1: \documentclass[12pt,preprint]{aastex}
2: %\documentclass[onecolumn]{emulateapj}
3: % \documentclass[numberedappendix]{emulateapj}
4: \usepackage{rotating}
5: % \usepackage{amsmath}
6: \usepackage{bm}
7:
8: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
9: % New commands
10: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
11:
12: \newcommand{\threej}[6]
13: { \left(\begin{array}{ccc}
14: #1\\
15: #4
16: \end{array}\right) }
17:
18: % \newcommand{\threej}[6]
19: % {\begin{pmatrix}
20: % #1\\
21: % #4
22: % \end{pmatrix}}
23:
24: \newcommand{\sixj}[6]
25: { \left\{\begin{array}{ccc}
26: #1\\
27: #4
28: \end{array}\right\} }
29:
30: % \newcommand{\sixj}[6]
31: % {\begin{Bmatrix}
32: % #1\\
33: % #4
34: % \end{Bmatrix}}
35:
36:
37: \begin{document}
38: \title{Advanced Forward Modeling and Inversion of
39: Stokes Profiles Resulting from the Joint Action of the Hanle and Zeeman Effects}
40:
41: \author{A. Asensio Ramos, J. Trujillo Bueno\altaffilmark{1}}
42: \affil{Instituto de Astrof\'{\i}sica de Canarias, V\'\i a L\'actea s/n, E-38205 La Laguna, Tenerife, Spain}
43: \and
44: \author{E. Landi Degl'Innocenti}
45: \affil{Universit\`a degli Studi di Firenze,
46: Dipartimento di Astronomia e Scienza dello Spazio, Largo Enrico Fermi 2, I-50125 Florence, Italy}
47: \altaffiltext{1}{Consejo Superior de Investigaciones Cient\'{\i}ficas (Spain)}
48: \email{aasensio@iac.es,jtb@iac.es,landie@arcetri.astro.it}
49:
50: \begin{abstract}
51: A big challenge in solar and stellar physics in the coming years will be
52: to
53: decipher the magnetism of the solar outer atmosphere (chromosphere and corona)
54: along with
55: its dynamic coupling with the magnetic fields of the underlying photosphere. To
56: this end, it
57: is important to develop rigorous diagnostic tools for the physical
58: interpretation of
59: spectropolarimetric observations in suitably chosen spectral lines. Here we
60: present a
61: computer program for the synthesis and inversion of Stokes profiles caused by
62: the joint
63: action of atomic level polarization and the Hanle and Zeeman effects in some
64: spectral lines of diagnostic interest, such
65: as those of the He {\sc i} 10830 \AA\ and 5876 \AA\ (or D$_3$) multiplets. It is
66: based
67: on the quantum theory of spectral line polarization, which takes into account in
68: a rigorous
69: way all the relevant physical mechanisms and ingredients
70: (optical pumping, atomic level polarization, level crossings and repulsions,
71: Zeeman, Paschen-Back and Hanle effects). The influence of radiative transfer on
72: the emergent spectral line radiation is taken into account through a suitable
73: slab model. The user can either calculate the emergent intensity and polarization
74: for any given magnetic field vector or infer the dynamical and magnetic
75: properties from the observed Stokes profiles via an efficient inversion
76: algorithm based on
77: global optimization methods. The reliability of the forward modeling and
78: inversion code presented here is
79: demonstrated through several applications, which range from the inference of the
80: magnetic field vector in solar active regions to determining whether or not it is
81: canopy-like in quiet chromospheric regions. This user-friendly diagnostic tool
82: called ``HAZEL" (from HAnle and ZEeman Light) is offered to
83: the astrophysical community, with the hope that it will facilitate new advances in solar and
84: stellar physics.
85: \end{abstract}
86:
87: \keywords{magnetic fields --- polarization --- radiative transfer --- scattering
88: --- Sun: chromosphere --- methods: data analysis, numerical}
89:
90: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
91: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
92: % INTRODUCTION
93: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
94: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
95: \section{Introduction}
96:
97: The present paper describes a computer program for the synthesis and inversion
98: of Stokes profiles resulting from the joint
99: action of the Hanle and Zeeman effects in some spectral lines of diagnostic
100: interest, such as those pertaining to the \ion{He}{1}
101: 10830 \AA\ and 5876 \AA\ (or D$_3$) multiplets. The effects of radiative
102: transfer on the emergent spectral
103: line radiation are taken into account through a suitable slab model.
104: Our aim is to provide the solar and stellar physics communities with a robust
105: but user-friendly tool for understanding
106: and interpreting spectropolarimetric observations, with the hope that this will
107: facilitate new advances in solar and stellar
108: physics.
109:
110: In particular, the lines of neutral helium at 10830 \AA\ are of great interest
111: for empirical investigations of the
112: dynamic and magnetic properties of plasma structures in the solar chromosphere
113: and corona, such as active
114: regions \citep[e.g.,][]{harvey_hall71,ruedi96,Lagg04,centeno06}, filaments
115: \citep[e.g.,][]{lin98,trujillo_nature02}, prominences \citep[e.g.,][]{merenda06}
116: and
117: spicules \citep[e.g.,][]{trujillo_merenda05,socas_elmore05}. The same applies to
118: the lines of the
119: \ion{He}{1} D$_3$ multiplet at 5876 \AA\, which have been used for investigating
120: the magnetic field vector in solar
121: prominences and spicules
122: \citep[e.g.,][]{landi_d3_82,querfeld85,bommier94,casini03,lopezariste_casini05,ramelli06_2,
123: ramelli06}
124:
125: Such helium lines result from transitions between terms of the triplet system of
126: helium (ortho-helium), whose
127: respective $J$-levels (with $J$ the level's total angular momentum)
128: are far less populated than the ground level of helium (that is, than the
129: singlet level $^1$S$_0$), except
130: perhaps in flaring regions. On the other
131: hand, the lower term (2s$^3$S$_1$) of the \ion{He}{1} 10830 \AA\ multiplet is the
132: ground level of ortho-helium, while
133: its upper term (2p$^3$P$_{2,1,0}$) is the lower one of 5876 \AA\ (whose upper term
134: is 3d$^3$D$_{3,2,1}$). Therefore, the
135: significant difference in the ensuing optical thicknesses of the observed solar
136: plasma structure implies that when
137: the radiation in these spectral lines is observed on the solar disk it is much
138: easier to see structures in
139: 10830 \AA\ than in 5876 \AA, while both lines are clearly seen in emission when
140: observing off-limb structures
141: such as prominences and spicules. The additional fact that the Hanle effect in
142: forward scattering creates
143: measurable linear polarization signals in the lines of the \ion{He}{1} 10830
144: \AA\ multiplet when the magnetic
145: field is inclined with respect to the local solar vertical direction
146: \citep{trujillo_nature02}, and that
147: there is a nearby photospheric line of Si {\sc i}, makes the 10830 \AA\ spectral
148: region very suitable for
149: investigating the coupling between the photosphere and the corona.
150:
151: While the Stokes $I$ profiles of the 10830 \AA\ and 5876 \AA\ helium lines
152: depend mainly on the
153: distribution of the populations of their respective upper ($J_u$) and lower
154: ($J_l$) levels along the
155: line-of-sight (LOS), their Stokes $Q$, $U$ and $V$ profiles depend on the
156: strengths
157: and wavelength positions of the $\pi$ ($\Delta{M}=M_u-M_l=0$),
158: $\sigma_{\rm blue}$ ($\Delta{M}=+1$)
159: and $\sigma_{\rm
160: red}$ ($\Delta{M}=-1$) transitions, which can only be calculated correctly
161: within the framework of the
162: Paschen-Back effect theory. Moreover, the $Q$, $U$ and $V$ profiles are also
163: affected by the atomic level polarization
164: induced by anisotropic pumping
165: processes \citep[e.g.,][]{landi_landolfi04}. An atomic level of
166: total angular
167: momentum $J$ is said to be polarized when its magnetic sublevels are unequally
168: populated and/or when there
169: are quantum coherences between them. The radiative transitions induced by the
170: anisotropic illumination of the helium atoms in the solar atmosphere
171: are able to create a significant amount of atomic polarization in the helium
172: levels, even in the metastable
173: lower level of the \ion{He}{1} 10830 \AA\ multiplet
174: \citep{trujillo_nature02}.
175: If the net circular polarization of the incident radiation at the wavelengths of
176: the helium transitions is
177: negligible, as it uses to be the case,
178: the radiatively induced atomic level polarization is such that the
179: populations of substates with different values of $|M|$ are different (non-zero
180: atomic alignment), while substates
181: with magnetic quantum numbers $M$ and $-M$ are equally populated (zero atomic
182: orientation). On the other
183: hand, elastic collisions with the neutral hydrogen atoms of the solar
184: chromospheric and coronal structures
185: are unable to destroy the atomic polarization of the He {\sc i} levels.
186: As a result, even in the
187: absence of magnetic fields,
188: linearly polarized spectral
189: line radiation would be produced, simply because the population imbalances among
190: the magnetic sublevels imply
191: more or fewer $\pi$-transitions, per unit volume and time, than $\sigma$
192: transitions. The atomic polarization
193: of the upper level of the line transition under consideration is thus
194: responsible of a {\em selective emission}
195: of polarization components, while that of the lower level may give rise to a
196: {\em selective absorption} of
197: polarization components (``zero-field'' dichroism). In order for this type of
198: dichroism to produce a measurable
199: contribution to the emergent linear polarization it is necessary to have a
200: substantial line-center optical
201: thickness along the LOS or, assuming a small but non-negligible optical
202: thickness, that the plasma structure under
203: consideration is observed against the bright background of the solar disk
204: \citep{trujillo_nature02,trujillo_asensio07}. Therefore, the observable effects
205: of dichroism are easier to detect in \ion{He}{1} 10830 \AA\ than in 5876 \AA.
206:
207: In the presence of a magnetic field the emergent polarization changes because of
208: the following two reasons.
209: First, because a magnetic field modifies the atomic level polarization, not only
210: by producing the Hanle-effect
211: relaxation of the quantum coherences pertaining to each individual $J$-level,
212: but also through possible
213: interferences between the magnetic sublevels pertaining to different $J$-levels,
214: which give rise to a variety of
215: remarkable effects such as the transfer of atomic alignment to atomic
216: orientation in the $J$-levels of the
217: upper term of the \ion{He}{1} D$_3$ multiplet \citep{landi_d3_82} or the
218: enhancement of the scattering
219: polarization in the D$_2$ line of \ion{Na}{1} by a vertical magnetic field
220: \citep{trujillo_casini02}.
221: Second, because the magnetic
222: splitting of the atomic energy levels give rise to significant wavelength shifts
223: between the $\pi$ and
224: $\sigma$ transitions (as compared with the spectral line width) and,
225: consequently, to the generation of
226: measurable linear and/or circular polarization (i.e., the familiar transverse
227: and longitudinal Zeeman effects,
228: respectively). Obviously, a correct modeling of the spectral line polarization
229: that results from the joint
230: action of the Hanle and Zeeman effects requires
231: the application of the
232: quantum theory of spectral line polarization \citep[see the monograph by][]{landi_landolfi04}, as done by several researchers for interpreting spectropolarimetric observations of solar plasma structures in the \ion{He}{1} D$_3$ multiplet \citep[e.g.,][]{landi_d3_82,bommier94,casini03,lopezariste_casini05} and in the \ion{He}{1} 10830 \AA\ triplet \citep[e.g.,][]{trujillo_nature02,trujillo_merenda05,merenda06}.
233:
234: Over the last few years new computer programs for the synthesis and inversion of Stokes profiles induced by the joint action of atomic level polarization and the Hanle
235: and Zeeman effects have been developed and applied to the interpretation of spectropolarimetric observations.
236: For example, \cite{landi_landolfi04} have developed some forward modeling codes with which they have calculated several Hanle effect diagrams and theoretical Stokes profiles of lines from complex atomic models, while the computer program of
237: \cite{casini_manso05} can treat even the case of a hyperfine structured multiplet taking into account the quantum interferences between the $F$ levels belonging to the $J$ levels of different terms. Concerning Stokes inversion techniques we should
238: mention that \cite{casini03} applied an inversion code based on principal component analysis \citep{arturo_casini02} to spectropolarimetric observations of solar prominences in the \ion{He}{1} D$_3$ multiplet, providing two-dimensional maps of the magnetic field vector and further evidence for strengths significantly larger than average. On the other hand,
239: \cite{merenda06} opted for an inversion strategy for the \ion{He}{1} 10830 \AA\ multiplet
240: in which the longitudinal component of the magnetic field
241: vector is obtained from the measured Stokes $V$ profiles (which are dominated by
242: the Zeeman effect for the lines of the \ion{He}{1} 10830 \AA\ multiplet), and its
243: orientation from the observed Stokes $Q$ and $U$ profiles (which in solar prominences are due to the presence of atomic level
244: polarization). In the just mentioned computer programs and
245: applications the optically thin
246: assumption was used, which was a suitable approximation for the particular
247: prominences observed, but an
248: unreliable one in general (and especially for the interpretation of observations
249: in the \ion{He}{1} 10830 \AA\ multiplet).
250:
251: A simple but suitable model for taking into account radiative transfer effects is the constant-property
252: slab model used by \cite{trujillo_nature02,trujillo_merenda05}
253: for the interpretation of spectropolarimetric observations in solar
254: filaments and spicules,
255: which has been recently extended by \cite{trujillo_asensio07} to include
256: magneto-optical effects.
257: These authors have applied this ``cloud" model for the interpretation of
258: spectropolarimetric observations in order to point out that the atomic
259: polarization of the helium levels
260: may have an important impact on the emergent linear polarization of the
261: \ion{He}{1} 10830 \AA\ multiplet, even
262: for magnetic field strengths as large as 1000 G. Therefore, inversion codes that
263: neglect the influence of
264: atomic level polarization, such as the Milne-Eddington codes of \cite{Lagg04} and \cite{socas_trujillo04},
265: should ideally be used
266: only for the inversion of Stokes profiles emerging from strongly magnetized
267: regions (with $B>1000$ G), or
268: when the observed Stokes $Q$ and $U$ profiles turn out to be dominated by the
269: transverse Zeeman effect
270: \citep[e.g., as it happens with some active regions filaments as a result of the
271: particular illumination conditions explained in][]{trujillo_asensio07}.
272:
273: It is also necessary to mention that \cite{manso_trujillo03a} developed a general
274: radiative transfer computer program for solving the so-called non-LTE problem of the
275: 2nd kind --that is, the multilevel scattering polarization problem including the Hanle
276: effect of a weak magnetic field. These authors considered the multilevel atom
277: approximation (which neglects quantum interferences between the sublevels
278: of {\em different} $J$-levels), but took fully into account the effects of radiative
279: transfer in realistic atmospheric models and the role of elastic and inelastic
280: collisions in addition to all the relevant optical pumping mechanisms. An interesting
281: application using a semi-empirical model of the solar atmosphere can be found in
282: \cite{manso_trujillo03b}. The recent advances in the development of efficient numerical
283: methods for the solution of non-LTE polarization transfer
284: problems \citep[e.g., the review by][]{trujillo03} and in computer technology
285: make now possible even the performing of three-dimensional radiative transfer
286: simulations of the Hanle effect in convective atmospheres \citep[e.g.,][]{trujillo_shchukina07}.
287:
288: The previous introductory paragraphs strongly suggest that it would be of great
289: interest to develop
290: a robust but user-friendly computer program for the synthesis and inversion of
291: Stokes profiles resulting from
292: the joint action of atomic level polarization and the Hanle and Zeeman
293: effects. We have done this
294: by implementing an efficient global optimization method for the solution of the
295: inversion problem and
296: by calculating at each iterative step the emergent spectral line polarization
297: through the solution of
298: the Stokes-vector transfer equation in a slab of constant physical properties in
299: which the
300: radiatively-induced atomic level polarization is assumed to be dominated by the
301: photospheric continuum
302: radiation. At each point of the observed field of view the slab's optical
303: thickness is chosen to fit
304: the Stokes $I$ profile, which is a strategy that accounts implicitly for the
305: true physical mechanisms
306: that populate the triplet levels of \ion{He}{1}
307: \citep[e.g., the photoionization-recombination mechanism, as shown
308: by][]{centeno08b}. The observed Stokes $Q$, $U$ and $V$ profiles are then used
309: to infer the magnetic field vector along with a few extra physical quantities.
310:
311: The outline of this paper is as follows. The formulation of the problem is
312: presented in \S2, where
313: we review the relevant equations
314: within the framework of the quantum theory of spectral line polarization. The
315: forward modelling option of
316: our computer program is described in \S3, including some interesting
317: examples of possible
318: applications. \S4 deals with a detailed description of the Stokes
319: inversion option, while \S5
320: considers a variety of applications aimed at an in-depth testing of this
321: diagnostic tool. The
322: important issue of the possible ambiguities and degeneracies is addressed in
323: \S6, with emphasis
324: on the Van Vleck ambiguity and on the possibility of inferring the atmospheric
325: height at which the
326: observed on-disk plasma structure is located. Finally, in \S7 we summarize
327: the main conclusions and
328: comment on some ongoing developments.
329:
330: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
331: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
332: % BASIC EQUATIONS
333: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
334: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
335: \section{Formulation of the problem}
336:
337: We consider a constant-property slab of \ion{He}{1} atoms, located at a height
338: $h$ above the
339: visible solar ``surface", in the presence of a deterministic magnetic field of
340: arbitrary strength $B$,
341: inclination $\theta_B$ and azimuth $\chi_B$ (see Fig. 1). The slab's optical
342: thickness at the wavelength
343: and line of sight under consideration is $\tau$.
344: We assume that
345: all the \ion{He}{1} atoms inside this slab are illuminated from below by the
346: photospheric solar continuum radiation field, whose center-to-limb variation has
347: been tabulated by \cite{pierce00}. The ensuing anisotropic radiation pumping
348: produces population imbalances and quantum coherences between pairs of magnetic
349: sublevels, even among those pertaining to the different $J$-levels of the
350: adopted \ion{He}{1}
351: atomic model. This atomic level polarization and the Zeeman-induced
352: wavelength shifts between the $\pi$ ($\Delta{M}=M_u-M_l=0$), $\sigma_{\rm blue}$
353: ($\Delta{M}=+1$) and $\sigma_{\rm red}$ ($\Delta{M}=-1$) transitions produce
354: polarization in the emergent spectral line radiation.
355:
356: In order to facilitate the reading and understanding of this paper, in
357: the following subsections we summarize the basic equations which allow us to
358: calculate the spectral line polarization taking rigorously into account the
359: joint action of the Hanle and Zeeman effects. To this end, we have applied the
360: quantum theory of spectral line polarization, which is described in great detail
361: in the monograph by \cite{landi_landolfi04}. We have also applied several methods
362: of solution of the Stokes-vector transfer equation, some of which can be
363: considered as particular cases of the two general methods explained in \S6 of \cite{trujillo03}.
364:
365: \subsection{The radiative transfer approach}
366: \label{sec:radiative_transfer}
367: The emergent Stokes vector $\mathbf{I}(\nu,\mathbf{\Omega})=(I,Q,U,V)^{\dag}$
368: (with $\dag$=transpose, $\nu$ the frequency and $\mathbf{\Omega}$ the unit vector indicating the direction of propagation of the ray) is obtained by solving the radiative transfer equation
369:
370: \begin{equation}
371: \frac{d}{ds}\mathbf{I}(\nu,\mathbf{\Omega}) =
372: \bm{\epsilon}(\nu,\mathbf{\Omega}) - \mathbf{K}(\nu,\mathbf{\Omega})
373: \mathbf{I}(\nu,\mathbf{\Omega}),
374: \label{eq:rad_transfer}
375: \end{equation}
376: where $s$ is the geometrical distance along the ray under consideration,
377: $\bm{\epsilon}(\nu,\mathbf{\Omega})=({\epsilon}_I,{\epsilon}_Q,{\epsilon
378: }_U,{\epsilon}_V)^{\dag}$ is the emission vector and
379: \begin{equation}
380: \mathbf{K} = \left( \begin{array}{cccc}
381: \eta_I & \eta_Q & \eta_U & \eta_V \\
382: \eta_Q & \eta_I & \rho_V & -\rho_U \\
383: \eta_U & -\rho_V & \eta_I & \rho_Q \\
384: \eta_V & \rho_U & -\rho_Q & \eta_I
385: \end{array} \right)
386: \label{eq:propagation}
387: \end{equation}
388: is the propagation matrix. Alternatively, introducing the optical distance along the ray,
389: ${\rm d}{\tau}=-{\eta_I}{\rm d}s$, one can write the Stokes-vector transfer Eq. (\ref{eq:rad_transfer}) in the following two ways:
390:
391: \begin{itemize}
392:
393: \item The first one, whose formal solution requires the use of the evolution operator introduced by \cite{landi_landi_85}, is
394: \begin{equation}
395: {{d}\over{d{\tau}}}{\bf I}\,=\,{\bf K}^{*}
396: {\bf I}\,-\,{\bf S},
397: \label{eq:rad_transfer_peo}
398: \end{equation}
399: where ${\bf K}^{*}={\bf K}/{\eta_I}$ and ${\bf S}=\bm{\epsilon}/{\eta_I}$.
400: The formal solution of this equation can be seen in eq. (23) of \cite{trujillo03}.
401:
402: \item The second one, whose formal solution does not require the use of the above-mentioned evolution operator is \citep[e.g.,][]{rees89}
403: \begin{equation}
404: {{d}\over{d{\tau}}}{\bf I}\,=\,{\bf I}\,-\,{\bf S}_{\rm eff},
405: \label{eq:rad_transfer_delo}
406: \end{equation}
407: where the effective source-function vector
408: $\,{\bf S}_{\rm eff}\,=\,{\bf S}\,-\,
409: {\bf K}^{'}{\bf I},\,\,\,$ being $\,{\bf K}^{'}={\bf K}^{*}-{\bf 1}$
410: (with $\bf 1$ the unit matrix). The formal solution of this equation can be seen in eq. (26) of \cite{trujillo03}.
411:
412: \end{itemize}
413:
414: Once
415: the coefficients $\epsilon_I$ and $\epsilon_X$ (with
416: $X=Q,U,V$) of the emission vector
417: and the coefficients $\eta_I$, $\eta_X$, and
418: $\rho_X$ of the $4\times4$ propagation matrix are known
419: at each point within the medium it is possible to solve formally Eq.
420: (\ref{eq:rad_transfer_peo}) or Eq.
421: (\ref{eq:rad_transfer_delo}) for
422: obtaining the emergent Stokes profiles for any desired line of sight.
423: Our computer program considers the following levels of sophistication for the solution of the radiative transfer equation:
424:
425: \begin{itemize}
426:
427: \item {\em Numerical Solutions}.
428: The most general case, where the properties of the slab vary
429: along the ray path, has to be solved numerically. To this
430: end, two efficient and accurate methods of solution of
431: the Stokes-vector transfer equation are those proposed by \cite{trujillo03} (see his eqs. (24) and (27), respectively). The starting points for the development of these two numerical methods were Eq. (\ref{eq:rad_transfer_peo}) and Eq. (\ref{eq:rad_transfer_delo}), respectively. Both methods can be considered as generalizations, to the Stokes-vector transfer case, of the well-known short characteristics method for the solution of the standard (scalar) transfer equation.
432:
433: \item {\em Exact analytical solution of the problem of a constant-property slab including the magneto-optical terms of the propagation matrix}. For the general case of a constant-property slab of arbitrary optical thickness we actually have the following analytical solution, which can be easily obtained as a particular case of eq. (24) of \cite{trujillo03}:
434:
435: \begin{equation}
436: {\bf I}={\rm e}^{-{\mathbf{K}^{*}}\tau}\,{\bf I}_{\rm sun}\,+\,\left[{\mathbf{K}^{*}}\right]^{-1}\,
437: \left( \mathbf{1} - {\rm e}^{-{\mathbf{K}^{*}}\tau} \right) \,\mathbf{S},
438: \label{eq:slab_peo}
439: \end{equation}
440: where $\mathbf{I}_{\rm sun}$ is the Stokes
441: vector that illuminates the slab's boundary that is most distant from the
442: observer. We point out that the exponential of the propagation
443: matrix ${\mathbf{K}^{*}}$ has an analytical expression similar to eq. (8.23) in \cite{landi_landolfi04}.
444:
445: \item {\em Approximate analytical solution of the problem of a constant-property slab including the magneto-optical terms of the propagation matrix}. An approximate analytical solution to the constant-property slab problem can be easily obtained as a particular case of eq. (27) of \cite{trujillo03}:
446:
447: \begin{equation}
448: \mathbf{I} = \left[ \mathbf{1}+\Psi_0 \mathbf{K}' \right]^{-1} \left[ \left(
449: e^{-\tau} \mathbf{1} - \Psi_M \mathbf{K}' \right) \mathbf{I}_{\rm sun} +
450: (\Psi_M+\Psi_0) \mathbf{S} \right],
451: \label{eq:slab_delo}
452: \end{equation}
453: where the coefficients $\Psi_M$ and $\Psi_0$ depend only on the optical thickness of the slab at the frequency and line-of-sight under consideration, since their expressions are:
454: \begin{eqnarray}
455: \Psi_M&=& \frac{1-e^{-\tau}}{\tau} - e^{-\tau},\nonumber \\
456: \Psi_0 &=&1-\frac{1-e^{-\tau}}{\tau}.
457: \end{eqnarray}
458:
459: Note that Eq. (\ref{eq:slab_delo}) for the emergent Stokes vector is the one used by \cite{trujillo_asensio07} for
460: investigating the impact of atomic level polarization on the Stokes profiles of the He {\sc i} 10830 \AA\ multiplet.
461: We point out that, strictly speaking, it can be considered only as the exact analytical solution of the optically-thin
462: constant-property slab problem\footnote{More precisely, when the optical thickness of the slab is small in comparison
463: with the eigenvalues of the matrix $\mathbf{K}'$.}. The reason why Eq. (\ref{eq:slab_delo}) is, in general, an approximate
464: expression for calculating the
465: emergent Stokes vector is because
466: its derivation assumes that the Stokes vector within the slab varies linearly with the optical distance. However, it provides
467: a fairly good approximation to the emergent Stokes profiles (at least for all the problems we have investigated in this paper).
468: Moreover, the results of fig. 2 of \cite{trujillo_asensio07} remain also virtually the same when using instead the exact
469: Eq. (\ref{eq:slab_peo}), which from a computational viewpoint is significantly less efficient than the approximate Eq. (\ref{eq:slab_delo}).
470:
471: \item {\em Exact analytical solution of the problem of a constant-property slab when neglecting the second-order terms of the Stokes-vector transfer equation}. Simplified expressions for the emergent Stokes vector can be obtained when
472: $\epsilon_I{\gg}\epsilon_X$ and $\eta_I{\gg}(\eta_X,\rho_X)$, which justifies to neglect the second-order terms of Eq. (\ref{eq:rad_transfer}). The resulting approximate formulae for the emergent Stokes parameters are given by eqs. (9) and (10) of \cite{trujillo_asensio07}, which are identical to those used by \cite{trujillo_merenda05} for modeling the Stokes profiles observed in solar chromospheric spicules. We point out that there is a typing error in the sentence that introduces such eqs. (9) and (10) in \cite{trujillo_asensio07}, since they are obtained only when the above-mentioned second-order terms are neglected in Eq. (\ref{eq:rad_transfer}), although it is true that there are no magneto-optical terms in the resulting equations.
473:
474: \item {\em Optically thin limit}. Finally, the most simple solution
475: is obtained when taking the optically thin limit ($\tau{\ll}1$) in the equations reported in the previous point, which lead to the equations (11) and (12) of \cite{trujillo_asensio07}. Note that if $\mathbf{I}_{\rm sun}=0$ (i.e., $I_0=X_0=0$), then such optically thin equations imply that ${X/I}\,{\approx}\,{\epsilon_X}/{\epsilon_I}$.
476:
477: \end{itemize}
478:
479: The coefficients of the emission vector and of the propagation matrix
480: depend on the multipolar components, $\rho^K_Q(J,J^{'})$, of the atomic density
481: matrix. Let us recall now the meaning of these physical quantities and how to
482: calculate them in the presence of an arbitrary magnetic field under given
483: illumination conditions.
484:
485: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
486: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
487: \subsection{The multipolar components of the atomic density matrix}
488:
489: We quantify the atomic polarization of the \ion{He}{1} levels using the multipolar components of the atomic density matrix. The
490: \ion{He}{1} atom can be correctly described under
491: the framework of the $L$-$S$ coupling \citep[e.g.,][]{condon_shortley35}. As
492: illustrated in Fig. 2, the
493: different $J$-levels are grouped in terms with well defined values of the
494: electronic
495: angular momentum $L$ and the spin $S$. Since the $^4$He atoms are devoid of
496: nuclear
497: angular momentum, we do not have to consider hyperfine structure\footnote{See
498: \cite*{belluzzi07} for
499: the formulation and solution of an interesting problem where hyperfine structure
500: is important.}. The energy
501: separation between the $J$-levels pertaining to
502: each term is very small in comparison with the energy difference between
503: different terms. For example, the energy separation between the $J=3$ and
504: $J=2$ levels of the term 3d$^3$D (the upper term of the D$_3$ multiplet) is of
505: the order of
506: $0.0025\,{\rm cm}^{-1}$, which is $\sim$2$\times$10$^{5}$ times smaller than
507: the separation between the 3d$^3$D and 3p$^3$P terms. On the other hand,
508: although the energy separations between the $J$-levels of the upper terms of the
509: 10830 \AA\ and D$_3$ multiplets are much larger than their natural widths, such
510: $J$-levels suffer
511: crossings and repulsions for the typical magnetic strengths encountered in the
512: solar atmospheric plasma (e.g., the $J=2$ and $J=1$ levels of the upper term of
513: the \ion{He}{1} 10830 \AA\ multiplet cross for magnetic strengths between 400 G
514: and 1600 G, while the $J=3$ and $J=2$ levels of the upper term of the D$_3$
515: multiplet cross
516: for strengths between 10 G and 100 G, approximately). This can be seen clearly in Fig. \ref{fig:splitting}.
517: Therefore, it turns out to be fundamental to allow for coherences between
518: different
519: $J$-levels pertaining to the same term but not between the $J$-levels
520: pertaining to
521: different terms. As a result, we can represent the atom under the
522: formalism of the multi-term atom discussed by \cite{landi_landolfi04}.
523:
524: In the absence of magnetic fields the energy eigenvectors can be written using
525: Dirac's notation as $|\beta L S J M\rangle$, where
526: $\beta$ indicates a set of inner quantum numbers specifying the electronic
527: configuration. In general, if a magnetic field of
528: arbitrary strength is present, the vectors $|\beta L S J M\rangle$ are no longer
529: eigenfunctions of the total Hamiltonian and $J$ is no longer a good quantum
530: number. In this
531: case, the eigenfunctions of the full Hamiltonian can be written as the following
532: linear combination:
533: \begin{equation}
534: \label{eq:eigenfunctions_total_hamiltonian}
535: |\beta L S j M\rangle = \sum_J C_J^j(\beta L S, M) |\beta L S J M\rangle,
536: \end{equation}
537: where $j$ is a pseudo-quantum number which is used for labeling the energy
538: eigenstates belonging to the subspace corresponding to assigned values of the
539: quantum numbers $\beta$, $L$, $S$, and $M$, and where the coefficients $C_J^j$
540: can be chosen to be real.
541:
542: In the presence of a magnetic field sufficiently weak so that the magnetic
543: energy is much smaller than the energy intervals between the $J$-levels, the energy eigenvectors are still
544: of the form $|\beta L S J M\rangle$ ($C_J^j(\beta L S, M) \approx \delta_{Jj}$), and the
545: splitting of the magnetic sublevels pertaining to each $J$-level is linear with the magnetic field strength. For stronger magnetic fields, we enter the incomplete Paschen-Back effect regime in which the energy eigenvectors are
546: of the general form given by Eq. (\ref{eq:eigenfunctions_total_hamiltonian}),
547: and the splitting among the various $M$-sublevels is no longer linear with the
548: magnetic strength. This regime is reached for magnetic strengths of the order of
549: 10 G for the \ion{He}{1} D$_3$ multiplet and of the order of 100 G for the 10830
550: \AA\ multiplet (see Fig. \ref{fig:splitting}). If the magnetic field strength is further increased we
551: eventually reach the so-called complete Paschen-Back effect regime, where the
552: energy eigenvectors are of the form $|L S M_L M_S\rangle$ and each $L$-$S$ term
553: splits into a number of components, each of which corresponding to particular
554: values of ($M_L+2M_S$).
555: %Figure 3 shows the energies of the magnetic sublevels as
556: %a function of the magnetic field strength, for both the upper term of the 10830 \AA\
557: %multiplet (left panels) and the upper term of the D$_3$
558: %multiplet (right panels).
559:
560: Within the framework of the multi-term atom model the atomic polarization of the
561: energy levels is described with the
562: aid of the density matrix elements
563: \begin{equation}
564: \rho^{\beta L S}(jM,j'M') = \langle \beta L S j M | \rho | \beta L S j' M'\rangle,
565: \end{equation}
566: where $\rho$ is the atomic density matrix operator. Using the expression of the
567: eigenfunctions of the
568: total Hamiltonian given by Eq. (\ref{eq:eigenfunctions_total_hamiltonian}), the
569: density matrix
570: elements can be rewritten as:
571: \begin{equation}
572: \rho^{\beta L S}(jM,j'M') = \sum_{JJ'} C_J^j(\beta L S, M) C_{J'}^{j'}(\beta L
573: S, M') \rho^{\beta L S}(JM,J'M'),
574: \end{equation}
575: where $\rho^{\beta L S}(JM,J'M')$ are the density matrix elements on the basis of
576: the eigenvectors $| \beta L S J M\rangle$.
577:
578: Following \cite{landi_landolfi04}, it is helpful to use the spherical
579: statistical tensor
580: representation, which is related to the previous one by the following linear
581: combination:
582: \begin{eqnarray}
583: {^{\beta LS}\rho^K_Q(J,J')} &=& \sum_{jj'MM'} C_J^j(\beta L S, M)
584: C_{J'}^{j'}(\beta L S, M') \nonumber \\
585: &\times& (-1)^{J-M} \sqrt{2K+1} \threej{J}{J'}{K}{M}{-M'}{-Q}
586: \rho^{\beta L S}(jM,j'M'),
587: \end{eqnarray}
588: where the 3-j symbol is defined as indicated by any
589: suitable textbook on Racah algebra.
590: This alternative representation has some advantages. Firstly, the well-known
591: results
592: obtained when atomic polarization effects are disregarded are easily
593: recovered by
594: considering only the elements of the density matrix with $K=0$ and $Q=0$.
595: Secondly,
596: the transformation law under rotations is much simpler because it involves only
597: one rotation matrix. Finally, each $\rho^K_Q$ element
598: has a clear physical interpretation: the $\rho^2_Q$ elements (with $Q=0,{\pm}1, {\pm}2$) are
599: called the atomic alignment components, while the $\rho^1_Q$ elements (with $Q=0,{\pm}1$) are
600: the atomic orientation components. The
601: ${^{\beta LS}\rho^K_Q(J,J')}$ elements are, in general, complex quantities
602: (except for the elements with
603: $Q=0$ and $J=J'$, that are real quantities), so that, taking into account that
604: the density matrix is an hermitian operator, the number of
605: real quantities required to describe the
606: atomic polarization properties of a given $L$-$S$ term is $(2S+1)^2(2L+1)^2$. This makes a total of 405 real quantities to describe the atomic polarization of the 5-term model atom shown in Fig. \ref{fig:helium_atom}.
607:
608: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
609: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
610: \subsection{Statistical equilibrium equations}
611: In order to obtain the
612: ${^{\beta LS}\rho^K_Q(J,J')}$ elements we have to solve the
613: statistical equilibrium equations. These equations, written in a reference
614: system in which
615: the quantization axis ($Z$)
616: is directed along the
617: magnetic field vector and
618: neglecting the
619: influence of collisions, can be written as \citep{landi_landolfi04}:
620: \begin{eqnarray}
621: \frac{d}{dt} {^{\beta LS}\rho^K_Q(J,J')} &=& -2\pi \mathrm{i} \sum_{K' Q'}
622: \sum_{J'' J'''} N_{\beta LS}(KQJJ',K'Q'J''J''') {^{\beta LS}\rho^{K'}_{Q'}(J'',J''')}
623: \nonumber \\
624: &+& \sum_{\beta_\ell L_\ell K_\ell Q_\ell J_\ell J_\ell'} {^{\beta_\ell L_\ell
625: S}\rho^{K_\ell}_{Q_\ell}(J_\ell,J_\ell')}
626: \mathbb{T}_A(\beta L S K Q J J', \beta_\ell L_\ell S K_\ell Q_\ell J_\ell
627: J_\ell') \nonumber \\
628: &+& \sum_{\beta_u L_u K_u Q_u J_u J_u'} {^{\beta_u L_u
629: S}\rho^{K_u}_{Q_u}(J_u,J_u')}
630: \Big[ \mathbb{T}_E(\beta L S K Q J J', \beta_u L_u S K_u Q_u J_u J_u') \nonumber \\
631: & &\qquad \qquad \qquad \qquad \qquad + \mathbb{T}_S(\beta L S K Q
632: J J', \beta_u L_u S K_u Q_u J_u J_u') \Big] \nonumber \\
633: &-& \sum_{K' Q' J'' J'''} {^{\beta L S}\rho^{K'}_{Q'}(J'',J''') } \Big[
634: \mathbb{R}_A(\beta L S K Q J J' K' Q' J'' J''') \nonumber \\
635: & & + \mathbb{R}_E(\beta L S K Q J J' K'
636: Q' J'' J''') + \mathbb{R}_S(\beta L S K Q J J' K' Q' J'' J''') \Big].
637: \label{eq:see}
638: \end{eqnarray}
639: The first term in the right hand side of Eq. (\ref{eq:see}) takes into account
640: the
641: influence of the magnetic field on the atomic level polarization. This term has
642: its simplest expression in the chosen magnetic field
643: reference frame \citep[see eq. 7.41 of][]{landi_landolfi04}.
644: In any other reference system, a more complicated expression
645: arises.
646: The second, third and fourth terms account, respectively, for coherence transfer due
647: to
648: absorption from lower levels ($\mathbb{T}_A$), spontaneous emission from upper
649: levels
650: ($\mathbb{T}_E$) and stimulated emission from upper levels ($\mathbb{T}_S$).
651: The remaining terms account for the relaxation of coherences due to absorption to
652: upper
653: levels ($\mathbb{R}_A$), spontaneous emission to lower levels ($\mathbb{R}_E$)
654: and stimulated emission to lower levels ($\mathbb{R}_S$), respectively.
655:
656: The stimulated emission and absorption transfer and relaxation rates depend explicitly on
657: the radiation field properties \citep[see eqs. 7.45 and 7.46 of][]{landi_landolfi04}.
658: The symmetry properties of the
659: radiation
660: field are accounted for by the spherical components of the radiation field
661: tensor:
662:
663: \begin{equation}
664: J^K_Q(\nu) = \oint \frac{d\Omega}{4\pi} \sum_{i=0}^3
665: \mathcal{T}^K_Q(i,\mathbf{\Omega}) S_i(\nu,\mathbf{\Omega}).
666: \label{eq:jkq}
667: \end{equation}
668: The quantities $\mathcal{T}^K_Q(i,\mathbf{\Omega})$ are spherical tensors that
669: depend
670: on the reference frame and on the
671: ray direction $\mathbf{\Omega}$. They are given by
672: \begin{equation}
673: \mathcal{T}^K_Q(i,\mathbf{\Omega}) = \sum_P t^K_P(i) \mathcal{D}^K_{PQ}(R'),
674: \label{eq:tkq}
675: \end{equation}
676: where $R'$ is the rotation that carries the reference system defined by
677: the line-of-sight $\mathbf{\Omega}$ and by the polarization unit vectors $\mathbf{e}_1$ and
678: $\mathbf{e}_2$ into the reference system of the magnetic field,
679: while $\mathcal{D}^K_{PQ}(R')$
680: is the usual rotation matrix \citep[e.g.,][]{edmonds60}.
681: Table 5.6 in \cite{landi_landolfi04} gives the
682: $\mathcal{T}^K_Q(i,\mathbf{\Omega})$ values for each Stokes parameter $S_i$ (with $S_0=I$, $S_1=Q$, $S_2=U$ and $S_3=V$).
683:
684: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
685: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
686: \subsection{Emission and absorption coefficients}
687: Once the multipolar components ${^{\beta L S}\rho^{K}_{Q}(J,J') }$ are known, the
688: coefficients $\epsilon_I$ and $\epsilon_X$ (with
689: $X=Q,U,V$) of the emission vector
690: and the coefficients $\eta_I$, $\eta_X$, and
691: $\rho_X$ of the propagation matrix
692: for a given transition between an upper
693: term $(\beta L_u S)$ and an lower term $(\beta L_\ell S)$ can be
694: calculated with the expressions of \S7.6.b in
695: \cite{landi_landolfi04}.
696: These radiative transfer coefficients are proportional to the number density of \ion{He}{1} atoms, $\mathcal{N}$. Their defining expressions contain also the Voigt profile and the Faraday-Voigt profile \citep[see \S5.4 in][]{landi_landolfi04}, which involve the following parameters: $a$ (i.e., the reduced damping constant), $v_\mathrm{th}$ (i.e., the velocity that characterizes the thermal motions, which
697: broaden the line profiles), and $v_\mathrm{mac}$ (i.e., the velocity of possible bulk motions in the plasma, which produce a Doppler shift).
698:
699: It is important to emphasize that the expressions for
700: the emission and absorption coefficients and those of the statistical
701: equilibrium equations are written in the reference system whose quantization
702: axis is parallel to the magnetic field. The following equation indicates how to
703: obtain the density matrix elements in a new reference system:
704: \begin{equation}
705: \left[ {^{\beta L S}\rho^{K}_{Q}(J,J') } \right]_\mathrm{new} = \sum_{Q'} \left[
706: {^{\beta L S}\rho^{K}_{Q'}(J,J') } \right]_\mathrm{old}
707: \mathcal{D}^K_{Q' Q}(R)^*,
708: \end{equation}
709: where $\mathcal{D}^K_{Q' Q}(R)^*$ is the complex conjugate of the rotation matrix for the rotation $R$ that carries the
710: old reference system into the new one.
711:
712:
713: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
714: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
715:
716: \subsection{The atomic model}
717:
718: The atomic model we have used in our calculations
719: includes the following five terms of the
720: triplet system of neutral helium: 2s$^3$S, 3s$^3$S, 2p$^3$P, 3s$^3$P and 3d$^3$D
721: (see Fig. \ref{fig:helium_atom}).
722: It has been concluded that the inclusion of these five terms is sufficient for a
723: reliable calculation of the atomic polarization that the anisotropic pumping of
724: the photospheric continuum radiation produces in the lower and upper $J$-levels
725: of the D$_3$ multiplet \citep{bommier80}. Since the only level (with $J=1$) of
726: the lower term 2s$^3$S is metastable, the adopted atomic model should be also
727: satisfactory for the 10830 \AA\ multiplet. In order to check this
728: we have compared the results obtained using two different atomic models. The
729: first one is that of
730: Fig. \ref{fig:helium_atom}. The second one is
731: a simplified version in which only the 2s$^3$S, 2p$^3$P and 3d$^3$D terms are taken into account.
732: We have verified, for a large number of possible configurations, that
733: differences
734: in the resulting values of the multipolar components of the atomic density
735: matrix
736: are never larger than $\sim$5\%
737: for the terms involved in the 10830 \AA\ transitions.
738: Although the selection rules
739: allow six transitions among the 5 terms, we have only included the four
740: transitions indicated in Fig. \ref{fig:helium_atom} (i.e.,
741: we have neglected the influence of the infrared transitions 3d$^3$D-3p$^3$P
742: and 3p$^3$P-3s$^3$S). The Einstein
743: coefficients for the four included transitions shown in Table
744: \ref{tab:tab_einstein} and the energy of the levels shown in Fig.
745: \ref{fig:helium_atom}
746: have been obtained from the NIST
747: database\footnote{\texttt{http://physics.nist.gov/PhysRefData/ASD/index.html}}
748: \citep[see also][]{wiese_nist66,drake_helium98}. Table \ref{tab:tab_einstein} also indicates the
749: value of the critical magnetic field strength for the operation of the
750: upper-level Hanle effect, which results from equating the Zeeman splitting of
751: the level with its natural width:
752:
753: \begin{equation}
754: B_\mathrm{critical} \approx 1.137 \times 10^{-7} / (t_\mathrm{life} g_L),
755: \label{eq:critical_field}
756: \end{equation}
757: where $t_\mathrm{life}$ is the level's lifetime in seconds, $g_L$ is its Land\'e factor and
758: $B_\mathrm{critical}$ is given in gauss.
759: Note that for obtaining the $B_\mathrm{critical}^\mathrm{upper}$ values of Table
760: \ref{tab:tab_einstein} we have used $t_\mathrm{life}{\approx}1/A_{ul}$, while
761: an estimation of the critical magnetic field strength for the operation of
762: the lower-level Hanle effect requires using
763: $t_\mathrm{life} \approx 1/(B_{lu}J^0_0)$ (which for the metastable lower-level
764: of the He {\sc i} 10830 \AA\ multiplet gives
765: $B_\mathrm{critical}^\mathrm{lower} \approx 0.1$ G).
766:
767: We point out that in our modeling we are not explicitly taking into account the
768: radiative mechanism that is thought to be responsible of the overpopulation of
769: the triplet levels of \ion{He}{1} required to
770: produce the
771: absorption or emission features observed in the spectral lines of such two
772: multiplets --that
773: is, ionizations from the singlet states of \ion{He}{1} caused by EUV ionizing
774: radiation coming downwards from the corona followed by recombinations towards
775: both the singlet and triplet states \citep[e.g.,][]{avrett94}. The fact that
776: most of the ionizations take place from the singlet states suggests that the
777: atomic polarization of the triplet states should be little affected by such EUV
778: coronal irradiation. This expectation is reinforced by the fact that the number
779: of photoionizations
780: per unit volume and time from the triplet levels of \ion{He}{1} is way smaller
781: than the number of bound-bound
782: transitions \citep[see Fig. 7 in][]{centeno08b}, which suggests that the atomic
783: polarization of the $J$-levels
784: of the 10830 \AA\ and D$_3$ multiplets is indeed dominated by optical pumping in
785: the line transitions
786: themselves. Following our approach, the key mechanism responsible of the
787: absorption or emission observed
788: in such lines of neutral helium, be it the above-mentioned
789: photoionization-recombination mechanism
790: and/or collisional excitation in regions with T$>20000$ K
791: \citep[cf.,][]{andretta_jones97}, is
792: unimportant in our forward modeling and inversion approach. The reason is that
793: they are
794: implicitly accounted for via the definition of the optical thickness of the slab
795: which, being
796: a free parameter in our modeling, is used to fit the observed intensity
797: profiles. This point has been clarified by \cite{centeno08b}.
798:
799: \subsection{The incident radiation field}
800:
801: As mentioned before, we consider a slab of \ion{He}{1} atoms
802: anisotropically
803: illuminated from below by the photospheric continuum radiation, which we assume
804: to have axial symmetry around the solar local vertical direction. Since the
805: illumination conditions are assumed to be known a priori, the radiation field
806: tensors can be calculated
807: directly from the given incident radiation.
808:
809: For symmetry reasons, it is advantageous to calculate the statistical tensors
810: of the radiation field in a reference frame in which the $Z$-axis is along the
811: vertical of the atmosphere. Since the incoming radiation is assumed to be
812: unpolarized, the only
813: non-zero spherical tensor components of the radiation field are $J^0_0$ and
814: $J^2_0$. They
815: quantify the mean intensity and the ``degree of anisotropy'' of the radiation
816: field, respectively. For the case of our plane-parallel slab model, their
817: expressions
818: reduce to:
819: \begin{eqnarray}
820: \label{eq:j00_j20}
821: J^0_0 &=& \frac{1}{2} \int_{-1}^{1} I(\mu) \mathrm{d}\mu \\
822: J^2_0 &=& \frac{1}{2\sqrt{2}} \int_{-1}^{1} (3\mu^2-1) I(\mu) \mathrm{d}\mu,
823: \end{eqnarray}
824: where $\mu=\cos \theta$ is the cosine of the heliocentric angle $\theta$.
825:
826: It is convenient to parameterize the radiation
827: field
828: in
829: terms of the number of photons per mode $\bar n$ and the anisotropy factor $w$:
830: \begin{equation}
831: \bar n = \frac{c^2}{2 h \nu^3} J^0_0, \qquad w = \sqrt{2} \frac{J^2_0}{J^0_0},
832: \label{eq:nbar_omega}
833: \end{equation}
834: where $c$ is the light speed, $h$ is the Planck constant and $\nu$ is the
835: frequency of the transition.
836: The anisotropy
837: factor
838: fulfills $-1/2 \leq w \leq 1$. It reaches $w=1$ when the radiation is
839: unidirectional
840: along the polar axis of the reference system, while $w=-1/2$ when the
841: radiation field
842: is azimuth-independent and confined to the plane perpendicular to the
843: quantization axis.
844: We have calculated the tensors of the radiation field by using the
845: center-to-limb variation and the wavelength dependence of the solar continuum
846: radiation
847: field
848: tabulated by \cite{pierce00}. Since the \ion{He}{1} lines originate in the
849: outer regions of the solar/stellar atmosphere, it is necessary to take into
850: account
851: the geometrical effect produced by the non-negligible height $h$ above the
852: surface of
853: the star. To this end, we have applied the strategy outlined in \S12.3 of
854: \cite{landi_landolfi04}.
855: Figure \ref{fig:nbar_omega} shows the sensitivity of $w$ and $\bar n$ to the
856: height in the solar atmosphere at which our slab is assumed to be located. The
857: number of photons
858: per mode decreases, while the anisotropy factor rapidly increases. This effect
859: is of a purely geometrical nature.
860:
861: It is important to note that the theory of spectral line
862: polarization
863: presented in the monograph of \cite{landi_landolfi04} is only valid if the
864: radiation
865: field illuminating the atomic system is spectrally flat (independent of
866: frequency) over a frequency interval larger than the Bohr frequencies connecting
867: the
868: levels that present quantum coherences. Fortunately, this is the case for all the lines
869: included in the atomic model of Fig. \ref{fig:helium_atom}, except for the
870: 3p$^3$P-2s$^3$S transition at 3888.6 \AA\ which
871: is situated in a quite crowded region of the spectrum.
872: Fortunately, this line is of
873: secondary importance in setting the statistical equilibrium, and according to
874: \cite{landi_landolfi04} the statistical tensors can be calculated by simply reducing the
875: photospheric continuum radiation intensity at that wavelength by a factor 5. In any case,
876: its inclusion
877: has a rather negligible impact on the
878: final results.
879:
880: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
881: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
882: % FORWARD MODELING
883: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
884: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
885: \section{The forward modeling code}
886:
887: By forward modeling we mean the calculation
888: of the emergent Stokes profiles for
889: given values of the height $h$ at which the slab
890: is located above the visible solar ``surface",
891: of the slab's optical thickness and of the magnetic field vector. We have
892: written this option of our computer program in a way such that it performs the
893: calculation at various levels of realism. We point out that similar type of
894: forward-modeling calculations can be carried out also with some of the computer
895: programs mentioned in \S1, which are likewise based on the density-matrix theory
896: of spectral line polarization. The main difference with ours is that we have taken
897: into account radiative transfer effects (without neglecting the magneto-optical
898: terms of the Stokes-vector transfer equation) and that we have developed a
899: user-friendly interface to facilitate the performing of numerical experiments
900: (see Fig. \ref{fig:front-end}). In what follows we first describe the various
901: options of our computer program and then show some illustrative examples of possible applications.
902:
903: \subsection{Description of the computing options}
904:
905: The solution of the statistical equilibrium equations, using the radiation field
906: tensors calculated from the given illumination conditions, provides the
907: mutipolar components of the atomic density matrix. The computer program
908: calculates such $\rho^K_Q(J,J^{'})$ elements in both the magnetic field
909: reference frame (if a deterministic
910: magnetic field is present) and in a reference system where the quantization
911: axis lies along the solar local vertical
912: direction (hereafter, ``vertical frame"). This option of the forward modeling
913: code helps to understand what's going on at the atomic level when an atomic
914: system is subjected to anisotropic radiative pumping processes in the absence
915: and in the presence of a magnetic field. Some illustrative examples for the
916: $J$-levels involved in the 10830 \AA\ and D$_3$ transitions are shown in \S3.2
917: below.
918:
919: The values of the multipolar components of the density
920: matrix is all we need for calculating the coefficients of the emission
921: vector and of the propagation matrix, which enter the Stokes vector transfer
922: equation (\ref{eq:rad_transfer}). Our forward modeling code can solve this
923: equation at the various
924: levels of approximation explained in \S\ref{sec:radiative_transfer}. Moreover,
925: it can also compute
926: the emergent Stokes profiles calculating the wavelength positions of the $\pi$
927: and $\sigma$ components assuming the linear Zeeman-effect regime (instead of
928: using the general Paschen-Back effect theory), incorporating or discarding the
929: influence of atomic level polarization.
930:
931: \subsection{Atomic level polarization in two reference systems}
932:
933: Figure \ref{fig:coherences} shows an illustrative
934: example of the population imbalances [$\rho^2_0(J)$ and $\rho^1_0(J)$]
935: and quantum coherences [$\rho^2_Q(J)$ and $\rho^1_Q(J)$, with $Q{\ne}0$] induced
936: by optical pumping processes in the levels of the \ion{He}{1} 10830 \AA\
937: multiplet that can carry atomic polarization (i.e., the lower level, with
938: $J_{l}=1$, and the upper levels with $J_{u}=2$ and $J_{u}=1$). Like in
939: \cite{trujillo_asensio07}, we
940: assume a slab with $\Delta{\tau}_{\rm red}=1$, where the label ``red'' indicates
941: that the optical
942: thickness of the slab along its normal direction
943: is measured at the frequency where the peak of the red blended
944: component of the 10830 \AA\ multiplet is located. The slab is assumed to be at a height
945: of only 3 arcseconds and in the presence of a horizontal magnetic field whose strength we can vary at will.
946:
947: Consider first the results for $J_{l}=1$ and $B=0$ G in a reference system with the quantization axis along the local vertical.
948: % (see
949: % the bottom left panel after recalling what is written in the last paragraph of \S2.2).
950: Since the pumping radiation is assumed to be unpolarized
951: and with axial symmetry around the vertical, for $B \approx 0$ G we see only population imbalances of the form $\rho^2_0(J)$.
952: % (i.e., for $B=0$ G
953: % the $\rho^K_Q(J_l)$ elements with $Q \ne 0$ are zero, while the $\rho^2_0(J)$
954: % atomic alignment value is non-zero).
955: Note that in the absence of magnetic fields
956: the $\rho^1_0(J_l) $ atomic orientation value is zero because there is no net
957: circular polarization in the incident radiation field.
958: The bottom right panel of
959: Fig. \ref{fig:coherences} shows that, for $B=0$ G, we have lower-level quantum
960: coherences of the form $\rho^2_Q(J)$ (with $Q{\ne}0$)
961: in the magnetic reference system, whose $Z$-axis is inclined with respect to the
962: symmetry axis of the pumping radiation field.
963: In fact, for given illumination conditions in the
964: absence of a magnetic field, whether or not we have such quantum coherences
965: depends only on the reference system.
966: In each
967: of the two bottom panels it is shown what happens with the population imbalances
968: and the
969: coherences of the $J_{l}=1$ lower level as the strength of the assumed
970: horizontal field is increased. Consider, for instance, the right panel results
971: corresponding to the magnetic field reference frame. Note that the lower-level
972: Hanle effect starts to operate for field strengths as low as 0.01 G, and that
973: for field intensities of the order or larger than 1 G all the lower-level coherences have been
974: relaxed. This happens because such a lower level
975: is metastable, which implies that its critical
976: Hanle field is only $0.1$ G (see Eq. \ref{eq:critical_field}).
977: It is also of interest to point out that
978: if the $J$-levels of our multiterm atomic model were isolated levels then the
979: $\rho^2_0(J)$ population imbalances would be constant in the magnetic field
980: reference frame, and the $\rho^1_0(J)$ atomic orientation value would be zero.
981: However, since the $J$-levels
982: suffer crossings and repulsions
983: non-zero $\rho^1_0(J)$ values appear through the alignment-to-orientation conversion
984: mechanism \citep[cf.][]{landi_d3_82}, which for the 10830 \AA\ triplet has
985: however a negligible impact on the emergent Stokes $V$ profiles. In addition, as
986: shown in the bottom right panel, the $\rho^2_0(J_l)$ alignment coefficient itself
987: starts to be modified as soon as the field strength is sensibly larger than 100
988: G. Although this lower level does not suffer any crossings with the other
989: $J$-levels of the model atom of Fig. \ref{fig:helium_atom}, its atomic polarization is modified
990: because it sensitively depends on that of the upper levels of
991: the 10830 \AA\ multiplet.
992:
993: Consider now the
994: results for the upper levels with $J_{u}=1$ (central panels of Fig.
995: \ref{fig:coherences}) and $J_{u}=2$ (top panels).
996: Obviously, we only see population imbalances of the form $\rho^2_0(J)$ at
997: zero field in the vertical frame. Around $B=10^{-2}$ G, the density matrix
998: elements start to be
999: affected by the magnetic field.
1000: This modification is due to the
1001: feedback effect that the alteration of the lower-level polarization has on the
1002: upper levels.
1003: It is also possible to note in the central and top panels of Fig.
1004: \ref{fig:coherences}
1005: the action of the upper level Hanle effect on the $J_{u}=1$ and $J_{u}=2$
1006: levels. In fact, there is a hint of a small plateau just above 0.1 G,
1007: as expected from the fact that the critical upper-level Hanle field value is 0.8
1008: G for the 10830 \AA\ multiplet (see Table \ref{tab:tab_einstein}). Similar
1009: features can be seen in the corresponding right panels of Fig.
1010: \ref{fig:coherences}.
1011: Probably, the most notable conclusion to highlight
1012: from this figure is that between approximately 10 and 100 G we only have
1013: population imbalances of the form $\rho^2_0(J)$ in the magnetic field reference
1014: frame (i.e., we can consider that between 10 and 100 G the He
1015: {\sc i} 10830 \AA\ multiplet is in the saturation regime of the Hanle effect,
1016: where the coherences are negligible and the atomic alignment values of the lower
1017: and upper levels are insensitive to the strength of the magnetic field).
1018:
1019: Finally, in Fig. \ref{fig:coherencesD3} we show similar results but for the upper levels of the He
1020: {\sc i} D$_3$ multiplet. The situation now is much more complicated, as
1021: evidenced by the fact that it is impossible to find a range of solar magnetic
1022: field strength values between which the coherences are zero in the magnetic
1023: field reference frame and, at the same time, the population imbalances are
1024: insensitive to the magnetic strength.
1025:
1026: \subsection{How to investigate empirically the possibility of magnetic canopies
1027: in the quiet solar chromosphere?}
1028: \label{sec:canopies}
1029: In the presence of an \emph{inclined} magnetic field forward scattering
1030: processes
1031: can produce linear polarization signatures in spectral lines
1032: \citep[e.g., the review by][]{trujillo01}. In this geometry, the polarization signal
1033: is \emph{created} by the Hanle effect, a physical phenomenon that has been
1034: clearly demonstrated via spectropolarimetry of solar coronal filaments in the
1035: \ion{He}{1} 10830 \AA\ multiplet \citep{trujillo_nature02}.
1036:
1037: Fig. \ref{fig:canopies} shows theoretical examples of the emergent fractional
1038: linear
1039: polarization in the lines of the \ion{He}{1} 10830 \AA\ multiplet
1040: assuming a constant-property slab of helium atoms permeated by a horizontal
1041: magnetic field of 10 G. As expected, the smaller the optical thickness of the
1042: assumed plasma structure the smaller the fractional polarization amplitude. In
1043: principle, the Tenerife Infrared
1044: Polarimeter \citep[TIP;][]{martinez_pillet99,collados_tipII07} mounted on the
1045: Vacuum Tower Telescope (VTT) of the Observatorio del Teide allows the detection
1046: of very low
1047: amplitude polarization signals, such as those corresponding to the
1048: $\Delta{\tau}_{\rm red}=0.1$ case of Fig. \ref{fig:canopies}. However, in order to be able
1049: to achieve
1050: this goal without having to sacrifice the spatial and/or temporal resolution we
1051: need a larger aperture solar telescope.
1052:
1053: We consider now the question of whether it is safer to interpret disk-center
1054: observations or to opt for a different scattering geometry. To this end, we have
1055: investigated how is the variation of $Q/I$ and $U/I$ at the central wavelengths
1056: of the red and blue components
1057: of the 10830 \AA\ multiplet for different inclinations ($\theta_B$) of the
1058: magnetic field vector and for different line-of-sights, assuming
1059: $\Delta{\tau}_{\rm red}=0.1$ and $B=10$ G (which implies that we are very close
1060: to the saturation regime of the upper level Hanle effect). Fig. \ref{fig:canopies_qi_peak} shows
1061: only the case of magnetic field vectors with azimuth $\chi_B=90^{\circ}$ (i.e.,
1062: contained in the Z-Y plane of Fig. \ref{fig:geometry}) and for line-of-sights with $\chi=0$
1063: (i.e., contained in the Z-X plane of Fig. \ref{fig:geometry}). It is interesting to note that for
1064: the case of a horizontal magnetic field (i.e., $\theta_B=90^{\circ}$, which
1065: implies that the magnetic field vector forms always an angle of $90^{\circ}$
1066: with respect to any of the line-of-sights of Fig. \ref{fig:canopies_qi_peak})
1067: Stokes $U=0$ and the
1068: Stokes $Q$ amplitudes of the emergent spectral line radiation are identical
1069: for all such line-of-sights. This is easy to understand by using
1070: the following approximate expressions for $\epsilon_Q/\epsilon_I$ and
1071: $\eta_Q/\eta_I$, which for the case of a deterministic magnetic field provide a
1072: suitable approximation if we are in the saturation regime of the upper-level
1073: Hanle effect \citep{trujillo03}:
1074:
1075: \begin{equation}
1076: {\frac{\epsilon_Q}{\epsilon_I}}\,{\approx}\,{\frac{3}{2\sqrt{2}}}\,({\mu}_B^2 -
1077: 1)\,{\cal W}\,[\sigma^2_0(J_u)]_B\, ,
1078: \label{eq:QIa}
1079: \end{equation}
1080:
1081: \begin{equation}
1082: {\frac{\eta_Q}{\eta_I}}\,{\approx}\,{\frac{3}{2\sqrt{2}}}\,({\mu}_B^2 -
1083: 1)\,{\cal Z}\,[\sigma^2_0(J_l)]_B\, ,
1084: \label{eq:QIb}
1085: \end{equation}
1086: where $[\sigma^2_0]_B=[\rho^2_0]_B/\rho^0_0$ quantifies the
1087: fractional atomic alignment in the magnetic
1088: field reference frame, while ${\cal W}$ and ${\cal Z}$ are numerical
1089: coefficients that depend on the $J_l$ and $J_u$
1090: values\footnote{Actually, ${\cal W}=w^{(2)}_{J_uJ_l}$ and ${\cal
1091: Z}=w^{(2)}_{J_lJ_u}$, with $w^{(2)}_{JJ^{'}}$ given by Eq. (10.12) of
1092: \cite{landi_landolfi04}. For instance, ${\cal W}=0$ and ${\cal Z}=1$ for a
1093: line transition with $J_l=1$ and $J_u=0$, such as that of the blue component of
1094: the 10830 \AA\ triplet.}. It is very important to note that in Eqs.
1095: (\ref{eq:QIa}) and (\ref{eq:QIb}) ${\mu}_B$ is {\em the cosine of the angle
1096: between the magnetic field vector and the line of sight}. Note also that in these
1097: expressions for $\epsilon_Q/\epsilon_I$ and $\eta_Q/\eta_I$, that are valid in the
1098: magnetic field reference frame, we have chosen the positive reference
1099: direction for Stokes $Q$ parallel to the projection of the magnetic field vector
1100: onto the plane perpendicular to the LOS, while in the similar Eqs. (16) and
1101: (17) of \cite{trujillo_asensio07} we chose it along the perpendicular direction. In both cases, we have $\epsilon_U=\eta_U=0$ (if we are really in the above-mentioned Hanle-effect saturation regime).
1102: Obviously, the observed $Q/I$ amplitude depends on the fractional atomic
1103: alignment of the upper and lower levels of the 10830 \AA\ line transitions
1104: calculated in the magnetic field reference frame, but their values are independent of the LOS.
1105: Actually, in the weak anisotropy limit they are given by
1106:
1107: \begin{equation}
1108: [\sigma^2_0]_{B}=[\sigma^2_0]_{V}{\,} \frac{1}{2} (3 \cos^2{\theta_B}-1),
1109: \end{equation}
1110: where $[\sigma^2_0]_{V}$ is the fractional atomic alignment in the vertical
1111: frame for the zero field case. Therefore, the dependency of the emergent Stokes $Q$ signal on the scattering angle is only through the factor $({\mu_B}^2 - 1)$, with
1112: ${\mu_B}=0$ for all the line-of-sights of Fig. \ref{fig:canopies_qi_peak} if
1113: $\theta_B=90^{\circ}$.
1114:
1115: As seen in Fig. \ref{fig:canopies_qi_peak}, for non-horizontal fields there are
1116: notable
1117: differences between the curves corresponding to each LOS, mainly for the cases
1118: with an
1119: inclination ($\theta_B$) of the magnetic field smaller than the Van-Vleck angle
1120: ($\theta_{VV}=54.73^{\circ}$, which corresponds to ${\rm
1121: cos}^{2}(\theta_{VV})=1/3$). For the forward-scattering case of a disk-center
1122: observation (the $\mu=1$ curves of
1123: Fig. \ref{fig:canopies_qi_peak}) Stokes $U \approx 0$ and Stokes $Q$ admits only
1124: one
1125: solution for $\theta_B > \theta_{VV}$ (with the exception of the well-known $180^{\circ}$ ambiguity of the Hanle effect)\footnote{The reason why Stokes $U$ in forward scattering geometry is
1126: not exactly zero for the red component is because for B=10 G some of the
1127: coherences are not completely insignificant.}.
1128: However, for $\theta_B < \theta_{VV}$ we may have two different magnetic
1129: field inclinations producing the same Stokes $Q$ values, each of them having
1130: its corresponding $180^{\circ}$ ambiguity of the Hanle effect. As seen in the figure, the range of
1131: $\theta_B$ values where two such solutions for Stokes $Q$ are possible depends
1132: on
1133: the $\mu$-value of the LOS.
1134: % For instance, for $\mu=0.1$ only one solution is
1135: % possible
1136: % if $\theta_B < 24^{\circ}$, while Stokes $Q$ admits two solutions for $\theta_B
1137: % > 24^{\circ}$.
1138: Stokes $U$ is clearly non-zero for $\mu < 1$, but the fact that
1139: $U \approx 0$ for the case of magnetic field with fixed inclination but with
1140: random-azimuth within the
1141: spatio-temporal resolution element of the observation, leads us to conclude
1142: that
1143: the best one can do for a reliable empirical investigation of the possibility
1144: of
1145: canopy-like fields in the quiet solar chromosphere is to interpret
1146: spectropolarimetric
1147: observations at the solar disk center, such as those considered in
1148: \S\ref{sec:internetwork_regions}.
1149:
1150: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1151: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1152: % INVERSION TECHNIQUES
1153: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1154: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1155: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1156: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1157: \section{The Inversion Code}
1158: \label{sec:inversion}
1159:
1160: Our inversion strategy is based on the minimization of a merit function
1161: that quantifies how well the Stokes profiles calculated in our atmospheric model
1162: reproduce the observed Stokes
1163: profiles. To this end, we have
1164: chosen the standard $\chi^2$--function, defined as:
1165: \begin{equation}
1166: \chi^2 = \frac{1}{4N_\lambda} \sum_{i=1}^4 \sum_{j=1}^{N_\lambda}
1167: \frac{\left[S_i^\mathrm{syn}(\lambda_j)-S_i^\mathrm{obs}(\lambda_j) \right]^2}{
1168: \sigma_i^2(\lambda_j)} ,
1169: \end{equation}
1170: where $N_\lambda$ is the number of wavelength points and $\sigma_i^2(\lambda_j)$ is the
1171: variance associated to the $j$-th wavelength point of the $i$-th Stokes profiles. The minimization
1172: algorithm tries to find the value of the parameters of our model that lead to
1173: synthetic Stokes profiles $S_i^\mathrm{syn}$ with the best possible fit to the
1174: observations.
1175: For our slab model, the number of
1176: parameters (number of dimensions of the $\chi^2$ hypersurface) lies between 5
1177: and
1178: 7, the maximum value corresponding to the optically thick case (see Table
1179: \ref{tab:parameters}).
1180: The magnetic field vector
1181: ($B$, $\theta_B$ and $\chi_B$), the thermal velocity ($v_\mathrm{th}$) and the
1182: macroscopic velocity ($v_\mathrm{mac}$) are always required. This set of
1183: parameters is enough
1184: for the case of an optically thin slab. In order to account for radiative
1185: transfer
1186: effects, we need to define the optical depth of the slab along its normal
1187: direction and at a suitable
1188: reference wavelength (e.g., the central wavelength of the red blended component
1189: for the \ion{He}{1} 10830 \AA\ multiplet). In addition,
1190: we may additionally need to include the damping parameter ($a$) of the Voigt profile if
1191: the wings of the observed Stokes profiles cannot be fitted
1192: using Gaussian line profiles.
1193:
1194: A critical problem in any inversion code is to identify possible degeneracies
1195: among different parameters of the model. When two or more parameters
1196: produce similar effects on the emergent Stokes profiles, the inversion algorithm
1197: is unable to distinguish between them. As a result, the emergent Stokes profiles
1198: corresponding to different combinations of the model parameters are
1199: indistinguishable within the noise level. Concerning this critical point, we
1200: investigate in detail in \S\ref{sec:invert_height} the possibility of
1201: using the observed Stokes profiles of the 10830 \AA\ triplet to obtain the value
1202: of the height $h$ at which the observed plasma structure is located.
1203:
1204: Ambiguities in the determination of the model's parameters can also result from
1205: the presence of degeneracies.
1206: However, this type of ambiguities occur only for a finite number of combinations
1207: of some
1208: parameters. Although the problem is complicated, it is possible to develop
1209: techniques that can help selecting one of the combinations as the most
1210: plausible. The well-known 180$^\circ$ ambiguity of the Hanle effect
1211: adds to the unfamiliar
1212: Van Vleck ambiguity \citep{house77,casini_judge99,casini05,merenda06}. The role
1213: of these ambiguities in the inferred model's parameters will be explored in \S\ref{sec:van_vleck},
1214: although some hints have been already given in
1215: \S\ref{sec:canopies}.
1216:
1217: It is also important to point out the interest of developing methods capable of
1218: providing reliable confidence intervals for all the inferred parameters. In a
1219: future development, we plan to implement Bayesian inference techniques
1220: \citep{asensio_martinez_rubino07}.
1221:
1222: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1223: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1224: \subsection{Levenberg-Marquardt}
1225: Probably, the most well-known
1226: procedure for the minimization of the $\chi^2$-function is the
1227: Levenberg-Marquardt (LM) method \citep[e.g.,][]{numerical_recipes86}.
1228: The minimization strategy uses the Hessian method when the
1229: parameters are close to the minimum of the $\chi^2$-function
1230: (a quadratic form approximately describes this function
1231: around the minimum) and the steepest descent method when the
1232: parameters are far from the minimum. The
1233: transition between both methods is done in an adaptive manner.
1234: % The LM method has become the standard numerical approach for the
1235: % minimization of nonlinear functions.
1236: Its main drawback
1237: (which applies also to the majority of the standard numerical methods of function minimization) is that it can easily get trapped in local minima of the
1238: $\chi^2$-function. Some alternatives are available to overcome this difficulty. The
1239: most straightforward but time consuming one is to restart the minimization
1240: process
1241: at different values of the initial parameters. If the obtained minimum is
1242: systematically the same, the probability that this is the global optimum is
1243: high. However, when the $\chi^2$ parameter hypersurface is complicated, this
1244: technique does not give any confidence on the validity of the global minimum.
1245: Other possibilities rely on the application of some kind of ``inertia'' to the
1246: method, so that the LM method can overcome such a local minimum problem when
1247: moving on the parameter hypersurface. Again, these methods do not guarantee the
1248: success of getting the global minimum of the function.
1249: On the contrary, the LM method turns out to be one of the fastest and simplest
1250: options when the initial estimate of the parameters is close to the
1251: absolute minimum.
1252:
1253: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1254: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1255: \subsection{Global Optimization techniques}
1256: In order to avoid the possibility of getting trapped in a local minimum of the
1257: $\chi^2$ hypersurface, global
1258: optimization methods have to be used. Several optimization methods have been
1259: developed to obtain the global minimum of a function
1260: \citep[e.g.,][]{global_optimization95}. The majority of them are based
1261: on stochastic optimization techniques. The essential philosophy of
1262: these methods is to sample efficiently the whole space of parameters
1263: to find the global minimum of a given function.
1264: % Each one of the available
1265: % methods
1266: % uses different techniques to avoid sampling the whole space of parameters.
1267: One of the most promising methods is genetic optimization (inspired by
1268: the fact of biological evolution), in
1269: which the parameters of the merit function are encoded in a gene.
1270: % Inspired by
1271: % the fact of
1272: % biological evolution, the algorithm performs crossings between the elements of a
1273: % given population of solutions, together with some degree of mutation.
1274: Although no mathematical proof of the convergence properties of these
1275: algorithms exists, recent advances suggest that the probability of
1276: convergence is very high \citep{gutowsky04}. Actually, they perform quite well
1277: in practice
1278: for the optimization of very hard problems. In solar spectropolarimetry,
1279: genetic optimization methods have been recently applied by \cite{Lagg04} to the
1280: inversion of Stokes profiles induced by the Zeeman effect in the \ion{He}{1}
1281: 10830 \AA\ triplet, neglecting the influence of atomic level polarization. The main
1282: disadvantage is that the computing
1283: time needed to reach convergence increases dramatically (by a factor $\sim$10
1284: with
1285: respect to standard methods based on the gradient descent like LM).
1286: % The
1287: % advantage is that genetic algorithms seem to produce much better
1288: % results in terms of continuity in the maps of magnetic field strength,
1289: % inclination and azimuth.
1290:
1291: Another different approach is based on deterministic algorithms
1292: \citep{global_optimization95}. Typically, these algorithms rely on a strong
1293: mathematical basis, so that their convergence properties are well known.
1294: We have chosen the DIRECT algorithm
1295: \citep{Jones_DIRECT93}, whose name derives from one of its main
1296: features: \emph{di}viding \emph{rect}angles. The idea is to recursively sample
1297: parts of the space of parameters, improving in each iteration the location of
1298: the part of the space
1299: where the global minimum is potentially located. The decision algorithm is based
1300: on the assumption that the function is Lipschitz continuous \citep[see][for details]{Jones_DIRECT93}.
1301: % Recall that a
1302: % one-dimensional function $f: M \to R$ is said to be Lipschitz continuous with
1303: % constant $\alpha$ in an interval $M=[a,b]$ when:
1304: % \begin{equation}
1305: % |f(x)-f(x')| \leq \alpha |x-x'|, \qquad \forall x, x'\in M.
1306: % \end{equation}
1307: % In particular, a consequence of the previous equation is that the two following
1308: % inequalities hold $\forall x \in M$:
1309: % \begin{eqnarray}
1310: % f(x) &\geq& f(a) - \alpha (x-a) \\
1311: % f(x) &\geq& f(b) + \alpha (x-b) .
1312: % \end{eqnarray}
1313: % These two straight lines
1314: % form a V-shape below the function $f(x)$, and their intersection
1315: % provides a first estimation, $x_1$, of the minimum of the function. Repetition
1316: % of
1317: % this procedure in the two new subintervals $[a,x_1]$ and $[x_1,b]$ leads to a
1318: % convergent algorithm. This algorithm is not as simple in complicated problems
1319: % with higher dimensions because of the difficulty in estimating the Lipschitz
1320: % constant.
1321: The method works very well in practice and can indeed find the minimum in
1322: functions that do not fulfill the condition of Lipschitz continuity. The reason
1323: is that the DIRECT algorithm does not require the explicit calculation of the
1324: Lipschitz constant but it uses all possible values of such a constant to determine
1325: if
1326: a region of the parameter space should be broken into subregions because of
1327: its potential interest \citep[see][for details]{Jones_DIRECT93}. A schematic
1328: illustration of the subdivision process for
1329: a function of two parameters is shown in Fig. \ref{fig:direct_method}.
1330:
1331: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1332: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1333: \subsection{Convergence}
1334: \label{sec:convergence}
1335: Taking into account that the dimension of our space of parameters is
1336: between 5 and 7, it seems unreasonable to use an algorithm like DIRECT to
1337: obtain a precise determination of the values of the model's parameters at the
1338: global minimum. The reason is that the precision
1339: in the values of the parameters decreases with the size of
1340: the hyperrectangles. Therefore, we would need to perform many divisions
1341: to end up with a reasonable precision. What we do is to let the DIRECT algorithm
1342: locate
1343: the global minimum in a region whose hypervolume is $V$. This hypervolume is
1344: obtained as the product of the length $d_i$ of each dimension associated with
1345: each of the $N$ parameters:
1346: \begin{equation}
1347: V = \prod_i^N d_i.
1348: \end{equation}
1349: When the hypervolume decreases by a factor $f$ after the DIRECT algorithm
1350: has discarded some of the hyperrectangles, its size along each dimension is
1351: approximately decreased by a factor $f^{1/N}$.
1352: In order to end up with a small region
1353: where the global minimum is located, many subdivisions are
1354: necessary, thus requiring many function evaluations.
1355: For this reason, it has been observed that although
1356: the DIRECT algorithm rapidly finds the region where the
1357: global minimum is located, its local convergence properties are rather poor
1358: \citep[see, e.g., ][for applications in the extremely hard problems of the
1359: design of
1360: high-speed civil transport, aircrafts and
1361: bioinformatics]{cox01,bartholomew02,ljungberg04}. In summary, the DIRECT
1362: method is an ideal candidate for its application as an estimator of the region
1363: where the global minimum is located, but not for determining it.
1364:
1365: The most time consuming part of any optimization procedure is the evaluation of
1366: the merit function. The DIRECT algorithm needs only a reduced number of evaluations
1367: of the merit function to find
1368: the region where the global minimum is located. For this reason, we have
1369: chosen it as the initialization part of the LM method. Since the initialization
1370: point is close to the global minimum, the LM method, thanks to its quadratic behavior,
1371: rapidly converges to the minimum.
1372:
1373: \subsection{Stopping criterium}
1374: A critical and fundamental problem in the optimization of functions (either
1375: local or global) is to identify when the method has converged to the solution.
1376: % In cases where the value of the function
1377: % at the global minimum is known, it is possible to stop the
1378: % convergence process when
1379: % the relative error is smaller than a fixed value. However, this is the case only
1380: % for the
1381: % optimization of simple functions.
1382: We have used two stopping criteria for the
1383: DIRECT algorithm. The first one is stopping when the ratio between the
1384: hypervolume where
1385: the global minimum is located and the original hypervolume is smaller than a
1386: given threshold.
1387: This method has been chosen when using the DIRECT
1388: algorithm as an initialization for the LM method, giving very good results. The
1389: other good
1390: option, suggested by \cite{Jones_DIRECT93}, is to stop after a fixed number of
1391: evaluations of the merit function.
1392:
1393: Since the
1394: intensity profile is not very sensitive to
1395: the presence of a magnetic field (at least for magnetic field
1396: strengths of the order of or smaller than 1000 G), we have decided to estimate
1397: the optical
1398: thickness of the slab, the thermal and the macroscopic velocity of the
1399: plasma and the damping constant by using only the Stokes $I$ profile, and then to determine the magnetic
1400: field
1401: vector by using the polarization profiles.
1402: The full inversion scheme, shown schematically in Table \ref{tab:inversion},
1403: begins by applying the DIRECT method to obtain a first estimation of the
1404: indicated four
1405: parameters by using only Stokes $I$. Afterwards,
1406: some LM iterations are carried out to refine the initial values of the
1407: model's parameters obtained in the previous step. Once the LM method
1408: has converged, the inferred values of $v_\mathrm{th}$, $v_\mathrm{mac}$
1409: (together with $a$ and $\Delta \tau$, when these are parameters of the model)
1410: are kept fixed in the next steps,
1411: in which the DIRECT method is used again for obtaining an initial approximation
1412: of
1413: the magnetic field vector
1414: ($B$,$\theta_B$,$\chi_B$).
1415: According to our experience,
1416: the first estimate of the magnetic field vector given by the DIRECT algorithm
1417: is typically very close to the final solution. Nevertheless, some iterations of
1418: the LM method are performed to refine the value of the magnetic field strength,
1419: inclination and azimuth.
1420: In any case, although we have found very good results with this procedure, the
1421: specific inversion scheme
1422: is fully configurable and can be tuned for specific problems.
1423:
1424: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1425: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1426: % APPLICATIONS
1427: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1428: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1429: \section{Applications}
1430: The main aim of this section is to illustrate the application of our inversion
1431: code to some selected spectropolarimetric
1432: observations in the \ion{He}{1} 10830 \AA\ multiplet, showing that it gives
1433: results that are
1434: in agreement with the published ones. In addition, in
1435: \S\ref{sec:internetwork_regions} we show a new application aimed
1436: at determining the strength and inclination of the magnetic field vector in the chromosphere above an
1437: internetwork region observed at solar disk center. Note that in the following
1438: applications $\Delta{\tau}_{\rm red}$ will continue denoting the optical thickness
1439: of the constant-property slab, along its normal direction, measured at the center
1440: of the red blended component of the \ion{He}{1} 10830 \AA\ multiplet.
1441:
1442: \subsection{Prominences}
1443: The first application is for the case of solar prominences, which are relatively
1444: cool and dense plasma structures embedded in the $T\sim 10^6$ K solar corona. In
1445: these objects the
1446: observed Stokes $Q$ and $U$ profiles of the \ion{He}{1} 10830 \AA\ multiplet are
1447: dominated by the presence of atomic level
1448: polarization, while the Stokes $V$ profile
1449: is dominated by the Zeeman effect. Recently,
1450: \cite{merenda06} have shown how to infer the
1451: magnetic field that confines the plasma of solar prominences via the inversion
1452: of the Stokes profiles induced by scattering processes and the Hanle and Zeeman
1453: effects in
1454: the lines of the \ion{He}{1} 10830 \AA\ multiplet. They analyzed in detail
1455: spectropolarimetric
1456: observations of the \ion{He}{1} 10830 \AA\ multiplet
1457: in a polar crown prominence and concluded that if the observed prominence
1458: was located in the plane of the sky
1459: the
1460: magnetic field had to be relatively strong ($B \approx 30 $ G) and inclined by only
1461: $25^{\circ}$ with respect to the local vertical.
1462:
1463: We have applied our inversion code to the
1464: spectropolarimetric observations
1465: shown in Fig. \ref{fig:prominence}, taken from Fig. 9 of \cite{merenda06}. We
1466: have assumed that
1467: the observed plasma structure was optically thin
1468: and that the prominence
1469: is located in the plane of the sky. The inversion code was
1470: used to infer the value of the thermal velocity $v_\mathrm{th}$, the
1471: macroscopic
1472: velocity shift $v_\mathrm{mac}$ (to allow for a shift in the wavelength
1473: calibration) and
1474: the magnetic field vector $(B,\theta_B,\chi_B)$. The atmospheric height was
1475: fixed to
1476: $h=20"$,
1477: the same value used by \cite{merenda06}.
1478: After the four steps summarized in
1479: Table \ref{tab:inversion}, we end up with a thermal velocity
1480: $v_\mathrm{th}=7.97$ km s$^{-1}$, a bulk velocity that is compatible with zero
1481: and a magnetic field vector characterized by $B=26.8$ G,
1482: $\theta_B=25.5^\circ$ and
1483: $\chi_B=161.0^\circ$. These values are in very good agreement with the results
1484: of \cite{merenda06}, namely $B=26$ G, $\theta_B=25^\circ$ and
1485: $\chi_B=160.5^\circ$. Note that since the prominence plasma was assumed to lie in the
1486: plane of the sky, the following solutions are also valid:
1487: $\theta_B^{*}=180^\circ-\theta_B$ and $\chi_B^{*}=-\chi_B$ (i.e., the
1488: well-known $180^{\circ}$ ambiguity of the Hanle effect).
1489:
1490: We point out that the total number of evaluations of the merit function was 132. For the inversion of the Stokes profiles corresponding to other points of the field of view, one can initialize the inversion using the model's parameters corresponding to the
1491: previous point. Using a few LM iterations, one should be able
1492: to reach the global minimum. In case this procedure does not work properly, one should return to the four-steps
1493: inversion scheme already presented in Table \ref{tab:inversion}.
1494:
1495: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1496: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1497: \subsection{Spicules}
1498: \label{sec:spicules}
1499: Another interesting problem is
1500: the determination of the magnetic field vector in solar
1501: chromospheric spicules. \cite{trujillo_merenda05} interpreted
1502: spectropolarimetric observations of spicules in the
1503: \ion{He}{1} 10830 \AA\ multiplet and concluded that the magnetic
1504: field of spicules in quiet regions of the solar chromosphere has a strength
1505: of the order of 10 G and is inclined by about 35$^\circ$ with respect to the
1506: local vertical. Their conclusion that the
1507: typical magnetic field strength is $\sim 10$ G required to obtain the
1508: longitudinal component of the magnetic field vector via
1509: some careful measurements of the Stokes
1510: $V$ profiles, such as that shown in Fig. 13 of \cite{trujillo_esa05}.
1511: This was needed because for field strengths larger than a few gauss
1512: the \ion{He}{1} 10830 \AA\ multiplet enters the saturation regime of the
1513: upper-level Hanle effect and the observed Stokes $Q$ and $U$ profiles provide only
1514: information on the orientation of the magnetic field vector.
1515: In fact, the application of our inversion code to the observed Stokes profiles
1516: shown in Fig. \ref{fig:spicules} (where the Stokes $V$ profile is at the noise level),
1517: assuming that the spicular material
1518: is located in the plane of the sky, provides several different magnetic field vectors that lead to equally
1519: reliable fits. One of them, given by $B=10$ G, $\theta_B=37^\circ$ and
1520: $\chi_B=172^\circ$ is similar
1521: to the one chosen by \cite{trujillo_merenda05}. Another possible fit is
1522: the one illustrated in Fig. \ref{fig:spicules}, which corresponds to $B=2.6$ G, $\theta_B=37^\circ$ and $\chi_B=35^\circ$.
1523:
1524: Fig. \ref{fig:spicules_chi2} gives the values of the $\chi^2$ function for
1525: all possible magnetic field inclinations and azimuths corresponding to
1526: the cases $B=10$ G (left panel) and $B=2.6$ G (right panel). In each of these
1527: panels we have indicated with white dots and numbers the four solutions that
1528: correspond to equally good best fits to the observed Stokes profiles of Fig. \ref{fig:spicules}.
1529: The pair of solutions $1$ and $2$ correspond to the Van-Vleck ambiguity\footnote{Information on this ambiguity,
1530: typical of the Hanle-effect saturation regime, can be found in \cite{casini05}, in \cite{merenda06} and in \S\ref{sec:van_vleck} below.}.
1531: The same applies to the $1'$ and $2'$ solutions.
1532: On the contrary, the pair of solutions
1533: $1$ and $1'$ or the $2$ and $2'$ are not strictly equivalent. The inversion code considers
1534: such pairs of solutions as equivalent because the observed Stokes $V$ profile is at
1535: the noise level and it is not able to differentiate between the two cases.
1536: Note that these pairs of solutions give exactly the same Stokes $Q$ and $U$ profiles, but their corresponding
1537: Stokes $V$ profiles have opposite signs.
1538: Concerning each pair of solutions in Fig. \ref{fig:spicules_chi2}, it is
1539: possible to verify that the projections on the plane of the sky of the magnetic fields
1540: corresponding to solutions $1$ and $2$ form an angle close to to 90$^\circ$, which is typical of the Van-Vleck ambiguity.
1541: The same happens for the magnetic fields corresponding to
1542: solutions $1'$ and $2'$. This holds for both cases, $B=10$ G and $B=2.6$ G. As pointed out above, when the observed Stokes $V$ signal is very small, it is
1543: very hard (or impossible) to differentiate between the two
1544: possibilities having azimuths $\chi_B$ and $180^\circ-\chi_B$. The
1545: magnetic field vectors $1$ and $1'$ (or those corresponding to the $2$ and $2'$
1546: solutions of Fig. \ref{fig:spicules_chi2}) have the same projection on the
1547: line of sight, except for a sign change. Therefore, the detection of Stokes $V$ turns out to be fundamental to
1548: determine which is the
1549: correct one \citep{trujillo_merenda05,merenda06}. Apart from the considered solutions, which are
1550: restricted to the interval $0^\circ < \theta_B < 180^\circ$, one has also
1551: to take into account the well-known ambiguity of the Hanle effect, which applies when
1552: the emitting plasma is located in the plane of the sky. In this case, we have to add to the possible
1553: set of solutions all the combinations fulfilling $\theta_B^{*}=180^\circ-\theta_B$ and $\chi_B^{*}=-\chi_B$ since both
1554: pairs produce the same Stokes profiles.\footnote{We point out that this ambiguity of the Hanle effect applies only to some particular scattering geometries (i.e., to those of 90$^\circ$ and of forward scattering). If the observed plasma structure is not located in the plane of the sky (which implies scattering processes at an angle different from 90$^\circ$), or if it is not located at the solar disk center, then one has a sort of quasi-degeneracy which can disappear when the angle $\theta$ of Fig. 1 is considerably different from 90$^\circ$ or from 0$^\circ$. This fact has been exploited by \cite{landi_bommier93} to propose a method for removing the azimuth ambiguity intrinsically present in vector magnetograms.}.
1555:
1556: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1557: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1558: \subsection{Filaments}
1559: We have also considered the inversion of the Stokes profiles presented in
1560: \cite{trujillo_nature02}, which were observed
1561: in a solar coronal filament at the solar disk center. Such profiles, which are
1562: reproduced in Fig. \ref{fig:filament_observation}, were used by those authors to
1563: demonstrate the presence of atomic polarization in the lower level of the 10830
1564: \AA\ multiplet
1565: and that the Hanle effect due to an inclined magnetic field creates
1566: linearly polarized radiation in forward scattering geometry. Note that the
1567: Stokes $Q$, $U$
1568: and $V$ profiles are normalized to the maximum depression in Stokes $I$
1569: (which is $0.4\,I_c$, approximately). The
1570: application of our inversion code using $h=40"$ confirms the conclusions of
1571: \cite{trujillo_nature02}, yielding the following values for the model's
1572: parameters: $\Delta \tau=0.86$, $v_\mathrm{th}=6.6$ km s$^{-1}$, $a=0.19$ and a
1573: magnetic field vector characterized by $B=18$ G and $\theta_B=105^\circ$.
1574:
1575: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1576: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1577: \subsection{Inter-network regions}
1578: \label{sec:internetwork_regions}
1579: As discussed in \S\ref{sec:canopies}, the investigation of the possibility of horizontal magnetic canopies in the quiet solar chromosphere above internetwork regions is feasible
1580: with TIP, especially when interpreting measurements of the polarization of the \ion{He}{1} 10830 \AA\ multiplet in forward scattering at the solar disk center\footnote{For a preliminary interpretation of an observation of a quiet region located at $\mu=0.5$
1581: see \cite{Lagg07}.}.
1582: We present in Fig. \ref{fig:canopy_observation} an observation carried
1583: out with TIP very close
1584: to the solar disk center ($\mu=0.98$). The slit was crossing an enhanced network
1585: region of circular shape.
1586: A time series of 50 steps with an integration time of 3 seconds was performed.
1587: The resulting polarimetric
1588: sensitivity after averaging over the 50 time steps and along a small spatial
1589: interval within the observed
1590: internetwork region is close to 6$\times$10$^{-5}$ in units of the continuum
1591: intensity. According to the results
1592: of the right panel of Fig. \ref{fig:canopies}, this is sufficient for detecting the
1593: linear polarization signal of a horizontal magnetic field provided the optical
1594: depth at the wavelength of the red
1595: component of the 10830 \AA\ multiplet is larger than $\sim 0.01$. The reference
1596: system has been rotated until
1597: Stokes $U$ is minimized. Since the inferred magnetic field
1598: strength implies that the \ion{He}{1} 10830 \AA\ is in the saturation
1599: regime of the upper-level Hanle effect, the resulting
1600: reference direction for Stokes $Q$ lies either
1601: along the projection of the magnetic
1602: field vector on the solar disk, or along the direction perpendicular to such a projection.
1603: We have applied our inversion
1604: code to the above-mentioned observed profiles assuming a
1605: slab located at a
1606: height of 3 arcsec and we have obtained the following
1607: results: $\Delta \tau=0.19$, $v_\mathrm{th}=9.2$ km s$^{-1}$, $a=0.62$ and a
1608: magnetic field vector characterized by
1609: $B=35$ G, $\theta_B=21^\circ$ and $\chi_B=0^\circ$. However, other possible
1610: solutions
1611: can be found with a similar goodness of the fit (e.g.,
1612: $B=47$ G, $\theta_B=47^\circ$ and $\chi_B=0^\circ$).
1613: These results obtained from a solar disk center observation
1614: suggest the presence of magnetic fields inclined by no more than
1615: $50^{\circ}$ in the observed quiet Sun chromospheric region.
1616:
1617: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1618: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1619: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1620: \subsection{Emerging magnetic flux regions}
1621: As pointed out by \citet{trujillo_asensio07}, the modeling of the emergent
1622: Stokes $Q$ and $U$ profiles of the \ion{He}{1} 10830 \AA\ multiplet should be done by
1623: taking into account the possible presence of
1624: atomic level polarization, even for magnetic field strengths
1625: as large as 1000 G.
1626: An example of a spectropolarimetric observation of
1627: an emerging magnetic flux region is shown by the circles of Fig.
1628: \ref{fig:lagg_emerging}.
1629: The solid lines show the best theoretical fit to these
1630: observations of \cite{Lagg04}. Here, in addition to the Zeeman effect, we
1631: took into account the influence of atomic level polarization. The
1632: dotted lines neglect the atomic level polarization that is induced by
1633: anisotropic radiation pumping
1634: in the solar atmosphere. Our results
1635: indicate the presence of atomic level polarization
1636: in a relatively strong field region (${\sim}$1000 G). However, it may be
1637: tranquilizing to point out that both inversions of the observed profiles yield, at least
1638: for this case, a
1639: similar
1640: magnetic field vector, in spite of
1641: the fact that the corresponding theoretical fit is much better for the case that
1642: includes atomic
1643: level polarization.
1644:
1645: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1646: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1647: \section{Ambiguity and degeneracies}
1648:
1649: In the previous subsections we have shown how our inversion code can be used for
1650: recovering the parameters of the
1651: assumed slab atmospheric model from the Stokes profiles observed in different
1652: solar plasma structures. The aim of this section is to discuss the Van Vleck
1653: ambiguity and to investigate
1654: whether we can infer the height of the
1655: observed plasma structure directly through the inversion approach.
1656:
1657: \subsection{Van Vleck Ambiguity}
1658: \label{sec:van_vleck}
1659: In general, the solution to any inversion problem is not unique --that is, it is often possible
1660: to detect several solutions which are compatible with the observations
1661: \cite[e.g.,][]{asensio_martinez_rubino07}.
1662: Some of the unicity problems are associated with the physics of the polarization
1663: phenomena (e.g., the 180$^\circ$ ambiguity of the Hanle effect for plasma structures located in the plane of the sky or the Van Vleck ambiguity).
1664: However, as seen in \S\ref{sec:spicules},
1665: in addition to this type of ambiguities, other degeneracies can appear because of the presence of
1666: noise in the observed Stokes profiles.
1667:
1668: The Van Vleck ambiguity occurs only for some
1669: combinations
1670: of the inclinations and azimuths. Moreover, it occurs mainly in the saturation regime
1671: of the Hanle effect. For example, Fig. 6 of \cite{merenda06}
1672: shows the region of parameters for which the Van
1673: Vleck ambiguity occurs in the Hanle-effect
1674: saturation regime of the \ion{He}{1} 10830 \AA\ triplet.
1675: Since two different
1676: magnetic field vectors give rise to exactly the same emergent Stokes
1677: profiles, it is impossible to distinguish between them
1678: using only the 10830 \AA\ multiplet
1679: (or four solutions, if the 180$^\circ$ ambiguity of the Hanle effect also applies). However,
1680: if more information is introduced
1681: in the inversion procedure (for instance, by using simultaneous observations in
1682: the 10830 \AA\
1683: and D$_3$ multiplets), it might be possible to distinguish between the two possible solutions.
1684:
1685: Unfortunately, it is not easy to determine
1686: the range of parameters in which we may have
1687: the Van Vleck ambiguity. One possibility \citep[used by][]{merenda06} is to
1688: consider the theoretical Hanle diagram of the red line of the He {\sc i} 10830 \AA\ multiplet and detect if the
1689: observed profiles fall in the ambiguity region. We propose another
1690: method based on the DIRECT algorithm implemented in our inversion code.
1691: The DIRECT algorithm can rapidly detect regions of the space of parameters
1692: where the global minimum may be located. Therefore, we can take advantage of
1693: this property to identify the two (or more) points in the space of parameters
1694: $(\theta_B,\chi_B)$ that
1695: produce the same emergent Stokes profiles for a given magnetic field strength.
1696:
1697: To this end, we have calculated the synthetic emergent Stokes profiles of the
1698: \ion{He}{1} 10830 \AA\ line from an optically thin prominence, located in the plane of
1699: the sky, with $v_\mathrm{th}=8$ km s$^{-1}$, $h=20"$, $B=25$ G, $\theta_B=40^\circ$ and
1700: $\chi_B=19^\circ$.
1701: According to the Hanle diagram shown by \cite{merenda06}, these profiles are
1702: indistinguishable from the ones given by the combination $B=22$ G,
1703: $\theta_B=100^\circ$ and $\chi_B=46^\circ$. To these two combinations,
1704: we have to add those corresponding to the 180$^\circ$ ambiguity: ($B=25$ G,
1705: $\theta_B=140^\circ$,
1706: $\chi_B=-19^\circ$) and ($B=22$ G, $\theta_B=80^\circ$, $\chi_B=-46^\circ$).
1707: Using
1708: the standard four steps inversion procedure explained in
1709: \S \ref{sec:inversion},
1710: the global minimum is rapidly located at position $B=22$ G,
1711: $\theta_B=100^\circ$ and
1712: $\chi_B=46^\circ$. Keeping fixed the value of all the thermodynamical
1713: properties
1714: and the field strength, the DIRECT algorithm is used to recover the inclination
1715: and
1716: azimuth of the magnetic field vector. We show in Fig \ref{fig:vanvleck_ambiguity} the position in the
1717: $(\theta_B,\chi_B)$ space of parameters of the $N$ evaluations performed by the
1718: DIRECT method. It has been possible to detect
1719: the two combinations of parameters that give the same emergent Stokes profiles,
1720: as stated above. With only $N=100$ evaluations of the merit function, the
1721: DIRECT algorithm has located and refined the position of the global minimum. It
1722: has also identified the second possible solution.
1723: When the
1724: number of function evaluations increases (even with only $N=200$), the DIRECT
1725: algorithm rapidly locates the two minima. For $N>200$, we face a
1726: degradation in the convergence rate as discussed in \S\ref{sec:convergence}.
1727:
1728: As already discussed, an interesting property of the DIRECT method is that no
1729: hyper-rectangle is ever discarded
1730: from the search. Therefore, a rectangle that in one iteration is not considered
1731: to be potentially interesting, can be chosen for division in a later iteration\footnote{This
1732: proves to be fundamental to demonstrate that the global minimum will always be found
1733: \citep{Jones_DIRECT93}.}.
1734: This behavior is shown in Fig.
1735: \ref{fig:vanvleck_ambiguity}. When $N=100$, only a part of the space of
1736: parameters
1737: has been sampled, with clear gaps for inclinations above 110$^\circ$. In spite
1738: of these gaps, the two global minima have been already
1739: found. However, when the number of function evaluation is increased, the
1740: numerical scheme finally evaluates the function in those regions with the
1741: aim of discarding the presence of additional global minima.
1742:
1743: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1744: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1745: \subsection{Can we infer the height of the observed plasma structure?}
1746: \label{sec:invert_height}
1747: In this section we briefly discuss the possibility of determining the height at
1748: which the \ion{He}{1} atoms are located by only using the information contained
1749: in
1750: the Stokes profiles of the 10830 \AA\ multiplet. For simplicity, we consider
1751: first the optically thin limit, the case of off-limb observations (i.e., $90^{\circ}$
1752: scattering geometry) and a magnetic field with a fixed
1753: azimuth and strength. The
1754: synthetic
1755: profiles correspond to the case $v_\mathrm{th}=8$ km s$^{-1}$ and
1756: $h=20"$, with the magnetic field vector given by $B=25$ G,
1757: $\theta_B=40^\circ$ and
1758: $\chi_B=19^\circ$. The aim of this experiment is to infer the inclination
1759: $\theta_B$ and height $h$ from synthetic Stokes profiles without noise.
1760: The positions where the merit
1761: function has been evaluated by the DIRECT algorithm are presented in the upper panels
1762: and in the bottom left panel of Fig.
1763: \ref{fig:height_degeneration}.
1764: % It clearly indicates that the
1765: % problem is intrinsically difficult.
1766: The $\chi^2$ surface is shown in the bottom right
1767: panel of Fig. \ref{fig:height_degeneration}. The presence of the vertical strip where the
1768: minimum is located makes it
1769: very difficult to converge to the minimum using gradient-based methods like the
1770: LM method. The derivatives cannot be correctly approximated when the $\chi^2$
1771: function has a large variation in such a small region of the space of
1772: parameters.
1773: % The DIRECT method needs about 1000 evaluations to locate the minimum,
1774: % in comparison with other cases in which it is found in less than $\sim$50
1775: % iterations.
1776: This shallow strip is produced by the quasi-degeneracy of the problem in both
1777: parameters. An infinity of combinations of both parameters give Stokes
1778: profiles
1779: that can approximately reproduce the observations. The difference in the
1780: $\chi^2$ merit function between
1781: these spurious cases and the exact one is very small. Two reasons produce
1782: such a behavior. On the one hand, the linear polarization signal is
1783: enhanced when the height is increased because the anisotropy of the radiation
1784: field
1785: increases (see Fig. \ref{fig:nbar_omega}, right panel).
1786: On the other hand, the Hanle effect turns
1787: out to be particularly efficient in reducing the atomic polarization when the magnetic field is significantly inclined with respect to the symmetry axis of the radiation field (the vertical
1788: direction).
1789: In a
1790: realistic case, the problem is much more complicated due to the presence of
1791: other additional
1792: parameters and the noise contamination.
1793:
1794: When the observed structure is off the limb, imaging techniques can be used to estimate $h$.
1795: On the contrary, the case of on-disk observations is much more complicated since no
1796: straightforward technique for estimating the height is available. One
1797: possibility is to follow the observed active region until it approaches the limb. The height
1798: can then be estimated if we assume that the height of the plasma structure has not changed
1799: between both observations. An even less precise procedure is to assume a given $h$ value
1800: based on the typical height of the solar structure type under study. Obviously, the ideal situation
1801: would be the one where $h$ could be inferred directly from the observed Stokes profiles.
1802: In order
1803: to investigate this possibility, we have performed an experiment in which the DIRECT method is used
1804: with disk-center ($\theta=0^\circ$) synthetic Stokes profiles. The emergent profiles have been calculated
1805: with $v_\mathrm{th}=8$ km s$^{-1}$, $\Delta \tau=0.8$, $h=20"$, $B=25$ G, $\theta_B=40^\circ$ and $\chi_B=19^\circ$,
1806: taking into account the effects of radiative transfer in the slab. We keep fixed
1807: all the parameters except for the inclination of the magnetic field $\theta_B$
1808: and the height $h$. The upper panels and the bottom left panel of Figure
1809: \ref{fig:height_diskcenter} present the points at which the
1810: DIRECT algorithm has evaluated the merit function, showing that it is possible
1811: to infer the height of the observation by only using the Stokes profiles. The shape
1812: of the $\chi^2$ surface is shown in the bottom right panel of Fig.
1813: \ref{fig:height_diskcenter}.
1814: In comparison with the off-limb case shown in Fig.
1815: \ref{fig:height_degeneration}, the minimum is located in a much less
1816: complicated region of the $\chi^2$ surface.
1817: The quasi-degeneracy present in
1818: the off-limb case is not present in the on-disk case.
1819: This is associated with the fact that the blue
1820: component gives no signal in the optically thin limit, while it does if an inclined field is
1821: present for the disk center case \citep{trujillo_nature02}.
1822:
1823: Interestingly, if one wants to infer the magnetic field vector and the height simultaneously
1824: from the observations, the code is unable to get a suitable global minimum, even in the noiseless
1825: case. However, an easily accessible global minimum exists when one of the parameters is kept
1826: fixed, thus inferring only the following combinations of
1827: parameters: ($B$, $\theta_B$, $h$), ($B$, $\chi_B$, $h$) and ($\theta_B$,
1828: $\chi_B$, $h$).
1829:
1830: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1831: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1832: % CONCLUSIONS
1833: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1834: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1835: \section{Conclusions}
1836: \label{sec:conclusions}
1837:
1838: The physical interpretation of spectropolarimetric observations of lines of
1839: neutral helium, such as those of the 10830 \AA\ and D$_3$ multiplets,
1840: represents a very important diagnostic window for investigating the dynamical
1841: behavior and the magnetic field of plasma structures in the solar chromosphere
1842: and corona, such as spicules, filaments, regions of emerging magnetic flux,
1843: network and internetwork regions, sunspots, flaring regions, etc. In order to
1844: facilitate this type of investigations we have developed a
1845: powerful forward modeling and inversion code that permits either to
1846: calculate the emergent spectral line intensity and polarization for any given
1847: magnetic field vector or to infer the dynamical and magnetic properties from the
1848: observed Stokes profiles. This diagnostic tool is based on the quantum theory of
1849: spectral line polarization \citep[see][]{landi_landolfi04}, which self-consistently
1850: accounts for the presence of atomic level polarization and the Hanle and Zeeman
1851: effects in the most general situation of the incomplete Paschen-Back effect regime. It is
1852: also of interest to mention that the same computer program can be easily applied
1853: to other chemical species apart from \ion{He}{1} (e.g., in order to
1854: investigate the magnetic sensitivity of the polarization caused by the joint
1855: action of the Hanle and Zeeman effects in many other spectral lines of
1856: diagnostic interest, both in the solar atmosphere and in other astrophysical plasmas).
1857:
1858: The influence of radiative transfer on the emergent spectral line radiation is
1859: taken into account by solving the Stokes-vector transfer equation in a slab of
1860: constant physical properties, including the magneto-optical terms of the
1861: propagation matrix. Although this ``cloud" model for the interpretation of
1862: polarimetric observations in such lines of \ion{He}{1} is suitable for inferring
1863: the magnetic field vector of plasma structures embedded in the solar
1864: chromosphere and corona, there are several interesting improvements and
1865: generalizations on which we are presently working on. The first one will be
1866: useful for improving the modeling of the Stokes profiles observed in
1867: low-lying optically-thick plasma structures embedded in the solar chromosphere,
1868: such as those of active region filaments. It consists in taking into account
1869: that in optically-thick plasma structures located at low atmospheric heights,
1870: the atomic level polarization is not going to be necessarily dominated by the
1871: anisotropic continuum radiation coming from the underlying solar photosphere (as
1872: we have assumed here), given that the radiation field generated by the
1873: optically-thick structure itself will tend to reduce the anisotropy factor of
1874: the true radiation field that pumps the helium atoms of the plasma structure
1875: under consideration \citep[see][]{trujillo_asensio07}. The second additional
1876: development consists in considering a Milne-Eddington atmospheric
1877: model, but determining consistently the height-dependent atomic level
1878: polarization induced by the anisotropic radiation field within the
1879: atmosphere model that provides the best fit to the observed
1880: Stokes profiles. Since the anisotropy factor is very sensitive to
1881: the source-function gradient \citep[e.g., Fig. 4 in][]{trujillo01} the solution of these type of problems in stratified model atmospheres may be facilitated by the application of efficient
1882: iterative schemes, such as those used by
1883: \cite{manso_trujillo03a,manso_trujillo03b} for developing a general multilevel radiative
1884: transfer program for modeling scattering line polarization and the Hanle effect in weakly
1885: magnetized stellar atmospheres.
1886:
1887: For the solution of the Stokes inversion problem we have applied an efficient
1888: algorithm based on global optimization methods, which permits a fast and
1889: reliable determination of the global minimum and facilitates the determination
1890: of the solutions corresponding to the unfamiliar Van-Vleck ambiguity. Our
1891: inversion approach is based
1892: on the application of the Levenberg-Marquardt (LM) method for locating
1893: the minimum of the merit function that quantifies the goodness of the fit between the observed and synthetic
1894: Stokes profiles. However, gradient-based methods suffer from convergence problems when the initial
1895: value of the parameters is not close to the minimum. In order to improve the convergence
1896: properties of the LM method, we have introduced a novel initialization technique. This
1897: method is based on the DIRECT algorithm, a deterministic global optimization scheme that
1898: performs very well. We have shown that a four-steps scheme using the DIRECT method to
1899: initialize the parameters and the LM method to refine the first estimation close to the
1900: minimum leads to a very robust technique.
1901:
1902: Our computer program has been
1903: developed with the aim of being computationally efficient and
1904: user-friendly. The relevant equations of the problem result from
1905: a general and robust theory, so that it is straightforward to
1906: treat limiting cases and include or discard several
1907: physical effects in a very transparent way. It is appropriate for
1908: its application to a wide variety of problems, from simple Zeeman-dominated Stokes profiles to more complex situations in which
1909: the influence of atomic level polarization cannot be neglected.
1910: The code is written in FORTRAN 90, and incorporates a user-friendly
1911: front-end based on IDL\footnote{\texttt{http://www.ittvis.com/idl}} which
1912: facilitates the execution and analysis of the synthesis and/or inversion calculations (see Fig. \ref{fig:front-end}).
1913:
1914: Obviously, our inversion strategy cannot
1915: compete in speed with algorithms based on look-up tables, like those applied by
1916: \cite{casini03} and \cite{merenda06}. At present, with a modern portable computer,
1917: we need of the order of 1 min. for the inversion of the Stokes profiles shown in
1918: Fig. \ref{fig:canopy_observation}. The strength of our approach is that it is very
1919: general and robust, and very suitable also to investigate the impact of the different
1920: physical mechanisms and parameters on the retrieved models. Concerning future improvements,
1921: we think that it would be worthwhile to treat the inversion problem within the framework of
1922: Bayesian inference techniques \citep[see][for a first application of such techniques
1923: to the inference of Milne-Eddington parameters from Stokes profiles induced
1924: by the Zeeman effect]{asensio_martinez_rubino07}. The
1925: aim is to sample the joint posterior probability distribution of the parameters of the model
1926: once the observation has been taken into account, and to carry out marginalizations
1927: to infer the probability distribution of each parameter. One of the main obstacles to overcome is to determine how
1928: to sample efficiently the full posterior probability distribution in the complex physical
1929: problem that we have investigated in this paper. A possible solution could be to rely on machine learning
1930: techniques for a fast solution of the forward problem, something that could be in perfect synergy
1931: with Markov Chain Montecarlo methods \citep{mackay03}.
1932:
1933: The reliability of the developments presented in this paper has been
1934: demonstrated through several model calculations and applications. Of particular
1935: interest is the investigation described in Section 3.4, which aimed at clarifying which is the
1936: optimum strategy for determining, from He {\sc i} 10830 \AA\
1937: spectropolarimetric observations, whether or not we have magnetic canopies with horizontal fields in the quiet solar chromosphere. The results of an aplication to an observation of
1938: a disk-center internetwork region can be found in
1939: \S5.4, which suggest the presence of magnetic fields inclined by no more than
1940: $50^{\circ}$ in the observed quiet chromospheric region.
1941:
1942: We have also discussed the potential problems that one may
1943: encounter. For example, we have investigated the presence
1944: of degeneracies, paying particular attention to the possibility of determining
1945: the height of the observed plasma structure from the observed Stokes profiles
1946: themselves and to demonstrate that the
1947: DIRECT method is a very efficient technique
1948: for detecting the solutions associated to the Van Vleck ambiguity.
1949:
1950: ``HAZEL" (an acronym from HAnle and ZEeman Light) is the name we have given to our IAC computer program for the synthesis and inversion of Stokes profiles resulting from the joint action of the Hanle and Zeeman effects in slabs of finite optical thickness. HAZEL will be continuously improved over the years (e.g., with extensions to more complicated radiative transfer models), but is now ready for systematic applications to a variety of spectropolarimetric observations in the spectral lines of the \ion{He}{1} 10830 \AA\ and D$_3$ multiplets. We offer it to the astrophysical community with the hope that it will help researchers to achieve new breakthroughs in solar and stellar physics. To get a copy, it suffices with making an e-mail request to the authors of this paper.
1951:
1952: \acknowledgments
1953: {\bf Acknowledgments}
1954: We thank Roberto Casini (HAO) for carefully reviewing of our paper.
1955: Finantial support by the Spanish Ministry of Education and Science through project AYA2007-63881 and by the European Commission through the SOLAIRE network (MTRN-CT-2006-035484) is gratefully acknowledged.
1956:
1957: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1958: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1959: % FIGURES
1960: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1961: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1962: \begin{figure}
1963: \plotone{f1.eps}
1964: \caption{The geometry for the scattering event. The $Z$-axis is placed along the vertical
1965: to the solar atmosphere. The magnetic field vector,
1966: $\mathbf{B}$,
1967: is characterized by its modulus $B$, the inclination angle $\theta_B$ and
1968: the azimuth $\chi_B$. The line-of-sight, indicated by the unit vector
1969: $\mathbf{\Omega}$,
1970: is characterized by the two angles $\theta$ and $\chi$.
1971: The reference direction for Stokes $Q$ is defined by the vector $\mathbf{e}_1$
1972: on the plane
1973: perpendicular to the line-of-sight. This vector makes an angle $\gamma$ with
1974: respect to the plane formed by
1975: the vertical and the line-of-sight. In the figures showing examples of the
1976: emergent Stokes profiles, our
1977: choice for the positive reference direction for Stokes $Q$ is $\gamma=90^\circ$, unless otherwise stated.
1978: For off-limb observations, we have $\theta=90^\circ$, while for observations
1979: on the solar disk, we have $\theta<90^\circ$. Note also that $\chi$ is generally taken to be $0^\circ$.
1980: \label{fig:geometry}}
1981: \end{figure}
1982:
1983: \clearpage
1984:
1985: \begin{figure}
1986: \plotone{f2.eps}
1987: \caption{Model atom of the triplet system of \ion{He}{1} used in this
1988: investigation. This work focuses on
1989: the polarization properties of the 10830 \AA\ multiplet
1990: between the 2p$^3$P and 2s$^3$S terms and on the D$_3$ multiplet between the
1991: 3d$^3D$ and
1992: 2p$^3$P terms. The energy of each $J$-level is taken from
1993: \cite{drake_helium98} and it
1994: is given in cm$^{-1}$ above the fundamental energy level (1s$^2 \, ^1S_0$). Note
1995: that the separation
1996: between the $J$-levels pertaining to each term is not drawn to scale.
1997: \label{fig:helium_atom}}
1998: \end{figure}
1999:
2000: \clearpage
2001:
2002: \begin{figure*}
2003: \plottwo{f3a.eps}{f3b.eps}
2004: \plottwo{f3c.eps}{f3d.eps}
2005: \caption{Energy splitting due to the presence of a magnetic field for the upper
2006: levels of the 10830 \AA\ multiplet
2007: (left panels) and for the upper levels of the D$_3$ multiplet at 5876 \AA\
2008: (right panels). Note that the
2009: regimes where level
2010: crossings and repulsions occur are different for the two terms. They can be
2011: better identified in the lower panels. The energy of each level is referred to
2012: the
2013: energy of the level with the smallest value of $J$ at zero magnetic field (i.e.,
2014: $J=0$ for the upper levels of the
2015: 10830 \AA\ multiplet and $J=1$ for the upper levels of the D$_3$ multiplet). The
2016: energy separation at $B=0$ G
2017: is obtained from the information presented in Fig. \ref{fig:helium_atom}.
2018: \label{fig:splitting}}
2019: \end{figure*}
2020:
2021: \clearpage
2022:
2023: \begin{figure*}
2024: \plottwo{f4a.eps}{f4b.eps}
2025: \caption{These two quantities (see Eqs. \ref{eq:nbar_omega}) characterize the radiation field that produces optical
2026: pumping processes in the \ion{He}{1} atoms.
2027: The number of photons per mode $\bar n$
2028: (\emph{left panel}) is proportional to the mean intensity, J$^0_0$, while the
2029: anisotropy factor
2030: $w$ (\emph{right panel}) is proportional to the $J^2_0$ tensor of
2031: the radiation field. Both quantities have been obtained using the
2032: limb-darkening data tabulated
2033: by \cite{pierce00}. The figures show also the variation of these quantities with
2034: the
2035: atmospheric height at which the slab of helium atoms is assumed to be located.
2036: Note that, while the number of photons per
2037: mode is almost insensitive to $h$, the geometrical effects produce a
2038: significant variation on the anisotropy factor.
2039: \label{fig:nbar_omega}}
2040: \end{figure*}
2041:
2042: \clearpage
2043:
2044: \begin{figure*}
2045: \includegraphics[width=\columnwidth]{f5.eps}
2046: \caption{Screen dump of the graphical front-end used for the synthesis.
2047: \label{fig:front-end}}
2048: \end{figure*}
2049:
2050: \clearpage
2051:
2052: \begin{figure}
2053: \plottwo{f6a.eps}{f6b.eps}
2054: \plottwo{f6c.eps}{f6d.eps}
2055: \plottwo{f6e.eps}{f6f.eps}
2056: \caption{Variation of the fractional population imbalances, $\rho^K_0(J)/\rho^0_0(J)$, and
2057: of the non-zero quantum coherences between
2058: magnetic sublevels pertaining to a given $J$-level, $\rho^K_Q(J)/\rho^0_0(J)$,
2059: for different values of the strength of a horizontal magnetic field. We show
2060: these quantities for the $J_u=1$ and $J_u=2$
2061: levels of the upper term of the 10830 \AA\ transition (upper and middle panel,
2062: respectively) and for the
2063: $J_l=1$ level of the lower term of the 10830 \AA\ transition (lower panel). The
2064: left panels
2065: show the results for the ``vertical'' reference frame in which the quantization
2066: axis is chosen
2067: along the symmetry axis of the radiation field, while the right panels show the
2068: results
2069: for the magnetic field reference frame in which the quantization
2070: axis is chosen along the (horizontal) magnetic field vector. Note that, at
2071: $B{\approx}0$ G, only population imbalances
2072: are present in the vertical reference frame, while both population imbalances
2073: and coherences are present in the
2074: magnetic reference frame. Note that in the magnetic field reference frame the
2075: quantum coherences are zero and $\rho^2_0(J)$ is constant for
2076: $10<B<100$ G.
2077: \label{fig:coherences}}
2078: \end{figure}
2079:
2080: \clearpage
2081:
2082: \begin{figure}
2083: \plottwo{f7a.eps}{f7b.eps}
2084: \plottwo{f7c.eps}{f7d.eps}
2085: \plottwo{f7e.eps}{f7f.eps}
2086: \caption{Same as Fig. \ref{fig:coherences}, but for the $J$-levels of the upper term of the He {\sc i} D$_3$ multiplet.
2087: \label{fig:coherencesD3}}
2088: \end{figure}
2089:
2090: \clearpage
2091:
2092: \begin{figure*}
2093: \plottwo{f8a.eps}{f8b.eps}
2094: \caption{Emergent fractional linear polarization at the solar disk
2095: center in the lines of the \ion{He}{1} 10830 \AA\ multiplet, assuming that a
2096: constant-property slab of helium atoms at a height of 3 arcseconds is permeated
2097: by a horizontal magnetic field of 10 G. The various $Q/I$ profiles of the left
2098: panel correspond to the slab's optical thickness indicated in the inset,
2099: calculated at the wavelength of the red blended component. The right panel
2100: indicates that the fractional polarization amplitudes of the red and blue
2101: components increase exponentially with the slab's optical thickness, till line saturation sets in.
2102: These calculations have been carried out using the exact solution of the
2103: radiative transfer equation given by Eq. (\ref{eq:slab_peo}). The
2104: positive reference direction for the definition of the Stokes $Q$ parameter lies along the horizontal magnetic field vector.
2105: \label{fig:canopies}}
2106: \end{figure*}
2107:
2108: \clearpage
2109:
2110: \begin{figure*}
2111: \plottwo{f9a.eps}{f9b.eps}
2112: \plottwo{f9c.eps}{f9d.eps}
2113: \caption{Variation of the linear polarization at the line center of the blue
2114: (left panel) and red (right panel)
2115: components of the \ion{He}{1} 10830 \AA\ multiplet for different inclinations of
2116: the magnetic
2117: field vector with respect to the local vertical and for different observing
2118: angles. The upper panels
2119: show Stokes $Q$ while the lower panels show Stokes $U$, normalized to the value
2120: of Stokes $I$ at the
2121: line center of each component. The calculations have been obtained
2122: assuming that a constant-property slab of helium atoms at a height of 3
2123: arcseconds is permeated
2124: by a magnetic field of 10 G with an inclination $\theta_B$ with respect to the
2125: solar
2126: local vertical direction. The slab's optical thickness at the wavelength of the
2127: red blended component is
2128: $\Delta \tau_\mathrm{red}=0.1$.
2129: The positive reference direction for Stokes $Q$ is along the projection of the
2130: magnetic field vector on the
2131: solar surface, which makes an angle of $90^{\circ}$ with any of the considered
2132: line-of-sights.
2133: \label{fig:canopies_qi_peak}}
2134: \end{figure*}
2135:
2136: \clearpage
2137:
2138: \begin{figure*}
2139: \plotone{f10.eps}
2140: \caption{This figure illustrates the philosophy of the DIRECT method for
2141: searching
2142: the region where the global minimum is
2143: located. In this case, we present an illustrative example in two dimensions.
2144: After the evaluation of
2145: the merit function at some selected points inside each region,
2146: the DIRECT algorithm decides, using the Lipschitz condition, which rectangles
2147: should
2148: be further subdivided in case they are potentially optimal (optimal means that a
2149: low value of the merit function has
2150: been found or that the rectangles are large and a finer sampling has to be
2151: carried out). This method rapidly
2152: finds the region where the minimum is located.
2153: \label{fig:direct_method}}
2154: \end{figure*}
2155:
2156: \clearpage
2157:
2158: \begin{figure}
2159: \plotone{f11.eps}
2160: \caption{The solar prominence case. Example of the best theoretical fit obtained
2161: with the inversion code to some of the
2162: observed Stokes profiles presented in \cite{merenda06}.
2163: The emergent Stokes profiles have been obtained by assuming an optically thin
2164: plasma. The inferred magnetic field vector is
2165: $B=26.8$ G, $\theta_B=25.5^\circ$ and $\chi_B=161^\circ$ and the
2166: inferred thermal velocity
2167: $v_\mathrm{th}=7.97$ km s$^{-1}$. All these values are in good agreement with
2168: the results obtained by \cite{merenda06}.
2169: The positive reference direction for Stokes $Q$ is the parallel to the solar
2170: limb.
2171: \label{fig:prominence}}
2172: \end{figure}
2173:
2174: \clearpage
2175:
2176: \begin{figure}
2177: \plotone{f12.eps}
2178: \caption{The case of solar chromospheric spicules. Example of the best
2179: theoretical fit obtained with the inversion code to the
2180: observed profiles presented in \cite{trujillo_merenda05}.
2181: The emergent Stokes profiles have been obtained by taking into account
2182: radiative transfer effects in a constant-property slab model. The inferred magnetic field vector
2183: is
2184: $B=2.6$ G, $\theta_B=37^\circ$ and $\chi_B=35^\circ$, although an equally good
2185: fit is obtained for other combinations like $B=10$ G, $\theta_B=37^\circ$ and $\chi_B=172^\circ$
2186: (see the text for more information). Furthermore, the
2187: inferred thermal velocity is
2188: $v_\mathrm{th}=13.9$ km s$^{-1}$, the slab's optical thickness in the red
2189: blended component is $\Delta \tau_{\rm red}=2.54$
2190: and the damping parameter is $a=0.22$. The positive reference direction for
2191: Stokes $Q$ is the parallel to the solar limb.
2192: \label{fig:spicules}}
2193: \end{figure}
2194:
2195: \clearpage
2196:
2197: \begin{figure*}
2198: \plottwo{f13a.eps}{f13b.eps}
2199: \caption{Values of the $\chi^2$-function for the possible fits of the Stokes
2200: profiles of the spicular material
2201: observed by \cite{trujillo_merenda05}, after considering several combinations of
2202: the inclination and azimuth of
2203: the magnetic field vector. We fixed the thermal velocity, the optical depth of
2204: the
2205: slab and the damping constant to the following values: $v_\mathrm{th}=13.9$
2206: km$s^{-1}$, $\Delta \tau_{\rm red}=2.54$
2207: and $a=0.22$. The magnetic field strength is 10 G in the left panel while it is
2208: 2.6 G in the right panel. Note
2209: the presence of several local
2210: minima. The white dots indicate the position of the global minimum of the
2211: $\chi^2$-function. The solutions $1$ and $2$ correspond to the Van-Vleck ambiguity. The same
2212: happens for the solutions $1'$ and $2'$.
2213: Additionally, the solutions $1$ and $1'$ are equivalent for the code even if they do produce a sign change in the
2214: Stokes $V$ signal because the observed Stokes $V$ profile is at the level of noise. The same happens for the
2215: solutions $2$ and $2'$.
2216: \label{fig:spicules_chi2}}
2217: \end{figure*}
2218:
2219: \clearpage
2220:
2221:
2222: \begin{figure*}
2223: \plotone{f14.eps}
2224: \caption{The case of a solar coronal filament. Example of the Stokes profiles of
2225: the \ion{He}{1} 10830 \AA\
2226: multiplet observed in a coronal filament
2227: at solar disk center. Our results confirm the conclusions of
2228: \cite{trujillo_nature02}, since we obtain
2229: $\Delta \tau=0.86$, $v_\mathrm{th}=6.6$ km s$^{-1}$, $a=0.19$,
2230: $B=18$ G and $\theta_B=105^\circ$.
2231: The positive direction of Stokes $Q$ is parallel to the projection of the
2232: magnetic field vector on the solar surface.
2233: \label{fig:filament_observation}}
2234: \end{figure*}
2235:
2236: \clearpage
2237:
2238: \begin{figure*}
2239: \plotone{f15.eps}
2240: \caption{The quiet chromosphere case. Example of the Stokes profiles of the
2241: \ion{He}{1} 10830 \AA\ multiplet
2242: observed in a disk center ($\mu=0.98$) internetwork
2243: region surrounded by an enhanced network region of almost circular shape. The
2244: observed Stokes profiles selected
2245: for the inversion correspond to a temporal and spatial average within the
2246: internetwork region. The solid
2247: line presents a
2248: fit to the observations which corresponds to a magnetic field vector with $B=35$
2249: G, $\theta_B=21^\circ$
2250: and $\chi_B=0^\circ$. The inferred optical depth in the red blended component is
2251: $\Delta {\tau}_{\rm red} \approx 0.2$, while $v_\mathrm{th}=9.2$ km s$^{-1}$ and $a=0.62$.
2252: An equally good fit is obtained for other field configurations like $B=47$ G,
2253: $\theta_B=47^\circ$ and
2254: $\chi_B=0^\circ$.
2255: The fact that with a ${\theta_B}<54.74^{\circ}$ the Stokes $Q$ signal of the
2256: red component is positive (after our rotation of the reference system to minimize Stokes
2257: $U$) indicates that the reference direction for Stokes $Q$ lies along the direction perpendicular to the
2258: projection of the magnetic field vector on the solar disk (see the $\mu=1$ case of Fig. 9).
2259: \label{fig:canopy_observation}}
2260: \end{figure*}
2261:
2262: \clearpage
2263:
2264: \begin{figure*}
2265: \plotone{f16.eps}
2266: \caption{The case of an emerging flux region. Example of the Stokes profiles of
2267: the \ion{He}{1}
2268: 10830 \AA\ multiplet in an emerging flux region located at $\mu=0.8$. The
2269: circles present the observed Stokes profiles obtained from Fig. 2 of
2270: \cite{Lagg04}, but after rotating the reference system by $-56^\circ$
2271: to have the reference direction
2272: for the Stokes $Q$ profile along the direction perpendicular to the straight line joining the
2273: disk center with the observed region. For this reason the inversion has been carried out
2274: using $\chi=0^{\circ}$ and $\gamma=90^{\circ}$ (see Fig. 1). The solid line presents
2275: the best fit obtained with our inversion code, taking into account the effect of atomic
2276: polarization on the emergent Stokes profiles, in addition to the Zeeman effect treated
2277: within the framework of the Paschen-Back effect theory. This full solution corresponds
2278: to a magnetic field vector with $B=1073$ G, $\theta_B=86^\circ$ and $\chi_B=170.7^\circ$,
2279: while $\Delta \tau_\mathrm{red} = 1.07$, $v_\mathrm{th}=7.24$ km s$^{-1}$ and $a=0.25$.
2280: The dotted line is the best fit obtained when neglecting atomic polarization. These
2281: results confirm the conclusion by \citet{trujillo_asensio07} that the presence of
2282: atomic polarization in the levels of the He {\sc i} 10830 \AA\ multiplet produces
2283: observable signatures on the emergent Stokes profiles even for fields as large as
2284: 1000 G. We point out that there is a typing error in the caption of figure 3 of
2285: \cite{trujillo_asensio07}, since the theoretical Stokes profiles in that figure
2286: were obtained for a LOS with $\mu=\cos \theta=0.8$ (i.e., not for the $\mu=1$
2287: forward scattering case) and with $\chi=\gamma=0^{\circ}$ (see Fig. 1).
2288: \label{fig:lagg_emerging}}
2289: \end{figure*}
2290:
2291: \clearpage
2292:
2293: \begin{figure*}
2294: \plotone{f17.eps}
2295: \caption{Value of $\theta_B$ and $\chi_B$ at which the merit function is
2296: evaluated
2297: for detecting possible ambiguities using the
2298: DIRECT method. The number $N$ of
2299: function evaluations is shown on the top of each panel. The two global minima
2300: seen in the figures result from the Van Vleck ambiguity,
2301: showing two different magnetic field vectors that produce the same emergent
2302: Stokes profiles. This figure illustrates the power of the
2303: DIRECT method, which allows to recover the two values of the magnetic field
2304: vector
2305: in this ambiguous case with less than 200
2306: function evaluations. We have shown only half of the space of parameters in
2307: order
2308: to avoid the 180$^\circ$ ambiguity.
2309: \label{fig:vanvleck_ambiguity}}
2310: \end{figure*}
2311:
2312: \clearpage
2313:
2314: \begin{figure*}
2315: \plottwo{f18a.eps}{f18b.eps}
2316: \plottwo{f18c.eps}{f18d.eps}
2317: \caption{Value of $\theta_B$ and $h$ at which the merit function is evaluated
2318: when
2319: using the
2320: DIRECT method for the inversion of the off-limb Stokes profiles indicated in the
2321: text
2322: (upper panels and bottom left panel). The number $N$ of
2323: function evaluations is shown on the top of each panel. The presence of a very
2324: deep and shallow strip where the minimum is located makes it difficult to obtain
2325: it.
2326: The DIRECT method needs at least 1000 function evaluations to find the
2327: region where the minimum is located.
2328: The bottom right panel shows the value of the $\chi^2$ merit function with the
2329: white dot
2330: indicating the combination of parameters that give the smallest value of the
2331: merit function.
2332: This is an example of a problem that poses severe difficulties to any
2333: gradient-based method like Levenberg-Marquardt.
2334: \label{fig:height_degeneration}}
2335: \end{figure*}
2336:
2337: \clearpage
2338:
2339: \begin{figure*}
2340: \plottwo{f19a.eps}{f19b.eps}
2341: \plottwo{f19c.eps}{f19d.eps}
2342: \caption{Value of $\theta_B$ and $h$ at which the merit function is evaluated
2343: when
2344: using the
2345: DIRECT method for the inversion of the on-disk Stokes profiles indicated in the
2346: text. The
2347: number $N$ of
2348: function evaluations is shown on the top of each panel (upper panels and bottom
2349: left panel). Even
2350: with only $N=50$, the DIRECT method is able to locate the
2351: region of the global minimum. When the number of evaluations increases, it
2352: converges towards the values $\theta_B=40^\circ$ and
2353: $h=20"$. The bottom right panel shows the value of the $\chi^2$ merit function
2354: with the white dot
2355: indicating the combination of parameters that give the smallest value of the
2356: merit function.
2357: \label{fig:height_diskcenter}}
2358: \end{figure*}
2359:
2360:
2361: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2362: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2363: % TABLES
2364: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2365: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2366: \clearpage
2367:
2368: \input{tab1.tex}
2369:
2370: \clearpage
2371:
2372: \input{tab2.tex}
2373:
2374: \clearpage
2375:
2376: \input{tab3.tex}
2377:
2378: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2379: % The bibliography
2380: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2381: % \bibliographystyle{apj}
2382: % \bibliography{apjmnemonic,../../biblio}
2383:
2384: \begin{thebibliography}{47}
2385: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
2386:
2387: \bibitem[{{Andretta} \& {Jones}(1997)}]{andretta_jones97}
2388: {Andretta}, V., \& {Jones}, H.~P. 1997, \apj, 489, 375
2389:
2390: \bibitem[{{Asensio Ramos} {et~al.}(2007){Asensio Ramos}, {Mart\'{\i}nez
2391: Gonz\'alez}, \& {Rubi\~no Mart\'{\i}n}}]{asensio_martinez_rubino07}
2392: {Asensio Ramos}, A., {Mart\'{\i}nez Gonz\'alez}, M.~J., \& {Rubi\~no
2393: Mart\'{\i}n}, J.~A. 2007, \aap, 476, 959
2394:
2395: \bibitem[{{Avrett} {et~al.}(1994){Avrett}, {Fontenla}, \& {Loeser}}]{avrett94}
2396: {Avrett}, E.~H., {Fontenla}, J.~M., \& {Loeser}, R. 1994, {Formation of the
2397: solar 10830 A line} (Infrared Solar Physics, IAU Symp.~No.~154,
2398: eds.~D.M.~Rabin, J.T.~Jefferies, and C.~Lindsey, Kluwer, Dordrecht,
2399: pp.~35-47), 35
2400:
2401: \bibitem[{{Bartholomew-Biggs} {et~al.}(2002){Bartholomew-Biggs}, {Parkhurst},
2402: \& {Wilson}}]{bartholomew02}
2403: {Bartholomew-Biggs}, M., {Parkhurst}, S., \& {Wilson}, S. 2002, Comp. Optim.
2404: Appl., 21, 311
2405:
2406: \bibitem[{{Belluzzi} {et~al.}(2007){Belluzzi}, {Trujillo Bueno}, \& {Landi
2407: Degl'Innocenti}}]{belluzzi07}
2408: {Belluzzi}, L., {Trujillo Bueno}, J., \& {Landi Degl'Innocenti}, E. 2007, \apj,
2409: 666, 588
2410:
2411: \bibitem[{{Bommier}(1980)}]{bommier80}
2412: {Bommier}, V. 1980, \aap, 87, 109
2413:
2414: \bibitem[{{Bommier} {et~al.}(1994){Bommier}, {Landi Degl'Innocenti}, {Leroy}, \&
2415: {Sahal-Brechot}}]{bommier94}
2416: {Bommier}, V., {Landi Degl'Innocenti}, E., {Leroy}, J.-L., \& {Sahal-Brechot}, S.
2417: 1994, \solphys, 154, 231
2418:
2419: \bibitem[{{Casini} \& {Judge}(1999)}]{casini_judge99}
2420: {Casini}, R., \& {Judge}, P.~G. 1999, \apj, 522, 524
2421:
2422: \bibitem[{{Casini} \& {Manso Sainz}(2005)}]{casini_manso05}
2423: {Casini}, R., \& {Manso Sainz}, R. 2005, \apj, 624, 1025
2424:
2425: \bibitem[{{Casini} {et~al.}(2005){Casini}, {Bevilacqua}, \& {L{\'o}pez
2426: Ariste}}]{casini05}
2427: {Casini}, R., {Bevilacqua}, R., \& {L{\'o}pez Ariste}, A. 2005, \apj, 622, 1265
2428:
2429: \bibitem[{{Casini} {et~al.}(2003){Casini}, {L{\'o}pez Ariste}, {Tomczyk}, \& {Lites}}]{casini03}
2430: {Casini}, R., {L{\'o}pez Ariste}, A., {Tomczyk}, S., \& {Lites}, B.~W. 2003, \apjl, 598, L67
2431:
2432: \bibitem[{{Centeno} {et~al.}(2006){Centeno}, {Collados}, \& {Trujillo
2433: Bueno}}]{centeno06}
2434: {Centeno}, R., {Collados}, M., \& {Trujillo Bueno}, J. 2006, \apj, 640, 1153
2435:
2436: \bibitem[{{Centeno} {et~al.}(2008){Centeno}, {Trujillo Bueno}, {Uitenbroek}, \&
2437: {Collados}}]{centeno08b}
2438: {Centeno}, R., {Trujillo Bueno}, J., {Uitenbroek}, H., \& {Collados}, M. 2008, \apj, 677, 742
2439:
2440: \bibitem[{{Collados} {et~al.}(2007){Collados}, {Lagg}, {D{\'{\i}}az
2441: Garc{\'{\i}}a}, {Hern{\'a}ndez Su{\'a}rez}, {L{\'o}pez L{\'o}pez}, {P{\'a}ez
2442: Ma{\~n}{\'a}}, \& {Solanki}}]{collados_tipII07}
2443: {Collados}, M., {Lagg}, A., {D{\'{\i}}az Garc{\'{\i}}a}, J.~J., {Hern{\'a}ndez
2444: Su{\'a}rez}, E., {L{\'o}pez L{\'o}pez}, R., {P{\'a}ez Ma{\~n}{\'a}}, E., \&
2445: {Solanki}, S.~K. 2007, in Astronomical Society of the Pacific Conference
2446: Series, Vol. 368, The Physics of Chromospheric Plasmas, ed. P.~{Heinzel},
2447: I.~{Dorotovi{\v c}}, \& R.~J. {Rutten}, 611
2448:
2449: \bibitem[{{Condon} \& {Shortley}(1935)}]{condon_shortley35}
2450: {Condon}, E.~U., \& {Shortley}, G.~H. 1935, The Theory of Atomic Spectra
2451: (Cambridge: Cambridge University Press)
2452:
2453: \bibitem[{{Cox} {et~al.}(2001){Cox}, {Haftka}, {Baker}, {Grossman}, {Mason}, \&
2454: {Watson}}]{cox01}
2455: {Cox}, S., {Haftka}, R., {Baker}, C., {Grossman}, B., {Mason}, W., \& {Watson},
2456: L. 2001, J. Global Optim., 21, 415
2457:
2458: \bibitem[{{Drake} \& {Martin}(1998)}]{drake_helium98}
2459: {Drake}, G. W.~F., \& {Martin}, W.~C. 1998, Can. J. Phys., 76, 679
2460:
2461: \bibitem[{{Edmonds}(1960)}]{edmonds60}
2462: {Edmonds}, A.~R. 1960, Angular Momentum in Quantum Mechanics (Princeton
2463: University Press)
2464:
2465: \bibitem[{{Gutowsky}(2004)}]{gutowsky04}
2466: {Gutowsky}, M.~W. 2004, in VII Domestic Conference on Evolutionary Algorithms
2467: and Global Optimization, cs.NE/0512019
2468:
2469: \bibitem[{{Harvey} \& {Hall}(1971)}]{harvey_hall71}
2470: {Harvey}, J., \& {Hall}, D. 1971, in IAU Symp. 43: Solar Magnetic Fields, ed.
2471: R.~{Howard}, 279
2472:
2473: \bibitem[{{Horst} \& {Pardalos}(1995)}]{global_optimization95}
2474: {Horst}, R., \& {Pardalos}, P.~M. 1995, Handbook of Global Optimization
2475: (Dordrecht: Kluwer Academic Publishers)
2476:
2477: \bibitem[{{House}(1977)}]{house77}
2478: {House}, L.~L. 1977, \apj, 214, 632
2479:
2480: \bibitem[{{Jones} {et~al.}(1993){Jones}, {Perttunen}, \&
2481: {Stuckmann}}]{Jones_DIRECT93}
2482: {Jones}, D.~R., {Perttunen}, C.~D., \& {Stuckmann}, B.~E. 1993, Journal of
2483: Optimization Theory and Applications, 79, 157
2484:
2485: \bibitem[{{Lagg} (2007)}]{Lagg07}
2486: {Lagg}, A. 2007, Advances in Space Research 39, 1734
2487:
2488: \bibitem[{{Lagg} {et~al.}(2004){Lagg}, {Woch}, {Krupp}, \& {Solanki}}]{Lagg04}
2489: {Lagg}, A., {Woch}, J., {Krupp}, N., \& {Solanki}, S.~K. 2004, \aap, 414, 1109
2490:
2491: \bibitem[{{Landi Degl'Innocenti}(1982)}]{landi_d3_82}
2492: {Landi Degl'Innocenti}, E. 1982, \solphys, 79, 291
2493:
2494: \bibitem[{{Landi Degl'Innocenti} \& {Landi Degl'Innocenti} (1985)}]{landi_landi_85}
2495: {Landi Degl'Innocenti}, E., {Landi Degl'Innocenti}, M. 1985, \solphys, 97, 239
2496:
2497: \bibitem[{{Landi Degl'Innocenti} \& {Bommier}(1993)}]{landi_bommier93}
2498: {Landi degl'Innocenti}, E., \& {Bommier}, V. 1993, \apjl, 411, L49
2499:
2500: \bibitem[{{Landi Degl'Innocenti} \& {Landolfi}(2004)}]{landi_landolfi04}
2501: {Landi Degl'Innocenti}, E., \& {Landolfi}, M. 2004, Polarization in Spectral
2502: Lines (Kluwer Academic Publishers)
2503:
2504: \bibitem[{{Lin} {et~al.}(1998){Lin}, {Penn}, \& {Kuhn}}]{lin98}
2505: {Lin}, H., {Penn}, M.~J., \& {Kuhn}, J.~R. 1998, \apj, 493, 978
2506:
2507: \bibitem[{{Ljungberg} {et~al.}(2004){Ljungberg}, {Holmgren}, \&
2508: {Carlborg}}]{ljungberg04}
2509: {Ljungberg}, L., {Holmgren}, S., \& {Carlborg}, {\"O}. 2004, Bioinformatics,
2510: 20, 1887
2511:
2512: \bibitem[{{L{\'o}pez Ariste} \& {Casini}(2002)}]{arturo_casini02}
2513: {L{\'o}pez Ariste}, A., \& {Casini}, R. 2002, \apj, 575, 529
2514:
2515: \bibitem[{{L{\'o}pez Ariste} \& {Casini}(2005)}]{lopezariste_casini05}
2516: ---. 2005, \aap, 436, 325
2517:
2518: \bibitem[{{MacKay}(2003)}]{mackay03}
2519: {MacKay}, D. J. C. 2003, Information Theory, Inference, and Learning Algorithms (Cambridge University Press)
2520:
2521: \bibitem[{{Manso Sainz} \& {Trujillo Bueno}(2003{\natexlab{a}})}]{manso_trujillo03a}
2522: {Manso Sainz}, R., \& {Trujillo Bueno}, J. 2003{\natexlab{a}}, in ASP Conf. Ser. 307: Solar Polarization 3, ed. J.~{Trujillo
2523: Bueno} \&
2524: J. ~{S\'anchez Almeida}, 251
2525:
2526: \bibitem[{{Manso Sainz} \& {Trujillo Bueno}(2003{\natexlab{b}})}]{manso_trujillo03b}
2527: {Manso Sainz}, R., \& {Trujillo Bueno}, J. 2003{\natexlab{b}}, \prl, 91, 111102
2528:
2529: \bibitem[{{Mart{\'{\i}}nez Pillet} {et~al.}(1999){Mart{\'{\i}}nez Pillet},
2530: {Collados}, {Bellot Rubio}, {Rodr{\'{\i}}guez Hidalgo}, {Ruiz Cobo}, \&
2531: {Soltau}}]{martinez_pillet99}
2532: {Mart{\'{\i}}nez Pillet}, V., {Collados}, M., {Bellot Rubio}, L.~R.,
2533: {Rodr{\'{\i}}guez Hidalgo}, I., {Ruiz Cobo}, B., \& {Soltau}, D. 1999, in
2534: Astronomische Gesselschaft Meeting Abstracts, vol. 15
2535:
2536: \bibitem[{{Merenda} {et~al.}(2006){Merenda}, {Trujillo Bueno}, {Landi
2537: Degl'Innocenti}, \& {Collados}}]{merenda06}
2538: {Merenda}, L., {Trujillo Bueno}, J., {Landi Degl'Innocenti}, E., \& {Collados},
2539: M. 2006, \apj, 642, 554
2540:
2541: \bibitem[{{Pierce}(2000)}]{pierce00}
2542: {Pierce}, K. 2000, in Allen's Astrophysical Quantities, ed. A. N. Cox (New
2543: York: Springer Verlag and AIP Press)
2544:
2545: \bibitem[{{Press} {et~al.}(1986){Press}, {Teukolsky}, {Vetterling}, \&
2546: {Flannery}}]{numerical_recipes86}
2547: {Press}, W.~H., {Teukolsky}, S.~A., {Vetterling}, W.~T., \& {Flannery}, B.~P.
2548: 1986, Numerical Recipes (Cambridge: Cambridge University Press)
2549:
2550: \bibitem[{{Querfeld} {et~al.}(1985)}]{querfeld85}
2551: {Querfeld}, C.~W., {Smartt}, R.~N., {Bommier}, V.,
2552: {Landi Degl'Innocenti}, E., {House}, L.~L. 1985, \solphys, 96, 277
2553:
2554: \bibitem[{{Ramelli} {et~al.}(2006{\natexlab{a}}){Ramelli}, {Bianda}, {Merenda},
2555: \& {Trujillo Bueno}}]{ramelli06_2}
2556: {Ramelli}, R., {Bianda}, M., {Merenda}, L., \& {Trujillo Bueno}, T.
2557: 2006{\natexlab{a}}, in ASP Conf. Ser., Vol. 358, Solar Polarization 4, ed.
2558: R.~{Casini} \& B.~W. {Lites}, 448
2559:
2560: \bibitem[{{Ramelli} {et~al.}(2006{\natexlab{b}}){Ramelli}, {Bianda}, {Trujillo
2561: Bueno}, {Merenda}, \& {Stenflo}}]{ramelli06}
2562: {Ramelli}, R., {Bianda}, M., {Trujillo Bueno}, J., {Merenda}, L., \& {Stenflo},
2563: J.~O. 2006{\natexlab{b}}, in ASP Conf. Ser., Vol. 358, Solar Polarization 4,
2564: ed. R.~{Casini} \& B.~W. {Lites}, 471
2565:
2566: \bibitem[{{Rees} et al. (1989)}]{rees89}
2567: {Rees}, D.~E., {Murphy}, G.~A. \& {Durrant}, C.~J. 1989, \apj, 339, 1093
2568:
2569: \bibitem[{{R{\"u}edi} {et~al.}(1996){R{\"u}edi}, {Keller}, \&
2570: {Solanki}}]{ruedi96}
2571: {R{\"u}edi}, I., {Keller}, C.~U., \& {Solanki}, S.~K. 1996, \solphys, 164, 265
2572:
2573: \bibitem[{{Socas-Navarro} \& {Elmore}(2005)}]{socas_elmore05}
2574: {Socas-Navarro}, H., \& {Elmore}, D. 2005, \apj, 619, L195
2575:
2576: \bibitem[{{Socas-Navarro} {et~al.}(2004)}]{socas_trujillo04}
2577: {Socas-Navarro}, H., {Trujillo Bueno}, J., \& {Landi Degl'Innocenti}, E. 2004, \apj, 612, 1175
2578:
2579: \bibitem[{{Trujillo Bueno}(2001)}]{trujillo01}
2580: {Trujillo Bueno}, J. 2001, in ASP Conf. Ser. 236: Advanced Solar Polarimetry --
2581: Theory, Observation, and Instrumentation, ed. M.~{Sigwarth}, 161
2582:
2583: \bibitem[{{Trujillo Bueno}(2003)}]{trujillo03}
2584: {Trujillo Bueno}, J. 2003, in Stellar Atmosphere Modeling, ed. I.~{Hubeny},
2585: D.~{Mihalas}, \& K.~{Werner}, ASP Conf. Ser. 288 (San Francisco: ASP), 551
2586:
2587: \bibitem[{{Trujillo Bueno}(2005)}]{trujillo_esa05}
2588: {Trujillo Bueno}, J. 2005, in ESA SP-600: The Dynamic Sun: Challenges for
2589: Theory and Observations, ed. D.~{Danesy}, S.~{Poedts}, A.~{De Groof}, \&
2590: J.~{Andries}, 7
2591:
2592: \bibitem[{{Trujillo Bueno} \& {Asensio Ramos}(2007)}]{trujillo_asensio07}
2593: {Trujillo Bueno}, J., \& {Asensio Ramos}, A. 2007, \apj, 655, 642
2594:
2595: \bibitem[{{Trujillo Bueno} \& {Shchukina}(2007)}]{trujillo_shchukina07}
2596: {Trujillo Bueno}, J., \& {Shchukina}, N. 2007, \apj, 664, L135
2597:
2598: \bibitem[{{Trujillo Bueno} {et~al.}(2002{\natexlab{a}}){Trujillo Bueno},
2599: {Casini}, {Landolfi}, \& {Landi Degl'Innocenti}}]{trujillo_casini02}
2600: {Trujillo Bueno}, J., {Casini}, R., {Landolfi}, M., \& {Landi Degl'Innocenti},
2601: E. 2002{\natexlab{a}}, \apjl, 566, L53
2602:
2603: \bibitem[{{Trujillo Bueno} {et~al.}(2002{\natexlab{b}}){Trujillo Bueno}, {Landi
2604: Degl'Innocenti}, {Collados}, {Merenda}, \& {Manso Sainz}}]{trujillo_nature02}
2605: {Trujillo Bueno}, J., {Landi Degl'Innocenti}, E., {Collados}, M., {Merenda},
2606: L., \& {Manso Sainz}, R. 2002{\natexlab{b}}, \nat, 415, 403
2607:
2608: \bibitem[{{Trujillo Bueno} {et~al.}(2005){Trujillo Bueno}, {Merenda},
2609: {Centeno}, {Collados}, \& {Landi Degl'Innocenti}}]{trujillo_merenda05}
2610: {Trujillo Bueno}, J., {Merenda}, L., {Centeno}, R., {Collados}, M., \& {Landi
2611: Degl'Innocenti}, E. 2005, \apj, 619, L191
2612:
2613: \bibitem[{{Wiese} {et~al.}(1966){Wiese}, {Smith}, \& {Glennon}}]{wiese_nist66}
2614: {Wiese}, W.~L., {Smith}, M.~W., \& {Glennon}, B.~M. 1966, Natl. Stand. Ref.
2615: Data Ser., Vol. I ({Natl Bur. Stand. (U.S.), NSRDS-NBS 4})
2616:
2617: \end{thebibliography}
2618:
2619:
2620: \end{document}
2621: