Stellar Objects: Equation of State 1 Equation of State

Our Goal here is to derive the equation of state (or the mutual dependencies among local thermodynamic quantities such as P,T,ρ, and Ni), not only for the classic ideal , but also for and fermions, for which quantum effects are important. The EoS together the thermodynamic equation allows to study how the stellar material react to the heat, changing density, etc.

1 The local thermodynamic equilibrium

We do this under the local thermodynamic equilibrium (LTE) assump- tion. The LTE means that at any position in a star, complete thermodynamic equilibrium is very nearly true: e.g., the Boltzmann populations of ion levels are consistent with the local kinetic of electrons.

This assumption is appropriate for a star, since the mean free path of particles is short generally short, compare to the variation scales of the thermodynamic −1 quantities; e.g., λph = (κρ) is on order of 1 cm. Exception cases are the stellar photosphere.

With such a short mean free path, how long does it take for a to diffuse out of a star?

Random walk: 1-D random walk theory.

2 2 The mean of N steps is zero, but the dispersion is not, σ ( i xi)= Nσ . The time scale to establish the LTE is also typically short enough,P except for nuclear reactions, compared to the evolutionary time scale of a star with the exception of various explosions.

2 Chemical reactions

Chemical reactions determine the species number density Ni may have the units of number per gram of material; i.e., Ni = ni/ρ. Stellar Objects: Equation of State 2

The thermodynamic equation with chemical reactions:

dE = TdS − P dV + µidNi (1) i X where µi is the associated with an ith species and clearly can be defined as ∂E µ = . (2) i ∂N  i S,V The LTE requires that

µidNi =0 (3) i X As an example,

+ − 0 H + e ↔ H + χH (4) where χH = 13.6 eV is the ionization potential from the ground state of H. For simplicity, we’ll assume that H has only one bound level and no other reactions that involve the above constituents and that the gas is pure hydro- gen.

We will neglect the photons here since its chemical potential is zero. Consider a classic cavity filled with radiation in thermodynamic equilibrium with the wall. The above equation can be written as

µidNi + µγdNγ =0 (5) i X Since photon number is not strictly conserved, in general dNγ =6 0. While dNi = 0 for the cavity wall (as a whole), we have µγ = 0. Now let us get back to ionization-recombination reaction, which can be writ- ten as 1H+ + 1e− − 1H0 =0 (6) And in general, a reaction can be written in such a symbolic form:

νiCi = 0, (7) i X Stellar Objects: Equation of State 3

where Ci is the names of the participating particle (except for photons). Clearly, dN = νi dN . Since dN is arbitrary, we have i ν1 1 1

µνi =0 (8) i X This the equation of chemical equilibrium. Here µi, containing the informa- tion of Ni, depending on T and ρ or equivalent thermodynamic quantities as well as on a catalog of the possible reactions. When they are known, we should be able to find all Ni. For a perfect (or ideal) gas, as will be assumed here, it is not difficult to establish such connection.

3 The Distribution Function

For a particular species of elementary nature (ions, photons, etc.) in the LTE, the occupation number in a certain quantum state, defined by the level j in the coordinate-momentum phase space is found from statistical mechanics to be 1 F (j,E(p)) = , (9) exp[−µ + Ej + E(p)]/kT ± 1 where Ej is the energy state j referred to the reference energy level, E(p) is the kinetic energy as a function of the momentum p, E(p)=(p2c2 + m2c4)1/2 − mc2. A “+” or “-” is used for Fermi-Dirac or Bose-Einstein parti- cles (fermions of half-integer or of zero and whole integer spin), respectively.

The number of states in the phase space, accounting for the degeneracy of the state gj, is g dpdr j . (10) h3

So the phase-space density can be defined as 1 g n(p)= j . (11) h3 exp[−µ + E + E(p)]/kT ± 1 j j X Stellar Objects: Equation of State 4

We can then calculate the physical space density n, kinetic P , and internal energy E as n = n(p)4πp2dp, (12) Zp This gives the connection between n, T , and µ.

1 P = n(p)vp4πp2dp, (13) 3 Zp ∂E(p) where v = ∂p . E = n(p)E(p)4πp2dp (14) Zp

4 Black body Radiation

Photons are mass-less bosons of unit spin. Since they travel at c, they have only two states (two spin orientations or polarization) for a given energy. Hence, 8π ∞ p2dp n = ∝ T 3. (15) γ h3 exp(pc/kT − 1) Z0 aT 4 4 Similarly, one can easily get Prad = 3 and Erad = aT = 3Prad, where a is the radiation constant.

The nice thing about the LTE is that all these quantities depend only on T .

Note that the above equation is a γ-law EoS P =(γ − 1)E with γ = 4/3.

For later use, it is easy to define the energy density per unit frequency (ν) as

8πhν3 1 u dν = erg cm−3Hz−1Hz. (16) ν c3 exp(hν/kT − 1) and the frequency-dependent Planck function c B (T )= u , (17) ν 4π ν Stellar Objects: Equation of State 5 which peaks at hν = 2.8kT .

The integrated Planck function is σ B(t)= B (T )dν = T 4 erg cm−2 s−1 sr−1. (18) ν π Z where σ is the Stefan-Boltzmann constant.

Reading assignment: CH 3.5-.6

5 Fermi-Dirac EoS

Elections, protons, and neutrons are fermions. They all have spin one half. 2 Hence g = 2. We choose an energy reference level of E0 = mc (no internal energy level; alternative choices would only lead to an additive constant in the definition of µ). The occupation number is 1 F (E)= , (19) exp ([E(p) − (µ − mc2)]/kT ) ± 1 −1/2 2 p 2 1/2 p p 2 where, in general, E(p)= mc (1 + mc ) − 1 and v(p)= m 1+ mc . h  i h  i 5.1 The Complete Degenerate Gas

Now let us examine the behavior of the above integral when T → 0. The exponential tends either to zero or infinity, depending on whether the kinetic energy E(p) > (µ−mc2) or E(p) < (µ−mc2). We call this critical kinetic en- 2 ergy, EF =(µ−mc ), as the “Fermi energy”, though we are yet to determine µ (= the total energy (including the rest mass energy) of the most energetic particle of the system). Fermions fill the energy range, 0 ≤ E ≤ EF — “filled Fermi sea”, which is capped by EF . Correspondingly, we have pF , as 2 2 1/2 pF defined by EF = mc (1 + x ) − 1 , where x = mc is the dimensionless momentum.    The number density can be calculated as − 8π pF 8π h 3 n = p2dp = x3, (20) h3 3 mc Z0   Stellar Objects: Equation of State 6

h −10 where mc is the Compton wavelength and is 2.4 × 10 cm for electrons and about 2000 smaller for neutrons (due to the mass difference). For the  complete degenerate electron gas, we then have ρ = Bx3, (21) µe −3 where B = 8πma h = 9.7 × 105 g cm−3. Therefore, once ρ is known, 3 mc µe we can derive x, and hence p and E .  F F The demarcation between relativistic and relativistic mechanics occur when 6 −3 pF ∼ mec or x ∼ 1. The corresponding density is ∼ 10 g cm , which, incidentally, is a typical central density of white dwarfs.

For typical densities in a neutron star (comparable to the nuclear densities of ρ ≈ 2.7 × 1014 g cm−3, it is easy to show x ≈ 0.35. Thus the neutron stars are non-relativistic.

The Pressure and internal energy density can be calculated as 8π m4c5 x x4 P = dx, (22) 3 h3 (1 + x2)1/2 Z0 and − h 3 x E = 8π mc2 x2[(1 + x2)1/2 − 1]dx, (23) mc   Z0 which have limiting forms of P ∝ E ∝ x5 ∝ ρ5/3 for x ≪ 1 and P ∝ E ∝ x4 ∝ ρ4/3 for x ≫ 1.

It is also easy to show E/P = 3/2(γ = 5/3) for x ≪ 1 and E/P = 3(γ = 4/3) for x ≫ 1.

For a γ-law EoS, the completely degenerate gas in a non-relativistic limit acts like a monatomic whereas, in the extreme relativistic limit, it behaves like a photon gas.

5.2 Application to White Dwarfs

The above P (x(ρ)) relation together with the mass conservation equation can be combined to for a second-order different equation (e.g., ρ vs. Mr). Its solution would give the structure of a WD. Stellar Objects: Equation of State 7

But here, we will only show the M − R scaling via the virial theorem in the hydrostatic form 3(γ − 1)U = −Ω (remember 2K = 3P ), assuming a white dwarf with constant density and composition throughout. Of course, the constant density (hence constant pressure) assumption does not satisfy the hydrostatic equation. But this does not change the above scaling, quali- tatively.

Because Ω ∼ GM 2/R and U = VE ∝ R3(M/R3)5/3 (in the non-relativistic limit), M ∝ R−3. This scaling shows the remarkable property that as mass increases, the radius decreases and is quit unlike the homology result for MS stars.

Similar dimensional analysis gives M ∝ R0 in the extreme relativistic limit. An exact analysis (which will be done later) yields

2 M M∞ 2 = = 1.456 , (24) M⊙ M⊙ µ  e  where M∞ is the Chandrasekhar limiting mass. A full virial analysis shows 1 d2I that 2 dt2 (= 3(γ − 1)U + Ω) becomes negative if the star mass exceeds M∞. The interpretation is that an electron degenerate object (of fixed µe) cannot have the mass exceeding the Chandrasekhar limit. The EoS is too soft! Increased density and pressure cannot halt the collapse because the relativistic limit (related to the limit of the speed of light) has already been reached. So a star needs either to start with a low mass or to rid itself of enough mass to eventually leave a WD with M < M∞.

A similar limit (∼ 3M⊙) can be found for a neutron star mass.

5.3 Temperature Effects

Now let us see what we really mean by stating T → 0. Suppose the tempera- ture is low — on some scale yet to be determined — but not zero. Fermions deep in the Fermi sea, at much less than EF , need roughly an addi- tional EF energy units to move around in energy. If this energy input, as mea- sured by kT , is much smaller than EF , the low-energy particles are excluded from promotion to already occupied upper energy levels. But Fermions near the surface of the sea, within a depth of ∼ kT , may find themselves elevated Stellar Objects: Equation of State 8

into energies greater an EF . The net effect is that the occupation distribution smooths out to higher energies. Apparently, the criterion for the transition between degeneracy and near- or degeneracy is EF ∼ kT . (The figure in the textbook shows the distribution for EF ∼ 20kT .)

The criterion gives a ρ/µe−T relation that roughly separates the two regimes, 2 2 as shown in Fig. 3.7. For a non-relativistic electron gas (EF = mec x /2), ρ = 6.0 × 10−9T 3/2 g cm−3. (25) µe

2 For a relativistic electron gas (EF = mec x),

ρ = 4.6 × 10−24T 3 g cm−3. (26) µe

The center of the present sun is non-degenerate, but close enough to the transition line that good solar model needs to include the effects of Fermi- Dirac statistics.

6 “Almost Perfect” EoS

The quantum mechanics effect on fermions or bosons affect the occupation distribution in the phase (or energy) space, but does not alter the energy of individual particles. In real gas, interactions need to be accounted for that modify the “perfect” results as discussed so far.

In addition, a stellar EoS might consist of many components with radiation, Maxwell-Boltzmann, and degenerate competing in importance. Fig. 3.9 summarizes some of the effects for a hypothetical gas composed of pure hydrogen.

A measure of the interaction energy between ions is the Coulomb potential. The ratio Z2e2 Γ ≡ (27) C akT

Read 3.7 and no class next Monday. Stellar Objects: Equation of State 9

7 Ideal Gas

The Boltzmann distribution for an ideal gas is characterized by µ/kT ≪ −1, i.e., the occupation number is very low.

8 The Saha Equation

Often the number densities of some species cannot be set a priori. This will affect the mean molecular weight, as discussed earlier, and other parameters (e.g., various derivatives), as will be discussed later. If the system is in thermodynamic equilibrium, however, then the chemical potentials of the reacting constituents depend on one another and this additional constraint is sufficient to determine the number densites.

Again consider the ionization-recombination reaction:

+ − 0 H + e ↔ H + χH (28) and assume no other reactions are taking place that involve the above con- stituents and that the gas is ideal and purely hydrogen with only one state. We establish the reference energy levels for all species by taking the zero + − energy as the just-ionized to be the H + e state. Thus E0 is zero for both electrons and protons, whereas = −χH for neutral hydrogen. To avoid the double counting, we may choose g− = 2 and g+ = 1 (or vise verse) as well as g0 = 2. Then it is easy to show 3/2 2[2πm kT ] − n = e eµ /kT (29) e h3 3/2 [2πm kT ] + n+ = p eµ /kT (30) h3 3/2 2[2π(me + mp)kT ] 0 n0 = eµ /kT eχH /kT (31) h3 − + 0 Notice that for the LTE, µ + µ + µ = 0 and [memp/(me + mp)] ≈ me, we have the Saha equation:

+ 3/2 n n 2πm kT H e = e e−χ /kT . (32) n0 h2   Stellar Objects: Equation of State 10

+ The electrical neutrality gives ne = n . Using y to parametrize the fraction n+ of all hydrogen that is ionized, i.e., y = n , the equation then becomes

2 3/2 y 1 2πm kT H = e e−χ /kT . (33) (1 − y) n h2   The dominant factor here is the exponential, while the dependence on the density is weak. The characteristic temperature for ionization-recombination is around 104 K , or roughly y ≈ 1/2 when χ/kT ∼ 10.

Neglecting any radiation contribution, the thermal pressure can be expressed as + 0 P =(ne + n + n )kT = (1+ y)NρkT, (34) where N is the total ion plus neutral atom number density per unit mass and is a constant, independent of density.

The specific internal energy is 3kT E = (1+ y)N + yNχ erg g−1, (35) 2 H in which the energetics of the reaction are accounted for and are needed to fully ionize the neutral gas. With the Saha equation, it is clear that P and E involve only the temperature and density.

For real calculation, of course, all species, energy levels, and reactions must be considered. For example, one also has the ionization-recombination for Helium at high (24.6 eV and 54.4 eV for the first and second electrons).

The presence of such zones of ionization has profound consequences for the structure of a star.

Later we will further use similar Saha equations for some fast nuclear re- actions under certain states where the thermodynamic equilibrium can be assumed. Stellar Objects: Equation of State 11

9 Adiabatic Exponents and Other Derivatives

Having the pressure and internal energy, we can in principle construct a simplified stellar model. But to construct realistic ones and to evolve them, we need several thermodynamic derivatives. We need them, for example, to study the stability of a model and to determine weather or not the convection needs to be considered.

Such derivatives include the specific heats (cρ and cp), power-law exponents ∂lnP ∂lnP of the EoS (χT = ∂lnT and χρ = ∂lnρ ), and adiabatic exponents ρ T ∂lnP   (Γ1 = ∂lnρ , Γ2, and Γ3) ad   9.1 Specific heat allowing for chemical reaction

∂Q ∂E As an example, we consider cρ = ∂T ρ = ∂T ρ for the above pure hydrogen gas. ∂E 3kT dT  2 3  χ 1 ∂y = (1+ y)N + + H . (36) ∂T 2 T 3 2 kT (1 + y) ∂T     Take a log of the Saha equation and then take the partial derivative of T , 2 1 ∂y 3 χH 1 + = + . (37) y 1 − y ∂T 2 kT T    

y(1−y) Define D(y)= (2−y)(1+y) ,

1 ∂y 3 χH 1 = D(y) + . (38) (1 + y) ∂T 2 kT T   Therefore,

3k 2 3 χ 2 c = (1+ y)N 1+ D(y) + H erg g−1 K−1. (39) ρ 2 3 2 kT "   # where the second term arises from the consideration of the ionization-recombination reaction. Note that D(1) = D(0) = 0 and, for general 0 ≤ y ≤ 1, D(y) ≥ 0. Stellar Objects: Equation of State 12

Thus the term is greater than zero, meaning that an increase of the tem- perature, keeping the density fixed, requires more energy when the reaction occurs than when it does not.

9.2 Maxwell’s relations

It is often convenient to express the derivatives in certain forms and to know their relationships. The Maxwell’s relations are often useful:

∂T ∂P = − (40) ∂V ∂s  s  V ∂T ∂V = (41) ∂P ∂s  s  P ∂s ∂P = (42) ∂V ∂T  T  V ∂s ∂V = − (43) ∂P ∂T  T  P These relations can be derived from the first law of thermodynamics, ex- pressed for E, H = E + P V (enthapy), F = E − Ts (free energy), or Φ= E +P V −Ts (the ), using s and V , s and P , T and V , or T and P as independent variables, respectively, using such mathematical relations as ∂2Φ ∂2Φ = (44) ∂P∂T ∂T∂P

9.3 Other relations

Such partial derivatives can be expressed in different forms, using the Jaco- bian theory, e.g., ∂(V,P ) ∂V ∂P = ∂(V,T ) = ∂T V (45) ∂T ∂(T,P ) − ∂P  P ∂(V,T ) ∂V T  Stellar Objects: Equation of State 13

This relation can also be derived from ∂V ∂V dV = dT + dP (46) ∂T ∂P  P  T by holding V constant, dV = 0.

The relation will be used shortly.

9.4 Application example

Consider the relation between cP and cρ or cVρ . Start by letting T and P be independent and write the specific heat content as

∂s ∂s dq = TdS = T dT + dP . (47) ∂T ∂P  P  T  and ∂P ∂P dP = dT + dV. (48) ∂T ∂V  V  T This gives

∂s ∂s ∂P ∂P dq = T dT + T dT + dV . (49) ∂T ∂P ∂T ∂V  P  T  V  T 

While holding V constant,

∂q ∂s ∂s ∂P = T dT + T (50) ∂T ∂T ∂P ∂T  V  P  T  V

∂s ∂P c − c = T . (51) p V ∂P ∂T  T  V Using one of the Maxwell’s relations

∂s ∂V = − (52) ∂P ∂T  T  P Stellar Objects: Equation of State 14 and then the relation we have just derived:

∂V ∂P = ∂T V (53) ∂T − ∂P  P ∂V T  we have − ∂P 1 ∂P 2 c − c = −T (54) p V ∂V ∂T  T  V

10 Review

Key Concepts: local thermodynamic equilibrium, random walk, chemical po- tential µ, chemical equilibrium, Fermi energy, Chandrasekhar limiting mass

What is an equation of state?

Why should the chemical potential of photon gas be equal to zero?

How to derive such quantities as density, pressure, and thermal energy for photon, ideal (Maxwell-Boltzmann), and complete degenerate gases, starting from the basic occupation distribution (with different µ limits)?

What is the degenerate pressure?

What is the Saha equation? Under which circumstance can the equation be used?

Get familiar with derivations of various adiabatic exponents and other deriva- tives.