Granulation in DA white dwarfs from CO5BOLD 3D model atmospheres

Similar documents
Oxygen in the Early Galaxy: OH Lines as Tracers of Oxygen Abundance in Extremely Metal-Poor Giant Stars

Astronomy. Astrophysics. Spectroscopic analysis of DA white dwarfs with 3D model atmospheres

3D MODEL ATMOSPHERES FOR EXTREMELY LOW-MASS WHITE DWARFS

CO5BOLD Workshop 2012

The instability strip of ZZ Ceti white dwarfs

3D hydrodynamical CO 5 BOLD model atmospheres of red giant stars

arxiv: v1 [astro-ph.sr] 23 May 2016

2. Stellar atmospheres: Structure

Model Atmospheres. Model Atmosphere Assumptions

Size-dependent properties of simulated 2-D solar granulation

Spots on the surface of Betelgeuse Results from new 3D stellar convection models

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

3D non-lte corrections for the 6 Li/ 7 Li isotopic ratio in solar-type stars

arxiv: v2 [astro-ph.sr] 31 Mar 2014

THE OBSERVATION AND ANALYSIS OF STELLAR PHOTOSPHERES

3D Model Atmospheres of Red Giant Stars

arxiv: v1 [astro-ph.sr] 7 Jan 2019

arxiv: v2 [astro-ph.sr] 6 Oct 2014

50F-1650, 1 Cyclotron Rd, Berkeley, CA , USA

New Dimensions of Stellar Atmosphere Modelling

A photometric analysis of ZZ Ceti stars: A parameter-free temperature indicator?

Observed Properties of Stars - 2 ASTR 2120 Sarazin

ASTRONOMY AND ASTROPHYSICS. Acoustic wave energy fluxes for late-type stars. II. Nonsolar metallicities

VLTI/AMBER studies of the atmospheric structure and fundamental parameters of red giant and supergiant stars

Limb Darkening. Limb Darkening. Limb Darkening. Limb Darkening. Empirical Limb Darkening. Betelgeuse. At centre see hotter gas than at edges

arxiv: v2 [astro-ph.sr] 19 Feb 2018

Measuring Radial & Tangential Velocity. Radial velocity measurement. Tangential velocity measurement. Measure the star s Doppler shift

ASTR Look over Chapter 15. Good things to Know. Triangulation

Magnetic flux tubes observed with THEMIS/MSDP

Chandra Spectroscopy of the Hot DA White Dwarf LB1919 and the PG1159 Star PG

Physical parameter determinations of young Ms

Chapter 15 Surveying the Stars Pearson Education, Inc.

Lecture 26 The Hertzsprung- Russell Diagram January 13b, 2014

Stars: basic observations

Spectroscopy of giants and supergiants! Maria Bergemann MPIA Heidelberg"

The Hertzprung-Russell Diagram. The Hertzprung-Russell Diagram. Question

Observed Properties of Stars - 2 ASTR 2110 Sarazin

A new extensive library of synthetic stellar spectra from PHOENIX atmospheres and its application to fitting VLT MUSE spectra

Lines of Hydrogen. Most prominent lines in many astronomical objects: Balmer lines of hydrogen

Turbulence models and excitation of solar oscillation modes

First MHD simulations of the chromosphere

From theory to observations

Mira variables They have very regular light curves, with amplitude V > 2.5 m, and periods Π > 100 days.

Binary Stars (continued) ASTR 2120 Sarazin. γ Caeli - Binary Star System

Chapter 9. Stars. The Hertzsprung-Russell Diagram. Topics for Today s Class. Phys1411 Introductory Astronomy Instructor: Dr.

arxiv:astro-ph/ v1 10 Nov 2006

ASTRONOMY AND ASTROPHYSICS. Synthetic photometry from ATLAS9 models in the UBV Johnson system. Fiorella Castelli

chapter 31 Stars and Galaxies

ASTR-1020: Astronomy II Course Lecture Notes Section III

arxiv:astro-ph/ v1 17 Nov 2005

Measuring Radial & Tangential Velocity. Radial velocity measurement. Tangential velocity measurement. Measure the star s Doppler shift

Mass-Luminosity and Stellar Lifetimes WS

UNIT 3: Astronomy Chapter 26: Stars and Galaxies (pages )

Stars and Galaxies. The Sun and Other Stars

Unit 2 Lesson 2 Stars. Copyright Houghton Mifflin Harcourt Publishing Company

A study of accretion disk wind emission

THE NEWLY DISCOVERED PULSATING LOW-MASS WHITE DWARFS: AN EXTENSION OF THE ZZ CETI INSTABILITY STRIP

STELLAR PARAMETERS OF M DWARFS FROM LOW AND HIGH-RESOLUTION SPECTRA TOGETHER WITH NEW MODEL ATMOSPHERES

Fundamental stellar parameters

ASTROPHYSICS. K D Abhyankar. Universities Press S T A R S A ND G A L A X I E S

The physics of stars. A star begins simply as a roughly spherical ball of (mostly) hydrogen gas, responding only to gravity and it s own pressure.

Stellar Astrophysics: The Classification of Stellar Spectra

L = 4 d 2 B p. 4. Which of the letters at right corresponds roughly to where one would find a red giant star on the Hertzsprung-Russell diagram?

L = 4 d 2 B p. 1. Which outer layer of the Sun has the highest temperature? A) Photosphere B) Corona C) Chromosphere D) Exosphere E) Thermosphere

From theory to observations

Energy transport: convection

Summer 2013 Astronomy - Test 3 Test form A. Name

Evolution Beyond the Red Giants

A NEW ZZ CETI WHITE DWARF PULSATOR: G30-20 Anjum S. Mukadam, 1,2 S. O. Kepler, 3 D. E. Winget, 1,2 and P. Bergeron 4

CASE STUDY FOR USE WITH SECTION B

Astronomy 1 Fall 2016

Stellar Structure and Evolution

Our sun is the star in our solar system, which lies within a galaxy (Milky Way) within the universe. A star is a large glowing ball of gas that

arxiv: v1 [astro-ph.sr] 4 Sep 2011

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

Chapter 7: From theory to observations

Stellar Spectra ASTR 2120 Sarazin. Solar Spectrum

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

Types of Stars and the HR diagram

Two-dimensional imaging of the He D 3 /Hβ emission ratio in quiescent solar prominences

A-star envelopes: a test of local and non-local models of convection

Chapter 15 Surveying the Stars

Chapter 15: Surveying the Stars

HS a DAZB white dwarf of very unusual composition

If a star is very hot, the electrons will be freed from the hydrogen atom. (Ionized) Once they are free, they act like particles and emit a

Limb darkening laws for two exoplanet host stars derived from 3D stellar model atmospheres

Massimo Meneghetti 1, Elena Torri 1, Matthias Bartelmann 2, Lauro Moscardini 3, Elena Rasia 1 and Giuseppe Tormen 1,

Calibrating Core Overshooting in Low-Mass Stars with Kepler Data

dp dr = GM c = κl 4πcr 2

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

Gravitational redshifts, and other wavelength shifts in stellar spectra

The Distance Modulus. Absolute Magnitude. Chapter 9. Family of the Stars

Chapter 8: The Family of Stars

Lecture 16 The Measuring the Stars 3/26/2018

Lecture Outlines. Chapter 17. Astronomy Today 8th Edition Chaisson/McMillan Pearson Education, Inc.

The Cosmic Perspective. Surveying the Properties of Stars. Surveying the Stars. How do we measure stellar luminosities?

LESSON 1. Solar System

Parallax: Measuring the distance to Stars

Exam #2 Review Sheet. Part #1 Clicker Questions

Irradiated brown dwarfs

Transcription:

Mem. S.A.It. Suppl. Vol. 24, 61 c SAIt 2013 Memorie della Supplementi Granulation in DA white dwarfs from CO5BOLD 3D model atmospheres P.-E. Tremblay 1, H.-G. Ludwig 1, B. Freytag 2, and M. Steffen 3 1 Zentrum für Astronomie der Universität Heidelberg, Landessternwarte, Königstuhl 12, D-69117 Heidelberg, Germany, e-mail: ptremblay@lsw.uni-heidelberg.de 2 CNRS, Université de Lyon, École Normale Supérieure de Lyon, 46 allée d Italie, F-69364 Lyon Cedex 7, France 3 Leibniz-Institut für Astrophysik Potsdam, An der Sternwarte 16, D-14482 Potsdam, Germany Abstract. Time-dependent 3D simulations of pure-hydrogen DA white dwarf atmospheres have been computed in recent years. Synthetic Balmer lines spectra drawn from these radiation-hydrodynamics (RHD) simulations have been shown to predict surface gravities significantly lower than the standard 1D models, in much better agreement with the expectation that white dwarfs cool at constant mass. We have now computed a grid of CO5BOLD pure-hydrogen 3D model atmospheres for surface gravities from log g = 7 to log g = 8.5 and effective temperatures from 6000 to 13,000 K. Over this range, we observe a significant variation of the intensity contrast of the surface granulation patterns, which indicates the strength of the 3D effects. Furthermore, the size and appearance of granules are also varying considerably. An explanation of these behaviours can lead to a better understanding of the physical processes responsible for the energy transfer in white dwarf atmospheres. Key words. convection line: profiles stars: atmospheres white dwarfs 1. Introduction The comparison of the observed and predicted Balmer line profiles has been the most successful method to derive the atmospheric parameters (T eff and log g) of DA white dwarfs (Bergeron et al. 1992; Finley et al. 1997). While hot white dwarfs (T eff > 13,000 K) have a mean spectroscopic mass in the range of 0.61-0.63 M (Tremblay et al. 2011a; Gianninas et al. 2011; Kleinman et al. 2012), cool white dwarfs below T eff < 13,000 K observed in the same surveys show a significant Send offprint requests to: P.-E. Tremblay increase of up to 20% in their mean mass. White dwarf stars are expected to cool at constant mass and almost constant radius, hence the increase in mass is likely caused by inaccuracies in the model atmospheres (Tremblay et al. 2010; Koester et al. 2009). Tremblay et al. (2011b) demonstrated that the inability of the 1D mixing-length theory (Böhm-Vitense 1958, hereafter MLT) to properly account for the convective energy transport was the source of the long-standing high-log g problem. In this first experiment, four 3D CO 5 BOLD (Freytag et al. 2012) models in a limited range of temperature (12,800 > T eff (K) > 11,300) were computed, and model spectra derived from

62 Tremblay et al.: Granulation in DA white dwarfs these RHD simulations were shown to predict lower masses than 1D models. We have recently expanded the CO 5 BOLD computations to a set of 48 DA model atmospheres in the range of 7.0 < log g < 8.5 and 13, 000 > T eff (K) > 6000 1. Fig. 1 presents the mass distribution of the SDSS spectroscopic sample (Eisenstein et al. 2006) adopting model spectra either from 1D (Tremblay et al. 2011a) or 3D structures (Tremblay et al. 2013). By relying on 3D spectra, the masses are clearly much more in agreement with the mean mass for hotter radiative white dwarfs (horizontal dotted line in Fig. 1). The 3D models therefore possess a great potential to study white dwarfs, although we are just starting to learn about the physical processes responsible for the differences between mean 3D and 1D atmospheric structures. As an initial step, we look in this work at the spatially resolved emergent intensity and the size of the convective cells. These quantities can not be derived from existing plane-parallel 1D structures and can provide insights on the convective processes in white dwarfs. We start in Sect. 2 by describing briefly the numerical setup of our CO 5 BOLD simulations. Then, we characterize in more detail the properties of the surface granulation in terms of intensity contrast (Sect. 3) and characteristic size (Sect. 4) as derived from our RHD models. The conclusion follows in Sect. 5 2. Grid of 3D model atmospheres Our CO 5 BOLD simulations rely on the box-ina-star setup and cover the entire atmospheres ( 5 < log τ R < 3). The lateral boundaries are periodic, and the top boundary is open to material flows and radiation. The bottom layer is closed to incoming material (but open to radiation) for the 6, 4, 3 and 1 hottest models at log g = 7.0, 7.5, 8.0 and 8.5, respectively. In contrast, the bottom layer is open to convective flows (and radiation) in all other models, and a zero total mass flux is enforced. We adopt a grid of 150 150 150 points for all models. 1 We use a step of 0.5 dex in log g and temperature steps of 500 K (above T eff = 9000 K) and 1000 K (below this temperature). Fig. 1: Mass distribution as a function of T eff for the SDSS sample (S/N > 15) derived from spectroscopic fits using 3D models (top panel) and 1D models (bottom panel). A horizontal dotted lines at the mean mass of the sample (0.613 M, for 40,000 > T eff (K) > 13,000) has been added as a reference. Details about the input microphysics and numerical parameters will be given elsewhere (Tremblay et al. 2013). We mention, however, that the microphysics is the same as in standard 1D models (Tremblay et al. 2011a). Furthermore, we have verified that the mean 3D structures were not sensitive to variations of the numerical parameters (e.g., size of the boxes and viscosity), unlike the 1D models which are very sensitive to the parametrization of the MLT. In Fig. 2 we present 3D structures at log g = 8, which are averages in time and space over surfaces of constant Rosseland optical depth, and compare them to 1D structures, adopting the ML2/α = 0.8 parametrization of the MLT. There are clearly systematic differences between both physical models of convection. The most striking difference is in the upper part of the atmospheres, where the temperatures are always much smaller in 3D models, which is the result of convective overshoot. Furthermore, the 1D models feature an abrupt transition from a convective zone to a purely radiative equilibrium at τ R 0.1, while

Tremblay et al.: Granulation in DA white dwarfs 63 3D models are smoother in this region because of convective overshoot. Fig. 2: Temperature structures versus log τ R for 3D (red, solid lines) and 1D (black, dashed lines) model atmospheres. From bottom to top, T eff values are from 6000 K to 8000 K (steps of 1000 K) and from 9000 K to 13,000 K (steps of 500 K). The temperature scale is correct for the 6000 K model, but other structures are shifted by 1 kk for clarity. 3. Intensity contrast The intensity contrast of the surface granulation is one measure of the strength of 3D effects. It must be understood, however, that one can not easily rely on the intensity contrast to predict the resulting 3D effects on the mean structures such as those presented in Fig. 2, since local effects may cancel during the averaging procedure. Nevertheless, a study of the granulation properties can provide insights on the physical processes responsible for the formation and evolution of convective cells and in an indirect way, on the energy transfer in the atmospheres. In Fig. 3, we present gray (bolometric) emergent intensity snapshots for 12 of our simulations. For each simulation, we rely on a time average of 250 snapshots similar to those in Fig. 3 to derive quantities presented in this work. The root-mean-square (RMS) relative intensity contrast is shown in Fig. 4 as a function of T eff (left panel). For all gravities, the intensity contrast resides in the 5-20% range, except for the low and high T eff extremes. This implies that the emergent flux varies significantly over position and time on the disk of these white dwarfs and that the 1D plane-parallel approximation is poor for the resolved atmospheres. Very similar values of the intensity contrast are found for A, F, G and K dwarfs (see Fig. 15 of Freytag et al. 2012). We find that the granulation is visually similar (see Fig. 3) in M dwarfs and 6000 K white dwarfs, and in A dwarfs and 12,500 K white dwarfs. From Fig. 4 (left panel), one could conclude from the shift of the curves at different log g values that the surface gravity has a significant impact on 3D effects. However, when looking at the intensity contrast as a function of the density in the photosphere, (Fig. 4, right panel), the relation becomes much more similar for all gravities. Therefore, the granulation properties appear to be better described by the characteristic photospheric density than the T eff and log g atmospheric parameters. This may also be part of the explanation why the intensity contrast is similar in (hot) white dwarfs and convective main-sequence stars, since the atmospheric densities are of the same order of magnitude. At the hot end of our grid, the intensity contrast is expected to decrease because the convective to radiative flux ratio becomes very small. At the opposite cool end of our sequence, convection is very efficient to transport the small total flux with only very small temperature variations. In between these two extremes, the increasing total flux contributes to enhance the contrast. Furthermore, as T eff is increasing, the Rosseland opacity is also increasing in the photospheres. This implies lower characteristic densities in the photospheres (see Fig. 4) and higher convective velocities. This effect also contributes in enhancing the fluctuations. The MLT qualitatively account for these effects when predicting the convective velocities.

64 Tremblay et al.: Granulation in DA white dwarfs 3 1.0 0.8 2 0.6 1 0.4 1 2 3 0.4 0.6 0.8 1.0 0.3 0.1 0.1 0.3 1.5 4 3 1.0 2 0.5 1 1 2 3 4 0.5 1.0 1.5 0.4 0.3 0.1 0.1 0.3 0.4 20 2.5 2.0 15 10 1.5 1.0 5 0.5 0 0 5 10 15 20 0.5 1.0 1.5 2.0 2.5 0.8 0.6 0.4

Tremblay et al.: Granulation in DA white dwarfs 65 Fig. 4: RMS intensity contrast divided by the mean intensity as a function of T eff (left panel) and log ρ (evaluated at τ R = 1, right panel) for our sequence of 3D models at log g = 7.0 (open points, red), 7.5 (filled points, green), 8.0 (squares, blue) and 8.5 (triangles, magenta). The points are connected for clarity. 4. Granulation size The 3D models have horizontal sizes, while such thing does not exist in 1D plane-parallel models. Calculations of stellar RHD models have shown that granular patterns have larger horizontal dimensions of the order of 10 times the local value of H p (Freytag et al. 1997). In Fig. 5, we present the power spectrum as a function of the horizontal wavenumber, for the emergent intensity of the 10,000 K and log g = 8 simulation. We did such Fourier transforms for all of our simulations, and we found that the structure of the power spectrum is very similar in all cases. The maxima of the power spectra are well fitted by a parabola, from which we derived the characteristic size of the granulation as the wavelength of the peak of the power spectrum. We note that this characteristic size is more sensitive to numerical parameters than other quantities derived from our simulations. For instance, we expect that grid resolution and viscosity will have an effect on the size of the Fig. 5: Mean power spectrum as a function of the horizontal wavenumber (2π/λ) averaged over 250 snapshots of the (bolometric) intensity map of the 10,000 K 3D simulation (log g = 8). granules. Therefore, our values should be taken as estimates only 2. 2 We discard the two hottest models at log g = 7 since convection is too weak to clearly differentiate

66 Tremblay et al.: Granulation in DA white dwarfs The ratio of the characteristic granulation size to the pressure scale height in the photosphere is presented in Fig. 6. It demonstrates that the characteristic cells dimension is not only a function of local the pressure scale height. The size of granules decreases much more rapidly with decreasing T eff than what we would expect from simple thermodynamics considerations. This effect may be part of the explanation, along with the effect from velocities, for the change in the characteristic size of the granulation. We note that the Péclet number is varying significantly in the photospheres, typically by one order of magnitude or more. Hence the values given in Fig. 7 for one reference layer should be taken as estimates only of the importance of radiative and advective processes in the full photospheres. Fig. 6: Ratio of the characteristic granular size (maximum of the mean power spectrum) to the pressure scale height at τ R = 1 as a function of T eff for our sequence of 3D models. The points are connected for clarity. The color code is the same as in Fig. 4. Fig. 7: Logarithm of the Péclet number in the photosphere (τ R = 1) of our simulations. The color code is the same as in Fig. 4. The Péclet number in the photosphere (τ R = 1) which is the ratio of the radiative timescale to the advective timescale (Ludwig & Kučinskas 2012), is shown on Fig. 7. The ratio is changing from a value below to above unity between the hotter and cooler models, respectively. This implies that the evolution of convective cells in the photosphere will be largely influenced by radiation in the hot models in our grid, whereas for the cool models, convective cells are more directly governed by quasi-adiabatic expansion and compression. granules from the effects of p-mode pulsations and possible unphysical numerical fluctuations. Fig. 8 presents a measure of the asymmetry between the dark and bright cells in our simulations, with the surface area fraction occupied by bright cells (I > Ī). It is clear that there is a trend for the bright cells to occupy more area in hotter models. This is also seen in the intensity snapshots of Fig. 3, with large bright cells and narrow dark lanes appearing in hot models. This effect can explain in part the increasing characteristic size for hotter models. A similar effect is observed for A-type stars and red supergiants when simulated with CO 5 BOLD (Freytag et al. 2012). This can be partly attributed to the faster moving downdrafts, that correspond to the small shallow surface granules.

Tremblay et al.: Granulation in DA white dwarfs 67 calculations have been performed on CALYS, a mini-cluster of 320 nodes built at Université de Montréal with the financial help of the Fondation Canadienne pour l Innovation. We thank Prof. G. Fontaine for making CALYS available to us. We are also most grateful to Dr. P. Brassard for technical help. This work was partially supported by Sonderforschungsbereich SFB 881 The Milky Way System (subproject A4) of the German Research Foundation (DFG). Fig. 8: Fraction of the geometrical surface (A) where the intensity is higher than the average (I + ). The color code is the same as in Fig. 4. 5. Conclusion We presented a grid of CO 5 BOLD 3D model atmospheres covering most of the range of observed atmospheric parameters for convective DA white dwarfs down to T eff = 6000 K. We relied on these simulations to look at the properties of the surface granulation. We demonstrated that the emergent intensity contrast is mostly a function of the density in the photosphere, and that granulation in white dwarfs is qualitatively similar to what is observed in the Sun. We have shown that the size of granules is not well scaled by the local pressure scale height, and we believe that this is due to increasing velocities (Mach number) and to the variation of the importance of the radiative versus advective transport of energy. Future works will compare in more details the properties of CO 5 BOLD simulations for white dwarfs, solar-type stars and giants. Acknowledgements. P.-E. T. is supported by the Alexander von Humboldt Foundation. 3D model References Bergeron, P., Saffer, R. A., & Liebert, J. 1992, ApJ, 394, 228 Böhm-Vitense, E. 1958, ZAp, 46, 108 Eisenstein, D. J., Liebert, J., Harris, H. C., et al. 2006, ApJS, 167, 40 Finley, D. S., Koester, D., & Basri, G. 1997, ApJ, 488, 375 Freytag, B., Holweger, H., Steffen, M., & Ludwig, H.-G. 1997, in Science with the VLT Interferometer, ed. F. Paresce (New York: Springer), 316 Freytag, B., Steffen, M., Ludwig, H.-G., et al. 2012, Journal of Computational Physics, 231, 919 Gianninas, A., Bergeron, P., & Ruiz, M. T. 2011, ApJ, 743, 138 Kleinman, S. J., Kepler, S. O., Koester, D., et al. 2012, arxiv:1212.1222 Koester, D., Voss, B., Napiwotzki, R., et al. 2009, A&A, 505, 441 Ludwig, H.-G., & Kučinskas, A. 2012, A&A, 547, A118 Tremblay, P.-E., Bergeron, P., Kalirai, J. S. & Gianninas, A. 2010, ApJ, 712, 1345 Tremblay, P.-E., Bergeron, P. & Gianninas, A. 2011a, ApJ, 730, 128 Tremblay, P.-E., Ludwig, H.-G., Steffen, M., Bergeron, P., & Freytag, B. 2011b, A&A, 531, L19 Tremblay, P.-E., Ludwig, H.-G., Steffen, M., & Freytag, B. 2013, submitted for publication in A&A