Conference PaperPDF Available

Computational Prediction of NASA Langley HYMETS Arc Jet Flow with KATS

Authors:
Computational Prediction of NASA Langley HYMETS
Arc Jet Flow with KATS
¨
Umran D¨uzel
, Olivia Schroeder
, Huaibao Zhang
, and Alexandre Martin§
University of Kentucky, Lexington, KY, 40506
Atmospheric entry occurs at very high velocities which makes the physics of the flow
complicated to predict. Although high enthalpy flow facilities such as arc jets and plasma
wind tunnels are widely used to examine the entry flow conditions, they can involve as-
sumptions which make unrealistic predictions. High-fidelity modeling tools are essential for
accurate predictions of the physical phenomena that occur during planetary entry as well
as to better interpret ground testing. The Kentucky Aerothermodynamic and Thermal-
response System - Fluid Dynamics (KATS-FD) is a research tool developed mainly to pre-
dict planetary entry conditions. In order to better understand and interpret the physics of
an arc jet flow, KATS-FD is used and compared to a set of experimental data. The latter is
provided by NASA Langley Hypersonic Material Environment Testing System (HYMETS)
arc jet facility. From the array of experiments a test case was selected which consisted of a
6.5 MJ/kg enthalpy flow that discharged into a vacuum chamber at 0.228 kPa. The simu-
lation parameters are compared to radial and axial velocity profiles measured using Planar
Laser Induced Fluorescence (PLIF). Surface pressure and heat flux profiles are analyzed.
The species concentration along the centerline is computed and presented. A parametric
study of the nozzle wall boundary conditions and variation of bulk enthalpy is carried out
in order to better assess the ground testing and numerical results.
Nomenclature
F
F
FConvective flux matrix
F
F
FdDiffusive flux matrix
SVector of source terms
QVector of conservative variables
PVector of primitive variables
J
J
JConvective flux Jacobian
HBulk enthalpy, MJ/kg
˙mMass flow rate, slmp
ρDensity, kg/m3
PPressure, Pa
TTemperature, K
uFreestream velocity, m/s
qHeat flux, W/cm2
sArc-length coordinate, m
Subscript
cChamber
wWall
iVariable number
tr Translational-rotational
ve Vibrational-electronic
Ph.D. Candidate, Department of Mechanical Engineering.
M.Sc. Student, Department of Mechanical Engineering.
Research Associate, currently at Computational Fluid Dynamics Research Center, Sun Yat-sen University.
§Associate Professor, Department of Mechanical Engineering, Associate Fellow AIAA.
1 of 13
American Institute of Aeronautics and Astronautics
I. Introduction
One of the biggest challenges of space flight is the ability of safely enter an atmosphere. Although the
duration of that phase is usually no more than a few minutes, entry vehicles encounter extreme conditions
in this short period of time, which can be damaging to the integrity of the vehicle and the payload. This
is caused by the high velocities that the vehicle has at the beginning of the entry phase. Typical entry
velocities can vary between 7 km/s and 13 km/s in Earth atmosphere.1,2At these velocities, a strong bow
shock is formed in front of the spacecraft. Most of the kinetic energy from the free stream is transferred
into thermal energy as it goes through the formed shock wave. The vehicle has to overcome this thermal
load. The encountered flow becomes chemically reactive and more complex to characterize. The extreme
environment must be overcome through design parameters so that the spacecraft can preserve its integrity.
Thermal Protective Systems (TPS) are developed so that the harsh entry conditions can be withstood.
The characteristics such as size, thickness, material composition, etc., of the TPS are chosen based on the
expected heat load of the atmospheric entry. One important remark on TPS is that they are described as
Single-Point-of-Failure (SPOF)3which indicates that if it fails, the complete system fails. Hence, accurate
prediction of the entry conditions is vital in order to select and size the TPS.
Figure 1. Schematic view of HYMETS test section (Image from
Inman et al.4)
In order to better understand the physics be-
hind entry, an experimental approach – flight
and ground testing – is commonly used. The
flight tests are the best option to understand
and predict atmospheric entry with high accu-
racy but they are rare due to their high cost.
The ground testing facilities such as arc jets
and inductively coupled plasma (ICP) wind tun-
nels are build mainly to match the enthalpy of
entry flow conditions. Although these facilities
are widely used for reproducing the flight condi-
tions, they maintain drawbacks such as high un-
certainties, difficulties in achieving target condi-
tions, cost, and deviation from flight conditions.
These facilities also use significant assumptions
in order to simplify the flow field and be able
to measure the targeted parameters.58High
fidelity real gas computational fluid dynamics
(CFD) solvers have been developed recently not
only to predict the flight conditions but also to
better understand the physics behind atmospheric entry phenomena, to assist in ground testing and to better
interpret experimental results. Currently used numerical tools for entry and ground testing simulations are
presented in review study.9With the aid of these CFD tool, it is also possible to predicted parameters
such as boundary layer thickness, boundary layer edge enthalpy, edge Mach number, and shear stress which
are note quite captured during testing but are essential to characterizing TPS. Moreover, if needed, these
parameters would be main inputs for material response10 and thermal stress simulations.11
The purpose of this paper is to model the NASA Langley Hypersonic Material Environment Testing
System (HYMETS) arc jet flow and, at the same time, assess the capability of the Kentucky Aerothermo-
dynamic and Thermal-response System - Fluid Dynamics (KATS-FD)12,13 to model such flow. The study
includes the HYMETS Mach 5 nozzle simulation which discharge of the flow into a vacuum chamber at a
pressure of 228 Pa. Moreover, a test specimen will be placed downstream of the nozzle exit in order to
determine the heat flux and surface pressure.
II. HYMETS arc jet facility and test data
The experiment used to verify the KATS-FD code was published in Inman et al.,4and conducted at
HYMETS facility. The HYMETS facility was developed at NASA Langley Research Center in the late 60s.
2 of 13
American Institute of Aeronautics and Astronautics
The facility was constructed to operate at 100 kW power. The plasma wind tunnel specifically was used
to conduct research for emissivity, catalycity, and dynamic oxidation testing of materials for hypersonic
vehicles.14 The facility was upgraded to operate up to 400 kW power in early 2000. Currently, the facility
is capable of heating a specimen surface up to 2800 K (4500 oF). Specimen stagnation pressure can vary
between 0.013 amd 0.079 atm, and the nominal free stream Mach number is 5. The bulk enthalpy of the
free stream can reach up to 27 MJ/kg (11500 Btu/lbm). Detailed information and historical development
of the facility is presented in Splinter et al.14
Figure 1 depicts the arc plasma generator, the nozzle, and the vacuum test chamber of the facility. The
laser sheet where the measurements are taken is illustrated by the dashed purple region. The mounted nozzle
is a Mach 5 converging/diverging nozzle, with a 12.7 mm throat diameter, and a 63.5 mm exit diameter. The
diverging section has 8ohalf angle. Depending on the atmosphere of the planet to replicate, the gases can
vary in content of Argon (Ar), Nitrogen (N2), Oxygen (O2) and Carbon Dioxide (CO2). The gas is heated in
the arc-plasma-generator section, accelerated through the nozzle and is discharged to 0.6 m diameter by 0.9
m diameter vacuum chamber. The facility is capable of using intrusive (pressure probes, heat flux probes,
etc.) and non-intrusive (flow visualization, spectroscopy, etc.) measurement instrument.
The provided test data4was obtained by Planar Laser-Induced Fluorescence (PLIF) which is a non-
intrusive measurement technique. PLIF was primarily used to visualize the arc jet flow, and the PLIF
molecular tagging velocimetry (MTV) was used to measure the axial velocity profile of the jet flow. Moreover,
Doppler-Shift based velocimetry was used to gather the radial velocity profiles. The provided test data is
intended to emulate Earth’s atmosphere, using a chemical composition of 75% N2, 20% O2and 5% Ar by
volume. Figure 2 illustrates the vacuum chamber before testing, the test specimen facing the plasma flow,
and the laser sheet for measuring the flow properties downstream of the nozzle exit.
(a) Test chamber (b) Test specimen (c) Nozzle flow
Figure 2. Images of the HYMETS facility at NASA Langley Research Center (Image from Inman et al.4)
Table 1 presents the test conditions for the provided data. Axial velocity profiles and radial velocity
profiles are presented in Fig. 3. The axial velocity profiles are presented as averaged and single-measurement
results. As it was explained in Inman et al.,4the center of the given points correspond to the measured mean
velocities, and the widths correspond to twice the associated uncertainty of the mean velocities. Stagnation
pressure was measured with a pitot tube and the 47 mm (1.85 inch) long, 3.175 mm (0.125 inch) edge radius
test specimen was located 50.8 mm (2 inch) downstream of the nozzle exit.
Table 1. Experimental conditions for provided test case (Earth atmosphere)
Enthalpy Mass Flow Rate Arc-Pressure Chamber Pressure
6.5 MJ/kg 403 slmp 109 kPa 0.228 Pa
III. Numerical approach
III.A. Numerical framework
The HYMETS Mach 5 non-equilibrium nozzle flow is simulated using the Fluid Dynamics (FD) of the Ken-
tucky Aerothermodynamics and Thermal-response System (KATS). KATS-FD is a high-fidelity, aerother-
modynamic, CFD solver developed at the University of Kentucky, used for the study and computation of
3 of 13
American Institute of Aeronautics and Astronautics
(a) Axial velocity – averaged (b) Axial velocity single-
measurement
(c) Radial velocity
Figure 3. Velo city profiles measured in the HYMETS facility (Image from Inman et al.4)
high enthalpy flows. KATS-FD has been used for the study of non-equilibrium re-entry flows, such as in
Stardust2and Mars Science Laboratory, as well as arc jet flows over test samples, and pure nozzle flow.12
The framework of KATS-FD is detailed in Ref.,12 and examples of its capability are shown in Refs.1517 The
code uses a second order finite-volume approach with implicit, first-order backward Euler time integration
scheme to solve the 3D, unsteady, compressible Navier-Stokes equations. In their conservative form, the
governing equations can be represented by
Q
∂t +∇ · (F
F
F − F
F
Fd) = S,(1)
In KATS, the conservative variables in the time derivative Qare changed to primitive variables Pthrough
the introduction of a flux Jacobian J
J
J=Q
∂t , making the solved system
J
J
JP
∂t +∇ · (F
F
F − F
F
Fd) = S,(2)
Here, Qis the vector of conservative variables, Pthe vector of primitive variables, Sis the source terms vector,
and F
F
Fand F
F
Fdare the convective and diffusive flux matrices, respectively. In this work, the convective flux
vector uses a Steger-Warming flux-splitting scheme for adding numerical dissipation to the shock in order
to stabilize the solution.18 While for equilibrium flows KATS-FD uses a single temperature to describe
the flow in the energy equation, for non-equilibrium problems, Park’s two temperature model is used. In
this model, the translational and rotational energy modes of the particles is described by one temperature
Ttr, while the vibrational and electronic energy modes are described by another Tve. The combination of
translational-rotational temperature and the vibrational-electronic temperature is also used to control the
dissociation reactions.
Finally, the transport properties of the gas are modeled using Wilke’s mixing rule,19 where viscosity is
calculated from Blottner’s curve fits.20 The species thermal conductivities are determined with Eucken’s
model.21
III.B. Mesh description
III.B.1. Nozzle
The computational prediction of the HYMETS arc jet flow first is assessed with the Mach 5 nozzle test
case. KATS-FD not only has been developed to compute external flow, but also to predict the internal flow
4 of 13
American Institute of Aeronautics and Astronautics
behavior. Since the problem is 2-dimensional, symmetry was used to reduce computational time, which makes
all of the computational domains axisymmetric. The nozzle geometrical constraint and the computational
domain is illustrated in Fig. 4. The nozzle mesh includes 210 nodal points in the stream-wise and 64 in the
radial directions. In order to ensure accurate simulation of the flow field, the mesh was refined at the wall.
Figure 4. HYMETS Mach 5 nozzle dimensions and computational domain
III.B.2. Nozzle connected to the vacuum chamber
Figure 5. HYMETS Mach 5 nozzle flow discharged into vacuum chamber
The nozzle is attached to a vacuum
chamber and the domain is con-
structed assuming there is no test
specimen. The vacuum chamber
domain is assembled according to
the real geometry , which has di-
mension of 0.6 m diameter by 0.9
m length. The mesh generated is
presented in Fig. 5. Although it is
computationally expensive, the en-
tire chamber is modeler to verify the
flow development at the nozzle exit.
III.B.3. Test specimen at the vac-
uum chamber
Having created the nozzle flow with
and without the vacuum chamber, the test specimen was then placed at the testing location. Fig. 6 represents
the HYMETS Mach 5 nozzle mounted to a vacuum chamber where a test specimen is located 1.5 inches
downstream of the nozzle exit. The domain is cut-off in the vertical direction in order to decrease the
computational time. The mesh domain includes 400 points in the stream-wise and 200 in the radial direction.
Figure 6. Test specimen located 51 mm downstream of the nozzle exit
III.B.4. Shock alignment and boundary layer refinement
Once a converged solution is achieved with a coarse grid, the shock in front of the sample can be traced.
The trace is used to generate a more refined mesh in the aft-shock region. The new mesh contains cells
that are re-oriented to follow the tangential direction of the the shock, so that the latter can be accurately
5 of 13
American Institute of Aeronautics and Astronautics
captured. A new spacing of 1 ×105m is also applied near the wall of the sample for better resolution in
the boundary layer and, ultimately, more accurate heat flux computation on the surface of the sample. The
mesh alignment process is manually repeated several time, until the shock is accurately captured. Figure 7
illustrates the process by showing one of the intermediate meshes obtained.
(a) Refined mesh near the shock (b) Shock capturing after refinement
Figure 7. Mesh alignment for accurate shock capturing
III.C. Initial and boundary conditions
In order to initialized the problem, the domain is split into two regions: upstream and downstream of the
nozzle throat. The downstream of the nozzle throat is initialized by applying a shrink factor to the pressure
and density. This helps the solution to converge. A shrink factor of 2.1×103is used. On the other hand,
the upstream of the nozzle is set to match inlet conditions. The nozzle inlet conditions are assigned based
on experimental data for the test case in conjunction with an equilibrium solver for computing mixture
and species properties. The parameters used from the experiment are presented in Table 1. The chemical
composition of an Earth’s atmosphere is used, which consists of a six species mixture (N2, O2, NO, N, O, and
Ar). Using the molar concentrations of N2, O2, and Ar, and mixture enthalpy of 6.5 MJ/kg, Mutation++22
is used to calculate the partial densities of the species at equilibrium. The translational-rotational as well
as vibrational-electronic temperature are determined by iteratively solving the equilibrium state until the
enthalpy is matched for the given arc-pressure. The same is done to determine the mixture density. Once
the mixture density is known, the inlet velocity is determined from the mass flow rate. The prescribed inlet
conditions are presented in Table 2.
Table 2. Inlet Conditions
T Tve u ρN2ρO2ρNO ρNρOρAr
3832 K 3832 K 208 m/s 61.7 g/m34.84 g/m34.50 g/m335.4 g/m31.22 ×102g/m36.06 g/m3
The outlet condition is set based on the pressure of the vacuum chamber, pc= 228 Pa. Finally, the
simulation uses a cold wall assumption of Tw= 350 K, and a non-catalytic, no-slip wall boundary condition.
IV. Numerical results and comparisons
IV.A. Nominal results
The experimental measurements were conducted at different locations for axial and radial velocity profiles.
The measurement window is represented in a sketch in Fig. 8 and the locations of each profile is stated
6 of 13
American Institute of Aeronautics and Astronautics
in Table 3. The locations are defined from the exit of the arc jet nozzle. Mach 5 nozzle configuration is
simulated first in order to asses the internal flow results. Figures 9(a) and 9(b) illustrate the nozzle flow
contour of pressure and Mach number, respectively. By design and according to quasi-1D inviscid flow
theory, the nozzle should generate a Mach 5 flow. The nozzle simulation resulted in a Mach number of 4.9
which is in good agreement with the nozzle design specification.
Figure 8. Graphical representation of the laser measurement sheet and profile locations
Table 3. Laser measurement profile locations
LiL1L2L3L4
LAxial 0.57 cm 1.97 cm 2.97 cm
LRadial 1 cm 2 cm 3 cm 4 cm
The species mass concentration along the nozzle centerline is given in Fig. 9(c). In arc jet nozzles, non-
equilibrium processes take place and species dissociate and recombine. Figure 10 illustrates the nozzle flow
discharged into the vacuum chamber which is at 0.228 kPa. It can be seen that the flow is under expansion.
Varying the pressure can reduce the expansion.23 The flow behavior is consistent with the nozzle flow studies
available in the open literature.2325 Nozzle discharge into the vacuum chamber simulation (without a test
sample placement) is carried out in order to observe the flow development.
The test sample described above is a 25 mm diameter SiC probe placed 51 mm away from the nozzle
exit. The SiC probe is represented in Figure 11(a) and the placement of the probe is shown in Figure 11(b).
Unfortunately, Inman et al.4– from which the vleocity measurements were taken – does not report the
stagnation heat flux values. However, Splinter et al.14 contains a wide range of measurement obtained with
different sets of probes, one of which is very close to the values used here (Earth’s atmosphere, 6.5 MJ, 400
slpm). This measured heat flus was obtained with a TeflonR
Slug Calorimeter. Although it is not the same
probe and exact test, the TeflonR
Slug Calorimeter can provide an estimated non-catalytic cold-wall heat
flux to be compared with the KATS-FD predictions.
In high enthalpy arc jet testing, the flow directly hits the sample and a bow shock is formed. Figures 12(a)
and 12(b) illustrate the Mach and velocity contours over the sample. The shock is accurately captured by
aligning the grid to the shock,26 as shown in Sec. III.B.4. The test specimen wall is assumed to be cold
(350 K) and non-catalytic. This is a strong assumption but it is sufficient for the heat flux evaluation and
comparison to the results obtained with the Teflon R
Slug Calorimeter.14 The nozzle wall is also assumed to
be cold (350 K) and non-catalytic.
Figures 13(a) and 13(b) represent the computed surface pressure and heat flux of the corresponding
test case with respect to arc length. Although the behavior of the profiles are consistent with previous
studies, the measured values have not yet been obtained for a comparison with a SiC probe configuration.
Thus, the profiles are compared to a Teflon R
Slug Calorimeter14 heat flux measurement under similar testing
conditions. Stagnation heat flux predictions are approximately 20% different from measured values. This can
be due to the use of two different probes which may have different geometries and diameters, or differences
in testing conditions. Also, it is stated that the uncertainty in non-catalytic cold wall heat flux can be up to
7 of 13
American Institute of Aeronautics and Astronautics
(a) Pressure contours
(b) Mach contours (c) Species concentration along nozzle centerline
Figure 9. Simulation of the HYMETS nozzle using KATS-FD
Figure 10. Nozzle discharging into vacuum, velocity contour
8 of 13
American Institute of Aeronautics and Astronautics
(a) SiC probe as a test sample (Image from Splin-
ter et al.14
(b) Location and configuration of test sample
Figure 11. Geometry and location of the test sample used for the simulation of the HYMETS facility
(a) Mach contours (b) Velocity contours
Figure 12. Simulation of the HYMETS nozzle and test-chamber, with specimen, using KATS-FD
(a) Pressure along the surface (b) Heat flux along the surface
Figure 13. Surface properties of a test specimen inside the HYMETS facility obtained using KATS-FD
9 of 13
American Institute of Aeronautics and Astronautics
64.5%.14 Hence, the results indicate that the simulations have a good agreement for heat flux and surface
pressure. Axial velocity and radial velocity profiles are presented in Figs. 14(a)-14(b). The experimental
profiles are measured as presented in Fig. 8. The shaded area indicates the minimum and maximum value
of single shot images taken with PLIF.4The profiles and the shaded area show that KATS results are
approaching the measured minimum axial velocity profiles. Also, the radial velocity profiles are in good
agreement with the measurements. Although the radial velocity profile follows the same flow behaviour as
the experimental one, including the uncertainty analysis to the plots can show perfect approximation with
the experiments.
(a) Axial velocity profile at location L1(b) Axial velocity profile at location L2
(c) Axial velocity profile at location L3(d) Radial velocity profiles
Figure 14. Comparison of the experimental velocity profiles and the numerical results obtained using KATS-FD for
the HYMETS facility
IV.B. Nozzle wall boundary conditions and enthalpy variations
The nozzle exit is assumed to have sharp edge in the simulations however in reality the edge can be trimmed
or rounded so that the expansion can be different from the simulations. In order to answer the effect of
nozzle exit edge being sharp or curved, a set of simulations under 6.5 MJ enthalpy were conducted, with the
mesh shown in Fig 15. The nozzle exiting area, for both sharp and rounded edge, was kept constant. The
results of the simulations are presented in Fig. 16(a). The profiles indicates that the effect is taking place in
radial position which indicates that it affects the flow expansion. The change in axial velocity is recorded to
be between 1% and 2%.
10 of 13
American Institute of Aeronautics and Astronautics
Figure 15. Rounded nozzle exit geometry
Another consideration was nozzle wall boundary
condition. In the results presented in the previous
section, KATS-FD under-predicts the axial velocity.
It was therefore hypothesized that the flow was los-
ing energy through the nozzle before reaching the
measurement location. To address this issue, a set
of simulations, also at 6.5 MJ, was performed where
the nozzle wall was adiabatic rather than set to a
cold-wall temperature. The results of this test are
shown in Fig. 16(a). It can be noted that while
the change in magnitude of the axial velocity at the
centerline is very small, the velocity profile is altered
significantly, making this a parameter worth considering in future works.
Lastly, uncertainty in the bulk enthalpy is an important concern both for ground testing and computa-
tional prediction. The measured bulk enthalpy uncertainty can be up to 14%.14 Therefore, it was of interest
to understand the degree to which the axial velocity profile changed by varying the enthalpy. A new set
of simulations were run at 9 MJ, and the comparison with the results of the previous section is shown in
Fig. 16(b). As expected, in the high enthalpy case, the magnitude of the velocity increases. Furthermore,
the profile of the velocity is practically unchanged, since only close to the centerline do the velocities differ
from each other. The centerline velocity gains 13% with increase of 2.5 MJ. Thus, the uncertainty in the
bulk enthalpy is an important aspect to consider when performing numerical predictions.
(a) Various nozzle wall boundary conditions (b) Enthalpy variation
Figure 16. Simulation results of the axial velocity profiles at L1for different configurations
V. Conclusion and remarks
High enthalpy flow facilities are commonly used for simulating conditions experienced during planetary
entries. However, these tests carry a high degree of uncertainty and are not fully capable of replicating
entry environments. Thus, there is a need for high fidelity modeling tools which can be applied to ground
testing as well as flight conditions. The scope of the current research is to asses the capability of KATS-FD
as a computational prediction tool by comparing results with a set of experiments conducted at HYMETS
facility at NASA Langley. In the future, KATS-FD can assist in ground testing by predicting unmeasured
parameters such as edge enthalpy, which is needed for surface recession investigation.
The current simulations show good agreement with the measurements. The solutions showed that the
axial and radial velocity profiles can be improved by correcting the bulk enthalpy, re-configuring the nozzle
exit edge and taking the nozzle wall boundary condition into consideration. The simulations with a cold-wall
nozzle, and sharp exit corner, at 6.5 MJ bulk enthalpy approaches the lower velocity limit of experiment
11 of 13
American Institute of Aeronautics and Astronautics
single shot. Increasing the enthalpy by 40%, results in an increase of centerline velocity by 13%.It is
therefore concluded that uncertainties associated with bulk enthalpy are a non-negligible consideration when
conducting simulations of arc jet flows, specifically concerning centerline velocity.
Changing the geometry of the exit of the nozzle can also influence the velocity profile as it can change the
flow expansion. However the change in magnitude of the velocity is not very significant. Another important
note from these simulations is that using the cold-wall assumption in the nozzle is conservative and it affects
the velocity profiles both in magnitude and shape. A more accurate option may be to use a temperature
profile instead.
The specimen is assumed to have a cold wall, non-catalytic surface. The resulting stagnation heat flux is
compared with Teflon R
Slug Calorimeter measurements whose testing conditions were assumed to be similar
to testing conditions. Even though two different probes were compared and knowing the fact that cold-wall
non-catalytic heat flux measurements can have up to 64.5% uncertainty, the results are in good agreement
with the experiment.
Acknowledgments
Financial support for this work was provided by SpaceTech-REDDI ESI Award NNX16AD18G, and
NASA Kentucky Space Grant Graduate Fellowship NNX15AR69H. The authors greatly appreciate the help
of R.S.C. Davuluri from Gas Surface Interaction Laboratory with the KATS framework. Many thanks to S.
Splinter from NASA Langley for providing details of the nozzle geometry. Special thanks to T. Gok¸cen from
NASA Ames Research Center for his guidance with the nozzle flow conditioning.
References
1Anderson Jr., J. D., Hypersonic and High-Temperature Gas Dynamics, AIAA Education Series, American Institute of
Aeronautics and Astronautics, Inc., Reston, VA, 2nd ed., 2006.
doi:10.2514/4.861956
2Kontinos, D. A. and Wright, M. J., “Introduction: Atmospheric Entry of the Stardust Sample Return Capsule,” Journal
of Spacecraft and Rockets, Vol. 47, No. 5, 2010, pp. 705–707.
doi:10.2514/1.51522
3Laub, B. and Venkatapathy, E., “Thermal Protection System Technology and Facility Needs For Demanding Future
Planetary Missions,” Planetary Probe Atmospheric Entry and Descent Trajectory Analysis and Science, edited by A. Wilson,
No. ESA SP-544, European Space Agency, Noordwijk, Netherlands, 2004, pp. 239–247.
4Inman, J. A., Bathel, B. F., Johansen, C. T., Danehy, P. M., Jones, S. B., Gragg, J. G., and Splinter, S. C., “Nitric-Oxide
Planar Laser-Induced Fluorescence Measurements in the Hypersonic Materials Environmental Test System,” AIAA Journal,
Vol. 51, No. 10, oct 2013, pp. 2365–2379.
doi:10.2514/1.j052246
5Chazot, O. and Panerai, F., Hypersonic Nonequilibrium Flows: Fundamentals and Recent Advances, chap. High-Enthalpy
Facilities and Plasma Wind Tunnels for Aerothermodynamics Ground Testing, Progress in Astronautics and Aeronautics,
American Institute of Aeronautics and Astronautics, 2014, pp. 471–521.
doi:10.2514/5.9781624103292.0471.0522
6Sakraker, I., Aerothermodynamics of Pre-Flight and In-Flight Testing Methodologies for Atmospheric Entry Probes, Phd
thesis, Universit´e de Li`ege and von Karman Institute for Fluid Dynamics, Li`ege, Belgium, February 2016.
7Viladegut, A., Duzel, U., and Chazot, O., “Diffusion effects on the determination of surface catalysis in Inductively
Coupled Plasma facility,” Chemical Physics, Vol. 485–486, 2017, pp. 88 – 97.
doi:https://doi.org/10.1016/j.chemphys.2017.02.002
8Duzel, U., “Validation Of Gas/Surface Interaction Modeling In Plasmatron Facility For In-Flight Experiment Develop-
ment,” Tech. rep., von Karman Institute for Fluid Dynamics, 2015.
doi:10.13140/rg.2.2.30219.28966
9Lachaud, J., Magin, T. E., Cazmuta, I., and Mansour, N. N., “A short review of ablative material response models and
simulation tools,” Tech. rep., European Space Agency, 2011, 7th Aerothermodynamics Symposium.
10Weng, H., Bailey, S. C., and Martin, A., “Numerical study of iso-Q sample geometric effects on charring ablative
materials,” International Journal of Heat and Mass Transfer, Vol. 80, No. Supplement C, 2015, pp. 570 – 596.
doi:10.1016/j.ijheatmasstransfer.2014.09.040
11Fu, R., Weng, H., Wenk, J., and Martin, A., “Two-way strongly coupled thermo-mechanical solver for ablation problems,”
55th AIAA Aerospace Sciences Meeting and Exhibit, AIAA Paper 2017-0439, Grapevine, TX, January 2017.
doi:10.2514/6.2017-0439
12Zhang, H., High Temperature Flow Solver for Aerothermodynamics Problems, Ph.D. Thesis, University of Kentucky,
Lexington, KY, July 2015.
13Weng, H. and Martin, A., “Multidimensional Modeling of Pyrolysis Gas Transport Inside Charring Ablative Materials,”
Journal of Thermophysics and Heat Transfer, Vol. 28, No. 4, December 2014, pp. 583–597.
doi:10.2514/1.T4434
12 of 13
American Institute of Aeronautics and Astronautics
14Splinter, S. C., Bey, K. S., Gragg, J. G., and Brewer, A. R., “Comparative Measurements of Earth and Martian Entry
Environments in the NASA Langley HYMETS Facility,” 49th AIAA Aerospace Sciences Meeting, AIAA Paper 2011-1014,
Orlando, FL, January 2011.
doi:10.2514/6.2011-1014
15Davuluri, R. S. C., Zhang, H., and Martin, A., “Effect of spalled particles thermal degradation on a hypersonic flow field
environment,” 54th AIAA Aerospace Sciences Meeting, AIAA Paper 2016-0248, San Diego, CA, January 2016.
doi:10.2514/6.2016-0248
16Davuluri, R. S. C., Zhang, H., and Martin, A., “Numerical Study of Spallation Phenomenon in an Arc-Jet Environment,”
Journal of Thermophysics and Heat Transfer, Vol. 30, No. 1, jan 2016, pp. 32–41.
doi:10.2514/1.t4586
17Li, Z., Zhang, H., Bailey, S. C., Hoagg, J. B., and Martin, A., “A data-driven adaptive Reynolds-averaged Navier–Stokes
k – w model for turbulent flow,” Journal of Computational Physics, Vol. 345, sep 2017, pp. 111–131.
doi:10.1016/j.jcp.2017.05.009
18Steger, J. L. and Warming, R. F., “Flux vector splitting for the inviscid gasdynamic equations with applications to finite
difference methods,” Journal of Computational Physics, Vol. 40, No. 2, April 1981, pp. 263–293.
doi:10.1016/0021-9991(81)90210-2
19Wilke, C. R., “A viscosity equation for gas mixtures,” Journal of Chemical Physics, Vol. 18, No. 4, 1950, pp. 517–519.
doi:10.1063/1.1747673
20Blottner, F. G., Johnson, M., and Ellis, M., “Chemically reacting viscous flow program for multi-component gas mixtures.”
Technical Report SC-RR–70-754, Sandia National Laboratories, Albuquerque, New Mexico, 1971.
doi:10.2172/4658539
21Vincenti, W. G. and Kruger, C. H., Introduction to physical gas dynamics, Wiley, New York, February 1965.
22Scoggins, J. B. and Magin, T. E., “Development of Mutation++: Multicomponent Thermodynamic and Transport
Properties for Ionized Plasmas written in C++,” 11th AIAA/ASME Joint Thermophysics and Heat Transfer Conference,
AIAA Paper 2014-2966, Atlanta, GA, 16-20 June 2014.
doi:10.2514/6.2014-2966
23Gokcen, T. and Alunni, A., “Flow Characterization Studies of the 10-MW TP3 Arc-Jet Facility: Probe Sweeps,” 46th
AIAA Thermophysics Conference, American Institute of Aeronautics and Astronautics, jun 2016.
doi:10.2514/6.2016-4153
24Noh, S. and Kim, K. H., “Numerical Prediction of Tungsten Ablation under Arc Heater Experimental Conditions,” 54th
AIAA Aerospace Sciences Meeting, American Institute of Aeronautics and Astronautics, jan 2016.
doi:10.2514/6.2016-0247
25Gokcen, T. and Skokova, K., “CFD Simulations of the IHF Arc-Jet Flow: Compression-Pad/Separation Bolt Wedge
Tests,” 47th AIAA Thermophysics Conference, AIAA Paper 2017-4451, June 2017.
doi:10.2514/6.2017-4451
26Kim, K., Kim, C., and Rho, O., “Accurate computations of hypersonic flows using AUSMPW+- scheme and shock-aligned
grid technique,” 29th AIAA Fluid Dynamics Conference, AIAA Paper 1998-2442, June 1998.
doi:10.2514/6.1998-2442
13 of 13
American Institute of Aeronautics and Astronautics
... KATS is a finite volume computational framework that features second order spatial accuracy and, first order implicit backward Euler temporal integration. KATS has been previously used to solve hypersonic aerothermodynamic CFD problems, as well as compute MR solutions [16][17][18][19][20][21][22][23]. In this work, KATS-US is developed based on merging KATS-CFD and KATS-MR together. ...
... A study of an arc jet test sample simulation is carried out under supersonic and hypersonic flow conditions in order to assess the capability of the implemented models. The arc jet test sample is chosen to be a 25 mm diameter probe similar to the ones used at HYMETS experiments and simulations [18,[34][35][36][37] . The material properties such as permeability, porosity, thermal conductivity, are chosen to show the capability of the new models. ...
... The inlet boundary conditions are obtained from the nozzle simulation study. Similar study is carried out in Ref. [18,39]. The velocity is obtained by using the mass flow rate and nozzle geometry. ...
Conference Paper
The KATS-US solver uses a modified Volume Averaged Navier-Stokes set of equations to model the interaction between ablative material and hypersonic flow. The approach uses a single set of equations for both fluid and porous domains. In the present work, a two-temperature model is added to the code. The implementation of the model was verified by comparing to published results. The thermal non-equilibrium effect within the porous medium is presented and discussed. A new equation is proposed that can be used to model vibrational accommodation inside the porous medium.
Conference Paper
View Video Presentation: https://doi.org/10.2514/6.2022-3948.vid Various sample geometries are proposed for future arc-jet experiments that are intended to physically capture spalled particles to more accurately analyze their size and shape. These samples were modeled using a computational fluid dynamics (CFD) solver and a spallation particle-tracking code to determine which geometry can best capture spalled particles ejected from the sample surface.
Conference Paper
View Video Presentation: https://doi.org/10.2514/6.2022-1501.vid Reentry and hypersonic vehicles are exposed to high temperatures during flight that require the use of thermal protection systems to protect the internal structure, payload, and crew. Simulating the response of a thermal protection system to different flight conditions provides time dependent, spatial information for all regions of the material, but such simulations are difficult to perform accurately and can incur large computational expense for geometrically and thermochemically complex materials. In this paper, we present a new high-order material response solver, named the Coupled Hypersonic Protection System (CHyPS) solver, based on the discontinous Galerkin formulation. CHyPS is shown to match results generated using the Porous material Analysis Toolbox based on OpenFOAM (PATO) for standard ablation test cases, including the effects of pyrolysis, ablation, and material recession.
Conference Paper
View Video Presentation: https://doi.org/10.2514/6.2022-1636.vid In this paper, the SU2-NEMO CFD solver is used to simulate conditions in the HyMETS arc jet test facility in order to test models of gas-surface interaction for catalytic heating augmentation in nonequilibrium flows. Simulation predictions of surface heat flux and pressure on a calorimeter probe are verified and validated against numerical results and experimental data. A sensitivity analysis with respect to surface chemistry modeling parameters is then conducted focusing on the impact of catalytic efficiency on surface heating rates.
Article
A blended drag coefficient model is constructed using a series of empirical relations based on Reynolds number, Mach number, and Knudsen number. When validated against experiments, the drag coefficient model produces matching values with a standard deviation error of 2.84% and a maximum error of 11.87%. The model is used in a Lagrangian code which is coupled to a hypersonic aerothermodynamic CFD code, and the particle velocity and trajectory are validated against experimental results. The comparative results agree well and show that the new blended drag coefficient model is capable of predicting the particle motion accurately over a range of Reynolds number, Mach number, and Knudsen number.
Preprint
A blended drag coefficient model is constructed using a series of empirical relations based on Reynolds number, Mach number, and Knudsen number. When validated against experiments, the drag coefficient model produces matching values with a standard deviation error of 2.84% and a maximum error of 11.87%. The model is used in a Lagrangian code which is coupled to a hypersonic aerothermodynamic CFD code, and the particle velocity and trajectory are validated against experimental results. The comparative results agree well and show that the new blended drag coefficient model is capable of predicting the particle motion accurately over a range of Reynolds number, Mach number, and Knudsen number.
Thesis
Full-text available
Atmospheric reentry occurs at hypersonic velocities. During this process, a strong bow shock is formed in front of the vehicle. Most part of the kinetic energy of the free stream flow is transferred into thermal energy across the shock and therefore extremely high temperatures are reached in this region. Atmospheric reentry vehicles are equipped with Thermal Protection Systems, TPS, in order to overcome the extreme heat flux that is developed during reentry. The heat flux determines the characteristics of TPS. Catalytic Recombination Coefficient is one of the main parameters that should be modeled for heat flux determinations. The objective of the research master project is to assess the catalytic model application and validation at off-stagnation point configuration and to address the flight extrapolation methodology for surface catalytic properties at off-stagnation region. A test campaign has been prepared and a at plate has been tested with Plasmatron facility at VKI. The wall heat flux along the flat plate has been evaluated from the test campaign. A numerical simulation study has been conducted with CFD++ over 2D flat plate model which represents the off-stagnation configuration. Experimental results have been checked with numerical simulation in order to assess a validation. Catalytic Recombination Coefficient has been determined at off-stagnation configuration based on heat flux measurements with the Plasmatron and CFD simulations. Finally, addressing to flight extrapolation methodology for surface catalytic properties at off-stagnation region has been discussed.
Article
Full-text available
This paper presents a new data-driven adaptive computational model for simulating turbulent flow, where partial-but-incomplete measurement data is available. The model automatically adjusts the closure coefficients of the Reynolds-averaged Navier-Stokes (RANS) turbulence equations to improve agreement between the simulated flow and the measurements. This data-driven adaptive RANS (D-DARK) model is validated with 3 canonical flow geometries: pipe flow, backward-facing step, and flow around an airfoil. For all test cases, the D-DARK model improves agreement with experimental data in comparison to the results from a non-adaptive RANS model that uses standard values of the closure coefficients. For the pipe flow, adaptation is driven by mean stream-wise velocity data from 42 measurement locations along the pipe radius, and the D-DARK model reduces the average error from 5.2% to 1.1%. For the 2-dimensional backward-facing step, adaptation is driven by mean stream-wise velocity data from 100 measurement locations at 4 cross-sections of the flow. In this case, D-DARK reduces the average error from 40% to 12%. For the NACA 0012 airfoil, adaptation is driven by surface-pressure data at 25 measurement locations. The D-DARK model reduces the average error in surface-pressure coefficients from 45% to 12%.
Article
Atomic recombination is an important process to consider when computing the heat flux transferred to the wall of a re-entry vehicle. Two chemical processes are influencing the species diffusion in the boundary layer surrounding a re-usable Thermal Protection System: gas phase reactions and catalytic recombination at the surface. The coupling between them is not normally taken into account when determining the catalytic recombination coefficient ( ) in plasma facilities. This work aims to provide evidence of such coupling based on both a theoretical analysis and an experimental campaign in the VKI-Plasmatron facility. Recombination coefficient measurements at off-stagnation point configuration on a linear copper calorimeter are provided. An evolution from a high-catalytic to a low-catalytic condition due to the boundary layer growth along the probe is observed. This result is consistent with a parametric analysis carried out using the in-house non-equilibrium boundary layer solver, which shows how the experimentally determined catalysis could be influenced by the amount of gas-phase recombination inside the boundary layer.
Conference Paper
Two-way coupling is performed between a spallation code and a hypersonic aerothermodynamics CFD solver to evaluate the effect of spalled particles on the flow field. Time accurate solutions are computed in argon and air flow fields. A single particle simulations and multiple particles simulations are performed and studied. The results show that the carbon vapor released by spalled particles tend to change the composition of the flow field, particularly the upstream region of the shock.