis, Int.~J.~Mod.~Phys. {\bf A14}, 1811 (1999).
2487:
2488: \bibitem{bf00}
2489: W.~Buchm\"uller and S.~Fredenhagen, Phys. Lett. {\bf B483}, 217 (2000).
2490:
2491: %\enlargethispage{0.5cm}
2492:
2493: \bibitem{bp96}
2494: W.~Buchm\"uller and M.~Pl\"umacher, Phys. Lett. {\bf B389}, 73 (1996).
2495:
2496: \bibitem{by99}
2497: W.~Buchm\"uller and T.~Yanagida, \pl{445}{1999}{399}
2498:
2499: \bibitem{Ghe}
2500: T.~Gherghetta and G.~Jungman, Phys.~Rev.~{\bf D48}, 1546 (1993).
2501:
2502: \bibitem{models}
2503: For recent discussions and references, see:
2504: R.~N.~Mohapatra, S.~Nasri and H.~Yu, Phys.~Lett.~{\bf B615}, 231 (2005);
2505: Z.-Z.~Xing, Phys.~Rev.~{\bf D70}, 071302 (2004);
2506: N.~Cosme, JHEP~{\bf 0408}, 027 (2004);
2507: W.~Rodejohann, Eur.~Phys.~J.~{\bf C32}, 235 (2004);
2508: W.~Grimus and L.~Lavoura, J.~Phys.~{\bf G30}, 1073 (2004);
2509: V.~Barger, D.~A.~Dicus, H.-J.~He and T.~Li, Phys.~Lett.~{\bf B583}, 173 (2004);
2510: P.~H.~Chankowski and K.~Turzy\'nski, Phys.~Lett.~{\bf B570}, 198 (2003);
2511: L.~Velasco-Sevilla, JHEP~{\bf 0310}, 035 (2003);
2512: E.~Kh. Akhmedov, M.~Frigerio and A.~Yu.~Smirnov, JHEP~{\bf 0309}, 021 (2003);
2513: G.~C. Branco, R.~Gonz\'alez~Felipe, F.~R.~Joaquim, I.~Masina, M.~N.~Rebelo
2514: and C.~A.~Savoy, Phys.~Rev.~{\bf D67}, 07025 (2003).
2515:
2516: \bibitem{KT} E.~W.~Kolb and M.~S.~Turner, {\bf The Early Universe}
2517: (Addison Wesley, New York, 1990).
2518:
2519: \bibitem{lut} M.~A.~Luty, \pr{45}{1992}{455}.
2520:
2521: \bibitem{plu}
2522: M.~Pl\"umacher, Z.~Phys.~{\bf C74}, 549 (1997).
2523:
2524: \bibitem{bdp02}
2525: W.~Buchm\"uller, P.~Di~Bari and M.~Pl\"umacher, \np{643}{2002}{367}.
2526:
2527: \bibitem{bcx00}
2528: R.~Barbieri, P.~Creminelli, A.~Strumia and N.~Tetradis, \np{575}{2000}{61}.
2529:
2530: \bibitem{bdp04}
2531: W.~Buchm\"uller, P.~Di~Bari and M.~Pl\"umacher,
2532: Ann.~Phys.~{\bf 315}, 303 (2005) [hep-ph/0401240].
2533:
2534: \bibitem{kw80}
2535: E.~W.~Kolb and S.~Wolfram, \np{172}{1980}{224}; \np{195}{1982}{542(E)}.
2536:
2537: \bibitem{pu04} A.~Pilaftsis and T.~E.~J.~Underwood, \np{692}{2004}{303}.
2538:
2539: \bibitem{gnx04} G.~F.~Giudice, A.~Notari, M.~Raidal, A.~Riotto and A.~Strumia,
2540: \np{685}{2004}{89}.
2541:
2542: \bibitem{crx98} L.~Covi, N.~Rius, E.~Roulet and F.~Vissani, \pr{57}{1998}{93}.
2543:
2544: \bibitem{bdp022} W.~Buchm\"uller, P.~Di~Bari, and M.~Pl\"umacher, \pl{547}{2002}{128}.
2545:
2546: %\enlargethispage{0.5cm}
2547:
2548: \bibitem{di} S.~Davidson and A.~Ibarra, \pl{535}{2002}{25}.
2549:
2550: \bibitem{hmy} K.~Hamaguchi, H.~Murayama and T.~Yanagida, \pr{65}{2002}{043512}.
2551:
2552: \bibitem{bdp2} W.~Buchm\"uller, P.~Di~Bari and M.~Pl\"umacher, \np{665}{2003}{445}.
2553:
2554: \bibitem{hlx} T.~Hambye, Y.~Lin, A.~Notari, M.~Papucci and A.~Strumia,
2555: \np{695}{2004}{169}.
2556:
2557: \bibitem{plu98}
2558: M.~Pl\"umacher, Nucl. Phys. {B530}, 207 (1998).
2559:
2560: \bibitem{db04}
2561: P.~Di~Bari, hep-ph/0406115.
2562:
2563: \bibitem{akx03} S.~Antusch, J.~Kersten, M.~Lindner and M.~Ratz,
2564: \np{674}{2003}{401}.
2565:
2566:
2567: \bibitem{bp} W. Buchm\"uller and M. Pl\"umacher, \pl{511}{2001}{74}.
2568:
2569: \bibitem{wet} G.~Lazarides, Q.~Shafi and C.~Wetterich, \np{181}{1981}{287};
2570: R.~N.~Mohapatra and G.~Senjanovi\'c, \pr{23}{1981}{165}; C.~Wetterich,
2571: \np{187}{1981}{343}.
2572:
2573: \bibitem{tri}
2574: T.~Hambye and G.~Senjanovi\'c, \np{582}{2004}{73};
2575: W.~Rodejohann, Phys.~Rev.~{\bf D70}, 073010 (2004);
2576: P.-H.~Gu and X.-J.~Bi, Phys.~Rev. {\bf D70}, 063511 (2004);
2577: G.~D'Ambrosio, T.~Hambye, A.~Hektor, M.~Raidal and A.~Rossi,
2578: Phys.~Lett.~{\bf B604}, 199 (2004).
2579:
2580: \bibitem{ak}
2581: S.~Antusch and S.~F.~King, Phys.~Lett~{\bf B597}, 199 (2004).
2582:
2583: \bibitem{fps} M.~Flanz, {\it et al.} and L.~Covi, {\it et al.}, in
2584: \cite{l-asymmetry}.
2585:
2586: \bibitem{rlm}
2587: S.~Dar, S.~Huber, V.~N.~Senoguz and Q.~Shafi, Phys.~Rev~{\bf D69}, 077701
2588: (2004);
2589: C.~H.~Albright and S.~M.~Barr, Phys.~Rev.~{\bf D70}, 033013 (2004);
2590: A. Pilaftsis, hep-ph/0408103.
2591:
2592: %\enlargethispage{0.5cm}
2593:
2594: \bibitem{rls}
2595: Y.~Grossman, T.~Kashti, Y.~Nir and E.~Roulet, Phys.~Rev.~Lett.~{\bf 91},
2596: 251801 (2003);
2597: G.~D'Ambrosio, G.~F.~Giudice and M.~Raidal, Phys.~Lett.~{\bf B575}, 75 (2003);
2598: T.~Hambye, J.~March-Russell and S.~M.~West, JHEP~{\bf 0407}, 070 (2004);
2599: L.~Boubekeur, T.~Hambye and G.~Senjanovi\'c, Phys.~Rev.~Lett.~{\bf 93},
2600: 111601 (2004);
2601: Y.~Grossman, T.~Kashti, Y.~Nir and E.~Roulet, JHEP~{\bf 0411}, 080 (2004);
2602: M.~C.~Chen and K.~T.~Mahanthappa, Phys.~Rev.~{\bf D70}, 113013 (2004);
2603: R.~Allahverdi and M.~Drees, Phys.~Rev.~{\bf D79}, 123522 (2004).
2604:
2605: \bibitem{FGY} P.~H.~Frampton, S.~L.~Glashow and T.~Yanagida,
2606: Phys.~Lett.~{\bf B548}, 119 (2002); see also T. Endoh {\it et al.},
2607: Phys.~Rev.~Lett.~{\bf 89}, 231601 (2002).
2608:
2609: \bibitem{KKM} M.~Kawasaki, K. Kohri and T.~Moroi,
2610: Phys.~Rev.~{\bf D71}, 083502 (2005).
2611:
2612: \bibitem{edm} P.~G.~Harris, {\it{et al.}},
2613: Phys.~Rev.~Lett.~{\bf 82}, 904 (1999).
2614:
2615: \bibitem{bound} V.~Baluni, Phys.~Rev.~{\bf D19}, 2227 (1979);
2616: R.~J.~Crewther, P.~di~Vecchia, G.~Veneziano and E.~Witten,
2617: Phys.~Lett.~{\bf B88}, 123 (1979) ; {\it ibid.} {\bf B91}, 487(E) (1980).
2618:
2619: \bibitem{RDPCP} See, for example, R.~D.~Peccei, in {\bf CP Violation}, p. 503
2620: ed. C.~Jarlskog (World Scientific, Singapore, 1989).
2621:
2622: \bibitem{PQ} R.~D.~Peccei and H.~R.~Quinn,
2623: Phys.~Rev.~Lett.~{\bf 38}, 1440 (1977); Phys.~Rev.~{\bf D16}, 1791 (1977).
2624:
2625: \bibitem{WW} S.~Weinberg, Phys.~Rev.~Lett.~{\bf 40}, 223 (1978);
2626: F.~Wilczek, Phys.~Rev.~Lett.~{\bf 40}, 271 (1978).
2627:
2628: \bibitem{Raffelt} See for example, G.~Raffelt,
2629: {\bf Stars as Laboratories for Fundamental Physics}
2630: (University of Chicago Press, Chicago, 1996).
2631:
2632: \bibitem{Cbounds} J.~Preskill, M.~Wise and F.~Wilczek,
2633: Phys.~Lett.~{\bf B120}, 127 (1983); L.~Abbott and P.~Sikivie,
2634: Phys.~Lett.~{\bf B120}, 133 (1983);
2635: M.~Dine and W.~Fischler, Phys.~Lett.~{\bf B120}, 137 (1983).
2636:
2637: %\enlargethispage{0.5cm}
2638:
2639: \bibitem{Sikivie} See, for example, P.~Sikivie,
2640: Nucl.~Phys.~Proc.~Suppl.~{\bf B87}, 41 (2000).
2641:
2642: \bibitem{SDM} See, for example, K.~A.~Olive, in {\bf Dark 2004}, Proc.
2643: 5th Int. Heidelberg Conf. on Dark Matter in Astro and Particle Physics,
2644: Springer Verlag, to be published. hep-ph/0412054.
2645:
2646: \bibitem{Coleman} S.~Coleman, Nucl.~Phys.~{\bf B262}, 263 (1985).
2647:
2648: \bibitem{DK} M.~Dine and A.~Kusenko, Rev.~Mod.~Phys. {\bf 76}, 1 (2004).
2649:
2650: \bibitem{KKT} S.~Kasuya, M.~Kawasaki and F.~Takahashi,
2651: Phys.~Rev.~{\bf D68}, 023501 (2003).
2652:
2653: \bibitem{FY02} M.~Fujii and T.~Yanagida, Phys.~Lett.~{\bf B542}, 80 (2002).
2654:
2655: \bibitem{Ara} J.~Arafune, {\it{ et al.}}, Phys.~Rev.~{\bf D62}, 105013 (2000).
2656:
2657: \bibitem {LPY} P.~Langacker, R.~D.~Peccei and T.~Yanagida,
2658: Mod.~Phys.~Lett.~{\bf A9}, 541 (1986).
2659:
2660: \bibitem{DFSZ} M.~Dine, W.~Fischler and M.~Srednicki,
2661: Phys.~Lett.~{\bf B104}, 199 (1981); A.~Zhitnitski,
2662: Sov.~Jour.~Nucl.~Phys.~{\bf 31}, 260 (1980).
2663:
2664: \bibitem{KSVZ} J.~E. ~Kim, Phys.~Rev.~Lett.~{\bf 43}, 103 (1979);
2665: M.~A.~Shifman, A.~I.~Vainshtein and V.~I.~Zakharov, Nucl.~Phys.~{\bf B166},
2666: 493 (1980).
2667:
2668: \bibitem{KL} M.~L.~Khlopov and A.~D.~Linde,
2669: Phys.~Lett.~{\bf B138}, 265 (1984); J.~R.~Ellis, J.~E.~Kim and
2670: D.~V.~Nanopoulos, Phys.~Lett.~{\bf B145}, 181 (1984).
2671:
2672: \bibitem{BBB} M.~Bolz, A.~Brandenburg and W.~Buchm\"uller,
2673: Nucl.~Phys.~{\bf B606}, 518 (2001).
2674:
2675: \bibitem{Falom} I.~V.~Falomkin, {\it et al.}, Nuovo~Cim.~{\bf A79}, 193 (1984)
2676: [Yad.~Fiz.~{\bf 39}, 990 (1984)].
2677:
2678: \bibitem{enhancement} A.~Pilafsis, Phys.~Rev.~{\bf D56}, 5431 (1997);
2679: J.~Ellis, M.~Raidal and T.~Yanagida, Phys.~Lett.~{\bf B546}, 228 (2002).
2680:
2681:
2682:
2683: \bibitem{BBP}
2684: M.~Bolz, W.~Buchm\"uller and M.~Pl\"umacher,
2685: Phys.~Lett.~{\bf B443}, 209 (1998).
2686:
2687: \bibitem{FIY}
2688: M.~Fujii, M.~Ibe and T.~Yanagida, Phys.~Lett.~{\bf B579}, 6 (2004);
2689: J.~R.~Ellis, K.~A.~Olive, Y.~Santoso and V.~C.~Spanos,
2690: Phys.~Lett.~{\bf B588 }, 7 (2004);
2691: J.~L.~Feng, S.~Su and F.~Takayama, Phys.~Rev.~{\bf D70}, 075019 (2004).
2692:
2693: \bibitem{RRA}
2694: L.~Roszkowski, R.~Ruiz~de~Austri and K.-Y.~Choi,
2695: JHEP~{\bf 0508}, 080 (2005) [hep-ph/0408227].
2696:
2697: \bibitem{frt03}
2698: J.~L.~Feng, A.~Rajaraman and F.~Takayama, Phys.~Rev.~Lett.~{\bf 91}, 011302-1
2699: (2003).
2700:
2701: \bibitem{bhr03}
2702: W.~Buchm\"uller, K.~Hamaguchi and M.~Ratz, Phys.~Lett.~{\bf B574}, 156 (2003).
2703:
2704: \bibitem{viel}
2705: M.~Viel {\it et al.}, Phys.~Rev.~{\bf D71}, 063534 (2005).
2706:
2707: \bibitem{FIY2} M.~Fujii and T.~Yanagida, Phys.~Lett.~{\bf B549}, 273 (2002);
2708: M.~Fujii, M.~Ibe and T.~Yanagida, Phys.~Lett.~{\bf B579}, 6 (2004).
2709:
2710: \bibitem{ay00}
2711: T.~Asaka and T.~Yanagida, Phys.~Lett.~{\bf B494}, 297 (2000).
2712:
2713: \bibitem{GGW}
2714: T.~Gherghetta, G.~F.~Giudice and J.~D.~Wells, Nucl.~Phys.~{\bf B559}, 27
2715: (1999).
2716:
2717: \bibitem{IKMY} See, for an explicit model, M.~Ibe, R.~Kitano, H.~Murayama
2718: and T.~Yanagida, Phys.~Rev.~{\bf D70}, 075012 (2004).
2719:
2720: \bibitem{BM} F.~Borzumati and A.~Masiero, Phys.~Rev.~Lett.~{\bf 57},
2721: 961 (1986); J.~Hisano, T.~Moroi, K.~Tobe, M.~Yamaguchi and T.~Yanagida,
2722: Phys.~Lett.~{\bf B357}, 579 (1995).
2723:
2724: \bibitem{di01}
2725: S.~Davidson and A.~Ibarra, JHEP~{\bf 0109}, 013 (2001).
2726:
2727: \bibitem{MS} For a recent discussion see, for example,
2728: I.~Masina and C.~A.~Savoy, Nucl.~Phys.~Proc.~Suppl. {\bf B143}, 70 (2005)
2729: [hep-ph/0410382].
2730:
2731: \bibitem{Shafi}
2732: G.~Lazarides and Q.~Shafi, Phys.~Lett.~{\bf B258}, 305 (1991);
2733: G.~Lazarides, C.~Panagiotakopoulos and Q.~Shafi, Phys.~Lett.~{\bf B315}, 325
2734: (1993).
2735:
2736: \bibitem{KMY}
2737: K.~Kumekawa, T.~Moroi and T.~Yanagida, Progr. Theor. Phys. {\bf 92},
2738: 437 (1994); T.~Asaka, K.~Hamaguchi, M.~Kawasaki and T.~Yanagida,
2739: Phys.~Lett.~{\bf B464}, 12 (1999); Phys.~Rev.~{\bf D61},
2740: 083512 (2000); G.~F.~Giudice, M.~Peloso, A.~Riotto and
2741: I.~Tkachev, JHEP~{\bf 08}, 014 (1999).
2742:
2743: \bibitem{MSYY}
2744: H.~Murayama, H.~Suzuki, T.~Yanagida and J.~Yokoyama,
2745: Phys.~Rev.~Lett.~{\bf 70}, 1912 (1993);
2746: J.~R.~Ellis, M.~Raidal and T.~Yanagida, Phys.~Lett.~{\bf B581}, 9 (2004).
2747:
2748: \bibitem{MrYa} H.~Murayama and T.~Yanagida, Phys.~Lett.~{\bf B322}, 349
2749: (1994); M.~Dine, L.~Randall and S.~Thomas, Nucl.~Phys.~{\bf B458}, 291
2750: (1996).
2751:
2752: \bibitem{MrYa2}
2753: K.~Hamaguchi, H.~Murayama and T.~Yanagida,
2754: Phys.~Rev.~{\bf D65}, 043512 (2002).
2755:
2756: %\enlargethispage{0.5cm}
2757:
2758: \bibitem{guth} A.~H.~Guth, Phys.~Rev.~{\bf D23}, 347 (1981).
2759:
2760: \bibitem{density-fluctuations}
2761: A.~H.~Guth and So-Y.~Pi, Phys.~Rev.~Lett.~{\bf 49}, 1110 (1982);
2762: S.~Hawking, Phys.~Lett.~{\bf B115}, 295 (1982); A.~A.~Starobinsky,
2763: Phys.~Lett.~{\bf B117}, 175 (1982).
2764:
2765: \bibitem{susycp}
2766: L.~Covi {\it et al.} in \cite{l-asymmetry}.
2767:
2768: \bibitem{linde} A.~Linde, Phys.~Lett.~{\bf B129}, 177 (1983);
2769: A.~Linde, Phys.~Scripta~{\bf T117}, 40 (2005).
2770:
2771: \bibitem{GL}
2772: A.~S.~Goncharov and A.~D.~Linde, Phys.~Lett.~{\bf B139}, 27 (1984);
2773: H.~Murayama, H.~Suzuki, T.~Yanagida and J.~Yokoyama,
2774: Phys.~Rev.~{\bf D50}, R2356 (1994).
2775:
2776:
2777: \bibitem{Q-ball2} A.~Kusenko and M.~Shaposhnikov, Phys.~Lett.~{\bf B418},
2778: 46 (1998).
2779:
2780: \bibitem{FHY} T.~Asaka, M.~Fujii, K.~Hamaguchi and T.~Yanagida,
2781: Phys.~Rev.~{\bf D62}, 123514 (2000); M.~Fujii, K.~Hamaguchi and
2782: T.~Yanagida, Phys.~Rev.~{\bf D63}, 123513 (2001) and references
2783: therein.
2784:
2785: \end{thebibliography}
2786:
2787:
2788: \end{document}
2789:
2790:
2791:
2792:
2793:
2794:
2795:
2796:
2797:
2798:
2799:
2800:
2801:
2802:
2803: