Download as pdf
Download as pdf
You are on page 1of 9
Chapter 14 Stellar Atmospheres ‘The phenomena that affect the transmission of light through planetary atmospheres also operate in stellar atmospheres. For planets, our interest in infrared light led us to focus on molecular vibrations. For stars, we are more interested in visible light so our attention shifts to electron excitation and ionization. In this chapter we study how those processes affect the absorption lines that appear in stellar spectra. 14.1 Atomic Excitation and Ionization Observed spectra of stars resemble blackbody spectra modified by absorption bands at specific wavelengths (see Figs. 14.1 and 14.2). The absorption lines are produced when light that originates from some modest depth passes through the star’s outer layers.! The gas in stars is predominantly hydrogen, so we might expect to see strong hydrogen absorption lines. Recall from Eq.(13.22) that the transition between hydrogen energy levels n and m corresponds to wavelength Yum = 911A (5s - ay ‘Thus the locations of hydrogen lines are as follows (with wavelengths in A): m=2 3 4 5 n=1 1216 1,026 973 949) “Lyman series” 2 6,563 4,861 4,340 mer series” 3 18,750 12,818 ‘Paschen series” tas sar photsphee, or apparent “surface comesponds to an optical depth of order unity (see 13.3.3). C. Keeton, Principles of Astrophysics: Using Gravity and Stellar Physics to Explore 285 the Cosmos, Undergraduate Lecture Notes in Physics, DOI 10.1007/978-1-4614-9236-8_14, © Springer Science+Business Media New York 2014 286 14 Stellar Atmospheres 06.5 Bo 86 At FO F5 co Gs ko Ks. oT Tia Mo Ms Fig. 14.1 Examples of absorption line spectra from different types of stars. The codes on the left refer to the spectral classification scheme discussed in Sect. 14.2 (the numbers indicate subcategories). The strong lines in the spectrum labeled A correspond to hydrogen. Moving down the sequence, the peak of the spectrum shifts to longer wavelengths and the hydrogen lines become less prominent (Credit: NOAO/AURA/NSF) intensity wavelangth (A) Fig. 14.2 A different view of spectra for different types of stars, from © (top) to K (bottom). M siars are not shown here. The curves show intensity versus wavelength, and are offset vertically for clarity (Data from Pickles [1]) 14.1 Atomic Excitation and Ionization 287 ‘The Balmer lines are the ones that appear at optical wavelengths. We do see them in some stellar spectra (e.g., the top half of Fig. 14.1), but not in others. Can we understand why? 14.1.1 Energy Level Occupation Consider an absorption line corresponding to an atomic transition from energy level 1 to energy level m > n. In order for this line to be strong, two conditions must hold: 1. The atom in question must be abundant. 2. In a non-negligible fraction of atoms, level n must be occupied but level m must not be full. Consider point #2. In isolation, an atom generally has all of its electrons in the lowest possible energy levels. In a gas, however, atoms occasionally bump into each other, and some of the kinetic energy can be transferred to internal energy. The temperature and associated kinetic energy in stellar atmospheres are high enough that collisions can excite electrons into higher energy levels. The amount of collisional excitation depends on the kinetic energy, which in turn depends on the temperature. To quantify this effect, recall from Eq. (12.1) that the probability for level n to be occupied is Py & gy C7 ER/AT where Ey is the energy and g, is the statistical weight, which in this case counts the number distinct quantum states in the energy level. The ratio of the number of atoms with level m occupied to the number with level n occupied is then fa tte tir aan This Boltzmann equation allows us to determine which energy levels are signifi- cantly occupied and analyze whether absorption lines will be strong or weak. 14.1.2 Ionization Stages A collision with sufficient energy can kick an electron out of an atom altogether. This influences absorption lines because electrons that are not in atoms cannot produce atomic transitions, and the loss of an electron may modify the energy levels for any remaining electrons in an atom. It is conventional to label different ionization stages with Roman numerals: 288 14 Stellar Atmospheres 1. Neutral Il. Singly-ionized Il. Doubly-ionized and so on, Consider a transition that starts with a neutral atom (stage 1) whose most loosely bound electron has energy Ey = —ys, where x; > 0 is the ionization energy, and ends with a stage II ion plus a free electron with speed v- or energy E1; = m,v*/2. The energy difference between the two states is AE = Ey; — E; = x1 +m,v"/2, so the relative abundance of ionized and neutral atoms has a Boltzmann factor of the form Nu Xr tmv/2 Wr on( kT (We will deal with statistical weights in a moment.) We do not actually care what the final speed of the electron is, so we can integrate over all velocities: Nu y ie f si _ da + mev?/2) PET 20 ox ek f emeFPET gary? ay lo ap « (=) emukT Me The key factors here are e~#"/*T and (kT)*/2. ‘There are a few more details that enter a complete analysis. First, the number density of free electrons must play a role, It appears in the denominator because free electrons can combine with ions and return them to a neutral stage. Second, where we used statistical weights to count states for the Boltzmann equation, we must now use a more general counting that is done with the partition function, ~ Z = Dan AT (14.2) where Ey is the energy of the ground state. This is basically the sum of the number ‘of ways the atom can arrange its electrons, with more energetic (and therefore less likely) configurations receiving less weight from the Boltzmann factor. Careful counting reveals that the ratio of the number of atoms in ionization stage II to the number in stage I is, 2Partition functions are often studied in courses on statistical mechanics. 14.1 Atomic Excitation and Ionization 289 Nu _ 2211 (mkT\? ar Monti (Ga) © ans This is called the Saha equation after Meghnad Saha. Hydrogen only has two ionization stages: neutral (H 1) and ionized (H Il). For heavier elements, higher ionization stages are possible, and the Saha equation can be applied to them as well. For example, the ratio of doubly ionized atoms (like He III) to singly ionized atoms (like He I) would be described by an equation like (14.3) but with the second ionization energy x11 and the partition functions Z1; and Z1 (see Problem 14.1), 14.1.3 Application to Hydrogen Let's use these ideas to study hydrogen Balmer lines. These occur when an electron jumps from level n = 2 up to level m > 2, so they are strong only if a reasonable fraction of hydrogen atoms have electrons in the n = 2 level. Before we can use the Boltzmann equation to compute that fraction, we need to specify the statistical weight for hydrogen: &n = 2n? The n? comes from the orbital quantum numbers, while the 2 comes from spin. Using To = 5,780 K for the Sun yields the following numbers: kT = 7.98 x 10-J = 0.498eV E, =—13.6eV Ey =-3.4eV gi =2 a= 8 = BAB er evrr = 5x 10 (144) m gi Only a tiny fraction of hydrogen atoms in the outer layers of the Sun are excited to n = 2, which is why Balmer lines are not very prominent in the spectrum of the ‘Sun. The excitation fraction increases with temperature, as shown in Fig. 14.3, so in general we would expect Balmer lines to be more prominent in hotter stars. However, as the temperature increases, so too does the ionization fraction. At some point ionization must overcome excitation and prevent hot stars from producing any hydrogen absorption lines. We can investigate this using the Saha equation, but first we need to determine the partition functions. The starting point is neutral hydrogen, whose partition function is 290 14 Stellar Atmospheres kT (eV) 04 0.6 08 12 no/ny for hydrogen 4000 6000-~—«8000-~—=«10000+~=«12000-~—=—«14000 temperature T (I) Fig. 14.3. Excitation ratio n2/n, for hydrogen, as a function of temperature. The value for the Sun is marked Zy = gr + Bye EDT 4. gy (EET 4 =n (: 4 82 EEDA Beteravr 4...) a gi ea(i+2e24..) mom We just found that at temperatures relevant for the surface of stars, most of the hydrogen atoms are in the ground state. Thus 12/7; is small, n3/7) is even smaller, and we have Z; © g, = 2. The ending point is a bare proton, which has only one possible state so Z;; = 1. ‘We want determine the ionization fraction, X. Let the number densities of hydrogen ions and neutral atoms be n ; and n;, respectively, and the total number density of hydrogen be ma = ny + 171. By the definition of the ionization fraction, nj; = Xi and ny = (1 — X)no. Then the left-hand side of the Saha equation (14.3) becomes n/n. X/(1 — X). The right-hand side has a factor of ne, and by charge conservation ne = 1); = XM for hydrogen. Collecting all factors of X on the left-hand side then yields an oy -2 ) epuke (45) 2ah? Mo Now we need to deal with 4 on the right-hand side. NASA’s Sun fact sheet gives the pressure and temperature at the bottom and top of the photosphere (see Table 14.1). We can compute a number density from the ideal gas law, n = P/kT (column 3). This includes contributions from all three constituents (neutral atoms, 14.1 Atomic Excitation and Ionization 21 ‘Table 14.1 Physical conditions at the bottom and top of the Sun’s photosphere, from NASA's Sun fact sheet. Page) pgm Bottom 1.25 x 108 6,600 1.37 x 10 2.29x 10-+ Top 8.68 x 10! 4,400 1.43 x 10 2.39 10-* ions, and electrons), so it depends on the ionization fraction. The mass density, by contrast, is much less sensitive to X (because electrons contribute so little mass). As we will see shortly, the ionization fraction in the Sun’s photosphere is quite low so to a good approximation we can take the mean particle mass to be the mass of a neutral hydrogen atom, 7 © mp, and then compute the mass density as p = #P/kT (column 4). Then we can use p = mp (again, the electrons are negligible in mass) to rewrite Eq. (14.5) as X?_ mp (mkT\? ar 1-x me (557) me (169) This is a quadratic equation that we can solve for X. Plugging in numbers for the bottom and top of the photosphere yields y = 462% 10 bottom 1.1.x 1075 top Very few hydrogen atoms in the outer layer of the Sun are ionized. ‘That would change if the temperature increased, of course. We can use Eq. (14.6) to estimate how the ionization fraction would increase if we assume the mass density remains fixed as we vary the temperature. This is not quite correct because the star would adjust its hydrostatic equilibrium for a different temperature, but it lets us obtain a useful estimate without getting too bogged down in details. Plugging numbers into Eq. (14.6) lets us write the temperature dependence as KT V3? 220x 10 (5) eB6eV/kT bottom an aarx 0 (7) TBST top Solving for X yields the curves shown in Fig. 14.4. At the top of the photosphere, the temperature would need to reach 10*K or more for the ionization fraction to become substantial. At the bottom of the photosphere, an even higher temperature would be required because the higher density makes it easier for electrons and ions to recombine into neutral atoms, which reduces the equilibrium ionization fraction at a given temperature. 292 14 Stellar Atmospheres £6 & . H ionization fraction, X § ° 10000 15000 temperature T (K) Fig. 144 Ionization fraction X for hydrogen, as a function of temperature, assuming the pressure at the top (solid) or bottom (dashed) of the Sun's photosphere ‘We can combine the Boltzmann and Saha analyses to determine the fraction of hydrogen atoms that are neutral and in the first excited state. If we assume that very few neutral atoms are excited beyond the first excited state, we can use ny © m1 +2 to write m _ mn n n m/n mmm 2 ( ) 2/ny = yA eR 1+m/m a-Xx) Me MY Mee MH Mot The first factor involves n2/n, from Fig. 14.3 while the second factor has X from Fig. 14.4. Putting them together yields Fig. 14.5, where we focus on the top of the photosphere because more absorption lines are produced there than at the bottom. For the assumed mass density, the excitation fraction n2/ Mx peaks for a photospheric temperature of around 12,000K. We need to be a little careful when interpreting this result because the mass density would not necessarily remain fixed as the temperature varies, and the density we have assumed for the Sun would not necessarily apply to stars with different masses. Nevertheless, our simplified analysis suggests that hydrogen Balmer lines will be most prominent in stars with photospheric temperatures in the ballpark of 10,000-15.000 K, which is in fact what we see. And we understand why: at cooler temperatures, too little hydrogen is excited into the state that can produce Balmer absorption lines; while at hotter temperatures, too much of the hydrogen is ionized. 142 Stellar Spectral Classification 293 np/Myoy for hydrogen ¥ 4 T o 1 1 4 10000 15000 temperature T (K) Fig. 14.5. Expected fraction of hydrogen atoms that are neutral and in the first excited state, as a function of temperature. We use the mass density at the top of the Sun's photosphere 14.2 Stellar Spectral Classification Now we understand that the appearance of spectral lines is governed not only by the composition of stellar atmospheres, which is to be expected, but also by the temperature, which is less obvious but no less important. Using the physics of atomic excitation and ionization, we can comprehend the key patterns observed in stellar spectra In the 1890s, astronomers at Harvard College Observatory amassed a large collection of stellar spectra. The Observatory employed many women (who were known at the time as “computers”) to analyze astronomical data. Williamina Fleming developed a taxonomy in which spectra were classified based on the strength of their hydrogen absorption lines. A stars had the strongest lines, B stars the next strongest, and so forth. Annie Jump Cannon then consolidated the spectral classes and discovered that the order “O B A F G K M” corresponds to a sequence in temperature, running from hot to cool.’ Figures 14.1 and 14.2 show that the peak of the spectrum shifts with temperature (Wien’s law) and the set of absorption lines varies as listed in Table 14.2. One additional pattern in star properties was discovered in the early 1900s by Ejnar Hertzsprung and Henry Norris Russell. Working independently, they compared the spectral classes and luminosities of different stars in a plot now called the Hertzsprung-Russell (HR) diagram (see Fig. 14.6). Stars do not scatter across 3People have invented a variety of mnemonics to remember the sequence. What's yours? ‘Recently the spectral sequence has been extended to include types L, T, and ¥ for low-mass stars known as brown dwarfs whose cores are not hot enough for normal hydrogen fusion to occur (see Problem 16.5).

You might also like