hep-ph0112158/KS1.tex
1: \documentclass[notoc,paper]{JHEP}
2: %
3: \usepackage{amssymb}
4: \usepackage{cite}
5: \usepackage{graphicx}
6: %
7: % -------------------- definitions ----------------------------
8: %
9: \newcommand{\nco}{\newcommand}
10: %
11: \def\phm{\phantom{-}}
12: %
13: \nco{\half}{\frac{1}{2}}
14: \nco{\shalf}{\ensuremath{\textstyle \frac{1}{2}}}
15: \newcommand{\ev}{\ensuremath{\mathrm{\; eV}}}
16: \newcommand{\mev}{\ensuremath{\mathrm{\; MeV}}}
17: 
18: %
19: % -------------------- title page ----------------------------
20: %
21: 
22: \title{Oscillation induced neutrino asymmetry growth in
23:        the early Universe}
24: 
25: 
26: \author{Kimmo Kainulainen \\ NORDITA \\Blegdamsvej 17
27:         \\ DK-2100 Copenhagen \O \\Denmark
28:         \\e-mail: \email{kainulai@nordita.dk}}
29: 
30: \author{Antti Sorri \\ Department of Physical Sciences
31:         \\ University of Helsinki
32:         \\ P.O. Box 64 \\ FIN-00014 University of Helsinki
33:         \\ e-mail: \email{antti.sorri@helsinki.fi}}
34: 
35: 
36: \abstract{We study the dynamics of active-sterile neutrino
37: oscillations in the early universe using full momentum-dependent
38: quantum-kinetic equations. These equations are too complicated to
39: allow for an analytical treatment, and numerical solution is greatly
40: complicated due to very pronounced and narrow structures in the
41: momentum variable introduced by resonances. Here we introduce a
42: novel dynamical discretization of the momentum variable which
43: overcomes this problem. As a result we can follow the evolution
44: of neutrino ensemble accurately well into the stable growing phase.
45: Our results confirm the existence of a ``chaotic region" of mixing
46: parameters, for which the final sign of the asymmetry, and hence
47: the SBBN prediction of $^4$He-abundance cannot be accurately
48: determined.}
49: 
50: \keywords{Physics of the Early Universe, Neutrino Physics }
51: \preprint{NORDITA 2001/76 HE}
52: 
53: %
54: % -------------------- the article ----------------------------
55: %
56: 
57: \begin{document}
58: 
59: \section{Introduction}
60: 
61: 
62: Recent observations from atmospheric and solar neutrinos from Super
63: Kamiokande (SK)\cite{SK1} and Sudbury Neutrino Observatory (SNO)
64: \cite{SNO} prefer conventional, mainly active-active mixings between
65: ordinary electron, muon and tau neutrinos as the solutions to the
66: observed flux deficits. However, while pure active-sterile solutions
67: are disfavoured, sizable admixtures of sterile states in the observed
68: fluxes are still allowed~\cite{smirnov}. New sterile states are also
69: needed in order to explain the LSND anomaly~\cite{LosAl} by neutrino
70: physics. Moreover, in a recent analysis of the so called
71: ``2+2''-models\footnote{Alternative ``3+1'' models fit less well
72: with short baseline reactor experiments~\cite{3p1schemes}.}
73: for four-neutrino mixing, the solutions with nonzero sterile
74: neutrino components were found to provide best global fits to the
75: solar, atmospheric and reactor data~\cite{stfluxes}. Corresponding
76: limits on active-sterile mixings are quite generous, for example
77: %
78: \begin{eqnarray}
79: \sin^2\theta_{\mu s} &\lesssim 0.48& \qquad \rm (Atmospheric) \nonumber\\
80: \sin^2\theta_{e s}   &\lesssim 0.72& \qquad \rm (Solar \;\; LMA),
81: \label{labfluxes}
82: \end{eqnarray}
83: %
84: where LMA refers to the Large Mixing Angle solution for the solar
85: neutrino deficit.
86: 
87: Active-sterile mixing, if realized, would have several interesting
88: effects in astrophysical settings~\cite{snova} and in particular for
89: the evolution of the early universe
90: \cite{dol,ektdo,ekm,ektBig,ektat,ekmL,old,ftv,shi,eks1,fSc,Sorri,shifuNc}.
91: In particular, sterile states could be brought into thermal equilibrium
92: by mixing before nucleosynthesis, so that the resulting anomalous increase
93: in the expansion rate of the universe would lead to overproduction of helium
94: in disagreement with the observations~\cite{dol,ektdo,ekm,ektBig,ektat,old}.
95: Excluding the parameter sets leading to equilibration provides useful bounds
96: on active-sterile mixing. Indeed, from results of ref.~\cite{ektBig} one
97: can infer that the sterile components in mass eigenstates responsible
98: for atmospheric anomaly and LMA are constrained by
99: %
100: \begin{eqnarray}
101: \sin^2\theta_{\mu s} &\lesssim 0.013& \qquad \rm (Atmospheric) \nonumber\\
102: \sin^2\theta_{e s}   &\lesssim 0.026& \qquad \rm (Solar \;\; LMA)
103: \label{cosmofluxes}
104: \end{eqnarray}
105: %
106: where we used a SBBN-limit of $N_\nu \lesssim 3.4$ for the number of
107: effective neutrino degrees of freedom~\cite{ektBig,recent}. These
108: numbers correspond to the ``light" case where the sterile component is
109: the heavier of the mixing states; in the opposite, ``dark" case, the
110: corresponding limits are even stronger by one to two orders of magnitude.
111: In either case the cosmological constraints are more stringent than
112: those obtained in terrestrial laboratories.
113: 
114: While quite generic, the cosmological bounds (\ref{cosmofluxes})
115: do depend on some simple prior assumptions, the most important of
116: which is that the primordial lepton asymmetry is not anomalously
117: large\cite{ekmL}.  It may be possible to
118: circumvent them in more complicated mixing schemes involving more
119: sterile states (for a recent discussion, see~\cite{dibSc}). In a
120: particular attempt it was shown in ref.~\cite{ftv} that for a large
121: negative $\delta m^2$ and a small enough mixing angle (so that
122: the equilibration bounds of~\cite{ektBig} can be avoided), say on
123: the $\nu_\tau-\nu_{s'}$-sector, resonant oscillations may trigger
124: a rapid growth of the tau lepton asymmetry. This asymmetry could then
125: become very large, $L_\tau \sim {\cal O}(1)$ and change the natural
126: prior condition of no significant initial lepton asymmetry for a
127: mixing in any other sectors.  Indeed, if created early enough, such
128: an asymmetry could suppress $\nu_\mu-\nu_s$-oscillations~\cite{fvL1,ekmL},
129: obviating the bounds (\ref{cosmofluxes}) for $\nu_\mu-\nu_s$-mixing.
130: However, as this scenario requires that the sterile state is the
131: {\em lighter} one (resonance condition), and that the mass splitting
132: $|\delta m^2|$ is large (to create $L_{\tau}$ early enough), it is
133: essentially excluded by the recent solar and atmospheric neutrinos
134: combined with the direct constraint on the electron neutrino
135: mass~\cite{mainz}.
136: 
137: Even with the mechanism of ref.~\cite{ftv} by and large excluded,
138: there can be important effects due to $L$-growth with smaller
139: $|\delta m^2|$.  For example, a large homogeneous electron neutrino
140: asymmetry $L_{\nu_e}$ would directly influence the nucleosynthesis
141: prediction for the helium-4 abundance through the reactions
142: %
143: \begin{eqnarray}
144:  n + e^+  &\leftrightarrow & p + \bar \nu_e \nonumber \\
145:  p + e^-  &\leftrightarrow & n + \nu_e \;,
146: \end{eqnarray}
147: %
148: which keep the neutron-to-proton ratio in equilibrium at early times.
149: Changing the electron neutrino abundance could tilt the balance of these
150: reactions with the possibility of either increasing $Y_{He}$ (negative
151: $L_{\nu_e}$), and hence strengthening the bounds (\ref{cosmofluxes}), or
152: decreasing $Y_{He}$ (positive $L_{\nu_e}$), leading to weakening of the
153: bounds (\ref{cosmofluxes}).
154: 
155: The physics involved with the resonant growth of the asymmetry is very
156: complicated because of several vastly different physical scales both
157: in the temporal direction and in momentum variable, which effect the
158: evolution of the ensemble in an essential way. The relevant QKE's are
159: nonlinear and strongly coupled internally via the asymmetry term, and
160: therefore no useful analytical approximation exists for the
161: problem\footnote{
162: Some confusion in the field was caused by a recent analytical
163: treatment~\cite{dolHan}, whose results contradicted earlier numerical
164: results~\cite{shi,eks1,fSc}. Disagreement was eventually clarified in
165: favour of numerical work in~\cite{DiBetal}.}.
166: Numerical solution is also greatly complicated by the various different
167: physical scales, and
168: while the phenomenon of asymmetry growth is fairly well established
169: by now, the determinacy of the final {\em sign} of the asymmetry is
170: still not well understood. The ambiguity was first observed in~\cite{shi},
171: and in~\cite{eks1,Sorri} it was shown that this ``chaoticity" occurs in
172: certain well defined region of mixing parameters.
173: References~\cite{shi,eks1,Sorri} employed a numerical solution for the
174: momentum averaged approximation of the QKE's however, and their results
175: were challenged by ref.~\cite{dolHan}, who claimed that the final
176: asymmetry is always fully determined by the initial conditions. While
177: the analysis of ref.~\cite{dolHan} is now discredited~\cite{DiBetal},
178: it still remains true that no numerically reliable momentum dependent
179: analysis has so far shown the existence of the chaotic
180: region.\footnote{
181: Although ref.~\cite{fdb} found support for chaoticity,
182: their conclusion is not firm because of the reported loss of the
183: numerical accuracy of their methods when approaching the potentially
184: chaotic region.}
185: It is important to settle the issue because, as mentioned above, large
186: $L_{\nu_e}$ could significantly alter the helium-4 abundance. Indeed,
187: if the sign of $L_{\nu_e}$ were found to be chaotic, then $Y_{He}$
188: could not be reliably determined from BBN~\cite{eks1}.
189: 
190: In this paper we study the dynamics of active-sterile neutrino
191: oscillations in the early universe using full momentum-dependent
192: quantum-kinetic equations (in homogeneous space). Our main technical
193: improvement over the previous works is the introduction of a novel
194: dynamical discretization method for the momentum variable which
195: enables us to model accurately the density matrices over the entire
196: momentum range, including the extremely pronounced and narrow
197: structures close to resonant momenta.  On physical context, our
198: results confirm the existence of a "chaotic" region of mixing
199: parameters, for which the final sign of the asymmetry is not
200: deterministic. It also confirms the expectation~\cite{eks1,fdb}, that
201: the size of the chaotic region is somewhat smaller than indicated by
202: the momentum averaged code~\cite{eks1}. These results can be seen as
203: giving more strength on the cosmological constraints on neutrino
204: mixing parameters.
205: 
206: The paper is organized as follows. In section 2 we set up the quantum
207: kinetic equations for the problem. In section 3 we introduce the novel
208: dynamically adjusted discretization of the momentum variable and in
209: section 4 we set up our kinetic equations with this parametrization.
210: The method we develop allows us to have enough resolution to solve the
211: problem with relatively small number of momentum bins (we use at most
212: 400 bins).  In section 5 we display our numerical results for the key
213: variables driving the oscillation and discuss the numerical stability
214: of our solutions.  Finally, in section 6 we present our results on the
215: asymmetry growth and oscillations as well as the interpretation of
216: the physical consequences, and the section 7 contains a summary and
217: outlook.
218: 
219: 
220: \section{Quantum Kinetic Equations}
221: 
222: In the early Universe the oscillating neutrinos experience frequent
223: scatterings which interrupt the coherent evolution of the state and
224: introduce collisional mixing between different momentum states. The
225: mathematical formalism which can incorporate all these features has
226: been developed elsewhere~\cite{stod,ekm,ektBig,sigl,TMcK}, and it takes
227: the form of the quantum kinetic equations for the reduced density
228: matrices for the neutrino and antineutrino ensembles. We parametrize
229: the density matrices by the Bloch vector presentation~\cite{ektBig}:
230: %
231: \begin{equation}
232: \rho_{\nu}      \equiv \frac{1}{2} f_0 (P_0
233:           + {\bf P} \cdot {\bf\sigma}) \; , \qquad
234: \rho_{\bar \nu} \equiv \frac{1}{2} f_0 (\bar P_0
235:           + {\bf \bar P} \cdot {\bf \sigma}),
236: \label{rho}
237: \end{equation}
238: %
239: where $\mathbf{\sigma}$ are the Pauli spin matrices and
240: $f_0 = (1+\exp(p/T))^{-1}$ is the Fermi-Dirac distribution
241: function without chemical potential. Then to the the lowest
242: order in the interaction~\cite{stod,ekm,sigl,TMcK}, the
243: QKE's for the density matrix take the form
244: %
245: \begin{eqnarray}
246: d_t P_0
247:      &=& \phm \frac{\Gamma}{f_0} \left[ f_{eq}
248:     - \rho_{\alpha \alpha} \right] \nonumber \\
249: d_t P_x
250:      &=& -    V_z P_y - D P_x \nonumber \\
251: d_t P_y
252:      &=& \phm V_z P_x - V_x P_z - D P_y \nonumber \\
253: d_t P_z
254:      &=& \phm V_x P_y + \frac{\Gamma}{f_0} \left[ f_{eq}
255:                                - \rho_{\alpha \alpha} \right] \, ,
256: \label{eq:master}
257: \end{eqnarray}
258: %
259: where $d_t \equiv \partial_t - H p \partial_p$ and $H$ is the
260: Hubble expansion factor. The rotation vector $\mathbf{V}$ has
261: the following components ($V_y \equiv 0$):  in $x$-direction
262: one has only the vacuum contribution
263: %
264: \begin{equation}
265:     V_x = \frac{\delta m^2}{2 p} \sin 2 \theta_0 \;,
266: \end{equation}
267: %
268: whereas in the $z$-direction also the matter induced effective
269: potential contributes:
270: %
271: \begin{equation}
272:     V_z  = V_0  + V_1  + V_L,
273: \label{vzeta}
274: \end{equation}
275: %
276: where
277: %
278: \begin{eqnarray}
279:   V_0 \phm       &=& - \frac{\delta m^2}{2 p} \cos 2 \theta_0, \\
280:   V_1^e \phm     &=& - 14 \sqrt{2} \frac{\zeta(4)}{\zeta(3)}
281:                              \frac{G_F}{M_W^2} N_{\gamma} p \; T
282:                    \left( 1 + {\textstyle \frac{1}{4}} \cos^2 \theta_W
283:                    \left[ n_{\nu_e} + n_{\bar \nu_e } \right] \right) \\
284:   V_1^{\mu,\tau} &=& - \frac{7 \sqrt{2}}{2} \frac{\zeta(4)}{\zeta(3)}
285:                              \frac{G_F}{M_Z^2} N_{\gamma} p \; T
286:                         \left[ n_{\nu_{\mu,\tau}}
287:                         + n_{\bar \nu_{\mu,\tau} } \right] \\
288:   V_L \phm       &=& \phm \sqrt{2} G_F N_{\gamma} L^{(\alpha)}.
289: \end{eqnarray}
290: %
291: Here $N_\gamma$ is the photon equilibrium number density,
292: $n_{\nu_{\alpha,(\bar\alpha)}}$ is the normalized to unity (in
293: equilibrium) actual neutrino (antineutrino) number density, and the
294: effective neutrino asymmetries $L^{(\alpha)}$ are given by
295: %
296: \begin{eqnarray}
297:     L^{(e)}    &=&  \left( \shalf + 2 \sin^2 \theta_W \right) L_e
298:                   + \left( \shalf - 2 \sin^2 \theta_W \right) L_p
299:                   - \shalf L_n + 2 L_{\nu_e}
300:                   + L_{\nu_{\mu}} + L_{\nu_\tau} \\
301:     L^{(\mu)}  &=&  L^{(e)} - L_e -L_{\nu_e} + L_{\nu_{\mu}} \\
302:     L^{(\tau)} &=&  L^{(e)} - L_e -L_{\nu_e} + L_{\nu_{\tau}} .
303: \end{eqnarray}
304: %
305: where $L_f \equiv (n_f - n_{\bar f})N_f/N_\gamma$.
306: 
307: The repopulation term $\Gamma (f_{eq} - \rho_{\alpha \alpha})$ in the
308: diagonal part of (\ref{eq:master}) is written in the relaxation time
309: approximation~\cite{aussiet}. The distribution $f_{eq}$ in the
310: repopulation term is the usual equilibrium Fermi-Dirac distribution
311: %
312: \begin{equation}
313:     f_{eq} = \frac{1}{1 + e^{\frac{p}{T}-\frac{\mu}{T}}}.
314: \end{equation}
315: %
316: and the reaction rates are
317: \begin{equation}
318:     \Gamma = C_\alpha G_F^2 p T^4
319: \end{equation}
320: with $C_e \simeq 1.27$ and $C_{\mu,\tau} \simeq 0.92$\cite{ektBig}.
321: The damping terms $D P_i$ correspond to the decohering scatterings,
322: which tend to destroy the coherence of the oscillation. In density
323: matrix language such terms appear as relaxation terms suppressing
324: the off-diagonals of the density matrix, or equivalently the
325: $P_{x,y}$-components of the Bloch vectors. The magnitude of the
326: decohering terms is just half of the corresponding scattering rate
327: %
328: \begin{equation}
329:     D = \half \Gamma.
330: \end{equation}
331: %
332: The equation of motion for anti-neutrinos can be found by substituting
333: $L^{(\alpha)} = -L^{(\alpha)}$ and $\mu_{\nu} = - \mu_{\nu}$ to the
334: above equations (the latter condition is true when neutrinos are in
335: chemical equilibrium).
336: 
337: The repopulation term used above is only an approximation for the
338: correct elastic collision integral, and unfortunately it breaks the
339: lepton number conservation. This is not a serious problem however,
340: and it can be circumvented by introducing an explicit lepton number
341: conserving evolution equation for $L^{(\alpha)}$.  The appropriate
342: equation, introduced in~\cite{fdb}, is
343: %
344: \begin{equation}
345:   d_t L^{(\alpha)}
346:          = \frac{1}{8\zeta(3)}
347:            \int\limits_0^\infty dp \;
348:                     p^2 (V_x P_y - \bar V_x \bar P_y).
349: \label{L-equation}
350: \end{equation}
351: %
352: Equation~(\ref{L-equation}) can be derived from the original equations
353: in the approximation where the $L$-violation due to the the approximate
354: repopulation term is neglected. So, when using (\ref{L-equation})
355: to evolve $L$, it is guaranteed that the repopulation term does not
356: {\em directly} affect the lepton asymmetry generation. The accuracy
357: of the method can be monitored by comparing the value of $L^{(\alpha)}$
358: derived from (\ref{L-equation}) and from directly integrating the
359: neutrino distribution functions. In practice the agreement was always
360: to be good.
361: 
362: Neutrino and antineutrino sectors are extremely strongly coupled
363: through the asymmetry $L^{(\alpha)}$ in equations~(\ref{eq:master}). In
364: order to compute $L^{(\alpha)}$ accurately enough despite the numerical
365: round-off errors, it is then necessary to avoid computing $L^{(\alpha)}$
366: through differences of large quantities, such as naively appear in the
367: equation (\ref{L-equation}).  To this end we will define the ``small"
368: and ``large" linear combinations of dynamical variables in particle
369: and antiparticle sector
370: %
371: \begin{equation}
372: P_i^\pm = P_i \pm \bar P_i .
373: \label{Pplusminus}
374: \end{equation}
375: %
376: For convenience we also separate active and sterile sectors
377: by defining
378: %
379: \begin{eqnarray}
380: P_a^\pm & = P_0^\pm + P_z^\pm & = 2 \rho_{\alpha \alpha}^\pm/f_0
381: \label{eq:newvar1}\\
382: P_s^\pm & = P_0^\pm - P_z^\pm & = 2 \rho_{ss}^\pm/f_0.
383: \label{eq:newvar2}
384: \end{eqnarray}
385: %
386: In terms of variables (\ref{eq:newvar1}-\ref{eq:newvar2}) the
387: equations take the form
388: %
389: \begin{eqnarray}
390: d_t P_a^\pm & = & \phm  V_x P_y^\pm + \Gamma  \left[2 f_{eq}^\pm/f_0
391:                       - P_a^\pm \right] ,
392: \label{aktiivi} \\
393: d_t P_s^\pm & = &   -   V_x P_y^\pm  ,
394: \label{steriili} \\
395: d_t P_x^\pm & = &   -  (V_0 + V_1) P_y^\pm  - V_L P_y^\mp - D P_x^\pm , \\
396: d_t P_y^\pm & = & \phm (V_0 + V_1) P_x^\pm  + V_L P_x^\mp
397:                   - \half V_x (P_a^\pm - P_s^\pm)  - D P_y^\pm ,
398: \label{eq:plusminus}
399: \end{eqnarray}
400: %
401: where we assumed $\Gamma = \bar \Gamma$ and defined
402: \begin{equation}
403: f_{eq}^\pm  =  f_{eq} (p,\mu) \pm f_{eq} (p,-\mu)
404: \end{equation}
405: %
406: Because $\bar V_x = V_x$, the equation for the asymmetry now reads
407: simply
408: %
409: \begin{equation}
410:   d_t L^{(\alpha)}
411:          = \frac{1}{8\zeta(3)}
412:            \int\limits_0^\infty dp \; p^2 V_x P_y^-.
413: \label{L-equation2}
414: \end{equation}
415: %
416: Finally, due to the particular form of the differential operator
417: $d_t = \partial_t - Hp\partial_p$, it is possible to reduce our
418: set of partial differential equations to ordinary differential
419: equations by changing variables $p \rightarrow x \equiv p/T$, and
420: $t \rightarrow T$, and correspondingly using the time-temperature
421: relation
422: %
423: \begin{equation}
424:  \frac{dT}{dt} = -HT.
425: \end{equation}
426: %
427: The differential operator then becomes simply
428: %
429: \begin{equation}
430:  d_t \rightarrow  - \frac{1}{HT} \partial_T.
431: \end{equation}
432: %
433: Let us stress that our introducing the small and large linear
434: combinations is an absolutely necessary step for obtaining reliable
435: solutions to
436: QKE's~(\ref{aktiivi}-\ref{eq:plusminus},\ref{L-equation2}).
437: This can best be appreciated from the figure (\ref{fig:rplus}) below,
438: where we show the energy spectrum in neutrino {\em and} antineutrino
439: sectors for a representative set of oscillation parameters; the
440: asymmetry corresponds to the integral over the difference (not even
441: visible!) of the neutrino and antineutrino distributions displayed.
442: 
443: %
444: %  SECTION:  Discretization of the momentum variable
445: %
446: \section{Discretization of the momentum variable}
447: %
448: 
449: While necessary, introducing parametrization (\ref{Pplusminus}) is
450: unfortunately not sufficient to tackle the momentum dependent QKE's.
451: Additional difficulties arise due to structures in the momentum
452: direction (in variable $x$) and in particular the very narrow ones
453: introduced by the resonances. In this paragraph we explain step by
454: step how we discretize the momentum variable such that these
455: difficulties can be overcome.
456: 
457: First, the momentum variable ranges over the semi-infinite range
458: $x \in [0,\infty]$, but in practice we will introduce cut-offs to both
459: ends of spectrum. First, the ultra-relativistic neutrino approximation
460: breaks down when $x\rightarrow 0$, formally appearing as a singularity
461: in at $x=0$ in our equations.  Second, it is useful to cut out very
462: large $x$, in the exponentially suppressed tail of the distribution.
463: So, we restrict $x$ to a range $x \in [x_{min},x_{max}]$, where in
464: practice it suffices to use\footnote{Note that even for
465: $x=10^{-4}$ still $p \gtrsim 1$ keV throughout our computation, so
466: that the ultrarelativistic approximation always remains valid.}
467: $x_{min} = 10^{-4}$ and $x_{max} = 100$.
468: 
469: Second, the variable $x$ is not ideal for discretization, because
470: most of the variation in the spectrum occurs at small $x \lesssim 3$.
471: We therefore have mapped the variable $x$ to $u(x)$ belonging to
472: range $[0,1]$ by a transformation
473: %
474: \begin{equation}
475:   u(x) = \frac{x(1+\epsilon_2 ) - x_{\rm ext} \epsilon_1 }{x+x_{\rm ext}}.
476: \label{mapping}
477: \end{equation}
478: %
479: where $\epsilon_1 \equiv k_1(1+\epsilon_2)$ and
480: $\epsilon_2 \equiv (1+k_1)/(k_2-k_1)$
481: with $k_1 \equiv x_{\rm min}/x_{\rm ext}$ and
482: $k_2 \equiv x_{\rm max}/x_{\rm ext}$.
483: Apart from a small correction due to the cut-off parameters
484: (\ref{mapping}) is designed so that the extremum point of the thermal
485: momentum distribution $x_{ext}$ gets mapped to $u \simeq 1/2$.
486: We will also need the inverse of this function, which is
487: %
488: \begin{equation}
489:    x(u) = \frac{x_{\rm ext}(u+\epsilon_1)}{1+\epsilon_2-u}.
490: \label{imapping}
491: \end{equation}
492: %
493: While clearly a significant improvement for binning the momentum
494: variable, (\ref{mapping}) is not clever enough to solve the numerical
495: problems arising from the the pronounced structures around the
496: resonant momenta. Fortunately, the resonance positions $x_{r_1}$ and
497: $x_{r_2}$ can be solved analytically from the conditions $V_z = 0$
498: (neutrinos) and $\bar V_z = 0$ (antineutrinos):
499: %
500: \begin{eqnarray}
501:     x_{r_1} &=& \sqrt{\varphi^2  + \chi} - \varphi
502:          \qquad \bar \nu{\rm -resonance}
503:     \label{eq:rescond1} \\
504:     x_{r_2} &=& \sqrt{\varphi^2  + \chi} + \varphi
505:          \qquad \nu{\rm -resonance}
506:     \label{eq:rescond2}
507: \end{eqnarray}
508: %
509: where in the case of $\nu_\mu$ and $\nu_\tau$ neutrinos
510: %
511: \begin{eqnarray}
512:     \varphi &=& \frac{\zeta(3) M_Z^2 L^{(\alpha)}}{14 \zeta(4) T^2}
513:     \approx 6.60 \times 10^8 L^{(\alpha)} T^{-2}_\mev  \\
514:     \chi    &=& \frac{\zeta(3) M_Z^2 |\delta m^2|
515:         \cos 2\theta_0}{14 \sqrt{2} \zeta(4)G_F N_{\gamma} T^3}
516:     \approx 1.64 \times 10^{8} |\delta m^2|_{\ev^2}
517:         \cos 2 \theta_0 T^{-6}_\mev.
518: \end{eqnarray}
519: %
520: Using (\ref{mapping}) one easily finds the corresponding resonant
521: values in the $u$-variable.
522: 
523: The resonance widths can be estimated from the distance over which
524: the matter mixing angle drops to a some fraction of its maximum value
525: of unity.  In this way one finds that the relative resonance width is
526: proportional to the vacuum mixing angle:
527: %
528: \begin{equation}
529: \frac{\Delta x}{x} \sim \tan 2\theta_0.
530: \label{reswidth}
531: \end{equation}
532: %
533: From (\ref{reswidth}) it is clear that for small mixing angles the
534: structures near resonances can only be resolved by an extremely
535: finely spaced momentum grid. For example\footnote{One is driven
536: to such small values of vacuum mixing to find acceptable phenomenology
537: for the asymmetry growth~\cite{ftv,fvSc}.}
538: with $\sin^22\theta_0 = 10^{-6}$, equation (\ref{reswidth}) predicts
539: structures in the scale $\Delta x/x \simeq 10^{-3}$. Assuming then
540: that at least 50 points are needed to describe the resonant region
541: accurately, one sees that a linear binning of the $x$-variable with
542: $x_{max}=100$ would require $5 \times 10^6$ grid points! Solving such
543: problem is clearly beyond capabilities of any computer we have today.
544: 
545: Our mapping (\ref{mapping}) helps the situation by roughly a factor
546: of hundred, but the remaining problem would still be way too large.
547: We therefore need to find a second transformation $u \rightarrow u(v)$
548: that would redistribute the points such that more grid points are
549: clustered around the resonances in the physical momentum variable.
550: There are obviously many ways to implement such a transformation. We
551: found it most convenient to construct it by using polynomials of the
552: form $a + b (v-v_r)^3$. Because we have two resonance points the
553: mapping function is a little complicated:
554: %
555: \begin{equation}
556: u(v) = \alpha v + (a + b(v-v_{r_1})^3)\, \theta(v_c-v)
557:         + (c + d(v-v_{r_2})^3)\, \theta(v-v_c),
558: \label{eq:v2u}
559: \end{equation}
560: %
561: where the two separate mappings around resonant values $v_{r_1}$
562: and $v_{r_2}$ are joined smoothly at the geometric mean of the
563: resonances $v_c \equiv \frac{1}{2}(v_{r_1} + v_{r_2})$. The linear
564: term $\alpha v$ was added in order to have a control how steep
565: is the distribution function of the grid points.
566: 
567: The mapping (\ref{eq:v2u}) has six free parameters $a$, $b$, $c$,
568: $d$, $v_{r_1}$ and $v_{r_2}$, which will be determined from boundary
569: conditions at $v=0$ and $v=1$ and continuity conditions at $v=v_c$.
570: First, the constants $a$, $b$, $c$ and $d$ are fixed by the boundary
571: conditions
572: %
573: \begin{eqnarray}
574:     u(0)       &=& 0, \\
575:     u(v_{r_1}) &=& u_{r_1}, \\
576:     u(v_{r_2}) &=& u_{r_2}, \\
577:     u(1)       &=& 1,
578: \end{eqnarray}
579: %
580: which results in
581: %
582: \begin{eqnarray}
583:     a &=& u_{r_1} - \alpha v_{r_1} \\
584:     b &=& \frac{u_{r_1}}{v_{r_1}^3} - \frac{\alpha}{v_{r_1}^2} \\
585:     c &=& u_{r_2} - \alpha v_{r_2} \\
586:     d &=& \frac{1-u_{r_2}}{(1-v_{r_2})^3} - \frac{\alpha}{(1-v_{r_2})^2}.
587: \end{eqnarray}
588: %
589: Note that $u_{r_i}$ are not unknowns, but can be solved analytically
590: from the resonance conditions (\ref{eq:rescond1}-\ref{eq:rescond2}) and
591: the inverse mapping (\ref{imapping}). Finally, the continuity conditions
592: %
593: \begin{eqnarray}
594:  \lim_{v \rightarrow v_c +} & u(v) &
595:                 = \lim_{v \rightarrow v_c -} u(v) \\
596:  \lim_{v \rightarrow v_c +} & \partial_v u(v) &
597:                 = \lim_{v \rightarrow v_c -} \partial_v u(v).
598: \end{eqnarray}
599: %
600: provide implicit equations for the remaining two unknowns $v_{r_i}$:
601: %
602: \begin{eqnarray}
603:         a + \frac{1}{8} b (v_{r_2} - v_{r_1})^3
604:     &=& c + \frac{1}{8} d (v_{r_1} - v_{r_2})^3
605:     \label{eq:match.cond1} \\
606:     b &=& d.
607:     \label{eq:match.cond2}
608: \end{eqnarray}
609: %
610: Unfortunately these are sixth order algebraic equations and can only
611: be solved by iterative numerical algorithms.
612: 
613: \begin{figure}
614: \centering
615: \includegraphics[height=7.5cm]{param.eps}
616:     \caption{The mapping $u=u(v)$ is shown with  $L^{(\alpha)} =
617:     4 \times 10^{-7}$, $T=11 \mev$ (solid line) and with
618:     $L^{(\alpha)} = 10^{-10}$, $T=15 \mev$ (dashed line).
619:     Mixing parameters are $\delta m^2 = -0.1 \ev^2$
620:     and $\sin^2 2 \theta_0 = 10^{-9}$. }
621:     \label{fig:param}
622: \end{figure}
623: 
624: \begin{figure}
625: \centering
626: \includegraphics[height=7.5cm]{param2.eps}
627:     \caption{Density of points (defined as the inverse of the step
628:     size) in the physical momentum variable $x$ for the transforms
629:     shown in figure~(\ref{fig:param}).}
630:     \label{fig:param2}
631: \end{figure}
632: 
633: The novel feature of the above change of variables is that it
634: automatically follows the resonances, placing majority of the grid
635: points dynamically in their vicinity. As an alternative to our approach,
636: one could think of using library routines with automatic grids to take
637: care of this distribution.  However, the resonance positions evolve as
638: a function of expansion scale of the universe and in particular as a
639: function of $L^{(\alpha)}$.  High point densities are thus needed at
640: different momenta at different times, and this movement can be fast.
641: None of the library routines we tried came close to being able to
642: cope with these requirements and failed completely to solve the
643: problem.  With our parametrization on the other hand, the numerical
644: program can be written for a fixed grid in the variable $v$, and the
645: redistribution of points is provided by the dynamical equations
646: themselves.
647: 
648: In Fig.~(\ref{fig:param}) we show the transformations $u \rightarrow v$
649: when $L^{(\alpha)}$ is almost zero (dotted line), and when $L^{(\alpha)}$
650: has grown to a substantial value (solid line). In the former case one
651: cannot distinguish the two resonances by eye, whereas in the second
652: case the structures have become very distinct. Fig.~(\ref{fig:param2})
653: shows the density of points as a function of $x$ (defined as the
654: inverse of separation of grid points in the $x$-variable).
655: 
656: %
657: %
658: \section{Changes in the QKE's due to the parametrization}
659: %
660: 
661: The changes of variables (\ref{mapping}) and (\ref{eq:v2u})
662: affect the partial derivatives appearing in
663: Eqns.~(\ref{aktiivi}-\ref{eq:plusminus},\ref{L-equation2}).
664: These equations have been written assuming the evolution of
665: $\rho$ and $L$ is along constant $x$ contours. However, we have
666: now fixed our grid in the variable $v$, and hence need to find the
667: evolution equations with $T$ and $v$ as the independent variables.
668: These are easily found from the old ones by the chain rule
669: partial differentiation
670: %
671: \begin{eqnarray}
672: \Big( \frac{\partial \rho\left(T,x(T,v)\right)}{\partial T} \Big)_v
673:     &=&
674:         \Big( \frac{\partial \rho}{\partial T} \Big)_x
675:     +   \Big(\frac{\partial x}{ \partial T} \Big)_v \,
676:         \Big( \frac{\partial \rho}{\partial x} \Big)_T \nonumber \\
677:     &=&
678:         \Big( \frac{\partial \rho}{\partial T} \Big)_x
679:     +   \Big( \frac{\partial x}{\partial T } \Big)_v
680:         \Big( \frac{\partial u}{\partial x} \Big)_T
681:         \Big( \frac{\partial v}{\partial u} \Big)_T \,
682:         \Big( \frac{\partial \rho}{\partial v} \Big)_T \nonumber \\
683:     &=&
684:         \Big( \frac{\partial \rho}{\partial T} \Big)_x
685:     +   \Big( \frac{\partial u}{\partial T } \Big)_v
686:         \Big( \frac{\partial v}{\partial u} \Big)_T \,
687:         \Big( \frac{\partial \rho}{\partial v} \Big)_T \,.
688:     \label{eq:master.changed}
689: \end{eqnarray}
690: %
691: In the last equality we used the fact that $u=u(x)$ only. The
692: differentials $(\partial_T \rho)_x$ are of course just given by the l.h.s.\
693: of equations (\ref{eq:plusminus}), and the last coefficient $(\partial_v \rho)_T$
694: is computed numerically as a part of the system of the partial differential
695: equations. To be precise, we are using the central derivative formula
696: %
697: \begin{equation}
698: \left( \frac{\partial \rho}{\partial v} \right)_T
699:                   = \frac{\rho(v+h) - \rho(v-h)}{2h}.
700: \end{equation}
701: %
702: Of the remaining differentials $(\partial_v u)_T$ is  also easily solved
703: from the parametrization Eq.~(\ref{eq:v2u}:
704: %
705: \begin{equation}
706:     \left( \partial_v u \right)_T
707:     =   \frac{1}{\alpha + 3[ b(v-v_{r_1})^2 \, \theta(v_c-v)
708:     +   d(v-v_{r_2})^2 \, \theta(v-v_c)]}.
709:     \label{eq:dvdu}
710: \end{equation}
711: %
712: However, solving the remaining differential $\left(\partial_T u \right)_v $
713: is somewhat more complicated. Using equation~(\ref{eq:v2u}) one can write
714: %
715: \begin{eqnarray}
716:     \left(\partial_T u \right)_v
717:     &=& \phm \left[ \partial_T a + \partial_T b (v-v_{r_1})^3
718:     -   3b (v-v_{r_1})^2 \partial_T v_{r_1} \right] \theta(v_c - v)
719:     \nonumber \\ &&
720:     +   \left[ \partial_T c + \partial_T d (v-v_{r_2})^3
721:     -   3d (v-v_{r_2})^2 \partial_T v_{r_2} \right] \theta(v-v_c),
722: \end{eqnarray}
723: %
724: where
725: %
726: \begin{eqnarray}
727:     \partial_T a &=& \partial_T u_{r_1} - \alpha \partial_T v_{r_1} \\
728:     \partial_T b &=& \frac{1}{v_{r_1}^3}
729:                    \left( \partial_T u_{r_1}
730:                     +\Big(2 \alpha - \frac{3 u_{r_1}}{v_{r_1}}\Big)
731:                                         \partial_T v_{r_1} \right) \\
732:     \partial_T c &=& \partial_T u_{r_2} - \alpha \partial_T v_{r_2}\\
733:     \partial_T d &=& - \frac{1}{(1 - v_{r_2})^3}
734:                    \left(\partial_T u_{r_2}
735:                    + \Big(2 \alpha - 3\frac{1-u_{r_2}}{1-v_{r_2}}\Big)
736:                                          \partial_T v_{r_2} \right).
737: \end{eqnarray}
738: %
739: Here the only unknowns are the temperature derivatives of the resonances
740: in the $v$-variable $\partial_T v_{r_i}$.  These can be solved from our
741: original matching conditions (\ref{eq:match.cond1}) and
742: (\ref{eq:match.cond2}), which can formally be written as
743: %
744: \begin{eqnarray}
745:     f_1 \left(v_{r_1}(T), v_{r_2}(T), T \right) &=& 0
746: \label{ekuf1} \\
747:     f_2 \left(v_{r_1}(T), v_{r_2}(T), T \right) &=& 0.
748: \label{ekuf2}
749: \end{eqnarray}
750: %
751: Differentiating equations (\ref{ekuf1}-\ref{ekuf2}) with respect to $T$
752: results in a {\em linear} set of equations for $\partial_T v_{r_i}$, which
753: after a little algebra can be brought to the form
754: %
755: \begin{equation}
756:     \partial_T v_{r_i} = U_{ij} \partial_T u_{r_j},
757: \label{partialTv}
758: \end{equation}
759: %
760: where the matrix $U_{ij}$ depends both on $u_{r_i}$ and
761: $v_{r_i}$.  Note that despite the apparent complexity of our
762: parametrization, we have been able to reduce everything except solving
763: for the resonance parameters $v_{r_i}$ to a simple linear algebra. Solving
764: $v_{r_i}$ from the matching equations at each time step would be very
765: time consuming however. This is so in particular because we need to
766: solve them with very high accuracy in order for not to lose the
767: continuity of the matching at $v=v_c$ and to avoid inducing random
768: numerical errors into the equations.  Fortunately, the equation
769: (\ref{partialTv}) itself offers the way out of this problem. Namely,
770: we only need to solve $v_{r_i}$ from the matching conditions
771: (\ref{eq:match.cond1}) and (\ref{eq:match.cond2}) once, in the
772: beginning of the iteration, as their value can later be {\em evolved}
773: by equations (\ref{partialTv}) as the solution proceeds. In this way
774: even solving for the parametrization from $v$ to $x$ becomes part of
775: the dynamical equations.  The accumulation of numerical error can be
776: easily traced by solving the matching conditions independently for
777: some temperatures. In practice the errors do not accumulate at all
778: and it turns out that using the differential equation is by far the
779: superior method for solving $v_{r_i}$ in comparison to using the
780: matching conditions at each temperature step.
781: 
782: %
783: %
784: \section{Numerical Results}
785: %
786: 
787: We have solved the evolution equations numerically for a range of
788: representative parameter sets. In figure (\ref{fig:rplus}) we show the
789: neutrino and antineutrino distribution functions in a relatively early
790: stage in the evolution for $\delta m^2 = -10 \; \rm eV^2$ and
791: $\sin^22\theta_0=10^{-6.1}$. The distributions show a very nice thermal
792: shape, and no visible difference between particle and antiparticle
793: sectors can be seen. This is so despite the fact that at the temperature
794: at which the distributions were plotted, the momenta $p \simeq 1.8T$ are
795: currently resonant (indicated by the gray line in the plot); the
796: amplitude of resonant structure is simply too small to be seen in
797: the scale of the ``large variables".
798: 
799: \begin{figure}
800: \centering
801: \includegraphics[height=7.5cm]{rhoplus.eps}
802:     \caption{Example of neutrino ($\rho_{\alpha \alpha}$) and
803:      antineutrino ($\rho_{\bar \alpha \bar \alpha}$) spectra as a
804:      function of $p/T$. Lines fall on top of each others. Gray
805:      line marks the resonance momenta.}
806:     \label{fig:rplus}
807: \end{figure}
808: 
809: The resonance does alter the asymmetry spectrum however. In figure
810: (\ref{fig:llx}) we show the distribution of the difference of neutrino and
811: antineutrino densities $\rho_{\alpha \alpha}-\rho_{\bar \alpha \bar \alpha}$
812: in the physical variable $x$. The prominent and extremely narrow resonance
813: structure (one cannot separate the neutrino and antineutrino resonances
814: here) is clearly visible.  The same spectrum is shown in the integration
815: variable $v$ in figure (\ref{fig:llv}). Now the sharp kink-like resonant
816: structure has broadened into a broad wave, which is easily represented
817: by the linearly discretized function in variable $v$ (grid points are
818: shown by the open circles). In this example we used just 400 grid points,
819: but essentially identical results were found with only 200 points.
820: Moreover, comparing the scales in figures~(\ref{fig:rplus}-\ref{fig:llx}),
821: one sees that the scale of variation of $L(x)$ is ten orders of magnitude
822: below unity. So, in order to follow the evolution of $L^{(\alpha)}$
823: accurately when using the original variables, one should be able to
824: integrate over distributions with an accuracy much better than ten
825: digits!  This seems totally impossible, and clearly indicates that
826: our use of variables $P^\pm$ is essential.
827: 
828: \begin{figure}
829: \centering
830: \includegraphics[height=7.5cm]{llx.eps}
831:     \caption{Difference of neutrino and antineutrino densities
832:     $\rho_{\alpha \alpha} - \rho_{\bar \alpha \bar \alpha}$
833:     as a function of $p/T$ for the same parameters as in
834:     figure~{\ref{fig:rplus}}.}
835:     \label{fig:llx}
836: \end{figure}
837: 
838: \begin{figure}
839: \centering
840: \includegraphics[height=7.5cm]{llv.eps}
841:     \caption{Same as figure~(\ref{fig:llx}), but shown against
842:     variable $v$.}
843:     \label{fig:llv}
844: \end{figure}
845: 
846: These examples prove the success of the key elements in our approach,
847: namely that by separating the small and large variables, and by
848: introducing the novel dynamical discretization of the momentum grid,
849: we can model the evolution of the ensemble of oscillating neutrinos
850: accurately.  Moreover, we can do so using small enough lattices such
851: that the code can be run even in a powerful modern workstation
852: (although a single run may then take as long as a couple of days).
853: 
854: \subsection{Regulating the sterile neutrino spectrum}
855: 
856: As one lets the system evolve further the resonance $x$ moves to
857: larger values while the integrated $L^{(\alpha)}$ decreases. This
858: continues until one reaches roughly the point when the main bulk of
859: the active neutrino distribution is getting through the resonance.
860: At this point one expects that the total asymmetry begins to oscillate
861: or grow, depending on the parameters. In our initial runs we did not
862: get that far however, but the program crashed due to an accumulation
863: of structures in sterile neutrino spectrum with momenta less than
864: $x_{res}T$.  How these structures arise is easy to understand.
865: Indeed, the active sector distributions are kept smooth away from
866: the resonances by diffusion induced by collisions with the background
867: particles. Sterile states interact only through the effective
868: interaction due to mixing however, and since this interaction
869: is neglibly small away from the resonances, the diffusion does not
870: work on sterile sector. As a result the moving resonance leaves
871: the sterile neutrino momentum distribution quite oscillatory for
872: $x < x_{res}$. While these developments in the sterile sector have
873: little effect on variables in the active sector (because active
874: and sterile states are essentially decoupled away from resonances),
875: they still pose a numerical problem: as the resonance moves forward
876: in $x$, the density of points reduces drastically in the region
877: $x\ll x_{res}$, until the point that the grid becomes too sparse
878: to model the oscillatory structures in the unsmoothed sterile
879: neutrino distribution. Eventually the amplitudes of these structures
880: overflow causing numerical instability and the solution breaks down.
881: 
882: The above explanation of the instability also suggests an obvious
883: cure for the problem. If one is not interested in studying the sterile
884: neutrino asymmetry spectrum, all we need to do is to add a small
885: regulatory interaction term to sterile neutrino sector, which will
886: smooth their distribution away from resonances, and yet does not
887: alter the evolution of the variables in the active sector. The latter
888: requirement can be met by a suitable form of the interaction term and
889: by making it small enough so that it does not change the physics in
890: the resonance region. The challenge is to do this so that the
891: regulatory term remains efficient enough to numerically stabilize the
892: system. To this end we have added the following ``sterile repopulation
893: terms" onto equation (\ref{steriili}):
894: %
895: \begin{eqnarray}
896:     R_{\nu_s}^+ &=&
897:       r_s \Gamma \left(n_{\nu_s}f_{eq}^+(x,\mu)
898:                      - \rho_{ss}^+  \right)
899: \label{rplusa} \\
900:     R_{\nu_s}^-  &=&
901:       r_s \Gamma \left( \rho_{aa}^{\rm init}
902:                      - \rho_{aa}^{-} - \rho_{ss}^- \right).
903: \label{regulators}
904: \end{eqnarray}
905: %
906: Here $\rho_{aa}^{\rm init}$ is the initial density matrix in the
907: active sector. We used the active sector interaction rate $\Gamma$
908: above, and let the parameter $r_s$ control the size of the regulatory
909: term.  The regulators (\ref{rplusa}-\ref{regulators}) differ from the
910: repopulation terms in the active sector in (\ref{aktiivi}) in that we
911: have added normalization factors to make sure that the integrals
912: over $R_{\nu_s}^\pm$ always vanish. (In this way we avoid spurious
913: sterile state equilibration even with large control parameter $r_s$.)
914: 
915: We obviously need to show that there exists a range of values
916: for which the results converge and become independent of $r_s$, while
917: the regulator still continues to stabilize the solution.  That we do
918: reach this goal is evidenced in Fig.~(\ref{fig:regulator}), where we
919: show the evolution of the total asymmetry for $r_s=0.1$, 0.01, 0.001,
920: $10^{-4}$ and $10^{-5}$, for $\delta m^2 = -0.1 \ev^2$ and
921: $\sin^2 2 \theta_0 = 10^{-6.1}$.
922: Unsurprisingly, for large $r_s$ there is a significant
923: effect, but for small $r_s$ the results do converge to the point that
924: no visible difference can be seen between the two curves corresponding
925: to two smallest $r_s$'s.  Let us stress that the mixing parameters
926: were here deliberately chosen such as to show a particularly large
927: sensitivity for the regulator.  For many other cases that we studied
928: it would have been difficult to show any effect at all; for example
929: the results displayed in figure~(\ref{fig:stable}) showed no
930: appreciable dependence on the regulator up to $r_s = 10$.
931: 
932: \begin{figure}
933:   \centering
934:   \includegraphics[height=7.5cm]{regulator.eps}
935:   \caption{The stability of asymmetry evolution against variation of
936:     parameter $r_s$. We have used values $r_s =0.0001$ (solid line),
937:     0.001 (dashed line), 0.01 (dotted line) and 0.1 (dash-dotted line).
938:     Solutions for values $r_s=10^{-4}$ and $r_s=10^{-5}$ (long dashed line)
939:     are already
940:     indistinguishable in the figure. Oscillation parameters are
941:     $\delta m^2 = -0.1 \ev^2$ and $\sin^2 2 \theta_0 = 10^{-6.1}$.}
942:   \label{fig:regulator}
943: \end{figure}
944: 
945: We have also studied the stability of our results against changes in
946: various numerical error handling parameters in our code. The stepwise
947: error tolerance for example can be relaxed by an order of magnitude
948: from what we actually employed, before any visible deviation from the
949: solution can be seen over the entire integration range. This is true also
950: for our examples corresponding to the oscillatory solutions in or
951: near the chaotic region.  Of course, the numerical accuracy would be
952: eventually compromised for parameters for which the asymmetry undergoes
953: very large number of oscillations before settling to the growing
954: curve, but as in~\cite{eks1}, that does not affect our main conclusions.
955: Also cut off parameters $x_{min}$ and $x_{max}$ and the tilt $\alpha$
956: which controls the density of points at the resonance can be varied
957: within reasonable limits without any observable effects on the results.
958: We therefore are confident that our observed pattern of asymmetry
959: oscillations is indeed physical, and not of numerical origin.
960: 
961: \subsection{Argument for chaoticity in $L$-growth}
962: 
963: Having introduced a method that can handle the evolution of momentum
964: dependent density matrices and in particular the asymmetry, and having
965: proven its numerical stability and accuracy, we now pursue the question
966: of the ``chaotic" behaviour of the final sign of the integrated asymmetry.
967: Note that while the use of word chaos is customary in this context, we
968: strictly speaking mean only sensitivity of the final sign of the asymmetry
969: on initial seed asymmetry and on oscillation parameters\footnote{
970: In fact the system {\em does} exhibit true chaoticity at a certain
971: level, but the related information loss is very small in
972: practice~\cite{Steen}.}.
973: 
974: In the context of our earlier work using the momentum averaged
975: equations~\cite{eks1,Sorri}, we explained what was the physical origin
976: for the sign sensitivity. Quite simply, changes in the boundary conditions
977: (initial baryon asymmetry and the oscillation parameters) change the
978: direction of the motion of the system, and the topography of the
979: attractor solutions in the phase space at the onset of resonance. Roughly
980: speaking, in the momentum averaged case one can draw an analogy between the
981: system and a somewhat sticky ball initially rolling down a single valley
982: floor that later branches to two, with a low maximum in between, at the
983: resonance point. Changing baryon asymmetry would then be analogous to
984: changing the speed and direction of the ball at the branching point and
985: changing oscillation parameters to that of changing the shape of the
986: valleys themselves. Sign sensitivity arises for those initial conditions
987: and valley forms for which the ball makes many oscillations between the
988: two degenerate valleys before stickiness (damping) eventually wins and
989: makes it to settle into either one of them.
990: 
991: In the momentum dependent case these structures become smeared out, and
992: to draw a mechanical analogy, one should rather think, instead of a ball,
993: of a string of varying size beads (largest ones in the middle, corresponding
994: to the peak in the thermal distribution) coupled by a very elastic cord,
995: wiggling down the valleys described above.  While this is obviously a
996: much more complicated system with qualitatively new features, one would
997: still expect to see oscillations in the weighted average position of beads
998: at least when the largest ones roll past the branching point (in $L$ when
999: the maximum of the distribution crosses the resonance). Moreover, if the
1000: average position of the string ($L$) makes many oscillations over the
1001: central hill before stickiness (damping) wins, one again should expect
1002: that the final valley that the string settles in (sign of $L$) strongly
1003: depends on initial and boundary conditions. While this analog is certainly
1004: bears only a crude resemblance to the true system, it should help convince
1005: the reader of the fact that there is nothing mysterious about the sign
1006: sensitivity in the $L$-evolution, but that it is rather something to be
1007: expected.
1008: 
1009: Obviously, all we need to do to prove the existence of chaoticity, is
1010: to find examples of almost identical sets of oscillation parameters for
1011: which the final asymmetry does show significantly different oscillation
1012: pattern, and leads to a different final sign. We show such a case in
1013: figure (\ref{fig:osc}). All three curves in the figure correspond to
1014: $\delta m^2 = -10 \ev^2$ but have slightly different vacuum mixing
1015: angles. The two curves which behave almost alike, have mixing angles
1016: $\sin^2 2 \theta_0 = 10^{-6.1}$ (solid curve) and $\sin^2 2 \theta_0
1017: = 10^{-6.099}$ (dotted curve) respectively. The dash-dotted curve
1018: with a very different behaviour corresponds to $\sin^2 2 \theta_0
1019: = 10^{-6.09}$. So, if such mixing were ever observed and the mixing
1020: angle was measured to a precision better than a tenth of a per cent,
1021: the final sign of the lepton asymmetry and its effect on nucleosynthesis
1022: could be predicted from our computation.  However, if the mixing angle
1023: was measured with less than about one per cent accuracy, we would not
1024: be able to say what the final asymmetry will be, and therefore what is
1025: the precise BBN prediction for helium abundance.
1026: 
1027: %
1028: \begin{figure}
1029:     \centering
1030:     \includegraphics[height=7.5cm]{osc.eps}
1031:     \caption{Neutrino asymmetry oscillations are shown near the boundary
1032:     of "chaotic" region using same $\delta m^2 = -10 \ev^2$ and three
1033:     slightly different mixing angles $\sin^2 2 \theta_0 = 10^{-6.1}$
1034:     (solid line) $\sin^2 2 \theta_0 = 10^{-6.099}$ (dotted line) and
1035:     $\sin^2 2 \theta_0 = 10^{-6.09}$ (dash-dotted line).}
1036:     \label{fig:osc}
1037: \end{figure}
1038: %
1039: 
1040: The sign sensitivity does not occur for all oscillation parameters,
1041: however. In the figure (\ref{fig:stable}) we show results from ``stable"
1042: region of parameters (same mass difference as above, but much smaller
1043: mixing angle). In this case, while visible differences in the evolution
1044: become visible when mixing angle is changed by a factor or three, the
1045: asymmetry never becomes oscillatory, and the final sign is determined
1046: by the initial sign of the asymmetry.
1047: 
1048: \begin{figure}
1049:   \centering
1050:   \includegraphics[height=7.5cm]{stable.eps}
1051:   \caption{Asymmetry evolution in the stable region of oscillation
1052:   parameters.  All curves corresponds to mass splitting $\delta m^2
1053:   = -0.1\;\rm eV^2$ and the mixing angles are $\sin^22\theta_0 =
1054:   10^{-9}$ (solid curve), $\sin^2 2\theta_0 = 10^{-8.9}$ (dotted curve)
1055:   and $\sin^2 2\theta_0 = 10^{-8.5}$ (dashed curve).}
1056:   \label{fig:stable}
1057: \end{figure}
1058: 
1059: Our first example was chosen from close to the edge of the chaotic
1060: region, and just as making $\sin^22\theta_0$ smaller made system more
1061: stable, increasing it has the opposite effect. The number of oscillations
1062: and also degree of sensitivity of the final sign to the oscillation
1063: parameters increases very rapidly with $\sin^22\theta_0$. Nevertheless,
1064: the degree of sensitivity to parameters is weaker and thus the chaotic
1065: region smaller than what was found in the momentum averaged
1066: treatment~\cite{eks1}. Quantitatively our new results agree roughly
1067: with the boundaries of the chaotic region inferred in~\cite{fdb}, but
1068: since the numerical work is, even with the improvements we have made,
1069: very time consuming, we will not attempt a precise mapping of the
1070: chaotic region here.
1071: 
1072: Let us finally to try to understand a little more deeply what precisely
1073: drives the oscillations of the total asymmetry in the momentum dependent
1074: case.  A clue can be found from Figure~(\ref{fig:py}) which shows the
1075: variable $P_y^-$ in the region where the asymmetry oscillates significantly.
1076: (At earlier times, when the total $L$ does not oscillate, the structure
1077: of the $P_y^-$-spectrum is similar to that of $\rho_{\alpha\alpha}^-$ shown
1078: in Fig.\ (\ref{fig:llv}).)  The two new peaks seen outside the resonance
1079: grow slowly in time, while their sign oscillates rapidly. Because these
1080: structures are asymmetric around the resonance (in the physical variable
1081: $x$), the integral over $P_y^\pm(x)$ is nonzero and oscillates. As a
1082: result also $L$, which is solved from (\ref{L-equation2}), becomes
1083: oscillatory.
1084: 
1085: What causes these structures, is a delicate interplay between the
1086: coherence length and the matter oscillation length. The former is simply
1087: given by the inverse of the damping rate $\ell_{\rm coh} = 1/D$ and the
1088: latter is
1089: %
1090: \begin{equation}
1091: \ell_{\rm m} = \frac{l_{\rm vac}}{\sqrt{1-2 \chi\cos2\theta_0 + \chi^2}}
1092: \end{equation}
1093: %
1094: where $l_{\rm vac} \equiv 4\pi p/|\delta m^2|$ and
1095: $\chi \equiv 2p|V|/|\delta m^2|$ where $|V|$ is the magnitude of
1096: the matter contribution to the $V_z$ in equation (\ref{vzeta}).
1097: Near the resonance the matter oscillation length is very long, and in
1098: particular at high temperatures it exceeds the coherence
1099: length by a large margin.  In such a situation the resonance is strongly
1100: {\em overdamped}, and despite the maximal mixing angle oscillations
1101: do not have time to develop because the state is continuously projected
1102: to the active direction. This in part explains why the resonant features
1103: have so modest amplitudes at early times (cf.\ Fig.\ (\ref{fig:llx})).
1104: 
1105: %
1106: \begin{figure}
1107:     \centering
1108:     \includegraphics[height=7.5cm]{py.eps}
1109:     \caption{Lower panel: Variable $P_y^-$ as a function of $v$ at
1110:             temperature $T=27.8 \mev$ for oscillation parameters
1111:             $\delta m^2 = -10 \ev^2$, $\sin^2 2\theta_0 = 10^{-6.099}$.
1112:             Upper panel: The coherence length (dashed line) and
1113:             $l_{\rm m}/4$ (solid line) for the same parameters.
1114:             Physical momentum at the resonance is $x_{\rm res} = 1.88$.}
1115:     \label{fig:py}
1116: \end{figure}
1117: 
1118: In the upper panel of the figure (\ref{fig:py}) we have plotted
1119: the coherence length and $\ell_{\rm m}/4$, which
1120: corresponds to the distance over which $P_y^\pm$ first rises to
1121: maximum due to oscillation. It is clear that the new structures
1122: correspond to the case when the matter oscillations first time
1123: break the overdamping situation.
1124: Further away from the resonance the damping no more can block the
1125: oscillations, but since the matter mixing angle decreases very fast
1126: away from the resonance, the amplitude of the oscillation features
1127: dies off quickly. As the temperature decreases, the new structures move
1128: away from the resonance, and they also become broader (one can show
1129: that their separation scales like $\delta x \sim T^{-3}$). This is why
1130: their amplitude slowly increases in time (with decreasing $T$) causing
1131: larger and larger amplitude oscillations in the total asymmetry
1132: $L$.  The lesson to bring home from these observations is that the
1133: dynamics of $L$-growth and oscillations is very dependent on the
1134: detailed fine structure of the various components of the density
1135: matrix.  Such fine details on the hand can only be studied in a
1136: numerical approach using both separated small and large variables
1137: ($P_i^\pm$) and dynamically adjusting momentum grids.
1138: 
1139: %
1140: \section{Conclusions}
1141: 
1142: Active-sterile mixing, while constrained by the present solar
1143: and atmospheric data, are nevertheless required if one wishes to
1144: incorporate the LSND-neutrino anomaly into the neutrino models.
1145: Moreover, the active-sterile neutrino mixing would have a rich
1146: phenomenology in the early universe, where it provides a unique
1147: theoretical challenge in the form of a system for which the
1148: quantum effects play an essential role in the kinetic equations.
1149: 
1150: Here we have studied the phenomenon of neutrino asymmetry growth in
1151: the early universe as a result of active sterile neutrino oscillations.
1152: Our main new contribution is the introduction of a novel method of
1153: discretizing the momentum variable such that the sharply pronounced
1154: structures near the resonances, which are here shown to drive the
1155: oscillation phenomena, can be treated numerically accurately.
1156: The only concession to rigor we made along the way to solution of
1157: the problem, was adding collision terms to regulate the sterile
1158: neutrino spectrum away from the resonances. However, we demonstrated
1159: that our regulators had no effect on the active neutrino evolution,
1160: and hence for the results presented here. Nevertheless, our treatment
1161: is obviously not adequate if the precise form of the sterile neutrino
1162: spectrum is important.  This might be the case for the models where
1163: sterile neutrinos are invoked to provide a non-thermal component for
1164: the dark matter~\cite{nonThDM}.
1165: 
1166: We have demonstrated that even tiny changes in the oscillation
1167: parameters may drastically alter the oscillation pattern, and even
1168: change the sign of the final asymmetry.  This behaviour does not
1169: occur for all oscillation parameters, but instead the dependence of
1170: the sign of $L$ on oscillation parameters is weak in the ``stable"
1171: region (in general small $\sin^22\theta_0$ and large negative
1172: $\delta m^2$) and strong in the chaotic region (in general large
1173: $\sin^22\theta_0$ and small negative $\delta m^2$). These results are in
1174: qualitative agreement with our earlier findings, which were based on
1175: a momentum averaged treatment~\cite{eks1}.  Obviously, if
1176: active-sterile mixing were to be observed with parameters residing
1177: in the chaotic region, it would be, due to inevitable errors in the
1178: experimentally measured parameters, impossible to reliably determine
1179: the sign of the neutrino asymmetry created by that mixing in the
1180: early universe.  Moreover, since large neutrino asymmetries (and
1181: electron neutrino asymmetry in particular) directly affect, in a
1182: $sign(L)$-dependent way, the weak interaction rates that
1183: determine proton-to-neutron ratio in the early universe, this
1184: indeterminacy would undermine our ability to accurately compute the
1185: BBN prediction for light element abundances for chaotic oscillation
1186: parameters.
1187: 
1188: These conclusions hold in the spatially homogeneous calculation.
1189: However, the sensitivity of $L$-growth on initial conditions also
1190: lends to a speculation that in reality the initially inhomogeneous
1191: seed asymmetry might give rise to an inhomogeneous texture of domains
1192: of very large lepton asymmetries with oscillating sign. This phenomenon
1193: has been shown to occur in a one-dimensional, momentum averaged
1194: model~\cite{eks2}, and could plausibly occur in a realistic three
1195: dimensional world.  The implications of such a scenario for SBBN would
1196: obviously be very different, including the possibility of very efficient
1197: equilibration of sterile neutrinos via MSW-effect within the domain
1198: boundaries~\cite{shiFuDo,eks2}. If this was the case, then the asymmetry
1199: growth mechanism would lead to even {\em stronger} bounds on mixing than
1200: what is displayed in equations~(\ref{cosmofluxes}).  It is beyond the scope
1201: of this work (and the reach of the present computers) to study this
1202: phenomenon quantitatively in the momentum dependent case, however.
1203: 
1204: Let us finally comment on the effects of large homogeneous lepton
1205: asymmetries for cosmic microwave background radiation (CMBR). The
1206: possibility of measuring $L$ by the Planck satellite data has been
1207: considered for example in~\cite{Kinney:1999pd} and
1208: in~\cite{CMBconstraints}.  However, the only effect of $L$ on CMBR
1209: comes through the associated fluctuations in the energy density,
1210: and correspondingly only the total asymmetry has relevance.
1211: In the oscillation scenarios, such as discussed in this paper,  the
1212: total asymmetry (here the sum of the sterile and active sector
1213: asymmetries) is conserved however, and hence the oscillation-induced
1214: asymmetries, in contrast to the situation with SBBN, would have
1215: no  direct effect on CMBR.
1216: 
1217: 
1218: \acknowledgments
1219: 
1220: We thank Kari Enqvist for useful conversations and a collaboration
1221: at earlier stages of this project. We also thank Steen Hannestad
1222: for discussions on the effects of $L$ on CMBR. AS wishes to thank
1223: Nordita for hospitality during several visits in the course of
1224: completing this project.
1225: 
1226: %
1227: %---------------------------- Journals----------------------------------
1228: %
1229: 
1230: \begin{thebibliography}{99}
1231: 
1232: \bibitem{SK1}
1233:     S.~Fukuda {\it et al.}  [Super-Kamiokande Collaboration],
1234:     %``Constraints on neutrino oscillations using 1258 days of  Super-Kamiokande solar neutrino data,''
1235:     Phys.\ Rev.\ Lett.\  {\bf 86} (2001) 5656
1236:     [arXiv:hep-ex/0103033]; \\
1237:     %%CITATION = HEP-EX 0103033;%%
1238:     S.~Fukuda {\it et al.}  [Super-Kamiokande Collaboration],
1239:     %``Tau neutrinos favored over sterile neutrinos in atmospheric muon  neutrino oscillations,''
1240:     Phys.\ Rev.\ Lett.\  {\bf 85} (2000) 3999
1241:     [arXiv:hep-ex/0009001].
1242:     %%CITATION = HEP-EX 0009001;%%
1243: 
1244: \bibitem{SNO}
1245:     Q.~R.~Ahmad {\it et al.}  [SNO Collaboration],
1246:     %``Measurement of the charged of current interactions produced by B-8  solar neutrinos at the Sudbury Neutrino Observatory,''
1247:     Phys.\ Rev.\ Lett.\  {\bf 87} (2001) 071301
1248:     [arXiv:nucl-ex/0106015].
1249:     %%CITATION = NUCL-EX 0106015;%%
1250: 
1251: \bibitem{smirnov}
1252:     O.~L.~Peres and A.~Y.~Smirnov,
1253:     %``(3+1) spectrum of neutrino masses: A chance for LSND?,''
1254:     Nucl.\ Phys.\ B {\bf 599}, 3 (2001) [arXiv:hep-ph/0011054].
1255:     %%CITATION = HEP-PH 0011054;%%
1256: 
1257: \bibitem{LosAl}
1258:     C.~Athanassopoulos {\it et al.}  [LSND Collaboration],
1259:     %``Evidence for nu/mu $\to$ nu/e neutrino oscillations from LSND,''
1260:     Phys.\ Rev.\ Lett.\  {\bf 81} (1998) 1774
1261:     [arXiv:nucl-ex/9709006]; \\
1262:     %%CITATION = NUCL-EX 9709006;%%
1263:     G.~B.~Mills  [LSND Collaboration],
1264:     %``Neutrino Oscillation Results From Lsnd,''
1265:     Nucl.\ Phys.\ Proc.\ Suppl.\  {\bf 91} (2001) 198.
1266:     %%CITATION = NUPHZ,91,198;%%
1267: 
1268: \bibitem{3p1schemes}
1269:     M.~Maltoni, T.~Schwetz and J.~W.~Valle,
1270:     %``Cornering (3+1) sterile neutrino schemes,''
1271:     Phys.\ Lett.\ B {\bf 518} (2001) 252
1272:     [arXiv:hep-ph/0107150].
1273:     %%CITATION = HEP-PH 0107150;%%
1274: 
1275: \bibitem{stfluxes}
1276:     M.~C.~Gonzalez-Garcia, M.~Maltoni and C.~Pena-Garay,
1277:     %``Solar and atmospheric four-neutrino oscillations,''
1278:     Phys.\ Rev.\ D {\bf 64} (2001) 093001
1279:     [arXiv:hep-ph/0105269];
1280:     %%CITATION = HEP-PH 0105269;%%
1281:     M.~C.~Gonzalez-Garcia, M.~Maltoni and C.~Pena-Garay,
1282:     %``Update on solar and atmospheric four-neutrino oscillations,''
1283:     arXiv:hep-ph/0108073.
1284:     %%CITATION = HEP-PH 0108073;%%
1285: 
1286: \bibitem{snova}
1287:     D.~O.~Caldwell, G.~M.~Fuller and Y.~Z.~Qian,
1288:     %``Sterile neutrinos and supernova nucleosynthesis,''
1289:     Phys.\ Rev.\ D {\bf 61} (2000) 123005
1290:     [arXiv:astro-ph/9910175].
1291:     %%CITATION = ASTRO-PH 9910175;%%
1292: 
1293: \bibitem{dol}
1294:     R.~Barbieri and A.~Dolgov,
1295:     %``Bounds On Sterile-Neutrinos From Nucleosynthesis,''
1296:     Phys.\ Lett.\ B {\bf 237} (1990) 440;\\
1297:     %%CITATION = PHLTA,B237,440;%%
1298:     R.~Barbieri and A.~Dolgov,
1299:     %``Neutrino Oscillations In The Early Universe,''
1300:     Nucl.\ Phys.\ B {\bf 349} (1991) 743.
1301:     %%CITATION = NUPHA,B349,743;%%
1302: 
1303: \bibitem{ektdo}
1304:     K.~Kainulainen,
1305:     %``Light Singlet Neutrinos And The Primordial Nucleosynthesis,''
1306:     Phys.\ Lett.\ B {\bf 244} (1990) 191;
1307:     %%CITATION = PHLTA,B244,191;%%
1308:     K.\ Enqvist, K.\ Kainulainen and J.\ Maalampi,
1309:     %``Resonant Neutrino Transitions And Nucleosynthesis,''
1310:     Phys.\ Lett.\ B {\bf 249}, 531 (1990);
1311:     %%CITATION = PHLTA,B249,531;%%
1312: 
1313: \bibitem{ekm}
1314:     K.~Enqvist, K.~Kainulainen and J.~Maalampi,
1315:     %``Refraction And Oscillations Of Neutrinos In The Early Universe,''
1316:     Nucl.\ Phys.\ B {\bf 349} (1991) 754.
1317:     %%CITATION = NUPHA,B349,754;%%
1318: 
1319: \bibitem{ektBig}
1320:     K.~Enqvist, K.~Kainulainen and M.~J.~Thomson,
1321:     %``Stringent cosmological bounds on inert neutrino mixing,''
1322:     Nucl.\ Phys.\ B {\bf 373} (1992) 498.
1323:     %%CITATION = NUPHA,B373,498;%%
1324: 
1325: \bibitem{ektat}
1326:     K.~Enqvist, K.~Kainulainen and M.~J.~Thomson,
1327:     %``Atmospheric Neutrino Fluxes And Sterile-Neutrinos,''
1328:     Phys.\ Lett.\ B {\bf 288} (1992) 145.
1329:     %%CITATION = PHLTA,B288,145;%%
1330: 
1331: \bibitem{old}
1332:     J.~M.~Cline,
1333:     %``Constraints on almost Dirac neutrinos from neutrino - anti-neutrino oscillations,''
1334:     Phys.\ Rev.\ Lett.\  {\bf 68}, 3137 (1992).
1335:     %%CITATION = PRLTA,68,3137;%%
1336:     X.~Shi, D.~N.~Schramm and B.~D.~Fields,
1337:     %``Constraints on neutrino oscillations from big bang nucleosynthesis,''
1338:     Phys.\ Rev.\ D {\bf 48}, 2563 (1993) [arXiv:astro-ph/9307027].
1339:     %%CITATION = ASTRO-PH 9307027;%%
1340: 
1341: \bibitem{ekmL}
1342:     K.~Enqvist, K.~Kainulainen and J.~Maalampi,
1343:     %``Neutrino Asymmetry And Oscillations In The Early Universe,''
1344:     Phys.\ Lett.\ B {\bf 244} (1990) 186.
1345:     %%CITATION = PHLTA,B244,186;%%
1346: 
1347: \bibitem{ftv}
1348:     R.~Foot, M.~J.~Thomson and R.~R.~Volkas,
1349:     %``Large neutrino asymmetries from neutrino oscillations,''
1350:     Phys.\ Rev.\ D {\bf 53} (1996) 5349
1351:     [arXiv:hep-ph/9509327].
1352:     %%CITATION = HEP-PH 9509327;%%
1353: 
1354: 
1355: \bibitem{shi}
1356:     X.~d.~Shi,
1357:     %``Chaotic Amplification of Neutrino Chemical Potentials by Neutrino Oscillations in Big Bang Nucleosynthesis,''
1358:     Phys.\ Rev.\ D {\bf 54} (1996) 2753
1359:     [arXiv:astro-ph/9602135].
1360:     %%CITATION = ASTRO-PH 9602135;%%
1361: 
1362: \bibitem{eks1}
1363:     K.~Enqvist, K.~Kainulainen and A.~Sorri,
1364:     %``On chaoticity of the amplification of the neutrino asymmetry in the  early universe,''
1365:     Phys.\ Lett.\ B {\bf 464} (1999) 199 [arXiv:hep-ph/9906452].
1366:     %%CITATION = HEP-PH 9906452;%%
1367: 
1368: \bibitem{fSc}
1369:     R.~Foot,
1370:     %``Implications of the nu/mu $\to$ nu/s solution to the atmospheric neutrino  anomaly for early universe cosmology,''
1371:     Astropart.\ Phys.\  {\bf 10} (1999) 253
1372:     [arXiv:hep-ph/9809315].
1373:     %%CITATION = HEP-PH 9809315;%%
1374: 
1375: \bibitem{Sorri}
1376:     A.~Sorri,
1377:     %``Physical origin of 'chaoticity' of neutrino asymmetry,''
1378:     Phys.\ Lett.\ B {\bf 477} (2000) 201
1379:     [arXiv:hep-ph/9911366].
1380:     %%CITATION = HEP-PH 9911366;%%
1381: 
1382: \bibitem{shifuNc}
1383:     X.~d.~Shi, G.~M.~Fuller and K.~Abazajian,
1384:     %``Neutrino-mixing-generated lepton asymmetry and the primordial He-4  abundance,''
1385:     Phys.\ Rev.\ D {\bf 60} (1999) 063002
1386:     [arXiv:astro-ph/9905259]; \\
1387:     %%CITATION = ASTRO-PH 9905259;%%
1388:     K.~Abazajian, G.~M.~Fuller and X.~Shi,
1389:     %``The increase to the primordial He-4 yield in the two-doublet  four-neutrino mixing scheme,''
1390:     Phys.\ Rev.\ D {\bf 62} (2000) 093003
1391:     [arXiv:astro-ph/9908081].
1392:     %%CITATION = ASTRO-PH 9908081;%%
1393: 
1394: \bibitem{recent}
1395:     S.~Burles, K.~M.~Nollett and M.~S.~Turner,
1396:     %``Big-Bang Nucleosynthesis Predictions for Precision Cosmology,''
1397:     Astrophys.\ J.\  {\bf 552} (2001) L1
1398:     [arXiv:astro-ph/0010171];
1399:     %%CITATION = ASTRO-PH 0010171;%%
1400:     R.~H.~Cyburt, B.~D.~Fields and K.~A.~Olive,
1401:     %``Primordial Nucleosynthesis with CMB Inputs: Probing the Early Universe and Light Element Astrophysics,''
1402:     arXiv:astro-ph/0105397.
1403:     %%CITATION = ASTRO-PH 0105397;
1404: 
1405: \bibitem{dibSc}
1406:     P.~Di Bari,
1407:     %``Update on neutrino mixing in the early universe,''
1408:     arXiv:hep-ph/0108182.
1409:     %%CITATION = HEP-PH 0108182;%%
1410: 
1411: \bibitem{fvL1}
1412:     R.~Foot and R.~R.~Volkas,
1413:     %``Reconciling sterile neutrinos with big bang nucleosynthesis,''
1414:     Phys.\ Rev.\ Lett.\  {\bf 75} (1995) 4350
1415:     [arXiv:hep-ph/9508275].
1416:     %%CITATION = HEP-PH 9508275;%%
1417: 
1418: \bibitem{mainz}
1419:     J.~Bonn {\it et al.},
1420:     %``Newest Results From The Mainz Neutrino Mass Experiment,''
1421:     Phys.\ Atom.\ Nucl.\  {\bf 63} (2000) 969 [Yad.\ Fiz.\  {\bf 63} (2000) 1044].
1422:     %%CITATION = PANUE,63,969;%%
1423: 
1424: \bibitem{dolHan}
1425:     A.~D.~Dolgov, S.~H.~Hansen, S.~Pastor and D.~V.~Semikoz,
1426:     %``Neutrino oscillations in the early universe: How large lepton asymmetry  can be generated?,''
1427:     Astropart.\ Phys.\  {\bf 14} (2000) 79
1428:     [arXiv:hep-ph/9910444].
1429:     %%CITATION = HEP-PH 9910444;%%
1430: 
1431: \bibitem{DiBetal}
1432:     P.~Di Bari, R.~Foot, R.~R.~Volkas and Y.~Y.~Wong,
1433:     %``Comment on 'Neutrino oscillations in the early universe: How can large  lepton asymmetry be generated?',''
1434:     Astropart.\ Phys.\  {\bf 15} (2001) 391
1435:     [arXiv:hep-ph/0008245].
1436:     %%CITATION = HEP-PH 0008245;%%
1437: 
1438: \bibitem{fdb}
1439:     P.~Di Bari and R.~Foot,
1440:     %``On the sign of the neutrino asymmetry induced by active-sterile  neutrino oscillations in the early universe,''
1441:     Phys.\ Rev.\ D {\bf 61} (2000) 105012
1442:     [arXiv:hep-ph/9912215].
1443:     %%CITATION = HEP-PH 9912215;%%
1444: 
1445: \bibitem{stod}
1446:     L.~Stodolsky,
1447:     %``On The Treatment Of Neutrino Oscillations In A Thermal Environment,''
1448:     Phys.\ Rev.\ D {\bf 36} (1987) 2273.
1449:     %%CITATION = PHRVA,D36,2273;%%
1450: 
1451: \bibitem{sigl}
1452:     G.~Sigl and G.~Raffelt,
1453:     %``General kinetic description of relativistic mixed neutrinos,''
1454:     Nucl.\ Phys.\ B {\bf 406} (1993) 423.
1455:     %%CITATION = NUPHA,B406,423;%%
1456: 
1457: \bibitem{TMcK}
1458:     B.~H.~McKellar and M.~J.~Thomson,
1459:     %``Oscillating doublet neutrinos in the early universe,''
1460:     Phys.\ Rev.\ D {\bf 49} (1994) 2710.
1461:     %%CITATION = PHRVA,D49,2710;%%
1462: 
1463: \bibitem{aussiet}
1464:     N.~F.~Bell, R.~R.~Volkas and Y.~Y.~Wong,
1465:     %``Relic neutrino asymmetry evolution from first principles,''
1466:     Phys.\ Rev.\ D {\bf 59} (1999) 113001
1467:     [arXiv:hep-ph/9809363].
1468:     %%CITATION = HEP-PH 9809363;%%
1469: 
1470: \bibitem{fvSc}
1471:     R.~Foot and R.~R.~Volkas,
1472:     %``Studies of neutrino asymmetries generated by ordinary sterile neutrino  oscillations in the early universe and implications for big bang nucleosynthesis bounds,''
1473:     Phys.\ Rev.\ D {\bf 55} (1997) 5147
1474:     [arXiv:hep-ph/9610229].
1475:     %%CITATION = HEP-PH 9610229;%%
1476: 
1477: \bibitem{Steen} P.~E.~Braad and S.~Hannestad,
1478:     %``On the chaoticity of active-sterile neutrino oscillations in the early  universe,''
1479:     arXiv:hep-ph/0012194.
1480:     %%CITATION = HEP-PH 0012194;%%
1481: 
1482: \bibitem{nonThDM} X.~d.~Shi and G.~M.~Fuller,
1483:     %``A new dark matter candidate: Non-thermal sterile neutrinos,''
1484:     Phys.\ Rev.\ Lett.\  {\bf 82} (1999) 2832
1485:     [arXiv:astro-ph/9810076]; \\
1486:     %%CITATION = ASTRO-PH 9810076;%%
1487:     K.~Abazajian, G.~M.~Fuller and M.~Patel,
1488:     %``Sterile neutrino hot, warm, and cold dark matter,''
1489:     Phys.\ Rev.\ D {\bf 64} (2001) 023501
1490:     [arXiv:astro-ph/0101524];\\
1491:     %%CITATION = ASTRO-PH 0101524;%%
1492:     K.~Abazajian, X.~Shi and G.~M.~Fuller,
1493:     %``Transformation induced nonthermal neutrino spectra and primordial  nucleosynthesis,''
1494:     arXiv:astro-ph/9909320.
1495:     %%CITATION = ASTRO-PH 9909320;%%
1496: 
1497: \bibitem{eks2}
1498:     K.~Enqvist, K.~Kainulainen and A.~Sorri,
1499:     %``Creation of large spatial fluctuations in neutrino asymmetry by  neutrino oscillations,''
1500:     JHEP {\bf 0104} (2001) 012
1501:     [arXiv:hep-ph/0012291].
1502:     %%CITATION = HEP-PH 0012291;%%
1503: 
1504: \bibitem{shiFuDo}
1505:     X.~d.~Shi and G.~M.~Fuller,
1506:     %``Leptonic domains in the early universe and their implications,''
1507:     Phys.\ Rev.\ Lett.\  {\bf 83} (1999) 3120 [arXiv:astro-ph/9904041].
1508:     %%CITATION = ASTRO-PH 9904041;%%
1509: 
1510: \bibitem{Kinney:1999pd}
1511:     W.~H.~Kinney and A.~Riotto,
1512:     %``Measuring the cosmological lepton asymmetry through the CMB anisotropy,''
1513:     Phys.\ Rev.\ Lett.\ {\bf 83} (1999) 3366 [arXiv:hep-ph/9903459].
1514:     %%CITATION = HEP-PH 9903459;%%
1515: 
1516: \bibitem{CMBconstraints}
1517:     S.~Hannestad,
1518:     %``New constraints on neutrino physics from Boomerang data,''
1519:     Phys.\ Rev.\ Lett.\  {\bf 85} (2000) 4203 [arXiv:astro-ph/0005018].
1520:     %%CITATION = ASTRO-PH 0005018;%%
1521: 
1522: 
1523: 
1524: \end{thebibliography}
1525: 
1526: \end{document}
1527: