Dense Hydrogen Thing

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

197 5ApJ. . .199. .

255G

The Astrophysical Journal, 199:255-264, 1975 July 1


© 1975. The American Astronomical Society. All rights reserved. Printed in U.S.A.

THERMODYNAMICS OF DENSE HYDROGEN-HELIUM FLUIDS*


Harold C. Graboske, Jr., and Robert J. Olness
Lawrence Livermore Laboratory, University of California, Livermore
AND
Allan S. Grossman
Iowa State University, Ames
Received 1973 July 13
ABSTRACT
The equilibrium thermodynamic properties of hydrogen-helium fluids are studied in the
density-temperature regime appropriate to the interiors of evolving very low mass stars and giant
planets. The low-density molecular fluid is treated by a free energy minimization method, utilizing
several theoretical models and recent high-pressure experimental results for hydrogen. The
high-density metallic fluid is treated by an extension of Thomas-Fermi theory. These approaches
are combined to yield a detailed thermodynamic description for two fluids: pure hydrogen
(X = 1.) and a solar mixture (X = 0.74, Y = 0.24, Z = 0.02). Equations of state are given for
the density range 10_4gcm“3to lOgcm-3 and a temperature range 200-60,000 K.
Subject headings: equation of state — interiors, planetary
I. INTRODUCTION
The equilibrium thermodynamic properties for fluid hydrogen and hydrogen-helium mixtures are extremely
difficult to calculate in regions of interest for the study of giant planet interiors. A detailed and complete survey
of this topic has been carried out by Hubbard and Smoluchowski (1973, hereafter referred to as HS). The diffi-
culties encountered in studies of the physics for static models of present epoch Jupiter are increased in an
evolutionary study, as a much wider range of density p and temperature T is required.
Two astrophysical materials will be considered here : pure hydrogen, and a solar mixture which is 90 percent
hydrogen by number. Because of its numerical dominance, hydrogen is the crucial element in the thermodynamics
of the giant planets. Fortunately, for hydrogen a significant number of experimental data are now becoming
available, and the theoretical treatment has fewer difficulties than any fluid mixture. A mixture with chemical
composition similar to that of the Sun is a priori the most probable material for the giant planets. The elemental
composition studied here consists of hydrogen, helium, and the solar distribution of metals as given by Allen
(1963), with abundances by mass for hydrogen, helium, and metals given by Y = 0.74, Y = 0.24, Z = 0.02.
The range of density and temperature required to encompass evolving giant planet models can be estimated
from the behavior of low-mass stars in their pre-main-sequence stages, together with the static-fluid planetary
models of Hubbard. This gives a temperature range of 60 K (hydrogen, hereafter H) or 100 K (solar mixture,
hereafter SI) to 25,000 K, an upper limit later extended to 60,000 K, and a density range from 10-7 g cm-3 to
10 gem-3. The thermodynamic properties for H and SI were calculated for a skewed grid within these limits.
The phase diagram for hydrogen is not clearly delineated even in a qualitative manner. Recently, the actual
existence of the solid metallic phase, long considered the most likely condition of the Jovian interior, has been
questioned (HS). With one exception, we shall consider only fluid phases. The discussion below separates naturally
into two parts: a low-density regime, p ^ 10_1gcm_3 which has a low-temperature region (60-4000 K) containing
the molecular fluid statè of H2, followed by a dissociation-ionization zone, succeeded in turn by the high-tempera-
ture Coulomb plasma; the high-density regime, p ^ 1 gem“3, which has a low-temperature region (~2000-
20,000 K) containing the metallic hydrogen fluid followed by the high-temperature, strongly coupled Coulomb
plasma. As discussed by Hubbard and Smoluchowski, at the lower temperatures there will exist a fluid-fluid phase
transition between the molecular and metallic phases, while at the higher temperatures (above some presently
unknown critical temperature) a smooth change occurs from a weakly coupled to a strongly coupled Coulomb gas.
At lower densities, a molecular solid will form as temperature is decreased below its (unknown) melting tem-
perature; at higher densities, a metallic solid may form similarly. Hence three more phase boundaries—molecular
fluid-solid, metallic fluid-solid, and molecular solid-metallic solid-will exist. The existence of these phase bound-
aries is very uncertain at present, although considerable study has been devoted to the problem (HS). It is possible
that the lowest temperature regions of our fluid systems really lie in regions where the material has solidified.
However, this possibility can be dealt with in subsequent studies because our planetary model sequences do not
* Work performed under the auspices of the U.S. Atomic Energy Commission.
255

© American Astronomical Society • Provided by the NASA Astrophysics Data System


197 5ApJ. . .199. .255G

256 GRABOSKE, OLNESS, AND GROSSMAN Vol. 199


approach the low temperatures at which hydrogen freezes. Hubbard and Smoluchowski note that the physics of
the metallic fluid is well approximated by several theoretical approaches. However, their conclusion concerning
the considerable uncertainties in the physics of the molecular fluid is no longer valid. Due to new theoretical
techniques and to new accurate experimental data on hydrogen in the 0.1-0.9 megabar (Mb) region, the thermo-
dynamics of the molecular fluid are at least as well known as the metallic fluid.
The thermodynamics of the fluid solar mixture are treated in the same manner as the pure hydrogen system.
Although the accuracy of the theory of hydrogen will give a good representation for the 90 percent of the mixture
that is hydrogenic, the interactions of the other 10 percent, especially in the most dense part of the molecular
fluid regime, are not as well known. In addition, the problems of solidification in the mixture are complicated by
the possibility of separation by diffusion (Smoluchowski 1967) or immiscibility (Salpeter 1973) of the primary
constituents, hydrogen and helium. The theory of such phase conditions is considerably more difficult than the
case of pure hydrogen, which, as described above, is still theoretically incomplete.

II. HYDROGEN LOW-DENSITY REGION


The low-density region for hydrogen can be considered in three parts: at low temperatures a molecular or
atomic fluid regime, at intermediate temperatures a dissociation-ionization regime, and at highest temperature a
Coulomb gas or plasma. All these regions involve many-body interactions, which are best studied by perturbation
techniques. Since even the pure hydrogen system has six separate components (H2, H2 + , H“, H, H + , e-1) while
the SI solar mix adds the atoms and ions of He, C, N, O, and Ne, the technique chosen is an equilibrium technique
involving minimization of a thermodynamic state function, the Helmholtz free energy F({Ni}, V, T). All approxi-
mations are made in the construction of the free energy function (or the equivalent canonical partition function)
of the system involving translational, configurational, and internal contributions. Once the free energy is deter-
mined, the subsequent analysis is thermodynamically consistent. The ideal free energies of Fermi-Dirac electrons,
classical atoms, ions, and nuclei, and photons represent the unperturbed basis, and to this are added terms
representing the interactions.
In the low-density region, interactions of three basic types can be identified, corresponding to the three regions
described above. In the ionization-dissociation regime, pair interactions between ions and electrons or between
two atoms lead to the formation of atoms and molecules. These interactions are represented by the internal
partition function (Rogers, Graboske, and DeWitt 1971) including all atomic, ionic, and molecular species:

-internal = 2 (2/ + CX
P ~ ^ +
Âf j0 PdPhÂP)™V '

The summation over all bound (negative energy) states contains the energy eigenvalues for all levels i of all
species y, while the integral over the scattering (positive energy) states includes the phase shifts §n for all states /
of all species j. This free energy term, with scattering contributions neglected and the eigenvalues taken from the
isolated atom (IA) or unperturbed Schrôdinger equation, when minimized in combination with the ideal free
energies, will yield a Saha ionization-dissociation equation set. For molecular species, vibration and rotation
contributions to the internal free energy must be included.
At high temperatures and densities where a nonnegligible fraction of the atoms are ionized, many-body Coulomb
interactions are encountered. The Coulomb free energy used here rests on the use of the first-order perturbation
term in the weak-interaction limit combined with a Monte Carlo numerical result for the strong-interaction limit.
The weak-interaction free energy is the generalized (multicomponent) ring term with Fermi-Dirac statistics in-
cluded and a short-range cutoff to approximate the effect of the higher-order direct terms. The strongly coupled
plasma free energy is taken from the numerical results of Brush, Sahlin, and Teller (1966, hereafter BST) which
are valid for a single-charge classicial ion plasma immersed in a uniform electron background. The numerical
results for this strong coupling theory go over smoothly to the corresponding ring (Debye) limit, as represented
by the interpolation formula:

^Coulomb 2>(0-2887r3,2)[(i + O.iossr)1'2 +


(1 + 0.3566r)3'2] ’

23
r = A '
3173 where A= Ap-2 = Anße2 2 Zi2-/V¡-/-1'2^
47777 ÀF3

where Jn(ae) is the Fermi-Dirac function of index n, ÀF is the screening length, ae = jS/xe, ß = l/kT, and /ze is the
chemical potential per electron. This combination is exact in the limit (A « 1, cce < 0) and is accurate for the
range (A > 0.1, ae ^ 10). It is least valid where the electrons are moderately degenerate; under such circumstances
the accuracy of the model is undetermined as there is no theory to compare with. An exchange interaction term is

© American Astronomical Society • Provided by the NASA Astrophysics Data System


197 5ApJ. . .199. .255G

No. 1,1975 THERMODYNAMICS OF DENSE H-He FLUIDS 257


also necessary. It is represented by an exact analytical expression for the first-order electron exchange, which is
only important for ae » + 1 :
NeYl ( K \/X^-1/2V)
~ßFlx = Ye2 = KlUZ/NJ^ZßNd112
2 [ZZ^nJ -A/22(«e)
The third effect at low density results from the neutral interatomic and intermolecular interactions. These
interactions are characterized by binary encounters exhibiting strong repulsion at short range and weak attraction
at long range. Like the Coulomb effects, they become small at low densities or at high temperatures, and are most
important in the high-density, low-temperature limit. Low temperature here means well below the effective dis-
sociation or ionization temperature. In contrast to the unresolved difficulties of the Coulomb fluid, several theo-
retical techniques have been developed recently which yield excellent results for neutral fluids. These include
fluid perturbation theory, a fluid variational method, and empirical techniques using analytical models in combina-
tion with experimental data. The latter method was chosen for the case of the Jovian molecular fluid.
The theoretical model of the equilibrium fluid is based on an analytical study of fluids and fluid mixtures called
scaled particle theory (Lebowitz, Helfand, and Praestgaard 1965). This model has been shown to be formally
equivalent to the equation of state obtained by using the compressibility relation in combination with the direct
correlation function resulting from an exact solution of the Percus-Yevick integral equation for a hard sphere
mixture (Lebowitz 1964). Furthermore, the scaled particle theory gives good agreement with numerical results of
Monte Carlo and molecular dynamics studies of hard-sphere fluids, both simple and binary species. The free
energy for a fluid mixture of particles of differing radii i?* undergoing repulsive interactions is given by
18^2 9f22 1
=^l- 6£0 ln (1 - £3) + (1 - f3) +
(1 - Í3)2J ’ ^ = jfZx¿2RJn-
The free energy corresponding to the attractive interactions is given by a van der Waals form:

^aur =
i 3
The theory to this point gives a model of a fluid with a very stiff repulsive potential (hard sphere) plus a soft
attractive term, of the order of r-6. This was used in Grossman et al. (1972) to investigate the feasibility of a stellar
structure study of Jupiter and to delineate the p-T regions traversed by the planetary models. It proved inadequate,
and is here replaced by a much improved form. The repulsive interaction for real molecules is not given by a hard-
sphere potential but is best represented by a function with exponential or inverse twelfth power radial dependence.
Such potentials, exponential-six (e-6), e-6-8, or 12-6, can be established empirically by requiring them to reproduce
experimental data. The difficulty here is that the only data available for H and H-He fluids (or indeed any other
fluids) are in regions that are orders of magnitude lower in pressure than the Jovian interior. The dominant inter-
action in the low-temperature hydrogen fluid is that for H2-H2; a potential for this interaction derived from
recent high pressure experimental studies in the 0.1-0.9 Mb range (van Thiel and Alder 1966; Dick 1971; van
Thiel and Hord 1973) allows an accurate theory to be produced up to densities near the molecular-metallic phase
transition.
The modification of scaled particle theory (SPT), which permits treatment of real gas interactions, involves the
replacement of a fixed hard-sphere radius by a temperature-dependent hard-sphere radius, also called a soft-
sphere model. This is done by determining the effective intermolecular or interatomic potential, and then deter-
mining an effective hard-sphere radius by equating it to one-half the mean distance of closest approach in a binary
encounter. This mean distance is obtained by finding the classical turning point, where the effective potential
F(2rf) equals the kinetic energy which is taken to be kT. The resulting temperature-dependent radii are then used
in the repulsive free energy term, producing a system that has smaller effective radii at higher temperatures. If the
procedure is carried out properly, it is self-consistent in that the potential derived from the experimental data is
used to compute effective radii which, in turn, are inserted in the theoretical model which should predict thermo-
dynamic properties that agree with the experimental data.
The definition of the effective potential for static data points is direct since P, F, and T are all known. For shock
data, P, F, and E are known, so that T must be found by theoretical methods. (In this context, it should be noted
that magnetic compression experiments [such as those quoted by Slattery and Hubbard 1973] yield direct measures
of F only, with S assumed. This incomplete thermodynamic description and the present lack of precision make
this difficult type of experiment unusable for comparison with theory.] Instead of determining temperatures by a
lengthy iterative technique, we adopted the temperatures assigned by Ross (1974), whose study of the dense
hydrogen molecular fluid represents the most complete and careful analysis presently available. The resulting
theoretical predictions of the SPT model gave satisfactory agreement with these initial temperature choices.
The effective potential is determined by equating the pressure obtained for each experimental point P*(Fi? ZJ)
to the pressure from the theoretical model P^F*, P*, r*). The pressure component arising from the repulsive inter-
action is then adjusted to give agreement with the experimental value by varying the effective hard-sphere radius.

© American Astronomical Society • Provided by the NASA Astrophysics Data System


197 5ApJ. . .199. .255G

258 GRABOSKE, OLNESS, AND GROSSMAN Vol 199


TABLE 1
Interatomic and Intermolecular Potentials
van der Waals
Form Ce K Cq Cq FcoreC^V) rcoret(Â) du
H2-H2 e-6-8* 203.3 3.150 6.0 10.0 1.0 1.39 0.30
H2-H e-6-8 671.8 3.713 5.007 24.69 1.4 1.36
H2-He e-6 470.1 3.832 2.878 ... 10.8 0.8
H-He e-6 71.4 3.106 1.74 ... 1.455 1.0
H-H F(eV) = 1.32e" 1-4053â;(1 + 0.1988* - 0.0795*2) , 6.4 0.33 0.0753
* = r/fl0 — 2.5 (a0 = Bohr radius)
He-He. F(eV) = 4€[(a/r)12 - (a/r)46]; o ... ... 0.0477
e = 8.806 x 10" eV = (10.22 K), o = 2.555 Â
* F(eV) = Cee~Kr - C6r~6 - C8r~e;r in angstroms.
t The parameter r corresponds throughout to a collision diameter.

The resultant set of effective radius-temperature points ^(7^) from 33 K to 7000 K is transformed, using the
relations fa = kTx and di = 2ru into a corresponding set of points which define the H2-H2 interparticle potential
curve. This curve is then approximated by an analytical function of the exponential-six-eight form.
An additional effect is involved, the negative pressure produced by the attractive part of the potential; but this
term, which is dominant in the lower supercritical region, becomes very small at the high densities and temperatures
encountered in the shock experiments. The parameter for the atractive term (van der Waals constant) was initially
chosen equal to the value found by Hoover et al. (1972), and then varied slightly to yield a final value which gave
best agreement with the low-temperature (33-300 K) experimental data.
In addition to the H2-H2 potential which produces good agreement with experimental data up to 0.9 Mb, the
H-H and H-H2 potentials must be defined to produce effective interaction radii for the corresponding particle
collisions. The H-H potential used here was derived by constructing a fit to the Hirschfelder-Linnett and to the
James-Coolidge-present calculations shown in Hirschfelder, Curtiss, and Bird (1954). The resulting potential
represents the triplet state interactions which account for three-quarters of the H-H interactions, and the resulting
radii were weighted accordingly; the singlet encounters were omitted since the singlet interaction has a very small
repulsive core, and is complicated by the existence of bound states. A potential for H2-H was constructed by adding
an orientation-averaged repulsive potential from Mason and Hirschfelder (1956) to an attractive part from
Margenau and Kestner (1971).
Table 1 gives the forms and associated parameters for the set of hydrogen interaction potentials. The forms
containing exponentials have singularities at the origin which are eliminated by using a hard-core radius. This
region is of no importance in that the real fluid particles do not approach this density while in the molecular phase.
If these radii are reached, the system will have crossed the phase boundary into the metallic region and these
potentials will be replaced by Coulombic forms. The table includes the collision diameters and corresponding
temperatures (energies) at which the hard cores are encountered. The van der Waals attractive coefficients are listed
for intraspecies interactions; for interspecies interactions, the coefficients are computed from % = (%%)1/2.
III. HYDROGEN HIGH-DENSITY REGION
There are only two valid theoretical approaches for a high-density, strongly coupled metallic fluid: the Thomas-
Fermi (TF) model, and the Monte Carlo method. The approach used in this study is based on the TF model which
is the standard high-density theory, due to its ease of use and to its exact nature in the high-density limit. A study
of zero-temperature TF theory and its extensions (Salpeter and Zapolsky 1967) discusses the background of the
approach and represents the current standard astrophysical form. Salpeter and Zapolsky (hereafter SZ) modify
the Thomas-Fermi-Dirac (TFD) theory by including the electron correlation energy in the total electron energy.
A different approach, carried out by several Soviet theoreticians (Kirzhnits 1957, 1959; Kompaneets and Pav-
lovskii 1957; Kalitkin 1960) includes an additional quantum correction term of the same order (Ä2) as the exchange
term of Dirac. The Soviet model, referred to as the TFK model, is the basis used in this work, and is obtained by
applying the variational principle to the form J With T* equal to the antisymmetrized many-electron
wave function and H given by

^ = 2i w + i i,k±i
2

one obtains the following expression for the “best” single-particle wave functions:

U(q)>Pi(q) + B(q)Hq) ~ 2 BM'l'kiq) + 2 ^Mq) = 0 . (1)


k

© American Astronomical Society • Provided by the NASA Astrophysics Data System


197 5ApJ. . .199. .255G

No. 1, 1975 THERMODYNAMICS OF DENSE H-He FLUIDS 259


The q arguments include both space and spin variables. The interactions, however, are taken as spin independent;

The aik incorporate the normalization constraint, J = 1. If we multiply equation (1) by ^*(<7) and subtract
the complex conjugate multiplied by ^i(q), the explicit reference to wave functions can be eliminated in favor of the
density matrix, leading to the following expression :

J dq"[(U^ + Biq. - Aqq~)p(q", q’) - P(q, q"){Uq.,q. + Bq„q. - Aq„q.)] = 0 (2)

with the definitions


P(q, q') s 2 MqW(q'), Uqq. ^ 8(q - q")U(q), Bqq., S 8{q - q") J p(q'\ q'”)V(q, q”)dq"',

Aqq,.= p(q,q")V(q,q")-
Equation (2) has the form of a commutator, or Poisson bracket, between the effective Hamiltonian and the
density matrix,
Hp — pH = 0 , (3)
with H = Uqq» + Bqq„ — Aqq». Bqq» represents the electron self-consistent field operator while Aqq» is the exchange
term. In the quasi-classical TF approximation, the solution of equation (3) is carried out by successive approxima-
tions, with the density matrix replaced by the Fermi-Dirac distribution function. By omitting the exchange term,
and ignoring the noncommutativity of the potential and kinetic energy operators, one obtains the zeroth-order or
Thomas-Fermi solution. The TFK model corresponds to inclusion of the exchange operator Aqq» and a proper
treatment of the commutators to order ñ2 in the resulting solution for the electron density. The corrections
correspond to a term in Aqq- (the exchange correction) and an additional term in Bqq» (the quantum correction).
This basic TFK model is augmented to include contributions from an interactive nuclear (ion) component and
an additive volume calculation for mixtures, the full treatment being contained in a computer code TFCMIX. The
validity of the additive volume treatment has been substantiated recently by Hubbard (1972) using dielectric
function theory and Monte Carlo methods for a metallic fluid mixture of hydrogen and helium. For a mixture 80
percent hydrogen and 20 percent helium (by number) he finds excellent agreement with the additive volume law
at high pressures, and good agreement for pressures as low as a few megabars. The interactive ion contribution is
similar in derivation to the Coulomb free energy term in the low-density theory. The ion contribution is a com-
posite form, starting from a weak coupling ring term (going to an ideal gas as the interaction goes to zero) smoothly
joined to the BST Monte Carlo result for the intermediate and strongly coupled Coulomb fluid, terminating in the
solid regime for which a harmonic lattice model is used. The nuclear contribution is important everywhere except
at highest densities.
The one exception to a completely fluid thermodynamic system was added for the lowest temperatures in the
high-density region. Recent calculations by Hansen (1972) using an improved Monte Carlo technique to extend
the BST results have located a fluid-solid phase transition in a classical one-component ion fluid with a uniform
electron background. This crystallization line is given by F = 143, which for metallic hydrogen ranges from 1587 K
at 1 g cm-3 to 2519 K at 4 g cm-3. The electron degeneracy at 1 g cm-3 is about +190 and increases along the
F = 143 locus. This indicates that the assumption of a uniform degenerate electron background is quite reasonable.
Using this result, we have included a solid metallic region for hydrogen, arbitrarily limiting it to /) > 1 gem-3
and T < 3000 K. The theory is that of Neece, Rogers, and Hoover (1972), a modification of the Wigner-Hunting-
ton calculation employing the Hartree-Fock self-consistent field technique along with the most recent theoretical
treatment of electron exchange and correlation contributions. This phase has been included to allow us to in-
vestigate possible solidification of the pure hydrogen model, although its inclusion is admittedly arbitrary since the
actual existence of the solid metallic phase is not certain. It has no influence on the evolutionary calculations as
the hydrogen models remain at temperatures well above those at which the fluid may be expected to crystallize
for periods of order 1010 years.

IV. COMPOSITE THERMODYNAMICS FOR HYDROGEN AND THE SI SOLAR MIXTURE


At this point a complete set of thermodynamic properties for hydrogen can be obtained by combining the two
theoretical approaches. The low-density theory is usable for p ^ 0.1 gem-3, a limit based on comparison with
experimental data and on theoretical validity limits inherent in the perturbation method. The high-density theory
is valid for ^ 10 g cm-3 but was extended to ~ 1 g cm-3, an extension which proved acceptable when compared
with a different theoretical model ( § V). At the higher temperatures (T > 20,000 K) the two theories are in reason-
able agreement, as is expected for the Coulomb fluid. At lower temperatures the intermediate zone (0.1-1.0 g cm-3)

© American Astronomical Society • Provided by the NASA Astrophysics Data System


197 5ApJ. . .199. .255G

260 GRABOSKE, OLNESS, AND GROSSMAN Vol. 199


contains the molecular-metallic phase transition. The two theories approach this transition region in very different
ways. The intermediate zone was treated by simply interpolating between the two regions, requiring smooth be-
havior of the first and second derivatives of pressure and energy throughout. Although this zone is small in extent,
the uncertainty of the thermodynamics is of sufficient importance that we include an error analysis (§ V).
The thermodynamics of the SI solar mixture were calculated using the same physical models as hydrogen. The
bound state (internal partition function), Coulomb plasma, and TF procedures are completely general and can be
extended directly to treat the He, C, N, O, and Ne atoms and ions. Two regions require different treatment. The
low-temperature, high-density solid is of course omitted, as there is no theory for this material. The molecular
fluid region contains nonnegligible amounts of helium. The small abundance by number allows neglect of the non-
ideal contribution from elements with Z > 2 in the neutral fluid, although they are nonnegligible in the Coulomb
interaction regions where their influence is felt through a Z2/^ dependence. The presence of the atoms requires

TABLE 2
Thermodynamic Properties for Dense Fluids
A. Hydrogen
Temperature (K)
-3
p(g cm ) 200 1000 6000 15,000 25,000 60,000
-2
LOoP(dyn cm )
1.0E-4 5.916 6.616 7.644 8.111 8.452 8.949
1.0E-3...... 6.918 7.618 8.542 9.093 9.368 9.857
1.0E-2 7.941 8.643 9.469 10.06 10.33 10.77
1.0E-1 9.511 9.956 10.55 11.03 11.32 11.73
2.0E-1 10.52 11.01 11.40 11.67 12.05
1.0E + 0 12.22 12.48 12.58 12.68 12.93
2.0E+0 13.03 13.10 13.16 13.21 13.37
1.0E+1 ... 14.45 14.46 14.48 14.53
log H(ergs g-1)
1.0E-4 10.45 11.16 12.46 12.79 13.21 13.57
1.0E-3 10.45 11.16 12.30 12.75 13.10 13.50
1.0E-2 10.46 11.16 12.16 12.69 13.04 13.45
1.0E-1 10.66 11.27 12.10 12.60 12.98 13.43
2.0E-1 11.51 12.22 12.60 12.97 13.39
1.0E + 0 12.50 12.81 12.94 13.05 13.32
2.0E+0 12.97 13.09 13.17 13.24 13.42
1.0E+1 ... 13.75 13.77 13.79 13.85
B. SI Solar Mixture
Temperature (K)
-3
p(g cm ) 200 1000 6000 15,000 25,000 60,000
log P (dyn cm-2)
1.0E-4 5.853 6.552 7.561 8.016 8.360 8.861
1.0E-3 6.854 7.554 8.473 9.000 9.281 9.773
1.0E-2 7.874 8.574 9.402 9.974 10.24 10.68
1.0E-1 9.292 9.822 10.47 10.95 11.23 11.64
2.0E— 1...... ... 10.48 10.89 11.29 11.56 11.94
1.0E + 0 ... 12.20 12.26 12.34 12.44 12.70
2.0E+0. ... 12.90 12.94 12.98 13.03 13.17
1.0E + 1 ... 14.31 14.32 14.33 14.35 14.39
log H (ergs g-1)
1.0E-4...... 10.38 11.08 12.36 12.68 13.11 13.48
1.0E-3 10.38 11.08 12.22 12.64 13.00 13.42
1.0E-2 10.38 11.08 12.07 12.60 12.94 13.36
1.0E— 1...... 10.50 11.16 12.01 12.51 12.88 13.34
2.0E-1 ... 11.35 12.03 12.49 12.87 13.33
1.0E+0 ... 12.42 12.59 12.74 12.95 13.34
2.0E + 0 12.86 12.92 13.00 13.11 13.39
1.0E + 1 ... 13.60 13.62 13.64 13.65 13.72

© American Astronomical Society • Provided by the NASA Astrophysics Data System


197 5ApJ. . .199. .255G

No. 1, 1975 THERMODYNAMICS OF DENSE H-He FLUIDS 261


additional effective potentials for He-He, He-H, and He-H2 interactions, which are used in same manner as the
hydrogen potentials discussed in § II. The He-He and He-H2 potentials were obtained from Margenau and Kestner
(1971). The repulsive part of the He-H potential was constructed from exponential forms of the He-He and H-H
repulsions by applying the atomic distortion model methods of Smith (1972). The attractive part was taken from
Karplus and Kolker (1964). Parameters for all six potentials are listed in Table 1.
With this addition, the thermodynamic properties for both the low- and high-density regions were calculated
as for hydrogen and combined in the same manner. The resultant pressures and enthalpies for H and SI composi-
tions are tabulated in Table 2 for six selected isotherms and a restricted density range. The low-temperature
(200 K, 1000 K) isotherms enter regions where the repulsive and attractive molecular fluid interactions are important.
The 6000 K isotherm passes through the H2 dissociation zone, while the 15,000 K and 25,000 K isotherms traverse
the H and He ionization zones, respectively.
V. COMPARISON WITH EXPERIMENT AND OTHER THEORETICAL MODELS
Figure 1 shows a plot of the pressure isotherms for the composite hydrogen equation of state in the density-
temperature region of the experimental data. The inset to the upper right shows the agreement of the low-density

Fig. 1.—Pressure-volume plot for hydrogen theoretical curves and experimental data. The point near 0.2 Mb represents a
composite of 6 separate measurements. The same is true of the point near 0.9 Mb, for which data became available after the
equation of state was established.

© American Astronomical Society • Provided by the NASA Astrophysics Data System


197 5ApJ. . .199. .255G

262 GRABOSKE, OLNESS, AND GROSSMAN Vol. 199

a-a central core track

T(eV)
Fig. 2.—Density-temperature limits (—I—) for the solar mixture equation of state. The indicated experimental data {filled
circles) are for hydrogen. The upper set of broken lines ( ) describes a TFCMIX-Monte Carlo pressure comparison; the lower
set indicates the relative importance of nonideal effects in low density theory. The line rH = 143 represents the locus of melting for
a pure hydrogen metallic solid. The solid ( ) and dot-dash ( ) lines are evolutionary tracks and p-T loci at fixed times,
respectively, from evolutionary calculations for a model of Jupiter.

theory with the 100 K and 300 K experimental data of Michels et al (1959), and Goodwin et al (1963). The inset
figure to the lower left shows the low-density theory, with the final value of a = 0.30, near the critical point (12.97
bar, 32.5 cm"3 g_1, 33.3 K). The dashed curve, obtained with an initial value oí a = 0.326, is included to
demonstrate the sensitivity to changes in the van der Waals constant.
The main figure illustrates the agreement of the composite model with the experimental shock data. (The zero
kelvin static data for solid hydrogen [Stewart 1956] are included only to indicate a lower bound on the pressure
isotherms for P ^ 20 kb.) The low-temperature metallic solid model is indicated above about 3 Mb. Even though
all these shock data points lie in the intermediate zone where the two models are combined through energy and
pressure interpolations, the agreement with the shock data near 200 and 900 kb is good. Based on Ross’s detailed
analysis (1974), these points correspond to 4500 K and 7000 K, respectively. The agreement at 900 kb is en-
couraging in that this point, at much higher pressure, was not available and therefore was not used to determine
the intermolecular potential. It should also be pointed out that these two points (200 and 900 kb) are each a
statistically reduced result of six separate measurements, which is the reason for their high precision. The inter-
polated fit to the 39.5-130 kb points is not as good, probably due to a lack of consistency among these shock data
points, which have much lower precision than the high-pressure points. The theoretical method of § II, to con-
struct a potential using all available experimental data for hydrogen, seems validated in that it gives best agreement
with the most accurate data points, i.e., the static data and the two high-pressure shock points. Figure 2 is a
diagram of (/>, T^-space for the SI solar mixture, showing the entire range of the final thermodynamic properties.
The experimental points for hydrogen are indicated, as well as an approximate boundary which indicates the upper
limit of validity for the molecular fluid model. The lower set of dashed lines (PniIP ^ n) are intended to demonstrate
the importance of low-density nonideal effects. These lines indicate boundaries below which the combined non-ideal
pressure terms, Pni, are less than the fraction n of the total pressure P. The nonideal terms include the intermolecular

© American Astronomical Society • Provided by the NASA Astrophysics Data System


197 5ApJ. . .199. .255G

No. 1,1975 THERMODYNAMICS OF DENSE H-He FLUIDS 263

Fig. 3.—Zero kelvin pressure comparison of Thomas-Fermi models for a range of elements from helium to uranium. The
Salpeter and Zapolsky (SZ) data are for the TFD model (curve a) and a TFD plus average correlation energy model (curve b). The
TFCMIX data correspond to the TFK model (TF corrected for quantum and exchange effects).

and interatomic repulsive and attractive terms and the Coulomb fluid effects. The results suggest that the low-
density model should be quite good for p < 0.1, the upper limit employed in this work.
For analysis of the high-density theory, there are no experimental data. There do exist, however, two inde-
pendent theoretical models which can be compared with our theory. The first comparison is with the zero tempera-
ture TF model of Salpeter and Zapolsky, shown in Figure 3. The SZ models give the zero kelvin pressure isotherms
for five elements ranging from helium to uranium. Their two models are {a) TFD (that is, Thomas-Fermi plus
exchange only), and (6) TFD plus electron correlation. The “average” correlation model is obtained by computing
the correlation energy as a function of the electron density averaged over the Wigner-Seitz sphere. This is stated
by SZ to be physically more realistic, in most cases, than the use of a local electron density. The curves {b) are
then to be compared with the TFCMIX results, and as shown in Figure 3, there is excellent agreement for low Z
and high Z at P > 1 Mb. For intermediate Z (carbon and magnesium), agreement is poor. The agreement for
P < 1 Mb becomes increasingly poor for all values of Z. Not only are the fundamental assumptions underlying
the Thomas-Fermi model becoming questionable in this pressure regime, but the correction terms predicted by
these different versions of the basic theory become large, so that details of the higher-order terms dominate the
results. Of the comparisons shown here, the most important region for study of the giant planet metallic interiors
is the case of He for P > 1 Mb, for which the agreement is quite good.
A second, completely independent theoretical approach which uses the Monte Carlo numerical method for a
system of charged classical nuclei with a uniform electron background was initiated by Brush et al. (1966). It was

© American Astronomical Society • Provided by the NASA Astrophysics Data System


197 5ApJ. . .199. .255G

264 GRABOSKE, OLNESS, AND GROSSMAN


extended by Hubbard and Slattery (1971) to include two nuclear species and the effects of nonuniformity of the
electron gas. In principle, this method should give the best theoretical results for mixtures of elements that interact
strongly in the presence of a highly degenerate electron gas. However, it requires large amounts of computer time
to obtain one F, T point for a single chemical composition, and its accuracy is dependent on an accurate potential,
plus sufficient system and configuration sizes. Hubbard and Slattery (1971) and Hubbard (1972) have published
Monte Carlo results for hydrogen and for H-He mixtures, including X = 0.7, Y = 0.3, which differs little from
the SI mixture used here. A comparison of the TFCMIX theory for the composition {X = 0.7, Y = 0.3) with
Hubbard’s results yields a set of lines delineating the agreement between the two theories. In Figure 2, the dashed
lines at p > 1 g cm-3 labeled 1, 10, and 20 percent mark the corresponding pressure differences between TFCMIX
and Hubbard’s data. The 20 percent line is probably quite close to the limit where all metallic theories cease to be
valid; that is, it lies close to 1 g cm-3 which is near the phase transition to the molecular state. The region below
the 20 percent line and above the 0.9 Mb experimental point, containing the molecular-metallic fluid phase
transition, is the region of uncertainty for predominantly hydrogen dense fluids. From this analysis, it seems
probable that a complete theoretical description of the H-He fluid can be made which is accurate to better than 10
percent over most of the dense fluid region, with a maximum uncertainty not much larger than 20 percent in the
vicinity of the fluid-fluid phase transition.
To relate the theoretical and experimental thermodynamic properties to the density-temperature region of
interest in evolutionary studies of giant planet structure, structure lines for Jupiter are included in Figure 2. The
lines (Graboske et al 1975), a-a and b-b identify the (/>, r)-path of the center and core boundary of the Jovian
model as it evolves from an age of 104 years to 1010 years. The broken lines c, d, and e are interior structure lines
at early, intermediate, and late epochs, while f-f is an inner envelope structure line for the late epoch model. A
comparison of the planetary structure lines with the thermodynamic information illustrates the areas where good
models can be expected and areas where future improvements are desirable.

One of the authors (A. S. G.) wishes to acknowledge support for this work from NSF grant GP-25949.
REFERENCES
Allen, C. W. 1963, Astrophysical Quantities (London: Athlone). Kompaneets, A. S., and Pavlovskii, E. S., 1957, Soviet
Brush, S. G., Sahlin, H. L. and Teller, E. 1966, J. Chem. Phys.—JETP, 4, 328.
Phys., 45, 2102 (BST). Lebowitz, J. L. 1964, Phys. Rev., 133, A895.
Dick, R. 1971, reported in Kerley, G. I., 1972, Phys. Earth Lebowitz, J. L., Helfand, E., and Proestgaard, E. 1965,
Planet. Interiors, 6, 78. /. Chem. Phys., 43, 774.
Goodwin, R. D., Diller, D., Roder, H. and Weber, L. 1963, Margenau, H., and Kestner, N. R. 1971, Theory of Inter-
J. Res., NBS, A2, 173. molecular Forces (Oxford: Pergamon Press).
Graboske, H. C. Jr., Pollack, J. B., Grossman, A. S., and Mason, E. A., and Hirschfelder, J. O. 1956, J. Chem. Phys.,
Olness, R. J. 1975, Ap. J., 199, 265. 26, 756.
Grossman, A. S., Graboske, H. C. Jr., Pollack, J. B., Rey- Michels, A., de Graaf, W., Wassenaar, T., Levelt, J., and
nolds, R., and Summers, A. 1972, Phys. Earth Planet. Louwerse, P. 1959, Physica, 25, 25.
Interiors, 6, 91. Neece, G. A., Rogers, F. J. and Hoover, W. G. 1971, J.
Hansen, J. P. 1972, Phys. Letters, 41 A, 213. Chem. Phys., 7, 621.
Hirschfelder, J. O., Curtiss, C. F., and Bird, R. B. 1954, Rogers, F. J., Graboske, H. C. Jr., and DeWitt, H. E. 1971,
Molecular Theory of Gases and Liquids (New York: Wiley), Phys. Letters, 34A, 127.
1063. Ross, M. 1974, /. Chem. Phys., 60, 3634.
Hoover, W. G., Ross, M., Bender, C. F., Rogers, F. J. and Salpeter, E. E. 1973, Ap. J. {Letters), 181, L83.
Olness, R. J. 1972, Phys. Earth Planet. Interiors, 6, 60. Salpeter, E. E., and Zapolsky, H. S. 1967, Phys. Rev., 158,
Hubbard, W. B. 1932, Ap. J., 176, 525. 876 (SZ).
Hubbard, W. B., and Slattery, W. L. 1971, Ap. J., 168, 131. Slattery, W. L., and Hubbard, W. B. 1973, Ap. J., 181, 1031.
Hubbard, W. B., and Smoluchowski, R. 1973, Space Sei. Smith, F. T. 1972, Phys. Rev. A, 5, 1708.
Rev., 14, 599 (HS). Smoluchowski, R. 1967, Nature, 215, 691.
Kalitkin, N. N. 1960, Soviet Phys.—JETP, 11, 1106. Stewart, J. W. 1956, J. Chem. Phys. Solids, 1, 146.
Karplus, M., and Kolker, H. J. 1964, J. Chem. Phys., 26, van Thiel, M., and Alder, B. 1966, Molec. Phys., 10, 427.
756. van Thiel, M., and Hord, B. L. 1973, unpublished data.
Kirzhnits, D. A. 1957, Soviet Phys.—JETP, 5, 64.
—. 1959, ibid., 8, 1081.

H. C. Graboske and R. J. Olness: Lawrence Livermore Laboratory, L-504, P.O. Box 808, Livermore, CA 94550
A. S. Grossman : Department of Physics, Iowa State University, Ames, IA 50010

© American Astronomical Society • Provided by the NASA Astrophysics Data System

You might also like