Sunteți pe pagina 1din 7

49th AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition AIAA 2011-248

4 - 7 January 2011, Orlando, Florida

Viscous Shock Layer Calculation of Stagnation-Region Heating Environment


in Neptune Aerocapture
Chul Park∗
Korea Advanced Institute of Science and Technology, Daejeon, Korea

The theoretical model describing the nonequilibrium The need to fly through the atmosphere of Neptune
and radiation phenomena in a hydrogen-helium-methane will arise if one hopes to be captured into an orbit around
mixture developed earlier by the present author is applied the planet that passes by Neptune’s moon Triton. This
to carry out a viscous shock layer calculation of the stag- aerocapturing flight will occur at an entry speed of up to
nation region of a blunt body. The calculation is made for about 30 km/s. At such a flight speed, the shock layer
the vehicles performing aerocapturing maneuver in Nep- flow will be partly ionized. Radiation, which is strongly
tune, including those considered earlier by Hollis et al and related to ionization, may be substantial. In order to pre-
by Jits et al. The present heat flux values are substan- dict the radiative heating rate, one must know the thermo-
tially higher than those obtained by the earlier investiga- chemical processes producing electrons. It is well known
tors. The radiative heat flux value is calculated over a that, in general, collisions of electrons with atoms are the
range of flow parameters. most efficient in producing ionization. But the mecha-
nism to produce the first electrons varies considerably by
Nomenclature gas mixtures. For the hydrogen-helium-methane mixture
under consideration, this mechanism is not yet fully un-
A : Reaction rate constant, cm3 /s. derstood.
e : Energy feed back factor.
kf : Forward reaction rate coefficient, cm3 /s. In the 1970s, shock tube experiments were carried out
n : Pre-exponential power in rate coefficient, dim- with a hydrogen-helium mixture, to a shock speed of 18
: ensionless. km/s by Leibowitz [4] with a helium-rich mixture, and to
p : Pressure, Pascal or Torr. 38 km/s by Livingston and Poon [5] with a hydrogen-rich
qr : Radiative heat flux, W/cm2 or kW/cm2 . mixture. The times to ionization equilibration was mea-
Ta : Effective temperature of reaction, K. sured in both experiments. Livingston and Poon mea-
Td : Reaction temperature, K. sured the electron density also. Leibowitz theorized that
Us : Shock speed in a shock tube, m/s. the ionization process is started with collisions of hydro-
ρ : Density, kg/m3 . gen and helium atoms with hydrogen atoms, and deduced
the cross section controlling the process. But these two
sets of results on ionization equilibration time did not
Subscript
agree with each other. Moreover, the electron densities
1 : Freestream. determined by Livingston and Poon were higher than
the equilibrium Rankine-Hugoniot values. More recently,
Introduction Bogdanoff and Park [6] carried out an experiment similar
to that by Leibowitz. Very fast ionization and very large
The chemical kinetics and radiation phenomena occur- electron densities were observed. It was determined that
ring in the shock layer over a body entering the atmo- the driver of the shock tube heated by an electric dis-
spheres of outer planets have been studied in [1] through charge reached a high temperature and emitted a strong
[3]. The atmospheres of all these outer planets are simi- radiation. This radiation radiatively heated the driven
lar: they consist mostly of hydrogen, a small concentra- gas to cause early and high ionization.
tion of helium, and a trace concentration of methane. For
the entry flight into Jupiter, the shock layer flow can be In [1], a thermochemical model is developed to rec-
considered to be in thermochemical equilibrium because oncile these inconsistencies by considering the radiative
of the relatively high stagnation pressures. In the flights heating by the hot driver gas. In [3], the effect of trace
into other outer planets, peak heating is likely to occur at concentration of methane on ionization process was in-
lower stagnation pressures, and therefore, the shock layer vestigated theoretically. By interpreting the published
flow may not be in equilibrium. results of several atomic beam experiments, the rate co-
efficients of the reactions involving carbon and hydrogen
were deduced. Through this procedure, a complete set of
reaction rates controlling the hydrogen-helium-methane
mixture was derived. According to the model, the first
∗ electrons are produced by the collisions of carbon atoms
Visiting Professor, Department of Aerospace Engi-
neering; Fellow, AIAA; cpark216@kaist.ac.kr. and hydrogen atoms. Ref. [3] used this reaction model

Copyright © 2011 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
2

to predict the radiative heat transfer rates occurring in Method of Calculation


the entry flights into the planet Neptune. The conditions
Reaction Kinetics of H2 +He+CH4 Mixtures
considered were those by Hollis et al [7] and those by Jits
et al [8]. The stagnation streamline flow was considered Neptune’s atmosphere [5] is believed to consist of, by
to be an inviscid constant-area one-dimensional flow as in mol, about 80 to 82% H2 , 17 to 18% He, and 1 to 2%
a shock tube. hydrocarbon, mostly CH4 . The present work assumes
it to be 81% H2 , 17.5% He, and 1.5% CH4 . As men-
The main shortcoming in [3] is that the stagnation- tioned, the reaction kinetics of this mixtures behind the
region flow is different from a shock tube flow in that the bow sock wave are proposed in [3]. The proposed scheme
former provides a longer flow residence time: very roughly, does not include C3 or C2 H. However, in a shock layer
in the inviscid region of the shock layer, the flow speed is over a blunt body, one must also account for C3 and C2 H
proportional to the distance from the wall. Therefore, the because they are formed as a result of ablation of the
flow residence time approaches a logarithmic infinity. It heat-shield. Therefore three exchange reactions involving
is the purpose of the present work to remedy this flaw in C3 and C2 are included in the present work. The needed
[3]. Here a viscous shock layer calculation was carried out rate coefficients are drawn from [10] to [17]. The reaction
for the stagnation region. The chemistry and radiation scheme so constructed is shown in Table 1. Therein, the
calculations are the same as in [3]. The calculation is ap- quantities A, n, and Td are the three parameters in the
plied to the entry flight conditions of Hollis et al and Jits expression for the forward reaction rate coefficient
et al. The results show that the radiative heating rates
vary between those of the corresponding shock tube flow kf = ATan exp(−Td /Ta ) cm3 /s.
and twice those values depending on the condition. The
calculation results were presented for a range of possible The quantity e in Table 1 is the energy feedback factor,
flight conditions, with the hope that the radiative heat- i.e., the amount of the energy fed back to vibrational-
ing rate for all practical Neptune entry missions could be electron-electronic mode during a reaction, expressed as
predicted to an engineering approximation through inter- a fraction of the reaction energy.
polation of these values.
Table 1. Reaction mechanisms and rate coefficients. A is in the units of cm3 /s.

No. Reaction A n Td e Remarks

R1 H2 + H2 → H + H + H2 -4.48 eV 2.967−7 -0.5165 52,530 0.73 Kim et al(2009)


R2 H2 + H → H + H + H - 4.48 eV 3.18−4 -1.0735 55,105 0.65 Kim et al (2009)
R3 H2 + He → H + H + He - 4.48 eV 9.272−4 -1.4773 55,105 0.63 Kim et al (2009)
R4 CH3 + H2 → CH4 + H - 0.00 eV 3−20 3 4,045 0 Sutherland et al(2001)
R5 CH3 + H → CH2 + H2 - 0.22 eV 1−10 0 7,600 0 Baulch et al(1994)
R6 CH2 + H → CH + H2 + 0.12 eV 1−11 0 900 0 NIST(2003)
R7 C + H2 → CH + H - 0.99 eV 6.6−10 0 11,700 0 Dean et al(1991)
R8 C + H → CH+ + e - 7.15 eV 1.26−21 2 90,300 0 Present work
R9 C+ + H2 → CH+ + H - 0.37 eV 6.86−10 -0.115 5086 0 Park (2010c)
R10 C+ + H → C + H+ - 2.34 eV 4.31−31 3.26 13,018 0 Park(2010c)
R11 H + M → H+ + e + M - 13.6 eV ∗ ∗ ∗
0.8 Park(2010a)
R12 C + e → C+ + e + e - 11.26 eV 1.63−8
0.330 98,810 0.8 Dunn (1971)
R13 C2 H + H → C2 + H2 - 1.44 eV 3.50−4 0. 30,000 0 Kruse and Roth (1997)
R14 C3 + H → C2 H + C - 1.64 eV 1.50−13 0. 19,000 0 Kruse and Roth (1997)
R15 C3 + C → C2 + C2 - 1.41 eV 1.44−11 0.75 16,000 0 Kruse and Roth (1997)

Not expressible as a simple function, see Park (2010a).

Flow-field stant. The code allows two temperatures, i.e., heavy par-
The viscous shock layer code developed earlier and ticle translational-rotational temperature and vibrational-
used by the present author for calculating several dif- electron-electronic temperature, and assumes a laminar
ferent entry flows [18,19] is used in the present work flow. At the shock wave, the flow conditions are deter-
also. It solves the viscous shock layer equation along the mined by the Rankine-Hugoniot condition accounting for
stagnation streamline assuming the pressure to be con- the shock slip. At wall, ablation and pyrolysis gas in-
3

jection phenomena are described totally kinetically, i.e., Calculation was made first for a case considered in [3],
the specie production rate at wall is equated with the which is a flow in a shock tube with a test gas consisting
rate of removal by diffusion. In calculating the diffusion of a 82.7%H2 +15.8%He+1.5%CH4 . The ambient tem-
flux of species, the bifurcation model is used. Transport perature and pressure are 300 K and 1 Torr, respectively.
properties necessary for this model are obtained from the The shock wave moves at a speed of 30 km/s. In [3],
work of Kim et al [20]. The bifurcation parameters were this condition was calculated to a distance from the shock
determined from the assumption that they are inversely wave of 1.2 cm. In the viscous shock layer calculation, the
proportional to the square-root of the species molecular freestream condition is the same, but the flow-field stud-
weight. All calculations in this work are carried out as- ied is the stagnation region of a spherical body of 14 cm
suming no ablation, though the code is capable of han- radius. The shock stand off distance becomes 1.4 cm for
dling ablation phenomenon. this flow.
Because of the assumption of constant pressure, the In Fig. 1, the calculated electron densities are com-
method overestimates shock stand-off distance by nearly pared. As seen, the distance to ionization equilibrium is
20% for a given nose radius. Therefore, nose radius nearly the same in the two calculations. The present cal-
smaller than the true nose radius is used in order to pro- culation exhibits a more gradual rise to the equilibrium
duce the desired shock stand-off distance. value compared with the shock tube flow. This can be at-
In the calculation of radiation, only H and C are con- tributed to the effect of diffusion: as mentioned, the shock
sidered. Radiation of other species is neglected because [3] tube flow calculation was made assuming an inviscid flow.
showed that they are unimportant. Radiation is loosely The viscous diffusion and conduction phenomena tend to
coupled with the flow-field. Radiation is calculated from smear the variations in properties.
60 nm to 800 nm at 39160 selected wavelengths. Twenty In Fig. 2, temperatures are compared between the
iterations between the flow-field calculation and radiation present viscous shock layer calculation and the inviscid
calculation were performed in order to obtain a converged shock tube calculation. As seen here, difference is small.
solution.
Calculation was performed on a personal computer 4
4x10
equipped with the quadri-core I7 architecture. Typically,
5 hours was expended to obtain a converged solution. Calculated temperatures
p1=1 Torr, Us=30 km/s
Shock layer, present work
3 Shock tube, Park (2010c)

1.0x10
17 Heavy particle
Temperature, K

translational-rotational
Calculated electron density
p1=1 Torr, Us=30 km/s
82.7%H2-15.8%He-1.5%CH4 2
0.8 Shock layer, present work
Shock tube, Park(2010)
-3
Electron density, cm

Equilibration
0.6 distance 1

Vibrational-electron-electronic
0.4
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Distance from shock wave, cm
0.2
Figure 2. Comparison of the calculated temperatures
Wall
between the inviscid shock tube flow
and the present viscous shock layer flow.
0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
In Fig. 3, the radiative heat flux in the downstream
Distance from shock wave, cm direction are compared between the two cases. As the fig-
Figure 1. Comparison of the calculated electron ure shows, the rise in the radiative heat flux starts earlier
densities between the inviscid shock tube flow in the shock layer flow, corresponding to the early rise in
and the present viscous shock layer flow. electron density seen in Fig. 1. In the shock layer solu-
tion, radiative heat flux reaches a peak at the boundary
Results
layer edge, and decreases toward the wall, understandably
Shock Tube Flow due to boundary layer absorption. The peak value at the
4

boundary layer edge is nearly the same as the peak value 8


Calculated radiative heat flux
in the shock tube flow. Hollis et al (2004) trajectory, no ablation
2
Present model, wall (133 kJ/cm )
2
Present model, boundary layer edge (185 kJ/cm )
2
Shock tube, Park (2010c) (104 kJ/cm )
3000 2
6 Hollis et al (2004) (46 kJ/cm )

2
Radiative heat flux, kW/cm
Calculated radiative heat flux
p1=1 Torr, Us=30 km/s
2500
82.7%H2-15.8%He-1.5%CH4
Shock layer, present work
2
Radiative heat flux, W/cm

Shock tube, Park (2010) 4


2000

1500
2

1000

500 0
Wall 150 160 170 180 190 200 210
Flight time, sec
0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 Figure 4. Radiative heat fluxes and heat loads
Distance from shock wave, cm in the entry flight into Neptune for the
Figure 3. Comparison of the calculated radiative trajectory of Hollis et al (2004).
heat flux between the inviscid shock tube flow
and the present viscous shock layer flow. 4x10
4
1.0x10
17

Neptune Entry Calculated temperatures and electron density


Hollis et al (2004) trajectory, 180 sec point
Temperature
In Fig. 4, the radiative heat fluxes in the entry flight Electron density 0.8
into Neptune are shown for the trajectory calculated 3

by Hollis et al [7] and are compared with their values

Electron density, cm
and with the values calculated assuming the stagnation Heavy particle
Temperature, K

translational-rotational 0.6
streamline flow to be a shock tube flow. The peak value
2
occurring at the boundary layer edge and the wall value
are shown. The calculation was made assuming a nose 0.4
radius of 0.28 meter, which reproduces the shock stand-

-3
off distance of 2.8 cm obtained by Hollis et al. The heat
1
loads, i.e., the time integration of the heat fluxes, are Vibrational-electron- 0.2
shown in the parentheses also. Both sets of present values electronic
are significantly higher than the other two sets of values.
In particular, the present heat load value at the wall is 0 0.0
three times higher than the value calculated by Hollis et 0.0 0.5 1.0 1.5 2.0 2.5 3.0

al. The present values are higher than the shock tube flow Distance from wall, cm

values at low densities where the nonequilibrium phenom- Figure 5. Flow-field properties for the 180 sec
ena are strong: the logarithmic increase in flow residence point in the entry trajectory of Hollis et al
time toward the wall is allowing equilibrium to be reached (2004). (a) Temperatures and electron density.
in this low density regime. Because of this phenomenon,
the highest radiative heat transfer rate occurred at 170 sec In Fig. 5(b), species mol fractions are shown. One
time point; in the solution by Hollis et al and in the shock notes in this figure that C+ exists in a respectable con-
tube calculation, the highest heat transfer rate occurs at centration. This implies that the excited states of C are
180 sec time point. populated significantly, and, consequently, radiation of C
The flow-field variables are shown for the 180 sec point could be strong.
in the Hollis et al’s trajectory in Figs. 5(a) to (c). In Fig. In Fig. 5(c), radiative heat fluxes in two directions,
5(a), the variations in temperature and electron density toward the wall and toward the shock wave, are shown.
are shown. Qualitatively, temperature variation in the The heat flux toward the wall reaches its peak value at the
shock layer is the same as that in a shock tube shown in distance of 2 cm. From Figs. 5(a) and (b) one sees that
Fig. 2. flow is not yet reached equilibrium at the peak-radiation
5

point: the two temperatures are slightly higher than the stand-off distance of 1 cm. Jits et al did not calculate
equilibrium temperature, and C+ concentration is slightly the radiative heat transfer rates in their work. Therefore,
higher than the equilibrium value. That is, the peak in the present calculation is compared only with the calcu-
radiation represents nonequilibrium radiation overshoot. lation made for a shock tube flow earlier by the author
0
[3]. As seen, the shock layer flow produces a stronger
10
H He
radiation compared with the shock tube flow in the low
10
-1
density regime: the peak value at 174 sec point is more
+
-2
H C
+ pronounced than in a shock tube flow, as was for the Hollis
10
et al’s case shown in Fig. 4.
-3
C
10
5000
Species mol fraction

Calculated species molfraction


-4 Hollis et al (2004) trajectory, 180 sec point CH2
10

-5
10 H2 4000
CH4

2
Radiative heat flux, W/cm
-6 +
10 CH CH
C2

10
-7
3000
CH3
-8
10 C3

10
-9 2000
C2H
Calculated radiative heat flux
Hollis et al (2004) trajectory, 180 sec point
0.0 0.5 1.0 1.5 2.0 2.5 3.0
with nose radius
Distance from wall, cm 1000 Wall
Boundary layer edge
Figure 5. (b) Species mol fractions.
7000 0
5 6 7 8 9 2 3 4 5 6 7 8 9
0.1 1
6000 Nose radius, m
Calculated radiative heat flux
Hollis et al (2004) trajectory, 180 sec point Figure 6. Dependence of radiative heat flux
Toward shock wave
2

on nose radius for the 180 sec point.


Radiative heat flux, W/cm

5000 Toward wall


5
4000 Calculated radiative heat flux
Jits et al (2005) trajectory, no ablation
2
Present work, wall (62.5 kW/cm )
2
3000 4 Present work, boundary layer edge (88.4 kJ/cm )
2
Shock tube flow, Park (2010c) (42.2 kJ/cm )
2
Radiative heat flux, kW/cm

2000
3

1000

2
0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Distance from wall, cm
1
Figure 5. (c) Radiative heat fluxes.
In order to explore how the nonequilibrium radiation
overshoot phenomenon manifests itself in radiative heat 0
transfer rates, calculation was repeated for the 180 sec 150 160 170 180 190 200 210

point with different nose radii. Fig. 6 shows the results. Flight time, sec

The figure shows that there is an unusually large increase Figure 7. Radiative heat fluxes and heat loads
in heat flux between the nose radii of 0.07 and 0.14 m. in the entry flight into Neptune for the
Seemingly, the overshoot phenomenon occurs prominently trajectory of Jits et al (2005).
in this regime for this particular case.
In Fig. 7, the radiative heat fluxes are shown for the Parametric Study
trajectory calculated by Jits et al [8]. For this calculation, The calculation was made by systematically varying
nose radius was taken to be 0.9 m to produce a shock the nose radius, flight speed, and freestream density. The
6

results are presented in Table 2. In the parentheses, the Conclusions


calculated shock stand-off distances are shown also. As
The viscous shock layer calculations made for the stag-
mentioned, the present method is inaccurate in predicting
nation region of a blunt body show approximately the
shock stand-off distance. For a comparison with a method
same relaxation phenomena as in a constant-area flow. In
more accurate than the present method in shock stand-
an entry flight into the planet Neptune, the viscous shock
off distances, the shock stand-off distance values should be
layer calculation shows a higher radiative heat transfer
used instead of nose radii. Only the boundary layer edge
rates than a shock tube flow in the low density regime. For
values of the radiative heat flux are given, because these
the entry flight considered by Hollis et al, the present cal-
are thought to be more meaningful: the wall values are
culation predicts substantially higher heat transfer rates.
affected by ablation phenomena. By interpolating these
Radiative heat transfer rates are presented parametrically
values, one should be able to predict the radiative heating
over a range of possible flight conditions.
rate to a Neptune aerobraking vehicle to an engineering
accuracy.
References
Table 2. Radiative heat fluxes calculated at
the boundary layer edge for different nose radii, [1] Park, C., “Nonequilibrium Ionization and Radia-
flight speed, and freestream density. Heat flux is tion in Hydrogen-Helium Mixtures,” AIAA Paper 2010-
in kW/cm2 . 814 (2010a); submitted to Journal of Thermophysics and
Heat Transfer.
ρ1 , kg/m3 3.16−6 1.0−5 3.16−5 1.0−4 3.16−4 1.0−3 [2] Park, C., “Stagnation-Point Radiative Heat Fluxes
in Neptune aerobraking,” paper presented at 7th Interna-
Nose radius = 10 cm tional Planetary Probe Workshop, Barcelona, June 14-18,
U s=30 0.579 3.223 5.882 11.420 2010 (2010b).
(1.254) (1.159) (1.043) (1.005) [3] Park, C., “Nonequilibrium Chemistry and Radia-
29 0.417 2.258 4.234 8.129 tion for Neptune Entry,” AIAA Paper 2510-4510, 2010
(1.275) (1.157) (1.052) (0.984) (2010c).
28 0.308 1.318 2.854 6.167 [4] Leibowitz, L. P., “Measurement of the Structure
(1.251) (1.150) (1.057) (1.039) of Ionizing Shock Wave in a Hydrogen-Helium Mixture,”
27 0.0439 0.733 1.808 3.785 The Physics of Fluids, Vol. 16, No. 1, 1973, pp. 59-68.
(1.222) (1.135) (1.058) (1.033) [5] Livingston, F. R., and Poon, T. Y., “Relaxation Dis-
26 0.0176 0.305 0.971 2.204 tance and Equilibrium Electron Density Measurements in
(1.191) (1.114) (1.051) (1.038) Hydrogen-Helium Plasmas,” AIAA Journal, Vol. 14, No.
Nose radius = 31.62 cm 9, 1976, pp. 1335-1337.
U s=30 0.521 2.988 4.897 6.497 [6] Bogdanoff, D. W., and Park, C., ”Radiative Inter-
(3.976) (3.656) (3.239) (3.046) action Between Driver and Driven Gases in an Arc-Driven
29 0.308 1.806 3.242 4.858 Shock Tube,” Journal of Shock Waves, Vol 12, 2002, pp
(4.044) (3.631) (3.265) (3.090) 205-214.
28 0.153 1.081 1.931 4.063 [7] Hollis, B. R., Wright, M. J., Olejniczak, J.,
(3.894) (3.614) (3.287) (3.185) Takashima, N., Sutton, K., and Prabhu, D., “Preliminary
27 0.049 0.623 1.205 2.504 Convective-Radiative Heating Environments for a Nep-
(3.894) (3.586) (3.302) (3.206) tune Aerocapture Mission,” AIAA Paper 2004-5177, 2004.
26 0.008 0.273 0.612 1.470 [8] Jits, R., Wright, M., and Chen, Y-K., “Closed-Loop
(3.800) (3.529) (3.292) (3.209) Trajectory Simulation for Thermal Protection System De-
Nose rdius = 100 cm sign for Neptune Aerocapture,” Journal of Spacecraft and
U s=30 0.534 3.095 4.564 3.754 Rockets, Vol. 42, No. 6, November-December 2005, pp.
(12.70) (11.72) (10.16) (9.230) 1025-1034.
29 0.267 1.893 2.947 2.846 [9] Gautier, D., Conrath, B. J., Owen, T., de Pater,
(12.78) (11.56) (10.23) (9.390) I., and Atreya, S. K., “The Troposphere of Neptune,” in
28 0.130 0.964 1.782 2.608 Neptune and Triton, edited by D. P. Cruikshank, The
(12.60) (11.40) (10.30) (9.777) University of Arizona Press, Tucson, Arizona, 1995, pp.
27 0.052 0.535 0.951 1.640 547-610.
(12.38) (11.28) (10.31) (9.869) [10] Sutherland, J. W., Su, M. C., and Michael, J. V.,
26 0.010 0.245 0.444 0.944 “Rate Constants for H + CH4 , CH3 + H2 , and CH4 Dis-
(12.10) (11.17) (10.33) (9.934) sociation at High Temperature,” International Journal of
Chemical Kinetics, Vol. 33, No. 11, 2001, pp. 669-684.
7

[11] Dean, A. J., Davidson, D. F., and Hanson, R. K.,


“A Shock Tube Study of Reactions of C Atoms with H2
and O2 Using Eximer Photolysis of C3 O2 and C Atom
Atomic Resonance Absorption Spectroscopy,” The Jour-
nal of Physical Chemistry, Vol. 95, No. 1, 1991, pp.
183-181.
[12] http://kinetics.nist.gov/CKMech/Links.jsp, NIST
Chemical Kinetics Database, Ver. 7.0, 2003.
[13] Baulch, D. L., Cobos, C. J., Cox, R. A., Prank,
P., Hayman, G., Just, Th., Kerr, J. A., Murrels, T.,
Pilling, M. J., Troe, J., Walker, R. W., and Warnatz,
J., “Evaluated Kinetic Data for Combustion Modelling,
Supplement I,” Journal of Physical Chemistry Reference
Data, Vol. 26, No. 6, 1994, pp. 847-1033.
[14] Kim, J. G., Kwon, O. J., and Park, C., “Master
Equation Study and Nonequilibrium Chemical Reactions
for H + H2 and He + H2 ,” Journal of Thermophysics and
Heat Transfer, Vol. 23, No. 3, July-September 2009, pp.
443-453.
[15] Kim, J. G., Kwon, O. J., and Park, C., “State-
to-State Rate Coefficients and Master Equation Study for
H2 + H2 ,” AIAA Paper 2009-1023, January 2009; to be
published in Journal of Thermophysics and Heat Trans-
fer.
[16] Dunn, M. G., “Measurement of C+ + e + e and
CO+ + e Recombination in Carbon Monoxide Flows,”
AIAA Journal, Vol. 9, No. 11, Vol. 9, No. 11, November
1971, pp. 2184-2191.
[17] Kruse, T., and Roth, P., “Kinetics of C3 Reac-
tions During High-Temperature Pyrolysis of Acetylene,”
Journal of Physical Chemistry, A, Vol. 101, 1997, pp.
2138-2146.
[18] Park, C., “Dissociative Relaxation in Viscous Hy-
personic Shock Layers,” AIAA Journal, Vol. 2, No. 7,
July 1964, pp. 1202-1207.
[19] Park, C., “Calculation of Stagnation-Point Heat-
ing Rates Associated with Stardust Vehicle,” Journal of
Spacecraft and Rockets, Vol. 44, No. 1, January-February
2007, pp. 24-32.
[20] Kim, J. G., Kwon, O. J., and Park, C., “A High
Temperature Elastic Collision Model for DSMC Based on
Collision Integrals,” AIAA Paper 2006-3803, 2006.

S-ar putea să vă placă și