The inverse C-shape of solar chromospheric line bisectors

Similar documents
The Solar Chromosphere

Chapter 1. Introduction. 1.1 Why study the sun?

The Solar Temperature Minimum and Chromosphere

Evidence for a siphon flow ending near the edge of a pore

THE NON-MAGNETIC SOLAR CHROMOSPHERE MATS CARLSSON. Institute of Theoretical Astrophysics, P.O.Box 1029 Blindern, N{0315 Oslo, Norway.

THE MYSTERIOUS SOLAR CHROMOSPHERE

Supporting Calculations for NASA s IRIS Mission. I. Overview

Chromospheric heating and structure as determined from high resolution 3D simulations

The Solar Chromosphere

Molecular hydrogen in the chromosphere IRIS observations and a simple model

Influence of Mass Flows on the Energy Balance and Structure of the Solar Transition Region

Surface Convection INTRODUCTION

arxiv:astro-ph/ v1 28 Feb 2003

Advances in measuring the chromospheric magnetic field using the He triplet

The accuracy of the center-of-gravity method for measuring velocity and magnetic field strength in the solar photosphere

What do we see on the face of the Sun? Lecture 3: The solar atmosphere

2 The solar atmosphere

Observable consequences

B.V. Gudiksen. 1. Introduction. Mem. S.A.It. Vol. 75, 282 c SAIt 2007 Memorie della

Scales of solar convection

Three Dimensional Radiative Transfer in Winds of Massive Stars: Wind3D

2. Stellar atmospheres: Structure

The Sun s Dynamic Atmosphere

Solar Astrophysics with ALMA. Sujin Kim KASI/EA-ARC

A STUDY OF TRANSITION REGION AND CORONAL DOPPLER SHIFTS IN A SOLAR CORONAL HOLE. M. D. Popescu 1,2 and J. G. Doyle 1

Investigating Molecular Hydrogen in Active Regions with IRIS

arxiv:astro-ph/ v1 21 Dec 2006

Observations of Umbral Flashes

Results from Chromospheric Magnetic Field Measurements

The impact of solar surface dynamo magnetic fields on the chemical abundance determination

pre Proposal in response to the 2010 call for a medium-size mission opportunity in ESA s science programme for a launch in 2022.

The solar atmosphere

Solar photosphere. Michal Sobotka Astronomical Institute AS CR, Ondřejov, CZ. ISWI Summer School, August 2011, Tatranská Lomnica

CNO abundances in the Sun and Solar Twins

THE OBSERVATION AND ANALYSIS OF STELLAR PHOTOSPHERES

IRIS views on how the low solar atmosphere is energized

Solar-B. Report from Kyoto 8-11 Nov Meeting organized by K. Shibata Kwasan and Hida Observatories of Kyoto University

Propagating waves in the sunspot umbra chromosphere. N. I. Kobanov and D. V. Makarchik

The dynamics of the solar chromosphere: comparison of model predictions with millimeter-interferometer observations ABSTRACT

Solar and Stellar Atmospheric Modeling Using the Pandora Computer Program

Time-dependent hydrogen ionisation in 3D simulations of the solar chromosphere. Methods and first results

O 5+ at a heliocentric distance of about 2.5 R.

DETERMINATION OF THE FORMATION TEMPERATURE OF Si IV IN THE SOLAR TRANSITION REGION

RESEARCH PROFILE OF DR. SVEN WEDEMEYER

THE SOLAR CHROMOSPHERE AND CORONA: QUIET SUN

Small-scale structure and dynamics of the chromospheric magnetic eld

The Interior Structure of the Sun

Eugene H. Avrett and Rudolf Loeser

Solar radiation and plasma diagnostics. Nicolas Labrosse School of Physics and Astronomy, University of Glasgow

Power halo and magnetic shadow in a solar quiet region observed in the Hα line ABSTRACT

Doppler Shifts of the Hα Line and the Ca II nm Line in a Quiet Region of the Sun Observed with the FISS/NST

Name: Partner(s): 1102 or 3311: Desk # Date: Spectroscopy Part I

Results from Chromospheric Magnetic Field Measurements

Energy transport: convection

Chapter 8 The Sun Our Star

1. INTRODUCTION 2. OBSERVATIONS AND DATA REDUCTION. The Astrophysical Journal, 502:L85 L90, 1998 July 20

The Sun. The Sun Is Just a Normal Star 11/5/2018. Phys1411 Introductory Astronomy. Topics. Star Party

Learning Objectives. wavelengths of light do we use to see each of them? mass ejections? Which are the most violent?

Chapter 6: Granulation

The Structure of the Sun. CESAR s Booklet

Chapter 14 Lecture. The Cosmic Perspective Seventh Edition. Our Star Pearson Education, Inc.

6. Stellar spectra. excitation and ionization, Saha s equation stellar spectral classification Balmer jump, H -

NLTE solar flare models with stationary velocity fields

The Sun. The Sun is a star: a shining ball of gas powered by nuclear fusion. Mass of Sun = 2 x g = 330,000 M Earth = 1 M Sun

ASTR-1010: Astronomy I Course Notes Section IV

The Sun. Basic Properties. Radius: Mass: Luminosity: Effective Temperature:

On the outflow at coronal heights or the way I learned to listen to Philippe

Phys 100 Astronomy (Dr. Ilias Fernini) Review Questions for Chapter 8

Predicting the Extreme-UV and Lyman-α Fluxes Received by Exoplanets from their Host Stars

arxiv: v1 [astro-ph.sr] 26 Jan 2010

Line profiles in moustaches produced by an impacting energetic particle beam

Search for photospheric footpoints of quiet Sun transition region loops

Section 11.5 and Problem Radiative Transfer. from. Astronomy Methods A Physical Approach to Astronomical Observations Pages , 377

A Closer Look at the Sun

Atomic Spectral Lines

y [Mm] y [arcsec] brightness temperature at λ = 1.0 mm [103 K] x [arcsec]

Astronomy 421. Lecture 14: Stellar Atmospheres III

The Sun: Our Star. The Sun is an ordinary star and shines the same way other stars do.

arxiv: v1 [astro-ph] 6 Mar 2008

Example: model a star using a two layer model: Radiation starts from the inner layer as blackbody radiation at temperature T in. T out.

Properties of Electromagnetic Radiation Chapter 5. What is light? What is a wave? Radiation carries information

A revolutionizing new view of our Sun with ALMA

First MHD simulations of the chromosphere

Size-dependent properties of simulated 2-D solar granulation

The Persistence of Apparent Non-Magnetostatic Equilibrium in NOAA 11035

Astronomy Chapter 12 Review

Our sole source of light and heat in the solar system. A very common star: a glowing g ball of gas held together by its own gravity and powered

SISD Training Lectures in Spectroscopy

Next quiz: Monday, October 24 Chp. 6 (nothing on telescopes) Chp. 7 a few problems from previous material cough, cough, gravity, cough, cough...

Announcements. - Homework #5 due today - Review on Monday 3:30 4:15pm in RH103 - Test #2 next Tuesday, Oct 11

The Sun. the main show in the solar system. 99.8% of the mass % of the energy. Homework due next time - will count best 5 of 6

The Sun Our Star. Properties Interior Atmosphere Photosphere Chromosphere Corona Magnetism Sunspots Solar Cycles Active Sun

(c) Sketch the ratio of electron to gas pressure for main sequence stars versus effective temperature. [1.5]

Chapter 1. Introduction. Abstract.

Model Atmosphere Codes: ATLAS12 and ATLAS9

Stellar coronae and the Sun

High-speed coronal rain

Model Atmospheres. Model Atmosphere Assumptions

Minicourse on Stellar Activity Stellar Chromospheres

Radiative Transfer and Stellar Atmospheres

Transcription:

The inverse C-shape of solar chromospheric line bisectors H. Uitenbroek National Solar Observatory/Sacramento Peak 1, P.O. Box 62, Sunspot, NM 88349 huitenbroek@nso.edu ABSTRACT Spatially averaged intensity profiles of the chromospheric Na I D and Ca II infrared lines exhibit a pronounced red asymmetry in their cores with bisectors in the shape of an inverse C. This shape stands in stark contrast to the regular C-shape of photospheric spectral line bisectors, which on average exhibit a blue shift as a result of the asymmetry in surface area subtended by convective upflows over downflows. The nature of the inverse chromospheric C-shape is investigated by comparing spatially averaged profiles of the Na I D and Ca II infrared lines with mean profiles computed through three-dimensional snapshots of a hydrodynamic convection simulation and a one-dimensional simulation of chromospheric radiation hydrodynamics. In part the red asymmetry is the result of the asymmetry in time the atmosphere spends in downward motion compared to upward motion when it is traversed by acoustic shocks. Profiles from convection simulations without shocks suggest that convective motions play a limited role in shaping the chromospheric line asymmetry. Further simulations that include effects of both convection and shockwave formation are needed to reach a definitive conclusion on the origin of the inverse C-shaped bisectors. Subject headings: line: formation line: profiles methods: numerical radiative transfer shock waves Sun: chromosphere 1. Introduction With a density that is typically four orders of magnitude lower than that of the underlying photosphere, the solar chromosphere is mostly transparent at optical wavelengths, 1 Operated by the Association of Universities for Research in Astronomy, Inc. (AURA), for the National Science Foundation

2 except for the centers of a handful of strong spectral lines. It becomes optically thick at continuum wavelengths only in the much less accessible regions of microwave and ultraviolet radiation long-ward and short-ward of the optical window. Few viable diagnostics, therefore, are available for routine chromospheric temperature, velocity and magnetic field measurements. Interpretation of these diagnostics is problematic because the low density environment favours radiative transitions over collisional ones, affecting both line excitation and ionization equilibria. As a result, chromospheric lines generally require non-lte radiative transfer solutions to determine the population of their upper and lower levels, their formation heights, and their possible contributions to the radiative gains and losses in the chromospheric energy balance. Furthermore, typical time scales for ionization and recombination (Carlsson & Stein 22) and chemical evolution (Asensio Ramos et al. 23) in the chromosphere are longer than dynamical time scales, so that ionization balances and chemical concentrations can be expected to be out of statistical equilibrium. Finally, whereas the kinetic energy contained in gas motions dominates over magnetic energy in the photosphere, except perhaps in sunspots, the chromosphere encompasses the transition to a low plasma β (8πP gas /B 2 < 1) regime, where plasma motions are dominated by the magnetic field. These theoretical complications combine with the observational difficulties mentioned above to explain why so much detail of chromospheric structure and dynamics is still in dispute. The chromosphere was originally characterized as the layer of the solar atmosphere that gives rise to the crimson off-limb emission in the combined wavelengths of H α and β observed at solar eclipses, but is now often referred to as the layer in which the initial (monotonic, and by assumption ubiquitous) temperature rise from the cool photosphere to the hot corona occurs in one-dimensional hydrostatic semi-empirical models of the solar atmosphere (Gingerich et al. 1971; Vernazza et al. 1973, 1976, 1981; Maltby et al. 1986; Fontenla et al. 199, 1991, 1993). This association with a specific shape of the temperature structure, rather than a temperature regime, has given rise to controversy since the notion of an ubiquitous temperature rise starting at about 5 km above the visible surface, as predicted by the standard models, seems untenable considering the failure of these models to explain the formation of the dark cores of vibration-rotation of the CO molecule observed towards the solar limb (Ayres & Testerman 1981; Ayres 1981; Ayres et al. 1986; Ayres & Wiedemann 1989; Uitenbroek 2; Asensio Ramos et al. 23; Wedemeyer et al. 24). Further doubt was cast on the uniqueness of the hydrostatic models as valid inversions of the spatially and temporally averaged solar spectrum by the results of radiation-hydrodynamic simulations describing the behavior of acoustic waves traveling through solar atmosphere. These simulations clearly indicate that it is possible to have chromospheric emission from an atmosphere which, on average, has no chromospheric temperature rise at all (Carlsson & Stein 1995).

3 In contrast to the semi-empirical hydrostatic models, which essentially constitute multiparameter fits to the observed spatially and temporally averaged spectrum, dynamic ab-initio simulations start out by specifying all the relevant physical processes, and further only specify the appropriate upper and lower boundary conditions to let the equations determine the structure and dynamics of the simulated atmosphere. Excellent examples are hydrodynamic simulations of the solar convection (Nordlund 1982; Stein & Nordlund 1989; Stein & Nordlund 1998) and the radiation-hydrodynamic simulations of the chromosphere (Rammacher & Ulmschneider 1992; Carlsson & Stein 1995, 1997; Skartlien et al. 2; Carlsson & Stein 22; Wedemeyer et al. 24). Since these simulations are usually intended to simulate the general behavior of the solar atmosphere, rather than specific events in it, the validity of such simulations has to be judged by comparing statistical properties of the emergent spectrum, of which the average spectrum is one of the most important ones, as it is for semi-empirical modeling. Because the atmospheric structure and dynamics in simulations directly depend on the details of the included physics these are very helpful in the investigation of what the relevant processes are that shape the solar atmosphere. A clear indication of the realism in current state-of the-art hydrodynamic simulations of solar convection is the close agreement between simulated and observed bisectors of photospheric Fe I lines in the average spectrum (Asplund et al. 2b). Calculated bisectors match the well-known C-shapes of the observed spatially averaged profiles very accurately when a sufficiently fine numerical mesh is employed (Asplund et al. 2a). The characteristic shapes of these bisectors are the result of differences in line strength, continuum level, and area coverage between up- and downflows in the convective motions. Typically, such numerical simulations with spatial resolutions of 2 3 km are able to reproduce both the residual intensity and the FWHM of weak Fe I lines to better than 1%, without any free parameters. One of the most convincing results of chromospheric hydrodynamics calculations is the reproduction of the so-called K 2V phenomenon and the concomitant demonstration that it is the result of shock wave dynamics Carlsson & Stein (1995). This phenomenon represents a very characteristic pattern in the evolution of the line profiles of the H and K resonance lines of singly ionized calcium in non-magnetic supergranulation cell interiors (Rutten & Uitenbroek 1991). It consists of a gradual shift of the K 3 central absorption towards the red, followed by a short brightening of the violet K 2 emission reversal and a sudden shift towards the blue of K 3. This pattern may repeat itself for several cycles with a typical period of three minutes. In Ca II H and K line-core filtergrams the sudden brightenings of the K 2V emission reversal appear as isolated bright points with a typical scale of several arcsecond. The hydrodynamic simulations faithfully reproduce the K 2V phenomenon both qualitatively, and quantitatively (Carlsson & Stein 22), and demonstrate that it is the result of pass-

4 ing acoustic shock waves, with the gradual redshift caused by (almost ballistically) falling material behind the shock, the sudden brightening the result of a combination of the linecore opacity being shifted to the blue and the heating directly behind the shock, and the blueshift of K 3 the direct result of the shock front passage. The K 2V phenomenon results in an asymmetric spatially averaged profile of the H and K lines in the quiet Sun, with a K 2V reversal that is slightly enhanced over K 2R, indicating that the waves that give rise to it are an integral part of chromospheric dynamics. Further proof of this was provided by von Uexkuell & Kneer (1995) who found that, although the K 2V bright points appear to be part of a horizontally extended wave field, the high amplitude oscillations (with intensity 1.5 times the average) that give rise to distinguished bright point features occur only 5-1% of the time in supergranulation cell interiors. Because of the small surface area to which the highest amplitude variations seem to be confined the bright point occurrences appear in only about 2% of the internetwork when integrated over one hour, giving the impression that the wave phenomenon is rare, despite the fact that the general pattern of which the bright points are part is much more common. The notion that these phenomena are part of a larger pattern agrees with the observations of Carlsson et al. (1997) who find that about 5% of the internetwork is covered with threeminute type oscillations as judged from intensity variations of lines and continua in the UV observed with the SUMER instrument on board the SoHO satellite, and the observations of the C II lines at 13.7 nm reported by Wikstøl et al. (2), who find that much of the internetwork region is taking part in these oscillations in both Doppler shift and intensity. Krijger et al. (21) discuss Fourier modulation maps of observations in the TRACE 155, 16, 17 nm passbands, which show extensive area coverage by the 3-min type oscillations in quiet-sun areas. These bands respond to temperature variations in the upper photosphere and temperature minimum region (3 5 km Fossum & Carlsson 24). In this paper we explore the possibility to further constrain chromospheric dynamics by comparing the bisectors of spatially averaged line profiles of the Na I D and Ca II infrared lines with those obtained from dynamic simulations. In this we build on the very successful comparison of observed and computed bisectors of photospheric absorption lines started by Dravins (1987a,b); Dravins & Nordlund (199) and continued by Asplund et al. (2b,a). The pertinent observations are described in Section 2, the model calculations in Section 3, and conclusions are discussed in Section 4.

5 2. Observations of chromospheric redshift The spatially averaged profiles of some chromospheric spectral lines exhibit a pronounced asymmetry towards the red in their cores, especially at disk center. As examples the atlas profiles of the Na I D 1 and Ca II 854.2 nm lines at disk center and towards the limb (at µ =.2) are shown in Figures 1 and 2. The profiles are taken from the Kitt Peak Preliminary Atlas, which tabulates high-quality disk center and µ =.2 spectra of relative intensity from 295 to 18 nm taken with the NSO Fourier Transform Spectrometer by Brault (Brault 1978, 1993). The atlas is provided by Kurucz 2, and is hereafter called the KPNO atlas. These observed profiles are compared with Non-LTE profiles of these lines calculated through the one-dimensional solar atmosphere model FALC (model C of Fontenla et al. 199), which is designed to reproduce the average quiet-sun spectrum. Because this model is in hydrostatic equilibrium, the calculated profiles are exactly symmetric, highlighting by comparison the asymmetry on the red side of the cores of the observed profiles. Since the atlas spectrum is given in relative intensity, it was scaled arbitrarily in intensity to fit the continuum level outside the calculated profiles in all four cases. The red asymmetry of the chromospheric line cores is particularly evident in the shape of the line bisectors drawn with thick curves in Figures 1 and 2 with their difference in wavelength with the synthetic bisectors, the dashed lines which define line centre, amplified by a factor of 2. While the total offset of the observed line bisectors (solid curves) with respect to the computed ones (dashed curves) indicates there may be a discrepancy between the wavelength calibration of the lines in the KPNO atlas and the rest wavelengths calculated from the energy levels in our atomic models, the curved shape of the bisectors most likely represents an intrinsic property of the solar chromosphere. This is certainly the case for the calcium infrared line. In the case of the Na I D 1 line the situation is less clear, but the presence of structure in the disk-center bisector of that line also points in the direction of a true solar origin. The relative shift of the profile of this line close to the limb is more suspicious, since we may expect a zero Doppler shift at the true limb. Photospheric lines show a small redshift at the limb, the so-called limb effect, which is the result of the inversion of the granular temperature contrast in the layer just above the photosphere and the preference of observing receding horizontal motion against the higher temperature over intergranular lanes in this layer (Balthasar 1985). For the rest, the line-of-sight projection of vertical velocities, which decreases with the cosine µ of the viewing angle, should vanish, and line shifts due to horizontal flows should mostly average out. The magnitude of 1.24 pm of the Na I D 1 µ =.2 bisector, corresponding to a redshift 638 m s 1 (comparable to the gravitational redshift of 2 http://kurucz.cfa.harvard.edu/sun/kpnoprelim

6 2. 1 8 Intensity [J m 2 s 1 Hz 1 sr 1 ] 1.5 1 8 1. 1 8 5. 1 9 FALC KPNO limb 4 1 8 Intensity [J m 2 s 1 Hz 1 sr 1 ] 3 1 8 2 1 8 1 1 8 589.5 589.6 589.7 Wavelength [nm] FALC KPNO atlas Fig. 1. Comparison of the calculated Na I D 1 line profile through the semi-empirical model of the average quiet Sun (FAL C) with the disk center (bottom panel) and limb (top panel, corresponding to µ =.2) profile of the KPNO atlas. The line bisectors of the synthetic and atlas profiles are plotted with the thick central curves. The wavelength difference between the atlas bisector and the synthetic one (dashed line), which defines line center, is amplified by a factor of 2 in both panels.

7 2.5 1 8 Intensity [J m 2 s 1 Hz 1 sr 1 ] 2. 1 8 1.5 1 8 1. 1 8 5. 1 9 3.5 1 8 FALC KPNO limb Intensity [J m 2 s 1 Hz 1 sr 1 ] 3. 1 8 2.5 1 8 2. 1 8 1.5 1 8 1. 1 8 5. 1 9 FALC KPNO atlas KPVT 854. 854.1 854.2 854.3 854.4 Wavelength [nm] Fig. 2. Similar to Figure 2, but for the Ca II 854.2 infrared line. The curve labeled KPVT in the bottom panel is the spatially averaged profile from a quiet-sun area observed with the Kitt Peak Vacuum Tower (see text).

8 633 m s 1 ), therefore strongly suggests the presence of an offset between the calculated hydrostatic profile and the atlas introduced by an error in the wavelength calibration of the atlas or in the theoretical rest wavelength (converted to wavelength in air) of the lines, or a combination of both. For each of the two lines separately we can estimate a potential artificial wavelength offset in the following way. If we assume the wavelength mismatch is the same for both µ values, and if we assume that the solar redshift scales linearly with µ, then we can estimate the artificial offset by fitting a straight line through the two maximum values of the bisectors for µ = 1. and µ =.2 and calculate what the intercept is for µ =. In this way we find that the Na I D 1 line in the atlas is shifted by 586 m s 1 towards the red when compared to the calculated FALC profile, and that the Ca II 854.2 line is shifted by 337 m s 1 towards the red. In further comparisons with calculated profiles we will shift the atlas profiles by these amounts to provide a better estimate of the absolute redshift of the two lines. The maximum redshift in the 854.2 bisector, corrected by the estimated artificial offset is 4.58 pm, which amounts to a downward velocity of 1.6 km s 1. The third curve in Figure 2 represents the spatially averaged profile of the λ854.2 line evaluated from a 15 27 section of quiet Sun in a spectroheliogram obtained with the Kitt Peak Vacuum Tower (KPVT) on 21, Apr 18 at a viewing angle corresponding to µ =.93. The data was corrected for dark current, flat fielded and corrected for a 9% scattered light contribution. A line-core intensity map of the observed area is shown in Figure 3, with the rectangle on the left marking the area over which the line profile was averaged to obtain the curve plotted in Figure 2. Like the atlas profiles, the spatially averaged observed profile was scaled to the absolute intensity of the calculated Ca II profile. It closely matches the atlas profile and exhibits the same red asymmetry, confirming that the atlas profile is indeed representative of the average quiet Sun. Since the dispersion relation in these observations was derived from the KPNO atlas we have to shift the average profile to the blue by the same amount as the atlas, namely 337 m s 1, or.96 pm. To confirm that the inverse C-shape of the 854.2 nm bisector is not a peculiarity of this line we plot the bisectors of the other IR triplet lines at 849.8 and 866.2 nm in Figure 4. Since the 866.2 nm line is severely blended in the blue part of its core we can only show the bottom part of its bisector. The 849.8 line is the weakest of the with a gf value that is 9 times smaller than that of the 854.2 line. Clearly all three IR lines show the same pattern. The stronger H and K resonance lines of singly ionized calcium are so much broader than the IR lines that line asymmetries are more difficult to detect, and their average profiles are much more affected by contributions of the magnetic network where the K 2 emission is much enhanced. The average network cell interior profile is asymmetric with a stronger K 2V

9 y [arcsec] 25 2 15 1.9.8.7.6 1 I core /<I cont > 5.5 1 2 3 4 x [arcsec] Fig. 3. Map of the 21 Apr 1 KPVT observation in the core of the Ca II 854.2 nm line. The rectangle on the left indicates the area from which the spatially averaged spectrum in Figure 2 was obtained. For clarity the map shows 1 I core / < I cont >, rather than relative intensity I core / < I cont >, where < I cont > is the continuum intensity averaged over the whole scan. The brightness scale is inverted, so that the brightest regions appear dark in the figure. North is to the right. emission reversal than its counterpart on the red side. This is the result of the wave pattern that gives rise to the K 2V phenomenon (Rutten & Uitenbroek 1991). Of chromospheric lines in the visible other than those of sodium and calcium, solely the hydrogen Balmer α line bisector shows a hint of the pattern, with an amplitude of only.5 pm, corresponding to a downward motion of 228 m s 1. The Mg I b lines do not show the inverse C-shape in their atlas profiles, nor do the Hβ and higher Balmer lines. In the near IR the He I 183. nm line shows mostly outflow in coronal holes, only spatially compact regions with downflows of approximately 2 km s 1, and and no predominant flow velocity in the average quiet Sun. (Dupree et al. 1996). This line tends to form higher than the Ca II lines when the atmosphere is irradiated by EUV and soft X-Ray radiation from the transition region and corona.

1.6.5 866.2 849.8 854.2 Relative Intensity I/I cont.4.3.2.1. 1 2 3 4 5 6 λ [pm] Fig. 4. Bisectors of the disk-centre atlas profiles of all three Ca II IR lines. Only the bottom part of the 866.2 nm line is shown because this line is severely blended with an Fe I line on the blue side of its core. 3. Model calculations 3.1. Non-LTE numerical code and model atoms Because both the Na I D 1 and Ca II 854.2 line emanate from the relatively low density of the chromosphere, their formation has to be treated under the general assumption of Non- LTE, which requires the sodium and calcium population numbers to be solved consistently with the radiation field. The equations of transfer and statistical equilibrium for both lines were solved numerically with the code described by Uitenbroek (21). This code extends the Multi-Level Accelerated Lambda Iteration (MALI) scheme of Rybicki & Hummer (1991, 1992) to include the effects of Partial Frequency Redistribution (commonly abbreviated to PRD), which is needed in the case of the solar Ca II H and K lines, but not for the Na I lines,

11 except very close to the solar limb (Uitenbroek & Bruls 1992). The equations of Non-LTE radiative transfer can be solved with this code in one-, two-, and three-dimensional Cartesian geometry. In multi-dimensional geometry the short characteristics method is used for the formal solution of the transfer equation, as described by Auer et al. (1994) for periodic boundary conditions. In one-dimensional geometry the Feautrier method is used as it was formulated by Nordlund (1982) and Rybicki & Hummer (1991) for the static case (i.e., in case of the profiles from hydrostatic models shown in Figures 1 and 2), and the short characteristics method was used for the dynamic calculation (Section 3.3). For Ca II modeling in one-dimensional geometry a standard five-level plus continuum model (Uitenbroek 1989) was used, including PRD in the H and K resonance lines. The already relatively simple Ca II atomic model was further reduced to a two-line, three-level plus continuum model containing only the K and 854.2 nm lines and their upper and lower levels for solutions in three-dimensional geometry. The profiles of the Na I D lines were calculated with a 12-level, 27-line atomic model (Bruls & Rutten 1992) in one-dimensional geometry. In three-dimensional geometry a smaller Na I atom was employed to remedy the large demand on computer memory placed by such a calculation. For these cases the atom was reduced to a 4-level, 2-line model which included only the D lines and their upper and lower levels ( 2 P and 2 S), and the Na II ground state. Taking out all the Na I levels above the 2 P state results in a severe underestimate in the population number of the sodium 2 S ground state, because it eliminates the level s main repopulation route that has to balance radiative ionization by the super-thermal radiation field in the 2 P bound-free continuum (Bruls & Rutten 1992). To compensate for this underpopulation the abundance of sodium was raised in the three-dimensional calculations by.4 dex from its nominal value of 6.33 (Grevesse & Anders 1991). This value was determined by matching the Na I D profiles calculated with the small atom in the one-dimensional quiet-sun model FALC to the atlas spectrum. In all cases all bound-free continua in the active Ca II and Na I models (i.e., the models that were treated under general Non-LTE conditions) were treated in detail, while bound-free transitions in background atoms were assumed to be in LTE. PRD in the Ca II H and K lines was treated in the angle-averaged approximation, which is valid even in the strong velocity gradients of chromospheric shock dynamics (Uitenbroek 22). 3.2. Line formation in three-dimensional geometry In order to investigate a possible connection between convective motions and the observed redshift in the chromospheric line cores of Na I and CaII the profiles of the Ca II 854.2 and Na I D 1 lines were computed through a snapshot from a simulation of solar convection

12 by Asplund et al. (2b). This simulation extends from approximately 2.8 Mm below to 1 km above the visible surface (the latter defined as the average height of optical depth unity at 5 nm) and had a horizontal grid size of 2 2 points with a spacing of 3 km. Ionization of hydrogen and other elements is treated in LTE, which leads to an underestimate of the hydrogen ionization and the electron density in the upper photosphere and lower chromosphere because the effect of the superthermal Balmer continuum is neglected. For the required Non-LTE transfer computations in calcium and sodium a single snapshot was re-interpolated to a 1 2 82 grid, which extended in height from 35 below to 1 km above the visible surface, and had a horizontal equidistant grid size of 6 km and an equidistant vertical spacing. An angular grid of 3 inclination angles, on a Gauss Legendre quadrature, and 1 azimuth angle per octant was used. The intensity in the vertical direction (which is not part of the angular quadrature) was calculated in a separate formal solution starting from the angular mean radiation field and population numbers given by the full angular solution. In the Asplund et al. (2b) simulation the time-averaged horizontal mean vertical velocity reaches up to 4 m s 1 in the downward direction at the top of the numerical domain of their hydrodynamic simulation (their figure 14, at 1 km). This results from the imposed zero net mass flux at the upper boundary of the simulation box and the typically smaller density and hence larger velocity of the downflowing material. Asplund et al. (2b) considered the net downflow and the resulting predicted red shift of lines forming in the upper part of their simulation (i.e., above approximately 5 km) to be artificial and possibly the result of insufficient numerical resolution. The spatially averaged profile of the Ca II 854.2 line over the whole field-of-view of the snapshot is plotted in Figure 5 with its bisector (thick curve, amplitude enhanced by a factor 2), and with the KPNO atlas profile and its bisector (solid curves). The calculated profile is much darker in its core than the observed one because the hydrodynamic simulation does not have a chromospheric temperature rise anywhere, and underestimates the electron density in the upper part of the atmosphere because hydrogen ionization is treated in LTE. The computed spatially averaged profile is redshifted with respect to the central wavelength of the line, but the amplitude of the bisector is about one fifth of that of the observed one. The convective motions alone, therefore cannot explain the observed red asymmetry in the core of the Ca II 854.2 line. Figure 6 shows the emergent continuum intensity in the vertical direction calculated from the hydrodynamic snapshot (top panel), the velocity determined from the Center-Of- Gravity (COG) position of the whole line (middle panel), and the velocity determined from the position of the line-core minimum of the Na I D 1 line (bottom). The line-core shift

13 3 1 8 Intensity [J m 2 s 1 Hz 1 sr 1 ] 2 1 8 1 1 8 KPNO atlas 3 D avg 854. 854.1 854.2 854.3 854.4 854.5 Wavelength [nm] Fig. 5. Calculated vertically emergent intensity in the Ca II 854.2 nm line averaged over the field-of-view of the (Asplund et al. 2b) snapshot. Line center rest wavelength is indicated with dotted vertical line. Solid curve represents the KPNO atlas profile, and amplitudes of bisectors are multiplied by a factor of 2 for clarity. was determined by fitting a parabola to the central part of the line. The COG method measures the shift of the whole line, weighted towards the core. In the convective flow field the COG determination appears to be mostly sensitive to photospheric motions, since it clearly outlines the granular pattern. By contrast, in photospheric lines the COG method provides good estimates for the actual line-of sight velocity at the depth of line-core formation (Uitenbroek 23). Averaged over the surface of the snapshot the COG method shows an upward velocity of 76 m s 1. For comparison, the COG of the spatial average in the atlas indicates a downward motion of -25 m s 1, when the atlas is corrected for the artificial line offset (Section2), a serious discrepancy. The line-center shift (Figure 6, bottom panel), on the other hand is mostly sensitive to motions in the upper part of the snapshot and shows a pattern that is not easily correlated with the photospheric granulation pattern. The surface

14 y [Mm] 5 4 3 2 1 4.5 1 8 4. 1 8 3.5 1 8 3. 1 8 2.5 1 8 Intensity [J m 2 s 1 Hz 1 sr 1 ] y [Mm] 5 4 3 2 1 4 2 2 4 Velocity [km/s] y [Mm] 5 4 3 2 1 2 2 Velocity [km/s] 1 2 3 4 5 x [Mm] Fig. 6. Calculated emergent continuum intensity (top panel), velocity determined from center-of-gravity of whole line (middle panel), and velocity determined from line-center shift of the Na I D 1 line (bottom). Lighter shades represent upward velocity (blue shift).

15 average of this velocity measure is about 2 m s 1 in the upward direction, half the upward velocity of the mean upward motion at the upper boundary of the simulation box. 4 1 8 Intensity [J m 2 s 1 Hz 1 sr 1 ] 3 1 8 2 1 8 1 1 8 3 D average KPNO atlas 589.54 589.56 589.58 589.6 589.62 589.64 589.66 Wavelength [nm] Fig. 7. Comparison of the average profile of the Na I D 1 line over the surface of the three-dimensional snapshot and the atlas profile, and their respective bisectors (thick curves, amplitude enhanced by a factor of 2). The spatial average of vertically emergent profile of the Na I D 1 line over the whole simulation snapshot, shown in Figure 7 with the dashed curve. The observed and calculated bisectors are very similar in shape, although the calculated has a larger amplitude towards the blue in the photosphere, and towards the red in the upper layers of the simulation. Since we consider the spatially averaged profile of only a single snapshot, we cannot be certain that the calculated profile is offset as a whole by oscillations that are present in the hydrodynamic simulation. These have a larger spatial scale than the granular pattern and move the whole atmosphere up and down by typically ±3 m s 1 (Asplund et al. 2b), basically providing macroscopic broadening to the longterm time averaged profile without changing its shape.

16 3.3. Ca II 854.2 nm line formation in a 1-D radiation-hydrodynamics simulation The three-dimensional radiation-hydrodynamic simulations of Skartlien et al. (2) extend even higher than the ones presented by Asplund et al. (2b), and partly solve the problem of Non-LTE continuum formation by applying a multi-group method for background line scattering (Skartlien 2). Bound-free continua are still treated in LTE, by solving the Saha-Boltzmann equations at the local temperature. From these simulations, which include convection and which have relaxed numerical viscosity to better allow for wave propagation and shock formation, it is clear that above.7-.8 Mm wave motions start to dominate over convective pressure dominated motions (e.g, figures 1 and 11 in Skartlien et al. 2). Thus, it makes sense to investigate what the influence is of these acoustic wave motions on the average chromospheric line profiles considered here. A time series of 376 snapshots from a one-dimensional radiation-hydrodynamic simulation of chromospheric dynamics by Carlsson & Stein (1999) was used to compute the temporally averaged profile of the Ca II 854.2 nm line under the influence of acoustic waves. These simulations consider the evolution and upward propagation of acoustic waves generated at the bottom of the atmosphere. When the acoustic waves travel upward through the gravitationally stratified atmosphere they steepen into shocks which heat the atmosphere and give rise to chromospheric emission. The series of 376 resulting 854.2 nm line profiles is rendered as a time-slice in Figure 8. This particular time series has oscillations of moderate amplitude: the co-temporal evolution of the K line only show clear K 2V grains at 9, 12, 32, 59, and 62 minutes in the sequence. Note that very little intensity variation is present in the wings of the line because the amplitude of the waves is small near the photosphere and because the waves are nearly isothermal due to the very efficient radiative cooling at these heights in the atmosphere. The bisector of the mean profile over the series is given in Figure 9 (dash-dotted curve). It shows that the mean profile of the 854.2 nm line is almost symmetric and does not have any significant redshift. The mean profile of a shorter sequence with much higher amplitude modeling a bright K 2V grain is shown in Figure 1. The profile has a net redshift, but the its shape is qualitatively different from the observed profile having a regular C-shaped bisector, which is however still shifted towards the red as a whole. Examination of a subset of the long lowamplitude sequence containing the K 2V bright points at 9 and 12 min. shows that a high amplitude is a prerequisite for a redshifted mean profile. The mean profile of this partial sequence is still almost symmetric. The reason for the red asymmetry of the high-amplitude mean profile is the behavior of the acoustic wave when it steepens into a shock. How this process explains bright grains in the Ca II K line is explained in detail in Carlsson & Stein

17 6 3. Time [min] 5 4 3 2 2.5 2. 1.5 1. Intensity [1 8 J m 2 s 1 Hz 1 sr 1 ] 1.5 854. 854.2 854.4 854.6 Wavelength [nm] Fig. 8. Evolution of the Ca II 854.2 nm profile in a series of 376 one-dimensional snapshots of a chromospheric dynamic simulation (Carlsson & Stein 1999, top panel).

18 4 1 8 Intensity [J m 2 s 1 Hz 1 sr 1 ] 3 1 8 2 1 8 1 1 8 KPNO atlas Dynamic mean 854. 854.1 854.2 854.3 854.4 854.5 Wavelength [nm] Fig. 9. The bottom panel shows the mean spectrum over 376 snapshots of a time sequence with low amplitude oscillations, and the KPNO atlas profile and their respective bisectors, amplified by a factor of 2. The dotted line indicates line center. (1997). When the shockwave passes through the height in the atmosphere from which the intensity in the inner core of the 854.2 nm line comes this region will experience a short burst of upward motion, resulting in a blue shift. After passage of the shock the material that was thrown up by it falls back accelerating downward until it is swept up by the next shock. This results in a gradually increasing redshift in the inner line core. Depending on the strength of the shock, there is a larger asymmetry between the times the atmosphere is in upward or downward motion. If the shock is weak the wave is almost sinusoïdal and results in a nearly symmetric line profile when averaged over time. Despite the large asymmetry in the times the atmosphere spends in up- and downflow during the passage of a strong shock, the flow is not in conflict with conservation of mass, because the upflow occurs in the dense phase of the wave, while the downflowing material is rarefied. Thus, it seems that the presence of high-amplitude shock waves could explain the red asymmetry in the observed mean chromospheric line profiles.

19 4 1 8 Intensity [J m 2 s 1 Hz 1 sr 1 ] 3 1 8 2 1 8 1 1 8 KPNO atlas Dynamic mean 854. 854.1 854.2 854.3 854.4 854.5 Wavelength [nm] Fig. 1. The bottom panel shows the mean spectrum over 2 snapshots of a short sequence with strong oscillations. The dotted line indicates line center. 4. Discussion and conclusions The spatially averaged profiles of chromospheric absorption lines, especially those of the Ca II IR triplet, and to a lesser degree the sodium D lines, have a central bisector in the shape of an inverse C (Section 2). This is in sharp contrast to the bisectors of photospheric lines, whose regular C-shape is the result of the area asymmetry between up- and downflow in the solar granulation, and the correlation between these flows and intensity. The numerical modeling of sodium and calcium line formation in a three-dimensional hydrodynamic convection simulation in the present paper shows that the inverse C shape may similarly be in part the result of convective motions. The spatially averaged profile of the Na I D 1 line computed through a single snapshot from such a simulation shows a substantial redshift (Section 3.2). However, there is a significant mismatch between the COG velocity of the average observed profile of 25 km s 1 downward, and the average over the FOV of the convection simulation snapshot of 76 km s 1 upward. Moreover, the computations seem to underestimate the amplitude of the stronger Ca II 854.2 nm line bisector towards the red by a factor of five (Figure 5). Asplund et al. (2b) argue that the average mean redshift at the top of their computational is artificial, and probably the result of insufficient numerical resolution and problems

2 with the upper boundary conditions. They note that the downflow is noticeably larger in simulations that extend to only.6 Mm and might be reduced if the simulations could be extended beyond the 1. Mm of the present one. Indeed, their assessment seems to be confirmed by the absence of redshifts in upper-photospheric lines (Allende Prieto & Lopez 1998). Above the chromosphere, in the transition region and corona, ultraviolet emission lines are redshifted by an amount that decreases from about 1 km s 1 for the O IV lines that form at a temperature of 1.86 1 5 to an approximately vanishing shift of the He I line at 58.4 nm which forms at 3.2 1 4 K in the middle to upper chromosphere (Peter & Judge 1999). The maximum observed excursion of the Ca II 854.2 nm bisector of 1.6 km s 1 to the red at diskcenter is probably not inconsistent with this because of the uncertainties in absolute wavelength calibration in both cases, and the fact that the UV shifts are determined from the whole line profile and represents an average over the whole line forming region, whereas the bisector value determined here represents a maximum over such region. Images in diagnostics that arise from 3 5 km above the photosphere are dominated by the so-called inverse granulation pattern Rutten et al. (24, and references therein). This pattern is the result of convective overshoot when upflowing granular material runs into the strong density decline of the gravitationally stratified atmosphere above the convection. The overshooting into the low density overlying region causes rapid horizontal expansion and cooling. The horizontal expansion is slowed down over the intergranular lanes where compression heating takes place, hence the inversion of the granular contrast compared to photospheric layers. The flows in this layer are therefore mostly horizontal and are unlikely to contribute to the asymmetry in the chromospheric lines since it is mostly due to vertical motions (Figure 2). A second possibility is that the inverse C-shape of the chromospheric lines is the result of wave motions of the type that produce the Ca II K 2V phenomenon and the related asymmetry in the spatially averaged H and K line core in non-magnetic areas. When the upward traveling acoustic waves steepen into shocks they give rise to an asymmetry in the time the traversed material spends in upward and downward motion. Thus, the inverse C-shape towards the red in bisectors of chromospheric lines could be the result of this asymmetry in time rather than the asymmetry in surface area which gives rise to the regular C-shaped bisectors towards the blue in photospheric lines. Indeed, the mean profile of the Ca II 854.2 nm line calculated over a series of 2 snapshots from a chromospheric dynamics simulation containing a large amplitude oscillation shows an average redshift, but the central portion of the profile displays a prominent increase in redshift with formation height where a decrease is observed (Section 3.3). An intriguing clue towards the origin of the red asymmetry in the Ca II 854.2 nm

21 line is presented by the behavior of that line in so-called circumfacular regions (Harvey 25). In these areas surrounding active regions on the Sun the 854.2 nm line profile is on average deeper and much more symmetric. This seems to rule out a substantial convective contribution to the average shape of the line since the convective pattern is normal in the areas surrounding active regions, even though its amplitude is suppressed in the active regions themselves. Near active regions however, the plasma β = 1 surface is generally lower (McIntosh & Judge 21). Since acoustic waves suffer conversion to other wave modes at this surface it is possible that in the circumfacular regions acoustic waves are converted before they reach the high amplitudes that preferentially give rise to the red asymmetry in the chromospheric line cores. However, judging from the simulated Na I and Ca II line profiles presented here it seems that neither convective motions nor acoustic waves alone give satisfactory mean line shapes. Since both the three-dimensional convection simulations, in which shockwave formation is suppressed with numerical viscosity, and the one-dimensional radiation-hydrodynamics simulations, which lack the area asymmetry in up and downflows, miss important ingredients that shape the form of chromospheric lines we cannot reach a definitive conclusion about the origin of the inverse C-shape in these lines from the line formation calculations presented here. It is clear that for this purpose we need simulations that include both ingredients, that reach to low enough densities to adequately cover the range in formations heights of the chromospheric lines discussed here, and that include the proper Non-LTE radiative contributions to the energy balance in the chromosphere. I am grateful to Martin Asplund for providing the hydrodynamic snaphshots and to Mats Carlsson and Bob Stein for providing a time-series of snaphots from their chromospheric dynamics simulation. This research has made use of NASA s Astrophysics Data System Bibliographic Services (ADS). REFERENCES Allende Prieto, C., & Lopez, R. J. G. 1998, A&AS, 131, 431 Asensio Ramos, A., Trujillo Bueno, J., Carlsson, M., & Cernicharo, J. 23, ApJ, 588, L61 Asplund, M., Ludwig, H.-G., Nordlund, Å., & Stein, R. F. 2a, A&A, 359, 669 Asplund, M., Nordlund, Å., Trampedach, R., Allende Prieto, C., & Stein, R. F. 2b, A&A, 359, 729

22 Auer, L., Fabiani Bendicho, P., & Trujillo Bueno, J. 1994, A&A, 292, 599 Ayres, T. R. 1981, ApJ, 244, 164 Ayres, T. R., & Testerman, L. 1981, ApJ, 245, 1124 Ayres, T. R., Testerman, L., & Brault, J. W. 1986, ApJ, 34, 542 Ayres, T. R., & Wiedemann, G. R. 1989, ApJ, 338, 133 Balthasar, H. 1985, Sol. Phys., 99, 31 Brault, J. W. 1978, in Future solar optical observations needs and constraints, p. 33 Brault, J. W. 1993, Solar Fourier transform spectroscopy (in Future Solar Optical Obervations Needs, and Constraints 1978), 273 +, Selected Papers on Instrumentation in Astronomy Bruls, J. H. M. J., & Rutten, R. J. 1992, A&A, 265, 257 Carlsson, M., Judge, P. G., & Wilhelm, K. 1997, ApJ, 486, L63 Carlsson, M., & Stein, R. E. 1999, in S. R. Habbal, R. Esser, J. V. Hollweg, P. A. Isenberg (eds.), AIP Conf. Proc. 471: Solar Wind Nine, AIP Conf. Proc., New York, p. 23 Carlsson, M., & Stein, R. F. 1995, ApJ, 44, L29 Carlsson, M., & Stein, R. F. 1997, ApJ, 481, 5 Carlsson, M., & Stein, R. F. 22, ApJ, 572, 626 Dravins, D. 1987a, A&A, 172, 2 Dravins, D. 1987b, A&A, 172, 211 Dravins, D., & Nordlund, Å. 199, A&A, 228, 23 Dupree, A. K., Penn, M. J., & Jones, H. P. 1996, ApJ, 467, L121 Fontenla, J. M., Avrett, E. H., & Loeser, R. 199, ApJ, 355, 7 Fontenla, J. M., Avrett, E. H., & Loeser, R. 1991, ApJ, 377, 712 Fontenla, J. M., Avrett, E. H., & Loeser, R. 1993, ApJ, 46, 319

23 Fossum, A., & Carlsson, M. 24, in ESA SP-547: SOHO 13 Waves, Oscillations and Small-Scale Transients Events in the Solar Atmosphere: Joint View from SOHO and TRACE, p. 125 Gingerich, O., Noyes, R. W., Kalkofen, W., & Cuny, Y. 1971, Sol. Phys., 18, 347 Grevesse, N., & Anders, E. 1991, in A. Cox, W. Livingston, M. Matthews (eds.), Solar interior and atmosphere., University of Arizona Press, Tucson, AZ, p. 1227 Harvey, J. 25, AGU 25 Joint Assembly, New Orleans, Poster SP41B 5 Krijger, J. M., Rutten, R. J., Lites, B. W., Straus, T., Shine, R. A., & Tarbell, T. D. 21, A&A, 379, 152 Maltby, P., Avrett, E. H., Carlsson, M., Kjeldseth-Moe, O., Kurucz, R. L., & Loeser, R. 1986, ApJ, 36, 284 McIntosh, S. W., & Judge, P. G. 21, ApJ, 561, 42 Nordlund, Å. 1982, A&A, 17, 1 Peter, H., & Judge, P. G. 1999, ApJ, 522, 1148 Rammacher, W., & Ulmschneider, P. 1992, A&A, 253, 586 Rutten, R. J., de Wijn, A. G., & Sütterlin, P. 24, A&A, 416, 333 Rutten, R. J., & Uitenbroek, H. 1991, Sol. Phys., 134, 15 Rybicki, G. B., & Hummer, D. G. 1991, A&A, 245, 171 Rybicki, G. B., & Hummer, D. G. 1992, A&A, 262, 29 Skartlien, R. 2, ApJ, 536, 465 Skartlien, R., Stein, R. F., & Nordlund, Å. 2, ApJ, 541, 468 Stein, R. F., & Nordlund, Å. 1989, ApJ, 342, L95 Stein, R. F., & Nordlund, Å. 1998, ApJ, 499, 914 Uitenbroek, H. 1989, A&A, 213, 36 Uitenbroek, H. 2, ApJ, 536, 481 Uitenbroek, H. 21, ApJ, 557, 389

24 Uitenbroek, H. 22, ApJ, 565, 1312 Uitenbroek, H. 23, ApJ, 592, 1225 Uitenbroek, H., & Bruls, J. H. M. J. 1992, A&A, 265, 268 Vernazza, J. E., Avrett, E. H., & Loeser, R. 1973, ApJ, 184, 65 Vernazza, J. E., Avrett, E. H., & Loeser, R. 1976, ApJS, 3, 1 Vernazza, J. E., Avrett, E. H., & Loeser, R. 1981, ApJS, 45, 635 von Uexkuell, M., & Kneer, F. 1995, A&A, 294, 252 Wedemeyer, S., Freytag, B., Steffen, M., Ludwig, H.-G., & Holweger, H. 24, A&A, 414, 1121 Wikstøl, Ø., Hansteen, V. H., Carlsson, M., & Judge, P. G. 2, ApJ, 531, 115 This preprint was prepared with the AAS L A TEX macros v5.2.