1: \documentclass[12pt,a4paper]{article}
2: \usepackage{fullpage}
3: %\usepackage{showkeys}
4: \usepackage{cite}
5: \usepackage[centertags]{amsmath}
6: \allowdisplaybreaks[4] \numberwithin{equation}{section}
7: \usepackage{amssymb}
8: \usepackage{bm}
9: \usepackage{bbm}
10: \usepackage{mathrsfs}
11: \usepackage{float}
12: \usepackage{graphicx}
13: \usepackage{url}
14: \usepackage[dvips,hyperindex]{hyperref}
15: \usepackage{tocbibind}
16: \usepackage{subfigure}
17: \renewcommand{\subfigtopskip}{0pt}
18: \renewcommand{\subfigcapskip}{0pt}
19: \renewcommand{\subfigbottomskip}{0pt}
20: \renewcommand{\subcapsize}{\small}
21: \renewcommand{\subcaplabelfont}{\sffamily}
22: \pagestyle{plain}
23: %
24: \newcommand{\comment}[1]{
25: \begin{center}
26: \begin{minipage}[t]{0.8\textwidth}
27: \hrule \textbf{ \centerline{\fbox{\textsf{Comment}}} #1
28: \smallskip
29: } \hrule
30: \end{minipage}
31: \end{center}
32: }
33: \newcommand{\bos}[1]{\overset{\scriptscriptstyle(-)}{#1}}
34: \newcommand{\boss}[2]{\ensuremath{\rlap{\kern-2.5pt\ensuremath{\overset{\scriptscriptstyle(-)}{\phantom{#1}}}}{\ensuremath{{#1}_{#2}}}}}
35: \newcommand{\vet}[1]{\ensuremath{\hskip-1pt\vec{\hskip1pt#1}}}
36: %
37: \begin{document}
38:
39: \begin{flushright}
40: %\textsf{\today}
41: \textsf{1 September 2009}
42: \\
43: \textsf{arXiv:0810.5443v3}
44: \\
45: \textsf{Phys. Rev. D 80 (2009) 053009}
46: \end{flushright}
47:
48: \vspace{1cm}
49:
50: \begin{center}
51: \large \textbf{Bayesian Constraints on $\bm{\vartheta_{13}}$ from Solar and KamLAND Neutrino Data} \normalsize
52: \\[0.5cm]
53: \large H.L. Ge$^{a}$, C. Giunti$^{b}$, Q.Y. Liu$^{a}$ \normalsize
54: \\[0.5cm]
55: \setlength{\tabcolsep}{2pt}
56: \begin{tabular}{ll}
57: $^{a}$ & Department of Modern Physics, University of Science and
58: \\
59: & Technology of China, Hefei, Anhui 230026, China
60: \\
61: \\
62: $^{b}$ & INFN, Sezione di Torino, Via P. Giuria 1, I--10125 Torino, Italy
63: \end{tabular}
64: \\[0.5cm]
65: \begin{minipage}[t]{0.8\textwidth}
66: \begin{center}
67: \textbf{Abstract}
68: \end{center}
69: We present the results of a Bayesian analysis of solar and
70: KamLAND neutrino data in the framework of three-neutrino mixing.
71: We adopt two approaches for the prior probability distribution of the oscillation parameters
72: $\Delta{m}^{2}_{21}$, $\sin^{2}\vartheta_{12}$, $\sin^{2}\vartheta_{13}$:
73: 1)
74: a traditional flat uninformative prior;
75: 2)
76: an informative prior which describes the limits on $\sin^{2}\vartheta_{13}$ obtained in
77: atmospheric and long-baseline accelerator and reactor neutrino experiments.
78: In both approaches, we present the allowed regions in the
79: $\sin^{2}\vartheta_{13}$--$\Delta{m}^{2}_{21}$ and
80: $\sin^{2}\vartheta_{12}$--$\sin^{2}\vartheta_{13}$ planes,
81: as well as the marginal posterior probability
82: distribution of $\sin^{2}\vartheta_{13}$.
83: We confirm the $1.2\sigma$ hint of $\vartheta_{13}>0$ found in \texttt{hep-ph/0806.2649}
84: from the analysis of solar and KamLAND neutrino data.
85: We found that the statistical significance of the hint is reduced to about $0.8\sigma$
86: by the constraints on $\sin^{2}\vartheta_{13}$ coming from atmospheric and long-baseline accelerator and reactor neutrino data,
87: in agreement with \texttt{hep-ph/0808.2016}.
88: \end{minipage}
89: \end{center}
90: % \begin{flushleft}
91: % \textbf{PACS Numbers:} \texttt{26.65.+t, 14.60.Pq, 14.60.Lm, 02.50.Tt}
92: % \\
93: % \textbf{Keywords:} \textsf{Solar Neutrinos, Neutrino Oscillations, Neutrino Mass, Neutrino Mixing, Bayesian Probability Theory}
94: % \end{flushleft}
95:
96: \section{Introduction}
97: \label{Introduction}
98:
99: Neutrino oscillation experiments have shown that
100: neutrinos are massive and mixed particles
101: (see
102: Refs.~\cite{hep-ph/9812360,hep-ph/0310238,Giunti-Kim}).
103: %Refs.~\cite{Bilenky:1978nj,Bilenky:1987ty,hep-ph/9812360,hep-ph/0202058,hep-ph/0211462,hep-ph/0310238,hep-ph/0405172,hep-ph/0506083,hep-ph/0606054,Giunti-Kim}).
104: Solar and KamLAND neutrino experiments
105: observed
106: $\boss{\nu}{e}\to\boss{\nu}{\mu,\tau}$
107: transitions due to neutrino oscillations generated by
108: a squared-mass difference
109: \begin{equation}
110: \Delta{m}^{2}_{\text{SOL}}
111: \simeq 8 \times 10^{-5} \, \text{eV}^{2}
112: \,.
113: \label{SOL}
114: \end{equation}
115: Atmospheric and long-baseline accelerator neutrino experiments
116: measured
117: $\boss{\nu}{\mu}\to\boss{\nu}{\tau}$
118: transitions due to neutrino oscillations generated by
119: a squared-mass difference
120: \begin{equation}
121: \Delta{m}^{2}_{\text{ATM}}
122: \simeq
123: 2.5 \times 10^{-3} \, \text{eV}^{2}
124: \,.
125: \label{ATM}
126: \end{equation}
127: Hence, there is a hierarchy of squared-mass differences:
128: \begin{equation}
129: \Delta{m}^{2}_{\text{ATM}}
130: \simeq
131: 30 \, \Delta{m}^{2}_{\text{SOL}}
132: \,.
133: \label{102}
134: \end{equation}
135: This hierarchy is easily accommodated in the framework of three-neutrino mixing,
136: in which there are two independent squared-mass differences.
137: We label the neutrino masses in order to have
138: \begin{align}
139: \null & \null
140: \Delta{m}^{2}_{\text{SOL}}
141: \equiv
142: \Delta{m}^{2}_{21}
143: \,,
144: \label{103}
145: \\
146: \null & \null
147: \Delta{m}^{2}_{\text{ATM}}
148: \simeq
149: |\Delta{m}^{2}_{31}|
150: \simeq
151: |\Delta{m}^{2}_{32}|
152: \,.
153: \label{104}
154: \end{align}
155: The two possible schemes are illustrated in Fig.~\ref{m008}.
156: They differ by the sign of
157: $
158: \Delta{m}^{2}_{31}
159: \simeq
160: \Delta{m}^{2}_{32}
161: $.
162:
163: \begin{figure}[t!]
164: \begin{center}
165: \renewcommand{\arraystretch}{2}
166: \setlength{\tabcolsep}{2cm}
167: \begin{tabular}{cc}
168: \begin{minipage}[c]{0.25\textwidth}
169: \includegraphics*[bb=175 469 415 779, width=\linewidth]{3nun.eps}
170: \end{minipage}
171: &
172: \begin{minipage}[c]{0.25\textwidth}
173: \includegraphics*[bb=180 469 420 779, width=\linewidth]{3nui.eps}
174: \end{minipage}
175: \\
176: \underline{NORMAL}
177: &
178: \underline{INVERTED}
179: \end{tabular}
180: \end{center}
181: \caption{ \label{m008}
182: The two three-neutrino schemes allowed by the hierarchy
183: $\Delta{m}^{2}_{\text{SOL}} \ll \Delta{m}^{2}_{\text{ATM}}$.
184: }
185: \end{figure}
186:
187: For the $3\times3$ unitary mixing matrix of neutrinos
188: we adopt the standard parameterization in Eq.~(\ref{f035}) of Appendix~\ref{Regeneration} \cite{Chau:1984fp,PDG-2006}.
189: The negative results of the
190: Chooz \cite{hep-ex/0301017} and Palo Verde \cite{hep-ex/0107009} long-baseline neutrino oscillation experiments,
191: together with the evidence of neutrino oscillations in atmospheric and long-baseline accelerator neutrino experiments,
192: imply that the mixing angle
193: $\vartheta_{13}$ is small
194: \cite{hep-ph/9802201}
195: (see Ref.~\cite{hep-ph/0808.2016} for updated bounds).
196: On the other hand, the values of the other two mixing angles
197: are known to be large from
198: the results
199: of solar and KamLAND experiments ($\vartheta_{12}$)
200: and
201: the results
202: of atmospheric and long-baseline accelerator neutrino experiments ($\vartheta_{23}$).
203:
204: In Ref.~\cite{hep-ph/0605195}
205: we presented the results of a Bayesian analysis
206: of the solar and KamLAND neutrino data
207: in the framework of two-neutrino mixing,
208: which is obtained from three-neutrino mixing in the approximation
209: of negligible $\vartheta_{13}$.
210: In this paper,
211: we extend our Bayesian analysis to the framework of three-neutrino mixing,
212: aiming at the determination of the constraints on the value of the small mixing angle
213: $\vartheta_{13}$
214: implied by solar and KamLAND neutrino data.
215:
216: The plan of the paper is as follows. In Section~\ref{chi2 analysis}
217: we present the constraints on the value of $\vartheta_{13}$ in a
218: standard $\chi^{2}$ analysis, to be compared with the Bayesian
219: results with a uninformative prior presented in Section~\ref{Bayesian Analysis}.
220: In Section~\ref{Informative} we present the results obtained with an informative prior
221: which represents information on $\vartheta_{13}$
222: obtained in atmospheric and long-baseline accelerator and reactor neutrino experiments,
223: independently from solar and KamLAND neutrino data.
224: The conclusions are given in
225: Section~\ref{Conclusions}.
226:
227: \section{$\chi^{2}$ Analysis}
228: \label{chi2 analysis}
229:
230: \begin{figure}[t!]
231: \includegraphics*[scale=0.44]{kai1.eps}
232: \hfill
233: \includegraphics*[scale=0.44]{kai2.eps}
234: \caption{ \label{f1chi} The 90\%, 95\%, and 99.73\% C.L. regions in
235: the $\sin^{2}\vartheta_{13}$--$\Delta{m}^{2}_{21}$ plane obtained
236: in the least-squares analysis. The solid and dotted lines enclose,
237: respectively, the regions obtained with solar data and KamLAND data.
238: The shadowed areas are obtained from the combined analysis of solar
239: and KamLAND data. The figure on the right is an enlargement of the
240: interesting area of the figure on the left.
241: The dot, cross, and asterisk indicate, respectively, the best-fit points of the solar, KamLAND, and combined analyses.}
242: \end{figure}
243:
244: \begin{figure}[t!]
245: \begin{center}
246: \includegraphics*[scale=0.6]{kai3.eps}
247: \end{center}
248: \caption{ \label{f2chi} The 90\%, 95\%, and 99.73\% C.L.
249: regions in the
250: $\sin^{2}\vartheta_{12}$--$\sin^{2}\vartheta_{13}$ plane
251: obtained in the least-squares analysis. The solid and dotted lines
252: enclose, respectively, the regions obtained with solar data and
253: KamLAND data. The shadowed areas are obtained from the combined
254: analysis of solar and KamLAND data.
255: The dot, cross, and asterisk indicate, respectively, the best-fit points of the solar, KamLAND, and combined analyses.}
256: \end{figure}
257:
258: \begin{figure}[t!]
259: \begin{center}
260: \includegraphics*[scale=0.6]{kai5.eps}
261: \end{center}
262: \caption{ \label{f3chi} $\Delta\chi^2$
263: as a function of $\sin^{2}\vartheta_{13}$. The straight horizontal
264: lines show the levels corresponding to 90\%, 95\%, and 99.73\% C.L.. }
265: \end{figure}
266:
267: The traditional way to extract information on the neutrino mixing parameters from solar neutrino data
268: is based on the standard least-squares method,
269: also called
270: ``$\chi^{2}$ analysis''.
271: The least-squares function $\chi^{2}$ is given by
272: $ \chi^{2} = - 2 \ln\mathcal{L} + \text{constant} $,
273: where $\mathcal{L}$ is the likelihood function.
274: In this section we present our results in a traditional least-squares analysis of solar and KamLAND neutrino data.
275: The least-squares function described in this Section will be used in the Bayesian analysis
276: presented in Section~\ref{Bayesian Analysis} for the calculation of the sampling probability distribution,
277: which is proportional to the likelihood function.
278:
279: We consider the data of the following solar neutrino experiments:
280: Homestake \cite{Cleveland:1998nv}, GALLEX/GNO \cite{hep-ex/0504037},
281: SAGE \cite{astro-ph/0204245}, Super-Kamiokande
282: \cite{hep-ex/0508053}\cite{hep-ex/0803.4312}, and SNO
283: \cite{nucl-ex/0502021}. The least-squares function of solar neutrino
284: data is given by
285: \begin{equation}
286: \chi^{2}_{\text{S}}
287: =
288: \sum_{i,j=1}^{N_{\text{S}}}
289: (R_{i}^{\text{exp}}-R_{i}^{\text{th}}) \, (V_{\text{S}}^{-1})_{ij} \, (R_{j}^{\text{exp}}-R_{j}^{\text{th}})
290: \,.
291: \label{chiS}
292: \end{equation}
293: Here $R_{i}^{\text{exp}}$ are the solar data points, whose number is
294: $N_{\text{S}}=80$, accounted as follows:
295: \begin{itemize}
296: %
297: \item
298: the rate of the Homestake ${}^{37}\text{Cl}$ experiment \cite{Cleveland:1998nv};
299: %
300: \item
301: the combined rate of the ${}^{71}\text{Ga}$ experiments GALLEX/GNO \cite{hep-ex/0504037} and SAGE \cite{astro-ph/0204245},
302: %
303: \item
304: the day and night energy spectra of the Super-Kamiokande experiment
305: \cite{hep-ex/0508053}($21+21$ bins) and Super-Kamiokande
306: experiment\cite{hep-ex/0803.4312} ;
307: %
308: \item
309: the day and night energy spectra of charged-current events in the SNO experiment \cite{nucl-ex/0502021}
310: ($17+17$ bins);
311: %
312: \item
313: the neutral-current event rate in the salt phase of the SNO experiment \cite{nucl-ex/0502021}.
314: %
315: \item
316: the NCD neutral-current event rate in the SNO experiment \cite{nucl-ex/0806.0989}.
317: %
318: \end{itemize}
319: The corresponding theoretical rates $R_{i}^{\text{th}}$ depend on the neutrino oscillation parameters.
320: The covariance error matrix $V_{\text{S}}$ takes into account the
321: correlations of theoretical uncertainties, according to the discussions in
322: Refs.~\cite{Fogli:1994nn,hep-ph/9912231,hep-ph/0006026,hep-ph/0108191}.
323: In our analysis, the initial flux of $^8\text{B}$ solar neutrinos is considered
324: as a free parameter to be determined by the fit,
325: mainly through the SNO neutral-current data.
326: For the other solar neutrino fluxes, we assume the
327: BP04 Standard Solar Model \cite{astro-ph/0402114}.
328: The transition probability in the Sun
329: is calculated using the standard method \cite{Shi:1992zw}
330: based on the hierarchy of squared-mass differences in Eq.~(\ref{102}),
331: which implies that the oscillations generated by the large mass-squared difference
332: $\Delta{m}^{2}_{\text{ATM}}$ are averaged out
333: (see
334: Refs.~\cite{hep-ph/9812360,hep-ph/0310238,Giunti-Kim}).
335: %Refs.~\cite{hep-ph/9812360,hep-ph/0405172,hep-ph/0506083,hep-ph/0606054,Giunti-Kim}).
336: For the calculation of the regeneration of solar $\nu_{e}$'s in the Earth,
337: we use Eq.~(\ref{e001}),
338: derived in Appendix~\ref{Regeneration}.
339:
340: Neutrino oscillations due to the same mixing parameters which
341: generate the oscillations of solar neutrinos have been observed in
342: the KamLAND very-long-baseline reactor neutrino oscillation
343: experiment \cite{hep-ex/0801.4589}.
344: The KamLAND least-squares function is\footnote{
345: In Ref.~\cite{hep-ph/0605195} and in the first version of this paper
346: (\texttt{arXiv:0810.5443v1})
347: we adopted a different least-squares function, which is appropriate for a Poisson distribution
348: (see Refs.~\cite{Baker:1984tu,PDG-2006}).
349: We think that the Gaussian least-squares function in Eq.~(\ref{chiK}) is more appropriate
350: for the analysis of KamLAND data, because it allows us to take into account
351: the systematic uncertainty in each energy bin,
352: as discussed in Ref.~\cite{hep-ph/0212129}.
353: }
354: \begin{equation}
355: \chi^{2}_{\text{S}}
356: =
357: \sum_{i,j=1}^{N_{\text{K}}}
358: (N_{i}^{\text{exp}}-N_{i}^{\text{th}}) \, (V_{\text{K}}^{-1})_{ij} \, (N_{j}^{\text{exp}}-N_{j}^{\text{th}})
359: \,,
360: \label{chiK}
361: \end{equation}
362: where
363: $N_{\text{K}}=17$ is the number of energy bins,
364: $N_i^{\text{exp}}$ is the number of events measured in the
365: $i\text{th}$ bin and $N_i^{\text{th}}$ is the corresponding
366: theoretical value, which depends on the neutrino oscillation
367: parameters.
368: The covariance error matrix $V_{\text{K}}$ takes into account the statistical uncertainties and the
369: correlated and uncorrelated systematic
370: uncertainties, all added in quadrature.
371:
372: The global least-squares function is
373: \begin{equation}
374: \chi^{2}_{\text{T}}
375: =
376: \chi^{2}_{\text{S}}
377: +
378: \chi^{2}_{\text{K}}
379: \,.
380: \label{chi2}
381: \end{equation}
382: We minimized $\chi^{2}_{\text{T}}$ with respect to the three mixing
383: parameters $\Delta{m}^{2}_{21}$, $\sin^{2}\vartheta_{12}$, and
384: $\sin^{2}\vartheta_{13}$. We found the best-fit point
385: \begin{equation}
386: \Delta{m}^{2}_{21} = 7.58\times 10^{-5} \, \text{eV}^{2} \,, \quad
387: \sin^{2}\vartheta_{12} = 0.31 \,, \quad \sin^{2}\vartheta_{13} =
388: 0.021\,. \label{chi2bestfit}
389: \end{equation}
390: The 90\%, 95\%, and 99.73\% C.L. regions in the
391: $\sin^{2}\vartheta_{13}$--$\Delta{m}^{2}_{21}$ and
392: $\sin^{2}\vartheta_{12}$--$\sin^{2}\vartheta_{13}$ planes are
393: shown, respectively, in Figs.~\ref{f1chi} and \ref{f2chi}.
394: These regions correspond to 2 degrees of freedom.
395: The third parameter ($\vartheta_{12}$ in Fig.~\ref{f1chi} and $\Delta{m}^2_{12}$ in Fig.~\ref{f2chi})
396: is marginalized by minimizing $\chi^{2}_{\text{T}}$.
397:
398: From Figs.~\ref{f1chi} and \ref{f2chi},
399: one can see that KamLAND data constrain $\vartheta_{13}$ more than solar data.
400:
401: Figure~\ref{f3chi} shows the difference $\Delta\chi^2$ of $\chi^2$
402: from its minimum value as a function of $\sin^{2}\vartheta_{13}$.
403: The resulting 90\%, 95\%, and 99.73\% C.L. upper bounds for
404: $\sin^{2}\vartheta_{13}$, determined by the intersection of the
405: $\Delta\chi^2$ curve with the straight horizontal lines in
406: Fig.~\ref{f3chi}, are, respectively,
407: \begin{equation}
408: \sin^{2}\vartheta_{13} < 0.051 \, (90\%) \,,\quad 0.057 \, (95\%)
409: \,,\quad 0.076 \, (99.73\%) \,. \label{t01}
410: \end{equation}
411:
412: It is interesting to note that the best-fit point for
413: $\sin^{2}\vartheta_{13}$ in Eq.~(\ref{chi2bestfit}) is slightly
414: larger than zero, in agreement with the value obtained in
415: Ref.~\cite{hep-ph/0806.2649}
416: (see also Refs.~\cite{hep-ph/0804.3345,hep-ph/0808.2016}), $ \sin^{2}\vartheta_{13} = 0.021 \pm
417: 0.017 $. Since we have $ \sin^{2}\vartheta_{13} = 0.021 \pm 0.018 $,
418: our hint of $\vartheta_{13}>0$ is at the $1.2\sigma$ level
419: (the precise value of $\Delta\chi^2$ for $\vartheta_{13}=0$ is 1.33, corresponding to $1.15\sigma$).
420:
421: \section{Bayesian Analysis}
422: \label{Bayesian Analysis}
423:
424: \begin{figure}[t!]
425: \includegraphics*[scale=0.44]{bay1.eps}
426: \hfill
427: \includegraphics*[scale=0.44]{bay2.eps}
428: \caption{ \label{f01} The 90\%,
429: 95\%, and 99.73\% Bayesian credible regions in the
430: $\sin^{2}\vartheta_{13}$--$\Delta{m}^{2}_{21}$ plane obtained with
431: an uninformative constant prior probability distribution. The solid
432: and dotted lines enclose, respectively, the credible regions
433: obtained with solar data and KamLAND data. The shadowed areas are
434: obtained from the combined analysis of solar and KamLAND data. The
435: figure on the right is an enlargement of the interesting area of the
436: figure on the left.
437: The dot, cross, and asterisk indicate, respectively, the best-fit points of the solar, KamLAND, and combined analyses.}
438: \end{figure}
439:
440: \begin{figure}[t!]
441: \begin{center}
442: \includegraphics*[scale=0.6]{bay3.eps}
443: \end{center}
444: \caption{ \label{f02} The 90\%, 95\%, and 99.73\% Bayesian credible
445: regions in the
446: $\sin^{2}\vartheta_{12}$--$\sin^{2}\vartheta_{13}$ plane
447: obtained with an uninformative constant prior probability
448: distribution. The solid and dotted lines enclose, respectively, the
449: regions obtained with solar data and KamLAND data. The shadowed
450: areas are obtained from the combined analysis of solar and KamLAND
451: data.
452: The dot, cross, and asterisk indicate, respectively, the best-fit points of the solar, KamLAND, and combined analyses.}
453: \end{figure}
454:
455: In the Bayesian approach, the analysis of the data allows us to
456: calculate the posterior probability distribution of the mixing
457: parameters, assuming a prior probability distribution which
458: quantifies the prior knowledge. Denoting with $ \mathcal{M} = \{
459: \Delta{m}^{2}_{21}, \sin^{2}\vartheta_{12}, \sin^{2}\vartheta_{13}
460: \} $ the set of mixing parameters to be determined by the analysis,
461: the normalized posterior probability distribution of the mixing
462: parameters is given by
463: \begin{equation}
464: p(\mathcal{M}|\mathcal{D},\mathcal{I})
465: =
466: \frac{
467: p(\mathcal{D}|\mathcal{M},\mathcal{I})
468: \,
469: p(\mathcal{M}|\mathcal{I})
470: }{
471: \int
472: \text{d}\mathcal{M}
473: \,
474: p(\mathcal{D}|\mathcal{M},\mathcal{I})
475: \,
476: p(\mathcal{M}|\mathcal{I})
477: }
478: \,,
479: \label{B1}
480: \end{equation}
481: where $p(\mathcal{D}|\mathcal{M},\mathcal{I})$ is the sampling
482: probability distribution, $p(\mathcal{M}|\mathcal{I})$ is the prior
483: probability distribution, and $ \text{d}\mathcal{M} \equiv
484: \text{d}\Delta{m}^{2}_{21} \, \text{d}\sin^{2}\vartheta_{12} \,
485: \text{d}\sin^{2}\vartheta_{13} $. The symbols $\mathcal{D}$ and
486: $\mathcal{I}$ represent, respectively, the experimental data and all
487: the prior general knowledge and assumptions on solar and neutrino
488: physics.
489:
490: The sampling probability distribution is given by
491: \begin{equation}
492: p(\mathcal{D}|\mathcal{M},\mathcal{I})
493: \propto
494: \left( |V_{\text{S}}| |V_{\text{K}}| \right)^{-1/2} e^{-\chi^{2}_{\text{T}}/2}
495: \,,
496: \label{sampling}
497: \end{equation}
498: with the least-squares function $\chi^{2}_{\text{T}}$ given in Eq.~(\ref{chi2}).
499: The normalization factor is irrelevant,
500: since it cancels in Eq.~(\ref{B1}).
501: We retained only the coefficient $\left( |V_{\text{S}}| |V_{\text{K}}| \right)^{-1/2}$,
502: which depends on the neutrino mixing parameters in $\mathcal{M}$
503: (see Ref.~\cite{hep-ph/0108191}).
504:
505: In Ref.~\cite{hep-ph/0605195} we have shown that the choices of
506: constant uninformative priors in the
507: $\sin^{2}\vartheta_{12}$--$\Delta{m}^{2}_{21}$ or
508: $\log\sin^{2}\vartheta_{12}$--$\log\Delta{m}^{2}_{21}$ planes are
509: practically equivalent, because of the excellent quality of the
510: data. Hence, in the three-neutrino mixing analysis we assume a
511: constant prior in the three-dimensional space of the parameters
512: $\Delta{m}^{2}_{21}$, $\sin^{2}\vartheta_{12}$, and
513: $\sin^{2}\vartheta_{13}$.
514:
515: Figures~\ref{f01} and \ref{f02} show the
516: resulting credible regions with 90\%, 95\%, and 99.73\% probability in the
517: $\sin^{2}\vartheta_{13}$--$\Delta{m}^{2}_{21}$ and
518: $\sin^{2}\vartheta_{12}$--$\sin^{2}\vartheta_{13}$ planes,
519: respectively.
520: A credible region is the smallest region with the given integral posterior probability.
521: In practice, a credible region is calculated as the two-dimensional region
522: surrounded by an isoprobability contour
523: which contains the point of highest posterior probability.
524: In each plane of parameters, the probability distribution is calculated by integrating
525: $p(\mathcal{M}|\mathcal{D},\mathcal{I})$
526: over the third parameter
527: ($\sin^{2}\vartheta_{12}$ in the plane $\sin^{2}\vartheta_{13}$--$\Delta{m}^{2}_{21}$
528: and
529: $\Delta{m}^2_{12}$ in the plane $\sin^{2}\vartheta_{12}$--$\sin^{2}\vartheta_{13}$).
530:
531: The credible regions in Figs.~\ref{f01} and \ref{f02}
532: are similar but slightly lager than the $\chi^2$-allowed regions in Figs.~\ref{f1chi} and \ref{f2chi}.
533: The comparison of the two types of region is shown in Fig.~\ref{f1bc} and \ref{f2bc}.
534: It is fair to conclude that the Bayesian analysis with an uninformative prior
535: confirms the results obtained with the traditional $\chi^2$ analysis.
536:
537: Figure~\ref{f03} shows the marginal posterior probability
538: distribution of $\sin^{2}\vartheta_{13}$, which is given by
539: \begin{equation}
540: p(\sin^{2}\vartheta_{13}|\mathcal{D},\mathcal{I}) = \int
541: \text{d}\Delta{m}^{2}_{21} \int \text{d}\sin^{2}\vartheta_{12}
542: \, p(\mathcal{M}|\mathcal{D},\mathcal{I}) \,. \label{t02}
543: \end{equation}
544: The resulting credible upper bounds for $\sin^{2}\vartheta_{13}$ with
545: 90\%, 95\%, and 99.73\% probability are, respectively,
546: \begin{equation}
547: \sin^{2}\vartheta_{13} < 0.048 \, (90\%) \,,\quad 0.054 \, (95\%)
548: \,,\quad 0.075 \, (99.73\%) \,. \label{t03}
549: \end{equation}
550: These limits are similar to those in Eq.~(\ref{t01}),
551: in agreement with the above conclusion that
552: an uninformative-prior Bayesian analysis
553: confirms the results obtained with a $\chi^2$ analysis.
554:
555: We investigated also the hint of $\vartheta_{13}>0$ in the Bayesian approach.
556: Since the probability of the smallest posterior credible region which includes $\vartheta_{13}=0$
557: is 0.86,
558: considering the rescaled probability corresponding to a two-tailed posterior Gaussian distribution
559: we obtain a hint of $\vartheta_{13}>0$ at the $1.2\sigma$ level,
560: as in the $\chi^2$ analysis
561: (see the discussion at the end of Section~\ref{chi2 analysis}).
562:
563: \begin{figure}[t!]
564: \includegraphics*[scale=0.27]{comp1.eps}
565: \hfill
566: \includegraphics*[scale=0.27]{comp2.eps}
567: \hfill
568: \includegraphics*[scale=0.27]{comp3.eps}
569: \caption{
570: \label{f1bc} Comparison of the allowed regions in the
571: $\sin^{2}\vartheta_{13}$--$\Delta{m}^{2}_{21}$ plane obtained with
572: the $\chi^{2}$ analysis and the Bayesian approach. The light-shadowed and
573: light+dark-shadowed areas cover, respectively, the Bayesian credible regions
574: with $\xi$ probability and the $\chi^{2}$ region with $\xi$ C.L.. In
575: the three figures, from left to right, $ \xi = 90\%, 95\%, 99.73\%
576: $. }
577: \end{figure}
578:
579: \begin{figure}[t!]
580: \includegraphics*[scale=0.27]{comp4.eps}
581: \hfill
582: \includegraphics*[scale=0.27]{comp5.eps}
583: \hfill
584: \includegraphics*[scale=0.27]{comp6.eps}
585: \caption{ \label{f2bc} Comparison of the allowed regions in the
586: $\sin^{2}\vartheta_{12}$--$\sin^{2}\vartheta_{13}$ plane
587: obtained with the $\chi^{2}$ analysis and the Bayesian approach. The
588: light-shadowed and
589: light+dark-shadowed areas cover, respectively, the Bayesian credible
590: regions with $\xi$ probability and the $\chi^{2}$ region with $\xi$
591: C.L.. In the three figures, from left to right, $ \xi = 90\%, 95\%,
592: 99.73\% $. }
593: \end{figure}
594:
595: As a caveat on the comparison of frequentist and Bayesian results,
596: let us remind that the two theories are based on different definitions of probability.
597: Hence,
598: although
599: "numerical results tend to be the same for the two approaches in the asymptotic regime,
600: that is, when there are a lot of data, and statistical uncertainties are small
601: compared with the distance to the nearest physical boundary" \cite{James:2006zz},
602: the interpretation is different.
603:
604: In the frequentist theory the probability of a class of random events is the
605: relative frequency of occurrence of these events
606: when the total number of events tends to infinity.
607: In parameter estimation, a confidence interval with $\alpha$ C.L.
608: is an element of a hypothetical set of confidence intervals which have a frequentist probability
609: $\alpha$ of covering the true value of the parameter
610: (see Refs.~\cite{James:2006zz,PDG-2008}).
611: Notice that in frequentist statistics it is not allowed to make any statement about the true value of the parameter,
612: which is a fixed unknown number, not a random variable,
613: albeit in practice frequentist statistics is very often applied to quantities which are not random variables,
614: as systematic errors.
615: The correct frequentist statements in parameter estimation concern intervals in the parameter space
616: and the frequency of their coverage of the unknown true value in the asymptotic limit.
617: This is the meaning of the allowed regions in Figs.~\ref{f1chi} and \ref{f2chi}.
618: The $1.2\sigma$ hint of $\vartheta_{13}>0$ discussed at the end of Section~\ref{chi2 analysis} means that
619: the best-fit value of $\sin^2\vartheta_{13}$ is $1.2\sigma$ away from $\sin^2\vartheta_{13}=0$,
620: i.e. the confidence intervals obtained in the $\chi^2$ analysis with less than about 76\% C.L. do not include $\sin^2\vartheta_{13}=0$.
621:
622: In the Bayesian theory probability represents the degree of belief based on the available knowledge.
623: Hence it is possible to estimate a probability for any kind of event,
624: not only for random variables as in frequentist statistics.
625: In particular, systematic errors can be treated without any inconsistency.
626: Moreover, there is no need to consider hypothetical quantities, since the posterior
627: probability distribution is straightforwardly obtained from the prior probability distribution
628: and the sampling probability distribution using Bayes' theorem, as in Eq.~(\ref{B1}).
629: The only difficult task in Bayesian theory probability is the estimation of
630: the prior probability distribution on the basis of the available knowledge.
631: In parameter estimation, one can calculate the Bayesian probability
632: of the true value of the parameter to lie in an interval by integrating the posterior
633: probability distribution.
634: The $1.2\sigma$ hint of $\vartheta_{13}>0$ discussed above means that
635: the credible intervals obtained with a Gaussian approximation of the posterior probability density
636: having less than about 76\% probability do not include $\sin^2\vartheta_{13}=0$.
637: Note that the Gaussian approximation of the posterior probability density is defined on the whole real axis of $\sin^2\vartheta_{13}$
638: for the comparison with the analogous frequentist result
639: using the traditional terminology.
640: In fact, the least-squares analysis leads to correct frequentist confidence intervals
641: only in the case of a Gaussian likelihood in which the mean values of the data points are linear functions of the parameters.
642: In practice this requirement is approximately satisfied in a region around the minimum of the $\chi^2$ if the data are
643: abundant and the minimum of the $\chi^2$ lies far from any boundary of the parameters.
644: Since in the case under consideration we are close to the boundary $\sin^2\vartheta_{13}\geq0$,
645: we can compare the Bayesian result with the frequentist least-squares result in which the boundary has not been taken into account
646: only by relaxing the boundary restriction.
647:
648: \begin{figure}[t!]
649: \begin{center}
650: \includegraphics*[scale=0.6]{fig9.eps}
651: \end{center}
652: \caption{ \label{f03} Marginal posterior probability distribution
653: of $\sin^{2}\vartheta_{13}$ obtained with an uninformative
654: constant prior probability distribution. The straight vertical lines
655: show the levels of 90\%, 95\%, and 99.73\% integrated probability. }
656: \end{figure}
657:
658: \section{An Informative Prior}
659: \label{Informative}
660:
661: In the previous Section we analyzed the solar and KamLAND neutrino
662: data assuming a constant uninformative prior probability
663: distribution in the three-dimensional space of the parameters
664: $\Delta{m}^{2}_{21}$, $\sin^{2}\vartheta_{12}$, and
665: $\sin^{2}\vartheta_{13}$. However, as remarked in the introductory
666: Section~\ref{Introduction}, the value of $\vartheta_{13}$ was known
667: to be small
668: before the analysis of solar and KamLAND neutrino data from
669: the negative results of the Chooz \cite{hep-ex/0301017} and Palo
670: Verde \cite{hep-ex/0107009} long-baseline neutrino oscillation
671: experiments combined with the evidence of neutrino oscillations in
672: atmospheric and long-baseline accelerator neutrino experiments.
673: In the Bayesian approach it is natural to try to express this prior knowledge
674: through a prior probability distribution. The resulting posterior
675: probability distribution of the mixing parameters
676: $\Delta{m}^{2}_{21}$, $\sin^{2}\vartheta_{12}$, and
677: $\sin^{2}\vartheta_{13}$ is interpreted as our knowledge about
678: their values obtained from solar and KamLAND neutrino data, taking
679: into account the information on $\vartheta_{13}$ obtained in
680: atmospheric and long-baseline accelerator and reactor neutrino
681: experiments.
682:
683: Since we do not have the machinery for the fit of the data of
684: atmospheric and long-baseline accelerator and reactor neutrino experiments,
685: we constructed a prior probability distribution for $\vartheta_{13}$
686: using the $\chi^{2}$ reported in Fig.~24 of Ref.~\cite{hep-ph/0506083},
687: where such fit was performed.
688: In Fig.~24 of Ref.~\cite{hep-ph/0506083}
689: there are two slightly different curves
690: corresponding to the normal and inverted schemes (see Fig.~\ref{m008}),
691: which depict $\chi^{2}(\cos\delta\sin\vartheta_{13})$
692: for the two CP-conserving cases $\cos\delta=\pm1$,
693: where $\delta$ is the phase in the mixing matrix in Eq.~(\ref{f035}).
694: Since we do not have any information on the value of $\delta$,
695: for each scheme
696: we considered a prior probability distribution for $\vartheta_{13}$
697: marginalized over $\cos\delta=\pm1$:
698: \begin{equation}
699: p(\vartheta_{13}|\mathcal{I})
700: \propto
701: \sum_{\cos\delta=\pm1}
702: \exp\left( - \frac{\chi^{2}(\cos\delta\sin\vartheta_{13}) }{ 2 } \right)
703: \,.
704: \label{t05}
705: \end{equation}
706: For $\sin^{2}\vartheta_{12}$ and
707: $\Delta{m}^{2}_{21}$ we assumed constant uninformative priors as in
708: Section~\ref{Bayesian Analysis}.
709:
710: The prior distributions (\ref{t05}) in the normal and inverted schemes
711: are depicted by the dotted curves in Fig.~\ref{f04p},
712: from which one can see that they have a maximum for
713: $\sin^{2}\vartheta_{13}=0$.
714: Hence, they disfavor the hint of $\vartheta_{13}>0$.
715: The 90\%, 95\% and 99.73\% prior upper bounds for $\sin^{2}\vartheta_{13}$ are
716: \begin{equation}
717: \sin^{2}\vartheta_{13} <
718: \left\{
719: %\setlength{\arraycolsep}{2pt}
720: \begin{array}{lcl} \displaystyle
721: 0.030 \,,\quad 0.036 \,,\quad 0.051 & \qquad & \text{(normal
722: scheme)} \,,
723: \\ \displaystyle
724: 0.033 \,,\quad 0.039 \,,\quad 0.057 & \qquad & \text{(inverted
725: scheme)} \,.
726: \end{array}
727: \right.
728: \label{priorbounds}
729: \end{equation}
730: Notice that such disfavoring of the hint of $\vartheta_{13}>0$ obtained from
731: the $\chi^{2}$ in Fig.~24 of Ref.~\cite{hep-ph/0506083} in the Bayesian approach
732: is in contrast with a faint hint of $\vartheta_{13}>0$ which can be obtained
733: in the frequentist approach by considering the minimum of $\chi^{2}$ at
734: $ \cos\delta\sin\vartheta_{13} \simeq - 0.1 $
735: and $ \Delta\chi^{2} \simeq 0.2 $ at $ \vartheta_{13} = 0 $
736: (see the discussion in Ref.~\cite{hep-ph/0506083}).
737: The contrast is due to the different marginalization procedures in the frequentist and Bayesian theories:
738: in the frequentist theory only the minimum of $\chi^{2}$ with respect to the marginalized parameters is considered,
739: whereas in the Bayesian theory marginalization is implemented by integrating over the distribution of the marginalized parameters,
740: as we have done, for example, in Eq.~(\ref{t02}).
741: In the case of the marginalization over $\cos\delta=\pm1$,
742: the Bayesian procedure of summing the prior probability distribution over $\cos\delta=\pm1$ in Eq.~(\ref{t05}) for each value of $ \sin\vartheta_{13} $
743: is different from the frequentist consideration of $\chi^{2}(\sin\vartheta_{13})$
744: for $\cos\delta=-1$ only,
745: which is due to
746: $\chi^{2}(\sin\vartheta_{13}|\cos\delta=-1)<\chi^{2}(\sin\vartheta_{13}|\cos\delta=1)$.
747: In general,
748: the Bayesian marginalization procedure has the merit to take into account all the distribution of the marginalized parameters,
749: which gives more information than the single point of minimum of $\chi^2$.
750:
751: \begin{figure}[t!]
752: \includegraphics*[scale=0.44]{prior1.eps}
753: \hfill
754: \includegraphics*[scale=0.44]{prior2.eps}
755: \caption{ \label{f01p}
756: Shadowed areas: 90\%, 95\%, and 99.73\% Bayesian credible
757: regions in the $\sin^{2}\vartheta_{13}$--$\Delta{m}^{2}_{21}$
758: plane obtained with the informative prior probability distribution
759: in Eq.~(\ref{t05}).
760: The regions enclosed by solid lines correspond to those in Fig.~\ref{f01},
761: obtained with an uninformative prior.
762: The straight vertical dotted lines enclose,
763: respectively, the 90\%, 95\%, and 99.73\% prior credible regions of
764: $\sin^{2}\vartheta_{13}$. The left and right plots correspond,
765: respectively, to a normal and an inverted scheme (see
766: Fig.~\ref{m008}).
767: The cross and asterisk indicate, respectively, the best-fit points of the analyses with uninformative and informative priors.}
768: \label{prior1}
769: \end{figure}
770:
771: Using the informative prior on $\vartheta_{13}$ in Eq.~(\ref{t05}),
772: from the analysis of solar and KamLAND data
773: We found the best-fit point, corresponding to the maximum of the posterior probability distribution,
774: \begin{equation}
775: \Delta{m}^{2}_{21} = 7.58\times 10^{-5} \, \text{eV}^{2} \,, \quad
776: \sin^{2}\vartheta_{12} = 0.31 \,, \quad \sin^{2}\vartheta_{13} =
777: 0.012\,. \label{bestfit}
778: \end{equation}
779: The shadowed areas in Figs.~\ref{f01p} and \ref{f02p}
780: show the posterior credible regions with 90\%, 95\%, and 99.73\%
781: probability in the $\sin^{2}\vartheta_{13}$--$\Delta{m}^{2}_{21}$
782: and $\sin^{2}\vartheta_{12}$--$\sin^{2}\vartheta_{13}$ planes,
783: respectively.
784: The boundaries of the corresponding regions obtained with an uninformative prior,
785: shown in Figs.~\ref{f01} and \ref{f02},
786: are depicted with solid lines.
787:
788: Since the prior information constrains only
789: $\vartheta_{13}$, the best-fit values and allowed ranges of $\Delta m^2$ and $\vartheta_{12}$
790: are similar to those obtained in Section~\ref{Bayesian Analysis} with an uninformative prior.
791: A small change is due to the correlation with $\vartheta_{13}$.
792:
793: On the other hand,
794: one can see that the assumption of the informative prior in Eq.~(\ref{t05})
795: leads to a significant reduction of the allowed range of $\vartheta_{13}$
796: with respect to that obtained with an uninformative prior,
797: as should have been expected.
798:
799: A curious feature of Figs.~\ref{f01p} and \ref{f02p}
800: is that the 90\% and 95\% allowed ranges of $\sin^{2}\vartheta_{13}$
801: seem to be larger than those allowed by the prior distribution
802: (vertical straight lines in Fig.~\ref{f01p} and
803: horizontal straight lines in Fig.~\ref{f02p}).
804: Such a conclusion would be erroneous,
805: because the prior distribution in Eq.~(\ref{t05}) concerns only one parameter,
806: whereas the credible regions in Figs.~\ref{f01p} and \ref{f02p}
807: constrain two parameters
808: taking into account their correlation.
809:
810: The posterior probability
811: distribution of $\sin^{2}\vartheta_{13}$
812: obtained from the
813: marginalization in Eq.~(\ref{t02}) implies an
814: allowed range of $\sin^{2}\vartheta_{13}$
815: which is smaller than that given by the prior distribution,
816: as one can see from Fig.~\ref{f04p}.
817: We obtained the 90\%, 95\% and 99.73\% upper bounds
818: \begin{equation}
819: \sin^{2}\vartheta_{13} <
820: \left\{
821: %\setlength{\arraycolsep}{2pt}
822: \begin{array}{lcl} \displaystyle
823: 0.027 \,,\quad 0.030 \,,\quad 0.045 & \qquad & \text{(normal
824: scheme)} \,,
825: \\ \displaystyle
826: 0.030 \,,\quad 0.033 \,,\quad 0.048
827: &
828: \qquad
829: &
830: \text{(inverted scheme)}
831: \,,
832: \end{array}
833: \right.
834: \label{postbounds}
835: \end{equation}
836: which are smaller than the corresponding ones in Eq.~(\ref{priorbounds}).
837: These bounds are also about 60\% smaller than those obtained in Eq.~(\ref{t03})
838: with an uninformative prior.
839: Figure~\ref{f03p} shows the comparison of the posterior probability with the one in Fig.~\ref{f03},
840: which has been obtained with an uninformative flat prior.
841:
842: \begin{figure}[t!]
843: \includegraphics*[scale=0.44]{prior3.eps}
844: \hfill
845: \includegraphics*[scale=0.44]{prior4.eps}
846: \caption{ \label{f02p}
847: Shadowed areas: 90\%, 95\%, and 99.73\% Bayesian credible
848: regions in the
849: $\sin^{2}\vartheta_{12}$--$\sin^{2}\vartheta_{13}$ plane
850: obtained with the informative prior probability distribution in
851: Eq.~(\ref{t05}).
852: The regions enclosed by solid lines correspond to those in Fig.~\ref{f02},
853: obtained with an uninformative prior.
854: The straight horizontal dotted lines enclose,
855: respectively, the 90\%, 95\%, and 99.73\% prior credible regions of
856: $\sin^{2}\vartheta_{13}$. The left and right plots correspond,
857: respectively, to a normal and an inverted scheme (see
858: Fig.~\ref{m008}).
859: The cross and asterisk indicate, respectively, the best-fit points of the analyses with uninformative and informative priors.}
860: \label{prior2}
861: \end{figure}
862:
863: It is interesting to note that the bounds on
864: $\sin^{2}\vartheta_{13}$ in Eq.~(\ref{postbounds}) are similar to
865: those obtained with a global $\chi^{2}$ analysis of neutrino oscillation data
866: in Ref.~\cite{hep-ph/0808.2016}
867: (see, however, the caveat on the comparison of frequentist and Bayesian results
868: discussed at the end of Section~\ref{Bayesian Analysis}).
869: Our results also agree with the weakening of the hint of $\vartheta_{13}>0$
870: discussed in Ref.~\cite{hep-ph/0808.2016} coming from the addition of
871: atmospheric and long-baseline accelerator and reactor neutrino data
872: to the analysis of solar and KamLAND data:
873: using the method described in Section~\ref{Bayesian Analysis},
874: the significance of the
875: hint of $\vartheta_{13}>0$ is reduced from about $1.2\sigma$
876: to about $0.8\sigma$
877: (with 0.72 and 0.75 respective probabilities of the smallest posterior credible region which includes $\vartheta_{13}=0$
878: in the normal and inverted schemes).
879: The discrepancy with the $1.6\sigma$ reported in Ref.~\cite{hep-ph/0806.2649}
880: is probably due to the marginalization over $\cos\delta=\pm1$ in Eq.~(\ref{t05}).
881: In fact, for $\cos\delta=-1$
882: Fig.~24 of
883: Ref.~\cite{hep-ph/0506083}
884: implies a prior in favor of $\vartheta_{13}>0$,
885: which leads to a global hint of $\vartheta_{13}>0$ at the $1.5\sigma$ level
886: (with 0.93 probability of the smallest posterior credible region which includes $\vartheta_{13}=0$),
887: in agreement with the value in Ref.~\cite{hep-ph/0806.2649}
888: ($1.6\sigma$).
889: Let us however emphasize that,
890: since the marginalization over the unknown value of $\delta$ is the correct procedure in the Bayesian approach,
891: our result for the statistical significance of the global hint of $\vartheta_{13}>0$ is $0.8\sigma$.
892:
893: \begin{figure}[t!]
894: \includegraphics*[scale=0.44]{prior7.eps}
895: \hfill
896: \includegraphics*[scale=0.44]{prior8.eps}
897: \caption{ \label{f04p} Marginal posterior probability distribution
898: of $\sin^{2}\vartheta_{13}$ obtained with the informative prior
899: probability distribution in Eq.~(\ref{t05}). The dotted curve shows
900: the prior distribution in Eq.~(\ref{t05}).
901: The long (short) straight vertical lines show the 90\%, 95\%, and 99.73\% posterior (prior) probability levels.
902: The short straight dotted vertical lines have been slightly shifted to the right to avoid superposition
903: (compare with Eqs.~(\ref{priorbounds}) and (\ref{postbounds})).
904: }
905: \end{figure}
906:
907: Let us finally remark that the results presented in this Section
908: depend on the choice of the prior probability distribution for $\vartheta_{13}$
909: obtained from the fit of the data of
910: atmospheric and long-baseline accelerator and reactor neutrino experiments.
911: In Eq.~(\ref{t05}), instead of the $\chi^{2}$ of
912: Ref.~\cite{hep-ph/0506083} we could have used,
913: for example,
914: the $\chi^{2}$ of one of
915: Refs.~\cite{hep-ph/0808.2016,hep-ph/0704.1800,nucl-th/0805.2924,hep-ph/0804.4857}.
916: However,
917: since in these papers the same data have been fitted with similar assumptions and methods,
918: using one of these $\chi^{2}$'s would not change dramatically the numerical results presented above.
919: For example, we considered the $\chi^{2}$ in Fig.~3 of Ref.~\cite{hep-ph/0808.2016},
920: which corresponds to the prior upper bounds
921: \begin{equation}
922: \sin^{2}\vartheta_{13}
923: <
924: 0.030 \, (90\%)
925: \,,
926: \quad
927: 0.039 \, (95\%)
928: \,,
929: \quad
930: 0.063 \, (99.73\%)
931: \,.
932: \label{t03a}
933: \end{equation}
934: We obtained the best-fit values
935: \begin{equation}
936: \Delta{m}^{2}_{21} = 7.58 \times 10^{-5} \, \text{eV}^{2}
937: \,,
938: \quad
939: \sin^{2}\vartheta_{12} = 0.31
940: \,,
941: \quad
942: \sin^{2}\vartheta_{13} = 0.008
943: \,,
944: \label{bestfit1}
945: \end{equation}
946: and the posterior upper limits
947: \begin{equation}
948: \sin^{2}\vartheta_{13}
949: <
950: 0.030 \, (90\%)
951: \,,
952: \quad
953: 0.033 \, (95\%)
954: \,,
955: \quad
956: 0.051 \, (99.73\%)
957: \,.
958: \label{t03b}
959: \end{equation}
960: One can see that these values are close to the corresponding ones in Eqs.~(\ref{priorbounds})--(\ref{postbounds}).
961: For the hint of $\vartheta_{13}>0$ we have a $0.9\sigma$ statistical significance
962: (with 0.78 probability of the smallest posterior credible region which includes $\vartheta_{13}=0$),
963: in perfect agreement with Ref.~\cite{hep-ph/0808.2016}.
964:
965: \begin{figure}[t!]
966: \includegraphics*[scale=0.44]{prior5.eps}
967: \hfill
968: \includegraphics*[scale=0.44]{prior6.eps}
969: \caption{ \label{f03p} Marginal posterior probability distribution
970: of $\sin^{2}\vartheta_{13}$ obtained with the informative prior
971: probability distribution in Eq.~(\ref{t05}). The dotted curve shows
972: the marginal posterior distribution in Fig.~\ref{f03}, obtained with
973: an uninformative flat prior. The long straight vertical lines show
974: the 90\%, 95\%, and 99.73\% posterior probability levels in
975: Eq.~(\ref{postbounds}). The short straight vertical lines show the
976: corresponding probability levels in Eq.~(\ref{t03}), obtained with
977: an uninformative flat prior.
978: The short straight dotted vertical line in the figure on the right has been slightly shifted to the right to avoid superposition
979: (compare with Eqs.~(\ref{t03}) and (\ref{postbounds})).
980: }
981: \end{figure}
982:
983: \section{Conclusions}
984: \label{Conclusions}
985:
986: In this paper we presented the results of a Bayesian
987: analysis of the solar and KamLAND neutrino data
988: with the aim of determining the value of the unknown mixing angle
989: $\vartheta_{13}$ in the framework of three-neutrino mixing.
990:
991: We found that with an uninformative flat prior distribution in the
992: relevant mixing parameters $\Delta m^2_{12}$,
993: $\sin^2\vartheta_{12}$, $\sin^2\vartheta_{13}$, the Bayesian
994: credible regions in the $\sin^2\vartheta_{13}-\Delta m^2_{12}$ and
995: $\sin^2\vartheta_{12}-\sin^2\vartheta_{13}$ planes are only slightly
996: smaller than the allowed regions obtained with a traditional
997: least-squares analysis, implying a rather stringent upper bound for
998: $\sin^2\vartheta_{13}$. Our analysis confirms the $1.2\sigma$ hint
999: of $\vartheta_{13}>0$ found in Ref.~\cite{hep-ph/0806.2649}.
1000:
1001: We also performed an analysis with an informative prior which
1002: represents information on $\vartheta_{13}$ obtained in atmospheric
1003: and long-baseline accelerator and reactor neutrino experiments,
1004: independently from solar and KamLAND neutrino data. We found that
1005: such a prior implies a significant decrease of the upper bound on
1006: $\vartheta_{13}$ with respect to that obtained with an uninformative
1007: prior and the hint of $\vartheta_{13}>0$ is reduced to a
1008: $0.8\sigma$ level. Our results are similar to those obtained with a
1009: global $\chi^{2}$ analysis of neutrino oscillation data in
1010: Ref.~\cite{hep-ph/0808.2016}
1011: (see, however, the caveat on the comparison of frequentist and Bayesian results
1012: discussed at the end of Section~\ref{Bayesian Analysis}).
1013:
1014: Let us finally emphasize that Bayesian inference
1015: (see
1016: Refs.~\cite{Jeffreys-book-39,Loredo-90,Loredo-92,Jaynes-book-2003,DAgostini-book-95})
1017: is founded on a consistent theory and can always be implemented
1018: in a correct way (given enough computational power).
1019: On the other hand,
1020: the frequentist method is based on an unphysical definition of probability
1021: and in most cases of interest cannot be implemented in a correct way.
1022: In particular,
1023: a dramatic flaw of the frequentist method is that
1024: the frequentist definition of probability does not allow the treatment of
1025: theoretical and systematic errors as random variables.
1026: Hence, the aim of the frequentist statistics approach
1027: of extracting objective statistical information from data cannot be realized in practice.
1028: Since the Bayesian theory does not suffer from such shortcomings,
1029: we think that it is preferable for attaining reliable results
1030: from the analysis of experimental data.
1031:
1032: \appendix
1033:
1034: \section{Regeneration of Solar $\nu_{\lowercase{e}}$'s in the Earth}
1035: \label{Regeneration}
1036:
1037: Solar neutrinos arriving at a detector during night-time
1038: pass through the Earth, where the matter effect
1039: (also called ``MSW effect'' \cite{Wolfenstein:1978ue,Mikheev:1986wj})
1040: can cause a change in the
1041: flavor composition,
1042: which is called ``regeneration of solar $\nu_{\lowercase{e}}$'s in the Earth''.
1043: In this Appendix, we derive the connection between the averaged probability of
1044: survival of solar electron neutrinos passing through the Earth,
1045: $\overline{P}_{\nu_{e}\to\nu_{e}}^{\text{Sun}+\text{Earth}}$,
1046: which is measured during night-time,
1047: the averaged probability of $\nu_{e}$ survival from the core of the Sun to the surface of the Earth,
1048: $\overline{P}_{\nu_{e}\to\nu_{e}}^{\text{Sun}}$,
1049: which is measured during day-time,
1050: and
1051: the probability of
1052: $\nu_{2}\to\nu_{e}$ transitions in the Earth,
1053: $P_{\nu_{2}\to\nu_{e}}^{\text{Earth}}$.
1054: We also discuss the connection between $P_{\nu_{2}\to\nu_{e}}^{\text{Earth}}$
1055: in the case of three-neutrino mixing
1056: and the probability of
1057: $\nu_{2}\to\nu_{e}$ transitions in the Earth
1058: in the case of two-neutrino mixing.
1059:
1060: The mixing of neutrino states is given by
1061: \begin{equation}
1062: | \nu_{\alpha} \rangle
1063: =
1064: \sum_{k=1}^{3} U_{\alpha k}^{*} \, | \nu_{k} \rangle
1065: \qquad
1066: (\alpha=e,\mu,\tau)
1067: \,,
1068: \label{e003}
1069: \end{equation}
1070: where $U$ is the $3\times3$ unitary mixing matrix of the neutrino fields
1071: (see
1072: Refs.~\cite{hep-ph/9812360,hep-ph/0310238,Giunti-Kim}).
1073: We adopt the standard parameterization \cite{Chau:1984fp,PDG-2006}
1074: \begin{equation}
1075: U
1076: =
1077: \begin{pmatrix}
1078: c_{12}
1079: c_{13}
1080: &
1081: s_{12}
1082: c_{13}
1083: &
1084: s_{13}
1085: e^{-i\delta}
1086: \\
1087: -
1088: s_{12}
1089: c_{23}
1090: -
1091: c_{12}
1092: s_{23}
1093: s_{13}
1094: e^{i\delta}
1095: &
1096: c_{12}
1097: c_{23}
1098: -
1099: s_{12}
1100: s_{23}
1101: s_{13}
1102: e^{i\delta}
1103: &
1104: s_{23}
1105: c_{13}
1106: \\
1107: s_{12}
1108: s_{23}
1109: -
1110: c_{12}
1111: c_{23}
1112: s_{13}
1113: e^{i\delta}
1114: &
1115: -
1116: c_{12}
1117: s_{23}
1118: -
1119: s_{12}
1120: c_{23}
1121: s_{13}
1122: e^{i\delta}
1123: &
1124: c_{23}
1125: c_{13}
1126: \end{pmatrix}
1127: \,,
1128: \label{f035}
1129: \end{equation}
1130: where
1131: $ c_{ab} \equiv \cos\vartheta_{ab} $
1132: and
1133: $ s_{ab} \equiv \sin\vartheta_{ab} $.
1134: The three mixing angles
1135: $\vartheta_{12}$,
1136: $\vartheta_{13}$,
1137: $\vartheta_{23}$
1138: take values in the ranges
1139: $ 0 \leq \vartheta_{ab} \leq \pi/2 $.
1140: The CP-violating phase
1141: $\delta$
1142: is confined in the interval
1143: $ 0 \leq \delta < 2 \pi $.
1144: We neglected possible Majorana phases,
1145: which are irrelevant for neutrino oscillations
1146: \cite{Bilenky:1980cx,Doi:1980yb,Langacker:1986jv}.
1147:
1148: A solar neutrino, created in the core of the Sun as a $\nu_{e}$,
1149: arrives in a detector as a superposition of
1150: $\nu_{1}$, $\nu_{2}$, and $\nu_{3}$.
1151: However,
1152: since the neutrino squared-mass differences are relatively large
1153: (see Eqs.~(\ref{SOL})--(\ref{104})),
1154: the average of the oscillation probability
1155: over the energy resolution of the detector washes out the
1156: interference terms between the massive neutrinos \cite{hep-ph/9903329}.
1157: This is due to the fact that the vacuum oscillation lengths
1158: are much shorter than the Sun--Earth distance:
1159: \begin{equation}
1160: L^{\text{osc}}_{21}
1161: =
1162: \frac{ 4 \pi E }{ \Delta{m}^2_{21} }
1163: \simeq
1164: 30 \, \text{km} \left( \frac{E}{\text{MeV}} \right)
1165: \,,
1166: \qquad
1167: L^{\text{osc}}_{32} \simeq L^{\text{osc}}_{31}
1168: =
1169: \frac{ 4 \pi E }{ |\Delta{m}^2_{31}| }
1170: \simeq
1171: 1 \, \text{km} \left( \frac{E}{\text{MeV}} \right)
1172: \,,
1173: \label{e101}
1174: \end{equation}
1175: where $E$ is the neutrino energy,
1176: which in solar neutrino experiments varies in the interval
1177: \begin{equation}
1178: 0.2 \, \text{MeV}
1179: \lesssim
1180: E
1181: \lesssim
1182: 15 \, \text{MeV}
1183: \,.
1184: \label{e102}
1185: \end{equation}
1186: Then,
1187: the measurable averaged survival probability of solar electron neutrinos after crossing the Earth is given by
1188: \begin{equation}
1189: \overline{P}_{\nu_{e}\to\nu_{e}}^{\text{Sun}+\text{Earth}}
1190: =
1191: \sum_{k=1}^{3}
1192: P_{\nu_{e}\to\nu_{k}}^{\text{Sun}} \, P_{\nu_{k}\to\nu_{e}}^{\text{Earth}}
1193: \,,
1194: \label{j128}
1195: \end{equation}
1196: where
1197: $P_{\nu_{e}\to\nu_{k}}^{\text{Sun}}$
1198: is the probability of $\nu_{e}\to\nu_{k}$ transitions from the solar core to the surface of the Earth
1199: and
1200: $P_{\nu_{k}\to\nu_{e}}^{\text{Earth}}$
1201: is the probability of $\nu_{k}\to\nu_{e}$ transitions in the passage through the Earth.
1202:
1203: In matter,
1204: electron neutrinos feel a charged-current potential
1205: $ V_{\text{CC}} = \sqrt{2} G_{\text{F}} N_{e} $,
1206: where $G_{\text{F}}$ is the Fermi constant and $N_{e}$ is the electron number density.
1207: The quantity which gives the matter effect in the evolution equation of
1208: neutrino flavors is
1209: \begin{equation}
1210: A_{\text{CC}} = 2 E V_{\text{CC}}
1211: =
1212: 1.53 \times 10^{-7} \, \text{eV}^2
1213: \left(
1214: \dfrac{ N_{e} }{ N_{\text{A}} \, \text{cm}^{-3} }
1215: \right)
1216: \left(
1217: \dfrac{ E }{ \text{MeV} }
1218: \right)
1219: \,,
1220: \label{e004}
1221: \end{equation}
1222: where
1223: $N_{\text{A}}$ is the Avogadro number.
1224: The electron number density in the solar core is about
1225: $ 100 \, N_{\text{A}} \, \text{cm}^{-3} $.
1226: In the Earth, the electron number density varies from about
1227: $ 2.2 \, N_{\text{A}} \, \text{cm}^{-3} $ in the mantle
1228: to about
1229: $ 5.5 \, N_{\text{A}} \, \text{cm}^{-3} $ in the core.
1230: Thus,
1231: for solar neutrinos we have
1232: $ A_{\text{CC}} \lesssim 2.3 \times 10^{-4} \, \text{eV}^2 $,
1233: which is much smaller than the atmospheric
1234: squared-mass difference $\Delta{m}^{2}_{\text{ATM}}$
1235: (see Eq.~(\ref{ATM})).
1236: This means that the matter effect cannot induce transitions between
1237: $\nu_{3}$ and the two neutrinos $\nu_{1}$ and $\nu_{2}$,
1238: since the two groups are separated by the large atmospheric
1239: squared-mass difference $\Delta{m}^{2}_{\text{ATM}}$
1240: (see Eq.~(\ref{104}) and Fig.~\ref{m008}).
1241: In other words,
1242: the massive neutrino component $\nu_{3}$ propagates without disturbance
1243: from the core of the Sun to the detector
1244: and the corresponding transition probabilities in Eq.~(\ref{j128})
1245: are simply given by
1246: \begin{equation}
1247: P_{\nu_{e}\to\nu_{3}}^{\text{Sun}}
1248: =
1249: P_{\nu_{3}\to\nu_{e}}^{\text{Earth}}
1250: =
1251: | \langle \nu_{3} | \nu_{e} \rangle |^2
1252: =
1253: | U_{e3} |^2
1254: \,.
1255: \label{e005}
1256: \end{equation}
1257: Furthermore,
1258: taking into account the conservation of probability,
1259: we have
1260: \begin{align}
1261: \null & \null
1262: P_{\nu_{e}\to\nu_{1}}^{\text{Sun}}
1263: =
1264: 1
1265: -
1266: P_{\nu_{e}\to\nu_{3}}^{\text{Sun}}
1267: -
1268: P_{\nu_{e}\to\nu_{2}}^{\text{Sun}}
1269: =
1270: 1
1271: -
1272: | U_{e3} |^2
1273: -
1274: P_{\nu_{e}\to\nu_{2}}^{\text{Sun}}
1275: \,,
1276: \label{e006}
1277: \\
1278: \null & \null
1279: P_{\nu_{1}\to\nu_{e}}^{\text{Earth}}
1280: =
1281: 1
1282: -
1283: P_{\nu_{3}\to\nu_{e}}^{\text{Earth}}
1284: -
1285: P_{\nu_{2}\to\nu_{e}}^{\text{Earth}}
1286: =
1287: 1
1288: -
1289: | U_{e3} |^2
1290: -
1291: P_{\nu_{2}\to\nu_{e}}^{\text{Earth}}
1292: \,.
1293: \label{e007}
1294: \end{align}
1295:
1296: Let us now express
1297: the averaged survival probability $\overline{P}_{\nu_{e}\to\nu_{e}}^{\text{Sun}}$ of electron neutrinos
1298: from the solar core to the surface of the Earth
1299: in terms of the transition probabilities $P_{\nu_{e}\to\nu_{k}}^{\text{Sun}}$:
1300: \begin{equation}
1301: \overline{P}_{\nu_{e}\to\nu_{e}}^{\text{Sun}}
1302: =
1303: \overline{ | \langle \nu_{e} | \mathcal{S} | \nu_{e} \rangle |^2 }
1304: =
1305: \overline{
1306: \left|
1307: \sum_{k=1}^{3} \langle \nu_{e} | \nu_{k} \rangle \langle \nu_{k} | \mathcal{S} | \nu_{e} \rangle
1308: \right|^2
1309: }
1310: =
1311: \sum_{k=1}^{3} |U_{ek}|^2 \, P_{\nu_{e}\to\nu_{k}}^{\text{Sun}}
1312: \,,
1313: \label{e008}
1314: \end{equation}
1315: where $\mathcal{S}$ is the evolution operator.
1316: We neglected the interference terms for the reason discussed above,
1317: before Eq.~(\ref{j128}).
1318: Using Eqs.~(\ref{e005}), (\ref{e006}), and (\ref{e008}),
1319: we can express $P_{\nu_{e}\to\nu_{2}}^{\text{Sun}}$ in terms of
1320: $\overline{P}_{\nu_{e}\to\nu_{e}}^{\text{Sun}}$:
1321: \begin{equation}
1322: P_{\nu_{e}\to\nu_{2}}^{\text{Sun}}
1323: =
1324: \frac{ |U_{e1}|^2 \left( 1 - |U_{e3}|^2 \right) + |U_{e3}|^4 - \overline{P}_{\nu_{e}\to\nu_{e}}^{\text{Sun}} }{ |U_{e1}|^2 - |U_{e2}|^2 }
1325: \,.
1326: \label{e011}
1327: \end{equation}
1328: Finally, using Eqs.~(\ref{e005}), (\ref{e006}), (\ref{e007}), and (\ref{e011}),
1329: we obtain, from Eq.~(\ref{j128}),
1330: \begin{equation}
1331: \overline{P}_{\nu_{e}\to\nu_{e}}^{\text{Sun}+\text{Earth}}
1332: =
1333: \overline{P}_{\nu_{e}\to\nu_{e}}^{\text{Sun}}
1334: +
1335: \frac
1336: {
1337: \left[ \left( 1 - |U_{e3}|^2 \right)^2 - 2 \left( \overline{P}_{\nu_{e}\to\nu_{e}}^{\text{Sun}} - |U_{e3}|^4 \right) \right]
1338: \left[ P_{\nu_{2}\to\nu_{e}}^{\text{Earth}} - |U_{e2}|^2 \right]
1339: }
1340: { |U_{e1}|^2 - |U_{e2}|^2 }
1341: \,.
1342: \label{e001}
1343: \end{equation}
1344: Since in practice $ |U_{e1}|^2 > |U_{e2}|^2 $, because $ \sin^2
1345: \vartheta_{12} < 1 $ (see
1346: Refs.~\cite{hep-ph/0506083,hep-ph/0605195}), and $|U_{e3}|^2$ is
1347: small, there is a regeneration of electron neutrinos in the Earth if
1348: $ P_{\nu_{2}\to\nu_{e}}^{\text{Earth}} > |U_{e2}|^2 $. Note that in
1349: the absence of matter effects, we have $
1350: P_{\nu_{2}\to\nu_{e}}^{\text{Earth}} = | \langle \nu_{2} | \nu_{e}
1351: \rangle |^2 = |U_{e2}|^2 $ and $
1352: \overline{P}_{\nu_{e}\to\nu_{e}}^{\text{Sun}+\text{Earth}} =
1353: \overline{P}_{\nu_{e}\to\nu_{e}}^{\text{Sun}} $.
1354:
1355: Let us now discuss the calculation of
1356: $ P_{\nu_{2}\to\nu_{e}}^{\text{Earth}} $.
1357: The evolution of neutrino flavors in matter is governed by the Schr\"odinger equation
1358: (see
1359: Refs.~\cite{hep-ph/9812360,hep-ph/0310238,Giunti-Kim})
1360: \begin{equation}
1361: i \, \frac{\text{d}}{\text{d}x} \,
1362: \Psi_{\text{F}}
1363: =
1364: \mathbb{H}_{\text{F}}
1365: \,
1366: \Psi_{\text{F}}
1367: \,,
1368: \label{i058}
1369: \end{equation}
1370: with the effective Hamiltonian
1371: \begin{equation}
1372: \mathbb{H}_{\text{F}}
1373: =
1374: \frac{1}{2E}
1375: \left(
1376: U
1377: \,
1378: \Delta\mathbb{M}^{2}
1379: \,
1380: U^{\dagger}
1381: +
1382: \mathbb{A}
1383: \right)
1384: \,,
1385: \label{i059}
1386: \end{equation}
1387: and
1388: \begin{equation}
1389: \Psi_{\text{F}}
1390: \equiv
1391: \begin{pmatrix}
1392: \psi_{e}
1393: \\
1394: \psi_{\mu}
1395: \\
1396: \psi_{\tau}
1397: \end{pmatrix}
1398: \,,
1399: \qquad
1400: \Delta\mathbb{M}^{2}
1401: \equiv
1402: \begin{pmatrix}
1403: 0 & 0 & 0
1404: \\
1405: 0 & \Delta{m}^{2}_{21} & 0
1406: \\
1407: 0 & 0 & \Delta{m}^{2}_{31}
1408: \end{pmatrix}
1409: \,,
1410: \qquad
1411: \mathbb{A}
1412: \equiv
1413: \begin{pmatrix}
1414: A_{\text{CC}} & 0 & 0
1415: \\
1416: 0 & 0 & 0
1417: \\
1418: 0 & 0 & 0
1419: \end{pmatrix}
1420: \,.
1421: \label{i060}
1422: \end{equation}
1423: Here,
1424: $
1425: \psi_{\alpha}
1426: =
1427: \langle \nu_{\alpha} | \nu \rangle
1428: $
1429: is the amplitude of the flavor $\alpha$ in the state $| \nu \rangle$ which describes
1430: a propagating neutrino.
1431: The column matrix $\Psi_{\text{F}}$ of flavor amplitudes
1432: is related to the column matrix
1433: $
1434: \Psi_{\text{M}}
1435: \equiv
1436: ( \psi_{1} , \psi_{2} , \psi_{3} )^{T}
1437: $
1438: of mass amplitudes
1439: ($
1440: \psi_{k}
1441: =
1442: \langle \nu_{k} | \nu \rangle
1443: $)
1444: by
1445: \begin{equation}
1446: \Psi_{\text{F}}
1447: =
1448: U \, \Psi_{\text{M}}
1449: \,.
1450: \label{e191}
1451: \end{equation}
1452: In the calculation of $ P_{\nu_{2}\to\nu_{e}}^{\text{Earth}} $,
1453: the initial mass and flavor amplitudes are
1454: \begin{equation}
1455: \psi_{k}(0) = \delta_{k2}
1456: \,,
1457: \qquad
1458: \psi_{\alpha}(0) = U_{\alpha2}
1459: \,.
1460: \label{e192}
1461: \end{equation}
1462: The probability of $\nu_{2}\to\nu_{e}$ transitions at a distance $x$ from neutrino production
1463: is given by
1464: \begin{equation}
1465: P_{\nu_{2}\to\nu_{e}}(x) = |\psi_{e}(x)|^2
1466: \,.
1467: \label{e193}
1468: \end{equation}
1469:
1470: Taking into account the fact that the mixing matrix
1471: in the parameterization in Eq.~(\ref{f035})
1472: can be written as
1473: \begin{equation}
1474: U = R^{23} W^{13} R^{12}
1475: \,,
1476: \label{e201}
1477: \end{equation}
1478: with
1479: \begin{equation}
1480: R^{12}
1481: =
1482: \begin{pmatrix}
1483: c_{12} & s_{12} & 0
1484: \\
1485: - s_{12} & c_{12} & 0
1486: \\
1487: 0 & 0 & 1
1488: \end{pmatrix}
1489: \,,
1490: \quad
1491: R^{23}
1492: =
1493: \begin{pmatrix}
1494: 1 & 0 & 0
1495: \\
1496: 0 & c_{23} & s_{23}
1497: \\
1498: 0 & - s_{23} & c_{23}
1499: \end{pmatrix}
1500: \,,
1501: \quad
1502: W^{13}
1503: =
1504: \begin{pmatrix}
1505: c_{13} & 0 & s_{13} e^{-i\delta}
1506: \\
1507: 0 & 1 & 0
1508: \\
1509: - s_{13} e^{i\delta} & 0 & c_{13}
1510: \end{pmatrix}
1511: \,,
1512: \label{e204}
1513: \end{equation}
1514: it is convenient to work with the new column matrix of amplitudes
1515: $ \widehat{\Psi} \equiv ( \widehat{\psi}_{1} , \widehat{\psi}_{2} , \widehat{\psi}_{3} )^{T} $
1516: defined by
1517: \begin{equation}
1518: \widehat{\Psi}
1519: =
1520: {W^{13}}^{\dagger}
1521: \,
1522: {R^{23}}^{\dagger}
1523: \,
1524: \Psi_{\text{F}}
1525: =
1526: R^{12}
1527: \,
1528: \Psi_{\text{M}}
1529: \,,
1530: \label{m110}
1531: \end{equation}
1532: which follows the evolution equation
1533: \begin{equation}
1534: i \, \frac{\text{d}}{\text{d}x} \,
1535: \widehat{\Psi}
1536: =
1537: \widehat{\mathbb{H}}
1538: \,
1539: \widehat{\Psi}
1540: \,.
1541: \label{m111}
1542: \end{equation}
1543: Since $R^{23}$ commutes with the matter potential matrix $\mathbb{A}$,
1544: the new effective Hamiltonian $\widehat{\mathbb{H}}$ is given by
1545: \begin{equation}
1546: \widehat{\mathbb{H}}
1547: =
1548: \frac{1}{2E}
1549: \left(
1550: R^{12}
1551: \,
1552: \Delta\mathbb{M}^{2}
1553: \,
1554: {R^{12}}^{\dagger}
1555: +
1556: {W^{13}}^{\dagger}
1557: \,
1558: \mathbb{A}
1559: \,
1560: W^{13}
1561: \right)
1562: \,.
1563: \label{m112}
1564: \end{equation}
1565: Explicitly, we have
1566: \begin{equation}
1567: \widehat{\mathbb{H}}
1568: =
1569: \frac{1}{2E}
1570: \begin{pmatrix}
1571: s_{12}^{2} \Delta{m}^{2}_{21}
1572: +
1573: c_{13}^{2} A_{\text{CC}}
1574: &
1575: c_{12} s_{12} \Delta{m}^{2}_{21}
1576: &
1577: - c_{13} s_{13} e^{-i\delta} A_{\text{CC}}
1578: \\
1579: c_{12} s_{12} \Delta{m}^{2}_{21}
1580: &
1581: c_{12}^{2} \Delta{m}^{2}_{21}
1582: &
1583: 0
1584: \\
1585: - c_{13} s_{13} e^{i\delta} A_{\text{CC}}
1586: &
1587: 0
1588: &
1589: \Delta{m}^{2}_{31}
1590: +
1591: s_{13}^{2} A_{\text{CC}}
1592: \end{pmatrix}
1593: \,.
1594: \label{m113}
1595: \end{equation}
1596: From Eq.~(\ref{m110}), we have
1597: $ \widehat{\psi}_{3} = \psi_{3} $.
1598: Therefore,
1599: $ \widehat{\psi}_{3} $ is the amplitude of $\nu_{3}$.
1600: Since $ \Delta{m}^{2}_{31} \gg A_{\text{CC}} $,
1601: in practice the third eigenvalue of $\widehat{\mathbb{H}}$
1602: is equal to $ \Delta{m}^{2}_{31} / 2 E $
1603: and the matter effect cannot induce transitions between
1604: $\nu_{3}$ and the other two massive neutrinos,
1605: as discussed above.
1606: Furthermore, since
1607: $ \widehat{\psi}_{3}(0) = \psi_{3}(0) = 0 $
1608: (from Eq.~(\ref{e192})),
1609: in practice the contribution of $\nu_{3}$ is negligible
1610: and $ P_{\nu_{2}\to\nu_{e}}^{\text{Earth}} $ can be calculated
1611: by solving the effective two-neutrino evolution equation
1612: \begin{equation}
1613: i \, \frac{\text{d}}{\text{d}x} \,
1614: \widetilde{\Psi}
1615: =
1616: \widetilde{\mathbb{H}}
1617: \,
1618: \widetilde{\Psi}
1619: \,,
1620: \label{e221}
1621: \end{equation}
1622: with
1623: $
1624: \widetilde{\Psi}
1625: \equiv
1626: ( \widetilde{\psi}_{1} , \widetilde{\psi}_{2} )^{T}
1627: =
1628: ( \widehat{\psi}_{1} , \widehat{\psi}_{2} )^{T}
1629: $
1630: and
1631: \begin{equation}
1632: \widetilde{\mathbb{H}}
1633: =
1634: \frac{1}{2E}
1635: \begin{pmatrix}
1636: s_{12}^{2} \Delta{m}^{2}_{21}
1637: +
1638: c_{13}^{2} A_{\text{CC}}
1639: &
1640: c_{12} s_{12} \Delta{m}^{2}_{21}
1641: \\
1642: c_{12} s_{12} \Delta{m}^{2}_{21}
1643: &
1644: c_{12}^{2} \Delta{m}^{2}_{21}
1645: \end{pmatrix}
1646: \,.
1647: \label{e222}
1648: \end{equation}
1649: This effective Hamiltonian coincides with the
1650: effective Hamiltonian in the case of two-neutrino mixing
1651: (see Refs.~\cite{hep-ph/9812360,hep-ph/0310238,Giunti-Kim}),
1652: with the matter contribution $A_{\text{CC}}$ multiplied by the three-neutrino mixing factor
1653: $c_{13}^{2}$.
1654: The initial column matrix of amplitudes
1655: is explicitly given, from Eqs.~(\ref{e192}) and (\ref{m110}), by
1656: \begin{equation}
1657: \widetilde{\Psi}(0)
1658: =
1659: \begin{pmatrix}
1660: c_{12} & s_{12}
1661: \\
1662: - s_{12} & c_{12}
1663: \end{pmatrix}
1664: \begin{pmatrix}
1665: 0
1666: \\
1667: 1
1668: \end{pmatrix}
1669: =
1670: \begin{pmatrix}
1671: s_{12}
1672: \\
1673: c_{12}
1674: \end{pmatrix}
1675: \,.
1676: \label{e223}
1677: \end{equation}
1678: The probability of $\nu_{2}\to\nu_{e}$ transitions at a distance $x$ from neutrino production
1679: is given by
1680: \begin{equation}
1681: P_{\nu_{2}\to\nu_{e}}(x)
1682: =
1683: | [ R^{23} W^{13} \widehat{\Psi}(x) ]_{e} |^2
1684: =
1685: c_{13}^{2} \, |\widetilde\psi_{1}(x)|^2
1686: \,.
1687: \label{e224}
1688: \end{equation}
1689: Therefore,
1690: in practice, the probability of $\nu_{2}\to\nu_{e}$ transitions
1691: in the Earth is given by
1692: \begin{equation}
1693: P_{\nu_{2}\to\nu_{e}}^{\text{Earth}}
1694: =
1695: \left( 1 - |U_{e3}|^2 \right) P_{\nu_{2}\to\nu_{e}}^{\text{Earth};2\nu}
1696: \,,
1697: \label{e002}
1698: \end{equation}
1699: where $P_{\nu_{2}\to\nu_{e}}^{\text{Earth};2\nu}$
1700: is the probability of $\nu_{2}\to\nu_{e}$ transitions
1701: calculated in the case of two-neutrino mixing with an effective matter contribution
1702: multiplied by $ c_{13}^2 = 1 - |U_{e3}|^2 $.
1703:
1704: \section*{Acknowledgments}
1705:
1706: C. Giunti would like to thank the Department of Theoretical Physics of the University of Torino
1707: for hospitality and support.
1708:
1709: %\bibliographystyle{h-elsevier3}
1710: %\input{bibtex/bib.tex}
1711:
1712: \input{t13.bbl}
1713:
1714: \end{document}
1715: