hep-ph0502154/inv.tex
1: \documentclass[a4paper]{JHEP3}
2: 
3: \usepackage[english]{babel}
4: \usepackage{amsmath,amssymb}
5: \usepackage[dvips]{graphicx}
6: \usepackage{ifthen,array}
7: \usepackage[comma,square,sort&compress]{natbib}
8: \usepackage{amsfonts}
9: \usepackage{bbm}
10: 
11: \renewcommand{\baselinestretch}{1.25}
12: 
13: \allowdisplaybreaks
14: 
15: % Command definitions
16: 
17: \newcommand{\eg} {{\it e.g.}}
18: \newcommand{\ie} {{\it i.e.}}
19: \newcommand{\pnu}[1] {\overset{\smash{\scriptscriptstyle (-)}}
20: {\nu}_{\hskip-3pt #1}}
21: \newcommand{\CL}   {C.L.}
22: \newcommand{\dof}  {d.o.f.}
23: \newcommand{\eVq}  {\text{eV}^2}
24: \newcommand{\Sol}  {\textsc{sol}}
25: \newcommand{\SlKm} {\textsc{sol+kam}}
26: \newcommand{\Atm}  {\textsc{atm}}
27: \newcommand{\Chooz}{\textsc{chooz}}
28: \newcommand{\Dms}  {\Delta m^2_\Sol}
29: \newcommand{\Dma}  {\Delta m^2_\Atm}
30: \newcommand{\Dcq}  {\Delta\chi^2}
31: \newcommand{\EtAl}  {{\it et al.\/}}
32: \newcommand{\flux}[2][]{\ensuremath{\ifthenelse{\equal{#1}{}}{}{^{#1}\!}
33: \mathit{#2}}}
34: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
35: 
36: \newcommand{\AddrAHEP}{
37:   AHEP Group, Instituto de F\'{\i}sica Corpuscular --
38:   C.S.I.C./Universitat de Val{\`e}ncia \\
39:   Edificio Institutos de Paterna, Apt 22085, E--46071 Val{\`e}ncia, Spain}
40: 
41:  \title{Geotomography with  solar and supernova neutrinos}
42: 
43: \author{ E. Kh. Akhmedov\thanks{On leave from the National Research
44:     Centre Kurchatov Institute, Moscow, Russia} \\
45: 
46: The Abdus Salam International Centre for Theoretical Physics \\ 
47: Strada Costiera 11, 34014 Trieste, Italy \\
48:   E-mail:  \email{akhmedov@ictp.trieste.it}}
49: 
50: \author{
51: M.~A.~T{\'o}rtola and
52:   J.~W.~F.~Valle \\
53:   \AddrAHEP \\
54:   E-mail:  \email{mariam@ific.uv.es},
55:   \email{valle@ific.uv.es}}
56: 
57: \abstract{ We show how by studying the Earth matter effect on
58:   oscillations of solar and supernova neutrinos inside the Earth one
59:   can in principle reconstruct the electron number density profile of
60:   the Earth.  A direct inversion of the oscillation problem is
61:   possible due to the existence of a very simple analytic formula for
62:   the Earth matter effect on oscillations of solar and supernova
63:   neutrinos. From the point of view of the Earth tomography, these
64:   oscillations have a number of advantages over the oscillations of
65:   the accelerator or atmospheric neutrinos, which stem from the fact
66:   that solar and supernova neutrinos are coming to the Earth as mass
67:   eigenstates rather than flavour eigenstates. In particular, this
68:   allows reconstruction of density profiles even over relatively short
69:   neutrino path lengths in the Earth, and also of asymmetric profiles.
70:   We study the requirements that future experiments must meet to
71:   achieve a given accuracy of the tomography of the Earth.  }
72: 
73: \keywords{Neutrino mass and mixing; solar and supernova neutrinos;  
74:   Earth structure}
75: 
76: \preprint{
77: IFIC/05-15 \\
78:   hep-ph/0502154}
79:   
80: \begin{document}
81: 
82: \section{Introduction}
83: The idea to use neutrinos in order to probe the interior of the Earth was 
84: originally put forward more than 30 years ago \cite{Volkova}, and since then 
85: has undergone a number of improvements and modifications  \cite{DeRujula:1983ya,
86: Wilson:1983an,Askarian:ca,Borisov:sm,Jain:1999kp,Ermilova:pw,Chechin,
87: Nicolaidis:fe,Nicolaidis:1990jm,Ohlsson:2001ck,Ohlsson:2001fy,
88: Ioannisian:2002yj,Lindner:2002wm,Winter:2005we}. In the present paper we offer 
89: a new insight into this problem, based on some recent theoretical and 
90: experimental developments in neutrino physics. 
91: 
92: There are in general two possible approaches to the problem of probing 
93: the Earth's structure with neutrinos: neutrino absorption tomography 
94: \cite{Volkova,DeRujula:1983ya,Wilson:1983an,Askarian:ca,Borisov:sm,Jain:1999kp} 
95: and neutrino oscillation tomography \cite{Ermilova:pw,Chechin,Nicolaidis:fe,
96: Nicolaidis:1990jm,Ohlsson:2001ck,Ohlsson:2001fy,Ioannisian:2002yj,
97: Lindner:2002wm}. The absorption tomography is based on the attenuation of the 
98: flux of neutrinos due to their scattering and absorption inside the Earth and 
99: so is only sensitive to the cumulative effect of neutrino absorption 
100: and deflection along its trajectory. Therefore the absorption tomography 
101: cannot give information about the matter density distribution along a given 
102: neutrino trajectory inside the Earth and requires many baselines to 
103: reconstruct the Earth's density profile.
104: 
105: The oscillation tomography of the Earth is possible due to the fact that 
106: neutrino oscillations in matter are different from those in vacuum 
107: \cite{Wolfenstein:1977ue,Mikheev:gs}. By studying matter effect on neutrino 
108: oscillations one can therefore probe the matter density distribution 
109: along the neutrino path. Being based on an interference phenomenon, the 
110: neutrino oscillation tomography has a much richer potential for studying the 
111: structure of the Earth. In particular, it is in principle possible to use 
112: just one baseline and probe the Earth's density at various points along 
113: the neutrino trajectory. 
114: 
115: In most studies on neutrino oscillation tomography, accelerators were 
116: considered  as the neutrino source. However, solar and supernova neutrinos 
117: have a number of advantages over the accelerator neutrinos in this respect 
118: as the probe of the Earth's interior \cite{Ioannisian:2002yj,Lindner:2002wm}. 
119: %
120: First and foremost, the Sun and a supernova provide us with a free source 
121: of neutrinos. 
122: %
123: In addition, it is very important that, due to the loss of coherence on 
124: their way to the Earth, solar and supernova neutrinos arrive at the Earth as 
125: mass eigenstates rather than flavour eigenstates. For oscillations of mass
126: eigenstate neutrinos in a medium, matter effects fully develop at much
127: shorter distances than they do for flavour eigenstates
128: \cite{Akhmedov:2000cs}, therefore by using solar or supernova
129: neutrinos one can probe the Earth's density distribution even over
130: relatively short neutrino path lengths inside the Earth.  Another
131: virtue of studying the oscillations of mass eigenstate neutrinos in
132: matter is that in that case asymmetric density profiles can be
133: reconstructed
134: \cite{Ioannisian:2002yj,deHolanda:2004fd,Ioannisian:2004jk}. 
135: 
136: Although to first approximation the Earth's density profile can be
137: considered as spherically symmetric, some deviations from perfect
138: symmetry are possible, especially over relatively short scales.
139: Exploring such short-scale inhomogeneities would be of particular
140: interest from the point of view of the possibility of oil or gas
141: prospecting \cite{Ohlsson:2001fy,Ioannisian:2002yj}.  As was shown in
142: \cite{Akhmedov:2001kd}, studying asymmetric density profiles through
143: the usual oscillations of flavour eigenstate neutrinos is practically
144: impossible (see also the discussion in section \ref{sec:symasym}).
145: 
146: Direct inversion of the neutrino oscillations problem in order to reconstruct 
147: the matter density profile is in general a difficult and subtle mathematical 
148: problem -- it reduces to the reconstruction of the potential of the 
149: Schr\"odinger equation from its solution. In the two-flavour case, this in 
150: general requires the knowledge of the energy dependence of the absolute 
151: value of one of the components of the neutrino wave function and also of 
152: the relative phase between the two components \cite{Ermilova:pw,Chechin}, 
153: which is not measured in neutrino flavour oscillation experiments. Therefore, 
154: one usually has to resort to indirect methods, for example, by generating 
155: random density distributions and comparing the corresponding predictions for 
156: oscillation probabilities with simulated data for the ``true'' profile 
157: \cite{Nicolaidis:1990jm,Ohlsson:2001ck,Ohlsson:2001fy}. This is a complicated 
158: and time consuming procedure of limited accuracy. Indeed, one normally 
159: represents the matter density profile along the neutrino path as a relatively 
160: small number of layers of constant and randomly chosen densities, which gives 
161: rather poor resolution due to the obvious limitation on the number of layers. 
162: 
163: 
164: In the present paper we develop a novel {\em direct} approach to the
165: neutrino oscillation tomography of the Earth. It is based on a
166: recently found simple expression for the Earth matter effect on the
167: oscillations of solar and supernova neutrinos in the Earth.  The
168: similarity of this expression to the Fourier transform of the Earth's
169: density profile allows one to employ a modified inverse Fourier
170: transformation and reconstruct the density profile in a simple and
171: straightforward way.
172: 
173: The paper is organized as follows. In section~\ref{sec:gener} we
174: discuss some general features of the Earth matter effect on
175: oscillations of solar and supernova neutrinos inside the Earth and
176: present the main formulas which will be used for the inversion of the
177: oscillation problem. In section~\ref{sec:symasym} we consider the
178: advantages of oscillations of solar and supernova neutrinos over the
179: other neutrino oscillations in reconstructing asymmetric matter
180: density profiles. In section~\ref{sec:lin} we consider the inversion
181: procedure based on the simplest formula for the Earth regeneration
182: factor, valid for neutrino path length inside the Earth $L\ll 1700$ km
183: (linear regime). We also develop simple iteration procedures which
184: allow one to overcome the difficulty related to the lack of knowledge
185: of the regeneration factor in the domain of high neutrino energies. In
186: section~\ref{sec:nonlin} we briefly discuss the Earth density
187: reconstruction based on the more accurate expression for the Earth
188: regeneration factor, which allows one to study longer neutrino path
189: lengths in the Earth (non-linear regime).  The requirements to the
190: experimental setups which have to be met in order to achieve a given
191: accuracy of the reconstructed Earth density profile are considered in
192: section~\ref{sec:exptl}. In particular, we discuss the effects of
193: finite energy resolution of the detector on the accuracy of the
194: oscillation tomography of the Earth. Our main results are summarized
195: and discussed in section~\ref{sec:disc}. Some technical details of our
196: calculations are given in the Appendix.
197: 
198: 
199: \section{Generalities}
200: \label{sec:gener}
201: 
202: In the 3-flavour framework, neutrino oscillations are in general described
203: by two mass squared differences, $\Delta m_{21}^2$ and $\Delta m_{31}^2$, 
204: three mixing angles, $\theta_{12}$, $\theta_{13}$ and $\theta_{23}$, and
205: the Dirac-type CP-violating phase $\delta_{\rm CP}$. For oscillations
206: of solar or supernova neutrinos inside the Earth, the third mass
207: eigenstate essentially decouples, and the relevant parameters are
208: $\Delta m_{21}^2$, $\theta_{12}$ and $\theta_{13}$
209: \cite{Blennow:2003xw,Akhmedov:2004rq}.  The first two of these are
210: determined from the solar neutrino data and long-baseline reactor
211: experiment KamLAND \cite{solardata,KamLAND}, while for the mixing
212: angle $\theta_{13}$ only an upper bound exists.
213: For example, the recent global fit of neutrino 
214: data of ref.~\cite{Maltoni:2004ei} gives for the solar neutrino oscillation 
215: parameters the $3\sigma$ allowed ranges $\theta_{12}=(28.7\div 38.1)
216: ^\circ$ and $\Delta m_{21}^2=(7.1\div 8.9)\times 10^{-5}$ eV$^2$, 
217: with the best-fit values $\theta_{12}=33.2^\circ$ and $\Delta m_{21}^2 =
218: 7.9\times 10^{-5} $eV$^2$, while for the mixing angle $\theta_{13}$ 
219: one finds $\theta_{13} \lesssim 9.1^\circ~(13.1)^\circ$, or $\sin \theta_{13} 
220: \lesssim 0.16~(0.23)$ at 90\% C.L. (3$\sigma$). 
221: 
222: Consider a flux of solar or supernova neutrinos arriving at the Earth 
223: and traveling a distance $L$ inside the Earth before reaching the 
224: detector. Due to the loss of coherence on their way to the Earth,
225: the incoming neutrinos represent an incoherent sum of fluxes of 
226: mass-eigenstate neutrinos (see, e.g., ref.~\cite{Dighe:1999id}).
227: The Earth matter effect on oscillations of such neutrinos inside the 
228: Earth is fully described by the so-called regeneration factor 
229: $P_{2e}^{\oplus}-P_{2e}^{(0)}$. Here $P_{2e}^{\oplus}$ is the probability 
230: that a neutrino arriving at the Earth as a mass eigenstate $\nu_2$ is 
231: found at the detector in the $\nu_e$ state after having traveled a 
232: distance $L$ inside the Earth, and $P_{2e}^{(0)}$ is the projection of the 
233: second mass eigenstate onto $\nu_e$: $P_{2e}^{(0)}=|U_{e2}|^2$, $U$ being 
234: the leptonic mixing matrix in vacuum. Note that $P_{2e}^{(0)}$ is in fact 
235: the value of $P_{2e}^\oplus$ in the limit of vanishing matter density  
236: or zero distance traveled inside the Earth. 
237: 
238: 
239: As has been shown in \cite{Akhmedov:2004rq}, in the 3-flavour 
240: framework the Earth 
241: regeneration factor for solar and supernova neutrinos can be written as 
242: \begin{equation}
243: P_{2e}^\oplus-P_{2e}^{(0)}=\frac{1}{2}\,\cos^2\theta_{13}\,\sin^2 
244: 2\theta_{12}\,f(\delta)\,,
245: \label{reg}
246: \end{equation}
247: where
248: \begin{equation}
249: %P_{2e}^{\oplus}-P_{2e}^{(0)}= c_{13}^4 \sin^2 2\theta_{12}\frac{1}{2}
250: f(\delta)=\int_0^L\! dx \,V(x) \sin\left[2\int\limits_x^{L}\!\omega(x')
251: \,dx'\right]\,, 
252: \label{f1}
253: \end{equation}
254: with
255: \begin{equation}
256: \omega(x)=\sqrt{[\cos 2\theta_{12}\,\delta-V(x)/2]^2+\delta^2
257: \sin^2 2\theta_{12}}\,,\qquad\qquad 
258: \delta=\frac{\Delta m_{21}^2}{4E}\,. 
259: \label{omega}
260: \end{equation}
261: The effective matter-induced potential of neutrinos $V(x)$ in eqs. 
262: (\ref{f1}) and (\ref{omega}) is related to the charged-current 
263: potential 
264: $V_{\rm CC}(x)$ through
265: \begin{equation}
266: V(x)=\cos^2\theta_{13} V_{\rm CC}(x)=\cos^2\theta_{13} \sqrt{2}\,G_F\,
267: N_e(x)\,,
268: \label{Vcc}
269: \end{equation}
270: where $G_F$ is the Fermi constant and $N_e(x)$ is the electron number density 
271: in matter, $x$ being the coordinate along the neutrino path in the Earth. 
272: The 2-flavour ($\theta_{13}=0$) version of eqs.~(\ref{reg})-(\ref{omega}) 
273: was derived in~\cite{Ioannisian:2004jk}, and similar formulas were also 
274: found in ref.~\cite{deHolanda:2004fd}.  
275: 
276: Equations (\ref{reg}) and (\ref{f1}) were obtained under the assumption 
277: $V(x)\ll 2\delta$, which is very well satisfied for oscillations of solar 
278: neutrinos in the Earth. It is also satisfied with a good accuracy for 
279: supernova neutrinos (except for very high energy ones, which are on the 
280: tail of the supernova neutrino spectrum).  If, in addition, one also 
281: requires $V L\ll 1$, eq.~(\ref{f1}) simplifies to \cite{Akhmedov:2004rq} 
282: \begin{equation}
283: f(\delta)=\int_0^L V(y)\sin 2\delta (L-y) dy\,.
284: \label{f2}
285: \end{equation}
286: 
287: Equation (\ref{f2}) is very suggestive: it has a Fourier integral 
288: form and actually means that in the small $V$ limit the function 
289: $f(\delta)$ is just the Fourier transform of matter-induced neutrino 
290: potential $V(x)$ 
291: \footnote{The finite range of integration in (\ref{f2}) is related to 
292: the fact that the function $V(x)$ is defined on the finite interval 
293: $0\le x \le L$.}.
294: Therefore, if $f(\delta)$ is determined experimentally through 
295: eq.~(\ref{reg}), one can employ the inverse Fourier transformation to 
296: reconstruct the effective matter-induced potential  $V(x)$: 
297: \begin{equation}
298: V(x)=\frac{4}{\pi}\int_0^\infty f(\delta)\sin 2\delta (L-x) d\delta\,.
299: \label{inv1}
300: \end{equation}
301: The electron number density profile of the Earth $N_e(x)$ can then be 
302: found from eq.~(\ref{Vcc}). 
303: 
304: The Earth density profile could in principle be exactly reconstructed 
305: from the solar and supernova neutrino data through eq. (\ref{inv1}) 
306: under the following conditions: 
307: 
308: \begin{enumerate}
309: \item
310: Eqs. (\ref{reg}) and (\ref{f2}) are exact;
311: 
312: \item
313: The function $f(\delta)$ is precisely measured in the whole interval 
314: $0\le\delta < \infty$ (i.e. in the infinite interval of neutrino energies 
315: $0\le E < \infty$);
316: 
317: \item The $\delta$-dependence of the function $f(\delta)$ is known
318:   precisely, i.e the detectors have perfect energy resolution, and can
319:   determine the energy of incoming neutrinos from those of the
320:   secondary particles. In addition, the neutrino parameters $\Delta
321:   m_{21}^2$, $\theta_{12}$ and $\theta_{13}$ are precisely known.
322: \end{enumerate}
323: 
324: In reality, none of these conditions is satisfied: eqs. (\ref{reg}) and 
325: (\ref{f2}) are only valid in the limit $V(x)\ll 2\delta$, $\,V L\ll 1$, 
326: the regeneration factor $P_{2e}^\oplus-P_{2e}^{(0)}$ (and so the function 
327: $f(\delta))$ can only be measured in a finite interval of energies and with 
328: some experimental errors, 
329: the detectors have finite energy resolution and can only 
330: give limited information on the energy of incoming neutrinos, and 
331: the neutrino parameters are only known with certain experimental 
332: uncertainties. In what follows we will study the constraints that these 
333: limitations put on the accuracy of the reconstructed potential $V(x)$, 
334: by relaxing conditions (1) -- (3) one by one. Conversely, we shall discuss 
335: the requirements that are put on the experimental installations by the 
336: condition of reconstructing the potential $V(x)$ with a given accuracy. 
337: We will also discuss the ways in which some of the above-mentioned 
338: limitations can be overcome. 
339: 
340: \section{Symmetric versus asymmetric density profiles}
341: \label{sec:symasym}
342: 
343: By symmetric density profiles we mean the profiles that are symmetric with 
344: respect to the midpoint of the neutrino trajectory inside the Earth. 
345: They give rise to the potentials that have the same property, i.e. 
346: \begin{equation}
347: V(L-x)=V(x)\,. 
348: \label{sym}
349: \end{equation}
350: If the electron number density of the Earth $N_e$ was 
351: exactly spherically symmetric, the corresponding neutrino potential $V(x)$ 
352: would have satisfied eq.~(\ref{sym}). However, this symmetry is only 
353: approximate; in particular, it is violated by inhomogeneities of the Earth's 
354: density distribution on short length scales. Studying these inhomogeneities  
355: may be especially interesting, e.g., from the point of view of possible oil or 
356: gas prospecting. 
357:         
358: Effects of asymmetric density profiles on oscillations of solar
359: neutrinos in the Earth have been previously discussed in
360: \cite{Ioannisian:2002yj,deHolanda:2004fd, Ioannisian:2004jk}. Here we
361: give a more detailed discussion of neutrino oscillations in asymmetric
362: matter and also compare in this context mass-to-flavour and pure
363: flavour neutrino oscillations.
364: 
365: It is easy to show that in the two-flavour (2f) framework the probabilities 
366: of oscillations between neutrinos of different flavour are the same for the
367: potentials $V(x)$ and \mbox{$V(L-x)$}. Indeed, the 2f neutrino oscillation 
368: probabilities are invariant under the time reversal transformation $P_{ab}\to 
369: P_{ba}$. This follows from the unitarity relations 
370: \begin{eqnarray}
371: P_{aa}+P_{ab} &=& 1\,, \nonumber \\
372: P_{aa}+P_{ba} &=& 1\,,
373: \label{unit1}
374: \end{eqnarray}
375: which enforce $P_{ab}=P_{ba}$ \cite{deGouvea:2000un,Akhmedov:2001kd}. On the 
376: other hand, for an arbitrary number of flavours, time reversal transformation 
377: of the probabilities of neutrino oscillations in matter is equivalent to 
378: flipping the sign of the Dirac-type CP-violating phases $\{\delta_{\rm CP}\}$ 
379: and replacing the potential $V(x)$ with the reverse potential $V(L-x)$ 
380: \cite{Akhmedov:2001kd}. Since Dirac-type CP-violation is absent in the 2f 
381: case, from $P_{ab}=P_{ba}$ one immediately finds that 2f oscillation 
382: probabilities are invariant under the transformation $V(x)\to V(L-x)$. 
383: 
384: This can also be expressed in the following way. The evolution matrix $S$ of 
385: a 2f neutrino system is a $2\times 2$ unitary matrix which 
386: can be written in the flavour eigenstate basis as 
387: %
388: \begin{equation} 
389: S=\left( \begin{array}{cc} ~\alpha & ~\beta
390: \\ -\beta^* & ~~\alpha^* \end{array} \right) 
391: \label{S} 
392: \end{equation}
393: %
394: with $|\alpha|^2+|\beta|^2=1$. In terms of the elements of $S$, the 
395: oscillation probabilities are given as $P_{ab}=|S_{ba}|^2$. Hence, time 
396: reversal of the evolution matrix (which in the 2f case is equivalent to the 
397: transformation $V(x)\to V(L-x)$) reduces to the transposition $S_{ab}\to 
398: S_{ba}$, i.e.  
399: \begin{equation} 
400: \alpha \to \alpha\,, \qquad\qquad \beta \to -\beta^*\,. 
401: \label{Trev} 
402: \end{equation}
403: Since the transition probabilities $P_{ab}=|\beta|^2$, as well as the survival 
404: probabilities $P_{aa}=|\alpha|^2$, are invariant under the transformation 
405: (\ref{Trev}), they cannot discriminate between the potentials $V(x)$ and 
406: $V(L-x)$. This means that flavour oscillations cannot be used for a unique   
407: reconstruction of asymmetric density profiles: there will always be a two-fold 
408: ambiguity.
409: 
410: The situation is drastically different when one considers the transitions 
411: between the mass and flavour eigenstates, as is the case for oscillations 
412: of solar and supernova neutrinos inside the Earth. In that case the 
413: unitarity conditions read
414: \begin{eqnarray}
415: P_{1e} \,&+& \, P_{2e} = 1\,, \nonumber \\
416: P_{1e}^{(0)} &+& P_{2e}^{(0)} = 1\,,
417: \label{unit2}
418: \end{eqnarray}
419: from which one only finds $P_{2e}-P_{2e}^{(0)}=-(P_{1e}^{(0)}-P_{1e}^{(0)})$, 
420: and no restrictions on the behaviour of the probabilities under the 
421: replacement $V(x)\to V(L-x)$ follow. In fact, it was shown in 
422: \cite{Akhmedov:2004rq} that in the 2f case the Earth regeneration factor 
423: $P_{2e}-P_{2e}^{(0)}$ can be expressed through the elements of the matrix 
424: $S$ as
425: \begin{equation}
426: P_{2e}-P_{2e}^{(0)}=\cos 2\theta_{12} |\beta|^2+2 \sin 2\theta_{12}{\rm Re}
427: (\alpha^* \beta)\,.
428: \label{TT}
429: \end{equation} While the first term on the right-hand side of eq.~(\ref{TT}) is
430: invariant with respect to the transformation (\ref{Trev}), the second
431: is in general not \footnote{It is only invariant when $\beta$ is pure
432:   imaginary, which is the case for symmetric density profiles
433:   \cite{Akhmedov:2001kd}. }.  Thus, unlike the flavour oscillations,
434: the oscillations of mass eigenstate neutrinos into flavour eigenstate
435: ones can be used to uniquely reconstruct asymmetric density profiles
436: even in the 2f framework. As was shown in
437: refs.~\cite{Blennow:2003xw,Akhmedov:2004rq}, in the case of
438: oscillations of solar or supernova neutrinos inside the Earth the
439: third neutrino flavour essentially decouples, and to a very high
440: accuracy the problem is reduced to an effective 2-flavour one.
441: Therefore, our conclusion about the impossibility of using the neutrino
442: flavour oscillations inside the Earth for an unambiguous
443: reconstruction of asymmetric density profiles holds also in the
444:   3-flavour case, hence the superiority of mass-to-flavour
445:   oscillations.  This is in accord with the previous findings that
446: matter-induced T violation in neutrino flavour oscillations inside the
447: Earth is too small to be measured~\cite{Akhmedov:2001kd}.
448: 
449: \section{Linear regime}
450: \label{sec:lin}
451: 
452: The linear regime of the inverse problem of neutrino oscillations in the Earth is 
453: based on the simple formula (\ref{f2}) for the function $f(\delta)$, which was 
454: derived under the assumptions
455: \begin{equation}
456: V/2\delta \ll 1\,,\qquad\quad V L\ll 1\,. 
457: \label{cond2}
458: \end{equation} 
459: For the matter density in the upper mantle of the Earth,
460: $\rho\simeq 3$ g/cm$^3$, the second of these conditions leads to the
461: upper limit on the allowed neutrino path lengths in the Earth \begin{equation} L\ll
462: 1700 ~\mbox{km}\,,
463: \label{upper}
464: \end{equation}
465: which we will now assume to be satisfied. This condition will be relaxed in 
466: the discussion of the non-linear regime in section~\ref{sec:nonlin}. 
467: %
468: \subsection{Integration over finite energy intervals}
469: \label{sec:finiteInt}
470: 
471: We shall first assume that eqs. (\ref{reg}) and (\ref{f2}) are exact, 
472: but the function $f(\delta)$ is only known in a finite interval 
473: $[\delta_{min}, \delta_{max}]$ (i.e. in a finite interval of neutrino 
474: energies $E_{min}\le E\le E_{max}$). To study the effects of finite 
475: $\delta_{min}$ and $\delta_{max}$, consider the integral of the type 
476: (\ref{inv1}) in the finite limits:
477: \begin{equation}
478: \frac{4}{\pi}\int_{\delta_{min}}^{\delta_{max}}\!f(\delta)\sin 2\delta 
479: (L-x) d\delta=\frac{1}{\pi}\int_0^L\!dy V(y)\left.\!\left\{\frac{\sin 
480: 2\delta (x-y)}{x-y}-\frac{\sin 2\delta(2L-x-y)}{2L-x-y}\right\}
481: \right|_{\delta_{min}}^{\delta_{max}}.
482: \label{aux1}
483: \end{equation}
484: Here we have used eq. (\ref{f2}), changed the order of integrations 
485: and performed the integral over $\delta$.
486: 
487: 
488: Ideally, one would like to have $\delta_{min} L\ll 1$ and $\delta_{max} L\gg 1$ 
489: in order that the integral in eq.~(\ref{aux1}) approach the integral over the 
490: infinite interval $0\le \delta < \infty$ in eq.~(\ref{inv1}) as closely as 
491: possible. As we shall see, having large enough $\delta_{max}$ in principle does 
492: not pose a problem.
493: In contrast to this, in most situations of practical interest $\delta_{min}
494: \gtrsim L^{-1}$, i.e. the condition $\delta_{min} L\ll 1$ is not satisfied, 
495: which could be a serious problem. 
496: We shall show, however, that this difficulty can be readily overcome.  
497: 
498: Let us first study the effect of finite $\delta_{max}$ in eq.~(\ref{aux1}). 
499: In the limit $k\to\infty$, the function $\sin kx/x$ goes to $\pi\delta(x)$ 
500: \footnote{Note that here $\delta(x)$ denotes Dirac's delta-function, not to 
501: be confused with the parameter $\delta$ defined in eq. (\ref{omega}).}, 
502: therefore for $\delta_{max}\to \infty$ the upper limit of the integral in 
503: (\ref{aux1}) would yield $V(x)-V(2L-x)$, i.e. essentially the potential $V(x)$ 
504: (note that $V(2L-x)=0$ for all $x\ne L$). The function $g(x)=\sin kx/x$ for 
505: finite $k>0$ is plotted in fig.~\ref{fig:g}. The width of its central peak is 
506: $\simeq \pi/k$. With increasing $k$, the peak becomes higher and narrower, and 
507: the amplitude of the side oscillations quickly decreases. It is therefore 
508: clear that finite $\delta_{max}$ leads to the finite coordinate resolution 
509: $\Delta x\simeq \pi/\delta_{max}$ of the reconstructed potential $V(x)$ as 
510: well as to small oscillations of the reconstructed potential around the 
511: true one. 
512: %
513: \begin{figure}[htbp]
514:   \centering
515: \includegraphics[height=7.5cm,width=.7\linewidth,clip]{fig1.eps}
516:   \caption{Function $g(x)=\sin(kx)/x$}
517:  \label{fig:g}
518: \end{figure}
519: %
520: 
521: Large values of $\delta$ correspond to small values of neutrino energy 
522: $E$ (see eq.~(\ref{omega})), therefore, the smaller the neutrino energy, 
523: the shorter the coordinate scale on which the Earth density profile can 
524: be probed. This may look somewhat counter-intuitive; however, one 
525: should remember that we are probing the Earth density distribution with 
526: neutrino oscillations rather than with direct neutrino-matter 
527: interactions. The smaller the neutrino energy, the shorter the 
528: oscillation length, and so the finer the density structures that can 
529: be probed. 
530: %
531: In future low-energy solar neutrino experiments, neutrinos of energies
532: as small as a hundred keV  to 1 MeV can probably be detected
533: \cite{Raghavan,Ejiri:1999rk,LOWE}; this would correspond to
534: $\delta_{max}\simeq (2\div 20)\times 10^{-11}$ eV, or $\Delta x\sim
535: \pi/\delta_{max}\simeq (3\div 30)$ km, which is a very good coordinate
536: resolution \footnote{This holds in the idealized case of perfect
537:   energy resolution of neutrino detectors. Finite energy resolution of
538:   the detectors is expected to reduce the coordinate resolution of the
539:   reconstruction procedure, see section~\ref{sec:finiteE}.}. For the
540: neutrino path lengths $L\gtrsim 100$ km one finds $\delta_{max} L
541: \gtrsim (10\div 100)$, i.e. the condition $\delta_{max} L \gg 1$ can
542: be easily met.
543: 
544: At the same time, as we have already mentioned, having a sufficiently
545: small $\delta_{min}$ may be a fundamental problem. There are two
546: reasons for that.  First, small $\delta_{min}$ implies large neutrino
547: energies, and there are upper limits to the available neutrino
548: energies. The spectrum of solar neutrinos extends up to 15 MeV ($hep$
549: neutrinos have slightly higher energies, but their flux is very
550: small); the average energy of supernova neutrinos is $\sim 20$
551: MeV~\cite{Keil:2002in}.  Yet, this problem can to some extent be
552: alleviated, at least in principle.  It might be possible to measure
553: the Earth matter effect for supernova neutrinos of energies up to
554: $\sim 100$ MeV, which are on the high-energy tail of the spectrum, but
555: still not too far from the mean energy. Thus, for a nearby supernova
556: and large enough detectors one can probably have sufficient
557: statistics.
558: 
559: The second obstacle is of more fundamental nature. Our expressions (\ref{f1}) 
560: and (\ref{f2}) are only valid in the approximation $V/2\delta \ll 1$, which 
561: may break down for too small $\delta_{min}$. This gives a lower limit on the 
562: values of $\delta_{min}$ one can use, depending on the accuracy of the 
563: reconstructed potential one wants to achieve. In principle, one could attempt 
564: at deriving a more general expression for $f(\delta)$, not relying on the 
565: approximation $V/2\delta \ll 1$; in particular, it is fairly easy to study the 
566: opposite case $V/2\delta \gg 1$. However, for $V\gtrsim 2\delta$ the 
567: expression for $f(\delta)$ does not have a simple dependence on the 
568: potential $V(x)$, and solving the inverse problem of neutrino oscillations 
569: becomes a difficult task. 
570: 
571: For an estimate of the constraint on $\delta_{min}$ imposed by the condition 
572: of small $V/2\delta$, let us require $V/2\delta_{min} < 1/5$. For the matter 
573: density in the upper mantle of the Earth, $\rho\simeq 3$ g/cm$^3$, this gives 
574: $\delta_{min}\gtrsim 2.8\times 10^{-13}$ eV, which for $\Delta m_{21}^2 
575: \simeq 7.9\times 10^{-5}$ eV$^2$ leads to $E_{max}\lesssim 70$ MeV. For 
576: neutrinos passing through the core of the Earth, the constraints on 
577: $\delta_{min}$ and $E_{max}$ will be a factor of 3 -- 4 more stringent.  
578: 
579: Let us now estimate the magnitude of $\delta_{min}L$ corresponding to $L\simeq 
580: 300$ km, which is a representative value of neutrino potholing inside the Earth, 
581: satisfying condition (\ref{upper}). For solar neutrinos ($E_{max}
582: \simeq 15$ MeV) we find $\delta_{min}L\simeq 2$, while for supernova neutrinos, 
583: taking $E_{max}\simeq 70$ MeV, we obtain $\delta_{min}L\simeq 0.43$. 
584: Thus, except for very small values of $L$, we have to deal with situations 
585: when $\delta_{min}L\gtrsim 1$. 
586:  
587: How large is the error introduced by non-vanishing $\delta_{min}$ in
588: the reconstruction of the density profile $V(x)$?  To study that, let
589: us consider the integral of the type (\ref{inv1}) with the finite
590: lower integration limit $\delta_{min}= L^{-1}$.  In
591: fig.~\ref{fig:step-dmin1} the lowest curve gives the result of such a
592: calculation for the step-function density profile (shown by the dashed
593: line) and $\delta_{max}=300 L^{-1}$. One can see several interesting
594: features of the result. First, the deviation from the exact profile is
595: relatively small near the detector ($x\simeq L$), but reaches about a
596: factor of three far from it ($x\simeq 0$). Second, despite a
597: significant deviation from the exact profile $V(x)$, the positions and
598: the magnitudes of the jumps in $V(x)$ are reproduced very accurately.
599: Both these features can be easily understood (see sections
600: \ref{sec:iterI} and \ref{sec:small} below).
601: 
602: Thus, we have seen that the error in the reconstructed profile due to the lack 
603: of the knowledge of the Earth matter effect in the domain of low $\delta$ (high 
604: energies) can be quite substantial. We shall show now how this problem can be 
605: cured by invoking simple iteration procedures. 
606:  
607: 
608: 
609: 
610: \subsection{Iteration procedure I}
611: \label{sec:iterI}
612: 
613: Let us study the effect of non-vanishing $\delta_{min}$ in more detail. In 
614: order to do so, we consider the limit $\delta_{max}\to \infty$, which is 
615: justified by the preceding discussion. From eq.~(\ref{aux1}) one then readily 
616: finds 
617: \begin{equation}
618: V(x)=\frac{4}{\pi}\int_{\delta_{min}}^\infty f(\delta)\sin 2\delta (L-x) 
619: d\delta +\frac{1}{\pi}\int_0^L V(y) F(x,y; 2\delta_{min}) dy\,,
620: \label{inv2}
621: \end{equation}
622: where the function $F(x,y;k)$ is defined as  
623: \begin{equation}
624: F(x,y;k)=\frac{\sin k(x-y)}{x-y}-\frac{\sin k(2L-x-y)}{2L-x-y}\,.
625: \label{F1}
626: \end{equation}
627: It is symmetric with respect to its first two arguments: $F(x,y; k)=F(y,x; k)$. 
628: 
629: Equation (\ref{inv2}) is exact provided that eq. (\ref{f2}) is exact. 
630: By comparing it with eq.~(\ref{inv1}), we find that the second integral in 
631: (\ref{inv2}) can be considered as compensating for an error introduced in 
632: eq. (\ref{inv1}) by having a non-zero lower limit in the integral over 
633: $\delta$. However, this compensating integral cannot be calculated directly 
634: because it contains the unknown potential $V(x)$. Thus, we have traded 
635: one unknown quantity -- the function $f(\delta)$ in the domain $\delta < 
636: \delta_{min}$ -- for another.
637: At first sight, this does not do us any good. This is, however, incorrect:
638: eq.~(\ref{inv2}) allows a simple iterative solution.  
639: 
640: 
641: We first note that in the limit $\delta_{min}\to 0$ the second integral in 
642: eq.~(\ref{inv2}) disappears, while the first one yields $V(x)$. Therefore, for 
643: not too large values of $\delta_{min}$ the first term in (\ref{inv2}) is 
644: expected to give a reasonable first approximation to $V(x)$. One can then use 
645: the result in the second integral to obtain the next approximation to $V(x)$, 
646: and so on. Thus, we define
647: \begin{eqnarray}
648: V_0(x) &=& \frac{4}{\pi}\int_{\delta_{min}}^\infty f(\delta)\sin 
649: 2\delta (L-x) d\delta\,, 
650: \label{V0m}\\ 
651: I_0(x) &=&\frac{1}{\pi}\int_0^L V_0(y) F(x,y; 2\delta_{min}) dy\,,
652: \qquad\quad V_1(x)=V_0(x)+I_0(x)\,, 
653: \label{I0}
654: \\
655: & & \nonumber \\
656: & & \dots\dots \nonumber \\
657: & & \nonumber \\
658: I_{n-1}(x) &=&\frac{1}{\pi}\int_0^L V_{n-1}(y) F(x,y; 2\delta_{min}) dy\,,
659: \qquad\quad \!\!V_n(x)=V_0(x)+I_{n-1}(x)\,.
660: \label{iter1}
661: \end{eqnarray}
662: This gives a sequence of potentials $V_0(x),\,V_1(x),\dots,\,V_n(x),
663: \dots$ which, for small enough $\delta_{min}$, converges to $V(x)$. It 
664: should be noted that, while the exact eq.~(\ref{inv2}) is independent 
665: of $\delta_{min}$,\footnote{Its left-hand side is $\delta_{min}$-independent, 
666: so must be the right-hand side. Actually, by {\em requiring} that the 
667: derivative of the right-hand side of (\ref{inv2}) with respect to 
668: $\delta_{min}$vanish, one can recover eq.~(\ref{f2}) for $f(\delta)$.} 
669: our iteration procedure involves 
670: the approximate potentials and so depends on it. The smaller 
671: the chosen value of $\delta_{min}$, the faster the convergence of $V_n(x)$ 
672: to $V(x)$; for $\delta_{min}$ exceeding some critical value (which in 
673: general depends on the profile $N_e(x)$) the iteration procedure fails.
674: 
675: This is illustrated in figs.~\ref{fig:step-dmin1} - \ref{fig:step-asym}. 
676: In fig.~\ref{fig:step-dmin1} we show the potential $V(x)$ 
677: for the step-function model of the Earth density profile, along with the 
678: zeroth-order reconstructed potential $V_0(x)$ and the results of the first, 
679: second and fourth iterations $V_1(x)$, $V_2(x)$ and $V_4(x)$ ($V_3(x)$ is 
680: not shown in order to avoid crowding the figure). The calculations were 
681: performed for $\delta_{min}=L^{-1}$, $\delta_{max}=300 L^{-1}$.  One can see 
682: that already the fourth iteration gives an excellent agreement with the exact 
683: potential. We have checked that for $\delta_{min}\ll L^{-1}$ already the 
684: zeroth-approximation potential $V_0(x)$ gives a very good accuracy (see 
685: also section~\ref{sec:small}). The wiggliness of the reconstructed potentials in 
686: fig.~\ref{fig:step-dmin1} is due to the finiteness of $\delta_{max}$; with 
687: increasing $\delta_{max}$ it decreases. 
688: 
689: 
690: \begin{figure}[htbp]
691:   \centering
692:   \includegraphics[width=.9\linewidth,clip]{step-dm1-dM300.eps}
693:   \caption{Step-function potential $V(x)$ (dashed line), zeroth order 
694:    reconstructed potential $V_0(x)$ and the results of the first, second and 
695:    fourth iterations $V_1(x)$, $V_2(x)$ and $V_4(x)$. The vertical scale is 
696:    that of the corresponding matter density in g/cm$^3$, assuming $Y_e=0.5$ and 
697:    $\theta_{13}=0$. The following values of the parameters were chosen: 
698:    $\delta_{min}=L^{-1}$, $\delta_{max}=300 L^{-1}$.
699:    } 
700:  \label{fig:step-dmin1}
701: \end{figure}
702: 
703: If $\delta_{min}$ exceeds the critical value (which for the chosen profile is 
704: approximately equal to $2.4 L^{-1}$ ), the successive iterations, instead of 
705: approaching the true potential, yield the potentials which more and more deviate 
706: from it. Thus, in this case the iteration procedure fails.  
707: 
708: \begin{figure}[htbp]
709:   \centering
710:   \includegraphics[width=.9\linewidth,clip]{prem-dmin1-4.eps}
711:   \caption{Same as in fig.\ref{fig:step-dmin1}, but for PREM-like profile 
712: (shown by dashed line), 
713: }  
714:  \label{fig:step-prem}
715: \end{figure}
716: 
717: 
718: 
719: The iteration procedure works very well not only for simple profiles like the 
720: step-function profile in fig.~\ref{fig:step-dmin1}, but also for more 
721: complicated ones. This is demonstrated in fig.~\ref{fig:step-prem}, which is 
722: similar to fig.~\ref{fig:step-dmin1}, but was plotted for the PREM-like density 
723: profile of the Earth \cite{Dziewonski:xy}. It should be stressed that this 
724: profile was used in fig.~\ref{fig:step-prem} for illustrative purposes only, 
725: since, in the form we used it, it is a realistic Earth density profile for 
726: $L=2R_\oplus \simeq 12742$ km and not for $L=300$ km (for which the Earth 
727: density is actually better approximated by the step-function profile of 
728: fig.~\ref{fig:step-dmin1}). 
729: %
730: Fig.~\ref{fig:step-asym} is similar to figs.~\ref{fig:step-dmin1} and 
731: \ref{fig:step-prem}, but was produced for an asymmetric Earth 
732: density profile (shown by the dashed line). It clearly demonstrates that 
733: asymmetric profiles can also be reconstructed very well, and so the 
734: inhomogeneities of matter distribution in the Earth can be studied by 
735: the method under consideration.   
736: 
737: \begin{figure}[htbp]
738:     \centering
739:     \includegraphics[width=.9\linewidth,clip]{step-asym.eps}
740:     \caption{Same as in fig.~\ref{fig:step-dmin1}, but for asymmetric
741:     step-function potential (dashed line). } \label{fig:step-asym}
742: \end{figure}
743: 
744: 
745: Finally, we note that if $\delta_{max} L\gg 1$ but still not very large, one 
746: can devise an iterative procedure correcting for the corresponding error in the 
747: reconstructed potential (see Appendix \ref{app:Improved}). 
748: 
749: \subsubsection{Convergence properties of iterations}
750: \label{sec:conv}
751: 
752: We shall now discuss some properties of the iteration procedure
753: (\ref{V0m}) -- (\ref{iter1}). First, we note that the function $F(x,y;
754: 2\delta_{min})$ is positive definite for all $0<x,y<L$ provided that
755: $\delta_{min}<2.246 L^{-1}$ (see Appendix B for the proof). One can
756: then show that for \begin{equation} \delta_{max} \to \infty~~~\mbox{and}~~~
757: \delta_{min}<{\rm min}\{2.246 L^{-1},\,\delta_1\}\,,
758: \label{cond1} 
759: \end{equation}
760: where $\delta_1$ will be defined shortly, the following string of
761: inequalities is satisfied:
762: \begin{equation}
763: V_0(x)<V_1(x)<V_2(x)<... <V(x)\,.
764: \label{ineq1}
765: \end{equation} 
766: This can be proven by induction. Consider first eq.~(\ref{inv2}).
767: {}From the positivity of $V(x)$ and $F(x,y; 2\delta_{min})$ it follows
768: that the second integral in (\ref{inv2}) (which we denote $I(x))$ is
769: positive, and so the first integral, which is the zeroth-approximation
770: potential $V_0(x)$, satisfies $V_0(x)<V(x)$. Generally, the larger
771: $\delta_{min}$, the smaller $V_0(x)$; for large enough $\delta_{min}$
772: the potential $V_0(x)$ may even become negative for some values of $x$
773: in the interval $[0,\,L]$. However, if $\delta_{min}$ does not exceed
774: certain limiting value $\delta_1$ (which we assume), the integral
775: $I_0(x)$ defined in eq.~(\ref{I0}) will still be positive. It is this
776: limiting value $\delta_1$ that appears in eq.~(\ref{cond1}). From the
777: definition of the integral $I_0(x)$ and the obtained condition
778: $V_0(x)<V(x)$ we then find $0<I_0(x)<I(x)$.  Therefore
779: $V_1(x)=V_0(x)+I_0(x)$ satisfies $V_0(x)<V_1(x)<V(x)$. Then, from the
780: definition of $I_1(x)$ we find $I_0(x)<I_1(x)<I(x)$, so that
781: $V_2(x)=V_0(x)+I_1(x)$ satisfies $V_1(x) <V_2(x)<V(x)$. Continuing
782: this procedure, we arrive at eq.~(\ref{ineq1}).
783: 
784: Thus, under the conditions of eq. (\ref{cond1}), each iteration produces the 
785: potential which is larger than that of the previous iteration. The potentials 
786: $V_n(x)$ approach the exact potential $V(x)$ from below and never exceed it. 
787: This is well illustrated by figs.~\ref{fig:step-dmin1} - \ref{fig:step-asym} 
788: (we recall that the wiggliness of the curves disappears in the limit 
789: $\delta_{max} \to\infty$). Conditions (\ref{cond1}) are thus sufficient for 
790: the convergence of the iteration procedure 
791: \footnote{It can be shown that the iteration potentials converge uniformly 
792: to $V(x)$ even for larger values of $\delta_{min}L$. Indeed, a sufficient 
793: condition for the uniform convergence is $\int_0^L F^2(x,y; 2\delta_{min})\,dx
794: \,dy < \pi^2$ (see, e.g., \cite{Goursat}), which is satisfied for 
795: $\delta_{min}L < 2.34$.}.
796: 
797: How can one estimate the critical value of $\delta_{min}$ 
798: above which the iteration procedure would diverge? From the preceding 
799: discussion it follows that the convergence of the iterations to the exact 
800: potential relies on the positivity of the ``correction terms'' $I_n(x)$. For 
801: values $\delta_{min}> \delta_1$ the integral $I_0(x)$ will become negative for 
802: some values of $x$ in the interval $[0,\,L]$, and so for those values of $x$ 
803: the potential $V_1(x)$ will deviate from $V(x)$ more than $V_0(x)$ does. 
804: If this propagates  into the further iterations, the iteration procedure 
805: would fail. However, it is possible that even if some $I_k(x)$ are not 
806: positive definite, higher-order correction integrals $I_n(x)$ with $n>k$ 
807: are. In this case the iteration procedure would still be convergent. 
808: 
809: In practice, it is difficult to determine the critical value of $\delta_{min}$ 
810: precisely. The closer (from below) $\delta_{min}$ to its critical value 
811: $\delta_{crit}$, the larger the number of iterations which is necessary to 
812: achieve a given accuracy of the reconstructed potential. Therefore, with 
813: $\delta_{min}$ approaching $\delta_{crit}$ it becomes more and more 
814: difficult to check if the procedure would converge. In addition, for 
815: $\delta_{min}$ close to $\delta_{crit}$ the behaviour of the iteration 
816: potentials $V_n(x)$ with increasing $n$ does not depend much on whether 
817: $\delta_{min}<\delta_{crit}$ or $\delta_{min}>\delta_{crit}$, so that it is 
818: difficult to decide if the critical value has already been exceeded. 
819: It is therefore reasonable to adopt some ``practical'' definition of 
820: $\delta_{crit}$, for example, as a value of $\delta_{min}$ for which the 
821: number of iterations necessary to reach a 10\% accuracy of the reconstructed 
822: potential reaches 100. For all the density profiles that we studied this 
823: value turned out to be $\delta_{crit}\simeq 2.4 L^{-1}$. We elaborate further 
824: on the convergence of the iterations in the next subsection.
825: 
826: Let us now return to the question of why the zeroth order potential
827: $V_0(x)$ correctly reproduces the positions and magnitudes of the
828: jumps in the exact potential $V(x)$. In eq.~(\ref{inv2}) (which is
829: exact in the linear regime), the second term on the right-hand side is
830: the integral over $y$ of the product of the continuous function
831: $F(x,y; 2\delta_{min})$ and the exact potential $V(y)$, which may have
832: discontinuities, but no $\delta$-function type singularities. Hence,
833: this integral is a continuous function of $x$. This immediately means
834: that all possible jumps in $V(x)$ are contained in the first integral
835: in eq.~(\ref{inv2}), i.e. $V_0(x)$.
836: 
837: 
838: \subsubsection{The limit of small $\delta_{min} L$}
839: \label{sec:small}
840: 
841: It is very instructive to consider the iteration procedure in the limit 
842: $\delta_{min} L\ll 1$. Since $x,y\le L$, the expression for $F(x,y; 
843: 2\delta_{min})$ in this case simplifies to 
844: \begin{equation}
845: F(x,y; 2\delta_{min})\simeq 
846: \frac{16}{3}\,\delta_{min}^3 \,(L-x)(L-y)\,.
847: \label{F2}
848: \end{equation}
849: Then from eq.~(\ref{iter1}) we find a very simple result:
850: \begin{equation}
851: I_n(x)\simeq \frac{16}{3\pi}\,\delta_{min}^3 \,(L-x)\int_0^L V_n(y) (L-y) dy\,.
852: \label{In1}
853: \end{equation}
854: It means that the ``correction terms'' $I_n(x)$, which compensate for 
855: $\delta_{min} \ne 0$ in the inverse Fourier transformation, have a very 
856: simple coordinate dependence $\propto (L-x)$ for all $n$ and scale with 
857: $\delta_{min} L$ as $(\delta_{min} L)^3$. The same is true for the ``exact''
858: correction term $I(x)$ (the second integral in eq.~(\ref{inv2})), which is 
859: obtained from eq.~(\ref{In1}) by replacing in the integrand $V_n(y)$ by the 
860: exact potential $V(y)$. The approximately linear coordinate dependence of the 
861: deviation of the iteration potentials from the exact one can be seen even for 
862: $\delta_{min} L\sim 1$ (see figs.~\ref{fig:step-dmin1} - \ref{fig:step-asym}). 
863: It explains, in particular, why the deviations of $V_n(x)$ from the exact 
864: potential are small at $x\simeq L$ and largest at $x=0$.    
865: 
866: 
867: Using eqs.~(\ref{iter1}) and (\ref{F2}), it is easy to show by 
868: induction that in the limit $\delta_{min} L\ll 1$ 
869: \begin{equation}
870: I_n(x)\simeq \frac{3}{2}\,v_0 \left(\frac{16}{9\pi}\,\delta_{min}^3 
871: L^3\right) 
872: \left[1-\left(\frac{16}{9\pi}\,\delta_{min}^3 L^3\right)^{n+1}\right] 
873: \,\frac{L-x}{L}\,,
874: \label{In2}
875: \end{equation}
876: \begin{equation}
877: V_n(x)\simeq V(x)-\frac{3}{2} v_0 \left(\frac{16}{9\pi}\,\delta_{min}^3 
878: L^3\right)^{n+1}\frac{L-x}{L}\,,~ ~~~~~~~~~~~~~~~~~
879: \label{Vn}
880: \end{equation}
881: where the constant $v_0$ is defined as
882: \begin{equation}
883: v_0=\frac{2}{L^2}\int_0^L V(y) (L-y) dy\,.
884: \label{V0}
885: \end{equation}
886: Note that these equations are in accord with eq.~(\ref{In1}). In the case 
887: of matter of constant density $V(x)=C_0=const$, one has $v_0=C_0$, and 
888: eq.~(\ref{Vn}) takes an especially simple form.   
889: 
890: {}From eq.~(\ref{Vn}) it follows that for $\delta_{min} L\ll 1$ the 
891: deviation of the $n$th-iteration potential $V_n(x)$ from the exact one 
892: scales as $[(16/9\pi)\,\delta_{min}^3 L^3]^{n+1}$, i.e. the convergence of 
893: $V_n(x)$ to the exact potential is very fast. 
894: 
895: Although eqs.~(\ref{In2}) and (\ref{Vn}) were obtained for $\delta_{min} L\ll 
896: 1$, one can expect that they give correct order of magnitude estimates even 
897: for $\delta_{min} L\sim 1$. This allows one to estimate the number 
898: of iterations which is necessary to achieve a given accuracy of the 
899: reconstructed potential in the whole range $\delta_{min}\le \delta_{crit}$. 
900: First, from eq.~(\ref{Vn}) we find that in the 
901: case under consideration the critical value of $\delta_{min}$, above which 
902: the iteration procedure diverges, is $\delta_{crit}=(9\pi/16)^{1/3} L^{-1}$, 
903: so that (\ref{Vn}) can be rewritten as 
904: \begin{equation}
905: V_n(x)=V(x)-(3v_0/2)(\delta_{min}/\delta_{crit})^{3(n+1)} (L-x)/L\,. 
906: \label{Vn2}
907: \end{equation}
908: Assume now that we want the relative error in the reconstructed potential 
909: (which we define here as $|V(x)-V_n(x)|/v_0$)  
910: to be below $\varepsilon$. Then from (\ref{Vn2}) we find that the 
911: necessary number of iterations is 
912: \begin{equation}
913: n\simeq \frac{\ln(2\varepsilon/3)}{3\ln(\delta_{min}/\delta_{crit})}-1\,.
914: \label{n}
915: \end{equation}
916: As an example, take $\varepsilon=0.01$. Then for 
917: $\delta_{min}/\delta_{crit}=0.3$ already the zeroth-approximation potential 
918: $V_0(x)$ has the desired accuracy; for $\delta_{min}/\delta_{crit}=0.9$ 
919: eq.~(\ref{n}) gives $n\simeq 15$, for $\delta_{min}/\delta_{crit}=0.99$ it 
920: gives $n\simeq 165$, and in the limit $\delta_{min}\to\delta_{crit}$ one 
921: finds $n\to \infty$, as expected. It should be remembered, however, that for 
922: $\delta_{min}/\delta_{crit} \sim 1$ eq.~(\ref{n}) gives only a rough 
923: estimate of the necessary number of iterations because eq.~(\ref{Vn2}) was 
924: obtained in the limit $\delta_{min} L \ll 1$, which essentially coincides 
925: with $\delta_{min}/\delta_{crit} \ll 1$.
926: 
927: We have considered in this subsection the reconstruction of the potential 
928: $V(x)$ in the limit of small $\delta_{min} L$ iteratively only in order to 
929: illustrate some general features of the iteration procedure. In fact, 
930: for $\delta_{min} L\ll 1$ it is easy to find the closed-form solution of 
931: eq.~(\ref{inv2}) \footnote{Eq.~(\ref{inv2}) is a linear Fredholm integral 
932: equation of the second kind with the symmetric kernel $F(x,y;2\delta_{min})$. 
933: In the limit $\delta_{min} L\ll 1$ the kernel becomes separable (see 
934: eq.~(\ref{F2})), and the equation is trivially solved.}. Indeed, since in 
935: this case $I(x)\propto (L-x)$, the potential $V(x)$ can be written as 
936: $V(x)=V_0(x)+C_1 (L-x)$ with $C_1$ a constant. Substituting this into 
937: eq.~(\ref{inv2}), one readily determines $C_1$, which gives 
938: \begin{equation}
939: V(x)\simeq V_0(x)+C_2 \frac{\frac{16}{3\pi}\delta_{min}^3 L^3}
940: {1-\frac{16}{9\pi}\delta_{min}^3 L^3}\,\frac{L-x}{L}\, 
941: \label{V}
942: \end{equation}
943: with  
944: \begin{equation}
945: C_2=\frac{1}{L^2}\int_0^L V_0(y)(L-y) dy \,.
946: \label{C2a}
947: \end{equation}
948: Since the function $V_0(x)$ is known, the problem is solved. 
949: 
950: 
951: \subsection{Iteration procedure II}
952: \label{sec:iterII}
953: 
954: In the iteration procedure considered in section \ref{sec:iterI} we
955: assumed that nothing is known \mbox{\em a priori} about the Earth
956: density profile, and the only experimental data available were the
957: neutrino data.  If some (even very rough) prior knowledge of the
958: matter density distribution inside the Earth exists, one can employ a
959: much better and faster iteration procedure to reconstruct the Earth
960: density profile. As an example, we consider here the case when the
961: average matter density along the neutrino path is known. The
962: corresponding average potential 
963: \begin{equation} \overline{V} = \frac{1}{L}
964: \int_0^L V(x)\,dx\,.
965: \label{Vbar}
966: \end{equation} 
967: is therefore known as well. 
968: The new iteration procedure is again described by eqs.~(\ref{I0}) 
969: - (\ref{iter1}), but the zeroth order approximation potential is now 
970: $V_0(x)=\overline{V}=const$. In fig.~\ref{fig:iterII} we plot this potential 
971: (horizontal line) and the first iteration potential $V_1(x)$ for the 
972: asymmetric step-function profile of fig.~\ref{fig:step-asym} (shown by the 
973: dashed line). One can see that, although $V_0$ is completely structureless, 
974: already the first iteration reproduces the exact profile extremely well. This 
975: should be compared with the results of iteration scheme I, for which only 
976: the fourth iteration gives similar accuracy (see fig.~\ref{fig:step-asym}).
977: 
978: %
979: \begin{figure}[htbp]
980:   \centering
981:   \includegraphics[width=.9\linewidth,clip]{step-asym-V0.eps}
982:   \caption{Asymmetric step-function profile (dashed line), zeroth 
983:   approximation potential $V_0(x)=\overline{V}$ and first order
984:   iteration potential $V_1(x)$. Vertical scale and values of 
985:   parameters are the same as in fig.~\ref{fig:step-dmin1}.} 
986:   \label{fig:iterII}
987: \end{figure}
988: %
989: To understand why this iteration scheme works so well, it is instructive to 
990: consider it in the limit $\delta_{min} L\ll 1$. First, we note that 
991: for an arbitrary $\delta_{min}$ and $n\ge 1$ eqs.~(\ref{inv2}) and 
992: (\ref{iter1}) yield 
993: \begin{equation}
994: V(x)-V_n(x) = \frac{1}{\pi}\int_0^L [V(y)-V_{n-1}(y)] 
995: F(x,y; 2\delta_{min}) 
996: dy\,.
997: \label{iter1a}
998: \end{equation}
999: In the limit $\delta_{min} L\ll 1$, the function $F(x,y; 2\delta_{min})$ is 
1000: given by (\ref{F2}); substituting this into (\ref{iter1a}), for $n=1$ 
1001: we find 
1002: \begin{equation}
1003: V(x)-V_1(x)\simeq \left(\frac{8}{3\pi}\,\delta_{min}^3 
1004: L^3\right)\frac{L-x}{L} (v_0-\overline{V})\,,~ ~~~~~~~~~~~~~~~~~
1005: \label{V1}
1006: \end{equation}
1007: where $v_0$ was defined in (\ref{V0}). For $n$th iteration ($n\ge 1$)
1008: we find by induction
1009: \begin{equation}
1010: V(x)-V_n(x)\simeq \frac{3}{2} \left(\frac{16}{9\pi}\,\delta_{min}^3 
1011: ~L^3\right)^n\frac{L-x}{L} (v_0-\overline{V})\,.~ ~~~~~~~~~~~~~~~~~
1012: \label{VnII}
1013: \end{equation}
1014: Comparing this with (\ref{Vn}), we find that in iteration scheme II the 
1015: difference $V(x)-V_n(x)$ contains one less power of $[(16/9\pi)\,
1016: \delta_{min}^3 L^3$], but is proportional to $v_0-\overline{V}$ rather 
1017: than to $v_0$. We shall show now that for symmetric density profiles
1018: ($V(x)=V(L-x)$) the difference $v_0-\overline{V}$ vanishes, and therefore 
1019: already the first iteration reproduces the 
1020: exact profile. Indeed, from eqs.~(\ref{V0}) and (\ref{Vbar}) we have 
1021: \begin{equation}
1022: v_0-\overline{V}=\frac{1}{L^2}\int_0^L V(y)[2 (L-y)-L]\,dy=
1023: \frac{2}{L^2}\int_{-L/2}^{L/2} V(z) z \,dy\,,
1024: \label{Vz}
1025: \end{equation}
1026: where in the last integral we changed the integration variable to $z=L/2-y$.
1027: For symmetric density profiles, $V(z)$ is an even function of $z$, and the 
1028: integral in (\ref{Vz}) vanishes. 
1029: 
1030: What if the density profile is not symmetric? In that case one 
1031: can write $V(z)=V_s(z)+V_a(z)$, where $V_s(z)$ and $V_a(z)$ are the symmetric 
1032: and antisymmetric parts of $V(z)$. The symmetric part $V_s(z)$ does not 
1033: contribute to the integral in (\ref{Vz}), whereas the contribution of $V_a(z)$ 
1034: is small because the inhomogeneities of the Earth density distribution, 
1035: responsible for $V_a$, occur on short coordinate scales \footnote{Generally, 
1036: the contribution of $V_a$ to $v_0-\overline{V}$ is of order $\Delta V x_1/L$, 
1037: where $\Delta V$ is the magnitude of the asymmetric inhomogeneity of the 
1038: potential, and $x_1$ is the coordinate scale of the inhomogeneity. While 
1039: $\Delta V$ may be of order of $V$, $x_1/L$ is always $\ll 1$.}. 
1040: Thus, in scheme II, even for asymmetric density profiles the first iteration 
1041: gives a very good approximation to the exact profile (see 
1042: fig.~\ref{fig:iterII}).
1043: 
1044: It should be noted that, while the above discussion referred to the case 
1045: $\delta_{min}L \ll 1$, fig.~\ref{fig:iterII} actually corresponds to 
1046: $\delta_{min} L=1$. However, our consideration approximately applies to that 
1047: case as well. For $\delta_{min} L\sim 1$, the integrand of the driving term for 
1048: the second iteration in scheme II will be proportional, instead of $V(y)
1049: (L/2-y)$, to the product of $V(y)$ and the function $[F(x,y; 2\delta_{min})-
1050: (1/L)\int_0^L F(x,x'; 2\delta_{min})dx']$. This function, though not exactly 
1051: antisymmetric with respect to $y\to L-y$, is almost antisymmetric (for all 
1052: $x\in [0, L]$). This explains why iteration scheme II works so well even in 
1053: the case $\delta_{min} L \sim 1$.   
1054: 
1055: 
1056: 
1057: \section{Non-linear regime} 
1058: \label{sec:nonlin}
1059: 
1060: 
1061: Our previous discussion of the reconstruction of the Earth density
1062: profile was based on the simple formula (\ref{f2}) for the function
1063: $f(\delta)$, which was derived under the assumptions (\ref{cond2}).
1064: The second of these conditions led to a rather stringent upper limit
1065: on the neutrino path lengths in the Earth (\ref{upper}), or $L\lesssim
1066: 300$ km.  In order to be able to reconstruct the potential $V(x)$ over
1067: larger distances, one has to employ an inversion procedure based on
1068: the more accurate expression (\ref{f1}), which only requires the first
1069: condition in (\ref{cond2}) for its validity. Let us now discuss this
1070: improved expression.
1071: 
1072: Since the function $\omega(x)$ in eq.~(\ref{f1}) depends on $V(x)$, we face a 
1073: non-linear problem now. The first condition in (\ref{cond2}) and the 
1074: non-linearity condition $V L\gtrsim 1$ imply $\delta_{min} L\gg 1$. This, in 
1075: turn, means that the matter density profile cannot be found by invoking an 
1076: iteration procedure
1077: %
1078: \footnote{In sec.~\ref{sec:lin} we showed that in the linear regime
1079:   the critical values of $\delta_{min}$, above which the iteration
1080:   approach fails, correspond to $\delta_{min} L={\cal O}(1)$. The
1081:   situation does not improve in the non-linear regime.}, and one
1082: should resort to different methods of solving eq.~(\ref{f1}).  The
1083: simplest possibility would be to discretize the problem and reduce the
1084: non-linear integral equation (\ref{f1}) to a set of non-linear
1085: algebraic equations, which can then be solved numerically.  However,
1086: eq.~(\ref{f1}) is a non-linear Fredholm integral equation of the first
1087: kind \footnote{We recall that integral equations of the first kind
1088:   involve an unknown function only under the integration sign, whereas
1089:   in equations of the second kind it is also present outside the
1090:   integral.}, and equations of this type are notoriously difficult to
1091: solve. Integral equations of the first kind belong to the so-called
1092: ill-posed problems: their solutions are very unstable, and to arrive
1093: at a reliable result one has to invoke special regularization
1094: procedures. For linear integral equations, such procedures are well
1095: developed (see section \ref{sec:finiteE} and Appendices C and D for
1096: more details); however, no universal regularization techniques exist
1097: for non-linear integral equations of the first kind. The situation is
1098: further complicated by the fact that for $\delta_{min} L\gg 1$ the
1099: potential $V(x)$ enters into the integrand of eq.~(2.2) being
1100: multiplied by a fast oscillating function of the coordinate.  All this
1101: makes the non-linear regime of the inverse problem of neutrino
1102: oscillations in matter very difficult to explore. This regime
1103: therefore requires a dedicated study, which goes beyond the scope of
1104: the present paper.
1105: 
1106: \section{Experimental considerations}
1107: \label{sec:exptl}
1108: 
1109: Up to now we ignored completely the experimental questions, such as the 
1110: effects of the errors in experimental data and of finite energy resolution of 
1111: neutrino detectors on the accuracy of the reconstructed matter density 
1112: distributions. We now turn to these issues. We shall discuss here only the 
1113: linear regime of the density profile reconstruction; the experimental 
1114: questions pertaining to the non-linear regime will be considered 
1115: elsewhere. 
1116: 
1117: 
1118: \subsection{Effects of experimental errors}
1119: \label{sec:errors}
1120: 
1121: For solar neutrinos, the night-day asymmetry of the signal can 
1122: be written as \cite{Akhmedov:2004rq}
1123: \begin{equation}
1124: A_{ND}=2\frac{N-D}{N+D}\simeq -c_{13}^2 \left[\frac{\sin^2 2\theta_{12}\,
1125: \overline{\cos 2\hat{\theta}_{12}}}{1+\cos 2\theta_{12}\,
1126: \overline{\cos 2\hat{\theta}_{12}}}\right] (f(\delta)/c_{13}^2)\,,
1127: \label{And}
1128: \end{equation}
1129: where $\overline{\cos 2\hat{\theta}_{12}}$ is cosine of twice the effective 
1130: 1-2 mixing angle in matter, averaged over the neutrino production coordinate 
1131: inside the Sun \cite{Blennow:2003xw}. The quantity $f(\delta)/c_{13}^2$ is 
1132: independent of $c_{13}^2$ in the linear regime (see eqs.~(\ref{f2}) and 
1133: (\ref{Vcc})). Thus, the relative error of the experimentally determined value 
1134: of $f(\delta)/c_{13}^2$ is the sum of the relative errors of $A_{ND}$, 
1135: $c_{13}^2$ and of the $\theta_{12}$ - dependent expression in the square 
1136: brackets in eq.~(\ref{And}). The dependence of the latter on the error of 
1137: $\theta_{12}$ is somewhat involved (mainly, because of the $\theta_{12}$ 
1138: dependence of the effective mixing angle in matter $\hat{\theta}_{12}$). 
1139: It can be approximated by 
1140: \begin{equation}
1141: \epsilon_{\theta_{12}}\simeq \left[1.74-\frac{1.83\,[1-(0.4-V_{av}/2\delta)^2]}
1142: {(0.4-V_{av}/2\delta)[1+0.4(0.4-V_{av}/2\delta)]}\right] \Delta \theta_{12}\,,
1143: \label{errtheta}
1144: \end{equation}
1145: where $V_{av}$ is the averaged over the neutrino production region value of 
1146: the matter-induced neutrino potential in the Sun. The values of $V_{av}$ for 
1147: various components of the solar neutrino spectrum can be found in table 1 
1148: of ref.~\cite{deHolanda:2004fd}. 
1149: 
1150: The next question is how the error in the experimentally determined quantity 
1151: $f(\delta)/c_{13}^2$ affects the accuracy of the reconstructed electron 
1152: number density profile $N_e(x)$. Since, in the linear regime, to obtain a 
1153: solution we invoke an iteration procedure, one might expect that the 
1154: corresponding errors in the reconstructed density profile would accumulate 
1155: with increasing iteration order. This is, however, not the case: at each 
1156: iteration,  the relative error of the solution is the same as the relative error 
1157: of $f(\delta)/c_{13}^2$, and the same applies to the exact profile $N_e(x)$. 
1158: This follows from the fact that each iteration profile and the exact solution 
1159: depend linearly on $f(\delta)/c_{13}^2$, which is a consequence of the linearity 
1160: of eq.~(\ref{inv2}). 
1161: 
1162: 
1163: The main contribution to the error of $f(\delta)/c_{13}^2$ (and thus, of the 
1164: reconstructed electron number density profile) is expected to come from the 
1165: error in $A_{ND}$, which is by far the largest one (at the moment, more than 
1166: 100\%). Therefore, an accurate reconstruction of the matter density 
1167: distribution inside the Earth would require very large detectors, capable of 
1168: measuring the solar neutrino day-night effect with an accuracy commensurate 
1169: with the desired accuracy of the matter density reconstruction. 
1170: 
1171: {}From eqs.~(\ref{f1}) and (\ref{omega}) it follows that the errors in the 
1172: parameter $\Delta m_{21}^2$ and in the energy scale of neutrino detectors go 
1173: linearly to the shifts in the reconstructed coordinate.
1174: 
1175: 
1176: \subsection{Finite energy resolution of detectors}
1177: \label{sec:finiteE}
1178: 
1179: Let us consider now the effects of finite energy resolution of neutrino 
1180: detectors. We shall be assuming that neutrinos are detected through a 
1181: charged-current capture reaction, so that the energy of the emitted electron 
1182: in the final state directly gives the energy of the incoming neutrino. The 
1183: electron energy, however, is not exactly measured because of the finite energy 
1184: resolution of the detector, characterized by the resolution function $R(T_e, 
1185: T_e')$. Here $T_e$ and $T_e'$ are the observed and true electron kinetic 
1186: energies, respectively. Because of the one-to-one correspondence between the 
1187: electron and neutrino energies, the resolution function can be written 
1188: directly in terms of the ``observed'' and true neutrino energies.  In our 
1189: discussion it proved more convenient to use $\delta=\Delta m_{21}^2/4E$ 
1190: instead of the neutrino energy $E$, therefore we will be considering the 
1191: detector resolution functions expressed in terms of $\delta$ and $\delta'$. 
1192: In many cases the resolution function can be approximated by a Gaussian 
1193: \begin{equation}
1194: R(\delta, \delta')=\frac{1}{\sqrt{2\pi}\,\sigma(\delta)}\,
1195: \exp\left\{-\frac{(\delta-\delta')^2}
1196: {2\sigma^2(\delta)}\right\}
1197: \label{Gauss1}
1198: \end{equation}
1199: with $\delta$-dependent width, e.g., $\sigma(\delta)=k_0\delta$ or
1200: $\sigma(\delta)=k_0\sqrt{\delta_{max}\delta}$. 
1201: 
1202: When finite energy resolution of detectors is taken into account, the function 
1203: $f(\delta)$ in the expression for the Earth regeneration factor in 
1204: eq.~(\ref{reg}) has to be replaced by 
1205: \begin{equation}
1206: {\cal F}(\delta)=\int_0^\infty R(\delta, \delta') f(\delta')\,d\delta'\,,
1207: \label{smooth}
1208: \end{equation}
1209: which describes the smoothing of $f(\delta)$ with the resolution function. In 
1210: order to proceed with the density profile reconstruction, we have first to extract 
1211: $f(\delta')$ from the experimentally measured quantity ${\cal F}(\delta)$. 
1212: Once this has been done, one can employ the procedures described in sections 
1213: \ref{sec:lin} or \ref{sec:nonlin}. Thus, instead of one-step inverse problem, 
1214: we now have a two-step one. 
1215: 
1216: \subsubsection{Finite energy resolution effects on the Earth regeneration 
1217: factor}
1218: 
1219: Let us first 
1220: discuss the physical effects of the finite energy resolution of detectors  
1221: on the observed Earth matter effect. 
1222: In ref. \cite{Ioannisian:2004jk} 
1223: it was shown that in the simplified case of the box-shaped resolution function 
1224: with the energy width $\Delta E$, eq.~(\ref{f1})  has to be replaced by
1225: \begin{equation}
1226: {\cal F}(\delta)\simeq \int_0^L V(x)\frac{\sin\Delta(\delta)(L-x)}{\Delta(\delta)
1227: (L-x)}\sin\left[2\int\limits_x^{L}\!\omega(x')\,dx'\right]dx \,, 
1228: \label{f1a}  
1229: \end{equation}
1230: where $\Delta(\delta)$ is the resolution width in terms of $\delta$: 
1231: $\Delta(\delta)=\delta \cdot(\Delta E/E)$. Eq.~(\ref{f1a}) differs from 
1232: (\ref{f1}) by the extra factor 
1233: $\sin \Delta(\delta)(L-x)/[\Delta(\delta)(L-x)]$ in the integrand. This factor 
1234: equals unity when $\Delta(\delta)(L-x)=0$ (which corresponds to perfect energy 
1235: resolution or $x=L$), and quickly decreases with increasing $\Delta(\delta)
1236: (L-x)$ (see fig.~\ref{fig:g}). Thus, finite energy resolution of detectors 
1237: leads to an attenuation of the contributions to the Earth matter effect coming 
1238: from the density structures which are far from the detector, the attenuation 
1239: length (distance from the detector) being
1240: \begin{equation}
1241: l_{att}~\simeq~ \frac{1}{\delta}\, \frac{E}{\Delta E}~=~\frac{1}{\pi}\, 
1242: l_{osc}(E)\, \frac{E}{\Delta E}\,.
1243: \label{atten}
1244: \end{equation}
1245: %
1246: {}From eq.~(\ref{atten}) it follows that for the attenuation to be negligible, 
1247: i.e. for $l_{att}$ to exceed considerably $L$ for all energies of interest, 
1248: one needs 
1249: \begin{equation}
1250: \delta_{max} L\,\ll \frac{E}{\Delta E}\,.
1251: \label{cond3}
1252: \end{equation}
1253: %
1254: This can be in conflict with the requirement of good coordinate resolution 
1255: $\delta_{max} L\gg 1$, unless the energy resolution is extremely good. 
1256: %
1257: Thus, finite energy resolution worsens the coordinate resolution of the 
1258: reconstruction procedure. 
1259: If one completely ignores the fact that finite energy resolution of neutrino 
1260: detectors modifies the observed matter effect and simply uses ${\cal F}(\delta)$ 
1261: instead of $f(\delta)$ for the density profile reconstruction, then in the 
1262: case when condition (\ref{cond3}) is not satisfied it is essentially the 
1263: finite energy resolution and not the value of $\delta_{max}$ that determines 
1264: the coordinate resolution of the inversion procedure. Increasing $\delta_{max}$ 
1265: beyond the limit (\ref{cond3}) would not lead to any significant improvement 
1266: of the coordinate resolution in that case. 
1267: 
1268: Eq.~(\ref{f1a}) exhibits a power-law attenuation of the matter density 
1269: structures far from the detector in the case of box-shaped energy 
1270: resolution function. However, in a more realistic case of the Gaussian  
1271: energy resolution, an even stronger (exponential) attenuation results. 
1272: To show that, we first recall that the function ${\cal F}(\delta)$ is measured 
1273: in a finite interval $\delta\in[\delta_{min},\delta_{max}]$. Although the 
1274: integral in eq.~(\ref{smooth}) extends over the whole interval $0\le \delta' 
1275: <\infty$, in fact the contributions to this integral coming from the domains 
1276: of $\delta'$ which are far outside the interval $[\delta_{min},\delta_{max}]$ 
1277: are strongly suppressed because of the presence of the resolution function 
1278: $R(\delta, \delta')$ in the integrand. In particular, if $\delta_{min}$ 
1279: exceeds a few widths of the Gaussian resolution function (\ref{Gauss1}), one 
1280: can formally extend the integration in eq.~(\ref{smooth}) to negative values 
1281: of $\delta'$ without affecting noticeably the value of the integral. 
1282: Extending the integration to the interval $-\infty < \delta' < \infty$, from 
1283: eqs.~(\ref{smooth}), (\ref{Gauss1}) and (\ref{f1}) one readily finds
1284: \begin{equation}
1285: {\cal F}(\delta)\simeq \int_0^L V(x)\,e^{-2(L-x)^2\sigma^2(\delta)}\,\sin
1286: \left[2\int\limits_x^{L}\!\omega(x')\,dx'\right] dx \,. 
1287: \label{f1b}  
1288: \end{equation}
1289: 
1290: 
1291: \subsubsection{Undoing the smoothing}
1292: 
1293: We now turn to the question of how to reconstruct the electron number
1294: density profile of the Earth from the Earth regeneration factor in the
1295: presence of a well-known but finite energy resolution.  As we already
1296: pointed out, if one ignores the fact the observed Earth regeneration
1297: factor is modified by the finite energy resolution of the detector and
1298: naively uses ${\cal F}(\delta)$ instead of $f(\delta)$ for the density
1299: profile reconstruction, the coordinate resolution of the
1300: reconstruction procedure will in general be very poor. We have checked
1301: numerically that for $\delta_{max} L\gtrsim 100$ and the energy
1302: resolution width $\Delta E/E=\Delta(\delta)/\delta$ exceeding 1\%, the
1303: reconstructed profile differs sizeably from the true one. However, if
1304: the resolution function is well known, one can do a much better job:
1305: first, recover the function $f(\delta')$ from the experimentally
1306: measured quantity ${\cal F}(\delta)$, and then reconstruct the Earth
1307: density profile from $f(\delta')$.  In other words, one could to some
1308: extent undo the smoothing of $f(\delta')$ caused by the finite energy
1309: resolution of the neutrino detectors and described by
1310: eq.~(\ref{smooth}). This is similar to the procedures one has to
1311: invoke in various remote sensing problems (e.g., in medical
1312: tomography).
1313: 
1314: In principle, if the resolution function $R(\delta, \delta')$  is exactly 
1315: known and the function ${\cal F}(\delta)$ is precisely determined from the 
1316: experiment, one could expect that the function $f(\delta')$ can be exactly 
1317: found from (\ref{smooth}). This is, however, not the case. The point is 
1318: that eq.~(\ref{smooth}) is a Fredholm integral equation of the first kind,
1319: which belongs to the class of ill-posed problems: small variations in 
1320: ${\cal F}(\delta)$ can lead to very large changes in the reconstructed function 
1321: $f(\delta')$. To illustrate this, let us note that for any integrable 
1322: $R(\delta,\delta')$ and an arbitrary constant $A$
1323: \[
1324: \lim_{k\to\infty}\, \int_0^\infty R(\delta, \delta')\,A \sin(k
1325: \delta')\,d\delta' ~=~0\qquad\quad
1326: \]
1327: by Riemann-Lebesgue lemma. Therefore, even for very large values of $A$,  
1328: changing $f(\delta')\to f(\delta')+ A\sin(k\delta')$ practically does not 
1329: modify the observable ${\cal F}(\delta)$ if $k$ is large enough. This means 
1330: that small variations in ${\cal F}(\delta)$ correspond to very large 
1331: high-frequency changes in the reconstructed function $f(\delta')$. The 
1332: solution does not depend continuously on the data, i.e. is highly unstable. 
1333: 
1334: The function ${\cal F}(\delta)$ is always determined experimentally with some 
1335: errors; the errors are drastically magnified in the process of solving   
1336: eq.~(\ref{smooth}) for $f(\delta')$. Even if one assumes ${\cal F}(\delta)$ to 
1337: be exactly known, the rounding errors, which are inherent to any numerical 
1338: calculation, would play the same role as the experimental errors and 
1339: destabilize the solution. Therefore, straightforward methods of solving 
1340: Fredholm integral equations of the first kind (such as discretization and 
1341: reduction to a system of linear algebraic equations) do not work. One has to 
1342: employ some regularization in order to suppress high-frequency noise in the 
1343: solution. 
1344: 
1345: There exist many regularization approaches for ill-posed problems and
1346: a vast literature on the subject. In our calculations we use two
1347: popular regularization schemes: the truncated singular value
1348: decomposition (TSVD) \cite{TSVD} and the Backus-Gilbert (BG) method
1349: \cite{BG,LE,num}, which are described in Appendices C and D,
1350: respectively.  The results of calculations for the step-function
1351: density profile and the Gaussian detector resolution function
1352: (\ref{Gauss1}) with 10\% resolution ($k_0=0.1$) are presented in
1353: figs.~\ref{fig:TSVD} and \ref{fig:BG}. In these figures shown are the
1354: exact profile (dashed line) and the reconstructed profiles in the case
1355: of Gaussian resolution: the naive calculation, which used ${\cal
1356:   F}(\delta)$ instead of $f(\delta')$, is shown by the dash-dotted
1357: curve, while the results of the corresponding regularization
1358: approaches are shown by the solid curves. In the case of the TSVD
1359: regularization, the contributions of the singular values $\lambda_i\le
1360: 10^{-10}$ were truncated (see Appendix C); in the BG approach,
1361: 1501-point calculation has been used. One can see from the figures
1362: that, while the naive calculation gives very poor reconstruction of
1363: the density profile far from the detector, undoing the smoothing
1364: caused by the finite energy resolution of the detectors improves the
1365: quality of the reconstruction drastically.
1366: 
1367: %
1368: \begin{samepage}
1369: \begin{figure}[tbp]
1370:   \centering 
1371:   \includegraphics[width=.9\linewidth,clip]{Fig-SVD.eps}
1372:   \caption{Density reconstruction for step-function density profile
1373:   using TSVD method. The dashed line is the exact profile, the
1374:   dash-dotted curve corresponds to a naive calculation using ${\cal
1375:   F}(\delta)$ instead of $f(\delta')$ for 10\% Gaussian resolution,
1376:   the solid curve is the result of TSVD regularization for Gaussian
1377:   resolution, truncation at $\lambda_i\le 10^{-10}$.} \label{fig:TSVD}
1378: \end{figure}
1379: %
1380: \begin{figure}[htbp]
1381:   \centering 
1382:   \includegraphics[width=.9\linewidth,clip]{Fig-BG.eps}
1383:   \caption{Same as in fig.~\ref{fig:TSVD}, but for Backus-Gilbert
1384:   method with $N=1501$ points. } \label{fig:BG}
1385: \end{figure}
1386: \end{samepage}
1387: %
1388: 
1389: Comparing figs.~\ref{fig:TSVD} and \ref{fig:BG} one can see that the BG 
1390: method gives in general more accurate reconstruction of the profile in the 
1391: regions that are far from the ends of the neutrino trajectory in the Earth, 
1392: whereas the TSVD method better reproduces the 
1393: exact profile near these endpoints, $x\simeq 0$ and $x\simeq L$, and the 
1394: jumps of the density (at the jumps, the TSVD curve is steeper than that of 
1395: the BG approach). One can employ both these methods of data treating for the 
1396: same set of experimental data and thus combine the advantages of each  
1397: approach. 
1398: 
1399: 
1400: \section{Discussion and outlook}
1401: \label{sec:disc}
1402: 
1403: \vspace*{2mm} We have developed a novel, {\em direct} approach to
1404: neutrino tomography of the Earth, based on a simple analytic formula
1405: for the Earth matter effect on oscillations of solar and supernova
1406: neutrinos inside the Earth.  These neutrinos have a number of
1407: advantages over the accelerator neutrinos from the viewpoint of the
1408: tomography of the Earth. Apart from coming from a free neutrino
1409: source, they are sensitive to the Earth matter effects even if the
1410: distances they travel in the Earth are as short as $\sim 50$ - 100 km.
1411: In addition, they can be used for reconstructing asymmetric density
1412: profiles and thus are sensitive to short scale inhomogeneities of
1413: matter distribution in the Earth, as possibly relevant in prospection
1414: applications.  Neutrino tomography allows the reconstruction of the
1415: electron number density of the Earth, i.e. is sensitive to its
1416: chemical composition. In this respect it can nicely complement
1417: geoseismological studies, which are based on the measurements of the
1418: spatial distributions of seismic wave velocities and can only probe
1419: the total density of the Earth's matter, but not its chemical
1420: composition.
1421: 
1422: 
1423: In the linear regime, which is valid for relatively short distances
1424: traveled by neutrinos in the Earth (up to several hundred km), the
1425: problem amounts to finding an inverse Fourier transform of the Earth
1426: regeneration factor.  However, to perform this transformation, one
1427: needs to know the Earth regeneration factor in the whole infinite
1428: interval of neutrino energies $0\le E < \infty$, whereas
1429: experimentally it can only be measured in a finite range $E_{min}\le
1430: E\le E_{max}$. Non-vanishing $E_{min}$ leads to a finite coordinate
1431: resolution $\Delta x\sim l_{osc}(E_{min})$ of the reconstruction
1432: procedure, while finite $E_{max}$ results in a systematic shift of the
1433: reconstructed profile compared to the true one.
1434: 
1435: Future experiments should be able to detect solar neutrinos with
1436: energies as small as $\sim 100$ keV, which means that a very good
1437: coordinate resolution can in principle be achieved; at the same time,
1438: finite $E_{max}$, i.e.  the lack of knowledge of the Earth
1439: regeneration factor in the high energy domain, could potentially be a
1440: serious drawback for the neutrino tomography of the Earth. We have
1441: shown, however, that this ignorance of the Earth matter effect at high
1442: neutrino energies can be compensated for by making use of simple
1443: iterative procedures. We have developed two such iteration schemes,
1444: one of them not using any prior information about the Earth density
1445: profile, and the other assuming the average matter density along the
1446: neutrino path to be known. Both schemes give very good results, the
1447: second one leading to a faster convergence.
1448: 
1449: In order to reconstruct the Earth matter density profiles over the
1450: distances of geophysical interest, i.e. exceeding a few hundred km,
1451: one needs to solve a complex non-linear problem. It should be noted,
1452: however, that even in the linear regime ($L\lesssim\,$ a few hundred
1453: km) one can obtain very interesting information about the structure of
1454: the Earth's crust and upper mantle. The maximal depth $d$ of the
1455: neutrino trajectory inside the Earth is \begin{equation}
1456: d=R_\oplus-\sqrt{R_\oplus^2-(L/2)^2}\,,
1457: \label{depth} 
1458: \end{equation}
1459: where $R_\oplus=6371$ km is the average radius of the Earth. For $L=300$ km 
1460: this gives $d\simeq 1.8$ km,  which could be quite sufficient, e.g., for 
1461: minerals, oil and gas prospecting. Learning more about the large-scale 
1462: structure of the Earth's interior would require calculations in the non-linear 
1463: regime. 
1464: 
1465: There is a number of very interesting physics and mathematics issues
1466: related to the problem of neutrino tomography of the Earth, which
1467: still are to be explored. We have studied in detail the linear regime
1468: of the Earth density profile reconstruction; attacking the non-linear
1469: regime remains a challenge for future investigations.
1470: %
1471: Another issue is recovering the true Earth regeneration factor,
1472: necessary for the inversion procedure, from the experimentally
1473: measured one, which is smoothed due to the finite energy resolution of
1474: the detector. Similar procedures are invoked in various remote sensing
1475: problems (e.g., in medical tomography, early vision, inverse problems
1476: of geo- and helioseismology, etc.).  We have used two relatively
1477: simple methods of undoing the smoothing effects, the truncated
1478: singular value decomposition and the Backus-Gilbert method, and only
1479: in the linear regime. At the same time, there is a variety of other
1480: methods developed for solving problems of this kind; it would be
1481: interesting to study some of them in application to the neutrino
1482: tomography of the Earth, and also to consider the de-smoothing
1483: procedure in the non-linear regime.
1484: 
1485:         
1486: Due to the rotation of the Earth, for a fixed position of the detector
1487: solar neutrinos will probe the Earth density distribution along a
1488: number of chords, ending at the detectors and spanning a segment
1489: inside the Earth. The size of the segment will depend on the
1490: geographic location of the detector. Because of the relatively low
1491: statistics, only density distributions averaged over certain angular
1492: bins of incoming neutrino directions will be probed. In order to
1493: reduce these averaging effects, i.e. to have small angular bins, one
1494: would need very large detectors and/or large overall detection time.
1495: %
1496: Supernova explosions being one-shot events, the Earth tomography with
1497: supernova neutrinos will allow the density profile reconstruction only
1498: along a single neutrino path for each detector and each supernova,
1499: depending on the detector and supernova positions at the time the
1500: supernova neutrinos arrive at the Earth.
1501: 
1502: 
1503: 
1504: The accuracy of the reconstruction of the Earth matter density
1505: distribution will depend crucially on the accuracy of the experimental
1506: data -- of the measured Earth regeneration factor and of the neutrino
1507: oscillation parameters. It will also depend on the accuracy of the
1508: reconstruction procedure itself. The largest error is expected to come
1509: from the errors in the measured Earth matter effect, mainly because
1510: this effect is rather small for solar and supernova neutrinos. To
1511: achieve a given accuracy of the density profile reconstruction, one
1512: should measure the Earth matter effect with similar or better
1513: accuracy. The present value of the night-day asymmetry of the solar
1514: neutrino signal, measured at Super-Kamiokande, is $2.1 \% \pm 2.0\%
1515: (\rm{stat}) \pm 1.3 \% (\rm{syst})$~\cite{Smy:2003jf}, i.e. the error
1516: constitutes more than 100\%. This is certainly a very important
1517: limiting factor for the reconstruction method described here.  Future
1518: megaton water Cherenkov detectors, such as UNO or Hyper-Kamiokande,
1519: are expected to reduce this error by a factor of 4 to 10; however,
1520: such detectors, unfortunately, are not very suitable for neutrino
1521: tomography of the Earth. The reason is that they are based on the $\nu
1522: e$ scattering process, in which the measured energy of the recoil
1523: electron gives only loose information on the energy of the incoming
1524: neutrino \footnote{In $\nu e$ scattering the neutrino
1525:   energy could be determined if, in addition to the energy of the
1526:   recoil electron, the direction of its momentum were measured.
1527:   Precise determinations of these directions are not possible with
1528:   water Cherenkov detectors, but can be realized in ICARUS-type liquid
1529:   argon detectors \cite{Amerio:2004ze}. }.  Detectors based on
1530: charged-current neutrino capture reactions, such as LENS or MOON
1531: \cite{Raghavan,Ejiri:1999rk,LOWE}, are probably more promising. It is
1532: not clear, however, whether sufficiently large (megaton or perhaps
1533: even tens of megaton scale) detectors of this kind can be constructed,
1534: and much work still has to be done to assess the feasibility of
1535: neutrino tomography of the Earth.  In any case, studying the Earth's
1536: interior with neutrinos is an exciting possibility, which can add yet
1537: another motivation for very large detectors of solar and supernova
1538: neutrinos.
1539: 
1540: 
1541: \vspace*{4mm}
1542: {\em Acknowledgements.} 
1543: The authors are grateful to G. Fiorentini and A. Yu. Smirnov for useful 
1544: discussions. This work was supported by Spanish grant BFM2002-00345 and by 
1545: the European Commission Human Potential Program RTN network 
1546: MRTN-CT-2004-503369. The work of E.A. was partially supported by the 
1547: sabbatical grant No. SAB2002-0069 of the Spanish MECD.   M.A.T.\ is
1548: supported by the M.E.C.D.\ fellowship AP2000-1953.
1549: 
1550: 
1551: 
1552: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1553: %%%%                      Appendix                              %%%%
1554: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1555: \begin{appendix}
1556: 
1557: \section{The case of not very large $\delta_{max} L$}
1558: \label{app:Improved}
1559: 
1560: If $\delta_{max} L\gg 1$ but still not very large, one can develop an
1561: iterative procedure correcting for the corresponding error in the
1562: reconstructed potential.
1563: 
1564: Let us add to and subtract from eq.~(\ref{aux1}) the corresponding integral 
1565: from $\delta_{min}$ to $\delta_0$, where $\delta_0\gg \delta_{max}$ is a 
1566: ``numerical emulation of infinity'', i.e. a large number for which one can 
1567: neglect the difference between $\sin (\delta_0 x)/x$ and $\pi\delta(x)$ (e.g., 
1568: $\delta_{0}=10^3 L^{-1}$). This gives 
1569: \begin{eqnarray}
1570: V(x)\simeq \frac{4}{\pi}\int_{\delta_{min}}^{\delta_{max}} 
1571: f(\delta)\sin 2\delta\,(L-x)\,d\delta\, +\frac{1}{\pi}\int_0^L V(y)
1572: \{F(x,y; 2\delta_{min})\qquad\quad \nonumber \\
1573: +\,F(x,y; 2\delta_{0})-F(x,y; 2\delta_{max})\}\,dy\,.
1574: \label{improved}
1575: \end{eqnarray}
1576: If eq.~(\ref{f2}) is exact, eq.~(\ref{improved}) becomes exact in the limit 
1577: $\delta_0\to\infty$. Note that the first term on the right hand side of 
1578: (\ref{improved}) contains integration over $\delta$ from $\delta_{min}$ to 
1579: $\delta_{max}$, 
1580: and not from $\delta_{min}$ to $\infty$. Thus, it properly takes into account 
1581: that the function $f(\delta)$ is only measured in the energy interval 
1582: $E_{min}\le E\le E_{max}$ with $E_{min}\ne 0$. 
1583: 
1584: Eq.~(\ref{improved}) can now be solved by iterations, the procedure 
1585: being quite analogous to the one used for solving eq.~(\ref{inv2}) in sections 
1586: \ref{sec:iterI} or \ref{sec:iterII}. If $\delta_{max} L$ is a very large 
1587: number, the difference between the solutions of eqs. (\ref{inv2}) and 
1588: (\ref{improved}) is numerically insignificant, but it can be noticeable when 
1589: $\delta_{max} L$ is not too large. In our calculation in the present paper we 
1590: always use $\delta_{max}L\ge 100$, so that there is no need to solve the 
1591: improved equation (\ref{improved}).
1592: 
1593: 
1594: \label{app:derivation}
1595: \renewcommand{\theequation}{\thesection\arabic{equation}}
1596: \setcounter{equation}{0}
1597: \section{Properties of $F(x,y; 2\delta_{min})$}
1598: \label{sec:properties}
1599: 
1600: 
1601: Let us prove that 
1602: \begin{equation}
1603: F(x,y; 2\delta_{min})>0 \quad \mbox{for all}\quad 0 < x,y <L\quad 
1604: \mbox{and}\quad \delta_{min}<2.246 L^{-1}\,. 
1605: \label{Fprop1}
1606: \end{equation}
1607: Indeed, the function $F(x,y; k)$ defined in eq.~(\ref{F1}) can be written as 
1608: $F(x,y; k)=g(x'-y')-g(x'+y')$, where $g(z)=\sin kz/z$, $x'=L-x$, $y'=L-y$. For 
1609: positive $k$, the function $g(z)$  reaches its first minimum at $kz\simeq 
1610: 4.493$ and is a decreasing function of $z$ for $0< kz < 4.493$ (see 
1611: fig.~\ref{fig:g}). Since $|x'-y'|\le x'+y'$, $F(x,y; k)$ is positive for all 
1612: $0<x,y<4.493/k$. Substituting $k=2\delta_{min}$ and requiring 
1613: $0<x,y<L<4.493/2\delta_{min}$, one arrives at (\ref{Fprop1}).
1614: 
1615: 
1616: \section{Truncated singular value decomposition (TSVD) }
1617: \label{sec:TSVD}
1618: 
1619: 
1620: Any square-integrable function $R(\delta, \delta')$ can be represented as 
1621: \cite{TSVD} 
1622: \begin{equation}
1623: R(\delta,\delta')~=~\sum_{i=1}^\infty \lambda_i\,u_i(\delta)\,v_i(\delta')\,.
1624: \qquad\quad
1625: \label{C1}
1626: \end{equation}
1627: where $\lambda_i$ are the so-called singular values, which converge to 0 
1628: as $i\to\infty$, and $u_i(\delta)$ and $v_i(\delta')$ are the singular 
1629: functions. They are orthogonal and can be normalized to satisfy 
1630: \begin{equation}
1631: (u_i\cdot u_j)~\equiv~\int u_i(\delta) u_j(\delta)\,d\delta~=~\delta_{ij}\,,
1632: \label{C2}
1633: \end{equation}
1634: and similarly for $v_i$. There exist standard programs for finding singular 
1635: values and singular functions of square-integrable kernels.
1636: 
1637: {}From eq.~(\ref{smooth}), which can be written in symbolic form as 
1638: ${\cal F}=R\cdot f$, and eqs.~(\ref{C1}) and (\ref{C2}) one finds 
1639: \begin{equation}
1640: f(\delta')~=~\sum_{i=1}^\infty \frac{(u_i\cdot{\cal F})}{\lambda_i}\,
1641: v_i(\delta')\,.\qquad\qquad
1642: \label{TSVDres}
1643: \end{equation}
1644: Since $\lambda_i\to 0$ as $i\to\infty$, the contributions of higher 
1645: harmonics to $f(\delta')$ are strongly enhanced, which leads to instabilities 
1646: due to small variations in ${\cal F}(\delta)$. To suppress these instabilities 
1647: (regularize the solution), one can truncate the series at certain value of 
1648: $i$, i.e. remove the contributions of very small singular values:
1649: \begin{equation}
1650: f(\delta')~\simeq~\sum_{i=1}^N \frac{(u_i\cdot{\cal F})}{\lambda_i}\,
1651: v_i(\delta')\,.\qquad\qquad
1652: \end{equation}
1653: In our calculations we truncate the series when $\lambda_i$ become smaller 
1654: than $10^{-10}$. 
1655: 
1656: A variant of the TSVD method employs a soft truncation, in which in the sum 
1657: in eq.~(\ref{TSVDres}) the following substitution is made: 
1658: \begin{equation}
1659: \frac{1}{\lambda_i}\to\frac{\lambda_i}{\lambda_i^2+\alpha^2}\,,\qquad\qquad
1660: \end{equation}
1661: where $\alpha$ is the regularization parameter. This procedure is equivalent 
1662: to using the Tikhonov regularization \cite{TSVD}. 
1663: 
1664: \section{Backus-Gilbert method }
1665: \label{sec:BG}
1666: 
1667: 
1668: This method \cite{BG,LE,num} is especially well suited for incorporating the 
1669: errors of experimental data into the inversion procedure.  
1670: 
1671: We want to solve eq.~(\ref{smooth}) for the function $f(\delta')$.
1672: First, we take into account that the actual experiments provide us with 
1673: the binned data, i.e. we will have not a function ${\cal F}(\delta)$, but 
1674: a finite set of discrete values ${\cal F}(\delta_i)\equiv{\cal F}_i$ 
1675: \mbox{($i=1,\dots,N$)}. 
1676: {}From eq.~(\ref{smooth}) we then have 
1677: \begin{equation}
1678: {\cal F}_i=\int_0^\infty R(\delta_i, \delta') f(\delta')\,d\delta'\,.
1679: \label{BG1}
1680: \end{equation}
1681: Then, we seek the solution of this equation in the form
1682: \begin{equation}
1683: f(\delta')~\simeq~ \sum_{i=1}^N a_i(\delta'){\cal F}_i\,,
1684: \label{BG2}
1685: \end{equation}
1686: where the coefficients $a_i(\delta')$ are to be determined. This is, actually, 
1687: the most general form for linear inversion. Substituting here ${\cal F}_i$ 
1688: from eq. (\ref{BG1}), we find
1689: \begin{equation}
1690: f(\delta')\simeq \int_0^\infty \hat{\delta}(\delta', \delta'') 
1691: f(\delta'') 
1692: \, d\delta''\,,
1693: \label{BG3}
1694: \end{equation}
1695: where the function $\hat{\delta}(\delta', \delta'')$ is given by
1696: \begin{equation}
1697: \hat{\delta}(\delta', \delta'')~=~\sum_{i=1}^N a_i(\delta') 
1698: R(\delta_i,\delta'')\,.
1699: \label{BG4}
1700: \end{equation}
1701: {}From eq. (\ref{BG3}) it is seen that in the ideal case the function 
1702: $\hat{\delta}(\delta', \delta'')$ should coincide with Dirac's 
1703: $\delta$-function. In reality it does not, but we can try to make it as close 
1704: to the $\delta$-function as possible. The simplest Backus-Gilbert prescription 
1705: for that is the following. Define the spread function $r(\delta')$ as
1706: \begin{equation}
1707: r(\delta')~=~ \int_0^\infty (\delta'-\delta'')^2 \,[\hat{\delta}(\delta', 
1708: \delta'')]^2 \, d\delta''\,. 
1709: \label{BG5}
1710: \end{equation}
1711: It is non-negative, and would vanish if the function $\hat{\delta}(\delta', 
1712: \delta'')$ were indeed Dirac's $\delta$-function. We now find the coefficients 
1713: $a_i(\delta')$ by minimizing $r(\delta')$, subject to the normalization 
1714: constraint 
1715: \begin{equation}
1716: \int_0^\infty \hat{\delta}(\delta', \delta'') \, d\delta'' ~=~ 1\,.
1717: \label{BG6}
1718: \end{equation}
1719: This would lead to $\hat{\delta}(\delta', \delta'')$ being closest to the 
1720: $\delta$-function for a given choice of the spread\footnote{Other choices of 
1721: the spread function are also possible.}. 
1722: The derivation is straightforward, and here we give the results.
1723: 
1724: Let us define the matrix $S(\delta')$ as 
1725: \begin{equation}
1726: S_{ik}(\delta') ~=~ \int_0^\infty (\delta'-\delta'')^2 \, R(\delta_i, 
1727: \delta'')\,R(\delta_k, \delta'') \, d\delta''\,. 
1728: \label{BG7}
1729: \end{equation}
1730: We also define $N$ numbers $u_i$:
1731: \begin{equation}
1732: u_i ~=~\int_0^\infty R(\delta_i, \delta_1)\,d\delta_1\,.
1733: \label{BG8}
1734: \end{equation}
1735: \noindent
1736: If the energy resolution function $R(\delta, \delta')$ is normalized 
1737: to unit integral, then all $u_i$ in eq.~(\ref{BG8}) are equal to 1, but 
1738: in general $u_i$ can be different from unity. Minimization of the function 
1739: $r(\delta')$, subject to the normalization constraint (\ref{BG6}), yields
1740: \begin{equation}
1741: a_i(\delta')~=~
1742: \frac{\sum\limits_{k=1}^N [S^{-1}(\delta')]_{ik}\,u_k}
1743: {\sum\limits_{j,k=1}^N u_j\,[S^{-1}(\delta')]_{jk}\, u_k}\,,
1744: \label{BG9}
1745: \end{equation}
1746: or, in matrix notation,
1747: \begin{equation}
1748: a(\delta')~=~\frac{S^{-1}(\delta')\, u}{u\, S^{-1}(\delta')\,u}\,.
1749: \label{BG10}
1750: \end{equation}
1751: Note that the matrix $S$ is symmetric and positive-definite, so it has an 
1752: inverse.
1753: 
1754: Using eq. (\ref{BG2}), one then finds the approximate solution of 
1755: eq.~(\ref{BG1}):
1756: \begin{equation}
1757: f(\delta')~\simeq~
1758: \frac{\sum\limits_{i,k=1}^N {\cal F}_i\,[S^{-1}(\delta')]_{ik}\,u_k}
1759: {\sum\limits_{j,k=1}^N u_j\,[S^{-1}(\delta')]_{jk}\, u_k}\,,
1760: \label{BG11}
1761: \end{equation}
1762: or, using the matrix notation,
1763: \begin{equation}
1764: f(\delta)'~\simeq~\frac{{\cal F}\, S^{-1}(\delta')\, 
1765: u}{u\, S^{-1}(\delta')\, u}\,.
1766: \label{BG12}
1767: \end{equation}
1768: 
1769: Some comments are in order. To have a good accuracy, one would like the 
1770: number of data bins $N$ to be large, but then $S(\delta')$ is a very large 
1771: matrix, and inverting it may be a time and memory consuming operation. It 
1772: is, however, not necessary to invert $S(\delta')$. It is enough to solve the 
1773: system of $N$ linear equations
1774: \begin{equation}
1775: \sum_{k=1}^N S_{ik}(\delta')y_k(\delta')=u_i
1776: \label{BG13}
1777: \end{equation}
1778: and find an $N$-component vector $y(\delta')$, which is a much simpler 
1779: problem. Then eqs.~(\ref{BG11}) and (\ref{BG12}) can be rewritten as
1780: \begin{equation}
1781: f(\delta')~\simeq~ \frac{\sum\limits_{i=1}^N {\cal F}_i\,y_i(\delta')}
1782: {\sum\limits_{k=1}^N u_k\, y_k(\delta')}~=~
1783: \frac{{\cal F} y(\delta')}{u\, y(\delta')}\,.
1784: \label{BG14}
1785: \end{equation}
1786: 
1787: 
1788: \end{appendix}
1789: 
1790: %\newpage
1791: 
1792: \begin{thebibliography}{999}
1793: 
1794: \bibitem{Volkova} L. V. Volkova, G. T. Zatsepin, Bull. Acad. Sci. 
1795: USSR, Phys. Ser. {\bf 38} (1974) 151. 
1796: 
1797: \bibitem{DeRujula:1983ya}
1798: A.~De Rujula, S.~L.~Glashow, R.~R.~Wilson and G.~Charpak,
1799: %``Neutrino Exploration Of The Earth,''
1800: Phys.\ Rept.\  {\bf 99} (1983) 341.
1801: %%CITATION = PRPLC,99,341;%%
1802: 
1803: \bibitem{Wilson:1983an}
1804: T.~L.~Wilson,
1805: %``Neutrino Tomography: Tevatron Mapping Versus The Neutrino Sky,''
1806: Nature {\bf 309} (1984) 38.
1807: %%CITATION = NATUA,309,38;%%
1808: 
1809: \bibitem{Askarian:ca}
1810: G.~A.~Askarian,
1811: %``Investigation Of The Earth By Means Of Neutrinos. Neutrino Geology,''
1812: Sov.\ Phys.\ Usp.\  {\bf 27} (1984) 896
1813: [Usp.\ Fiz.\ Nauk {\bf 144} (1984) 523].
1814: %%CITATION = SOPUA,27,896;%%
1815: 
1816: \bibitem{Borisov:sm}
1817: A.~B.~Borisov, B.~A.~Dolgoshein and A.~N.~Kalinovsky,
1818: %``Direct Method For Determination Of Differential Distribution Of The Earth
1819: %Density By Means Of High-Energy Neutrino Scattering. (In Russian),''
1820: Yad.\ Fiz.\  {\bf 44} (1986) 681.
1821: %%CITATION = YAFIA,44,681;%%
1822: 
1823: \bibitem{Jain:1999kp}
1824: P.~Jain, J.~P.~Ralston and G.~M.~Frichter,
1825: %``Neutrino absorption tomography of the Earth's interior using isotropic
1826: %ultra-high energy flux,''
1827: Astropart.\ Phys.\  {\bf 12} (1999) 193
1828: [hep-ph/9902206].
1829: %%CITATION = HEP-PH 9902206;%%
1830: 
1831: \bibitem{Ermilova:pw}
1832: V.~K.~Ermilova, V.~A.~Tsarev and V.~A.~Chechin,
1833: %``Restoration Of The Density Distribution Of Material Based On Neutrino
1834: %Oscillations,''
1835: %\href{http://www.slac.stanford.edu/spires/find/hep/www?irn=2007819}{SPIRES 
1836: %entry}
1837: Bull. Lebedev Phys. Inst. {\bf 3} (1988) 51.
1838: 
1839: \bibitem{Chechin} V.A. Chechin and  V.K. Ermilova, in: Proc. of LEWI'90 
1840: School, Dubna, 1991, p. 75. 
1841: 
1842: \bibitem{Nicolaidis:fe}
1843: A.~Nicolaidis,
1844: %``Neutrinos For Geophysics,''
1845: Phys.\ Lett.\ B {\bf 200} (1988) 553.
1846: %%CITATION = PHLTA,B200,553;%%
1847: 
1848: \bibitem{Nicolaidis:1990jm}
1849: A.~Nicolaidis, M.~Jannane and A.~Tarantola,
1850: %``Neutrino Tomography Of The Earth,''
1851: J.\ Geophys.\ Res.\  {\bf 96} (1991) 21811.
1852: %%CITATION = JGREA,96,21811;%%
1853: 
1854: \bibitem{Ohlsson:2001ck}
1855: T.~Ohlsson and W.~Winter,
1856: %``Reconstruction of the earth's matter density profile using a single neutrino
1857: %baseline,''
1858: Phys.\ Lett.\ B {\bf 512} (2001) 357
1859: [hep-ph/0105293].
1860: %%CITATION = HEP-PH 0105293;%%
1861: 
1862: \bibitem{Ohlsson:2001fy}
1863: T.~Ohlsson and W.~Winter,
1864: %``Could one find petroleum using neutrino oscillations in matter?,''
1865: Europhys.\ Lett.\  {\bf 60} (2002) 34
1866: [hep-ph/0111247].
1867: %%CITATION = HEP-PH 0111247;%%
1868: 
1869: \bibitem{Ioannisian:2002yj}
1870: A.~N.~Ioannisian and A.~Y.~Smirnov,
1871: %``Matter effects of thin layers: Detecting oil by oscillations of solar
1872: %neutrinos,''
1873: hep-ph/0201012.
1874: %%CITATION = HEP-PH 0201012;%%
1875: 
1876: \bibitem{Lindner:2002wm}
1877: M.~Lindner, T.~Ohlsson, R.~Tomas and W.~Winter,
1878: %``Tomography of the earth's core using supernova neutrinos,''
1879: Astropart.\ Phys.\  {\bf 19} (2003) 755
1880: [hep-ph/0207238].
1881: %%CITATION = HEP-PH 0207238;%%
1882: 
1883: \bibitem{Winter:2005we}
1884: W.~Winter,
1885: %``Probing the absolute density of the Earth's core using a neutrino beam,''
1886: hep-ph/0502097.
1887: %%CITATION = HEP-PH 0502097;%%
1888: 
1889: \bibitem{Wolfenstein:1977ue}
1890: L.~Wolfenstein,
1891: %``Neutrino Oscillations In Matter,''
1892: Phys.\ Rev.\ D {\bf 17} (1978) 2369.
1893: %%CITATION = PHRVA,D17,2369;%%
1894: 
1895: \bibitem{Mikheev:gs}
1896: S.~P.~Mikheev and A.~Y.~Smirnov,
1897: %``Resonance Enhancement Of Oscillations In Matter And Solar Neutrino
1898: %Spectroscopy,''
1899: Sov.\ J.\ Nucl.\ Phys.\  {\bf 42} (1985) 913
1900: [Yad.\ Fiz.\  {\bf 42} (1985) 1441].
1901: %%CITATION = SJNCA,42,913;%%
1902: 
1903: \bibitem{Akhmedov:2000cs}
1904: E.~K.~Akhmedov,
1905: %``Matter effects in oscillations of neutrinos traveling short distances  
1906: %in matter,''
1907: Phys.\ Lett.\ B {\bf 503} (2001) 133
1908: [hep-ph/0011136].
1909: %%CITATION = HEP-PH 0011136;%%
1910: 
1911: \bibitem{Ioannisian:2004jk}
1912: A.~N.~Ioannisian and A.~Y.~Smirnov,
1913: %``Neutrino oscillations in low density medium,''
1914: Phys.\ Rev.\ Lett.\  {\bf 93} (2004) 241801 [hep-ph/0404060].
1915: %%CITATION = HEP-PH 0404060;%%
1916: 
1917: \bibitem{deHolanda:2004fd}
1918: P.~C.~de Holanda, W.~Liao and A.~Y.~Smirnov,
1919: %``Toward precision measurements in solar neutrinos,''
1920: Nucl.\ Phys.\ B {\bf 702} (2004) 307 [hep-ph/0404042].
1921: %%CITATION = HEP-PH 0404042;%%
1922: 
1923: \bibitem{Akhmedov:2001kd}
1924: E.~K.~Akhmedov, P.~Huber, M.~Lindner and T.~Ohlsson,
1925: %``T violation in neutrino oscillations in matter,''
1926: Nucl.\ Phys.\ B {\bf 608} (2001) 394
1927: [hep-ph/0105029].
1928: %%CITATION = HEP-PH 0105029;%%
1929: %\bibitem{}
1930: 
1931: \bibitem{Blennow:2003xw}
1932: M.~Blennow, T.~Ohlsson and H.~Snellman,
1933: \newblock Phys. Rev. {\bf D69} (2004) 073006 [hep-ph/0311098].
1934: %%CITATION = HEP-PH 0311098;%%
1935: 
1936: \bibitem{Akhmedov:2004rq}
1937: E.~K.~Akhmedov, M.~A.~T\'ortola and J.~W.~F.~Valle,
1938: %``A simple analytic three-flavour description of the day-night effect in the
1939: %solar neutrino flux,''
1940: JHEP {\bf 0405} (2004) 057 [hep-ph/0404083].
1941: %%CITATION = HEP-PH 0404083;%%
1942: 
1943: \bibitem{solardata} See, e.g., talks by C. Cattadori, M. Nakahata and J. 
1944: Wilkerson at the 21st International Conference on Neutrino Physics and 
1945: Astrophysics (Neutrino 2004), Coll\`ege de France, Paris, June 14 - 19, 
1946: 2004 (transparencies at http://neutrino2004.in2p3.fr/). 
1947: 
1948: \bibitem{KamLAND} 
1949: %\cite{Eguchi:2002dm}
1950: %\bibitem{Eguchi:2002dm}
1951: KamLAND Collaboration, K.~Eguchi {\it et al.},   
1952: %``First results from KamLAND: Evidence for reactor anti-neutrino
1953: %disappearance,''
1954: Phys.\ Rev.\ Lett.\  {\bf 90} (2003) 021802 [hep-ex/0212021]; 
1955: %%CITATION = HEP-EX 0212021;%%
1956: %
1957: %\bibitem{Araki:2004mb}
1958: T.~Araki {\it et al.},  
1959: %[KamLAND Collaboration],
1960: %``Measurement of neutrino oscillation with KamLAND: Evidence of spectral
1961: %distortion,''
1962: hep-ex/0406035.
1963: %%CITATION = HEP-EX 0406035;%%
1964: 
1965: 
1966: \bibitem{Maltoni:2004ei}
1967: M.~Maltoni, T.~Schwetz, M.~A.~T\'ortola and J.~W.~F.~Valle,
1968: %``Status of global fits to neutrino oscillations,''
1969: New J.\ Phys.\ {\bf 6} (2004) 122 [hep-ph/0405172, version 4, 
1970: which includes an appendix with updated results taking into
1971: account a new background source in KamLAND data].
1972: %%CITATION = HEP-PH 0405172;%%
1973: 
1974: 
1975: \bibitem{Dighe:1999id}
1976: A.~S. Dighe, Q.~Y. Liu and A.~Y. Smirnov,
1977: \newblock hep-ph/9903329.
1978: %%CITATION = HEP-PH 9903329;%%
1979: 
1980: \bibitem{deGouvea:2000un}
1981: A.~de Gouv\^ea,
1982: %``The oscillation probability of GeV solar neutrinos of all active species,''
1983: Phys.\ Rev.\ D {\bf 63} (2001) 093003 [hep-ph/0006157].
1984: %%CITATION = HEP-PH 0006157;%%
1985: 
1986: \bibitem{Raghavan} R. Raghavan, {\em Discovery potential of low energy  
1987: solar neutrino experiments},\\
1988: http://www.hep.utexas.edu/sno/apsstudy/LONUDISC.pdf.
1989: 
1990: %\cite{Ejiri:1999rk}
1991: \bibitem{Ejiri:1999rk}
1992: H.~Ejiri {\it et al.,}
1993: %J.~Engel, R.~Hazama, P.~Krastev, N.~Kudomi and R.~G.~H.~Robertson,
1994: %``Spectroscopy of double-beta and inverse-beta decays from Mo-100 for
1995: %neutrinos,''
1996: Phys.\ Rev.\ Lett.\  {\bf 85} (2000) 2917 [nucl-ex/9911008].
1997: %%CITATION = NUCL-EX 9911008;%%
1998: 
1999: \bibitem{LOWE} M. Nakahata, talk given at 5th Workshop on "Neutrino 
2000: Oscillations and their Origin" (NOON2004), Tokyo, Japan, Febr. 11-15, 
2001: 2004. Transparencies at http://www-sk.icrr.u-tokyo.ac.jp/noon2004/.
2002: 
2003: \bibitem{Keil:2002in}
2004: See, e.g., M.~T.~Keil, G.~G.~Raffelt and H.~T.~Janka,
2005: %``Monte Carlo study of supernova neutrino spectra formation,''
2006: Astrophys.\ J.\  {\bf 590} (2003) 971 [astro-ph/0208035].
2007: %%CITATION = ASTRO-PH 0208035;%%
2008: 
2009: \bibitem{Dziewonski:xy}
2010: A.~M.~Dziewonski and D.~L.~Anderson,
2011: %``Preliminary Reference Earth Model,''
2012: Phys.\ Earth Planet \ Interiors {\bf 25} (1981) 297.
2013: %%CITATION = PEPIA,25,297;%%
2014: 
2015: 
2016: \bibitem{Goursat} E. Goursat, {\em A course in mathematical analysis}, 
2017: New York, Dover, 1959, v. III, \mbox{part 2}, $\S$103.  
2018: 
2019: \bibitem{TSVD} See, e.g., P.~C.~Hansen, Inverse Problems {\bf 8} (1992) 849. 
2020: 
2021: \bibitem{BG} G. Backus and F. Gilbert,
2022: %The resolving power of gross earth data,
2023: Geophys. J. R. Astron. Soc. {\bf 16} (1968) 169; 
2024: %Uniquenss in the inversion of inaccurate gross earth data,
2025: Philos. Trans. R. Soc. London {\bf 266} (1970) 123.
2026: 
2027: \bibitem{LE} T. J. Loredo and R. I. Epstein,
2028: %Analyzing gamma-ray bursts spectral data,
2029: Astrophys. J. {\bf 336} (1989) 896.
2030: 
2031: \bibitem{num}
2032: W.~H~. Press, S.~A.~Teukolsky, W.~T.~Vetterling and B.~P.~Flannery,  
2033: {\em Numerical recipes in C++: the art of scientific computing}, 2nd ed.,
2034: Cambridge University Press, Cambridge, 2002.
2035: 
2036: \bibitem{Smy:2003jf}
2037: Super-Kamiokande Collaboration, M.~B. Smy {\em et~al.},
2038: \newblock Phys. Rev. {\bf D69} (2004) 011104.
2039: %[hep-ex/0309011].
2040: %%CITATION = HEP-EX 0309011;%%
2041: 
2042: %\cite{Amerio:2004ze}
2043: \bibitem{Amerio:2004ze}
2044: ICARUS Collaboration, S.~Amerio {\it et al.},   
2045: %``Design, construction and tests of the ICARUS T600 detector,''
2046: Nucl.\ Instrum.\ Meth.\ A {\bf 527} (2004) 329.
2047: %%CITATION = NUIMA,A527,329;%%
2048: 
2049: \end{thebibliography}
2050: 
2051: \end{document}
2052: 
2053: