Sie sind auf Seite 1von 17

EPJ manuscript No.

(will be inserted by the editor)

A simple thermodynamic model of diluted


hydrogen gas/plasma for CFD applications
L. Quartapelle1a and A. Muzzio2
1
Dipartimento di Scienze e Tecnologie Aerospaziali, Politecnico di Milano
Via La Masa 34, 20156 Milano, Italy;
2
Dipartimento di Energia, Politecnico di Milano
Via Lambruschini 4, 20156 Milano, Italy.

Received: date / Revised version: date

Abstract. This work describes a simple thermodynamic model of the hydrogen gas at low densities and for
temperatures going from those involving quantum rotations of ortho- and para-hydrogen up to the fully
ionized state. The closed-form energy levels of Morse rotating oscillator given in Harris and Bertolucci [14]
(but not those in Morse’s original paper) are shown to provide an internal partition function of H2 that
is a sufficiently accurate representation of that exploiting the state-of-the-art spectrum of roto-vibrational
levels calculated by Pachucki and Komasa [13]. A system of two coupled quadratic equations for molecular
dissociation and atomic ionization at equilibrium is derived according to the standard statistical mechanics
by assuming that the system is an ideal mixture containing molecules, neutral atoms and noninteracting
protons and electrons. The system of two equations reduces to a single quartic equation for the ionization
unknown, with the coefficients dependent on the temperature and the specific volume. Explicit relations
for specific energy and entropy of the hydrogen ideal gas/plasma model are derived. These fully compatible
equations of state provide a complete thermodynamic description of the system, uniformly valid from low
temperatures up to a fully ionized state. The comparison with results of other models developed in the
framework of the physical picture shows that the proposed elementary model is adequate for computational
fluid dynamics purposes, in applications with the hydrogen gas under diluted conditions and when the
dissociation and ionization can be assumed at the thermodynamic equilibrium.

PACS. PACS-key 51.30.+i

1 Introduction vibrational temperature of the molecule, which makes it


impossible the fundamental relation to be expressible ex-
Monatomic ideal gases at low densities and for tempera- plicitly in analytical form.
tures lower than the ionization energy are described prop- On the other hand, the harmonic model is conflicting
erly by the pressure equation of state P V = N kT together with the possibility of molecular dissociation occuring at
with the specific heat equation cv = constant, both deriv- T > Tv . This difficulty is attenuated in some models where
able from a single fundamental thermodynamic relation char
v (T ) is halved, as proposed by Lighthill [3], which cor-
S = S(E, V, N ), available in an explicit analytic form, responds to take the vibrational degrees of freedom of the
see e.g., [1, p. 373, eq. (16.73)]. Molecular gases do not molecules always half-excited at any temperature, see also
admit a similar compact representation due to the pres- [4, Sec. 5.3, p. 157]. A physically sounder procedure for
ence of the internal rotational and vibrational modes. In overcoming the harmonic approximation consists in mak-
the simplest case of a diatomic gas, molecular vibrations ing recourse to the Morse potential [5] which accounts for
are often described within the harmonic approximation, the vibrational nonlinearity and allows for the molecu-
by an infinite ladder of the uniformly distributed energy lar disintegration into atoms. The Morse vibrational spec-
levels of the Einstein quantum oscillator, see [1, p. 336] or trum is still discrete but with a nonuniform spacing of the
[2, p. 116]. This model yields a nonconstant specific heat energy levels of increasing density, and has a finite number
cv due to the temperature dependent vibrational contri- of excited states, before the molecular dissociation. A ther-

Tv 2 eTv /T
bution char
v (T ) = R T (eTv /T −1)2
, with Tv denoting the modynamic model for diatomic gases based on the rotat-
ing Morse oscillator was developed by Gordillo-Vázquez
Send offprint requests to: and Kunc [6] and was compared for several diatomic gases
a
e-mail address: luigi.quartapelle@polimi.it to that relying upon the Tietz–Hua rotating oscillator.
2 L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications

The molecular rotations present however an additional alternative closed-form expression for the roto-vibrational
difficulty: rotations and vibrations are in principle cou- energy levels is also derived from the classical reference
pled together. For most diatomic molecules, the moment Harris&Bertolucci [14]. These energy levels differ from
of inertia is so large and the rotational temperature Tr so Morse’s original ones [5] only for the term representing
small, of the order of only a few kelvins, that the rota- the interaction between rotations and vibrations. Rather
tional states may be uncoupled from the vibrational ones. unexpectedly, the Harris&Bertolucci spectrum is found to
Consequently, the molecular rotations can be treated ac- give the internal partition function H2 notably different
cording to a semi-classical description of fully excited rigid from that provided by Morse spectrum and a fairly ac-
rotations valid for T > Tr , resulting in a constant contri- curate approximation to the partition function based on
bution R to cv . On the contrary, the hydrogen molecule Pachucki and Komasa spectrum. The Harris&Bertolucci
H2 (and its isotopes) has a very low moment of inertia, as approximate spectrum is employed here to formulate a
noted by Landau & Lifshitz [7, p. 139], yielding the fairly simple closed-form analytical model of the thermodynamic
high value Tr = 88 K, so that the roto–vibration coupling properties of the hydrogen gas, including dissociation and
must be taken into account. ionization, a result that seems to be new, to the best of
In the range of high temperatures, say T ≫ 5000 K, a our knowledge.
thermodynamic description faces with another difficulty— As far as the atomic partition function of H is con-
hydrogen ionization. For this element, atomic ionization cerned, its divergence is circumvented by means of a sim-
is most simple since the spectrum of the electron states ple modification introduced originally by Fermi [15]. The
around the proton nucleus is known exactly and can be entire analysis is conducted in the framework of the statis-
used to compute the electronic contribution to the par- tical mechanics of noninteracting particles. A final thermo-
tition function of H. Unfortunateley, the series diverges dynamic model is obtained which describes the hydrogen
due to the long-range character of the electric interaction: gas/plasma at equilibrium by means of fully compatible
therefore atomic ionization requires a subtler mathemat- simple equations of state and is at the same time uniformly
ical treatment than molecular dissociation. In principle, valid for any temperature in the limit of low density.
the equations of states of the ionized hydrogen gas can be
derived by statistical mechanics starting from the com-
plete Hamiltonian of the system, including the interaction 2 Molecular partition function
energy between the charges, within the framework called
The starting assumption is that of an ideal gas accord-
physical picture, see, e.g., the classical studies of Hummer,
ing to the classical Maxwell–Boltzmann statistics. This is
Mihalas and Däppen [8] and [9], or the work of Alastuey 
V 1/3 h
et al. [10] on exact results for the hydrogen plasma at possible provided λT ≪ N , where λT = (2πmkT )1/2
low temperatures. Along the same rigorous lines, the re- is the thermal de Broglie wavelength and N is the number
cent work of Alastuey and Ballenegger [11] has derived the of identical particles of mass m contained in a volume V
equations of state covering molecular, atomic and plasma and at temperature T , see, e.g., [1, p. 405]. The Helmholtz
phases by means of the Ebeling function and a new four free energy F of a system of noininteracting particles con-
body partition function. sisting of NH2 molecules and NH atoms of hydrogen is [12,
Within a reduced setting, for low densities, say in the p. 185]
range ρ < 1 kg/m3 , the electrical charges are only weakly  NH2 NH 
coupled, see, e.g., [12, p. 216], and one may attempt to Z H2 Z H
F (T, V, NH2 , NH ) = −kT ln . (2.1)
formulate a simpler thermodynamic model by assuming NH2! NH!
that, after ionization, all charges behave as noninteracting Such Helmholtz potential is based on the assumption that
particles. the gas is a mixture of ideally noninteracting molecules
The aim of this work is to present a complete ther- and neutral atoms, which holds only in the limit of a low
modynamic description of the diluted hydrogen gas in- density.
cluding the rotational–vibrational coupling active at low The partition functions of the hydrogen molecule H2
T , the molecular dissociation process occurring for T > and atom H in (2.1) are given respectively by
1000 K, and the atomic ionization which may occur for
T > 5000 K. Consistent expressions for the specific energy ZH2 (T, V ) = ZHtr2 (T, V ) ZHrv,nuc
2
(T ) ZHel2 (T ),
and entropy of the gas/plasma system at thermodynamic (2.2)
equilibrium are derived in an explicit form, involving the ZH (T, V ) = ZHtr (T, V ) ZHnuc,el (T ).
internal partition function of the molecule H2 and a suit- The translational partition function of the molecule is
able reduced version of the partition function of the hy-  3
drogen atom. tr mH2 kT 2
ZH2 (T, V ) = V (2.3)
The former is first evaluated from the accurate nu- 2π~2
merical energy levels of all the roto-vibrational states be-
fore dissociation determined recently by Pachucki and Ko- and similarly for the atomic one ZHtr (T, V ), while
masa including nonadiabatic corrections [13]. These ener- ZHel2 (T ) = gHel2 e−EH2 ,0 /kT ,
gies are used here to compute the internal partition func- (2.4)
tion needed to model molecular dissociation of H2 . An ZHnuc,el (T ) = gHn gHel e−EH,0 /kT ,
L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications 3

where gHn = 2sH + 1 = 2 is the multiplicity of the spin 104


states of the atomic nucleus of H of spin sH = 21 and g el
is the statistical weight of the electronic state. Due to the Harris–Bertolucci (1989)
(2009)
Pachucki–Komasa
homonuclear nature of H2 , its internal partition function
cannot be factored in a nuclear and a rotational part, cf.

zrv
[2, p. 185]. Moreover, since rotations and vibrations are Rotating Morse oscillator (1929)
coupled together, they are considered jointly in the factor
ZHrv,nuc
2
(T ). By summarizing, the complete partition func-
tions of the molecule and the atom will be, respectively,
 3
mH2 kT 2 rv,nuc 0
ZH2(T, V ) = ZH2 (T ) gHel2 e−EH2 ,0 /kT V, 104 3 104 5 104 7 104 9 104
2π~2
 3 T [K]
mH kT 2 n el −EH,0 /kT
ZH (T, V ) = gH gH e V. (2.5) Fig. 1. Internal partition function zrv (T ) of H2 for differ-
2π~2
ent roto-vibrational spectra. Continuous line: numerical en-
The internal partition function of the diatomic molecule ergy levels of Pachucki and Komasa [13]. Dotted curve: closed-
must include a summation over its different electronic form approximate spectrum (2.9) of Harris and Bertolucci
states, and the summations over the rotational and vi- [14]. Dashed curve: closed-form approximate spectrum (2.17)
brational states, which are both dependent on the elec- of Morse [5].
tronic configuration, see, e.g., Babou et al. [16]. Initially,
one can consider only the ground state of the molecule
and disregard any electronic excitation. This assumption It is interesting to compare this partition function with
is used to simplify the analysis but can be removed by in- the one provided by approximate analytical solutions to
cluding later higher electronic states of the molecule with the Schrödinger equation for the nuclear motion with the
~2 j(j+1)
the associate roto-vibrational states. By taking into ac- effective Morse potential Vjeff (r) = V (r) + 2µ r2 which
count the combined effect of the multiplicity of the hy- includes the centrifugal rotational term of the molecule,
drogen nuclear spin sH = 12 and of the degeneracy of the µ being its reduced mass. In particular, the energy levels
rotational quantum
 number
 j = 0, 1, . . . through the co- can be expressed in closed form according to the energy
efficient ∆j ≡ 2 − (−1)j (2j + 1), the internal partition spectrum given in Harris and Bertolucci [14, p. 115, eq.
function of the hydrogen molecule is assumed in the fol- (3.45)], see also [17, p. 56, eq. (7.11)]. The expression is
lowing form made transparent by introducing the dimensionless pa-
rameters
JX Nj
max X
rv,nuc
ZH2 (T ) = ∆j e−Ej,n /kT . (2.6) ~β 1
χe ≡ √ and κ≡ , (2.8)
j=0 n=0 2µDe βre

The external sum in (2.6) limited to j even or to j odd the former being twice the usual anharmonicity constant
gives the para- and ortho-hydrogen, respectively. To eval- xe = χe /2. In terms of these parameters, the complete
uate this function, the upper bounds Jmax and Nj of j expression of the Harris–Bertolucci approximate spectrum
and n in the two sums must be known. reads
The roto-vibrational energy levels Ej,n of H2 before Ej,n
HB
 
dissociation are taken from the work of Pachucki and Ko- = −1 + 1 − j(j + 1)(κ2 χe )2 j(j + 1)(κχe )2
masa [13]. In Figure 1 we plot the function zrv (T ) = De    
+ 2 − n + 12 χe n + 21 χe (2.9)
e−De /kT ZHrv,nuc (T ), which should be indicated more pre-
2
nuc
cisely by zrv (T ). The energy De is the depth of the em- 
− 3j(j + 1) n + 21 (1 − κ)(κχe )3 .
pirical Morse potential [5]
 2  The internal partition function based on the Harris–Ber-
V (r) = De 1 − e−β(r−re ) − 1 , (2.7) tolucci spectrum will be written in the following form

where r is the internuclear distance, re is its equilibrium JX


max Nj
X j
value and β −1 is the length scale of the Morse elastic force. HB
zrv (T ) = ∆j e−bj Tr /T e−an Tv /T , (2.10)
The values for the hydrogen molecules are given in Tables j=0 n=0
1 and 2 and De = 4.75 eV. The continuous curve of Figure
1 represents the partition function based on the full finite where kTr = ǫ0,r = ~2/(2µre2 ) and kTv = ~ωe = 2De χe .
set of accurate dissociation energy values for 301 roto- In this form of the sum, an exponential term dependent
vibrational states of the ground state 1 Σg+ calculated in only on the angular quantum number j and the rotational
[13], before the molecule disintegrates. The bounds for the temperature Tr is factored out from the internal sum on
summations are Jmax = 31 and Nj = 14, 13, . . . , 1, 0, for the vibrational states. The coefficients in the exponentials
j = 0, 1, . . . , 31, with repetitions. are given by the approximate spectrum (2.9) and are found
4 L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications

Table 1. Morse parameters of H2 [18]. The range of the rotational quantum number j is limited
vib rot
by the maximum values Jmax and Jmax which are obtained
µ (amu) β (Å−1 ) β −1 (Å) re (Å) κ = 1/βre by solving the equations N (j) = 0 and N rot (j) = 0 for
vib

j, respectively. The latter are simple quadratic equations


H2 0.504 1.905 0.525 0.741 0.7085
of the type j(j + 1) + constant = 0, cf. [6]. Then we define
vib rot

Table 2. Other Morse parameters for H2 [18]. Jmax ≡ min Jmax , Jmax . (2.14)
vib rot
χe Nmax De (eV) U (eV) For H2 , it results Jmax = 43 and Jmax = 24, so that Jmax =
24 and the cutoff curve has actually two parts. The two
H2 0.0571 17 4.75 4.48 curves may intersect for j ∗ solution of
2 κ2 − κ + 31
Table 3. Temperatures of H2 in kelvin. j ∗ (j ∗ + 1) = . (2.15)
χ2e κ4 (κ − 31 )(5 − 3κ)
Tcr Tr Tv TD Ti
For H2 , the solution is found to be j ∗ = 16 and n∗ = 14.
H2 33.04 88.3 6300 55 121 157 800 The first (lower) part is due to stationarity with re-
spect to vibrations while the second (upper) part is due
to stationarity with respect to rotations. Thus, the cutoff
to be value Nj for any admissible rotational mode j of the H2
molecule described to the Harris–Bertolucci spectrum will
h 
ajn = 1 − n + 21 χ2e be defined by
i  (2.11) ( vib
− 3j(j + 1)(1 − κ) κχe TTvr n + 12 , N (j) for 0 ≤ j ≤ j ∗
Nj = (2.16)
  N rot (j) for j ∗ ≤ j ≤ Jmax
bj = 1 − j(j + 1)(κ2 χe )2 j(j + 1), (2.12)
for 0 ≤ j ≤ Jmax and 0 ≤ n ≤ Nj , with Jmax and Nj to be and gives a total of 342 states.1 The function zrv
HB
(T ) is the
defined below. Notice the occurrence of the temperature dotted curve in Figure 1, with the values of the Morse pa-
ratio Tr /Tv as a factor in the term coupling rotations with rameters of H2 given in Tables 1 and 2. The curve is found
vibrations. Thanks to the factor Tr /Tv , the evaluation of to provide a fairly good approximation of zrv (T ) that is
zrv
HB
(t) with the inner summation on n and the outer on based on the accurate numerical energy levels of Pachucki
j turns out to be more stable numerically than the other and Komasa [13]. Admittedly, near the cutoff the Harris–
way around. Bertolucci energy levels are of decreasing precision, but
To define the upper bounds in the two sums of (2.10) the accuracy of the remaining admissible states is such
it is necessary to establish a cutoff for the admissible val- that an overall fairly good representation of the partition
ues of j and n, within the considered approximation of function is achieved for temperatures which are not too
the spectrum. For the nonrotating Morse oscillator, i.e., large. This quite positive behaviour is not shared by the
for j = 0, the maximum vibrational mode is easily found spectrum of the Morse rotating oscillator [5, p. 64]. Using
to be Nmax = 1/χe − (1/2). According to the nonrotating our parametrization χe -κ, the original Morse spectrum as-
Morse model, vibrational states with n > Nmax are impos- sumes the same expression (2.9) of Harris–Bartolucci ex- 
sible: the molecule disappears and is replaced by its two cept for the coefficient of the coupling term j(j +1) n+ 12 ,
constituent atoms. When rotations are allowed within the which is replaced as follows
empirical Morse model, the limit of molecular existence
−3(1 − κ)(κχe )3 → −(κχe )2 χe . (2.17)
depends on the rotational state and can be defined by the
stationarity of the energy of the vibro-rotational levels. In In this case the conditions ∂Ej,n /∂n = 0 and ∂Ej,n /∂j =
M M

particular, by continuity, at low rotational numbers j the 0 give relations NvibM


(j) and Nrot
M
(j) similar to those in
energy stationarity will be achieved by ∂Ej,n /∂n = 0, a (2.13), but with always Nvib (j) > Nrot
M M
(j) and moreover
criterion employed also in [16, p. 420]. However, for the en- Nvib (0) = Nrot (0). Then, the cutoff for the Morse approx-
M M

ergy levels (2.9) of H2 at higher j this derivative does not imate spectrum can be based on either bounds Nvib M
(j)
vanish and a cutoff is attainable only via the stationarity or Nrot
M
(j), but in both cases the corresponding partition
at fixed n, namely by the condition ∂Ej,n /∂j = 0. Thus, function is the dashed curve reported in Figure 1. The dif-
the curve of energy stationarity that delimits the region ference with respect to the curve based on Pachucki and
of allowed quantum states before the hydrogen molecule Komasa energy levels is much lager than for the Harris–
breaks into atoms consists of two portions, associated with Bertolucci levels.
the conditions ∂Ej,n /∂n = 0 or ∂Ej,n /∂j = 0, respec- 1
tively. One finds the bounds of n for fixed j The combined rotational–vibrational character of the cutoff
(2.16) is a distinctive feature of the Harris–Bertolucci spectrum
of the molecular hydrogen H2 , a peculiarity shared only by the
N vib (j) = Nmax − 23 (1 − κ)κ3 χe j(j + 1),
diatomic molecules LiH, with j ∗ = 39 & n∗ = 20, and Li2 ,
1 1 2κ3 χe (2.13) with j ∗ = 106 & n∗ = 37. For any other diatomic molecule, the
N rot (j) = − − j(j + 1). cutoff of the Harris–Bertolucci spectrum is merely vibrational.
3(1 − κ)κχe 2 3(1 − κ)
L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications 5

0.09 troduced, namely,


0.08
0.08
JX Nj
0.07 max X j
−b̃j /t
HB
zrv (t) = ∆j e e−an /t , (2.19)
∆Ej,n /De

0.06
0.06
Morse
0.05 j=0 n=0
j=2n
0.04
0.04
0.03
where b̃j = bj Tr /Tv . Then, we introduce three sets of func-
0.02
0.02
j=n tions of the dimensionless temperature
0.01
Harris–Bertolucci
j=n    
j=2n Aj (t) Nj 1
00 X  j  −ajn /t
0 10 14
Bj (t) ≡  an  e , 0 ≤ j ≤ Jmax , (2.20)
n
Cj (t) n=0 (aj )2
n
Fig. 2. Dimensionless energy difference ∆Ej,n /De between
the Pachucki–Komasa numerical energy levels Ej,n and the whose derivatives satisfy the simple relations dAj (t)/dt =
HB
Harris–Bertolucci levels Ej,n of equation (2.9) (dotted curves) Bj (t)/t2 and dBj (t)/dt = Cj (t)/t2 . Then, in terms of the
M
or the Morse levels Ej,n of equation (2.17) (dashed curves), definitions (2.20), we introduce the following three func-
along two representative straight lines j = n and j = 2n. tions
 HB   
zrv (t) JXmax
Aj (t)
xHB   b̃j Aj (t) + Bj (t)  −b̃ /t
The origin of this difference can be clarified by plotting rv (t) ≡   ∆j e j
in Figure 2 the energy difference in dimensionless form yrv
HB
(t) j=0 b̃2 A (t) + 2b̃ B (t) + C (t)
j j j j j
∆Ej,n /De along two straight lines j = n and j = 2n be-
tween the Pachucki–Komasa numerical energy levels Ej,n ,
assumed as reference values, and the approximate ana- 3 Dissociating hydrogen without ionization
lytical values of the Harris–Bertolucci spectrum (dotted
curves) or the Morse spectrum (dashed curves). For not For temperatures before the occurrence of ionization, say,
too small n, the error of Morse energy levels along the T < 5000 K, a simple thermodynamic model of the hy-
bisecting line j = n is substantially larger than that of drogen gas is easily obtained by finding the equilibrium
the Harris–Bertolucci levels. The energy difference is due composition of the dissociating gas that minimizes the
to the respective terms coupling rotations and vibrations, Helmholtz free energy, see, e.g. [12, p. 185]. Let NH2 be
according to (2.17). This difference increases further along the total number of H2 molecules in the volume V when
the second straight line j = 2n, thus demonstrating that all molecules are undissociated. The degree of dissociation
the representation of the interaction by Morse spectrum is α of the gas in a state with NH2 molecules is defined by
indeed too crude. On the contrary, the closed-form approx-
imate spectrum of Harris–Bartolucci (2.9) together with N H 2 − NH 2
the cutoff criterion defined by (2.13)–(2.16) represents a α= , (3.1)
NH2
viable and analytically convenient alternative to the use of
the full set of accurate numerical energy levels of Pachucki so that NH2 = (1 − α)NH2 and, thanks to the conserva-
and Komasa, at least for the limited scope of deriving ther- tion of the total number of hydrogen nuclei, 2NH2 + NH =
modynamic properties. In all our calculations, the vari- NH = 2NH2 , also NH = 2αNH2 = ςαNH2 . Here, the con-
ous thermodynamic equations of state evaluated starting stant ς is the symmetry factor of the molecule, ς = 2 for
from the roto-vibrational energy levels of Pachucki and homonuclear molecules and equal to 1 for heteronuclear
Komasa have been found to be virtually indistinguishable molecules. By minimizing the free energy (2.1), we obtain
from those derived from Harris and Bertolucci approxi- the equation for dissociation equilibrium
mate analytical spectrum (2.9).
To compute thermodynamic properties, the first and NH2 [ZH (T, V )]2
= . (3.2)
second derivatives of zrv (t) are needed. They can be evalu- NH 2 ZH2(T, V )
ated through the intermediate functions xrv (t) = t2 zrv ′
(t)
and yrv (t) = t2 x′rv (t) defined by Expressing NH and NH2 in terms of α, the equation for
the dissociation equilibrium becomes
  JX XNj  
xrv (t) max
εj,n −εj,n /t α2 K(T ) V
= ∆j e , (2.18) = 2 , (3.3)
yrv (t) ε2j,n 1−α ςH2 NH2
j=0 n=0
2

where εj,n = (Ej,n + De )/kTv . where the equilibrium constant K(T ) = [ZZHH(T,V )/V ]
(T,V )/V us-
2

When the approximate spectrum in closed form of Har- ing (2.5) with EH,0 = 0 and EH2 ,0 = 0, is defined by
ris and Bertolucci (2.9) is used, the derivatives can be eval-   32
uated more simply as follows. First, the function zrvHB
(t) ← T e−TD /T
zrv (T ) of the dimensionless temperature t = T /Tv is in-
HB K(T ) = Cd · , (3.4)
Tv zrv (T /Tv )
6 L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications

where TD = De /k = 55 121 K and


  32
(gHn )2 (gHel )2 mH kTv
Cd = √ . (3.5) 15
2 2 gHel2 2π~2
1
2
+ ǫhar
v (t)
The occurrence of the energy depth De = kTD in func-
1
tion K(T ) instead of the molecular dissociation energy ideal gas mixture 2
+ 2t
of the Morse anharmonic oscillator, U = −Ej=0, n=0 =

rot-vib energy
2 10
10
1 − χ2e De (UH2 = 4.483 eV), is only apparent. In fact, classical rotations
the denominator zrv (T/Tv ) contains the following factor & Morse vibrations
exp − 21 1 − χ4e Tv /T that combines, with e−TD /T , to t + ǫv (t)
give e−TU /T , with TU = U/k.
In terms of the dimensionless temperature t = T /Tv 5
coupled rotations
and of the specific volume v = V /(mH2 NH2 ), the equation and vibrations ǫrv (t)
for the dissociation equilibrium reads Harris and Bertolucci
or Pachucki and Komasa
3
α2 t 2 e−tD /t v
= , (3.6) 00
1−α zrv (t) ςH22 v̄d

where tD = TD /Tv = De /kTv = 8.749 and with the scale 0 1 2 3 4 5 6 7 8


constant vd for the variable v defined by t
Fig. 3. Dimensionless roto-vibrational energy of the H2
v̄d−1 = mH2 Cd . (3.7)
molecule as a function of the dimensionless temperature t =
T /Tv . Continuous curve: molecular energy ǫrv (t) of the ro-
Being gHel = 2 and gHel2 = 1 for the molecular ground
tating oscillator with Harris–Bertolucci or Pachucki–Komasa
state 1 Σg+ , we have v̄d−1 = 1.7655 × 106 kg/m3 and v̄d = spectrm. In the remaining three curves, the rotations are ac-
0.56641 × 10−6 m3/kg. counted for classically as fully excited. Dashed curve: t + ǫv (t)
The energy E of the hydrogen gas model is obtained with molecular vibrations described by the nonrotating Morse
from F (T, V, NH2 , NH ) by means of the relation E = F − anharmonic oscillator. Dotted curve: t + ǫhar
v (t) with harmonic
T (∂F/∂T )V,NH2 ,NH , which yields the dimensionless spe- molecular vibrations. Dot-dashed curve: function 21 + 2t which
cific energy ǫ = e/RH2 Tv (with e = E/mH2 NH2 and RH2 = is the internal energy contribution of H2 according to the ideal
4157.2 J/(kg · K)), gas mixture approximation, cf. equation (3.11) plus the zero-
point energy shift.
ǫ(t, α) = 23 (1 + α) t + (1 − α)[ǫrv (t) − tD ], (3.8)
where where tU = U/kTv . By including the zero-point energy
xrv (t) shift and using tD − tU = 12 1 − χ4e ≈ 12 , we have
ǫrv (t) = . (3.9)
zrv (t) 1 mix

2 +ǫ (t, α) = 23 (1 + α) t + (1 − α) 12 + 2t + αtD , (3.12)
At zero temperature ǫ(0, 0) = ǫrv (0) − tD . It is common
to normalize the energy by adding the energy depth De to be compared with relation (3.10). Thus, the term that
of the Morse potential so that the energy for T → 0 is should be compared with ǫrv (t) is simply 12 + 2t. The
positive and includes the zero point energy shift ~ωe /2, functions ǫrv (t) and 21 + 2t are represented in Figure 3
namely ǫ̃(t, α) = ǫ(t, α) + tD . The final expression of the by the continuous and dot-dashed curves, respectively.
new normalized energy, denoted by the same letter ǫ for Thus, the straight line 2t of the ideal gas mixture as-
notational simplicity, will be sumes a fully classical approximation of both rotations
and oscillations. For completeness, in the same figure we
ǫ(t, α) = 23 (1 + α) t + (1 − α) ǫrv (t) + αtD , (3.10) plot also the curves of two other possible approximations
and now ǫ(0, 0) = ǫrv (0), as required. for the molecular energy of H2 : the dashed curve corre-
Relation (3.10) for the internal energy must be com- sponds to the molecular energy t + ǫv (t) consisting of
pared with the analogous expression obtained by consid- fully excited rotations but quantum vibrations accord-
ering the mixture of ideal gases and summing directly the ing to the Morse anharmonic oscillator (detailed in Sec-
energy of the molecular and atomic components. Assum- tion 7); and finally the dotted curve is the molecular en-
1 e−1/t
ing that all energy contributions due to the electronic ex- ergy t + ǫhar
v (t) = t + 2 + 1−e−1/t for fully excited classical
citation can be neglected, the dimensionless expression of rotations and the Einstein quantum linear oscillator.
the energy of the ideal gas mixture assumes the form, see, The slope of the curves in Figure 3 at large tempera-
e.g., [27, p. 19], tures is related to the finite or infinite number of quantum
states of the molecule: the roto-vibrational model repre-
ǫmix (t, α) = 21 (7 − α) t + αtU sented by the continuous curve ǫrv (t) has a finite number
(3.11)
= 32 (1 + α) t + 2(1 − α) t + αtU , of states, leading to a zero slope; the Morse simpler model
L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications 7

77 20
20

66

ǫ
v = 103 m3/kg
10
10
55 102
10 1
0.1
ǫrv (t)
rot-vib energy

44

00
33
t + ǫhar
v (t)
0 1 2
t
22 1
+ 2t
2 Fig. 5. Dimensionless internal energy ǫ in the range of disso-
ciation temperatures for different values of v. Comparison of
11 the present thermodynamic model (continuous curve) with the
ideal gas mixture model of equation (3.12) (dot-dashed curve).

00

0 1 2 3 physically more complete model is due to the possibility


t of building a corresponding term for the entropy equation,
as it will be described below. In this way, two fully com-
Fig. 4. Particular of Figure 3 representing the dimensionless patible equations of state for energy and entropy can be
roto-vibrational energy of the H2 molecule. The same labeling built so as to guarantee a complete representation of the
for the curves used in the previous figure applies here. The dot-
fundamental thermodynamic relation of the gas system. In
ted curve corresponds to the model with fully excited classical
Figure 5 the internal energy e(T, v) of the present model
rotations combined with the harmonic molecular vibrations of
(continuous curves) is compared with the function (3.12)
the Einstein quantum oscillator, namely, t + ǫharv (t).
of the ideal gas mixture model (dot-dashed curves) for the
specific volume values v = 10ℓ m3 /kg, for ℓ = −1, 0, 1, 2, 3.
In the range of dissociation temperatures the two models
with ǫv (t) has a finite number of vibrational states but an agree fairly well except for small differences at the lower
infinite number of rotational states, implying a curve with tempertaures due to the straight line approximation of
slope 1, asymptotically; in the other two models (Einstein the vibrational energy of the ideal gas mixture. The dif-
quantum oscillator and the ideal gas mixture) both rota- ferences at the higher temperatures of the plot are instead
tional and vibrational states are infinite and the slope of due only to the fact that the continuous curves refer to a
the curve is 2 as t → ∞ and for any t, in the two cases. The thermodynamic mode including the possibility of ioniza-
great differences in the molecular energy in the range t > 2 tion (see below) which is not accounted in the ideal gas
3
and for a density not exceeding ρ ∼ 10 kg/m are however mixture approximation (3.12).
not important because in this range of temperatures the Similarly to the energy, the dimensionless specific en-
dissociation is complete and the contribution to the en- tropy σ = s/RH2 = S/mH2 NH2 /RH2 = S/(mH2 RH2 NH2 ) is
ergy of this term vanishes. For lower temperatures, in the obtained from F by S = −(∂F/∂T )V,NH2 ,NH = (E−F )/T ,
range of only partial dissociation, the differences are much to give
smaller and they are displayed more clearly in Figure 4   
which contains an enlargement of the bottom-left corner of 5 3 v
Figure 3. The initial unit slope of the curves of three mod- σ(t, v, α) = (1 + α) + ln t + ln
2 2 v̄d (3.13)
els corresponds to the rotational contribution t to the en- i=0
ergy which is built inside the function ǫrv (t) or is included + (1 − α) σrv (t) + Υς (α) + σ0 ,
explicitly in both the nonharmonic (nonrotating Morse)
and harmonic models. The comparison of the different en- where, cf. e.g., [19],
ergies for not too large temperatures reveals that the fully xrv (t)
classical approximation of the ideal gas mixture is a fairly σrv (t) = ln zrv (t) + ,
good approximation of the Einstein oscillator model ex- tzrv (t) (3.14)
cept for rather small t. Moreover, the gas mixture model Υς (α) = −(1 − α) ln(1 − α) − 2α ln(ςα).
gives an acceptable approximation of the energy and of its
first derivative, which is required to evaluated the specific The function σrv (t) represents the contribution of the rota-
heat and related thermodynamic quantites. The error in tional-vibrational entropy of the molecule while function
the energy values may become appreciably large only for Υς (α) is the entropy of the mixing of the molecular and
t > 2. The description of the molecular energy provided atomic species (function x ln x → 0 as x → 0 by de
by the roto-vibrating model ǫrv (t) is in general more sat- l’Hôpital rule). Finally σ0i=0 is an arbitrary constant. The
isfactory than the other models. The real interest of this presence of the contribution (1 − α)σrv (t) in the entropy
8 L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications

ensures the compatibility with the term (1−α)ǫrv (t) of the In a given generic thermodynamic state of the gas there
energy, similarly to nondissociating molecular ideal gases are NH2 molecules of H2 , together with NH neutral hydro-
with variable specific heats, see, e.g., in [20, p. 33]. In this gen atoms and Np protons and Ne = Np electrons, due
case, the full compatibility is guaranteed by the fact that to the overall charge neutrality of the gas. The degree of
the roto-vibrational contribution (1 − α)σrv (t) to the en- dissociation of the molecular hydrogen and the dgree of
tropy occurs together with the mixing entropy function atomic ionization are measured by the dissociation and
Υς (α). The leading part of entropy relation (3.13) is akin ionization coefficients
to the Sackur–Tetrode equation for monatomic gas [21]
but (3.13) applies to a roto-vibrating diatomic gas, with N H 2 − NH 2 Np
α= and i= . (4.2)
the dissociation α fixed by the equilibrum condition (3.6). NH 2 2αNH2
The solution to the latter, namely,
Thanks to the conservation of the total number of hydro-
p 
α(t, v) = 12 D0 (t, v) 1 + 4/D0 (t, v) − 1 , (3.15) gen nuclei, 2NH2 + NH + Np = NH = 2NH2 , and to the
charge neutrality, Np = Ne , the mole numbers of the gas
3 −tD /t components are easily found in terms of NH2
with D0 (t, v) = t 2zerv (t) ς 2vvd , is substituted into (3.10)
and (3.13) to obtain eventually the equations of state of NH2 (α) = (1 − α)NH2 ,
energy and entropy
 NH (α, i) = 2α(1 − i)NH2 , (4.3)
e(T, v) = RH2 Tv ǫ T /Tv , α(T /Tv , v) and (3.16) Np (α, i) = 2αiNH2 .

s(T, v) = RH2 σ T /Tv , v, α(T /Tv , v) , (3.17) The total number of moles is N (α, i) = (1 + α + 2αi)NH2
both being functions only of T and v. The two func- and the molar fractions are
tions provide a compatible parametric representation of NH2 (α) 1−α
the fundamental thermodynamic relation s = s(e, v) or XH2 (α, i) ≡ = ,
N (α, i) 1 + α + 2αi
e = e(s, v) of the roto-vibra-dissociating hydrogen gas.
The parametric form is strictly equivalent to the original NH (α, i) 2α(1 − i)
fundamental relation (2.1) of the free energy F (T, V, NH2 , NH ), XH (α, i) ≡ = , (4.4)
N (α, i) 1 + α + 2αi
we have started from. But the availability of the two ex-
plicit functions e(T, v) and s(T, v) is very useful in fluid Np (α, i) 2αi
Xp (α, i) ≡ = .
dynamics applications. In fact, the energy and entropy N (α, i) 1 + α + 2αi
relations are required in order to describe, for instance,
adiabatic irreversible transformations produced by shock The free energy of the system of noninteracting particles
waves and isentropic processes occurring in rarefaction comprising the four species H2 , H, p and e is given by
waves. On the other hand, for temperatures T > 5000 K  NH2 NH Np Ne 
hydrogen atoms can ionize and the effects of ionization Z H2 Z H Z p Z e
F = −kT ln , (4.5)
cannot be neglected anymore. A possible manner of in- NH2! NH! Np ! Ne !
cluding the atomic ionization effects in a thermodynamic
description of the hydrogen gas is proposed in the follow- where ZH2 = ZH2(T, V ), ZH = ZH (T, V ), Zp = Zp (T, V )
ing. and Ze = Ze (T, V ) are the partition functions of H2 ,
H, p and e, respectively. The assumption underlying this
Helmholtz potential is that any interaction between all
4 Dissociation and ionization at equilibrium particles can be neglected completely, as in an ideal gas
mixture. The assumption was acceptable at low densities
The model for the ionizable hydrogen gas proposed here for the dissociating hydrogen gas without ionization with
assumes that the gas is a mixture of molecular hydrogen free energy (2.1), but is always violated for a ionized gas,
H2 , atomic hydrogen H as well as free protons and elec- due to the electrostatic interaction between the charged
trons, resulting from the ionization of the neutral atoms. particles. However, when the gas density is such that the
The chemical model underlying the system consists in the Coulomb interaction energy between neighboring charges
dissociation and ionization reactions is small with respect to their kinetic energy, the ionized
gas can be considered as only weakly noindeal and the
H2 ⇋ 2H and H ⇋ p + e. (4.1) Helmholtz free energy (4.5) for noninteracting particles
can be assumed as the starting point, see, e.g., [12, p. 215].
The mixture comprises therefore the two chemical species, For example, in the hydrogen gas of interest here, with an
H2 and H, together with bare protons and electrons, p and atomic density N e and an average distance between the
e. Thus, the model excludes the possibility of the processes e
particles hri ≈ N −1/3 , the condition for an ideal behav-
of molecular ionization, H2 ⇋ H+ 2 + e and of formation 3
e kT
of negatively charged hydrogenic ions, H2 ⇋ H− + p. ior is N < qe2 /4πǫ0 which at T = 30000 K gives for the
As shown by Alastuey and Ballenegger [11], the reactions specific volume the condition v > 0.1 m3/kg. For higher
forming the ions H+ −
2 and H can be neglected. densities, corrections accounting for Coulomb interaction
L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications 9

are necessary, but their consideration and inclusion are The rule of Fermi consists in assigning an a priori prob-
beyond the scope of this work. ability of the quantum state, decreasing with the volume
The stationarity of the free energy (4.5) with respect of the excited atom and depending on the gas density, to
to dissociation and ionization gives the two equations give
X

NH2 Z 2 (T, V ) el
ZH (T, v) = gH el
n2 eTi /n
2
T −4B N̂ (v) n6
e (4.15)
= H , (4.6)
NH 2 ZH2(T, V ) n=1
X
∞ T
Np2 Zp (T, V ) Ze (T, V ) 1
1− n2 i 6
T − 4B N̂ (v) n ,
= . (4.7) = gHel eTi /T n2 e−
NH ZH (T, V ) n=1
4 3 4πǫ0 ~2
Using the expressions of NH2 , NH and Np in (4.3), the two where B = 3 πa0 , a0 = me qe2 = 5.29×10−11 m is the Bohr
relations give the following system of algebraic equations radius, and N̂ (v) = 1/(vmH ) is the density of hydrogen
( 2 nuclei. Thus, the definition ZHel (T, v) = gHel eTi /T zH (τ, v),
α (1 − i)2 = D(T, V, NH2 )(1 − α),
(4.8) allow us to introduce the atomic partition function, de-
αi2 = I(T, V, NH2 )(1 − i), pendent both on the dimensionless temperature τ = T /Ti ,
X
∞ 
1 1
where zH (τ, v) =
6
n2 e− 1− n2 τ − 4B N̂ (v) n , (4.16)
[ZH (T, V )]2 1 n=1
D(T, V, NH2 ) ≡ , (4.9)
4ZH2(T, V ) NH2 and on the specific volume v, due to the presence of N̂ (v)
in the Fermi cutoff (as in other similar rules, see below).
Zp (T, V ) Ze (T, V ) 1
I(T, V, NH2 ) ≡ . (4.10) The Fermi cutoff is a form of the occupation probability
2ZH (T, V ) NH 2 methods, that are used to have a finite internal partition
function zH , in the study of dense plasmas, see, e.g., [8]. In
The partition function for the hydrogen molecule is as
these applications the numbers Nn of atoms in the elec-
discussed so far, namely,
tronically excited state En are retained as independent
  32 thermodynamic variables. Some criticisms are presently
mH2 kT raised concerning the consistency of some expressions of
ZH2(T, V ) = zrv (T /Tv )
2π~2 (4.11) the free energy used by the occupation probability meth-
el
× g H2 e (De −EH2 ,0 )/kT
V. ods. These criticisms do not apply to the present deriva-
tion since our thermodynamic model is limited to diluted
The partition function of the hydrogen atom is taken in conditions and, above all, has a single variable NH to de-
the form scribe the total number of all neutral atoms, irrespective of
their state of electronic excitation. Stated in other terms,
ZH (T, V ) = ZHtr (T, V ) ZHnuc,el (T ) the Fermi cutoff is used here not to represent the occu-
 3 (4.12) pancy of the excited level, but only to reduce them to a
mH kT 2 n el finite number, so as to make a path toward the production
= gH ZH (T ) V.
2π~2 of pairs of (noninteracting) charges available, much in the
same manner of the molecular dissociation mechanism.
The electronic partition function ZHel (T ) must include the Alternatively, according to Zel’dovich [12, p. 199] an
energy levels of the electron in the excited bound quantum even simpler manner to avoid the divergence is to truncate
states2 of the hydrogen atom, namely, the energy values the sum by retaining only the electron states before the
size of the electronically excited atom is large enough to
IH
En = − , n = 1, 2, . . . (4.13) interfere with the neighboring atoms at the considered gas
n2 density. This leads to the truncated series, with τ = T /Ti ,
m q4 ncutoff (v) 1
where IH = 8he2 ǫe2 = 13.60 eV denotes the ionization energy X 1
1− n2
0
of the hydrogen. However, the partition function zHcutoff (τ, v) = n2 e− τ , (4.17)
n=1
X
∞ X

2 where
ZHel (T ) = gHel n2 e−En /kT = gHel n2 eTi /n T
, (4.14)
n=1 n=1
1  mH v  61
ncutoff (v) = (ncutoff ≥ 1). (4.18)
2 a30
with Ti = IH /k = 157 800 K, cannot be used directly
since the series is divergent. To circumvent this obstacle, Thus, the complete partition function for the hydrogen
a method originally introduced by Fermi is adopted [15]. atom for modelling the gas/plasma is assumed to be
 3
2
In this section, n denotes the principal quantum number mH kT 2 n el IH /kT
ZH (T, V ) = gH gH e zH (T /Ti , v) V.
of the atomic electron state, and is not the quantum number 2π~2
of the molecular vibrations, as elsewhere in the paper. (4.19)
10 L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications

Finally, the partition functions for the free proton and For all values v > 10−5 m3/kg and T < 106 K this equa-
electron are tion has been found to possess only one real root in the
 3 unit interval [0, 1]; for higher temperatures, a solution i
mp kT 2 −Ep,0 /kT extremely near to 1 is easily selected as the proper one
Zp (T, V ) = 2 e V,
2π~2 among three real admissible values. The dissociation α is
 3 (4.20) given by α = I(1 − i)/i2 .
me kT 2 −Ee,0 /kT A fourth-order equation for the electron partial pres-
Ze (T, V ) = 2 e V,
2π~2 sure Pe is considered in Capitelli’s monograph [27], whose
coefficients are defined in terms of the equilibrium con-
where the 2’s are due to the spin 12 of proton and electron. stants for dissociation and ionization. When they are known
By taking EH2 ,0 = −2IH and Ep,0 = Ee,0 = 0, the as a function of T , the gas total pressure P can be de-
equilibrium constants of dissociation and ionization are termined and subsequently also α and i. Actually, the
respectively equilibrium constants are function of two thermodynamic
 3 variables, and therefore the problem for the partial pres-
T 2 −TD /T [zH (T /Ti , v)]2 sure is nonlinear, beyond the algebraic quartic character.
Kd (T, v) = Cd · e , (4.21)
Tv zrv (T /Tv ) However, in practice the dependence of the solution on
  32 the second thermodynamic variable v is rather weak and
T e−Ti /T a couple of iterations are sufficient to guarantee a very
Ki (T, v) = Ci · , (4.22)
Ti zH (T /Ti , v) accurate solution to the pressure equation starting from
the initial value v0 = 1 m3/kg.
with the constant Cd is defined in (3.5) while On the contrary, the equation for i (4.28) allows the di-
  32 rect determination of the dissociation and ionization in the
2 mp me kTi gas for given thermodynamic conditions defined in terms
Ci = . (4.23)
gHn gHel 2πmH ~2 of variables T and v. The gas pressure is calculated only
subsequently from the values of T, v, α and i. From the nu-
Using the partition functions (4.11), (4.19) and (4.20), the merical viewpoint, we have solved the fourth-order equa-
system of the equilibrium equations is recast in the form tion for i and that for Pe by means of Ferrari algorithm,
( 2 see, e.g., [22], and have found that the solution of the for-
α (1 − i)2 = D(t, v)(1 − α),
(4.24) mer is much more stable numerically than the latter.
αi2 = I(t, v)(1 − i),

with the functions D = 41 mH2 Kd v/vd and I = mH2 Ki v/vi 5 Equations of state of hydrogen gas/plasma
controlling the coupled dissociation and ionization equi-
libria defined by The equation of state of the energy is obtained from E =
F −T (∂F/∂T )V, NH2, NH , Np , Ne which yields the dimension-
3 [zH (t/ti , v)]2 v less specific energy ǫ = e/RH2 Tv ,
D(t, v) = t 2 e−tD /t , (4.25)
zrv (t) vd
  32 ǫ(t, v, α, i) = 32 (1 + α + 2αi) t
t e−ti /t v
I(t, v) = , (4.26) + (1 − α)[ǫrv (t) − tD ] (5.1)
ti zH (t/ti , v) vi
+ 2(1 − i)[αǫH (t/ti , v) − 1] ti ,
where tD = T /TD = De /kTv , ti = Ti /Tv = IH /kTv and
where ǫH (τ, v) = xH (τ, v)/zH (τ, v) with also xH (τ, v) =
mH 2 C d τ 2 ∂zH (τ, v)/∂τ .
vd−1 = and vi−1 = mH 2 C i . (4.27)
4 The energy at zero temperature is ǫrv (0)−tD −2ti . It is
therefore convenient to normalize the energy by adding the
The calculation yields vd = 2.26564×10−6 m3/kg and vi =
depth energy De of the Morse potential for all molecules,
4.02157 × 10−3 m3/kg.
as well as the binding energy of all the hydrogen atoms
The two equilibrium equations (4.24) are both quadra-
2ti with the electron in its ground state n = 0. The nor-
tic, the first both in α and i, while the second only in i. The
malized energy is ǫ̃(t, v, α, i) = ǫ(t, v, α, i) + tD + 2ti and
system reduces to a single fourth order algebraic equation
its final expression, indicated by the same symbol ǫ for
i4 + ai3 + bi2 + ci + d = 0 (4.28) notational simplicity, will be

with the four coefficients a, b, c, d defined in terms of D(t, v) ǫ(t, v, α, i) = 32 (1 + α + 2αi) t


and I(t, v): + (1 − α) ǫrv (t) + αtD (5.2)
(D + 4I)I (D + 6I)I + 2α[(1 − i) ǫH (t/ti , v) + i]ti ,
a= , b=− ,
D − I2 D − I2 so that at zero temperature ǫ(0) = ǫrv (0).
(4.29)
4I 2 I2 The specific entropy is obtained in a manner similar to
c= , d=− . the energy from the relation S = −(∂F/∂T )V, NH2, NH , Np , Ne
D − I2 D − I2
L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications 11

which gives its dimensionless counterpart σ = s/RH2 : 106



σ(t, v, α, i) = (1 + α + 2αi) 25 + 32 ln t v = 103 m3/kg
  
v 102
+ (1 − α) σrv (t) + ln 105

cv [J/kg · K]
10
v1 1
   0.1
v
+ 2α(1 − i) σH (t/ti , v) + ln (5.3) equilibrium
v2 para
  104
v ortho
+ 4αi ln + Υ (α, i) + σ0 ,
v3
where the occurrence of the factor 4 in the third volu- 103
mic term must be noted. Here σH (τ, v) = ln zH (τ, v) + 10 102 103 104 105 106
xH (τ, v)/[τ zH (τ, v)], and T [K]
  32 Fig. 6. Specific heat cv (T, v) as a function of temperature
mH2 kTv
v1−1 = gHel2 mH2 , for different values of specific volume v in the hydrogen ideal
2π~2 gas/plasma model. The different curves for T < 103 K are valid
 3 only for low densities. In particular, the unlabeled dotted curve
n el mH kTv 2
v2−1 = g H g H mH 2 , (5.4) between the curves of the ortho- and para-hydrogen refers to
2π~2 the gas in a condition of “frozen” metastable equilibrium de-
√ 3 scribed in the text.
me mp kTv 2
v3−1 = 2mH2 .
2π~2
Moreover, function Υ (α, i) represents the mixing entropy several curves which are valid only in the limit of low den-
of the dissociating and ionizing gas/plasma, which is found sity and which correspond to different forms of the H2
to be gas. The continuous curve is for the molecular hydrogen
at equilibrium while the dashed and dot-dashed curves
Υ (α, i) ≡ − (1 − α) ln(1 − α) are for the ortho- and para-hydrogen forms, which are
(5.5)
− 2α(1 − i) ln[2α(1 − i)] − 4αi ln(2αi). obtained by limiting the sum in (2.19) to odd or even j,
respectively. The dotted curve of Figure 6 located between
Note that Υ (α, 0) = Υς=2 (α). The specific heat at constant those of the ortho- and para-hydrogen corresponds to the
volume cv = (∂e/∂T )v is calculated from e(T, v, α, i) by experimental results, with the gas in a state of “frozen”
      metastable equilibrium consisting of 34 ortho-hydrogen and
∂e ∂e ∂e 1
cv = + αT (T, v) + iT (T, v), (5.6) 4 para-hydrogen, as explained in [23, p. 472], such that
∂T α,i ∂α T,i ∂i T,α

where the partial derivatives αT = (∂α/∂T )v and iT = cfrozen


P = 43 cortho
P + 14 cpara
P . (6.1)
(∂i/∂T )v are evaluated from system (4.24) by implicit dif-
ferentiation. The specific heat at a constant
 pressure will The four curves reported in Figure 6 for T < 600 K are
be evaluated from cP = cv − T (∂P/∂T )2v (∂P/∂v)T (for found in fair agreement with those given in [23, p. 471].
the pressure equation P = P (T, v) see (6.2)). Moreover These curves of cv are valid only for a low gas density, say
the relations v > 0.1 m3/kg, since, for T about and less than 20 K, a
zrv (t) yrv (t) − x2rv (t) dense H2 becomes a liquid and undergoes a phase tran-
ǫ′rv (t) = (5.7) sition which is dependent on the para/ortho ratio. The
[tzrv (t)]2 failure of the ideal gas assumption at low temperatures
∂ǫH (τ, v) zH (τ, v) yH (τ, v) − x2H (τ, v) and high densities implies that real gas effects must be
= (5.8) included. A simple approach is to resort to the van der
∂τ [τ zH (τ, v)]2
Waals equation of state, with a critical temperature Tcr =
will be used, implying that the calculation of cv requires 33.04 K and a critical specific volume vcr = 0.033 m3/kg.
the moments yrv (t) and yH (τ, v) to assure a full consistency This would also lead to include a contribution to the spe-
in the determination of the involved second derivative. Of cific energy of the form −a/v, where a = 89 vcr RH2 Tcr is the
course, the energy equation of state will be defined by coefficient associated to van der Waals attractive forces.
e(T, v) ≡ RH2 Tv ǫ(t, v, α(t, v), i(t, v)), with t = T /Tv , and The study of real gas effects in H2 is beyond the scope
similarly for the entropy s(T, v). of the present work. The reader interested in these ef-
fects in H2 and in its allotrope forms of ortho- and para-
hydrogen is referred to [24] and [25]. At higher tempera-
6 Comparisons tures the specific heat cv depends also on v in consequence
of dissociation and ionization. The curves of cv (T, v) for
The curves of cv (T, v) are shown in Figure 6. For T < v = 10ℓ m3/kg, ℓ = −1, 0, 1, 2, 3 are shown in Figure 6:
600 K, the specific heat depends only on T but there are each has two distinct bells centered on the crossover tem-
12 L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications

1
40 H2

partial pressures
equilibrium
cP [J/K · mol]

para H
p
20 5

ortho

0
10
0 100 200 300 400 500 600 0 104 2 104 3 104
T [K] T [K]
Fig. 7. Specific heat cP (T ) of the three forms of H2 ideal gas Fig. 9. Dimensionless partial pressures as a function of T
for low T and small density. Continuous curves from Harris– with gas pressure P = 1 bar. Continuous curves: present work.
Bertolucci energy levels (2.9) and dotted curves from [26]. Dotted curves: Reference [27].

less partial pressures PH2 /P, PH /P and Pp /P have been


peratures, dependent on v, between the successive states,
plotted by the continuous curves in Figure 9. The refer-
namely molecular, atomic and ionized.
ence results [27, p. 12] are plotted as the dotted curves.
In Figure 7 the values of cP (T ) provided by the appox- The very small differences can be explained by the differ-
imate energy levels (2.9) in the range T < 600 K have been ent temperature dependence in the equilibrium constants
compared to those calculated in [26] by means of the nona- of the present model with respect to those in [27], which
diabatic eigenvalues of all bound and quasibound levels. are exponential functions. At this low pressure the disso-
The value of the proposed simple thermodynamic model ciation and ionization are to a good approximation in-
based on the Harris&Bertolucci approximate spectrum of dependent phenomena and the exponential dependence
Morse rotating anharmonic oscillator are found to be very on T is dominating over the other functions accounting
close to the accurate values of [26], with a maximum dif- for the translational energy of molecules and atoms, and
ference not exceeding 1% in the range 0 < T < 600 K. for their internal structure through the involved partition
The specific heat at constant pressure has been com- functions. For pressure P = 100 bar the partial pressures
pared also in the temperature range of ionization. In Fig- are reported in Figure 10 and the result of the two mod-
ure 8 we plot the cP (T, P ) as a function of T for three els are found to be appreciably different. At this higher
fixed values of pressure: P = 0.01, 1, and100 bar. The con- pressure the dissociation and ionization processes are cou-
tinuous curves refer to the present thermodynamic model pled and rather different partial pressure at equilibrium
while the dotted curves are reproduced from [27]. For the are found. The disagreement can be attributed to the dif-
intermediate pressure the agreement is satisfactory but at ferent temperature dependencies in the equilibrium con-
the other two pressures the differences are larger. stants, here by the functions of (4.25) and in [27] by purely
The equilibrium dissociation and ionization predicted exponential functions entering a single nonlinear equation
by the proposed model have been compared with the re- for the electron pressure. The role of the more complex
sults from the monograph of Capitelli, Colonna and D’An- temperature dependence in the coupled regime has been
gola [27]. For the fixed pressure P = 1 bar the dimension- verified by

5 105 1
P = 0.01 bar H2
partial pressures

4 105
H
3 105 p
P = 1 bar 5
2 105 P = 100 bar

105
0

104 2 104 3 104 4 104 0 2 104 4 104 6 104


T [K] T [K]

Fig. 8. Specific heat cP (T, P ) in the range of temperatures in- Fig. 10. Dimensionless partial pressures as a function of T
volving hydrogen ionization, for some values of P . Continuous with P = 100 bar. Continuous curves: present work. Dotted
curve: present thermodynamic model. Dotted curves: Refer- curves: Reference [27]. Dashed curves correspond to a merely
ence [27]. exponential dependence on temperature in functions D and I.
L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications 13

fixing the temperature in the functions other than the ex- To determine the limitation of the present model at
ponential of the proposed model. The corresponding par- higher densities due to not accounting for the electric in-
tial pressures for a fixed value T = 8000 K are plotted as teraction of the charges, we have calculated the pressure
the dashed curves in Figure 10. These curves are found to in the density range 1 < ρ < 103 kg/m3 at the three tem-
approximate fairly well the reference dotted curves of [27] peratures T = 104.1 K = 1.2589 × 104 K, T = 104.5 K =
in the full range of temperatures. 3.1623 × 104 K and T = 105.3 K = 1.9953 × 105 K. The
The behaviour of the thermodynamic model for the pressure values in this range provided by the proposed
hydrogen gas/plasma proposed here is checked at low den- method are compared with the values of the hydrogen
sities against the model of Alastuey and Ballenegger, in plasma model of Poteckhin, which includes the effect of
which the processes of recombination and formation of the electric interaction by means of an occupation proba-
molecules and atoms is taken into account within the phys- bility method [28]. The plots in Figure 12 show the three
ical picture via the Ebeling function and a four-body par- isotherms in the ρ-P plane with logarithmic scales. The
tition function [11]. The comparison is done along the iso- comparison of the two models for the highest tempera-
chore v = 1 m3/kg and considering the equation of state ture T = 1.9953 × 105 K shows clearly a fair agreement
of pressure P = −(∂F/∂V )T, NH2, NH , Np , Ne : in the considered density range. At lower temperatures,
the agreement is restricted to the domain of low densities;
RH2T in particular, for T < 104 K the electric interaction has
P (T, v) = [1 + α(T, v) + 2α(T, v) i(T, v)] (6.2)
v substantial effects on pressure even at the intermediate
3
made dimensionless with respect to its value in the purely density ρ ≈ 10 kg/m , namely already for v ≤ 0.1 m3/kg.
molecular state, to give p = P/(2RH2 T /v). The additional The comparison shows that the proposed method is suffi-
pressure term associated to the v-dependence of zH (τ, v) ciently accurate only in the limit of high temperatures and
is negligible since these variations are extremely small. of low density, while it is unsuited to describe the thermo-
The pressure as a function of T is reported in Figure 11: dynamics of dense hydrogen at not high temperatures.
the continuous line refers to the Fermi treament (4.16) to
make the series of the atomic partition function conver-
gent, the dotted curve is for the Zel’dovich cutoff (4.17),
107
and the dash-point curve is drawn from the reference re-
sult of [11].
The temperatures of crossover between the molecular
and atomic phases and between the atomic and ionized
ones are predicted with comparable accuracy by the meth-
ods. Along the considerd isochore the ionized charges are
weakly coupled and the present thermodynamic model 106 T = 1.9953 × 105 K
with Fermi treatment reproduces the results of [11] for
temperatures T > 5 × 104 K accurately. In this range
P [bar]

the cutoff of Zel’dovich overestimates the reference value


slightly. Both treatments give correct pressure values ap- T = 3.1623 × 104 K
preciably different from those of the Saha approximation
in the region of nearly complete ionization, not reported
105
in the figure, cf. [11]

104 T = 1.2585 × 104 K


p

0 103
103 104 105 106 1 10 100 1000
T [K] ρ [kg/m3 ]
Fig. 11. Dimensionless pressure along the isochore v = Fig. 12. Pressure depending on density ρ for the temperatures
1 m3/kg. Continuous line: present work with atomic partition T = 104.1 K = 1.2589 × 104 K, T = 104.5 K = 3.1623 × 104 K
function made convergent by the Fermi rule. Dotted curve: and T = 105.3 K = 1.9953 × 105 K. Continuous curves: present
present work with Zel’dovich truncation. Dot-dashed curve: thermodynamic model. Dashed curves: occupation probability
method of Alastuey and Ballenegger [11]. method for the hydrogen plasma of Poteckhin [28].
14 L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications

7 Gas/plasma model under classical rotations where ǫv (t) = xv (t)/zv (t). By adopting the same energy
normalization employed before, the addition of tD + 2ti
For T ≫ Tr , the classical approximation for the molecular gives the quantity ǫ̃⋆ = ǫ⋆ + tD + 2ti , and the final expres-
rotations of a homonuclear (symmetric) molecule Z rot (T ) = sion of the new normalized dimensionless energy, denoted
1
2 T /Tr can be exploited to avoid the calculation of the dou- by ǫ⋆ for notational simplicity, reads
ble sum of the roto-vibrational partition function zrv (t)
and replace it by the simpler function zv (t). The partition ǫ⋆ (t, v, α, i) = 12 (5 + α + 6αi) t
function for the hydrogen molecule under the assumption + (1 − α) ǫv (t) + αtD (7.9)
of fully excited rotations is
+ 2α[(1 − i) ǫH (t/ti , v) + i]ti
  23
mH2 kT T Here α and i are solution to the D⋆ -counterpart of system
ZH⋆2(T, V ) = (gHn )2 zv (T /Tv )
2π~2 2Tr (7.1) (4.24), through the solution of the quartic equation (4.28).
× g el e(De −EH2 ,0 )/kT V. The new equation for the entropy reads
H2

The (purely) vibrational partition function zv (t) can be σ ⋆ (t, v, α, i) = 21 (5 + α + 6αi)(1 + ln t) + (1 + α + 2αi)
  
evaluated starting from either the “nonrotating” eigenval- v
ues E0,n , with 0 ≤ n ≤ 14, of Pachucki and Komasa [13] + (1 − α) σv (t) + ln
v1⋆
or those of the nonrotating Morse oscillator. In the latter   
v
case the partition function and its moments are defined + 2α(1 − i) σH (t/ti , v) + ln (7.10)
quite simply by v2
 
    v
zv (t) N 1 + 4αi ln + Υ (α, i) + σ0⋆ ,
Xmax v3
xv (t) ≡ an  e−an /t , (7.2)
yv (t) n=0 a2 where σv (t) = ln zv (t) + xv (t)/tzv (t) and
n
  32
where Nmax = 1/χe − (1/2) and (gHn )2 gHel2 Tv mH2 kTv
(v1⋆ )−1 = mH2 , (7.11)
    2 Tr 2π~2
an = 1 − n + 21 χ2e n + 12 . (7.3)
while the constants v2 and v3 have already been defined
Using (7.1) with (4.19), the new dissociation equilibrium in (5.4).
)/V ]2
constant Kd⋆ (T, v) = [ZZH⋆ (T,V
(T,V )/V is found to be The two hydrogen gas models, one including the quan-
H2
tum treatment of rotation and the other assuming classical
1 zH2 (T /Ti , v) rotations, are compared by evaluating the sound speed by
Kd⋆ (T, v) = Cd⋆ · (T /Tv ) 2 e−TD /T , (7.4) means of the well-known expression
zv (T /Tv )
s  s  2  
where   32 ∂P T ∂P ∂P
1 (gHel )2 Tr mH kTv c= =v − . (7.12)
Cd⋆ =√ , (7.5) ∂ρ s cv ∂T v ∂v T
2 gHel2 Tv 2π~2
The values of c(T, v), based on the general gas/plasma
since the nuclear spin multiplicity of H and H2 in (4.19) model described in section 5, are shown in Figure 13 for
and (7.1), respectively, cancel in the equilibrium constant,
as usual in the approximation of fully excited rotations.
Correspondingly, the definition (4.25) of function D(t, v)
is replaced by 5 104
√ −t /t [zH (t/ti , v)]2 v 4 104 v = 0.1 m3/kg
D⋆ (t, v) = te D (7.6) 1
c [m/s]

zv (t) vd⋆ 3 104


10
fully ionized
monatomic ideal gas
ideal plasma 102
with the new scale vd⋆ for the specific volume defined by 2 10 4
10 3

rigid diatomic molecules


mH2 Cd⋆ 10 4
(vd⋆ )−1 = . (7.7)
4 0
The calculation yields vd⋆
= 0.3233 × 10 m /kg. The new −3 3
0 104 105
T [K]
equation of state for the energy is obtained in the form
Fig. 13. Sound speed c(T, v) as a function of T for different
ǫ⋆ (t, v, α, i) = 12 (5 + α + 6αi) t values of v. The dotted curves are for ideal gases with different
+ (1 − α)[ǫv (t) − tD ] (7.8) specific heat ratios γ. Rigid diatomic molecule γrig.dia. = 75 .
Monatomic: γa = 35 . Fully ionized ideal plasma: γplasma = 10 .
+ 2(1 − i) [αǫH (t/ti , v) − 1] ti , 3
L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications 15

different values of v. At low temperatures, the sound speed the Hugoniot relation is written in the form
of the hydrogen gas coincides with that of a diatomic
ideal gas with rigid molecules, with γrig.dia. = 57 , namely e(T, v) − e0 + 12 [P0 + P (T, v)](v − v0 ) = 0 (7.13)
p
crig.dia. (T ) = (7/5)RH2 T , for any density. Then, for T >
1000 K, the molecular oscillations come into play and make which represents an implicit definition of the function v =
the dissociation possible, so that the sound speed depends v RH (T ) along the Rankine–Hugoniot adiabat. This non-
also on v. About T ≈ Tv the dissociation tends to be com- linear equation can be solved by Newton method. Fig-
plete and the gas becomes monatomic, so all curves tend ure 15 contains the curves of the irreversible adiabats
to one and the √ same sound
p speed curve of the monatomic P = P RH (v) ≡ P (T RH (v), v), issuing from several initial
gas ca (T ) = γa RH T = (10/3)RH2 T . For higher T , the states, all with v0 = 103 m3/kg but for different dimen-
ionization sets in and the sound speed of the partially ion- sionless temperatures t0 = T0 /Tv = 0.05, 0.1, 0.25, 0.5,
ized gas depends substantially on its density, √ but tends 1, 2, 2.5, 3, 5.
uniformly
p to the limit value c plasma (T ) = γa 2RH T = Differently from the ordinary Rankine–Hugoniot adia-
2 (5/3)RH2 T which corresponds to a monatomic gas with bats of an ideal gas, which are always monotonic curves,
an average atomic mass half of that of H, the electron mass the adiabats in hydrogen may have one or three local ex-
being negligible with respect to the proton one. trema (function v = v RH (P )) depending on the initial
Figure 14 contains an enlargement of the curves of Fig- state, due to the fact that the pre-shock hydrogen condi-
ure 13 in the range of low temperatures and for the same tions may be very different. In fact, the gas can be initially
density values. The sound speed of the general thermody- either fully molecular, or mainly atomic after a great dis-
namic model (continuous curves) are compared to that of sociation, or also in a prevalently ionized states. According
the simplified gas/plasma model under classical rotations to each of these initial conditions, shocks of increasing in-
described in this section (dashed curves). The two models tensity may or may not involve dissociation or ionization
are found to agree fairly well at the lower densities and or both. This is at the origin of the very different be-
the differences do not exceed a few % at the highest con- haviours of the hydrogen adiabats shown in Figure 15. In
sidered density of v = 0.1 m3/kg, confirming the adequacy particular, the lowest curve starts at the smallest initial
of the simpler model for gasdynamic applications. temperature T0 = 315 K and the gas compression initially
follows the standard hyperbolic curve with P → ∞ tend-
As a last demonstration of the suitability of the pro- ing to a vertical asymptote.
posed thermodynamic model of the hydrogen gas/plasma
Three different possible asymptotes are plotted in the
for applications in computational fluid dynamics, we de-
figure as dotted vertical lines. They correspond to the
termine the Rankine–Hugoniot adiabats for shock waves
lower bound for the ratio v asym /v0 = (γ − 1)/(γ + 1) that
involving dissociation and ionization. Since the equations
the adiabat cannot overcome, since P → ∞, for three pos-
of state e(T, v) and P (T, v) in (7.9) and (6.2) are available,
sible ideal gas models. The first asymptote on the right in
the figure corresponds to a monatomic gas with γa = 53
and the value (v asym /v0 )a = 41 . The second intermediate
asymptote corresponds to a diatomic ideal gas with rigid
molecules and fully excited rotations such that γrig.dia. = 75
monatomic
and (v asym /v0 )rig.dia. = 16 . The third asymptote, on the
left, corresponds to a diatomic ideal gas with fully ex-
cited rotations and vibrations so that γvib.dia. = 79 and
104 10 1 v = 0.1 m3/kg (v asym /v0 )vib.dia. = 18 .
The compression adiabat lowest in the figure tends ini-
tially toward the most left asymptote (v asym /v0 )vib.dia. =
c [m/s]

rigid diatomic 1
8 , but, for higher P and T , the curve deviates as a conse-
103
quence of molecular dissociation. The dissociable molec-
ular hydrogen can become denser than permitted to its
undissociable ideal counterpart. The dissociation proceeds
further and a condition is reached in which compression
and temperature can increase while the gas density de-
creases: the molecular dissociation allows an augmentation
0 of the internal energy at a lower density due to duplica-
0 104 2 104 tion of the number of particles with their kinetic energy.
At higher and higher compression and temperature, the
T [K]
atomic hydrogen starts to ionize and the adiabatic curve
Fig. 14. Sound speed c(T, v) in the temperature range of dis- has a behaviour similar to that found in the dissociation
sociation for different values of v. Continuous curves: general zone. Eventually, for extremely strong compressions, the
thermodynamic gas/plasma model (5.2)–(5.3). Dashed curves: curve tends to the first vertical asymptote (v asym /v0 )a = 14
simpler model under classical rotations (7.9)–(7.10). The mean- pertaining to an atomic gas. However, it must be noted
ing of the dotted curves is as in Figure 13. that the adiabat tends to the monatomic asymptote from
16 L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications

10 respectively. These adiabats involving a progressive ioniza-


tion toward the fully ionized state tend to the monatomic
asymptote still from the left.
For the last two adiabats starting at the relatively high
temperatures T0 = 18900 K and T0 = 31500 K, the initial
1
ionization is higher i0 = 0.75 and i0 = 0.98 and the curves
have the usual monotonic behaviour, with the graph tend-
ing to the atomic vertical asymptote, this time from the
right, as in any ideal gas.
10−1

31500 K 8 Conclusions
P RH (v)/RH2 Tv

18900 K
We have presented a simple approximate thermodynamic
model for the hydrogen gas including the possibility of
15750 K molecular dissociation into neutral atoms and of their ion-
12600 K ization to form a gas/plasma. The physical model is re-
stricted to the regime of low densities and assumes that
6300 K the system consists of an ideal gas mixture of free non-
10−3 interacting particles. While this assumption is adequate
3150 K in representing molecular dissociation under diluted con-
ditions, it fails when ionization occurs, since the electric
1575 K
interaction between the charges is always present at any
density and temperature. However, for low densities up
10−4 to the moderate density ρ ≈ 1 kg/m3 , the ionized charges
630 K
remain weakly coupled [12, p. 216] and the ideal gas ap-
T0 = 315 K proximation can be accepted.
The proposed model is based on either the very accu-
rate spectrum of the roto-vibrational states of the molecu-
10−5
0 1 1 1
1
lar hydrogen calculated recently by Pachucki and Komasa
8 6 4
v/v0 [13] or the energy levels derived from the Morse rotating
oscillator by Harris and Bertolucci [14]. The latter lev-
Fig. 15. Dimensionless pressure P RH (v)/RH2 Tv along several els are given by a simple closed-form analytical expres-
Rankine–Hugoniot adiabats as a function of the ratio v/v0 for sion, that has been shown to provide an internal partition
shock waves in hydrogen gas/plasma producing dissociation function approximating rather closely that derived from
and ionization: initial states with v0 = 103 m3/kg and with Pachucki and Komasa energy levels, to the point of giv-
different values of the initial temperature T0 . ing virtually indistinguishable equations of state. Thus the
two spectra provide alternative descriptions equally suit-
able for thermodynamic purposes.
the left side, and not from right side, as occurs in a truly In the model, the degree of dissociation and ionization
monatomic ideal gas. α and i are determined by solving a very simple system of
Considering now different adiabats with higher initial two coupled second order equations, with coefficients de-
temperatures T0 = 630 K and T0 = 1575 K, their be- pendent on the thermodynamic state of the gas. The sys-
haviour is very similar to the first one, with three similar tem reduces to a single fourth order equation for i which
local extrema, since the initial dissociation degree of the has been found to have only one physically admissible so-
gas in the two initial conditions is very small, α0 ≈ 0 and lution i ∈ [0, 1] for densities in the range v > 10−5 m3/kg
α0 ≈ 0.22 × 10−4 , respectively. The initial temperature and for T < 106 K; for higher temperatures, a solution i
of the fourth adiabat is T0 = 3150 K with α0 = 0.43 and extremely near to 1 is easily selected as the proper one
i0 = 0 and the three local extremes are still present since among three admissible real values.
the dissociation has still to play an essential part of its The equations of state for energy, entropy and spe-
role, while the distance between the first minimum and cific heat have been given in closed form as analytical
the local maximum is decreasing. expressions which are useful for representing the thermo-
The next adiabat starts at T0 = 6300 K with an ini- dynamical processes encountered in fluid dynamic appli-
tial dissociation almost complete, α0 = 0.998 and a rather cations such as, for example, the Hugoniot adiabats and
small initial ionization i0 ≈ 0.16 × 10−4 . The correspond- the rarefaction waves. This thermodynamic model is very
ing curve has only one local extremum due to the ioniza- crude since it is based on the ideal gas hypotesis and disre-
tion occurring for increasing compression. This behaviour gards completely the Coulomb interactions in the plasma
is shared also by the next two adiabats starting at T0 = state. The model is therefore limited to low densities and
12600 K and T0 = 15750 K, which are characterized by an it excludes also molecular ionization and negative ion for-
appreciable initial ionization of i0 = 0.133 and i0 = 0.44, mation. However the model encompasses the full range
L. Quartapelle, A. Muzzio: A simple thermodynamic model of diluted hydrogen gas/plasma for CFD applications 17

of temperatures, from very low, where both ortho- and 21. W. Grimus, arXiv:1112.3748v1 [physics.hist-ph] 16
para-hydrogen forms of the gas may manifest, through Dec 2001.
the intermediate values where molecular dissociation sets 22. Don Herbison-Evans, Technical Report TR94-487, Basser
in, up to the high temperatures of partial and eventually Department of Computer Science, University of Sydney, Aus-
full ionization. tralia, Updated 31 March 2011.
The values of pressure in the presence of dissociation 23. T. L. Hill, Statistical Mechanics, Principles and Selected
and ionization at equilibrium have been compared with Applications, Dover, New York, 1987.
those calculated by a recent refined model formulated in 24. R. T. Jacobsen, J. W. Leachman, S. G. Penoncello and E.
the framework of the physical picture [11] and of the chem- W. Lemmon, Int. J. Thermophys., 28, 3, (2007) 758–772.
25. J. W. Leachman, R. T. Jacobsen, S. G. Penoncello and
ical picture [28]. The comparison has shown that the pres-
E. W. Lemmon, J. Phys. Chem. Ref. Data, 38, 3, (2009)
sure equations of the two models at the moderate density
721–748.
v = 1 m3/kg are consistent and of comparable accuracy. 26. R. J. Le Roy, S. G. Chapman and F. R. W. McCourt, J.
This confirms the potentialities of the new simple ther- Phys. Chem., 94, (1990) 923–929.
modynamic model for numerical simulations of supersonic 27. M. Capitelli, G. Colonna and A. D’Angola, Fundamen-
and hypersonic flows of the hydrogen at thermodynamic tal Aspects of Plasma Chemical Physics: Thermodynamics,
equilibrium, cf. e.g. [29]. Springer, New York, 2011.
28. A. Y. Poteckhin, Phys. Plasmas, 3, (1996) 4156–4165.
Acknowledgement. The authors are grateful to Laura Dalzini 29. M. MacLean, A. Dufrene, T. Wadhams and M. Holden,
for the assistance in the reproduction of some graphical results. AIAA Paper, (2010) 2010–1562.

References
1. H. B. Callen, Thermodynamics and an Introduction to Ther-
mostatistics, John Wiley & Sons, New York, 1988.
2. R. Kubo, Statistical Mechanics, North-Holland Publishing
Company, Amsterdam, 1965.
3. J. Lighthill, J. Fluid Mechanics, 2, 1, (1957) 1–32.
4. W. G. Vincenti and C. H. Kruger, Introduction to Physical
Gas Dynamics, John Wiley, New York, 1965.
5. P. M. Morse, Phys. Rev., 34, (1929) 57–64.
6. F. J. Gordillo-Vázquez and J. A. Kunc, J. Appl. Phys., 84,
9, (1998) 4693–4703.
7. L. D. Landau and E. M. Lifshitz, Statistical Physics, Part.
1, Elsevier, Oxford, 1980.
8. D. G. Hummer and D. Mihalas, Astrophysical J., 331,
(1988) 794–814.
9. D. Mihalas, W. Däppen and D. G. Hummer, Astrophysical
J., 331, (1988) 815–825.
10. A. Alastuey, V. Ballenegger, F. Cornu and Ph.A. Martin,
J. Stat. Phys., 130, (2008) 1119–1176.
11. A. Alastuey and V. Ballenegger, Contrib. Plasma Phys.,
52, (2012) 95–99.
12. Ya. B. Zel’dovich and Yu. P. Raizer, Physics of Shock
Waves and High-Temperature Hydrodynamic Phenomena,
Academic Press, New York, 1967.
13. K. Pachucki and J. Komasa, J. Chem. Phys., 130, 164113
(2009).
14. D. C. Harris and M. D. Bertolucci, Symmetry and Spec-
troscopy, Dover, New York, 1989.
15. E. Fermi, Zs. Phys., 26, (1924) 54–56.
16. Y. Babou, Ph. Rivière, M.-Y. Perrin and A. Soufiani, Int.
J. Thermphys., 30, (2009) 416–438.
17. G. A. Blake, www.gps.caltech.edu/~ gab/ch21b/lectures/
lecture07.pdf, Lecture # 7, Vibration-Rotation Spectra of
Diatomic Molecules (2009).
18. G. V. Yukhnevich, Doklady Physics, 45, 5, (2000) 201–204.
19. J. M. L. Martin, J. P. Francois and R. Gijbels, J. Chem.
Phys., 96, (1992) 7633–7645.
20. G. Emanuel, Advanced Classical Thermodynamics, AIAA
Education Series, Washington, D.C., 1987.

Das könnte Ihnen auch gefallen