0810.3613/ms.tex
1: \documentclass[12pt,preprint]{aastex}
2: 
3: %\slugcomment{Not to appear in Nonlearned J., 45.}
4: 
5: \shorttitle{Wave propagation in magnetic structures}
6: \shortauthors{Centeno et al.}
7: 
8: 
9: 
10: \begin{document}
11: 
12: \title{Wave propagation and shock formation in different magnetic structures}
13: 
14: \author{R. Centeno\altaffilmark{1,2}, M. Collados\altaffilmark{2}, J. Trujillo Bueno\altaffilmark{2,3}}
15: \altaffiltext{1}{High Altitude Observatory (NCAR) The National Center for Atmospheric Research is sponsored by the National Science Foundation, Boulder CO 80301, USA}
16: \altaffiltext{2}{Instituto de Astrof\'\i sica de Canarias, 38205, La Laguna, Tenerife (Spain)}
17: \altaffiltext{3}{Consejo Superior de Investigaciones Cient\'\i ficas (Spain)}
18: \email{rce@ucar.edu, mcv@iac.es, jtb@iac.es}
19: 
20: 
21: \begin{abstract}
22: 
23: 
24: Velocity oscillations "measured" simultaneously at the photosphere and
25: the chromosphere -from time series of spectropolarimetric data in the 
26: 10830 \AA\ region- of different solar magnetic features allow us to study 
27: the properties of wave propagation as a function of the magnetic flux of 
28: the structure (i.e. two different-sized sunspots, a tiny pore and a facular region). While photospheric oscillations have similar characteristics 
29: everywhere, oscillations measured at chromospheric heights show different 
30: amplitudes, frequencies and stages of shock development depending on the 
31: observed magnetic feature. The analysis of the power and the phase spectra, 
32: together with simple theoretical modeling, lead to a series of results 
33: concerning wave propagation within the range of heights of this study. 
34: We find that, while the atmospheric cut-off frequency and the propagation 
35: properties of the different oscillating modes depend on the magnetic feature, 
36: in all the cases the power that reaches the high chromosphere above the 
37: atmospheric cut-off comes directly from the photosphere by means of linear 
38: vertical wave propagation rather than from non-linear interaction of modes.
39: 
40: \end{abstract}
41: 
42: 
43: \keywords{Sun : photosphere, Sun : chromosphere, Sun : magnetic fields, techniques : polarimetric, shock waves}
44: 
45: 
46: 
47: \section{Introduction}
48: 
49: 
50: Encrypted in the oscillatory behavior of the solar atmosphere lies crucial information
51: for understanding its dynamical and physical properties. The stratification caused by gravity, together with the presence of
52: magnetic fields, leads to a variety of magneto-gravity-acoustic modes \citep{khomenko2006, abdelatif, thomas}.
53: By analyzing the local properties of oscillations and wave propagation we can 
54: infer information about the stratification and dynamics of different atmospheric 
55: structures. 
56: 
57: \noindent Wave propagation is an efficient means of carrying energy 
58: between different layers of the atmosphere and of dissipating it efficiently
59: through the formation and breaking of shocks \citep{mihalas}. Although acoustic (slow-mode)
60: heating is not important for the upper atmosphere (Athay \& White, 1979; White \& Athay, 1979), 
61: the role that wave propagation plays in the heating problem is still
62: one of the most challenging open debates amongst solar physics researchers.  
63: 
64: 
65: The first measurements of oscillations in the quiet Sun 
66: \cite{leighton} came nearly a decade before the first 
67: detection of sunspot oscillations were reported \citep{beckers}.
68: Since then, many works have tried to put the pieces of the puzzle together
69: \cite[see][for a comprehensive review
70: of the literature of wave propagation in sunspots]{lites1992, bogdan2006}.
71: 
72: \noindent Oscillations above sunspots show a 5-minute periodicity in the photosphere.
73: However, at chromospheric levels and higher up, this picture changes into a 3-minute
74: saw-tooth pattern \citep[see e.g.][]{lites1986, lites1986b, sn2000,collados2001, bry2003, bry2004}.
75: Several hypotheses have attempted to explain the origin of the chromospheric 3-min
76: oscillations. Zhugzhda, Locans \& Staude (1985) proposed a model with a resonant 
77: chromospheric cavity that would explain the presence of multiple peaks in the
78: chromospheric power spectrum together with the  3 minute period. On the other hand, Gurman
79: \& Leichbacher (1984) suggested an origin based on the non-linear interaction of photospheric modes. 
80: Centeno, Collados and Trujillo Bueno (2006a) showed that the chromospheric 3-minute signal in the umbrae of sunspots is a result
81: of linear wave propagation of the photospheric perturbations in the 6 mHz range,
82: thus ruling out the non-linear origin for these oscillations.
83: 
84: 
85: 
86: Observations above facular and network areas report 5-minute oscillations
87: at chromospheric heights \citep[see e.g.][]{krijger, lites93, depontieu2003a, centeno2006a}
88: and higher up. For this to happen, the atmospheric
89: cut-off frequency must be reduced, allowing the "evanescent" photospheric 
90: 5-minute oscillations propagate into the chromosphere. De Pontieu et al. (2003, 2004)
91: suggested that the inclination of magnetic fields plays an important role in
92: this p-mode leakage with enough energy to give rise to the dynamic
93: jets that are observed in active region fibrils. With this idea in mind,
94: Jefferies et al. (2006) suggested that inclined flux tubes might explain
95: the observed properties of waves at chromospheric heights.
96: 
97: The effective cut-off frequency is lowered by the cosine
98: of the inclination angle with respect to the local vertical. Thus, if the magnetic
99: field is sufficiently inclined, the flux tubes will channel the photospheric 
100: 5-minute perturbations all the way up into the chromosphere and corona. An alternate 
101: possibility takes into account a departure from adiabaticity due to radiative
102: losses, which results in a reduced cut-off frequency \cite{roberts, centeno2006a, khomenko2008} that allows the 5-minute modes to propagate.
103: 
104: 
105: 
106: Simultaneous time-series
107: observations of various spectral lines that sample different regions of the 
108: solar atmosphere is one of the most useful techniques for studying wave 
109: propagation. 
110: For the analysis that follows we measure simultaneously the full Stokes vector of the 
111: photospheric Si {\sc i} \hbox{10827 \AA} line and of the chromospheric 
112: He {\sc i} \hbox{10830 \AA} multiplet on different magnetic targets. The analysis of the photospheric and 
113: chromospheric LOS velocity oscillations and the relation between them give 
114: us information about the behavior of the atmosphere: the propagation of 
115: photospheric disturbances, the amplification of the oscillations as they 
116: travel towards higher layers of the atmosphere, the cut-off frequency below which the 
117: oscillation modes do not propagate, the development of shocks and so on. 
118: We carry out a comparative study among 4 magnetic structures with 
119: different magnetic fluxes: a rather big sunspot (analyzed in Centeno et al. 2006a,
120: hereafter Paper I), a smaller one, a pore that has developed no penumbra and a 
121: facular region.
122: The physical properties of the atmosphere in each case determines how the 
123: propagation of the photospheric perturbations takes place and viceversa.
124: 
125: 
126: The outline of the paper is as follows: after a brief description of the observations
127: and the inversion procedures in Section 2, Section 3 continues to describe the properties
128: of the oscillations measured at photospheric and chromospheric heights.
129: Section 4 analyzes the relations between them, which are determined by the propagation
130: properties of the atmosphere in each case. The comparison with a simple model of linear vertical 
131: wave propagation in a magnetized atmosphere with radiative losses will yield to a
132: qualitative understanding of the origin of the chromospheric oscillations in the
133: analyzed magnetic structures. 
134: A brief discussion of the results is presented in Section 5.
135: 
136: 
137: 
138: 
139: \section{Observations}
140: 
141: \clearpage
142: \begin{figure}[t!]
143: \center 
144: \includegraphics[width=5cm]{f1a.ps} 
145: \includegraphics[width=5cm]{f1b.ps} 
146: \includegraphics[width=5cm]{f1c.ps} 
147: \caption{White light slitjaw images for the small sunspot and the pore, and Ca K image for the
148: facular region. In all the cases, the vertical line corresponds to the shadow of the spectrograph
149: slit (which was kept fixed during each run) and horizontal lines delimit the field of
150: view of the polarimeter ($\sim 40 \arcsec$). Of the three pores showed in the second image, 
151: we chose the central one to carry out the analysis presented in this paper.}
152: \label{fig:slitjaws}
153: \end{figure}
154: \clearpage
155: 
156: The observations were carried out at the German Vacuum Tower Telescope (VTT) 
157: of the Observatorio del Teide between September 2000 and June 2002, using the 
158: Tenerife Infrared Polarimeter \cite[TIP,][]{vmp-tip}. 
159: This instrument allowed us to measure simultaneously the four Stokes profiles for
160: all the spatial positions along the spectrograph slit at each scanning step.
161: The slit (0\farcs5 wide 
162: and 40\arcsec\ long) was placed over the targets and was kept fixed during 
163: the entire observing run (approx. 1 hour for each target). The image stability throughout 
164: the time series was achieved by using a correlation tracker device
165:  (Schmidt \& Kentischer 1995; Ballesteros et al. 1996)
166: which compensated, to first order, for the solar rotation as well as for the image motion 
167: induced by the Earth's high frequency atmospheric variability, thus minimizing the image jitter. Despite of being observed at very different moments, the quality is similar throughout the 
168: four datasets, and the instrument components and configuration were identical in all cases. 
169: The spatial resolution varies roughly between $1\arcsec$ and $1\arcsec.5$ (the worst data-set
170: being the one that corresponds to the smalls sunspot).
171: Details on the 4 data-sets analyzed in this paper are given in Table \ref{tab:data}. Fig. 
172: \ref{fig:slitjaws} shows the slit-jaw images for the small sunspot, the pore and the facula.
173: The size of the structure refers to the umbra in the case of the sunspots and to the 
174: enhanced bright region as seen in the Ca K slit-jaw in the case of the facula.
175: The data-set corresponding to the big sunspot was one of those used throughout Paper I.
176: We chose the more regular and homogeneous of the two sunspots and we only included it here for 
177: a sake of completeness, to compare it with the results from the other 3 magnetic structures.
178: 
179: 
180: Flat-field and dark current measurements were performed at
181: the beginning and the end of all observing runs, and, in order to
182: compensate for the telescope instrumental polarization, we also
183: took a series of polarimetric calibration images. The calibration
184: optics \cite[see][]{collados99} allows us to obtain the Mueller matrix of
185: the light path between the instrumental calibration subsystem
186: and the polarimeter.
187: This process leaves a section of the telescope
188: without being calibrated, so further corrections of the residual
189: cross-talk among Stokes parameters were done: the I to Q, U, and
190: V cross-talk were removed by forcing the continuum polarization
191: to zero, and the circular and linear polarization mutual cross-talks
192: were estimated by means of statistical techniques \cite{collados2003}.
193: 
194: 
195: \clearpage
196: \begin{table}[!t]
197: \begin{center}
198: \begin{tabular}{c|cccc}
199:  & big sunspot & small sunspot & pore & facular region\\
200: \hline
201: Date & May 9, 2001 & Sep 30, 2000 & Jun 13, 2002 & Jun 14, 2002 \\
202: Pos X [$\arcsec$] & -392 & 557 & -339 & -90 \\
203: Pos Y [$\arcsec$] & -163 & 102 & 306 & -291 \\
204: $\mu = {\rm cos} \theta$ & 0.89 & 0.81 & 0.88 & 0.94 \\
205: Duration [s] & 4214 & 3922 & 4842  & 4873\\
206: Cadence [s] & 2.1 & 7.9 & 5.4 & 5.4\\
207: Noise & $2 \cdot 10^{-3}$ & $7 \cdot 10^{-3}$&$2 \cdot 10^{-3}$ & $ 10^{-3}$\\
208: Size [$\arcsec$] & 16 & 10 & 4 & 30 \\
209: \label{tab:data}
210: \end{tabular}
211: \caption{Details of the four data-sets obtained with TIP. Positions X and Y represent terrestrial E-W and N-S directions and are measured from sun center.
212: }
213: \end{center}
214: \end{table}
215: \clearpage
216: 
217: The observed spectral range spanned from 10825 to 10833 \AA, with a 
218: spectral sampling of 31 m\AA\ per pixel. This spectral region is a powerful 
219: diagnostic window for the solar atmospheric properties since it contains 
220: valuable information coming from two different layers in the atmosphere. 
221: It includes a photospheric Si {\sc i} line at 10827 \AA\ and a chromospheric 
222: He {\sc i} triplet centered around 10830 \AA. 
223: The Si line is formed in the high photosphere. The response function (Ruiz Cobo \& del
224: Toro Iniesta, 1994) of the intensity profile to the temperature shows a height of formation
225: between 300 and 540 km above the base of the photosphere (Bard and Carlsson, 2008).
226: The He multiplet is formed in the high chromosphere \cite{avrett, knoelker, bruno07, centeno2008},
227: although the exact location depends critically on the atmospheric stratification and
228: the coronal illumination coming from above, that triggers the formation of the multiplet. 
229: Thus, the difference in height between the photospheric
230: and the chromospheric indicators ranges between 1000 and 1500 km. 
231: The He triplet serves as a unique diagnostic tool for 
232: chromospheric magnetic fields \cite[see][for a recent review]{lagg2007}.
233: 
234: In order to infer the physical parameters of the magnetized 
235: atmosphere in which the measured spectral lines were generated, we carried 
236: out the full Stokes inversion of both the silicon line and the helium triplet 
237: for the whole time series of observations and for all four data-sets. The Si
238: line was treated in Local Thermodynamic Equilibrium (LTE) and inverted with
239: the code LILIA (Socas-Navarro 2001). This inversion code yields the line-of-sight (LOS) 
240: velocity, magnetic field, temperature, density and electron
241: pressure stratification of the atmosphere in the layers where the spectral line
242: radiation is generated.
243: The observations of the He {\sc i} triplet were interpreted with our Milne-Eddington 
244: inversion code of Stokes profiles induced by the Zeeman effect, which is a suitable
245: strategy for extracting information on the LOS velocity and gives a reliable estimation
246: of the field strength from the full Stokes vector (see Trujillo Bueno \& Asensio Ramos 2007; 
247: Centeno et al, in press). We did not consider the incomplete Paschen-Back effect
248: in the modeling, so the magnetic field strength is underestimated by up to a 20\%
249: \cite{sasso}. However, in the analysis carried out in this paper we focus on the velocities,
250: which are not affected by this.
251: By inverting the whole time series we were able to obtain the temporal variability of several physical quantities (line-of-sight velocity, magnetic field intensity and orientation, 
252: \dots) at the photosphere and chromosphere of the four magnetic structures.
253: Both inversion codes took into account only one atmospheric component (one velocity and one
254: magnetic field value).
255: Only in the case of the facula, a stray light component was included to account for the 
256: non-magnetic part of the spectral profiles.
257: 
258: 
259: \noindent The magnetic field values yielded by the inversions are given in the line-of-sight (LOS)
260: reference system (which depends on the position of the target on the solar disk).
261: The azimuth origin, defined by the polarimetric calibration optics of the telescope,
262: is referenced to the Earth's North-South direction.
263: In order to determine how vertical the inferred magnetic fields are, we have to transform them
264: to the local vertical reference frame. The $180^{\circ}$ ambiguity in the azimuth 
265: leads to two possible inclination values in the new reference system. The flotability of
266: strong magnetic flux tubes can be used as a physical argument to choose the more vertical 
267: solution over the other option.
268: The photospheric magnetic fields obtained from the inversion of the Si line
269: turn out to be very vertical (with a range of inclinations between $0^{\circ}$ and 
270: $20^{\circ}$) in all the cases.
271: The He lines show barely any linear polarization at all due to the weaker magnetic field
272: regime in the high chromosphere.
273: 
274: 
275: \noindent Throughout the rest of this paper we will focus on the results concerning the 
276: LOS velocity oscillations, which are the line-of-sight projection of the plasma movements
277: along the magnetic field lines. 
278: All the targets are relatively close to disk center (the farther one having a heliocentric
279: angle of $\mu={\rm cos}\,\theta=0.81$), so this means that the maximum projection effect would happen for the
280: small sunspot, for which the LOS forms an angle of $\sim 35^{\circ}$ with the local vertical.
281: Taking into account the estimated height difference between the formation of the photospheric 
282: and the chromospheric indicators ($\sim1000 - 1500$ km), this angle would lead to a 
283: maximum projected horizontal displacement (for two positions on the same vertical) of some
284: $900$ km. The spatial resolution in our data was limited by seeing, which
285: we estimated to be of the order of $\sim1 - 1.5\arcsec$, so the displacement due to projection 
286: effects will be barely noticeble in the worst case.
287: 
288: \section{Oscillations and Shock Waves}
289: 
290: \clearpage
291: \begin{figure}[t!]
292: \center 
293: \includegraphics[width=7cm]{f2.ps} 
294: \caption{Photospheric oscillations for one position inside the umbra of
295: sunspot \#1. The rest of the structures analyzed in this work show a 
296: qualitatively similar behavior at photospheric levels.}
297: \label{fig:photosphere}
298: \end{figure}
299: 
300: \begin{figure}[t!]
301: \center 
302: \includegraphics[width=7cm]{f3a.ps} 
303: \includegraphics[width=7cm]{f3b.ps} 
304: \includegraphics[width=7cm]{f3c.ps} 
305: \includegraphics[width=7cm]{f3d.ps} 
306: \caption{Chromospheric oscillations in different magnetic features as
307: seen by the He {\sc i} 10830 \AA\ triplet. 
308: From left to right and top to bottom, the panels depict a typical 
309: velocity profile for one position inside the umbra of a big sunspot, the 
310: umbra of small sunspot, a pore and a facula. In all the cases, 
311: asterisks represent the measured values, equispaced in time.}
312: \label{fig:chrom-oscil}
313: \end{figure}
314: \clearpage
315: 
316: Photospheric velocity\footnote{The sign convention used throughout this 
317: paper is such that negative velocities correspond to material approaching 
318: the observer along the line-of-sight.} oscillations 
319: (see Fig.~\ref{fig:photosphere}), 
320: retrieved from the inversion of the Si {\sc i} line, show the same 
321: characteristics everywhere: a typical 5-min period, 300--400 m\,s$^{-1}$ 
322: peak to peak amplitudes and fairly sinusoidal patterns. 
323: 
324: \noindent On the other hand, chromospheric velocity oscillations (encoded in
325: the Doppler shift of the He {\sc i} triplet) show very different
326: behavior depending on the magnetic structure. Fig.~\ref{fig:chrom-oscil} 
327: depicts a detail of the chromospheric oscillation pattern in the four
328: regions analyzed. From left to right and top to bottom, the
329: panels show the velocity variations in the umbrae of the big and medium-sized sunspots,
330: the pore and the facular region. The asterisks represent the measured 
331: values, equispaced in time. The departure from a sinusoidal behavior
332: is a signature of the passage of shock waves through this 
333: layer of the atmosphere.
334: It is clear from Fig.~\ref{fig:chrom-oscil} that, as the magnetic flux
335: of the structure decreases, the amplitude of the oscillations also becomes
336: smaller (from $\sim$15 km\,s$^{-1}$ in the big sunspot to $\sim$3--4 km\,
337: s$^{-1}$ in the facula). Note that the projection effects are larger for the three
338: sunspot-like features than for the facula because they are farther away from disk center. 
339: Assuming that the plasma movements are directed along the field lines - parallel to the
340: local vertical- then the actual oscillation amplitudes for the sunspots are even larger than
341: the values given above.
342: The steepness and the frequency of appearance of the shocks is also correlated with
343: the magnetic flux of the observed feature. 
344: 
345: \noindent The chromospheric oscillations in the facular region present a 
346: particularity that is not shared by the sunspot-like structures (i.e. both
347: sunspots and the pore). While the latter show a characteristic 3-minute pattern, 
348: the facular region presents, in contrast, an obvious 5-minute period at 
349: chromospheric heights.
350: 
351: \clearpage
352: \begin{figure}[t!]
353: \center 
354: \includegraphics[width=7cm]{f4a.ps} 
355: \includegraphics[width=7cm]{f4b.ps} 
356: \includegraphics[width=7cm]{f4c.ps} 
357: \includegraphics[width=7cm]{f4d.ps} 
358: \caption{Normalized LOS velocity power spectra of the four analyzed regions. 
359: From left to right
360: and top to bottom, the panels show the photospheric (dashed) and chromospheric
361: (solid) power spectra, (averaged over the length that the structure spanned
362: along the spectrograph slit and normalized to their maximum), for the big sunspot, the small sunspot, the pore
363: and the facular region.}
364: \label{fig:power-spectra}
365: \end{figure}
366: \clearpage
367: 
368: Fig.~\ref{fig:power-spectra} illustrates the difference between
369: the spectral composition of the oscillation patterns of the sunspot-like 
370: structures and the facular region.
371: From left to right and top to bottom, the panels represent the 
372: average photospheric ({\em dashed}) and chromospheric ({\em solid}) power
373: spectra in the four magnetic features. While in the sunspot-like
374: structures the majority of the oscillatory energy is concentrated around 
375: different frequency regimes in the photosphere and the chromosphere, the 
376: facular power spectra show a strikingly similar behavior at both heights 
377: (with the maximum power lying at 3.3\,mHz and the secondary peaks being 
378: co-located and showing pretty much the same distribution).
379: 
380: 
381: \noindent One could attribute this to the helium triplet being formed lower in the atmosphere
382: (closer to the region of formation of the Silicon line).
383: However, this can be ruled out for two reasons: 1) the oscillations derived from
384: the He lines have at least ten times the amplitude of those measured with the Si line,
385: indicating that they are clearly chromospheric. 2) if the He triplet is formed
386: by the triggering effect of the coronal irradiance \cite[see][]{avrett, centeno2008}, there is no way the lines can have a contribution function from the 
387: photosphere, but it can only come from the high chromosphere.
388: 
389: \noindent The 5-minute chromospheric oscillations appear as a consequence of the reduced
390: cut-off frequency in facular regions.
391: 
392: 
393: 
394: \section{Wave propagation}
395: 
396: For the wave propagation analysis we will follow the strategy adopted in
397: Paper I.
398: First, we will analyze the information given by the mean power, phase difference and amplification
399: spectra. This will allow us to identify the propagation regime of the wave modes in each case
400: and determine the cutoff frequency and the amplification of the oscillations as the waves travel
401: through the atmosphere. Then, we will try to reproduce these observables with a simple model 
402: of vertical linear wave propagation along constant magnetic field lines in a stratified 
403: atmosphere. The theoretical modeling will allow us to make a prediction about the time 
404: delay between the signals measured at the photosphere and the chromosphere, which will be
405: contrasted with what is obtained by mere cross-correlation of the measured velocity maps
406: at both heights.
407: 
408: \subsection{Phase difference and cutoff frequencies}
409: 
410: \clearpage
411: \begin{figure}[t!]
412: \center 
413: \includegraphics[width=7cm]{f5a.ps} 
414: \includegraphics[width=7cm]{f5b.ps} 
415: \includegraphics[width=7cm]{f5c.ps} 
416: \includegraphics[width=7cm]{f5d.ps} 
417: \caption{Phase difference between photospheric and chromospheric oscillations 
418: as a function of frequency for the four magnetic structures (from left to right and top to bottom
419: the panels correspond to the umbrae of the big and small sunspots, the pore and the facula). Each cross on 
420: the figure was computed as the difference between the phase of the 
421: chromospheric and the photospheric oscillation for a unique position along
422: the slit and a particular frequency. The small insets show the respective
423: amplification spectra. Solid lines correspond to the best fit to a model of
424: linear wave propagation in a stratified isothermal atmosphere with radiative losses,
425: permeated by a constant vertical magnetic field.}
426: \label{fig:phase}
427: \end{figure}
428: \clearpage
429: 
430:  
431: Fig.~\ref{fig:phase} shows the 
432: phase difference between the photospheric and the chromospheric 
433: oscillations for the four magnetic features analyzed in this paper.
434: Each cross on the panels was computed as the difference between the phase of 
435: the chromospheric and the photospheric oscillations for one position along
436: the slit and one frequency value.
437: The turn-off points in the phase spectra 
438: correspond to the effective cut-off frequencies of the 
439: atmosphere. Oscillations with lower frequencies will not be able to make it 
440: through the atmosphere and, thus, will remain trapped producing
441: stationary waves. Above the cut-off value, perturbations propagate upwards 
442: freely into the chromosphere.
443: Note that in the case of the sunspot-like structures the atmospheric 
444: cut-off stands around 4 mHz, inhibiting the propagation of lower
445: frequency modes. This has strong implications on what happens
446: to the photospheric oscillatory energy -which lies
447: below the cut-off value- and strikes the question of the origin of the 
448: 3-minute chromospheric power in such structures. Paper I addressed this question 
449: and proved that, inside the umbrae of sunspots, the chromospheric 
450: 3-minute oscillations come from the upward linear propagation of the photospheric
451: 3-minute power rather than from the non-linear interaction and 
452: redistribution of the energy stored in the 5-minute photospheric modes.
453: This means that, while the 5-minute photospheric component stays trapped in the
454: atmosphere giving rise to stationary waves, the 3-minute photospheric
455: component travels upward through the atmosphere driving the chromospheric 
456: oscillations.
457: 
458: \noindent The case of the facular region is somewhat different. Now, the
459: cut-off frequency is lower than that of the typical photospheric modes, 
460: thus allowing them to propagate through the atmosphere.
461: This would explain why the facular chromospheric power spectrum peaks around
462: 3.3 mHz, since all the 5-minute photospheric power can travel freely upward
463: into the high chromosphere. 
464: 
465: \noindent The small inset in the lower right corner of each panel of
466: Fig.~\ref{fig:phase} represents
467: the corresponding amplification spectrum (dashed line), this is, the ratio of the
468: chromospheric power over the photospheric power as a function of
469: frequency. For decreasing magnetic flux, the amplification of the power
470: in the propagation regime (i.e. above the cut-off frequency) also decreases.
471: 
472: \subsection{Model}
473: 
474: A simple model of upward linear wave propagation in an isothermal stratified 
475: atmosphere permeated by a constant vertical magnetic field (described in Paper I, 
476: but originally accounted for by Souffrin in 1972) was chosen to further explain the 
477: observations. Energy exchange by radiative losses is permitted by Newton's 
478: cooling law, which accounts for the damping of the temperature fluctuations
479: with a typical relaxation time $\tau_R$ \citep{spiegel57, kneer87}:
480: 
481: \begin{equation}
482: \tau_R=\frac{\rho c_v}{16 \chi \sigma_RT^3}
483: \end{equation}
484: 
485: \noindent where $\chi$ is the grey absorption coefficient and $\sigma_R$ the 
486: Stefan-Boltzmann constant. The solution $A(z) = D e^{z/2H_0} e^{ik_z z} e^{i\omega t}$ substituted
487: into the wave equation:
488: 
489: \begin{equation}
490: \hat c^2\frac{d^2 A(z)}{dz^2} -\hat \gamma g \frac{dA(z)}{dz} +\hat \omega^2 A(z) = 0
491: \end{equation}
492: 
493: \noindent yields a dispersion relation:
494: 
495: \begin{equation}
496: k_z^2 = \frac{\omega^2-\hat\omega_{ac}^2}{\hat c^2},
497: \end{equation}
498: 
499: 
500: \noindent where $k_z$ is the vertical wave number, $H_0$ the pressure scale height,
501: and $\hat\omega$ and $\hat c$ were defined in \cite{bunte}:
502: 
503: \begin{equation}
504: \hat \omega_{ac} = \hat c / 2H_0, \qquad \hat c^2 = \hat \gamma g H_0, \qquad \hat \gamma = \frac{1-\gamma i \omega \tau_R}{1 - i \omega \tau_R}
505: \end{equation}
506: 
507: 
508: \noindent This results in a 3-free-parameter model, with the 
509: temperature, $T$, the vertical distance between the measured oscillations, $\Delta z$, 
510: and the typical radiative cooling time, $\tau_R$, as the fitting coefficients.
511: For a more detailed description, we refer the reader to Paper I, B\"unte and Bogdan (1994), Mihalas \& Wiebel-Mihalas (1984) and Souffrin (1972).
512: 
513: \noindent The model assumes a magnetic field that is aligned with gravity. The full
514: Stokes inversion of the photospheric Si line in our data-sets shows that, in all cases,
515: the deviation of the magnetic field direction from the local vertical never exceeds 
516: $20^{\circ}$, being close to $0^{\circ}$ for most of the pixels (within the uncertainties). 
517: 
518: 
519: The solid lines (both in the bigger panels and in the insets) of Fig. \ref{fig:phase}
520: correspond to the best fit of the model to the data. There, all the pixels of the observed 
521: targets are plotted together. The free parameters of the model were adjusted to explain, simultaneously, the average observed phase difference and 
522: average amplification spectrum. The fit has to account for the turn-off point 
523: due to the cut-off frequency and the steepness of the phase spectrum together 
524: with the magnitude of the amplification of the power. The different parameters have very
525: distinct effects on the resulting curves. For instance, while the radiative cooling time controls
526: the position of the turn-off point, the height difference determines the magnitude of the
527: amplification of the signals from the photosphere to the chromosphere. 
528: Table \ref{tab:fitting-params} compiles the values of the resulting 
529: fitting parameters.
530: 
531: \clearpage
532: \begin{table}[!t]
533: \begin{center}
534: \begin{tabular}{c|cccc}
535:  & big sunspot & small sunspot & pore & facular region\\
536: \hline
537: $T$ [K] & 4000 & 4500 & 5000 & 9000 \\
538: $\Delta z$ [km] & 1000 & 1000 & 1000 & 1500 \\
539: $\tau_R$ [s] & 55 & 30 & 25 & 10 
540: \label{tab:fitting-params}
541: \end{tabular}
542: \caption{Fitting parameters used in linear wave propagation modeling}
543: \end{center}
544: \end{table}
545: \clearpage
546: 
547: In the sunspot-like structures, as the magnetic flux decreases, the typical
548: cooling time also decreases while the temperature grows. Although quantitatively
549: these numbers are difficult to justify due to the simplicity of the model, 
550: they do make sense in a qualitative way.
551: It is no doubt expected that the smaller the structure, the larger the temperature 
552: inside it, since its magnetic field becomes more and more incapable of inhibiting the 
553: convection process underneath the photosphere. On the other hand, the radiative
554: cooling time is related to the inhomogeneity, taking smaller values the less homogeneous
555: the structure \citep[see][]{spiegel57, kneer87}. The difference in heights remains essentially the same for the three sunspots,
556: while it turns out to be larger in the case of the facula. Sunspot-like features are evacuated
557: structures (due to the balance between magnetic and gas pressure), allowing the coronal
558: EUV irradiance to travel further down into the chromosphere and trigger the formation of
559: the He {\sc i} 10830 \AA\ triplet at lower layers \cite[see][]{centeno2008}.
560: 
561: 
562: 
563: \subsection{Theoretical prediction and measured time delays}
564: 
565: \clearpage
566: \begin{figure}[t!]
567: \center 
568: \includegraphics[width=7cm]{f6a.ps} 
569: \includegraphics[width=7cm]{f6b.ps} 
570: \includegraphics[width=7cm]{f6c.ps} 
571: \includegraphics[width=7cm]{f6d.ps} 
572: \caption{Expected and measured time delays. From left to right and top to bottom,
573: the panels correspond to the umbra of the big sunspot (extracted from Paper I), 
574: the umbra of the small sunspot, the pore and the facular region. The solid line shows 
575: the expected time delay, as a function  of frequency, predicted by the model. The asterisks correspond to the measured delays. }
576: \label{fig:theor-time-delay}
577: \end{figure}
578: 
579: 
580: \begin{figure}[t!]
581: \center 
582: \includegraphics[width=9cm]{f7a.ps} 
583: \includegraphics[width=9cm]{f7b.ps} 
584: \includegraphics[width=9cm]{f7c.ps} 
585: \includegraphics[width=9cm]{f7d.ps} 
586: \caption{Time delays in the umbra of the small sunspot. 
587: The panels show the velocities, for one pixel position inside the umbra, filtered in 1 mHz frequency bands around 4, 5, 6 and 7 mHz. Overplotted to the 
588: chromospheric velocity (dashed lines) is the photospheric signal (solid lines), filtered
589: in the same frequency band, but amplified and delayed to make it match the latter one. Similar plots for other pixels along the slit result in delays and amplifications that are consistent 
590: with those shown here.}
591: \label{fig:time-delay}
592: \end{figure}
593: 
594: \begin{figure}[t!]
595: \center 
596: \includegraphics[width=9cm]{f8a.ps} 
597: \includegraphics[width=9cm]{f8b.ps}  
598: \caption{Measured time delays inside the pore. Analogous to 
599: Fig. \ref{fig:time-delay} except for the filtering bands. From top to bottom, the velocity
600: signals have been filtered in 1 mHz bands around 5.5 and 6.5 mHz, respectively. The photospheric signal was shifted in time and amplified in order to match the chromospheric one.}
601: \label{fig:pore-time-delay}
602: \end{figure}
603: 
604: \begin{figure}[t!]
605: \center 
606: \includegraphics[width=9cm]{f9a.ps} 
607: \includegraphics[width=9cm]{f9b.ps} 
608: \includegraphics[width=9cm]{f9c.ps} 
609: \caption{Measured time delays inside the facular region. Analogous to 
610: Fig. \ref{fig:time-delay} except for the filtering bands. From top to bottom, the velocity
611: signals have been filtered in 1 mHz bands around 2.5, 3.5 and 4.5 mHz, respectively. The photospheric signal was shifted in time and amplified in order to match the chromospheric one.}
612: \label{fig:facula-time-delay}
613: \end{figure}
614: 
615: 
616: \clearpage
617: 
618: The solid lines in  Fig.~\ref{fig:theor-time-delay} show the expected time delays 
619: (in the four magnetic structures), as a function of frequency, between the oscillations measured at the photosphere and the chromosphere as derived from the 
620: theoretical model. The time delay, $\Delta t$, depends on the difference in heights, $\Delta z$
621: and the group velocity of the wave packet, $v_g$:
622: 
623: \begin{equation}
624: \Delta t = \frac{\Delta z}{v_g}
625: \end{equation}
626: 
627: 
628: \noindent where 
629: 
630: \begin{equation}
631: v_g = \frac{d \omega}{d k_z}
632: \end{equation}
633: 
634: \noindent Using the best fitting values for $\Delta z$, $T$ and $\tau_R$, 
635: we computed the theoretical time delay from the wave propagation model described above.
636: 
637: 
638: 
639: \noindent If propagation is mainly linear within the 2 - 7 mHz band (as suggested by the good fits of the model to the data)
640: and it takes place along the magnetic field lines, then we should expect to see a 
641: correlation of the photospheric and chromospheric oscillation patterns above the 
642: cutoff-frequency.
643: It should be possible to determine the time delay from the observations by simply
644: comparing these modulation patterns at both heights. 
645: The time shift that yields the maximum correlation will correspond to the measured delay. 
646: 
647: 
648: However, the theoretical time delays depicted in Fig. \ref{fig:theor-time-delay} show a very strong dependence on the 
649: frequency of the oscillating mode, so we have to take this fact into account
650: when doing the comparison. Following the approach taken in Paper I, we first filter 
651: both the photospheric and the chromospheric velocity maps in narrow 
652: frequency ranges (narrow enough that the expected time delay does not 
653: vary significantly within the bandwidth).
654: Then, we compare the photospheric and chromospheric filtered signals finding that
655: we have to apply a certain time shift between them in order to make their external
656: modulation schemes match. This shift corresponds to the time that a perturbation
657: (within the filtering frequency range) originated in at photospheric levels takes to reach
658: the chromosphere.
659: 
660: In the case of the sunspot-like structures, we filtered the velocity maps in 
661: several one-mHz bands close to 6 mHz (where the main contribution to the 
662: chromospheric power lies). 
663: We then compared each pair of filtered maps finding a time shift between them 
664: that depended on the frequency range in which the maps were filtered. 
665: The four panels of Fig. \ref{fig:time-delay} show the photospheric (solid) and 
666: chromospheric (dashed) velocities, 
667: for one position inside the umbra of the small sunspot, filtered in $\sim 1$ mHz
668: frequency bands around 4, 5, 6 and 7 mHz. In each case, the photospheric signal has been amplified 
669: and delayed to make it match the chromospheric one. 
670: 
671: 
672: \noindent This procedure was repeated for several pixels in the umbra.
673: There are coherent patches of a few arcseconds along the slit in 
674: which the velocity signals are very similar, so we chose 3 or 4 
675: positions far apart, and measured the delay for each of them finding consistent 
676: time shifts and amplification values at the different spatial locations. The 
677: uncertainty for the time delay is of the order of 1 - 2 time steps ($\sim$ 10 - 15s).
678: The measured delays extracted from this method are represented by the asterisks
679: over plotted to the top left panel of Fig. \ref{fig:theor-time-delay}. Analogous analyses
680: were carried out for the pore (shown in Fig. \ref{fig:pore-time-delay}) and for the big 
681: sunspot (in Paper I).
682: 
683: \noindent The pore is only 4 arcsecs wide and oscillations are quite 
684: coherent throughout the whole structure. Also, there is not a strong 
685: modulation of the velocity oscillation pattern, so, in certain 
686: frequency bands, it is not clear what time shift gives the larger 
687: correlation between photosphere and chromosphere. This is why we only show 
688: the results for two frequency bands in Fig. \ref{fig:pore-time-delay}.
689: 
690: \noindent In the case of the facular region we filtered the velocity maps 
691: in three ranges around 3 mHz (where both photospheric and chromospheric power spectra have their main contributions). The remaining analysis is parallel to the former 
692: case. The three panels of Fig. \ref{fig:facula-time-delay} show the chromospheric 
693: ({\em solid}) and photospheric ({\em dashed}) velocity signals 
694: filtered around 2.5, 3.5 and 4.5 mHz, respectively. Again, the photospheric signal has
695: been amplified and shifted in time to make it match the chromospheric one.
696: 
697: In all the cases, the amplification factors turn out to be consistent with
698: the values of the amplification spectrum in Fig. \ref{fig:phase},
699: and the measured time delays (represented by asterisks superposed to the theoretical 
700: time delays in Fig. \ref{fig:theor-time-delay}) obtained from the shifts are 
701: consistent with the expected values obtained from the model. 
702: Even though the theoretical curves predict a very strong variation of the time 
703: delay within the 1 mHz filtering bands, the measured delays agree surprisingly well 
704: with what is expected by the model.
705: 
706: \section{Discussion and conclusions}
707: 
708: In this paper we have investigated the wave propagation in the atmospheres of four solar 
709: magnetic structures with decreasing
710: flux (two different-sized sunspots, a pore and a facular region). 
711: Simultaneous and co-spatial measurements of the LOS velocity at the photosphere 
712: and the chromosphere of these structures allow us to infer information about 
713: the properties of wave propagation from one atmospheric layer to another. A simple model of 
714: linear vertical wave propagation in a magnetized stratified medium with radiative 
715: losses is enough to explain the observed phase difference and amplification spectra 
716: and the measured time delays.
717: 
718: The inversion of the full Stokes vector in the four structures reveals vertical (i.e. radial)
719: magnetic fields in all cases. The comparison of the photospheric and chromospheric velocity
720: maps, filtered in narrow frequency ranges, shows a pixel to pixel correlation along the slit 
721: of the external modulation of the 
722: wave pattern (after accounting for a global time shift and a global amplification of the 
723: photospheric signal). These two facts are enough to justify the election of a model of
724: linear wave propagation along vertical magnetic field lines.
725: 
726: In the case of sunspot-like structures the atmospheric cut-off lies around $\sim$4 mHz,
727: so the modes with frequencies below this one will not be able to reach the high 
728: chromosphere. Many authors have argued before about the possibility 
729: of the non-linear interaction among 5-min modes being the source of the
730: chromospheric 3-min oscillations \citep[see, e.g.][]{fleck}. But if this
731: were the case, the photospheric and chromospheric filtered velocity maps would show no
732: resemblance with each other. As a matter of fact, a clear correlation exists, 
733: indicating that most of the 6 mHz power observed at chromospheric heights 
734: in sunspot-like structures comes directly from the same frequency range in the
735: photosphere via upward linear wave propagation. This ratifies and extends the
736: conclusions of Paper I (which focuses on the umbrae of big sunspots) to smaller sunspots and
737: pores. As the size of the structure decreases, so does the typical radiative cooling time,
738: while the temperature, on the other hand, grows. As argued before, both these 
739: behaviors are in agreement with what is expected from a qualitative point of view.
740: 
741: In the case of facular regions, the cut-off frequency stands around 2 mHz.
742: This is accounted for in the model by introducing a shorter cooling time and a higher
743: temperature. This allows the 5-min power to
744: propagate through the atmosphere and reach the high chromosphere. Again, a clear correlation between photospheric and chromospheric
745: filtered velocity maps can be found, indicating that the propagation is mainly linear and vertical
746: within the 2--5 mHz range.
747: 
748: \noindent It is interesting to point out that, in this particular case, there is no need to 
749: invoke a large inclination of magnetic flux tubes to explain the p-mode leakage into the chromosphere \cite[][]{depontieu2004}. 
750: Furthermore, the comparison of the photospheric and chromospheric velocity maps shows a good 
751: co-spatial correlation (after applying a convenient time shift) indicating that
752: the propagation is essentially vertical. If we assume that the wave propagation takes place
753: along the field lines and we take into account a height difference of 1000 - 1500 km,
754: a magnetic field inclination of $40 - 45^{\circ}$ would result in a spatial displacement of
755: 3 - 5 pixels between the photospheric and the chromospheric oscillations (rather than 
756: happening on the same vertical). In the best case scenario, this displacement would be parallel to
757: the slit; however, in the worst case, it would happen along the direction perpendicular to
758: the slit, leading to a correlation between the photosphere and the chromosphere that is
759: marginally compatible with the spatial resolution of
760: our data. On the other hand, the measured Stokes profiles set an upper limit of 
761: $20^{\circ}$ on the inclination of the photospheric facular magnetic fields. Both these arguments point
762: towards a magnetic field structure that is incompatible with the inclinations needed to lower 
763: the cut-off frequency enough to allow p-mode leakage.
764: 
765: \noindent The numerical simulations of Khomenko et al. (2008) have confirmed this 
766: conclusion since they show how, in a more realistic atmosphere, it is possible to explain the propagation of 5-minute modes into 
767: the chromosphere through vertical thin flux tubes, using radiative losses as the
768: main ingredient to lower the cut-off frequency.
769: 
770: As the photospheric perturbations propagate upwards, their amplitude
771: increases due to the rapid decrease in density and they eventually develop
772: asymmetries that steepen more or less depending on the magnetic structure
773: (the lower the magnetic flux, the smaller the amplification and the steepness of the developing shock).
774: In all the cases, the time delay between photospheric and chromospheric
775: oscillations is around several minutes in the frequency range near where the
776: chromospheric power peaks.
777: 
778: 
779: 
780: 
781: \acknowledgments
782: This research was partially funded by the Spanish Ministerio de Educaci\'on 
783: y Ciencia through project AYA2007-63881.
784: The National Center for Atmospheric Research (NCAR) is sponsored by the 
785: National Science Foundation
786: 
787: 
788: 
789: \begin{thebibliography}
790: 
791: \bibitem[Abdelatif \& Thomas 1987]{abdelatif} Abdelatif, T.A, Thomas, J.H, 1987, ApJ, 320, 884
792: 
793: \bibitem[Athay \& White, 1979]{athay} Athay, G.R., White, O.R., 1979, ApJS, 39, 333.
794: \bibitem[Avrett et al. 1994]{avrett} Avrett, E.H., Fontenla, J.M., Loeser, R., 1994, IAUS, 154, 35	
795: \bibitem[Ballesteros et al. 1996]{ballesteros1996} Ballesteros, E., 
796: Collados, M., Bonet, J. A., Lorenzo, F., Viera, T., Reyes, M. \&
797: Rodriguez Hidalgo, I., 1996, A\&AS, 115, 353.
798: \bibitem[Bard \& Carlsson 2008]{silicon} Bard, S., Carlsson, M., 2008, ApJ, 682, 1376.
799: \bibitem[Beckers \& Tallant 1969]{beckers} Beckers, J.M., Tallant, P.E., 1969, Sol. Phys., 7, 351.
800: \bibitem[Bogdan \& Judge 2006]{bogdan2006} Bogdan, T.J., Judge, P.G., 2006, Phil. Trans. R. Soc. London A, 364, 313.
801: \bibitem[Brynildsen et al. 2004]{bry2004} Brynildsen, N.,Maltby, P., Fredvik, T., \& Kjeldseth-Moe, O., 2004, in SOHO 13: Waves, Oscillations, and Small-Scale Transient Events in the Solar Atmosphere, ed. H. Lacoste (ESA SP-547; Noordwijk: ESA), 45
802: \bibitem[Brynildsen et al. 2003]{bry2003}Brynildsen, N., Maltby, P., \& Kjeldseth-Moe, O. 2003, A\&A, 398, L15
803: \bibitem[B\"unte \& Bogdan 1994]{bunte} B\"unte, M., Bogdan, T.J., 1994, A\&A, 283 642
804: \bibitem[Centeno et al. 2006a]{centeno2006} Centeno, R., Collados, M., Trujillo Bueno, J. 2006a, ApJ, 640, 1153.
805: \bibitem[Centeno et al. 2006b]{centeno2006a} Centeno, R., Collados, M., Trujillo Bueno, J., 2006b, Proceedings of the Solar Polarization Workshop 4, ASP Conference Series, Vol. 358. Edited by R. Casini and B. W. Lites, p.465
806: \bibitem[Centeno et al. 2008]{centeno2008} Centeno, R., Trujillo Bueno, J., Uitenbroek, H., Collados, M., 2008, ApJ, 677, 742.
807: \bibitem[Centeno et al. in press]{centeno2008a} Centeno, R., Trujillo Bueno, J., Uitenbroek, H., Collados, M., to appear in the proceedings of the Solar Polarization Workshop 5.
808: \bibitem[Collados 1999]{collados99} Collados, M. 1999, in ASP Conf. Ser. 184, Third Advances in Solar Physics Euroconference: Magnetic Fields and Oscillations, ed. B. Schmieder, A. Hofmann, \& J. Staude (San Francisco: ASP), 3
809: \bibitem[Collados 2003]{collados2003} Collados, M., 2003, in "Polarimetry in Astronomy", edited by Silvano Fineschi, Proceedings of the SPIE, vol. 4843, pp. 55-65
810: \bibitem[Collados et al. 2001]{collados2001} Collados, M., Trujillo Bueno, J., Bellot Rubio, L.R., Socas-Navarro, H., 2001, INTAS Workshop on MHD Waves in Astrophysical Plasmas, Universitat de les Illes Balears, pp. 151-154, Eds. J.L. Ballester and B. Roberts.
811: \bibitem[De Pontieu et al. 2003]{depontieu2003} De Pontieu, B., Erdelyi, R., De Moortel, I., 2005, ApJ, 624, 61.
812: \bibitem[De Pontieu et al. 2003a]{depontieu2003a} De Pontieu, B., Erdelyi, R., de Wijn, A.G., 2003, ApJ, 595, L63
813: \bibitem[De Pontieu et al. 2004]{depontieu2004} De Pontieu, B., Erdelyi, R., James, S.P, 2004, Nature, 430, 536
814: \bibitem[Fleck \& Schmitz 1993]{fleck} Fleck, B. \& Schmitz, F.~1993, 
815: A\&A, 273, 671
816: \bibitem[Gurman \& Leichbacher 1984]{gurman} Gurman, J.B., Leichbacher, J.W., 1984, ApJ, 283, 859.
817: \bibitem[Jefferies et al. 2006]{jefferies2006} Jefferies, S.M., McIntosh, S.W.,
818: Armstrong, J.D., Bogdan, T., Thomas, J., Cacciani, A. \& Fleck, B. ApJ, 648,
819: L151
820: \bibitem[Khomenko et al. 2008]{khomenko2008} Khomenko, E., Centeno, R., Collados, M., Trujillo Bueno, J., 2008, ApJ, in press.
821: \bibitem[Khomenko \& Collados 2006]{khomenko2006} Khomenko, E., Collados, M., 2006, ApJ, 653, 755.
822: \bibitem[Kneer \& Trujillo Bueno 1987]{kneer87} Kneer, F., Trujillo Bueno, J., 1987, A\&A, 183, 91.
823: \bibitem[Krijger et al. 2001]{krijger} Krijger, J.M., Rutten, R.J., Lites, B.W., Straus, T., Shine, R.A., Tarbell, T.D., 2001, A\&A, 379, 1052.
824: \bibitem[Lagg 2007]{lagg2007} Lagg, A., 2007, Advances in Space Research 39, 1734.
825: \bibitem[Leighton et al. 1962]{leighton} Leighton, R.B., Noyes, R.W., Simon, G.W., 1962,
826: ApJ, 135, 474.
827: 
828: \bibitem[Lites 1986]{lites1986} Lites, B.W., 1986, ApJ, 301, 1005.
829: \bibitem[Lites 1986b]{lites1986b} Lites, B.W.,1986, ApJ, 301, 992.
830: \bibitem[Lites 1992]{lites1992} Lites, B.~W.~1992, in Sunspots: Theory and
831: Observations, ed. J.~H.~Thomas \& N.~O.~Weiss, NATO ASI ser. C 375,
832: Kluwer, Dordrecht, 261
833: \bibitem[Lites et al. 1993]{lites93} Lites, B.W., Rutten, R.J., Kalkofen, W., 1993, ApJ, 414, 354
834: \bibitem[Mart\'\i nez Pillet et al. 1999]{vmp-tip} Mart\'\i nez Pillet, V., Collados, M., S\'anchez Almeida, J. et al., 1999, High Resolution
835: Solar Physics: Theory, Observations and Techniques, ASP Conference Series 183,
836: (eds.) T.R. Rimmele, K.S. Balasubramaniam, and R. Radick, 264.
837: \bibitem[Mihalas \& Wiebel-Mihalas 1984]{mihalas} Mihalas, D., Wiebel-Mihalas, B., 1984, "Foundations of Radiation Hydrodynamics", Oxford University Press, Oxford.
838: \bibitem[Roberts 2006]{roberts} Roberts, B., 2006, Royal Society of London Transactions Series A, 364, Issue, 1839, 447.
839: %%\bibitem[R\"uedi et al. 1995]{ruedi} R\"uedi, I., Solanki, S.K., Livingston, W.C.,  1995, A\&A, 293, 252
840: \bibitem[Ruiz Cobo et al. 1994]{brc94} Ruiz Cobo, B., del Toro Iniesta, J.C., 1994, A\&A, 283, 129.
841: \bibitem[S\'anchez-Andrade Nu\~no et al. 2007]{bruno07} S\'anchez-Andrade Nu\~no, B., Centeno, R., Puschmann, K.G., Trujillo Bueno, J., Blanco Rodr\'\i guez, J., Kneer, F., 2007, A\&A Letters, in press.
842: \bibitem[Sasso et al. 2005]{sasso} Sasso, C., Lagg, A., Solanki, S.K., 2005, ESASP, 596E, 64.
843: \bibitem[Schmidt \& Kentischer 1995]{schmidt1995} Schmidt, W. \& Kentischer, T.,
844: 1995, A\&AS, 113, 363
845: \bibitem[Schmidt et al. 1994]{knoelker} Schmidt, W., Kn\"olker, M., Westendorp Plaza, C., 1994, A\&A, 287, 229.
846: \bibitem[Socas-Navarro Trujillo Bueno and Ruiz Cobo 2000]{sn2000} Socas-Navarro, H., Trujillo Bueno, J., \& Ruiz Cobo, B. 2000, Science, 288, 1396
847: \bibitem[Souffrin 1972]{souffrin} Souffrin, P., 1972, A\&A, 17, 458.
848: \bibitem[Spiegel 1957]{spiegel57} Spiegel, E.A., 1957, ApJ, 126, 202.
849: \bibitem[Thomas 1983]{thomas} Thomas, J.H., 1983, AnRFM, 15, 321.
850: \bibitem[Trujillo Bueno \& Asensio Ramos 2007]{tb2007} Trujillo Bueno, J., Asensio Ramos, A., 2007, ApJ, 655. 642.
851: \bibitem[White \& Athay 1979]{white} White, O.R., Athay, G.R., 1979, ApJS, 39, 347.
852: \bibitem[Zhugzhda et al 1985]{zhu} Zhugzhda, Y.D., Locans, V., Staude, J., 1985, A\&A, 143, 201.
853: 
854: \end{thebibliography}
855: 
856: 
857: \end{document}
858: