1: \documentclass[preprint,aps]{revtex4}
2: \usepackage{graphicx}% Include figure files
3: \begin{document}
4: %\draft
5: \title{Active-Sterile Neutrino Conversion: Consequences for the
6: $r$-Process and Supernova Neutrino Detection}
7: \author{J. Fetter,$^{1}$ G. C. McLaughlin$^{2}$,
8: A. B. Balantekin,$^{1}$ and G. M. Fuller,$^{3}$}
9: \affiliation{
10: $^{1}$Department of Physics, University of Wisconsin, Madison,
11: Wisconsin 53706\\
12: $^{2}$Department of Physics, North Carolina State University, Raleigh,
13: NC 27695-8202\\
14: $^{3}$Department of Physics, University of California, San Diego,
15: La Jolla, CA, 92093-0319\\
16: }
17: \date{\today}
18:
19: \begin{abstract}
20: We examine active-sterile neutrino conversion in the late time
21: post-core-bounce supernova environment. By including the effect of
22: feedback on the Mikheyev-Smirnov-Wolfenstein (MSW) conversion
23: potential, we obtain a large range of neutrino mixing parameters which
24: produce a favorable environment for the r-process. We look at the
25: signature of this effect in the current generation of neutrino
26: detectors now coming on line. We also investigate the impact of the
27: neutrino-neutrino forward scattering-induced potential on the MSW
28: conversion.
29:
30: \end{abstract}
31: \maketitle
32:
33: \newpage
34: \section{Introduction}
35: \label{sec:intro}
36:
37: One form of element synthesis which may take place in core collapse
38: supernovae ({\it e.g.,} Type II, Ib, Ic) is the $r$-process, or rapid
39: neutron capture process. The
40: $r$-process of nucleosynthesis has long been thought to be the
41: mechanism for producing many of the heavy nuclei (mass number, A $>$
42: 100) \cite{bbfh}. Successful production of the
43: r-process elements with anything approaching a solar system-like
44: abundance distribution requires a
45: neutron-rich environment, at least in the conditions
46: suggested by current supernova models and simple neutrino-driven wind
47: models without extremely rapid outflow and/or extremely relativistic
48: neutron star configurations.
49:
50: In this paper, we study a primary $r$-process which occurs in stages.
51: The material in a fluid element moving away from the neutron star
52: remnant left by a core collapse supernova explosion
53: will experience three stages of nuclear evolution.
54: First, only free nucleons are present, then alpha particles coexist with
55: free neutrons. At still lower temperatures the matter is
56: composed of alpha particles,
57: ``seed'' nuclei with $50 \lesssim A \lesssim 100$, and
58: neutrons. In the last stage the neutrons capture on the seed nuclei.
59:
60: The neutrino-driven-wind environment, which is thought to occur at late
61: time (time post bounce $t_{pb} \sim 10 {\rm s}$) in the supernova is a
62: promising candidate site for the production of the r-process elements
63: \cite{r1}. The wind is thought to arise well after the initial
64: collapse/explosion event; its successful generation of course
65: presupposes a successful explosion. In the classic neutrino driven wind
66: models, neutrinos transfer energy to material near the surface of the
67: protoneutron star. As this material flows outward, it expands and
68: cools. As it cools, weak and electromagnetic/strong nuclear reaction
69: rates drop out of equilibrium sequentially. If enough neutrons are
70: present after/during this freeze-out process, the $r$-process may take
71: place. The number of neutrons is determined by three factors, the
72: initial neutron-to-proton ratio, the outflow timescale and the entropy of
73: the material. Semianalytic models of the neutrino driven wind have shown
74: that it is difficult, though perhaps not impossible, to generate the
75: required numbers of neutrons per seed nucleus \cite{hwq}.
76:
77: Self-consistent inclusion of neutrino capture reactions into a reaction
78: network code has shown that the problem of obtaining the required number of
79: neutrons is exacerbated by the interplay between neutrino and nuclear
80: reactions, especially the formation of alpha particles as described below
81: \cite{mmf}. This is the so-called \lq\lq alpha
82: effect.\rq\rq\ The relative number of neutrons and protons is usually cast
83: in terms of the electron fraction,
84: $Y_e = 1/(1 + n/p)$, the net number of electrons (electrons minus
85: positrons) per baryon.
86: The electron fraction
87: is determined mostly by the rates of neutrino and antineutrino
88: capture on free neutrons and protons \cite{Qetal}. The most detrimental
89: effect, the alpha effect, takes place when material which would
90: eventually undergo heavy element nucleosynthesis passes through the
91: intermediate step of forming alpha particles. At this stage, all the
92: protons are locked into alphas, but any excess neutrons remain free.
93: Neutrino captures on the remaining neutrons decrease the ratio of
94: neutrons to protons (or equivalently raise the electron fraction).
95: Supernova neutrino energies are too low to permit compensating charged
96: current captures on alpha particles \cite{mmf}.
97: Without sufficient numbers of neutrons, there can be no rapid neutron
98: capture process in these sorts of wind scenarios.
99:
100: There are three potential solutions to this problem. One solution involves
101: hydrodynamics. For example, a very fast outflow may in principle cure the
102: problems associated with this environment \cite{hwq} and also the alpha
103: effect \cite{cf}.
104: Alternately, other scenarios which have rapid expansion (perhaps
105: followed by a slow down) accompanied by high entropy and temperature
106: may also cure the problem \cite{r1}. However, it is not clear that
107: such a fast outflow
108: or high entropy is present in this environment, or that either can be
109: achieved with simple neutrino heating.
110:
111: The second solution may be to look for another site, such as neutron star
112: mergers. Some neutron rich material may be ejected in the merger and
113: this has been shown to be capable of
114: producing r-process elements \cite{bsm89}. These
115: mergers are likely too rare to reproduce the total observed abundance
116: of r-process elements, unless a significant fraction of the neutron
117: star material is ejected \cite{thielemann}. There is significant
118: energy released in the form of neutrinos in these models.
119: Therefore, the alpha effect may
120: compromise the r-process in these environments as well, although the
121: neutrino average energies in neutron star -neutron star mergers could
122: be somewhat lower than in core collapse supernovae and neutron star -
123: black hole mergers, as discussed in e. g. Ref. \cite{janka99}.
124:
125: A third solution is the one that is investigated in this paper:
126: active-sterile ($\nu_e \leftrightarrow \nu_s$, $\bar{\nu_e}
127: \leftrightarrow \bar{\nu_s}$) neutrino transformation. The $\nu_s$ in
128: our study is defined as a particle which mixes with the $\nu_e$ (and
129: possibly also with $\nu_\mu$, and/or $\nu_\tau$) but does not contribute to
130: the width of the Z boson. One candidate for this particle, although
131: not the only one, is a right handed Dirac neutrino. A large number of
132: models exist for such light ``sterile'' neutrinos, for example, any
133: SU(2) Standard Model singlet.
134:
135: In fact, the existence of sterile neutrinos with large masses ({\it
136: e.g.,} of order the unification scale) would not be unexpected, as they
137: are suggested by many neutrino mass/mixing models. However, light
138: sterile neutrinos do not often occur naturally in these models.
139:
140: The Sudbury Neutrino Observatory
141: (SNO) and the SuperKamiokande (SuperK) experiments together observe high
142: energy solar neutrinos and SuperK can observe and characterize the
143: atmospheric neutrino flux. What is emerging is a picture in which
144: there is near
145: maximal mixing between the mu and tau flavor neutrinos at the atmospheric
146: neutrino mass-squared difference scale $\delta m^2 \approx
147: 3\times{10}^{-3}\,{\rm eV}^2$, and a convincing result that two-thirds of
148: the neutrinos coming from the sun are mu/tau neutrinos, narrowly favoring
149: again near maximal mixing between the electron flavor neutrinos produced
150: by nuclear reactions in the sun and mu/tau flavor neutrinos at the solar
151: neutrino mass-squared difference $\delta m^2 <{10}^{-4}\,{\rm eV}^2$
152: \cite{superk,sno,bah}. Is there any indication of light sterile degrees of
153: freedom?
154:
155: The Los Alamos Liquid Scintillator Neutrino Detector (LSND) experiment
156: remains the only possible indication of neutrino flavor oscillations (in
157: the $\nu_{\mu} \rightleftharpoons\nu_e$ channel at a neutrino
158: mass-squared scale $\delta m^2 \sim 0.5\,{\rm eV}^2$)
159: beyond those derived from solar and atmospheric
160: neutrino considerations \cite{lsnd}.
161: The mini-BooNE experiment at Fermilab will soon
162: check this result. If LSND holds up then it is tantamount to direct
163: evidence for the existence of a light sterile neutrino. Four neutrino
164: schemes designed to fit all of the existing neutrino oscillation data
165: \cite{gg} are, however, barely viable or are outright challenged by the data.
166: By contrast, we emphasize here that our considerations of the effect of a
167: sterile neutrino on supernova physics are completely independent of the
168: LSND result. For example, the sterile neutrino mass range which we find to
169: be efficacious in supernova r-process nucleosynthesis spans the entire
170: LSND-inspired range but also extends to considerably higher values of
171: $\delta m^2$ and to smaller vacuum mixing angles.
172:
173: The effect of active-sterile transformation differs from that of
174: active-active ({\it e.g.}, $\nu_e \leftrightarrow \nu_{\mu\tau}$)
175: transformation, in that active-active transformation usually tends to
176: decrease the number of neutrons in the supernova, while active-sterile
177: transformation can increase it. (Active-active transformation can
178: slightly decrease the electron fraction for certain narrow ranges of
179: neutrino mass-squared difference.) It has been shown that matter-enhanced
180: $\nu_e
181: \leftrightarrow \nu_{\mu,
182: \tau}$ could under some circumstances rather drastically affect the
183: neutron to proton ratio in the neutrino-driven wind and, thereby, affect
184: the prospects for the r-process there
185: \cite{Qetal}.
186:
187: Active-sterile transformation in the supernova environment must be
188: treated carefully for several reasons. The first is that, due to the
189: structure of the Mikheyev-Smirnov-Wolfenstein (MSW) conversion
190: potential, both neutrinos and antineutrinos may undergo transformation
191: in the same scenario. In fact, in the solutions discussed here both
192: types of transformations take place. The second is that the neutrinos
193: themselves in part determine the MSW potential. Any calculation of
194: the effect of oscillations must therefore include a feedback loop. The
195: neutrinos determine the electron fraction; the electron fraction
196: determines the potential; and the potential determines the spectrum of
197: neutrinos. The spectrum of neutrinos then sets the electron fraction.
198: Additionally, the $\nu \nu$-forward scattering contribution to the
199: MSW potential depends on the flavor states of the neutrinos.
200:
201: In section \ref{sec:mech} we review the calculations which illustrate the
202: feedback mechanism \cite{us}. Our study differs from a previous
203: study \cite{nunokawa} in that we track the thermodynamic and nuclear
204: statistical equilibrium evolution (NSE) of a mass element and update the
205: numbers of neutrons and protons at each time step directly from the weak
206: capture rates. This is essential to
207: accurately determine the number of neutrons available for the r-process.
208: In addition to producing a favorable $n/p$ ratio, this type of
209: transformation can suppress the population of $\nu_e$ and thereby defeat
210: the alpha effect. In section
211: \ref{sec:detect} we discuss the observational consequences of such a
212: neutrino flavor transformation scenario in current neutrino detectors.
213: In section
214: \ref{sec:back} we explore the importance of neutrino background terms, or
215: neutrino-neutrino scattering in the MSW potential. In section
216: \ref{sec:concl} we give conclusions.
217:
218: \section{The Mechanism}
219: \label{sec:mech}
220:
221: In this section, we briefly describe our calculations and then we
222: summarize the impact of active-sterile neutrino transformation on the
223: electron fraction in the neutrino driven wind. Our study in this
224: paper differs from that in our earlier work in Ref.\ \cite{us} in that
225: here we discuss the expected effects of active-sterile neutrino flavor
226: transformation on the expected supernova neutrino signal in several
227: different detectors and we also consider effects of the neutrino-neutrino
228: forward scattering-induced potential (the \lq\lq neutrino
229: background\rq\rq) on the neutrino flavor transformation problem.
230:
231: If we neglect $\nu$-$\nu$ forward scattering contributions to the
232: weak potentials, then the equation which governs the evolution of the
233: neutrinos as they pass though the material in the wind can be written as:
234:
235: \begin{equation}
236: i\hbar \frac{\partial}{\partial r} \left[\begin{array}{cc} \Psi_e(r)
237: \\ \\ \Psi_s(r) \end{array}\right] = \left[\begin{array}{cc}
238: \varphi_e(r) & \sqrt{\Lambda} \\ \\ \sqrt{\Lambda} & -\varphi_e(r)
239: \end{array}\right]
240: \left[\begin{array}{cc} \Psi_e(r) \\ \\ \Psi_s(r)
241: \end{array}\right]\,,
242: \label{eq:msw}
243: \end{equation}
244: where
245: \begin{equation}
246: \label{2} \varphi_e(r) = \frac{1}{4 E} \left( \pm
247: 2 \sqrt{2}\ G_F \left[
248: N_e^-(r) - N_e^+(r) - \frac{N_n(r)}{2} \right] E - \delta m^2
249: \cos{2\theta_v} \right)
250: \label{eq:potne}
251: \end{equation}
252: The upper sign is relevant for neutrino transformations; the lower one is
253: for antineutrinos. In these equations
254: \begin{equation}
255: \label{eq:offdia}
256: \sqrt{\Lambda} = \frac{\delta m^2}{4 E}\sin{2\theta_v},
257: \end{equation}
258: $\delta m^2 \equiv m_2^2 - m_1^2$ is the vacuum mass-squared
259: splitting, $\theta_v$ is the vacuum mixing angle, $G_F$ is the Fermi
260: constant, and $N_e^-(r)$, $N_e^+(r)$, and $N_n(r)$ are the
261: total proper number densities of electrons, positrons,
262: and neutrons respectively in the
263: medium. Again, this form of the potential, $\varphi_e(r)$,
264: is valid only in the absence of
265: neutrino background effects. The background effects will be explored in
266: detail in Section \ref{sec:back}, but we note that Eq. (\ref{eq:msw}) is
267: adequate for following $\nu$ flavor evolution well above the neutron star.
268:
269: Eq. (\ref{eq:potne}) can be rewritten in
270: terms of a potential,
271: \begin{equation}
272: V(r) = 2 \sqrt{2} G_F \left[ N_e^-(r) - N_e^+(r) - \frac{N_n(r)}{2} \right].
273: \label{v-def}
274: \end{equation}
275: Then the on-diagonal term in the Hamiltonian becomes
276: \begin{equation}
277: \varphi_e(r) = \frac{1}{4E}
278: \left( \pm V(r) E - \delta m^2 \cos 2\theta_v \right).
279: \label{eq:subst-v}
280: \end{equation}
281: Neutrinos undergo a resonance and transform primarily when the on-diagonal
282: term is zero, that is, at the resonant energy
283: \begin{equation}
284: E_{res}(r) = \pm\frac{\delta m^2 \cos 2\theta_v}{V(r)}.
285: \label{eq:eres}
286: \end{equation}
287: We see that neutrinos undergo resonance when the potential is positive, and
288: antineutrinos undergo resonance when it is negative. Finally, we rewrite
289: the potential in terms of the electron fraction,
290: \begin{equation}
291: \label{eq:potye}
292: V(r) = \frac{3 G_F \rho (r)}{2 \sqrt{2}m_N}
293: \left( Y_e - \frac{1}{3} \right),
294: \end{equation}
295: where
296: \begin{equation}
297: \label{eq:ye}
298: Y_e (r)= \frac{N_e^-(r)-N_e^+(r) }{N_p(r)+N_n(r)}.
299: \end{equation}
300: Here $\rho(r)$ is the density of the matter at distance $r$ from the
301: protoneutron star, and $m_N$ is the mass of a nucleon
302: and $N_p(r)$ is the total proper proton number density. We assume that
303: $N_p = N_e^- - N_e^+$ because of local electromagnetic charge neutrality.
304: Since the electron fraction can take on values between
305: zero and one, the potential can be either positive or negative.
306: Therefore, depending on the value of the electron fraction, either
307: neutrino or antineutrino transformation may be matter-enhanced.
308:
309: We solve Eq. (\ref{eq:msw}) numerically for survival probabilities
310: of neutrinos as they pass through the matter above the protoneutron
311: star. No approximations are employed in computing the survival
312: probabilities, for the adopted potential.
313: In the absence of oscillations and matter-enhanced flavor transformation,
314: the spectrum of neutrinos and antineutrinos can be parameterized as
315: approximately Fermi-Dirac in character. We
316: assume that no transformation has taken place within the protoneutron
317: star, so we begin with a full complement of each species of
318: active neutrino.
319: However, since neutrinos of different energies will have
320: different survival probabilities at each distance above the surface,
321: the distribution quickly departs from the Fermi-Dirac shape. We use
322: representative initial neutrino distributions with temperatures of
323: ${\rm T}_{\nu_e} = 3.22 \,
324: {\rm MeV}$ and ${\rm T}_{\bar{\nu}_e} = 4.5 \, {\rm MeV}$,
325: luminosities of $L_{\nu_e} = 1.08 \times 10^{51} {\rm ergs} \, {\rm
326: s}^{-1}$ and $L_{\bar{\nu}_e} = 1.3 \times 10^{51} {\rm ergs} \, {\rm
327: s}^{-1}$ and effective chemical potentials of zero. Though different
328: numerical simulations of neutrino transport in the hot protoneutron star
329: core differ in their predictions of these spectral values, we note that
330: our adopted values serve to illustrate the general qualitative behavior
331: that would be expected if the simulations included this neutrino flavor
332: transformation physics.
333:
334: In the neutrino driven wind, the neutrinos and antineutrinos
335: are the most important agents in determining the electron
336: fraction. However, near the surface of the protoneutron star there is
337: also a contribution from electrons and positrons:
338: \begin{equation}
339: \label{eq:ccf}
340: \nu_e + {\rm n} \rightleftharpoons {\rm p}+ e^{-};
341: \end{equation}
342: \begin{equation}
343: \label{eq:ccbf}
344: \bar{\nu}_e + {\rm p} \rightleftharpoons {\rm n} + e^{+}.
345: \end{equation}
346: Close to the surface of the protoneutron star, these weak capture rates are
347: fast in comparison with the outflow timescale. As we move far from the
348: surface, they become negligible in comparison with outflow. Therefore the
349: weak capture rates are in the process of freezing out of steady state
350: (chemical) equilibrium. We calculate the electron fraction by numerically
351: following the evolution of the number densities of neutrons and protons.
352:
353: Our calculations are performed by tracking the evolution of mass elements
354: in the neutrino driven wind. We use distance and density profiles from
355: analytic descriptions of the wind; $r \propto \exp(-t/\tau)$ and $\rho
356: \propto r^{-3}$ where $\tau$ is the outflow timescale \cite{qw}. For
357: illustrative purposes, we use a timescale of $ \tau = 0.3 \, {\rm s}$ and
358: an entropy per baryon of $s = 100$ in units of Boltzmann's constant. Close
359: to the surface of the protoneutron star, before the wind begins to operate,
360: we use the density profile of Wilson and Mayle \cite{wilson}. Calculations
361: done with different timescales show the same qualitative features that we
362: present here \cite{us} and variations in the entropy are expected to have a
363: similar effect.
364:
365: At each time step the distance and density are incremented. All other
366: thermodynamic quantities, including the number densities of positrons and
367: electrons, are calculated from the entropy and density. The mass fractions
368: of the neutrons, protons, alpha particles and heavy nuclei (A $>$ 40) are
369: calculated in NSE (Nuclear Statistical Equilibrium). The weak rates are
370: then computed and the electron fraction is updated. This new electron
371: fraction is used in the MSW equations to calculate new survival
372: probabilities for each neutrino energy bin. We terminate calculations
373: for a particular mass element at the point when heavy nuclei begin to
374: form. Since the outflow times considered in our calculations are
375: relatively short, $t \sim \tau
376: \lesssim 0.5 \, {\rm s}$, we have assumed that each mass element
377: experiences the same evolution as the previous ones.
378:
379: The consequences of the feedback effect are illustrated in Figure
380: \ref{fig:efraction1}. In this figure, the electron fraction is plotted
381: against distance as measured from the center of the protoneutron star. The
382: upper curve shows the evolution of the electron fraction in the absence of
383: neutrino transformation, while the lower curve shows the evolution of the
384: electron fraction for mixing parameters of $\delta m^2 = 20 \, {\rm eV}^2$,
385: $\sin^2 2 \theta_{v} = 0.01$. The initial rise in the electron
386: fraction is due to Pauli unblocking of neutrino capture on neutrons as the
387: density rapidly decreases at the edge of the protoneutron star. At such
388: high density we do not include feedback effects; the validity of this
389: approximation is discussed below. We begin the feedback effects at about
390: the distance where the wind solution begins to dominate the hydrodynamic
391: flow. In the case of no transformation the electron fraction shows a
392: slight drop, due to the decreasing importance of electron and positron
393: capture, and then a slight rise due to the alpha effect.
394:
395: In the presence of mixing, however, the situation is quite different. Low
396: energy electron neutrinos begin to transform slightly above the surface of
397: the protoneutron star, where the potential is large (Eq.~\ref{eq:eres}).
398: This transformation decreases the rate of neutrino capture on neutrons,
399: which lowers the electron fraction, and therefore the potential. Thus
400: neutrinos of higher and higher energy begin to transform. Eventually the
401: electron fraction drops below 1/3, so the potential becomes negative, and
402: the highest energy electron antineutrinos begin to transform. This slows
403: down the drop in $Y_e$, but does not halt it entirely, since low energy
404: electron antineutrinos are still present. As the electron fraction
405: continues to fall, the magnitude of $V$ increases, so lower energy electron
406: antineutrinos transform. Meanwhile, the mass density is falling, causing
407: the potential to reach a minimum and, eventually, return toward zero.
408: The minimum means that only antineutrinos above a certain energy will
409: undergo resonance, and since the potential goes back to zero, all these
410: neutrinos (originally active antineutrinos) are re-converted from sterile
411: to active species. By the time alpha particles form, there are many
412: electron antineutrinos present but few electron neutrinos. Therefore,
413: there is no alpha effect. The resulting electron fraction is so low
414: $\sim 0.1$, that conditions become favorable for the r-process. There is
415: no alpha effect because the $\nu_e$'s that would have captured on
416: neutrons during the epoch of alpha particle formation are, for the right
417: neutrino mass/mixing parameters, converted to
418: sterile species in large measure at the alpha formation epoch. A study of
419: many different neutrino driven wind conditions, shows that for a
420: timescale of 0.3~s and an entropy per baryon of $\sim 100$, the electron
421: fraction must be below 0.18 in order to produce the requisite neutron to
422: seed ratio
423: \cite{meyer97}.
424:
425: An analysis in Ref.\ \cite{nunokawa} assumed that $Y_e(r)$ was equal to
426: its equilibrium value, as set from the weak capture rates
427: \cite[Eq.~3.14]{us}. Ref. \cite{nunokawa} neglected electron and
428: positron capture in finding the equilibrium value of of $Y_e(r)$, and
429: did not take into account alpha particle formation and did not take
430: account of the feedback of the outflow rate on neutrino capture
431: processes and neutrino flavor transformation. Similarly,
432: Ref. \cite{peres} did not include this physics.
433: Without these effects, the
434: system finds a fixed point at
435: $Y_e = 1/3$; neutrino conversions bring the electron fraction to this
436: value. Once there, the neutrino-matter forward scattering-induced
437: potentials vanish and there is no further neutrino flavor transformation
438: in the channels $\nu_e\rightleftharpoons\nu_s$ and
439: $\bar\nu_e\rightleftharpoons\bar\nu_s$.
440: In Fig. \ref{fig:equ}, we have duplicated this result by making the
441: assumptions outlined above. The figure also shows the evolution of $Y_e$
442: when we take account of alpha formation, but keep the other
443: simplifications as above. In this case, $Y_e$ goes to 1/3 as before, but
444: then the alpha effect drives it to 1/2 eventually. These results hold
445: for a broad range of neutrino mixing parameters.
446:
447: Except in Fig. \ref{fig:equ} we have carefully tracked the actual
448: value of $Y_e$ as distinct from its equilibrium value; we have
449: also taken account of electron and positron captures and alpha
450: formation. In this case, $Y_e$ lags behind its equilibrium value.
451: As a result, $Y_e$ remains greater than 1/3, even when the equilibrium
452: electron fraction ${(Y_e)}_{\rm eq}$ drops below that value.
453: Therefore, neutrinos continue to transform, and eventually drive the system
454: to low electron fraction values, $Y_e < 1/3$.
455:
456: We investigate a range of mixing parameters in Figure \ref{fig:efraction2}.
457: This contour plot shows the electron fraction, measured at the point where
458: heavy nuclei begin to form, for various mixing parameters. Here we employ
459: the same neutrino driven wind model used in Figure \ref{fig:efraction1}.
460: In the bottom left corner of the plot, the solution is approaching the
461: case without neutrino transformation. In the middle of the plot, the
462: transformations of neutrinos produce a very neutron rich environment.
463:
464: In the upper right part of the plot, the electron neutrinos are undergoing
465: an additional resonance, at smaller distance than those described above, and
466: in fact close to the neutron star surface where the density scale height is
467: small. This resonance occurs at the first place where the electron fraction
468: passes close to 1/3, at very high density. To correctly model this range of
469: mixing parameters it is necessary to include feedback in the very dense
470: region near the neutron star surface.
471:
472: In addition, because this resonance is
473: so close to the neutrinosphere, some neutrinos travel through it along
474: extremely nonradial paths. In general neutrinos which travel nonradially
475: transform differently from those which travel radially, but for most of our
476: parameter space, the effects are small. The results presented here include
477: only radial paths. We have checked them against another (very slow)
478: calculation, which includes nonradial effects. The difference is negligible
479: except when neutrinos or antineutrinos transform at this inmost resonance.
480:
481: We have drawn a shaded area on the contour
482: plot. In this shaded area the inclusion of nonradial neutrino paths
483: would alter the results by more than 10\%. We show this also in Fig.
484: \ref{fig:nonradial}. Here the difference in survival probabilities of
485: electron neutrinos at 11 km is plotted. The same wind parameters are
486: used here as in Fig. \ref{fig:efraction2}.
487:
488: This shaded region of Fig. \ref{fig:efraction2} (and also
489: the unshaded region in the upper right corner above it)
490: would also require us to take
491: account of feedback effects in the steep density gradient region near the
492: neutrino sphere. We have not included a full nonradial calculation of the
493: results there. These regions of the plot require
494: further investigation before
495: definitive conclusions about the evolution of the electron fraction can be
496: drawn. The work of Ref. \cite{mitesh} has discussed the difficulties
497: inherent in treating the steep region.
498: We emphasize, however, that the rest of our parameter space presents a
499: favorable environment for the
500: $r$-process.
501:
502: \section{Detection}
503: \label{sec:detect}
504:
505: In this section we study the supernova neutrino signal that would be
506: seen at SNO or SuperKamiokande after active-sterile transformation.
507: We consider the $\nu_e$,
508: and $\bar{\nu}_e$ parameters used in section II and
509: additional individual neutrino parameters of
510: $T_{\nu_\mu,\bar{\nu}_\mu,\nu_\tau,\bar{\nu}_\tau}= 6 \, {\rm MeV}$ and
511: $L_{\nu_\mu,\bar{\nu}_\mu,\nu_\tau,\bar{\nu}_\tau} = 1 \times 10^{51}
512: {\rm ergs} \, {\rm s}^{-1}$ .
513:
514: A galactic supernova is expected to produce a large electron
515: antineutrino signal in SuperK. In fact, a supernova at 10 kiloparsecs is
516: estimated to see about 8000 events from
517: $\nu_e + n \rightarrow e^- + p $ \cite{superK}.
518: In principle, there will be an additional 300 events
519: from neutrino scattering on electrons. Since both charged and neutral
520: current processes contribute to this type of scattering, all types of
521: neutrinos, except sterile will contribute.
522:
523: SNO will see an estimated 500 events from neutral current
524: break-up of the deuteron, $ \nu_x + d \rightarrow n + p$ and an
525: additional 150 for each of the charged
526: current processes, $\nu_e + d \rightarrow p + p + e^{-}$ and
527: $\bar{\nu}_e + d \rightarrow n + n + e^{-}$ \cite{SNO}.
528: These break-up reactions may be
529: distinguished and tagged by the presence or absence of the emitted neutron(s)
530: and the Cerenkov light from the electron or positron.
531:
532: \subsection{Active-sterile transformation alone}
533:
534: Figure \ref{fig:sno1} shows the ratio of the expected charged current electron
535: neutrino-induced event rates in SNO for the case with active-sterile
536: transformation to the case without such transformation. In the region which
537: produced the lowest electron fraction in Figure \ref{fig:efraction2}, we find
538: that the charged current event rate is suppressed by 90\%. In the lower left
539: hand corner of the plot, the ratio approaches one as the solution
540: asymptotes to the case of no transformation. In Figs. \ref{fig:sno1} -
541: \ref{fig:elscat2}, the shaded region shows where the neutron to seed
542: ratio should be greater than 100 (i.e. favorable for the r-process)
543: for the given wind parameters.
544:
545: Figure \ref{fig:elscat1} shows the ratio of the expected electron
546: neutrino-induced scattering events from all processes at Superkamiokande
547: for the case with active-sterile
548: transformation to the case without such transformation. Because the
549: charged current electron neutrino scattering rate is so much larger than the
550: neutral current one, the absence of electron neutrinos has a significant
551: $\sim 40$\% effect. This may be surprising since the
552: $\nu_\mu$,$\bar{\nu}_\mu$,$\nu_\tau$, and $\bar{\nu}_\tau$ have two and a
553: half times the energy of the
554: $\nu_e$ on average and the cross section is approximately linear in
555: neutrino energy. However, the luminosities of the species are roughly
556: the same, so the number flux of the high energy neutrinos is two and
557: half times less than the number flux of the lower energy neutrinos.
558: Therefore most of the small energy dependence in the rate stems from
559: the 5 MeV detection threshold in SuperK.
560:
561: These figures demonstrate that because of the deficit of electron neutrinos
562: produced by the active-sterile solution, the SNO charged current rate will
563: be greatly reduced from the expected event rate. However, the SuperK electron
564: (antineutrino) capture rate will be for the most part unchanged. Note that
565: our model is, therefore, unconstrained by data from supernova 1987A.
566:
567: The calculations of detector signals have assumed no
568: other type of neutrino mixing until this point.
569: However, both the atmospheric
570: neutrino mixing solution and the solar neutrino solution will cause
571: additional oscillations. The effect of $3 \, \times \, 3$
572: active only mixing on
573: a supernova neutrino signal has been discussed in the literature
574: (ref. \cite{fhm,dighe}). The vacuum mixing of $\nu_\mu
575: \leftrightarrow \nu_\tau$ which may solve the atmospheric neutrino
576: problem will cause $\nu_\mu \leftrightarrow \nu_\tau$ mixing
577: in supernova neutrinos. Since
578: the $\nu_\mu$ neutrinos and the $\nu_\tau$ neutrinos have the same
579: energies and luminosities, this mixing will not directly
580: cause changes in the detector signal. In the following discussion,
581: we assume $\sin^2 2 \theta_{atm} \approx \sin^2 2 \theta_{23} = 1$.
582:
583: \subsection{SMA and active-sterile transformation}
584:
585: We also consider the situation where $\nu_e
586: \leftrightarrow \nu_\mu$ MSW is the solution to the solar neutrino
587: problem. The small mixing angle (SMA) parameters,
588: $\delta m^2_{solar} = 5 \times 10^{-6} \, {\rm eV}^2$ and $\sin^ 2
589: 2 \theta_{solar} = \sin^2 2 \theta_{12} \approx 10^{-2}$
590: will cause partial conversion in the hydrogen envelope of
591: the supernova progenitor star.
592: The exact survival probabilities will depend on the
593: electron density scale height:
594: \begin{equation}
595: R_s = - { N_e \over dN_e/dr}.
596: \end{equation}
597: We use densities and distances from Ref. \cite{woosley} to estimate $R_s
598: \approx 10^{10} {\rm cm}$. We assume here that the density and
599: composition will be largely unaffected by the core collapse event at the
600: time that the neutrinos move through it.
601: For 30 MeV neutrinos, near-total
602: conversion (electron neutrino survival probability less than 1\%) requires
603: the density profile to be much shallower: $R_s > 8 \times10^{10}$ cm. For
604: no conversion (electron neutrino survival probability more than 99\%), the
605: density profile must be much steeper: $R_s < 1 \times 10^{8}$ cm.
606: Therefore it is very likely that partial conversion of the electron neutrinos
607: will take place, even given a more detailed model of the hydrogen envelope.
608:
609: Using a fit to the points given in ref. \cite{woosley}, we obtain ratios
610: for the SNO charged current electron neutrino scattering rate (Figure
611: \ref{fig:sno2}) and the SuperK electron neutrino rate (Figure
612: \ref{fig:elscat2}), with both the active-sterile mixing and the
613: additional envelope conversion. Again, we show results in terms of
614: ratios: the expected detector supernova neutrino-induced event rates with two
615: types of mixing to the expected event rates in the case with no neutrino
616: flavor conversion. The SuperK electron antineutrino capture rate is
617: the same as in Figure 4, since no electron antineutrino conversion occurs in
618: the envelope.
619:
620: Figure \ref{fig:sno2} shows that the partial conversion of muon neutrinos to
621: electron neutrinos, after the initial conversion of electron neutrinos
622: to steriles, produces a ratio of order one.
623: Figure \ref{fig:elscat2}
624: shows that electron scattering
625: ratio of rates at SuperK with conversion. The partial conversion of the muon
626: neutrinos more or less reproduces the case with no transformation whatsoever.
627:
628: \subsection{LMA and active sterile transformation}
629:
630: In table \ref{tab:osc} we show a few of the possible
631: outcomes for final electron neutrino spectrum, as a result of the
632: active-active mixings. The relevant parameters are $\theta_{12}$,
633: $\theta_{13}$ and the associated $\delta m^2$s. The case of maximal
634: mixing for the atmospheric neutrinos and the large angle solution to the
635: solar neutrino problem, with $\sin^2 2 \theta_{12} \approx
636: \sin^2 2\theta_{solar} \sim 0.8$ is shown.
637: Here we implicitly assume a normal (as opposed to an inverted) neutrino mass
638: hierarchy.
639:
640: As in the case of small angle mixing, the Large Mixing Angle (LMA)
641: solar parameters will cause neutrino flavor conversion in the supernova at a
642: density comparable to where resonant neutrino flavor conversion takes place in
643: the sun. In the supernova progenitor, this resonance will be located in
644: the helium/hydrogen envelope for typical neutrino energies. If the
645: supernova progenitor density profile is fairly smooth, then there will likely
646: be an adiabatic level crossing at this location, so that
647: the $\nu_e$
648: will wind up with an energy spectrum which would be measured as
649: $\sin^2 \theta_{solar}$ of the original
650: $\nu_e$ spectrum and $\cos^2 \theta_{solar}$ of the originally ``hotter''
651: $\nu_\mu$ and $\nu_\tau$ type spectrum.
652: The \lq\lq hopping probability\rq\rq\ at
653: resonance will be much smaller for the large angle solution than in the case
654: of the SMA.
655: For the LMA solution, even if the resonance is
656: completely nonadiabatic so that there is no level crossing
657: at resonance, the electron neutrinos will still become more energetic on
658: account of their large vacuum mixing with mu and tau neutrinos.
659:
660: With the LMA included,
661: the difference between the active-sterile neutrino conversion case considered
662: here and the case with no steriles is that part of the $\nu_e$ spectrum
663: simply disappears at some point during the deleptonization \lq\lq
664: cooling\rq\rq\ epoch of the proto-neutron star. At early times, when the
665: density profile is too steep for much active-sterile transformation to occur,
666: the ``cold'' part (or the part that was originally $\nu_e$) of the
667: spectrum will be present, and at late times, when the transformation
668: begins and the r-process takes place,
669: this part will disappear. We show this pictorially in
670: Fig. \ref{fig:earlylate}.
671:
672: In Fig. \ref{fig:earlylate} the bottom panel shows what happens to the neutrino
673: spectrum if $\nu_e \leftrightarrow \nu_s$
674: resonant neutrino flavor conversion is significant
675: and no subsequent vacuum mixing or envelope conversion takes place.
676: The solid curve corresponds to the early time, when the
677: density profile is too steep for much flavor transformation to occur and
678: the dashed curve corresponds to the late time when the evolution is
679: adiabatic.
680:
681: The top panel of Fig. \ref{fig:earlylate}
682: shows the case where the
683: LMA parameters provide the solution to the solar neutrino problem. We assume
684: completely adiabatic
685: transformation in the supernova progenitor envelope in this case. In this
686: panel, the original $\nu_e$ spectrum becomes mixed with the $\nu_\mu$ and
687: $\nu_\tau$, through the solar parameters $\sin^2 2 \theta_{solar} \def \sin^2
688: 2\theta_{12} \approx 0.8$. This is evident in the figure: note, for
689: example, the
690: longer tail on the distribution function. The early time (solid line) case,
691: where the $
692: \nu_e
693: \leftrightarrow \nu_s$ transformation is inoperative, has many more counts
694: at low energy than does the late time scenario. From this curve it is
695: evident that measuring the low energy
696: part of the neutrino spectrum would be most useful for determining whether
697: large scale active-sterile transformation occurs. A dramatic change in the
698: number of low energy neutrinos relative to the number of high energy neutrinos
699: would be an indication that the type
700: of active-sterile transformation discussed here was taking place.
701:
702: \subsection{The third active-active mixing angle}
703:
704: An interesting case occurs if the unknown parameters $\theta_{13}$ and
705: $\delta m_{13}^2$ are such that a second resonance occurs at a density
706: between the $\nu_e \leftrightarrow \nu_s$ transformation density and the
707: resonance region associated with
708: $\delta m^2_{12}$. In the case of complete neutrino flavor transformation, the
709: $\nu_e$'s are almost completely transformed into
710: $\nu_{\mu,\tau}$ and vice versa. In this case,
711: neutrinos which were originally electron flavor when they left the neutrino
712: sphere are almost completely coincident with the $\nu_3$ state.
713: This state evolves separately and does not encounter subsequent mixing
714: with the other flavors \cite{bf,cfq}.
715: In this special case of complete transformation
716: at the
717: $\delta m^2_{13}$ resonance, it would be more difficult
718: to recognize the effects of
719: a $\nu_e \leftrightarrow \nu_s$ transformation at a higher
720: density.
721:
722: \subsection{Detection Summary}
723: If only active-sterile transformation takes place, there
724: would be a dramatic deficit in the expected charged current signal in SNO.
725: However, with $\nu_e \leftrightarrow \nu_{\mu,\tau}$ transformations
726: included as
727: well, whether the detected signal is distinguishable from the case with no
728: steriles depends on the oscillation parameters and the density and electron
729: fraction profiles in the supernova. One signal of the active-sterile
730: mixing and
731: the LMA would be a decrease in the number of low energy neutrinos with time.
732:
733: We note that we have not considered in this section the mixing
734: of the sterile neutrino with the other (mu and tau) active neutrino species.
735: Such mixing opens a large new range of parameter space. Depend on the
736: $\delta m^2$s and mixing angles, this additional unknown
737: mixing could alter the
738: results presented in this section. Finally, density fluctuations,
739: both near the protoneutron star and in the progenitor envelope
740: may also change these results \cite{loreti}.
741:
742: \section{Background}
743: \label{sec:back}
744:
745: In this section we consider the effects of the neutrino background
746: potential. The neutrino-neutrino forward scattering-induced \lq\lq
747: background\rq\rq\ potential adds an important new twist to resonant neutrino
748: flavor conversion, rendering the problem severely nonlinear. This potential can
749: be written
750: \begin{equation}
751: V_\nu = 2\sqrt{2}G_F N_\nu.
752: \label{eq:nunu-v-def}
753: \end{equation}
754: Here $N_\nu$ is the {\bf effective} (because a neutrino's individual
755: contribution depends on the angle it makes with the \lq\lq
756: test\rq\rq\ neutrino)
757: net (neutrinos minus antineutrinos) neutrino number density. In the case of
758: active-sterile mixing, only the flavor basis on-diagonal terms entering
759: into Eq.
760: (\ref{eq:msw}) are non-vanishing, although in the general case of active-active
761: mixing, both on- and off-diagonal matrix elements of the neutrino-neutrino
762: forward scattering potential in the flavor basis could be nonzero
763: \cite{pantaleone:offd,sigl-raffelt:kinetic}. We only need to take account of
764: electron neutrinos forward-scattering on active flavors, since we
765: assume that the
766: matrix elements for electron neutrinos to forward-scatter off steriles are
767: negligible, as are the matrix elements for sterile neutrino forward-scattering
768: off other sterile species
769: \cite{sigl-raffelt:kinetic}. See
770: Ref.~\cite{qf:background} for a discussion of the background effect in
771: active-active neutrino mixing.
772:
773: The effective neutrino number density is written as
774: \begin{equation}
775: N_\nu = \int d^3 q \left( 1 - { p \cdot q \over |p| |q|} \right)
776: (\rho_{q,aa} - \bar{\rho}_{q,aa}).
777: \end{equation}
778: Here $q$ is the momentum of the background neutrino, $p$ is the
779: momentum of the test neutrino, $\rho_{q,aa}$ is the density matrix
780: element for neutrino-neutrino forward scattering and $\bar{\rho}_{q,aa}$ is the
781: corresponding density matrix element for the antineutrinos. In these
782: expressions for the density matrix elements, the subscript $aa$ denotes the
783: flavor-diagonal matrix element, so that $a=e$,$s$, electron or sterile flavor,
784: respectively. We write the density matrix elements as
785: \begin{equation}
786: \rho_{q,aa} = N(r,R_\nu,E_q) P(r,E_q,\psi)
787: \end{equation}
788: where $E_q$ is the energy of a neutrino with momentum $q$.
789: $N(r,R_\nu,E_q)$ is the number density of neutrinos
790: emitted into a pencil of direction and energy interval $d^3q$, corresponding
791: to the neutrino number flux (divided by the speed of light $c$) at a
792: given energy
793: in the pencil of directions centered on the angle
794: $\psi$. Here
795: $\psi$ is the angle of emission measured from the normal to the
796: neutrinosphere.
797: The function
798: $P(r,E_q,\psi)$ is the survival probability as a function of given
799: energy and emission angle.
800:
801: We find that for a test neutrino traveling at an angle $\alpha_t$
802: to the radial,
803: then $N_\nu$ can be written as
804: \begin{equation}
805: N_\nu (r,\alpha_t) =
806: 2\pi N_\nu^0
807: \int_0^\infty dE_\nu f_\nu(E)
808: \left( Q_0(r,E_\nu) - \cos\alpha_t Q_1(r,E_\nu) \right).
809: \label{eq:nveff}
810: \end{equation}
811: where
812: \label{eqs:ang-moments}
813: \begin{eqnarray}
814: Q_0(r,E_\nu) & = & \int_\mu^1 d\cos\alpha_b \, P(r,E,\alpha_b), \\
815: Q_1(r,E_\nu) & = & \int_\mu^1 d\cos\alpha_b \, P(r,E,\alpha_b) \cos\alpha_b,
816: \end{eqnarray}
817: Here, $\mu$ is defined as $\mu = \sqrt{1 - (R_\nu/r)^2}$, $\alpha_b$ is
818: the angle of the background neutrino with respect to the radial, and
819: $N_\nu^0$ is the number density of the neutrinos at the neutrino sphere.
820:
821: From these expressions it is clear that when including background effects,
822: the potential depends on the survival probability of neutrinos which
823: travel at different angles and take different paths.
824:
825: In calculations with the neutrino background potential included,
826: it is most convenient to evolve bilinears of the
827: wave functions. For much of the evolution, it is actually more efficient
828: to change to the matter basis and evolve with the angle-phase parameterization
829: ~\cite{pantaleone:isotropic}.
830:
831: To compute the neutrino potential at each time step, we must find the
832: active-active neutrino survival probability at each neutrino energy. In
833: the matter basis, the survival probability depends on the matter angle; but
834: in turn, the matter angle depends on the survival probability through
835: $V_\nu$. For a given set of matter states, then, we must find a $V_\nu$
836: which gives a consistent set of survival probabilities and matter angles.
837: Through most of the evolution, there is a unique neutrino potential which
838: is consistent with the matter states. When large numbers of neutrinos
839: begin to transform, though, there can be multiple consistent potentials.
840: In this range we eliminate the ambiguity by changing to the matter basis;
841: we change back when there is little transformation.
842:
843: \subsection{Radial Results}
844:
845: To simplify the problem, we have assumed that the survival probability does
846: not depend on the neutrino's direction. The next section discusses some
847: implications of relaxing this assumption. We do not simply use
848: $\alpha_t = 0$, since this underestimates the potential.
849: Instead, we take the survival probability to be independent of direction:
850: \begin{equation}
851: Q_0(E_\nu) = P(E_\nu)(1-\mu)
852: \end{equation}
853: and
854: \begin{equation}
855: Q_1(E_\nu) = P(E_\nu)(1-\mu^2)/2.
856: \end{equation}
857: The neutrino potential is then given as
858: \begin{equation}
859: N_\nu^{\mathrm{avg}} = 2 \pi N_\nu^0 (1-\mu)^2 \left( \frac{3+\mu}{4} \right)
860: \int_0^\infty dE_\nu f_\nu(E_\nu) P(E_\nu).
861: \label{eq:navg}
862: \end{equation}
863: To leading order, the geometric factor $(1-\mu)^2 (3+\mu)/4$ is equal to
864: $(R_\nu/r)^4/4$. This is double the leading-order potential we would find
865: by taking $\cos\alpha_t = 1$.
866:
867: A sample run using this potential is shown in Figure \ref{fig:back}. In this
868: figure we reduce the neutrino luminosities by a factor of ten from the
869: calculations in the previous sections. We do this to make the calculation
870: more tractable. However, we also note that these luminosities are not
871: unrealistic at very late times, where the r-process elements are likely
872: still being made ~\cite{cfq}.
873:
874: There are several effects stemming from the additional neutrino potential
875: provided by the neutrino forward scattering-induced background. One is that
876: there is somewhat more transformation in the innermost resonance
877: relative to the
878: no neutrino background case. The neutrino background potential changes more
879: slowly than the external potential provided by neutrino forward scattering on
880: the electrons: the result is to make neutrino amplitude evolution through the
881: resonance more adiabatic. Smaller values of
882: $\delta m^2 \sin^2 2\theta_v$ will reduce the adiabaticity, however. Overall,
883: though the increase in adiabaticity afforded by a significant neutrino
884: background potential implies more transformation of neutrino flavors at
885: resonance and does suggest that the effect of the neutrino background will be
886: to extend the epoch of significant neutrino flavor transformation out to
887: relatively smaller luminosities and, hence, later times than would be the case
888: with a purely electron-driven neutrino flavor transformation potential in
889: operation.
890:
891: Another effect of a significant neutrino background arises in the wind
892: epoch/region. As in the purely electron-driven neutrino flavor
893: transformation case, but not so monotonically, the resonance energy will tend to
894: sweep from low energy to high energy through the
895: $\nu_e$ energy distribution. As
896: $\nu_e$'s begin to transform,
897: $V_\nu$ decreases. Its decrease becomes faster than that of the external
898: potential; as $V_\nu$ decreases, the higher-energy resonances become
899: nonadiabatic. Therefore, with a significant neutrino background we have less
900: flavor transformation than in the no-background case. Finally, $V$
901: becomes small
902: enough that
903: $E_{res}$ is larger than the maximum $\nu_e$ energy; at this point,
904: we have swept
905: through the entire
906: $\nu_e$ population.
907:
908: After $E_{res}$ has passed through the $\nu_e$ distribution,
909: significant neutrino
910: flavor transformation mostly ceases. Except for the slow decrease of $Y_e$ and
911: $\rho$, there is nothing more to change the potential, and $\bar{\nu}_e$
912: transformation is minimal; even though the potential becomes negative, the
913: $\bar{\nu}_e$ resonant energy is always above the $\bar{\nu}_e$ distribution.
914: The
915: survival probabilities, then, are essentially fixed. $Y_e$ continues to
916: fall towards its equilibrium value, and in this example
917: (Fig. \ref{fig:back}), ends up fairly
918: low. However, even in this example there is a non-negligible alpha effect,
919: visible not as an upturn but a flattening of the $Y_e$ profile. At higher
920: luminosities, fewer neutrinos need to transform in order for $V_\nu$ to
921: balance the external potential. The increased $\bar{\nu}_e$ population then
922: blocks the lowering of $Y_e$ and causes a larger alpha effect. A sizable
923: radial background effect thus blocks our mechanism, or at least changes the
924: optimal $\delta m^2$ and $\sin^2 2 \theta_v$ at which it would occur.
925:
926: The larger the luminosity, of course,
927: the fewer the number of neutrinos which must transform in order to drive the
928: potential
929: $V$ to zero. However, large luminosity means $V_\nu$ will fall faster once
930: $\nu_e$'s start to transform to sterile species, and therefore the resonances
931: will tend to be less adiabatic. The situation is, in some ways, analogous to
932: the case where
933: $Y_e$ is set equal to its equilibrium value. In that case, as here with a
934: significant neutrino background, there is immediate feedback between survival
935: probabilities and the MSW potential $V$; in contrast, feedback
936: without a neutrino
937: background is substantially delayed. Because there is no delay in the feedback
938: with a significant neutrino background, once $V$ is driven near zero,
939: there is no
940: more reason for it to change---the system has found an equilibrium. An
941: independent calculation of the active-sterile background effect~\cite{mitesh}
942: finds similar effects on
943: $Y_e$.
944:
945: \subsection{Nonradial Speculations}
946:
947: When we use a radial treatment for the background, neutrinos evolve until
948: $V=0$, then stop. Prior treatments of background have also used this
949: approximation. However, the dominance of the fixed point at $V=0$ makes a
950: radial treatment unsuitable in our situation, because neutrinos coming in
951: at different angles will see different potentials. Therefore, a full
952: treatment will not have the same fixed point we see in the radial
953: approximation.
954:
955: A complete treatment of the background, including nonradial neutrino paths,
956: may be possible by an extension of the treatment described above. In
957: addition to dividing neutrinos into bins of energy, one would bin them
958: according to their emission angle $\psi$. Because of the spherical
959: symmetry, neutrinos with the same $\psi$ will encounter the same potential
960: (including neutrino potential), and have the same survival probability.
961: A full nonradial treatment will likely give substantially different results
962: than the radial one.
963:
964:
965: \section{Conclusions}
966: \label{sec:concl}
967:
968: We have considered the effect of active-sterile neutrino transformation
969: on the r-process in the neutrino driven wind environment.
970: Preliminary calculations show that it is possible to
971: drive the material sufficiently
972: neutron-rich that rapid neutron capture is possible.
973:
974: The effect on the neutrino signal from a nearby supernova depends very
975: much on the additional, active-active mixing parameters. In the absence of
976: this additional mixing, the effect on the neutrino signal would be dramatic.
977: If the r-process in the neutrino driven wind is made possible by active-sterile
978: transformation, then their would be a dramatic lack of electron neutrinos
979: coming from the supernova at late times, which would correspond to a reduction
980: in the neutrino signal of $\sim 90\%$.
981:
982: However, recent data from
983: Superkamiokande and SNO indicates that the additional active-active
984: transformations probably do take place. The most relevant for the
985: discussion here are the transformations involving $\nu_e$. In the case
986: of the SMA, we estimate about a 50\% tranformation for $\nu_e$ with
987: $\nu_{\mu,\tau}$. In the case of the LMA, depending on the
988: exact value of $\sin^2 2 \theta_{solar}$, more than 50\% of the electron
989: neutrinos are likely to transformation to $\nu_{\mu,\tau}$ and vice-versa.
990: In either case, the signal of active-sterile neutrino transformation
991: is more difficult to identify, but is likely to marked by a decrease of
992: low energy neutrinos at late time.
993:
994: We have also discussed background effects. We have done calculations
995: for reduced luminosity and with radial neutrinos. The latter
996: is the approximation
997: that all neutrinos see the same potential, regardless of their path.
998: We find
999: that the background effects tend to prevent the electron fraction
1000: from dropping to values as low as in the nonbackground case. This
1001: implies that the effect is most successful for the producing the r-process
1002: elements at late times, when the luminosity is low. A detailed
1003: investigation of neutrino background effects,
1004: using different methods and a different range of
1005: luminosities has been undertaken recently in Ref. \cite{mitesh}.
1006: Finally, we note that nonradial background effects, which take
1007: into account the different potentials seen by neutrinos traveling on different
1008: paths, is likely to be quite important, and we speculate that this may
1009: significantly alter the results from the radial case.
1010:
1011: \acknowledgements
1012:
1013: This work was supported in part by the U.S. National Science
1014: Foundation Grants No.\ PHY-0070161 at the University of Wisconsin and
1015: No. PHY-0099499 at UCSD, and in part by the University of Wisconsin Research
1016: Committee with funds granted by the Wisconsin Alumni Research
1017: Foundation. We thank the ECT* for their hospitality during part of the
1018: completion of this work.
1019:
1020: \begin{references}
1021:
1022: \bibitem{bbfh}
1023: E.M. Burbidge, G.R. Burbidge, W.A. Fowler, and F. Hoyle, Rev. Mod.
1024: Phys. {\bf 29}, 547 (1957); A. G. W. Cameron,
1025: Proc.~Astron.~Soc.~Pacific {\bf 69}, 201 (1957).
1026:
1027: \bibitem{r1}
1028: S.E. Woosley, J.R. Wilson, G.J. Mathews, R.D. Hoffman, and B.S.
1029: Meyer, Astrophys. J. {\bf 433}, 229 (1994); B.S. Meyer, W.M. Howard,
1030: G.J. Mathews, S.E. Woosley, and R.D. Hoffman, Astrophys. J. {\bf
1031: 399}, 656 (1992). K. Takahashi, J. Witti, and H.-Th.
1032: Janka, Astron. Astrophys. {\bf 286}, 857 (1994).
1033:
1034: \bibitem{hwq}
1035: R. D. Hoffman, S. E. Woosley, and Y.-Z. Qian, Astrophys. J. {\bf
1036: 482}, 951 (1996).
1037:
1038: \bibitem{mmf}
1039: B. S. Meyer, G. C. McLaughlin, and G. M. Fuller, Phys. Rev. C,
1040: {\bf 58} 3698 (1998), G. M. Fuller and B. S. Meyer
1041: Astrophys. J. {\bf 453} 792-809 (1995), G. C. McLaughlin,
1042: G. M. Fuller and J. R. Wilson, Astrophys. J. {\bf 472}
1043: 440 (1996).
1044:
1045: \bibitem{Qetal}
1046: Y.-Z. Qian, G. M. Fuller, G. J. Mathews, R. Mayle, and J. R. Wilson,
1047: Phys. Rev. Lett. {\bf 71}, 1965 (1993).
1048:
1049: \bibitem{cf} C. Cardall and G. M. Fuller, Astrophys.
1050: J. {\bf 486} L111 (1997), J. Pruet, G. M. Fuller, and C. Cardall.
1051: Astrophys. J. {\bf 561} 957-963 (2001), T. Thompson and
1052: A. Burrows, Nucl. Phys. {\bf A688} 377-381 (2001).
1053:
1054: \bibitem{bsm89}
1055: B. S. Meyer, ApJ, {\bf 343}, 254, (1989), C. Freiburghaus,
1056: S. Rosswog, F.-K. Thielemann, ApJ {\bf 525} L121 (1999)
1057:
1058: \bibitem{thielemann}
1059: S. Rosswog, M . Liebendorfer, F.-K. Thielemann, M. B. Davies,
1060: W. Benz, and T. Piran, Astron. Astrophys. {\bf 341} 499 (1999).
1061:
1062: \bibitem{janka99}
1063: Janka, H.-Th., Eberl, T., Ruffert, M. and Fryer, C., ApJ {\bf 527}
1064: L39, (1999)
1065:
1066: \bibitem{superk}
1067: SuperKamiokande collaboration, Y. Fukuda et al., Phys. Rev. Let.
1068: {\bf 81}, 1562 (1998).
1069:
1070: \bibitem{sno}
1071: SNO collaboration, Q. R. Ahmad et al., Phys. Rev. Lett.
1072: {\bf 87} 071301 (2001).
1073:
1074: \bibitem{bah}
1075: J. N. Bahcall, P. Krastev, and A. Yu Smirnov, Phys. Rev {\bf D58},
1076: 096016 (1998).
1077:
1078: \bibitem{lsnd}
1079: C. Athanassaopolous et al., Phys. Rev. Lett. {\bf 75} 2560 (1995);
1080: {\bf 77} 3082 (1996); {\bf 81} 1774 (1998); Phys. Rev. {\bf C54}
1081: 2685 (1996); {\bf 54} 2685 (1996); {\bf 58} 2489 (1998).
1082:
1083: \bibitem{gg}
1084: M. C. Gonzalaz-Garcia, M. Maltoni, C. Pena-Garay, and J. W. F. Valle,
1085: Phys. Rev. {\bf D63}, 033005 (2001).
1086:
1087: \bibitem{us}
1088: G. C. McLaughlin, J. M. Fetter, A. B. Balantekin and G. M. Fuller,
1089: Phys. Rev. C {\bf 59}, 2873.
1090:
1091: \bibitem{nunokawa}
1092: H. Nunokawa, J. T. Peltoniemi, A. Rossi, and J. W. F. Valle, Phys.
1093: Rev. D {\bf 56}, 1704 (1997).
1094:
1095: \bibitem{qw}
1096: Y.-Z. Qian and S. E. Woosley, Astrophys. J., {\bf 471},
1097: 331 (1996).
1098:
1099: \bibitem{wilson}
1100: R. W. Mayle and J. R. Wilson, unpublished (1993).
1101:
1102: \bibitem{meyer97}
1103: B. S. Meyer and J. S. Brown, Astrophys. J. Suppl. {\bf 112}, 199
1104: (1997).
1105:
1106: \bibitem{peres}
1107: O. L. G. Peres and A. Yu. Smirnov, Nuclear Physics {\bf B599} 2
1108: (2001).
1109:
1110: \bibitem{mitesh}
1111: M. Patel, unpublished (2001), M. Patel and G. M. Fuller, hep-ph/0003034
1112: (2000).
1113:
1114: \bibitem{superK} J. F. Beacom and P. Vogel, Phys. Rev. D {\bf 58},
1115: 053010 (1998).
1116:
1117: \bibitem{SNO} J. F. Beacom and P. Vogel, Phys. Rev. D {\bf 58},
1118: 093012 (1998).
1119:
1120: \bibitem{fhm} G. M. Fuller, W. C. Haxton, G. C. McLaughlin, Phys. Rev.
1121: {\bf D59} 085005 (1999).
1122:
1123: \bibitem{dighe}
1124: A. S. Dighe and A. Y. Smirnov, Phys. Rev. D {\bf 62} 033007
1125: (2000).
1126:
1127: \bibitem{woosley}
1128: S. E. Woosley, D. H., Hartmann, W. C. Hoffman and W. C. Haxton,
1129: Astrophys. J. {\bf 356}, 272 (1990).
1130:
1131: \bibitem{bf} A. B. Balantekin and G. M. Fuller, Phys. Lett.
1132: {\bf B471} 195 (1999).
1133:
1134: \bibitem{cfq} D. Caldwell, G. Fuller, and Y.-Z. Qian,
1135: Phys. Rev. {\bf D61} 123005 (2000).
1136:
1137: \bibitem{loreti}
1138: F. N. Loreti \& A.B. Balantekin, Phys. Rev. {\bf D50} 4762 (1994);
1139: F. N. Loreti, Y.Z. Qian, G.M. Fuller, A.B. Balantekin,
1140: Phys. Rev. {\bf D52} 6664 (1995);
1141: A. B. Balantekin, J. M. Fetter, F. N. Loreti,
1142: Phys. Rev. {\bf D54} 394 (1996).
1143:
1144: \bibitem{pantaleone:offd}
1145: J. Panteleone, Phys. Lett. {\bf B287}, 128 (1992).
1146:
1147: \bibitem{sigl-raffelt:kinetic}
1148: G. Sigl and G. Raffelt, Nucl. Phys. {\bf B406}, 423 (1993).
1149:
1150: \bibitem{qf:background}
1151: Y.-Z. Qian and G. M. Fuller, Phys. Rev. D {\bf 51}, 1479 (1995).
1152:
1153: \bibitem{pantaleone:isotropic}
1154: T. K. Kuo and J. Panteleone, Rev. Mod. Phys. {\bf 61}, 937 (1989);
1155: J. Panteleone, Phys. Rev. D {\bf 58}, 073002 (1998).
1156:
1157: \end{references}
1158:
1159: \newpage
1160:
1161: \begin{table}
1162: \begin{tabular}{|c|c|c|l|c|c|c|c|}
1163: \hline\hline
1164: & \\
1165: Complete $\nu_e \leftrightarrow \nu_s$ transformation and
1166: &
1167: $\nu_e$ spectrum \\ \hline\hline
1168: $\theta_{13}$,$\delta m^2_{13}$ no conversion & \\
1169: $\theta_{12}$,$\delta m^2_{12}$ adiabatic level crossing
1170: & 3/4 ``hot'' + 1/4 ``nothing'' \\ \hline
1171: $\theta_{13}$,$\delta m^2_{13}$ no conversion & \\
1172: $\theta_{12}$,$\delta m^2_{12}$ nonadiabatic level crossing &
1173: 1/4 ``hot'' + 3/4 ``nothing'' \\ \hline
1174: $\theta_{13}$,$\delta m^2_{13}$ complete conversion & \\
1175: $\theta_{12}$,$\delta m^2_{12}$ adiabatic level crossing
1176: & all ``hot'' \\ \hline
1177: $\theta_{13}$,$\delta m^2_{13}$ complete conversion & \\
1178: $\theta_{12}$,$\delta m^2_{12}$ nonadiabatic level crossing
1179: & all ``hot'' \\ \hline
1180: \hline
1181: & \\
1182: No steriles and
1183: & $\nu_e$ spectrum \\ \hline\hline
1184: $\theta_{13}$,$\delta m^2_{13}$ no conversion & \\
1185: $\theta_{12}$,$\delta m^2_{12}$ adiabatic level crossing &
1186: 3/4 ``hot'' + 1/4 ``cold'' \\ \hline
1187: $\theta_{13}$,$\delta m^2_{13}$ no conversion & \\
1188: $\theta_{12}$,$\delta m^2_{12}$ nonadiabatic level crossing &
1189: 1/4 ``hot'' + 3/4 ``cold'' \\ \hline
1190: $\theta_{13}$,$\delta m^2_{13}$ complete conversion & \\
1191: $\theta_{12}$,$\delta m^2_{12}$ adiabatic level crossing &
1192: all ``hot'' \\ \hline
1193: $\theta_{13}$,$\delta m^2_{13}$ complete conversion & \\
1194: $\theta_{12}$,$\delta m^2_{12}$ nonadiabatic level crossing &
1195: all ``hot'' \\ \hline
1196:
1197: \end{tabular}
1198: \vspace*{0.5cm}
1199: \caption{
1200: \label{tab:osc}
1201: Examples of various scenarios of additional neutrino oscillations in
1202: the envelope. The $\nu_e$ spectrum in the second column is the
1203: spectrum of the neutrinos as they arrive at the earth. We assume that
1204: the $\nu_\mu, \bar{\nu}_\mu, \nu_\tau, \bar{\nu}_\tau$ neutrinos
1205: have spectra with higher average energy (``hot''), then the electron
1206: neutrino (``cold'').
1207: We have associated the angle $\sin^2 2 \theta_{12}$
1208: with $\sin^2 2 \theta_{solar} \sim 0.8$ and the angle
1209: $\sin^2 2 \theta_{23}$ with
1210: $\sin^2 2 \theta_{atmospheric} \sim 1$.
1211: The angle $\theta_{13}$ is unknown and limited only by
1212: the reactor neutrino data.
1213: In this table we have neglected the possible consequences of mixing between
1214: the sterile and the other active flavors. In the case where there is
1215: complete transformation in the $\delta m^2_{13}$ resonance,
1216: there is no observable effect on the $\nu_e$ spectrum from the subsequent
1217: $\theta_{12}, \delta m^2_{12}$ mixing. If there is no transformation in
1218: the $\theta_{13}$ channel, then the $\nu_e \leftrightarrow \nu_s$
1219: oscillation solution may be detected as a loss of the low energy neutrinos,
1220: see Fig. \ref{fig:earlylate}.}
1221: \end{table}
1222:
1223: \newpage
1224: \clearpage
1225:
1226:
1227: \begin{figure}
1228: \centerline{\includegraphics[angle=0,width=14cm]{fig01.ps}}
1229: \caption{The electron fraction is plotted against distance from the
1230: center of the neutron star. The upper lines shows the evolution with no
1231: transformation. The lower lines shows the evolution of active-sterile mixing
1232: parameters of $\sin^2 \theta_v = 0.01$ and $\delta m^2 = 20 \, {\rm eV}^2$.
1233: The dashed lines show the equilibrium electron fraction while the
1234: solid lines show the actual electron fraction.
1235: }
1236: \label{fig:efraction1}
1237: \end{figure}
1238:
1239: \begin{figure}
1240: \centerline{\includegraphics[angle=0,width=14cm]{fig02.ps}}
1241: \vspace*{1cm}
1242: \caption{The electron fraction is plotted as in Fig. \ref{fig:equ},
1243: under the same neutrino mixing parameters; however, we have assumed that
1244: the electron fraction goes immediately to its equilbrium value, as
1245: set by the neutrino capture rates. In this figure only, we have neglected
1246: electron and positron captures. In the lower line, we have further
1247: neglected the formation of alpha particles. The upper line shows the effect
1248: of including alpha particles. The brief dip below 1/3 is an artifact of
1249: the code's finite step size.
1250: \label{fig:equ}}
1251: \end{figure}
1252:
1253: \begin{figure}
1254: \centerline{\includegraphics[width=14cm]{fig03.ps}}
1255: \caption{Contour plot of electron fraction as measured at the
1256: point where heavy nuclei begin to form Neutrino driven wind parameters
1257: employed here are $s/k = 100$, $\tau = 0.3 {\rm s}$.}
1258: \label{fig:efraction2}
1259: \end{figure}
1260:
1261: \begin{figure}
1262: \includegraphics[angle=0,width=14cm]{fig04.ps}
1263: \caption{Shows the effect of including non radial paths. The same
1264: wind parameters are used here as in Fig. \ref{fig:efraction2}. This
1265: shows the difference in $\nu_e$ survival probability at 11 km
1266: with and without nonradial neutrino paths. For large $\delta
1267: m^2 \sin^2 2 \theta_v$ (the upper right corner), most neutrinos
1268: convert in the radial scenario, so the radial and nonradial results
1269: converge.}
1270: \label{fig:nonradial}
1271: \end{figure}
1272:
1273:
1274: \begin{figure}
1275: \includegraphics[angle=-90,width=14cm]{fig05.ps}
1276: \caption{
1277: Ratios of events that would be seen in SNO from charged
1278: current electron neutrino break-up of the deuteron with active-sterile
1279: transformation to without active-sterile transformation.}
1280: \label{fig:sno1}
1281: \end{figure}
1282:
1283: \begin{figure}
1284: \centerline{\includegraphics[angle=-90,width=14cm]{fig06.ps}}
1285: \caption{Ratios of events that would be seen in SuperKamiokande from
1286: all types of neutrino scattering on electrons with active-sterile
1287: transformation to without active-sterile transformation.}
1288: \label{fig:elscat1}
1289: \end{figure}
1290:
1291: \begin{figure}
1292: \centerline{\includegraphics[angle=-90,width=14cm]{fig07.ps}}
1293: \caption{Ratios of events that would be seen in SNO from charged
1294: current electron neutrino break-up of the deuteron with both active-sterile
1295: transformation and the active-active SMA solar solution to
1296: without any transformation.}
1297: \label{fig:sno2}
1298: \end{figure}
1299:
1300: \begin{figure}
1301: \centerline{\includegraphics[angle=-90,width=14cm]{fig08.ps}}
1302: \caption{Ratios of events that would be seen in SuperKamiokande from
1303: all types of neutrino scattering on electrons with both active-sterile
1304: transformation and the active-active SMA solar solution to
1305: without any transformation.}
1306: \label{fig:elscat2}
1307: \end{figure}
1308:
1309: \newpage
1310:
1311: \begin{figure}
1312: \vspace*{-0.5cm}
1313: \includegraphics[width=12cm]{fig09.ps}
1314: \vspace*{-1cm}
1315: \caption{The bottom panel shows an example of early (solid line) and late
1316: (dashed line) electron neutrino spectrum for the active-sterile transformation
1317: solution to the r-process. No other transformations with the electron neutrino
1318: are considered. The top panel shows an example of
1319: the early (solid line) and late (dashed line) electron neutrino spectrum
1320: when both the $\nu_e \leftrightarrow \nu_s$ and the LMA solution to the
1321: solar neutrino problem are considered. In this panel we assume completely
1322: adiabatic transformation through the $\delta m_{12}$ resonance region. The
1323: signature of this type of active sterile transformation is a relative
1324: decrease in the number of low energy neutrinos from early to late
1325: times when compared with the number of high energy neutrinos. }
1326: \label{fig:earlylate}
1327: \end{figure}
1328:
1329: \newpage
1330:
1331: \begin{figure}
1332: \includegraphics[angle=0,width=14cm]{fig10.ps}
1333: \caption{ A sample run with the background turned on. The top
1334: panel shows the actual (solid line) and equilibrium (dashed line)
1335: electron fraction. The middle panel shows the MSW potential and
1336: the bottom panel shows the average survival probabilities for neutrinos
1337: (solid line) and antineutrinos (dashed line).
1338: The parameters used in this run were $\delta m^2 = 20 \; {\rm eV}^2$ and
1339: $\sin^ 2 \theta_v = 10^{-2}$. The neutrino luminosities were
1340: $1.08 \times 10^{50}
1341: {\rm erg} {\rm s}^{-1}$ for the neutrinos and
1342: $1.3 \times 10^{50} {\rm erg} \, {\rm s}^{-1}$ for the antineutrinos.
1343: }
1344: \label{fig:back}
1345: \end{figure}
1346:
1347: \end{document}
1348:
1349: