You are on page 1of 29

Plasma Chemistry and Plasma Processing, Vol. 15, No.

2, 1995

Transport Coefficients of Air, Argon-Air,


Nitrogen-Air, and Oxygen-Air Plasmas
A. B. M u r p h y m

Received May 9, 1994; revised September 5, 1994

Calculated values of the viscosity, thermal conductivity, and electrical conductivity


of air and mixtures of air and argon, air and nitrogen, and air and oxygen at high
temperatures are presented, In addition, combined ordinary, pressure, and thermal
diffusion coeJficients are given for the gas mixtures. The calculations, whicb assume
local thermodynamic equilibrium, are performed for atmospheric pressure plasmas
in the temperature range from 300 to 30,000 K. The results for air plasmas are
compared with those of published theoretical and experimental studies. Significant
discrepancies are found with the other theoretical studies; these are attributed to
differences hi the collision integrals used in calculatbzg the transport coefficients. A
number of the collision integrals used here are significantly more accurate than
values used previously, resulting in more reliable values of the transport coefficients.

KEY WORDS: Transport coefficients; transport properties; viscosity; thermal


conductivity; electrical conductivity; diffusion coefficient; Chapman-Enskog
method; air; argon; nitrogen; oxygen; plasma.

1. INTRODUCTION
The transport coefficients of high-temperature gases and gas mixtures
are important inputs in numerical simulations of plasma phenomena. The
reliability of the simulations is dependent on accurate values for the transport
coefficients being used. Unfortunately, transport coefficients, such as viscos-
ity, thermal conductivity, electrical conductivity, and diffusion coefficients,
are difficult to calculate, and large uncertainties remain in their values. This
is mainly due to uncertainties in the values of the intermolecular potentials
from which the transport coefficients are derived, although uncertainties
in the gas composition and approximations made in the calculations also
contribute.
In a previous paper, (I) transport coefficients of argon, nitrogen, oxygen,
mixtures of argon and nitrogen, and mixtures of argon and oxygen were

tCSIRO Division of Applied Physics, P.O. Box 218, LindfieldN S W 2070, Australia.

0272.4324/95/0600-0279s07.50/0© 1995Plenum PublishingCorporation


280 Murphy

calculated, utilizing recently published intermolecular potential data for a


number of important interactions to significantly improve the reliability of
the transport coefficients compared to previously published values. In this
paper, the calculations are extended to include air and mixtures of air and
the other gases. Many important thermal plasma processes involve air or a
mixture of a pure gas and air. For example, air is used as the plasma gas
in many plasma cutting and mineral processing applications. In addition,
air will be entrained into the plasma gas in any plasma process performed
in the open atmosphere. This entrainment has been shown to be particularly
rapid in cases where the plasma is turbulent, such as in plasma torch
plumes ~2~). Since argon and nitrogen are widely used as plasma gases in
plasma spraying, and oxygen is often used in plasma cutting, mixtures of
these gases with air are frequently encountered; however, to the author's
knowledge, no transport data have been published for such mixtures.
The results are presented for a pressure of 101.3 kPa and for the tem-
perature range 300-30,000 K, conditions which are applicable to most ther-
mal plasma processes. It is assumed that the plasmas are in local
thermodynamic equilibrium (LTE), since this simplifies the calculations, and
since most computational modeling of atmospheric-pressure plasmas relies
on the same assumption.
We present values of viscosity, thermal conductivity, and electrical con-
ductivity for both the pure gases and the gas mixtures. In addition, for the
gas mixtures, combined ordinary, pressure, and thermal diffusion coefficients
are presented. These three combined diffusion coefficients, which have
recently been defined, t5"6) fully describe the diffusion of one gas relative to
another due to concentration, pressure, and temperature gradients,
respectively.
Recently published intermolecular potential data for a number of the
important interactions.have been used. This has significantly improved the
reliability of the calculated transport coefficients of air compared to previ-
ously published values, which have relied on inaccurate collision data for
some important interactions.
In Section 2, the method of calculation of the transport coefficients is
discussed. The choice of the intermolecular potentials and other data on
which the calculation of the transport coefficients is based is discussed in
Section 3. The results are presented and, in the case of air, compared with
values of the transport coefficients from other sources, in Section 4. Conclu-
sions are presented in Section 5.

2. METHOD OF CALCULATION
A discussion of the method of calculation of transport coefficients was
given in the previous paper, c~) Hence, only details that are specific to the
case of air will be presented here.
Transport Coefficients of Air Plasmas 281

Air is taken to consist of a mixture of 78.09% nitrogen, 20.94% oxygen,


0.93% argon, and 0.033% carbon dioxide by volume, tT) Water vapor and
trace species such as neon and helium are neglected. Species concentrations
are calculated by minimization of Gibbs free energy; details are given in
Ref. 1. The nitrogen species considered in the calculations are N2, N~, N,
N-, N ÷, N ÷+, N +÷+', the oxygen species considered are 03, 02, 0 2 , O +2,O,
O-, O +, O ++, O+++; the argon species considered are At, Ar +, Ar ++, Ar ÷++ ;
the carbon species are C(s), C, C-, C ÷, C +÷, C2, C~-, C3, C4, C5 ; and the
mixed species considered are NO, NO +, NO2, NO~-, NO3, NO~,
N20, N20 ÷, CN, CO, CO +, CO2, CO~-, C20, C302; the electron is also
considered.
The transport coefficients are calculated using the well-known Chap-
man-Enskog method, ts-'°) Collisions involving the following species are con-
sidered: the electron; all the nitrogen, o~ygen, and argon species listed above,
with the exception of N-, 03, and O~ ; and the carbon and mixed species
C, C +, C ++, CO, CO2, NO, NO +, NO2, and N20. Details of the method of
calculation of viscosity, thermal conductivity, and electrical conductivity
were given by Murphy. °~ The combined diffusion coefficients are defined by
an expression for the number flux of gas A relative to the mass-average
velocity~5):

__ n 2 __ __ ors
gA = - - m n ( D ~ s V x n + D ~ s V In P ) - _ _ V in T (1)
/9 ma
Here D ~ n , D ~ n , and D ~ s are, respectively, the combined ordinary, pressure,
and thermal diffusion coefficients, mA and m s are the number-density-
w__eightedaverage mass of the species present in gas A and gas B, respectively,
x n is the sum of the mole fractions of the species of gas B, P is the total
pressure, and T is the temperature. The combined diffusion coefficients of
gas A through gas B are given by~5~

~,A. =- s,
~
m,O~ ~
Oxj (2)
m s ~ = 2 j=, dxs

mn~=2 j=, \ p oF/

and
.. {pro

Here D~ and D r° are, respectively, the ordinary and thermal diffusion


coefficients, calculated taking into account ambipolar effects, n and p are,
7.82 Murphy

respectively, the number and mass densities of the gas mixture, and nj, pj,
xj, and mj are, respectively, the number density, mass density, mole fraction,
and mass of the jth species. Species 1 is the electron, species 2 to p belong
to gas A, and species p + 1 to q belong to gas B. The sa are stoichiometric
coefficients, equal to the number of atoms contained in a molecule of species
i, normalized to the number-density-weighted average number of atoms con-
tained in all the species of the gas of which species i is a component. The
combined diffusion coefficients obey the relations D~a = D[a, Daea= -DBeA
and DrB = - D r .
It is important to note that the application of the transport coefficients
calculated for air and for mixtures of air with argon, nitrogen, or oxygen
requires the assumption that the composition of the air component, in terms
of the relative mass fractions of the constituent elements N, O, Ar, and C,
is constant in space and time. In real gas mixtures, this assumption is not
completely justified, since some species present in air diffuse more quickly
through other gases than other species; for example, atomic oxygen generally
diffuses more rapidly than the CO2 molecule. However, measurements°)
have demonstrated that, at least in turbulent plasmas, this demixing effect
is small, and that the air composition is essentially constant. This is because
the turbulent mixing processes dominate the diffusive demixing effects. Since
most applications in which mixtures of air and other gases occur involve
turbulent plasmas, and since numerical modeling of plasma processes has
not advanced to the stage in which the components of air are treated separ-
ately, the presentation of transport coefficients for air and mixtures of air
and other gases is justified. The size of the inaccuracy introduced by any
variations in the air composition that do occur can be estimated from the
data presented in Sections 4.2 to 4.4 for different mixtures of air with argon,
nitrogen, and oxygen.
In this context, it should be noted that the combined diffusion
coefficients are, strictly speaking, only defined in mixtures of two homo-
nuclear nonreacting gases (an example of a gas under the definition used
here is nitrogen, including the species N2, N, N ÷, electrons derived from
ionization, etc.). Air is clearly not a homonuclear gas, and nitrogen and
oxygen can react to form nitric oxide, for example. However, if two
assumptions are made, combined diffusion coefficients can be calculated.
The first is the same assumption that is made in the calculation'of the other
transport coefficients, i.e., that the composition of the air component in
terms of the mass fractions of the constituent elements is constant. The
second is to treat air and the other gas as separate, nonreacting gases. This
has two consequences; the first is that reactions between the air and the
other gas cannot be considered; for example, in the case of nitrogen diffusing
through air, the nitrogen is excluded from reacting with the air to form oxides
Transport Coefl~dents of Air Plasmas 283

of nitrogen (although the oxides of nitrogen present in the air component are
still taken into consideration). The second consequence arises because a
species that is present in both air and the other gas is considered to be a
separate species in each gas; this has the effect of increasing the calculated
degree of dissociation of a molecular species present in both gases. The
reason is that the partial pressures of the atomic species derived from the
dissociation, being divided among the two gases, are lower, so the equilib-
rium degree of dissociation is changed.
The inaccuracies introduced by these consequences are significant
around the dissociation temperature of oxygen (3500 K) for mixtures of air
and oxygen, the dissociation temperature of nitrogen (7000 K) for mixtures
of air and nitrogen, and the temperature range in which nitric oxide is
produced in greatest quantities (also 3500 K) for both of these mixtures.
The error in the combined ordinary diffusion coefficient is always small,
typically less than 5%, and always less than 10%. However, because the
combined pressure and temperature diffusion coefficients are sums of both
positive and negative terms, they can be very sensitive to minor changes in
composition in conditions under which there is no dominant term. Although
the relative error can be large in such cases, the absolute error is still small,
since the diffusion coefficients are then small.
As an example, the upper limit to the inaccuracy introduced by the
assumption that air and the other gases do not react can be estimated
by calculating the diffusion coefficients for two different mole fractions
of NO in a mixture of air and oxygen. In the first case, which is the
standard case, only the NO present in the air component is considered;
in the second case, the mole fraction of NO is calculated taking into
account reactions between the air and the oxygen. In the second case it
is necessary in order to calculate combined diffusion coefficients to assign
all the NO, although some is formed by reactions between the two gases,
to either the air or the oxygen gas; for reasons of computational convenience,
air was chosen here. The calculation was performed for the conditions for
which the equilibrium NO mole fraction is largest, i.e., for mixtures of
between 50% air and 75% air by mole fraction with oxygen at a temperature
of 3500 K; the mole fraction of NO is then around 0.065 and 0.067, respec-
tively. It is found that including the NO molecules formed by reaction of
the oxygen and air alters the combined ordinary and thermal diffusion
coefficients by less than 10%. The combined pressure diffusion coefficient is
more strongly affected, undergoing a sign change; however, its absolute
value is very small at this temperature, so pressure diffusion will in any case
be insignificant.
The generally minor inaccuracies that the assumptions introduce are
outweighed in most applications by the simplicity of the combined diffusion
284 Murphy

coefficient formulation, compared to alternative approaches that require a


separate equation to describe diffusion of each species.

3. COLLISION INTEGRALS
Calculation of the transport coefficients is based on collision integrals,
which are averages over a Maxwellian distribution of the collision cross-
sections for pairs of species. The collision integrals .,,~ C~.(.t,O , 0--0'(2,2) , 0--,~0,2) ,

f~.c.,,3)
tj for interactions involving all pairs of species i and j, and additionally
the collision integrals f~ff"), f~ff.5), f~?.3,, and t i f f " ) f o r interactions
between electrons and other charged species, are required for the calculations
of the transport coefficients. The collision integrals f~,~t,,) were defined and
a general discussion of the methods used to calculate these integrals was
given in the previous paper. <') Further, details of the intermolecular poten-
tials and other data used in the calculations for interactions between pairs
of argon species, between pairs of nitrogen species, between pairs of oxygen
species, between argon species and nitrogen species, and between argon
species and oxygen species, were presented. In this paper, this set of data is
completed by considering interactions between nitrogen species and oxygen
species; between carbon species and argon, nitrogen, and oxygen species;
and between mixed species and all species. It should be noted that while only
approximate collision data are available for some of the minor interactions
considered here, accurate collision integrals have been used for practically
all of the most important interactions, as discussed previouslytin) and below.
In particular, use of the recent data published by Levin, Stallcop, and
Partridge (~'?'~) for the interactions involving N, O, N +, and O + has allowed
the reliability of transport coefficients calculated for high-temperature
nitrogen, oxygen, and air to be greatly improved.

3.1. Neutral-Neutral Interactions


Table I summarizes the sources of data from which the collision inte-
grals for neutral-neutral interactions have been derived. Accurate data are
available for the most important interactions, in the form of Buckingham-
Corner (6,8,10) intermolecular potentials for the N2-O2 and N2-O
interactions, (14) and collision integrals derived from a combination of ab
initio calculations and fits to experimental data for the N - O interaction. ° I~
Acceptable intermolecular potentials have also been published for many of
the less important interactions, such as N-O2, N-NO, N-CO, N-CO2,
O-NO, O-CO, and CO2-CO2. The Lennard-Jones (12,6) potentials used
for many of the interactions should be accurate at low temperatures, since
the parameters of the potentials are generally calculated to fit properties
Tramport Coefficients of Air Plasmas ?.85

Table i. Data Sources for Neutral-Neutral Interactions


Interaction Method References
N2-O2 Buckingham-Corner (6,8,10) potential 14
N2-O Buckingham-Corner (6,8,10)potential 14
N2-NO Exponential potential 15
N-O2 Exp-6 potential 16
N-O Collision integral tabulation 11
N-NO Exp-6 potential 16
N-CO Exp-6 potential 16
N-COs Exp-6 potential 16
O2-NO Exponential potential 15
O-NO Exp-6 potential 16
O-CO Exp-6 potential 16
NO-NO Lennard-Jones (12,6) potential 17
NO2-NO2 Lennard-Jones (12,6) potential 18
N20-N20 Lennard-Jones (12,6) potential 8, p. I111
C-C Lennard-Jones (12,6) potential~ 19
CO-CO Lennard-Jones (! 2,6) potential 17
CO2-CO~ 14-6-8 potential 20
°Lennard-Jones parameters calculated using empirical expression of Ref.
8, p. 245.

measured at temperatures below 1000 K; it is likely, however, that they


underestimate the strength of the repulsion at high collision energies. The
exponential potentials used for the N2-NO and O2-NO interactions were
calculated °s~ using an averaging procedure.
The collision integrals for those, predominantly minor, neutral-neutral
interactions that do not appear in Table I or in Table I of Ref. 1 are
calculated using the combination rule presented by Svehla and McBride. (2t)

3.2. Ion-Neutral Interactions

Resonant charge exchange is the dominant process in determining the


collision integrals f~ (1~> with 1 odd for interactions of the type X*-X and
X - - X , where X is an atomic or molecular species, particularly at high energ-
ies. At lower energies, and for interactions of the type X * - Y , where X and Y
are different atomic or molecular species, elastic interactions are important.
Collision integrals with 1 even are wholly determined by the elastic inter-
action. Table II summarizes the sources of data for both inelastic and elastic
ion-neutral interactions.
The required collision integrals f~o,~, f1,,2~, f~o.3j, and f~(2,2jwere calcu-
lated using a number of methods. The collision integrals tabulated by Par-
tridge et al. ('3~ were used directly for the N+-O and O+-N interactions; as
286 Murphy

Table Ii. Data Sources for Ion-Neutral Interactions


Interaction Elastic collisions Inelastic collisions
Method Reference Method Reference
N*-O Collision integral 13 Collision integral 13
tabulation tabulation
N~-O Polarization potential Charge-exchange 22
O÷-N Collision integral 13 Collision integral 13
tabulation tabulation
NO+-NO Polarization potential Charge-exchange 23
C*-C Polarization potential Charge-exchange 24

was the case for the N - O interaction, these values were derived from a
combination of ab initio calculations and experimental data.
For all other pairs of species considered here, the elastic interaction was
calculated using the polarization potential. Use of this potential has been
shown to lead to overestimates of the collision integralst25~; however, since
only minor interactions are involved, this is expected to lead to only very
small inaccuracies in the calculated transport coefficients.
For most o f the interactions of the type X+-Y, t h e / = ! collision inte-
grals were equated to those derived from the elastic interaction. However,
for interactions o f the type X ~ - X , and also for the N~-O interaction, both
the inelastic and elastic interactions were taken into account in calculating
the 1= ! collision integrals, using the expression
f~ ("~ = [(n,(. ',~ )~ + (n~, ''~ )~]'/~ (5)
where the subscripts in and el denote the collision integrals derived from the
inelastic and the elastic interactions respectively. Data for the charge-
exchange cross-section Qex were used to calculate O~.s~ in these cases. Pub-
lished values of Qex,C2~-24~and the equation ~26~
OtiS(g) =2Qex (6)

were used to numerically evaluate the momentum-transfer cross-section


Q(1~(g), from which f2i~I"~ may be derived, tl~

3.3. Electron-Neutral Interactions


The 1= 1 collision integrals for electron-neutral interactions were calcu-
lated by numerical integration of momentum-transfer cross-section data. It
was assumed that n(2.2)_n(~,u~
"°;~ - " ; s . The momentum-transfer cross-section
values given by Robinson and Geitmann (27~ were used for the e-C inter-
action. The values recommended by Itikawa (28~were used for the e - N O and
Transport Coefficients of Air Plasmas 287

e-N20 interactions, supplemented in the latter case by the experimental data


of Johnstone and Newell t29) for high-energy collisions. The measured data of
Land <3°Jand Lowke e t al. ~3~ were used for the e-CO and e-CO2 interactions
respectively. The total cross-section measurements of Szmytkowski e t al. ~32>
were used in place of momentum-transfer cross-section data for e-NO2 inter-
actions, since no measurements of the latter were available. The sources of
the momentum-transfer cross-section data for interactions of electrons with
the other neutral species were given in Ref. 1.

3.4. Charged-Charged Interactions

Collision integrals for interactions between charged species were calcu-


lated using the screened Coulomb potential, t~3J the screening distance being
set equal to the Debye radius.

4. RESULTS
4.1. Air

I now present results of calculations of the transport properties of pure


air, performed using the methods described above.
Figure 1 compares the calculated temperature dependence of the viscos-
ity of air with the results of other authors. Reasonable agreement is seen
with the calculations of Bacri and Raffanel <s4)at temperatures below 7000 K
and above 20,000 K. At intermediate temperatures, the values of viscosity
calculated by Bacri and Raffanei are significantly higher. The intermolecular
potentials used by Bacri and Raffanei for the most important interactions,
namely N-N, N+-N, O-O, and O+-O, are those suggested by Capitelli and
Devoto <3~ for the nitrogen interactions and Aubreton e t al. ~36) for the oxygen
interactions; generally good agreement is found between the values of viscos-
ity of nitrogen and oxygen calculated by these authors and the respective
values calculated °~ using the same potentials that are used here. However,
the potentials that Bacri and Raffanel used for the most important nitrogen-
oxygen interactions, namely N-O, N+-O, and O+-N, are inaccurate. Indeed,
Bacri and Raffanel demonstrated that replacing the rigid-sphere potential
that they used for the N - O interaction by a more accurate Morse potential
decreases the calculated viscosity by up to 50%. The polarization potential
they used for the N÷-O and O÷-N interactions generally gives lower collision
integrals, and hence higher viscosity values than more accurate potentials,
which would compensate the influence of the N - O potential to some extent.
Further, Bacri and Raffanei performed their calculations for air without
carbon dioxide. This, too, leads to significant inaccuracies in the calculated
288 Murphy

0.00040

/'N
0.00030 / \
I
/ \
!
/ c~
E
~o.ooo2o

~ 0.00010

(( \'~....o o o o o

0.00000
0 10000 20000 30000
temperature (K)
Fig. I. Viscosity o f air: - - this work; - - - Bacri and Raffanel °4~; • • - Elchinger eta/. ~37J8~;
O Yos.°9~

0.00030

T 0.00020
T
E

o
.~ 0.00010

0.00000
0 10000 20000 30000
temperature (K)
Fig. 2. Viscosity of: • air; • • • air without carbon dioxide.
Transport Coet~icients of Air Plasmas 289

viscosity. This is illustrated in Fig. 2, in which data for air are compared to
values calculated for air without carbon dioxide. Differences of up to 15%
are found. The omission of carbon dioxide has only negligible effects on the
thermal conductivity and electrical conductivity.
Better agreement is found with the values of Elchinger et al., t37) whose
results have been tabulated by Boulos et al. t3s~ Elchinger et al. omit the
influence of the CO2 component of air from their calculations; comparison
of Figs. 1 and 2 shows that their results agree closely with the present
calculations for air without CO2 at temperatures up to 10,000 K. At higher
temperatures, the results of Elchinger et al. overestimate the viscosity, sug-
gesting that their collision integrals for ion-neutral collisions are too small.
Fair agreement is found between the present results and the viscosities
calculated by Y o s (39) at temperatures up to 15,000 K. At higher tempera-
tures, larger discrepancies are seen. This is expected, since at these tempera-
tures, interactions between charged particles dominate, and Yos's treatment
of such interactions is known to be inaccurate: 34'4°)
In Fig. 3, the calculated thermal conductivity values are compared with
the results of other workers. As with the viscosity, significant differences are
found from the values of Bacri and Raffanel. t34~ The main discrepancies
occur at temperatures around 7000 K and 15,000 K, which correspond
respectively to the dissociation temperature of nitrogen, and the ionization
temperature of both nitrogen and oxygen. At these temperatures, the largest
component of the thermal conductivity is the reaction thermal conductivity,
as is shown in Fig. 4. In a pure gas, the reaction thermal conductivity
associated with the dissociation of X2 molecules, where X is an atomic
species, is approximately inversely proportional to the collision integral
f~o.t)
X,X ; similarly, the reaction thermal conductivity associated with the ioniz-
ation of X atoms is approximately inversely proportional to the collision
integral ncq~
~*x .x. Murphy and Arundeli t'~ showed that the N÷-N potential
used by Bacri and Raffanel, that of Capitelli and Devoto, ~35~gives values
of f ~ . that are much smaller than the more reliable values of Stallcop et
al., ~2) and thus leads to major overestimates of the thermal conductivity at
temperatures around 15,000 K. However, the values of ~ calculated using
the N - N potential of Capitelli and Devoto, also used by Bacri and Raffanel,
are very similar to those given by Levin et al., ~ i) which are used here. Hence,
the different values of thermal conductivity at temperatures around 7000 K
must be attributed to other factors, probably the inaccurate N - O potential
used by Bacri and Raffanel.
Excellent agreement is found with the calculated values of Elchinger et
al. (37'3s) at temperatures up to 10,000 K. This is consistent, in view of the
insensitivity of thermal conductivity calculations to the omission of carbon
dioxide from the air, with the good agreement for this temperature range
290 Murphy

6
l'X\ ~.
.-.5
//
"r
5; ./y o
E 4
j¢o
•~ 3

82

i
0

0 I i I i I
0 10000 20000 30000
temperature (K)
F i g . 3. T h e r m a l conductivity of air: - - this work: - - - Bacri and Raffanel°4~;. • • Elchinger
et al.°7"~s~; 0 YOSt391; ~ Devoto et al.t4m; [] Asinovsky et al. t4u~

that has already been noted between their viscosity data and the present
viscosity data for air without carbon dioxide. At higher temperatures, how-
ever, Elchinger et al. greatly overestimate the thermal conductivity; this may
be due to their values of •nt~
,N'~oN being too small.
Reasonable agreement is found with the calculated thermal conductivity
values of Yost39); as with the viscosity, the main differences are at high
temperatures. Generally good agreement is found with the experimental
values of Devoto et al. ~4°~ and Asinovsky et aL ~4~ A t temperatures above
10,000 K, the values of Devoto et al. are too large; this is because they
neglected the effects of radiation in calculating the thermal conductivity from
the experimental data.
The calculated values of the electrical conductivity of air, shown in Fig.
5, agree very closely with most of the calculations and the measurements of
other workers. Excellent agreement is found with the measurements of
Devoto et al. t4°) and Asinovsky et aL, t4° and with the calculations of Bacri
and Raffanelt34) and Devoto. t42) The measurements of Schreiber et al. t43~
tend to underestimate the electrical conductivity, as do the calculations of
Transport Coefficients of Air Plasmas 291

.-,5
'T

,_z,
-
.
/'/,.~
//~ ii

>3

I
b ; il
: g.

'.,

0 I . . ¢ z . . L - ' : ~ - ~ / - 7 - - ---. " - _ - . ~ . _ _ _ ;:'_'--..... ._=--_-';:",


0 10000 20000 30000
temperaiure (K)
Fig. 4. Components of the thermal conductivity of air: - - total; . . . . . heavy particle;
- - - electron; -- - internal; • • • reaction thermal conductivity.

at high temperatures. Fair agreement is found with the calculations


Y o s t39)
of Elchinger eta/. t37'38) at temperatures up to 15,000 K; at higher tempera-
tures, the values of Elchinger et al. are significantly larger than all other
values. This indicates that their Coulomb cross-sections for electron-ion
interactions, which are dominant at high temperatures, are smaller than
those used by other workers.

4.2. Argon-Air Mixtures


The temperature dependence of the viscosity of mixtures of argon and
air in three different proportions is compared in Fig. 6 with that of pure
argon and pure air. It is interesting to note that in the temperature range
6 6 0 0 - 9 0 0 0 K, shown in detail in Fig. 6b, the mixtures have lower viscosities
than either air or argon, and between 12,600 and 19,200 K, shown in detail
in Fig. 6c, the mixtures have viscosities higher than either of the pure gases.
292 Murphy

4 0 0 0 e° ""*

Z"
//oOoo0

+'t+ io
fo °

" 2

0 10000 20000 30000


temperature (K)
Fig. 5. E]ectrical c o n d u c t i v i t y o f a i r : - this w o r k ; - - - Bacri and Raffanel<341; • • • Elchinger
et aLI37"3s); 0 Yos°9); x Devoto(42); A Devoto et al.(4°); [] Asinovsky et al.(4u; <) Schreiber
et aL <4~

Similar behavior was observed for argon-nitrogen and, in the higher tem-
perature range, argon-oxygen mixtures. (t) The effect in the higher tempera-
ture range is mainly a result of the t~(2'2) collision integrals, with which the
viscosity varies inversely, for the X+-Y (e.g., Ar+-N) interactions being
lower than those for the X+-X (e.g., Ar+-Ar or N+-N) interactions. In the
6600-9000 K range, the viscosity of the argon-air mixture is lower than that
of air because the collision integrals for the Ar-N interaction are much
higher than those for the N - N interaction. The viscosity of the mixture is
lower than that of pure argon, despite the fact that the Ar-N collision
integrals are slightly lower than those for the Ar-Ar interaction, because of
the lower reduced mass of the species involved in the mixed interaction; the
viscosity increases with the reduced mass of the interacting species.
The thermal conductivity of argon-air mixtures is shown in Fig. 7. At
temperatures around 15,000 K, for which the reaction thermal conductivity
associated with ionization reactions makes a major contribution to the total
thermal conductivity, the thermal conductivity of most mixtures is greater
than that of air or argon. This effect occurs because the f~(t,~) collision
Transport Coefficients of Air Plasmas 293

0.00030 -(a)

~" 0.00020

==

i0.00010

0.00000
0 10000 20000 30000
temperature (K)

0.00025
"(b) ..

0.00023
. , " " ' / ' / S
I

E
=~0.00021

> 0.00019

0.00017 I = = • =
6000 7000 8000 9000 10000
temperature (K)
]Fig. 6. Viscosity of mixtures of air and argon: - - 100% air; - - - 75% air, 25% argon;
- - - 50% air, 50% argon; - - - 25% air, 75% argon; • • - 100% argon. Percentages refer to mole
fractions. (a) Full temperature range; (b) detail in temperature range 6000 to I0,000 K; (c)
detail in temperature range 12,000 to 20,000 K.
294 Murphy

0.00025
-(c)
0.00020

..~_.'~'~"
"~
~0.00015

-¢@0.00010
o
0

0.00005

0.00000 I ~ I , I ~ I
12000 14000 16000 18000 20000
temperature (K)
Fig. 6. C o n t i n u e d

.-.5

~4 ""
.~_
a f"Y"

° .........
0

° Jr, :!.'.,
o ~ ...... " ' , . , . ,
0 10000 20000 30000
temperature (K)
Fig. 7. Thermal conductivity of mixtures of air and argon. Symbols are as in Fig. 6.
Transport Coefficients of Air Plasmas 295

integrals for the nonresonant (X÷-Y) interactions are very much lower than
those for the resonant (X+-X) interactions.
The electrical conductivity, shown in Fig. 8, is practically independent
of composition at temperatures below 19,000 K for all argon-air mixtures.
The difference at higher temperatures can be attributed, noting that the
collision integrals for the e-X ~n+J)+ interaction are greater than those for
the e-X "+ interaction, to the greater proportion of multiply ionized species
in argon, compared to nitrogen and oxygen, at these temperatures.
Figure 9 shows the temperature dependence of the combined ordinary
diffusion coefficient for different mixtures of argon and air. For reasons
discussed by Murphy and Arundeli, °) the only dependence of the combined
ordinary diffusion coefficient on the relative concentrations of the two gases
is that due to the effect that these concentrations have on the degree of
ionization and dissociation of the gases. Thus, at temperatures below 3000 K,
for which significant dissociation of nitrogen and oxygen does not occur,
the combined ordinary diffusion coefficient is independent of the relative
concentration of the gases. At higher temperatures, particularly at those at
which ionization occurs, a dependence does exist, since the proportion of
Ar, N, and O atoms ionized, and to a lesser degree, the proportion of 02
and N2 molecules dissociated, varies with the relative concentration of argon
and air.
In contrast, the combined pressure and combined thermal diffusion
coefficients, shown in Figs. 10 and 11, respectively, exhibit a strong depend-
ence on fractional composition at all temperatures. This dependence was
also discussed in detail by Murphy and Arundell. It can be seen from Figs.
10 and 11 that both the combined pressure and combined thermal diffusion
coefficients are largest when the two gases are present in reasonably large
concentrations; when only a trace concentration of one gas is present, the
two coefficients are very small.

4.3. Nitrogen-Air Mixtures


The temperature dependence of the viscosity of three different mixtures
of nitrogen and air is plotted in Fig. 12, together with values for pure
nitrogen and pure air. As is expected, given the large proportion of nitrogen
in air, all the viscosities are similar. The viscosities of the mixtures lie between
those for nitrogen and air.
The thermal conductivity of nitrogen-air mixtures, shown in Fig. 13,
also lies between the values for nitrogen and air. The thermal conductivity
of nitrogen does not exhibit the peak associated with the dissociation of
oxygen at 3500 K; similarly, air and nitrogen-air mixtures have a lower
296 Murphy

14000-

12000

10000
Or)

.~' 8000
15
-o
t- 6000
O
0
4000

E 2000

0
0 10000 20000 30000
temperature (K)
Fig. 8. Electrical c o n d u c t i v i t y of m i x t u r e s o f air a n d argon. S y m b o l s are as in Fig. 6.

0.014

0.012

0.01

j 0.008

g° 0.006

~ 0.004

,- 0.002

10000 20000 30000


temperature (K)
Fig. 9. C o m b i n e d o r d i n a r y diffusion coefficient D~,.,~, o f m i x t u r e s o f air a n d a r g o n : 99%
0 0
air, ! % a r g o n ; - - - 75% air, 25% a r g o n ; - • - 50% air, 50% a r g o n ; - - - 25'/0 air, 75'A argon;
• • • 1% air, 99% argon. Percentages refer to mole fractions•
Transport Coellidents of Air Plasmas

0.004-

A
I

;\

/ f,',
i/\X
0.002
/i,'\\
~,i// ~\,,\
c
..9o

"o 0.001 ~,

Q.
//
_.,,,,." . ~ ....... , ...... ~,
0
0 10000 20000 30000
temperature (K)
Fig. 10. Combined pressure diffusion coefficient D~r.air Of mixtures of air and argon. Symbols
are as in Fig. 9.
0.2-

?,,
~o,~ ii
0.1

!~ I: iX
~o.o, t~ // ~,

\~/
"" -0.05 , I s I , I
0 10000 20000 30000
temperature (K)
Flg. 11. Combined thermal diffusion coefficient D~rJir Of mixtures of air and argon. Symbols
are as in Fig. 9.
298 Murphy

0.00030 I

~7"~"
0.00020

¢
~ 0.00010

0.00000
0 10000 20000 30000
temperature(K)
Fig. 12. Viscosityof mixtures of air and nitrogen: 100%air; - - - 75% air, 25% nitrogen;
- • - 50% air, 50% nitrogen; - - - 25% air, 75% nitrogen;. • • 100%nitrogen. Percentagesrefer
to mole fractions.

thermal conductivity at 7000 K, the temperature at which nitrogen dissoci-


ates. The thermal conductivity of air and the nitrogen-air mixtures is greater
than that of nitrogen at temperatures around 14,000 K, at which N and O
atoms ionize; this is because the reaction thermal conductivity of oxygen is
larger than that of nitrogen at this temperatureJ '~
The values of the electrical conductivity of nitrogen, air, and nitrogen-
air mixtures are almost equal at temperatures between 8000 and 22,000 K,
as shown in Fig. 14. The electrical conductivity of air is slightly higher at
lower and higher temperatures. The effect at low temperatures is due to the
electron-neutral momentum transfer cross-section for nitrogen molecules,
which are present in significant quantities up to 8000 K, being much larger
than those for oxygen molecules and atoms. The effect at high temperatures
is due to the greater proportion of multiply ionized species in oxygen.
The combined ordinary diffusion coefficients of the nitrogen-air mix-
tures, shown in Fig. 15, exhibit similar behavior to those of the argon-air
mixtures. In contrast, the combined pressure diffusion coefficients, shown in
Fig. 16, are an order of magnitude smaller in absolute value than those in
the argon-air mixtures. This is because the terms in the factor xj-pj/p in
Traaspert Coefficients of Air Plasmas 299

6,-

II

---5

T
E 4

•> 3
U
"0
t-
O
~2

ID
,--1

0 ~
0 10000 20000 30000
t e m p e r a t u r e (K)

Fig. 13. Thermal conductivity of mixtures of air and nitrogen. Symbols are as in Fig. 12.

the expression for DeAn[Eq. (3)], which, in most cases, "~ including that of
the argon-air mixture, are much larger than the terms in the factor P OxJOP,
are in this case small. This is because most of the important species in the
nitrogen-air mixture have similar masses, so that pj/p.~xi.
The magnitude of the combined thermal diffusion coefficients of the
nitrogen-air mixtures, given in Fig. 17, is somewhat smaller than those of
the argon-air mixtures. In both eases the term in TOxj/OT [see Eq. (4)]
dominates at most temperatures.

4.4. Oxygen-Air Mixtures


The values of viscosity of three different mixtures of oxygen and air are
plotted in Fig. 18, together with those of pure oxygen and of pure air. Over
most of the temperature range shown, oxygen has a significantly higher
viscosity than air and the mixtures, whose viscosities lie between those of
oxygen and air except at temperatures between 12,000 and 20,000 K.
The thermal conductivity of oxygen-air mixtures, shown in Fig. 19, lies
between the values for oxygen and air, except for temperatures between
3oo M=r~y

14000

12000

7"E 10000

~ 8000

6ooo
¢J
4o0o
"1,.

20OO"

0
0 10000 20O00 30O00
temperature (K)
Fig. 14. Electrical conductivity of mixtures of air and nitrogen. Symbols are as in Fig. 12.
0.018

0.016 f
0.014

0.012 I

~. 0.01

~ 0.008
"~ 0.OO6

'- 0.002
/ "

0 i I a
0 10000 20000 30000
temperature (K)
Fig. 15. Combined ordinary diffusion coemcient 17~2.,i, o f mixtures o f air and nitrogen: - -
99% air, 1% nitrogen; - - - 75% air, 25% nitrogen; - - . - - 50o/o air, 50% nitrogen; - - - 25%
air, 75% nitrogen; - • - I % air, 99% nitrogen. Percentages refer to mole fractions.
Transport Coefliclenls of Air Plasmas 301

0.0003

0.00O2
A
E 0.0001
c
Q
•~. 0 _ L.\l5~--~-1__ . .......

o
•- -0.0001 I i', /#~:~'--'='-'--'~"--'~
.2
~
"o
-0.0002
.,\ //v"
\ \ ~ "J!
/, 7
-0.0003
/
\.j°

-0.0004 , I = I = I
0 10000 20000 30000
temperature (K)
Fig. 16. Combined pressure diffusion coefficient DN2.air
P of mixtures of air and nitrogen. Symbols
are as in Fig. 15.
0.08 ,-

0.06

0.04

== 0.02

8 o

8
"~--0.02
~ -0.04
" --0.08 i I i I a I
0 10000 20000 30000
temperature (K)
Fig. 17. Combined thermal diffusion coefficient D~2.,~r of mixtures of air and nitrogen. Symbols
are as in Fig. 15.
302 Murphy

0.00030 -

o," ",
(:'~"

o.ooo o

i O.O0010
0.00000 , i , i , l
0 10000 20000 30000
temperature ( K )
Fig. 18. Viscosity o f m i x t u r e s o f a i r and o x y g e n : 100% a i r ; - - - 75% air, 25% o x y g e n ;
--. - - 50% air, 50% o x y g e n ; - - - 2 5 % air, 7 5 % o x y g e n ; • • . 100% oxygen. Percentages refer
to m o l e fractions.

10,000 and 21,000 K. Oxygen does not exhibit the peak associated with the
dissociation of nitrogen at 7000 K, and air and oxygen-air mixtures have a
lower thermal conductivity at 3500 K, the temperature at which oxygen
dissociates. The anomalously large thermal conductivity of the mixtures
between 10,000 and 21,000 K occurs, as for argon-air mixtures, because the
f~('") collision integrals for the nonresonant (X+-Y) interactions are very
much lower than those for the resonant (X+-X) interactions.
The electrical conductivity of oxygen-air mixtures is compared to that
for oxygen and air in Fig. 20. The electrical conductivity at low and high
temperatures increases as the proportion of air decreases, which is the oppo-
site trend to that seen for nitrogen-air mixtures in Fig. 14. The observed
changes may be explained on the basis of the different properties of oxygen
and nitrogen, using the arguments presented in Section 4.3.
The combined ordinary diffusion coefficients of oxygen-air mixtures,
shown in Fig. 21, exhibit similar behavior to those of argon-air and
nitrogen-air mixtures. The combined pressure diffusion coefficients, given
in Fig. 22, are somewhat larger than those of the nitrogen-air mixtures, but
still small compared to those of the argon-air mixtures. This is again because
Transport Coefficients of Air Plasmas 303

6.-

---5

T f.
E 4

•-> 3
"0
t-
¢/
O
~2
t~
E
,-1
000

0 ~
0 10000 20000 30000
t e m p e r a t u r e (K)

Fig. 19. Thermal conductivity of mixtures of air and oxygen. Symbols are as in Fig. 18.

the factors xj-pj/p are generally small. The combined thermal diffusion
coefficients are shown in Fig. 23; again the terms in T axj/dT dominate.

5. CONCLUSIONS
I have presented calculated values of the transport properties of air,
and mixtures of argon, nitrogen, and oxygen with air, in the temperature
range from 300 to 30,000 K. The calculations were performed using standard
approximations based on the Chapman-Enskog method, assuming the exist-
ence of local thermodynamic equilibrium.
Previously published calculations of the transport coefficients of air
agree closely with the values presented in the case of the electrical conductiv-
ity. However, significant discrepancies were found in the viscosity and ther-
mal conductivity values. These have been accounted for mainly in terms of
differences in the collision integrals used for certain interactions; also, the
neglect of the CO2 component ofair in previous calculations led to significant
errors in the viscosity. The collision integrals employed here in many cases
3O4 M~y

14OOO

12000

10OOO
v

_z, 8OOO
.P.
"o 6000- =
c
o
o P

o
4000 L
h.

2000

0
0 10000 20000 30000
temperature (K)
Fig. 20. Electrical conductivity of mixtures of air and oxygen. Symbols are as in Fig. 18.
0.025

~" 0.02
E

0.015
=::

c 0.01

"o
~ 0.005
o ~
c

o
0
0 10000 20000 30000
temperature (K)
Fig. 21. C o m b i n e d o r d i n a r y d i f f u s i o n coefficient D~2.=~, o f m i x t u r e s o f a i r a n d o x y g e n : - -
9 9 % air, I % o x y g e n ; - - - 7 5 % air, 2 5 % o x y g e n ; - - . - - 5 0 % air, 5 0 % o x y g e n ; - - - 2 5 % air,
7 5 % o x y g e n ; • • . 1% air, 9 9 % o x y g e n . P e r c e n t a g e s refer to m o l e f r a c t i o n s .
Transport Coel~dentsof Air Plasmas
0 . 0 0 0 8 ,-

.... 0.0006
F\
0.0004
E
0.0002

-0.0002

-0.0004
e-s
-0.0006 1
I
~i ~ I , I s ..,,. ii
0 10000 20000 30000
temperature (K)
Fig. 22, Combined pressure diffusion coefficient D~:.,~, of mixtures of air and oxygen. Symbols
are as in Fig. 21.
0.08 -
• i~
.-. 0.06 - if
.

~" 0.02-
.
i I/\
. \/ ' ~:\- - .;/
\ .11

I ~1~1... I" ~ I
o 10ooo 2ooo0 zoooo
temperature (K)
Fig. 23. Combined thermal diffusion coefficient Do2~.,,
r of mixtures of air and oxygen. Symbols
are as in Fig. 21.
306 Murphy

represent significant improvements on those previously used; accordingly,


the transport coefficients presented here are expected to be more accurate
than previously calculated values.
No published data for the transport coefficients of mixtures of air and
the other gases could be located. The data presented here for such mixtures
are expected to be useful in the modeling of the interaction of pure gas
plasmas with the surrounding atmosphere. In particular, the combined
diffusion coefficient data will allow the entrainment of air to be modeled
simply and accurately.
Tabulations of the data presented in this paper are available on diskette
on request from the author.

REFERENCES
I. A. B. Murphy and C. J. Arundell, Plasma Chent. Plasma Process. 14, 451 (1994).
2. E. Pfender, J. Fincke, and R. Spores, Plasma Chem. Plasma Process. I1,529 (19913.
3. J. R. Fincke, W. D. Swank, and D. C. Haggard, Plasma Chem. Plasma Process. 13, 579
(19933.
4. A. B. Murphy and P. Kovitya, J. Appl. Phys. 73, 4759 (19933.
5. A. B. Murphy, Phys. Rev. E48, 3594 (19933.
6. A. B. Murphy, J. Chem. Phys. 99, 1340 (19933.
7. L'Air Liquide, Division Scientifique, Gas Enc)'lopaedia, Elsevier, Amsterdam (1976), p. 61.
8. J. O. Hirschfelder, C. F. Curtiss, and R. B. Bird, Molecular Theor), of Gases and Liquids,
2nd edn., Wiley, New York (19643.
9. S. Chapman and T. G. Cowling, The Mathematical Theory of Non-Uniform Gases, 3rd
edn., Cambridge University Press, Cambridge, UK (1970).
10. J. H. Ferziger and H. G. Kaper, Mathematical Theory of Transport Processes in Gases,
North-Holland, Amsterdam (19723.
II. E. Levin, H. Partridge, and J. R. Stallcop, J. Thermophys. Heat Tran.qer 4, 469 (19903.
12. J. R. Stallcop, H. Partridge, and E. Levin, J. Chem. Phys. 95, 6429 (1991).
13. H. Partridge, J. R. Stallcop, and E. Levin, Chem. Phys. Lett. 184, 505 (19913.
14. B. Brunetti, G. Liuti, E. Luzzatti, F. Pirani, and F. Vecchiocattivi, J. Chem. Phys. 74, 6734
(19813.
15. K. S. Yun and E. A. Mason, Phys. Fluids 5, 380 (19623.
16. B. Amaee and W. B. Brown, Chem. Phys. 174, 351 (1993).
17. A. A. Clifford, P. Gray, and N. Platts, J. Chem. Soc. Faraday Trans. 173, 381 (1977).
18. R. S. Brokaw and R. A. Svehla, J. Chem. Phys. 44, 4643 (1966).
19. P. Kovitya, "Theoretical determination of material functions of plasmas formed from
ablated PTFE, alumina, PVC, and Perspex for the temperature range of 5000 to 30,000 K,"
Technical Memorandum No. 3, CSIRO Division of Applied Physics, Sydney, Australia
(1982).
20. H. J. M. Hanley and M. Klein, J, Phys. Chem. 76, 1743 (19723.
21. R.A. Svehla and B. J. McBride, "Fortran IV computer program for calculation of thermo-
dynamic and transport properties of complex chemical systems", Technical Note TN
D°7056, NASA, Washington, DC (19733.
22. J. A. Rutherford and D. A. Vroom, J. Chem. Phys. 61, 2514 (1974).
23. T. F. Moran, M. R. Flannery, and P. C. Cosby, J. Chem. Phys. 61, 1261 (19743.
Transport Coefficients of Air Plasmas 307

24. V. A. Belyaev, B. G. Brezhnev, and E. M. Erastov, Soo. Phys. JETP 27, 924 (1968).
25. M. Capitelli, J. Phys. Colloq. 38, C3-227 (1977).
26. A. Dalgarno, Philos. Trans. R. Soc. London 250, 426 (1958).
27. E. J. Robinson and S. Geltman, Phys. Reo. 153, 153 (1967).
28. Y. Itikawa, At. Data Nucl. Data Tables 21, 69 (1978).
29. W. M. Johnstone and W. R. Newell, J. Phys. B: At. Mol. Opt. Phys. 26, 129 (1993).
30. J. E. Land, J. Appl. Phys. 49, 5716 (1978).
31. J. J. Lowke, A. V. Phelps, and B. W. Irwin, J. Appl. Phys. 44, 4664 (1973).
32. C. Szmytkowski, K. Maci~tg, and A. M. Krzysztofowicz, Chem. Phys. Lett. 190, 141 (1992).
33. E. A. Mason, R. J. Munn, and F. J. Smith, Phys. Fluids 10, 1827 (1967).
34. J. Bacri and S. Raffanel, Plasma Chem. Plasma Process. 9, 133 (1989).
35. M. Capitelli and R. S. Devoto, Phys. Fluids 16, 1835 (1973).
36. J. Aubreton, C. Bonnefoi, and J. M. Mexmain, Reo. Phys. Appl. 21,365 (1986).
37. M. F. Elchinger, B. Pateyron, G. Delluc, and P. Fauchais, in Proceedings of the Ninth
International Symposium on Plasma ChemistiT, Pugnochiuso, Italy, 1989, Vol. I, Inter-
national Union of Pure and Applied Chemistry, Oxford, UK (1989), p. 127.
38. M. I. Boulos, P. Fauchais, and E. Pfender, Thermal Plasmas: Fundamentals and Applica-
tions, Vol. I, Plenum Press, New York (1994), pp. 413-417.
39. J. M. Yos, "Transport properties of nitrogen, hydrogen, oxygen, and air at 30,000 K",
Technical Memorandum RAD-TM-63-7, AVCO Corporation, Wilmington, Massachusetts
(1963).
40. R. S. Devoto, U. H. Bauder, J. Cailleteau, and E. Shires, Phys. Fluids 21,552 (1978).
41. E. 1. Asinovsky, A. V. Kirillin, E. P. Pakhomov, and V. I. Shabashov, Proc. IEEE 59, 592
(1971).
42. R. S. Devoto, Phys. Fluids 19, 22 (1976).
43. P. W. Schreiber, A. M. Hunter, and K. R. Benedetto, AIAA J. 11,815 (1973).

You might also like