Knight 2012

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Progress in Aerospace Sciences 48–49 (2012) 8–26

Contents lists available at SciVerse ScienceDirect

Progress in Aerospace Sciences


journal homepage: www.elsevier.com/locate/paerosci

Assessment of CFD capability for prediction of hypersonic shock interactions


Doyle Knight a,n, José Longo b,1, Dimitris Drikakis e, Datta Gaitonde c,2, Andrea Lani d, Ioannis Nompelis f,
Bodo Reimann b, Louis Walpot g
a
Rutgers - The State University of New Jersey, New Brunswick, NJ 08903, USA
b
German Aerospace Center DLR, 38108 Braunschweig, Germany
c
Air Force Research Laboratory, Wright-Patterson Air Force Base, OH 45433, USA
d
Von Karman Institute for Fluid Dynamics, Brussels, Belgium
e
Aerospace Sciences Department, Cranfield University, Cranfield, Bedfordshire MK43 0AL, United Kingdom
f
Department of Aerospace Engineering and Mechanics, University of Minnesota, Minneapolis, MN 55455, USA
g
Advanced Operations and Engineering Services, The Netherlands

a r t i c l e i n f o a b s t r a c t

Available online 10 January 2012 The aerothermodynamic loadings associated with shock wave boundary layer interactions (shock
Keywords: interactions) must be carefully considered in the design of hypersonic air vehicles. The capability of
Hypersonics Computational Fluid Dynamics (CFD) software to accurately predict hypersonic shock wave laminar
Computational fluid dynamics boundary layer interactions is examined. A series of independent computations performed by
Shock waves researchers in the US and Europe are presented for two generic configurations (double cone and
cylinder) and compared with experimental data. The results illustrate the current capabilities and
limitations of modern CFD methods for these flows.
& 2011 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2. Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1. Double cone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2. Cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3. Details of computations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.1. Governing equations and boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.1.1. Conservation of mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.1.2. Conservation of momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.1.3. Conservation of energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.1.4. Conservation of vibrational energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.1.5. Equation of state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.1.6. Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.2. Thermochemistry model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.3. Thermodynamic and transport properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.4. Numerical methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.4.1. Drikakis et al. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.4.2. Gaitonde . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.4.3. Lani . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.4.4. Nompelis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.4.5. Reimann . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

n
Corresponding author. Tel.: þ1 732 445 4464; fax: þ 1 732 445 3124.
E-mail address: [email protected] (D. Knight).
1
Currently at European Space Research & Technology Centre, Keplerlaan 1,
Postbus 299, 2200 AG Noordwijk, The Netherlands.
2
Currently at Department of Mechanical and Aerospace Engineering, Ohio
State University, Columbus, OH 43210, USA.

0376-0421/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.paerosci.2011.10.001
D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26 9

3.4.6. Walpot . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
4. Results for double cone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
4.1. Run 42 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4.1.1. Assessment of accuracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4.1.2. Flowfield structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.2. Run 40 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.2.1. Assessment of accuracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.3. Run 80 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
5. Results for cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
5.1. HEG I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
5.1.1. Assessment of accuracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
5.1.2. Flowfield structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5.2. HEG III . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5.2.1. Assessment of accuracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
6. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

1. Introduction
and base flows with and without plume interaction [2]. Subgroup
3 of WG10 examined CFD capability for shock wave–turbulent
Renewed interest in hypersonic air vehicles such as the Boeing
boundary layer interactions focusing on five generic configura-
X-51 (Fig. 1) has focused research on topics critical to hypersonic
tions: 2-D compression corner (Fig. 3a), 2-D expansion–compres-
flight. The design of hypersonic air vehicles involves numerous
sion corner (Fig. 3b), 2-D shock impingement (Fig. 3c), 3-D single
engineering disciplines including aerothermodynamic analysis. In
fin (Fig. 2a) and 3-D double fin (Fig. 2b). All cases were perfect gas
particular, the interaction of shock waves with the vehicle
flows. All 2-D shock interaction flow simulations were performed
boundary layers can lead to regions of enhanced aerothermody-
using either Direct Numerical Simulation (DNS) or Large Eddy
namic loading, and therefore accurate modeling of shock wave
Simulation (LES) models, and all 3-D shock interaction flow
boundary layer interactions (‘‘shock interactions’’) is essential.
simulations were performed using RANS methods. The report
During the past decade two NATO Research Technology Organiza-
concluded that DNS and LES results for 2-D shock wave turbulent
tion (RTO) Working Groups (WGs) have assessed the capabilities for
boundary layer interactions showed significant progress in pre-
prediction of aerothermodynamic loads in high speed flight. AGARD
dicting the flow; however, the Reynolds numbers of the simula-
Working Group 18 (WG18) examined the Computational Fluid
tions were relatively low due to computational resource
Dynamics (CFD) capability for prediction of 2-D and 3-D shock wave
requirements, and no comparison with experimental surface heat
laminar and turbulent boundary layer interactions for three generic
transfer measurements was performed. The report concluded that
configurations: single fin (Fig. 2a), double fin (Fig. 2b) and hollow
new RANS concepts for 3-D shock wave turbulent boundary layer
cylinder flare (Fig. 2c) [1]. The single and double fin configurations
interactions showed improvement in prediction of the flow;
involved shock wave turbulent boundary layer interactions, while the
however, heat transfer was not accurately predicted.
hollow cylinder configuration included both laminar and turbulent
This paper presents the results of the participants in RTO AVT
shock boundary layer interactions. All cases were perfect gas flows.
136 Subtopic No. 2. The focus is an assessment of CFD for the
All turbulent flow simulations were performed using Reynolds-
specific issue of shock interactions and control surfaces in non-
Averaged Navier–Stokes (RANS) models. The report concluded that
equilibrium laminar flows. The paper presents a comparison of
while turbulent perfect gas shock interaction predictions were
computed and experimental results for two new configurations.
accurate for 3-D mean pressure and primary separation locations,
The first configuration is a double cone (Fig. 4a) from Calspan
nevertheless the skin friction and heat transfer were poorly predicted.
University of Buffalo Research Center (CUBRC). The second con-
RTO Working Group 10 (WG10) conducted a detailed exam-
figuration is a cylinder (Fig. 4b) from the German Aerospace
ination of CFD capability for six areas relevant to hypersonic
Center (DLR). Experimental data are available for two different
flight: boundary layer instability and transition, real gas flows,
enthalpy conditions under laminar flow conditions.
laminar hypersonic viscous–inviscid interactions, shock–shock
interactions, shock wave–turbulent boundary layer interactions

2. Experiments

2.1. Double cone

The experiments were conducted in the LENS I shock tunnel at


CUBRC to obtain detailed surface and flow characteristics over a
double cone configuration with semi-angles of 251 and 551 and a
base diameter of 10.3 in. (Fig. 4a). Measurements were made in
nitrogen for total enthalpy conditions of 5.38 and 9.17 MJ/kg
creating conditions with negligible dissociation. A subsequent
experiment was performed at 5.28 MJ/kg. In allied studies, mea-
surements were made to determine the velocity and NO concen-
tration in the freestream for the airflows. The measured
freestream velocity was in excellent agreement with non-equili-
brium flow predictions using the nozzle code of Candler [3].
Fig. 1. Boeing X-51 waverider. However, the measured NO concentration was measurably less
10 D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26

Nomenclature uj Cartesian jth component of mass-averaged velocity


uaj Cartesian jth component of diffusion velocity of
Roman
species a
xj Cartesian jth coordinate
eav vibrational energy per unit mass of species a Ya mass fraction of species a
h static enthalpy per unit mass of mixture
ha static enthalpy per unit mass of species a
o Greek
hf a enthalpy of formation of species a at T ref
H total enthalpy per unit mass of mixture
kt thermal conductivity of translational heat flux
a index of species
dij Kronecker delta function
kr thermal conductivity of rotational heat flux
kv thermal conductivity of vibrational heat flux
e total energy per unit volume of mixture
Ma molecular weight of species a
ev vibrational energy per unit volume of mixture
yva characteristic harmonic oscillator temperature of spe-
p static pressure
cies a
qj Cartesian jth component of heat transfer vector in
energy equation
m molecular viscosity of mixture
qtj Cartesian jth component of translational heat flux
r density of mixture
qrj Cartesian jth component of rotational heat flux
ra density of species a
qvj Cartesian jth component of vibrational heat flux
ta vibrational–translational relaxation time
qatv translational–vibrational energy transfer
tij Cartesian viscous stress tensor
R universal gas constant
o_ a rate of production of species a per unit volume
t time
o_ v rate of production of vibrational energy per
unit volume
T translational–rotational (static) temperature
Tv vibrational temperature

than the predicted levels. The test conditions are shown in Table 1 3. Details of computations
where subscripts o and 1 represent reservoir and test section
conditions, respectively. The values represent best estimates The participants in the study, listed in Table 3, were solicited on
based on measurements made in the reservoir and freestream the basis of their expertise in CFD simulation of shock interactions in
in flow calibration studies. hypersonic flows. All details of the computations were determined
individually and independently by each participant including selec-
tion of the thermochemistry and transport models, inclusion or
omission of a separate vibrational energy model, choice of numerical
2.2. Cylinder algorithms, determination of the numerical grid and grid refinement
strategy, etc. In this section we present the governing equations and
A test campaign [4] to investigate the flow past a cylindrical boundary conditions, and the thermochemistry, thermodynamic and
model was performed at the high enthalpy shock tunnel (HEG) of transport models used by each participant. In addition, a brief
the German Aerospace Center (DLR). The cylindrical model, with a description of the numerical algorithms is provided.
radius of 45 mm and a span of 380 mm, was mounted on the
nozzle centerline with its axis transverse to the flow. It was
3.1. Governing equations and boundary conditions
equipped with 17 pressure transducers and the same number of
thermocouples to measure surface pressure and heat flux dis-
For convenience of comparison between the participants, the
tributions. The transducers were distributed along six rows
governing equations for unsteady, laminar, compressible flow of a
located close to the plane of symmetry at the mid-span location
mixture of reacting gases are written in Cartesian tensor notation
(10, 20 and 30 mm to the left and right of the plane of symmetry,
(using the modified Einstein summation convention3 where
covering a circumferential angle of 7601 with respect to the
convenient for readability). A different coordinate system may
inflow direction).
have been used for the governing equations by the participants.4
The large shock standoff distance of this configuration permitted
All computations assumed laminar flow.
the application of optical measurement techniques for the determi-
nation of the gas properties in the shock layer. Holographic inter-
ferometry and time resolved Schlieren were applied to measure 3.1.1. Conservation of mass
density distributions in the shock layer and the temporal evolution The conservation of mass for each specie is
of the bow shock shape. Freestream static and pitot pressures and
@ra @ra uj @
stagnation heat transfer on a sphere were recorded at each run for þ ¼o
_ aþ ðra uaj Þ ð1Þ
@t @xj @xj
calibration, normalization and statistical purposes.
The measurements on the cylinder were carried out at different for a ¼ 1, . . . ,n where ra is the density of species a, uj is jth
total enthalpies (HEG conditions I and III, 22.4 MJ/kg and 13.5 MJ/kg, Cartesian component of the mass-averaged velocity of the mix-
respectively) and with air as a test gas. The HEG reservoir and ture, and o _ a is the net rate of production of species a per unit
freestream data for the measurements are listed in Table 2.
At the experimental conditions the flow in the shock layer is
3
subject to non-equilibrium chemical relaxation processes that sig- The modified Einstein summation convention implies summation over all
values of a Roman index (e.g., j) when it appears twice in a given term. Summation
nificantly affect the density distribution and hence the shock standoff over a Greek index (e.g., a) is identified by a summation sign S.
distance. Therefore, this test case represents a useful basis for the 4
For example, the governing equations for the double cone configuration may
validation of the physico-chemical models used in CFD codes. be written in cylindrical polar coordinates.
D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26 11

Fig. 3. Configurations for WG10. (a) 2-D compression. (b) 2-D compression–
expansion. (c) Shock impingement.

where e is the total energy per unit volume


e ¼ rHp ð5Þ
with the total enthalpy per unit mass defined as
1
H ¼ hþ uu ð6Þ
2 i i
and the static enthalpy per unit mass defined as
X
n
h¼ Y a ha ð7Þ
Fig. 2. Configurations for WG 18. (a) Single fin. (b) Double fin. (c) Hollow cylinder. a¼1

where the mass fraction Y a of species a is


ua
volume. The species diffusion is ra j where j is the Cartesianua ra
Ya ¼ ð8Þ
jth component of the diffusion velocity of species a. r
The static enthalpy per unit mass for species a is
3.1.2. Conservation of momentum Z T
o
The conservation of mass-averaged momentum is ha ¼ hf a þ cpa dT þ eva ð9Þ
T ref
@rui @rui uj @p @tij o
þ ¼ þ ð2Þ where hf a is the enthalpy of formation of species a at T ref and eva is
@t @xj @xi @xj
the vibrational energy per unit mass of species a defined by
where the viscous stress tensor is defined by R ya v
  eva ¼ ð10Þ
2 @u @ui @uj Ma expðyva =T v Þ1
tij ¼  m k dij þ m þ ð3Þ
3 @xk @xj @xi
where R is the universal gas constant, M a is the molecular weight
v
where dij is the Kronecker delta. The mixture molecular viscosity of species a, ya is the characteristic harmonic oscillator tempera-
m and its dependence on the individual species’ molecular ture, and Tv is the vibrational temperature.
viscosities ma is described in Section 3.3. The heat flux vector is comprised of three terms
qj ¼ qtj þqrj þqvj ð11Þ
3.1.3. Conservation of energy
where qtj ,
and qrj , qvj
are the translational, rotational and vibra-
The conservation of energy is
tional heat flux, respectively,
!
@e @ @tij ui @qj @ X
n
@T
þ ðe þpÞuj ¼   ra ha uaj ð4Þ qtj ¼ kt ð12Þ
@t @xj @xj @xj @xj a¼1 @xj
12 D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26

Table 1
Test conditions for double cone.

Quantity Run 42 Run 40 Run 80

Ho (MJ/kg) 9.17 5.38 5.28


po (MPa) 17.9 17.3 8.48
To (K) 6611 4327 4269
U 1 (m/s) 4065 3094 3066
p1 (Pa) 121 129 63.6
r1 (kg/m3) 1:34  103 2:52  103 1:29  103
T 1 (K) 303 173 166
T 1 (K)
vib
3085 2735 2711
Y[N2] 0.9973 1.0000 0.9999
Y[O2] 0.0000 0.0000 0.0000
Y[NO] 0.0000 0.0000 0.0000
Y[N] 0.0027 0.0000 0.0001
Y[O] 0.0000 0.0000 0.0000
Re1 (m  1) 3:0  105 6:9  105 3:6  105

Table 2
Test conditions for cylinder.

Quantity I III

Ho (MJ/kg) 22.4 13.5


po (MPa) 35.0 48.3
To (K) 9200 7370
U 1 (m/s) 5956 4776
p1 (Pa) 476 687
r1 (kg/m3) 1:547  103 3:26  103
T 1 (K) 901 694
pp1 (kPa) 52.9 70.8
M1 8.98 8.78
Y½N2 1 0.7543 0.7356
Y½O2 1 0.00713 0.1340
Y½NO1 0.01026 0.0509
Y½N1 6:5  107 0.0000
Y½O1 0.2283 0.07955

Table 3
Participants.

Name Organization

Drikakis, Mosedale, Cranfield University, United Kingdom


Tissera
Gaitonde Air Force Research Laboratory, Wright-Patterson AFB,
OH, USA
Fig. 4. Configurations for AVT 136. (a) Double cone (dimensions in inches).
Lani Von Karman Institute for Fluid Dynamics, Brussels,
(b) Cylinder.
Belgium
Nompelis University of Minnesota, Minneapolis, MN, USA
@T
qrj ¼ kr ð13Þ Reimann DLR, Braunschweig, Germany
@xj Walpot Advanced Operations and Engineering Services, The
Netherlands
@T v
qvj ¼ kv ð14Þ
@xj

where the translational and rotational states of the gas are The heat flux is given by (14) and nd is the number of diatomic
assumed in equilibrium at temperature T and the vibrational species. The source term is
state is characterized by the vibrational temperature Tv. nd nd
X X
o_ v ¼ qatv þ o_ a eva ð17Þ
a¼1 a¼1
3.1.4. Conservation of vibrational energy
The conservation of vibrational energy is where qatv is the translational–vibrational energy transfer defined
! by the Landau–Teller model [5]
@ev @ev uj @qvj @ Xnd
v a
þ ¼  r e u þ o_ v ð15Þ
@t @xj @xj @xj a ¼ 1 a a j eav ðTÞeav ðT v Þ
qatv ¼ ra ð18Þ
ta
where ev is the vibrational energy per unit volume
nd
X and ta is the relaxation time [6,7].
ev ¼ ra eva ð16Þ An equation for vibrational energy was used in most, but not
a¼1 all, of the computations. Details are presented in Table 4.
D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26 13

Table 4 Table 6
Included vibrational energy equation. Mixture species and reactions.

Name Double cone Cylinder Configuration Species Reaction

Drikakis   Double conea N2 ,N N2 þ N2 ¼ N þ N þ N2


Gaitonde   N2 þ N ¼ N þ N þ N
Lani   Cylinder N2 ,O2 ,NO,N,O N2 þ N2 ¼ N þ N þ N2
Nompelis   N2 þ N ¼ N þ N þ N
Reimann  N2 þ NO ¼ N þ N þ NO
Walpot   N2 þ O2 ¼ N þ N þ O2
N2 þ O ¼ N þ N þ O
N2 þ O ¼ NO þ N
O2 þ N2 ¼ O þ Oþ N2
O2 þ N ¼ O þ O þ N
Table 5
O2 þ N ¼ NO þ O
Boundary conditions at wall.
O2 þ NO ¼ O þ O þ NO
O 2 þ O 2 ¼ O þ O þ O2
Boundary cond Drikakis Gaitonde Lani Nompelis Reimann Walpot
O2 þ O ¼ O þ Oþ O
NO þ N2 ¼ N þ O þ N2
Double cone
NO þ NO ¼ N þ O þ NO
uw 0 0 0 0 0 0
NO þ O2 ¼ N þ Oþ O2
Tw (K) 295.5a 294.7 294.0a 300 300 300
NO þ N ¼ N þ O þ N
294.1b 294.7b
NO þ O ¼ N þ O þ O
raw NOTE 1 NOTE 2 NOTE 1 NOTE 1 NOTE3 NOTE4
Tv Tv ¼ Tw Tv ¼ Tw Tv ¼ Tw Tv ¼ Tw Tv ¼Tw Tv ¼Tw a
The second reaction was not used by Walpot.
Cylinder
uw – 0 0 0 0 0
Tw (K) – 300 300 300 300 300
raw – NOTE 2 NOTE 1 NOTE 5 NOTE 3 NOTE6 Table 7
Tv – Tv ¼ Tw Tv ¼ Tw Tv ¼ Tw – Tv ¼Tw Reaction rates.

NOTE 1: Non-catalytic (i.e., @Y i =@n ¼ 0). Participant Reference


NOTE 2: Fully catalytic (i.e., equilibrium composition at Tw).
NOTE 3: Thermochemical equilibrium (maximum entropy) at surface. Drikakis Gnoffo et al. [27], Gupta et al. [28]
NOTE 4: Y½N2  ¼ 1:0, Y½N ¼ 0. Gaitonde Park [7]
NOTE 5: See Nompelis and Candler [22]. Lani Park [7]
Nompelis Park [29–32]
NOTE 6: Y½N2  ¼ 0:7624, Y½O2  ¼ 0:2376.
Reimann Park [7], Gupta et al. [28]
a
Run 40. Walpot Dunn and Kang [33], Park [7]
b
Run 42.

3.1.5. Equation of state Table 8


The equation of state is Dalton’s law Thermodynamic properties.

Xn
ra Participant Reference
p ¼ RT ð19Þ
M
a¼1 a Drikakis Gupta et al. [28]
Gaitonde Grossman and Cinnella [34], Chase et al. [35]
where R is the universal gas constant and M a is the molecular Lani Panesi [36]
weight of species a. Nompelis Blottner et al. [37]
Reimann Bottin [38]
Walpot Blottner et al. [37]

3.1.6. Boundary conditions


The boundary conditions used by the participants are detailed Table 9
in Table 5. Species thermal conductivity and viscosity models.

Participant Reference

3.2. Thermochemistry model Drikakis Gnoffo et al. [27], Gupta et al. [28]
Gaitonde Blottner et al. [37], White [39]
The same thermochemistry model was used by all participants Lani Barbante [40]
for each configuration. The double cone was modeled as a two- Nompelis Blottner et al. [37], Nompelis [41]
Reimann Gupta et al. [28]
component mixture (N2 and N). The cylinder was modeled as a
Walpot Blottner et al. [37]
five-component mixture (N2, O2, NO, N and O). The reactions used
by the participants are listed in Table 6. References for the
reaction rates are given in Table 7. Table 10
Mixture thermal conductivity and viscosity models.

Participant Conductivity Viscosity


3.3. Thermodynamic and transport properties
Drikakis Yos (see Gnoffo et al. [27]) Yos (see Gnoffo et al. [27])
The reference data for thermodynamic and transport proper- Gaitonde Wilke [42]
ties of the individual species are indicated in Tables 8 and 9, Lani Magin and Degrez [43,44], Hirschfelder et al. [45]
respectively. References for the models of mixture transport Nompelis Wilke [42]
Reimann Herning and Zipperer [46] Wilke [42]
properties are listed in Table 10. The models for species diffusion Walpot Wilke [42]
are indicated in Table 11.
14 D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26

Table 11 3.4. Numerical methods


Species diffusion models.
3.4.1. Drikakis et al.
Participant Model
The computations at Cranfield University were performed
Drikakis Gnoffo et al. [27] using the code CNS3D [8–10]. The code comprises a library of
Gaitonde Fick’s law [45], Le¼ 1.15 numerical methods, including second- and fifth-order MUSCL
Lani Stefan–Maxwell [47] schemes with low-Mach corrections and very high-order WENO
Nompelis Fick’s law [45], Le¼ 1.4
Reimann Fick’s law [45], Sc ¼0.7
schemes up to ninth-order accurate, for spatial discretization. The
Walpot Fick’s law [45], Le¼ 1.2 time-integration is obtained by TVD Runge–Kutta schemes, third
and fifth-order accurate. CNS3D is fully parallelized using MPI

7.0E+04 7.0E+04

6.0E+04 Experiment 6.0E+04


Computed Experiment
Computed (140976 cells)
5.0E+04 5.0E+04 Computed (563904 cells)

4.0E+04 4.0E+04
P (Nt/m )
P (Nt/m )

3.0E+04 3.0E+04

2.0E+04 2.0E+04

1.0E+04 1.0E+04

0.0E+00 0.0E+00
0 0.025 0.05 0.075 0.1 0.125 0.15 0 0.025 0.05 0.075 0.1 0.125 0.15
X (m) X (m)

Fig. 5. pw for Run 42. (a) Gaitonde. (b) Lani. (c) Drikakis et al. (d) Nompelis. (e) Reimann. (f) Walpot.
D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26 15

systems. CNS3D has been previously used in various hypersonic third-order upwind-biased MUSCL approach modulated by the
flows, with and without ablation effects [11–13]. Additional minmod limiter. Additional details are provided in Gaitonde [17].
details are provided in Tissera et al. [12].

3.4.3. Lani
The double cone Run 40 configuration was computed using a
3.4.2. Gaitonde cell-centered finite volume algorithm based on AUSM þ with
All results were obtained with the GASP code [14]. For the Least Squares reconstruction and the Venkatakrishnan limiter.
double-cone case, the Roe scheme was used (see, e.g., [15] for The double cone Run 42 was computed using a novel Contour
formulation with chemistry) with the van Albada limiter. For the Residual Distribution method with the BXC scheme. The cylinder
cylinder cases, the AUSMþ scheme [16] was employed with the configuration was simulated using a cell-centered finite volume

7.0E+06
7.0E+06 Run 42 - park I
Experiment 6.0E+06
6.0E+06 514x256 grid
1024x512 grid
5.0E+06
5.0E+06

Q (W/m2)
4.0E+06
Q (W/m2)

4.0E+06

3.0E+06
3.0E+06

2.0E+06
2.0E+06

1.0E+06 1.0E+06

0.0E+00 0.0E+00
0 0.025 0.05 0.075 0.1 0.125 0.15 0 0.025 0.05 0.075 0.1 0.125 0.15
X (m) X (m)

7.0E+06
7.0E+06 Max value 8.1 10
7

Experiment 6.0E+06
6.0E+06 Computed (48x128, VL)
Computed (48x128, WENO5th)
Computed (96x256, VL) Experiment
Computed (96x256, WENO5th) 5.0E+06 Computed
5.0E+06
Q (W/m2)

4.0E+06
Q (W/m2)

4.0E+06

3.0E+06 3.0E+06

2.0E+06 2.0E+06

1.0E+06 1.0E+06

0.0E+00 0.0E+00
0 0.025 0.05 0.075 0.1 0.125 0.15 0 0.025 0.05 0.075 0.1 0.125 0.15
X (m) X (m)

7.0E+06 7.0E+06

6.0E+06 Experiment 6.0E+06


Computed Experiment
Computed (140976 cells)
Computed (563904 cells)
5.0E+06 5.0E+06
Q (W/m2)

Q (W/m2)

4.0E+06 4.0E+06

3.0E+06 3.0E+06

2.0E+06 2.0E+06

1.0E+06 1.0E+06

0.0E+00 0.0E+00
0 0.025 0.05 0.075 0.1 0.125 0.15 0 0.025 0.05 0.075 0.1 0.125 0.15
X (m) X (m)

Fig. 6. qw for Run 42. (a) Gaitonde. (b) Lani. (c) Drikakis et al. (d) Nompelis. (e) Reimann. (f) Walpot.
16 D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26

algorithm based on AUSM þ with Least Squares reconstruction LU-SGS scheme have been used. Additional details are provided in
and the Venkatakrishnan limiter. Reimann and Hannemann [24].

3.4.4. Nompelis 3.4.6. Walpot


The computations were performed by a finite volume code The full unsteady laminar Navier–Stokes equations are solved
that uses the modified Steger–Warming flux splitting [18,19]. To using a finite volume algorithm. The algorithm is second order
achieve second order spatial accuracy a MUSCL [20] extrapolation accurate. The Van Albada limiter is used.
is done to the primitive variables. The equations are integrated
with a fully implicit line-relaxation method [21]. Additional
details are provided in Nompelis and Candler [22]. 4. Results for double cone

The test conditions are summarized in Table 1. The gas is


3.4.5. Reimann nitrogen. The stagnation enthalpy ho ¼9.17 MJ/kg (Run 42),
The flow solver is the DLR TAU code [23]. The Reynolds- 5.38 MJ/kg (Run 40) and 5.28 MJ/kg (Run 80). The stagnation
averaged Navier–Stokes equations are discretized by a finite pressure po ¼17.9 MPa (Run 42), 17.3 MPa (Run 40) and 8.48 MPa
volume technique. The AUSMDV second order upwind scheme (Run 80). The test section flow is nearly pure molecular nitrogen
with MUSCL reconstruction is used for the inviscid fluxes. For with a small (0.27%) concentration of atomic nitrogen (Run 42).
time discretization, a local time stepping with a three-stage The experimental data includes surface pressure and heat
Runge–Kutta algorithm, as well as an approximately factored transfer. The estimated uncertainty is 3% for surface pressure

Fig. 7. Flowfield structure for Run 42. (a) Overall flowfield. (b) Enlargement of flowfield. (c) Enlargement of flowfield.
D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26 17

and 4% for heat transfer relative to the individual measured indicate sensitivity of the upstream propagation (Fig. 5a) to the
values [25]. grid refinement; in particular, the predicted upstream propaga-
tion location for the 1024  512 grid are in excellent agreement
4.1. Run 42 with experiment, while the predicted upstream propagation for
the 512  256 grid underestimates the upstream propagation by
4.1.1. Assessment of accuracy approximately 4 mm.
The computed and experimental surface pressure are pre- The assessment of the accuracy of the predictions of ‘‘peak’’
sented in Fig. 5. All six simulations accurately predict the initial pressure is problematic for two reasons. First, it is not at all
plateau pressure (0:05 m ox o 0:09 m), although some of the evident that the experimental transducer at x ¼0.1133 m (the
simulations underestimate the upstream propagation of the first location of the highest measured p) corresponds to the peak
shock boundary layer interaction at x  0:05 m. Gaitonde’s results pressure in the experiment. Consequently, a computed peak

7.0E+04 Run 40 - 1024x512 mesh 7.0E+04

Experiment
6.0E+04 0.00300
6.0E+04
0.00325 Experiment
0.00350 Computed
5.0E+04 0.00370 5.0E+04
0.00400

P (Nt/m2)
P (Nt/m2)

4.0E+04 4.0E+04

3.0E+04 3.0E+04

2.0E+04 2.0E+04

1.0E+04 1.0E+04

0.0E+00 0.0E+00
0 0.025 0.05 0.075 0.1 0.125 0.15 0 0.025 0.05 0.075 0.1 0.125 0.15
X (m) X (m)

7.0E+04 7.0E+04 4
4 Max value 8.5 10
Max value 7.2 10

6.0E+04 Experiment 6.0E+04 Experiment


Computed (48x128, VL) 0.0031
Computed (48x128, WENO5th) increasing
0.0198 time
Computed (96x256, VL)
5.0E+04 Computed (96x256, WENO5th)
5.0E+04 0.0364
0.0530
0.0697
P (Nt/m2)
P (Nt/m2)

4.0E+04 4.0E+04 0.0863


0.1029
0.1195
3.0E+04 3.0E+04 0.1362
0.1528
0.1695
2.0E+04 2.0E+04 0.1861

1.0E+04 1.0E+04

0.0E+00 0.0E+00
0 0.025 0.05 0.075 0.1 0.125 0.15 0 0.025 0.05 0.075 0.1 0.125 0.15
X (m) X (m)

7.0E+04 7.0E+04

6.0E+04 Experiment 6.0E+04


Computed (t = 60 μs)
Experiment
Computed (t = 110 μs)
Computed (Level 0)
5.0E+04 Computed (t = 160 μs) 5.0E+04
Computed (Level 1)
Computed (t = 190 μs)
Computed (Level 2)
P (Nt/m2)

P (Nt/m2)

4.0E+04 4.0E+04

3.0E+04 3.0E+04

2.0E+04 2.0E+04

1.0E+04 1.0E+04

0.0E+00 0.0E+00
0 0.025 0.05 0.075 0.1 0.125 0.15 0 0.025 0.05 0.075 0.1 0.125 0.15
X (m) X (m)

Fig. 8. pw for Run 40 (y Time counter is notional and does not represent absolute time). (a) Gaitonde y . (b) Lani. (c) Drikakis et al. (d) Nompelis. (e) Reimann. (f) Walpot.
18 D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26

Run 40 - 1024x512 mesh


7.0E+06
7.E+06
Experiment
6.0E+06 0.00300
0.00325 6.E+06
0.00350 Experiment
5.0E+06 0.00370 Computed
0.00400 5.E+06
Q (W/m2)

4.0E+06

Q (W/m2)
4.E+06

3.0E+06 3.E+06

2.0E+06 2.E+06

1.0E+06 1.E+06

0.0E+00 0.E+00
0 0.025 0.05 0.075 0.1 0.125 0.15 0 0.025 0.05 0.075 0.1 0.125 0.15
X (m) X (m)

7.0E+06
7.0E+06
6.0E+06
Experiment Experiment
6.0E+06 Computed (48x128, VL) 0.0031
Computed (48x128, WENO5th) 0.0198
Computed (96x256, VL) 5.0E+06 0.0364
5.0E+06 Computed (96x256, WENO5th) 0.0530
0.0697
Q (W/m2)

4.0E+06 0.0863
Q (W/m2)

4.0E+06
0.1029
0.1195
3.0E+06 0.1362
3.0E+06
0.1528
0.1695
2.0E+06 2.0E+06 0.1861

1.0E+06 1.0E+06

0.0E+00 0.0E+00
0 0.025 0.05 0.075 0.1 0.125 0.15 0 0.025 0.05 0.075 0.1 0.125 0.15
X (m) X (m)

7.0E+06 7.0E+06

6.0E+06 6.0E+06 Experiment


Experiment Computed (Level 0)
Computed (t = 60 μs) Computed (Level 1)
5.0E+06 Computed (t = 110 μs) 5.0E+06 Computed (Level 2)
Computed (t = 160 μs)
Computed (t = 190 μs)
Q (W/m2)

Q (W/m2)

4.0E+06 4.0E+06

3.0E+06 3.0E+06

2.0E+06 2.0E+06

1.0E+06 1.0E+06

0.0E+00 0.0E+00
0 0.025 0.05 0.075 0.1 0.125 0.15 0 0.025 0.05 0.075 0.1 0.125 0.15
X (m) X (m)

Fig. 9. qw for Run 40 (y Time counter is notional and does not represent absolute time). (a) Gaitonde y . (b) Lani. (c) Drikakis et al. (d) Nompelis. (e) Reimann. (f) Walpot.

pressure greater than the experimentally measured peak pressure distribution would be needed. Consequently, a strict comparison
is not necessarily in error.5 Second, the transducer size is finite between the computed and experimental peak pressure is inap-
and therefore integrates the pressure over a small surface area. propriate in regions of extremely rapid changes. Nevertheless, a
Depending on the grid resolution, the lateral scale of the trans- visual comparison of the computed and experimental pressure
ducer may be equivalent to several times the grid spacing along profile in the vicinity of the peak pressure does indicate reason-
the surface, and therefore filtering of the computed pressure ably close agreement. Finally, the secondary plateau (x Z0:13 m)
is accurately predicted by all participants. All six simulations
resulted in steady flows in agreement with experiment.
5
The converse, of course, is not true. A computed peak pressure which is less
The computed and experimental surface heat transfer are
than the experimentally measured peak pressure (and outside the experimental presented in Fig. 6. The computed separation point at
error bars) is in error. The same holds true for heat transfer. x¼0.055 m is in good agreement with experiment for nearly all
D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26 19

simulations. It is interesting to note that while the refined grid peak heat transfer (Fig. 6) equals or exceeds the measured peak
(1024  512) results of Gaitonde show closer agreement with the heat transfer for all computations except Nompelis (Fig. 6d) and
surface pressure in the vicinity of separation, the coarser grid Riemann (Fig. 6e). The predicted heat transfer in the second
(512  256) results agree better with the experimental heat plateau region shows reasonable agreement with experiment
transfer in this region. This raises a question regarding the for all simulations. The convergence of all simulations to a steady
accuracy of the Park model for this type of flow. The computed state is in agreement with the experiment.

Fig. 10. qw for Run 40 vs t (experiment). (a) qw vs t for Run 40 (experiment). (b) qw vs t for Run 40 (experiment).

3.0E+04 3.0E+06

2.5E+04 2.5E+06
Experiment Experiment
Computed Computed
2.0E+04 2.0E+06
Q(W/m2)
P(N/m )
2

1.5E+04 1.5E+06

1.0E+04 1.0E+06

5.0E+03 5.0E+05

0.0E+00 0.0E+00
0 0.025 0.05 0.075 0.1 0.125 0.15 0 0.025 0.05 0.075 0.1 0.125 0.15
X(m) X(m)

Fig. 11. pw and qw for Run 80 (experiment and computation). (a) qw vs t for Run 80 (experiment). (b) qw for Run 80 (experiment). (c) pw for Run 80 (Nompelis). (d) qw for
Run 80 (Nompelis).
20 D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26

4.1.2. Flowfield structure structure illustrated in Fig. 7b and a rapid pressure rise on the
The structure of the double cone flowfield is illustrated in surface resulting in the formation of a separated region delimited
Fig. 7 for Run 42 based upon the computations by Gaitonde. by the separation and reattachment points at x  0:05 m and
Fig. 7a displays the entire flowfield. The first cone shock gener- x  0:11 m, respectively (Fig. 7a). The separation point is slightly
ated by the 251 cone is attached to the cone apex and is located downstream of the upstream influence point where the surface
very close to the first cone surface. A second cone shock is pressure begins to rise (Fig. 5a). Within the separation region the
generated by the 551 cone. The interaction of these two shock pressure is approximately constant (Fig. 5a) and the heat transfer
waves (shock–shock interaction) creates a complex wave is substantially reduced (Fig. 6a). The separation of the boundary

80000

70000

60000

50000
p (N/m )

40000

30000

20000 Experiment
Computed
10000

0
-80 -60 -40 -20 0 20 40 60 80
θ (deg)

80000

70000 Experiment
Computed (Park)
Computed (Dunn and Kang)
60000

50000
p (N/m )

40000

30000

20000

10000

0
-80 -60 -40 -20 0 20 40 60 80
θ (deg)

Fig. 12. pw for HEG I. (a) Gaitonde. (b) Lani. (c) Nompelis. (d) Reimann. (e) Walpot.
D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26 21

layer causes a significant deflection of the flow resulting in a shock interaction (Fig. 7b) with another contact surface and
separation shock (Fig. 7a). The separation shock and second cone reflected shocks. The downstream of the second cone shock is
shock interact (shock–shock interaction) (Fig. 7b) and form a subsonic while the flow beneath the contact surface emanating
triple point with a consequent reflected shock and a contact from the crossing shock interaction is supersonic. One of the
surface [26] (contact discontinuity). The reattachment of the reflected shocks from the crossing shock interaction further
boundary layer generates a reattachment shock (Fig. 7b) with a reflects from the surface and subsequently reflects from the sonic
consequent rapid increase in surface pressure and heat transfer at line as an expansion (Fig. 7c) which interacts with the cone
x  0:12 m (Figs. 5a and 6a). The interaction of the reflected shock surface thereby reducing the surface pressure and heat transfer at
from the triple point and reattachment shock results in a crossing x  0:12 m (Fig. 5a).

1E+07

9E+06

8E+06

7E+06

6E+06
Q (W/m )

5E+06

4E+06

3E+06 Experiment
Computed
2E+06

1E+06

0
-80 -60 -40 -20 0 20 40 60 80
θ (deg)

1E+07

9E+06

8E+06

7E+06
Q (W/m )

6E+06

5E+06

4E+06

3E+06

2E+06 Experiment
Computed (Park)
Computed (Dunn and Kang)
1E+06

0
-80 -60 -40 -20 0 20 40 60 80
θ (deg)

Fig. 13. qw for HEG I. (a) Gaitonde. (b) Lani. (c) Nompelis. (d) Reimann. (e) Walpot.
22 D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26

4.2. Run 40 heat transfer at several different points in time, while Fig. 9b
displays only one selected instant in time. The former simulations
4.2.1. Assessment of accuracy show significant unsteadiness in the flowfield as expected from
The computed and experimental surface pressure are pre- the surface pressure predictions. A similar conclusion was
sented in Fig. 8. Fig. 8(a), (c)–(f) show instantaneous surface obtained by Lani.
pressure at several different points in time, while Fig. 8b displays Nevertheless, the experimental heat transfer was observed to
only one selected instant in time. It is evident from the former five be steady as indicated in Fig. 10. The behavior of the experimental
simulations that the computed flowfield is unsteady and does not heat transfer at a location upstream of the separation region is
achieve a stationary (steady) configuration. A similar conclusion displayed in Fig. 10a and indicates that a steady heat transfer has
was obtained by Lani. been achieved. Similarly, the experimental heat transfer distribu-
The computed and experimental surface heat transfer are tion on the surface is displayed in Fig. 10b and indicates that a
presented in Fig. 9. Fig. 9(a), (c)–(f) show instantaneous surface steady distribution is achieved for the entire cone surface. By

8.0 5500.0
7.0 5000.0
6.0 4500.0
5.0 4000.0
4.0 3500.0
3.0 3000.0
2.0 2500.0
1.0 2000.0
0.0 1500.0
1000.0
500.0

13000
12000
12000
11000
11000
10000
10000
9000
9000
8000
8000
7000
7000
6000
6000
5000
5000
4000
4000
3000
3000
2000
2000
1000
1000
0

2000 0.30
1000 0.25
0 0.20
-1000 0.15
-2000 0.10
-3000 0.05
-4000 0.00
-5000
-6000
-7000
-8000

Fig. 14. Flow structure for HEG I (Walpot, Dunn and Kang Model). (a) Mach number. (b) Velocity (m/s). (c) T (K). (d) Tv (K). (e) T v T (K). (f) N.
D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26 23

analogy, a steady experimental surface pressure distribution is certainly not attributable to a particular grid since different grid
also achieved. convergence studies were performed by the participants.
This is a paradox. Six separate and independent simulations
have observed a significant unsteadiness in the flowfield, while 4.3. Run 80
the experiment shows steady flow. The discrepancy is certainly
not attributable to a particular flux algorithm, since a variety of A separate experiment (Run 80) was performed to assess the
different algorithms were used by the participants. Also, it is effect of the Reynolds number (Table 1). The Reynolds number is

80000

70000

60000

50000
P (N/m2)

40000

30000
Experiment
Computed
20000

10000

0
-80 -60 -40 -20 0 20 40 60 80
θ (deg)

80000

70000

60000

50000
P (N/m2)

40000

30000

20000
P (N/m )
Computed (Park)
10000 Computed (Dunn and Kang)

0
-80 -60 -40 -20 0 20 40 60 80
θ (deg)

Fig. 15. pw for HEG III. (a) Gaitonde. (b) Lani. (c) Nompelis. (d) Reimann. (e) Walpot.
24 D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26

1E+07
9E+06
8E+06
7E+06
6E+06

Q (W/m2)
5E+06
4E+06
3E+06
2E+06
1E+06
0
-80 -60 -40 -20 0 20 40 60 80
θ (deg)

1E+07
9E+06
8E+06
7E+06
6E+06
Q (W/m2)

5E+06
4E+06
3E+06
2E+06
1E+06
0
-80 -60 -40 -20 0 20 40 60 80
θ (deg)

Fig. 16. qw for HEG III. (a) Gaitonde. (b) Lani. (c) Nompelis. (d) Reimann. (e) Walpot.

3:6  105 m1 which is approximately one-half of the value for kg) and HEG III6 is a low enthalpy (ho ¼13.5 MJ/kg). Note that
Run 40. The development of the surface heat transfer distribution both cases have significantly higher enthalpy than the double
in time is shown in Fig. 11a, and the heat transfer distribution cone. The experimental data include surface pressure, heat
during the latter 5 ms of the test is shown in Fig. 11b. The transfer, holographic interferometry and Schlieren images. The
experimental heat transfer distribution reaches a steady state. estimated experimental uncertainty is shown in one of the five
The computed surface pressure and heat transfer for this config- sets of computed results for simplicity.
uration by Nompelis is shown in Fig. 11c and d, respectively, and
shows close agreement with experiment. The computed flowfield
reaches a steady state. 5.1. HEG I

5.1.1. Assessment of accuracy


5. Results for cylinder The computed and experimental surface pressure are presented
in Fig. 12. All five simulations display excellent agreement with
Two configurations of the cylinder were considered by the
participants (Table 2). Both flows are air with significant atomic 6
We have chosen to retain the numbering of the experimental datasets in
oxygen in the freestream. HEG I is a high enthalpy (ho ¼22.4 MJ/ agreement with the original publications of the experiment.
D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26 25

experiment. This is a particularly significant result since HEG I is the Reynolds number was reduced by an approximate factor of two.
highest enthalpy case considered in the study. Virtually identical The computed flowfield for this case was observed to be steady in
results are obtained by Nompelis (Fig. 12c) for surface accommoda- agreement with experiment, and the computed surface pressure
tion factor g ¼ 0 (non-catalytic) and g ¼ 1 (fully catalytic for radi- and heat transfer displayed close agreement with experiment.
cals). All computed flowfields converge to steady state. The predicted heat transfer and surface pressure for the low
The computed and experimental surface heat transfer are enthalpy cylinder are in good agreement with experiment. For the
presented in Fig. 13. There is a consistent underprediction of heat high enthalpy cylinder case, the predicted pressure profiles show
transfer for all simulations with the exception of the computation close agreement with experiment while the predicted heat
by Nompelis with g ¼ 1 (where g refers to the catalytic efficiency transfer underpredicts the experiment, thus indicating the need
at the wall). These results indicate that accurate modeling of for accurate modeling of the effects of surface catalycity.
surface catalysis is critical to the prediction of the peak heat
transfer for this case.

5.1.2. Flowfield structure Acknowledgments


The flowfield structure is illustrated in Fig. 14 for Case I based
upon the computations by Walpot using the Dunn and Kang The research by the first author (DK) is supported by the US Air
model. Results are shown on the symmetry plane. Fig. 14a dis- Force Office of Scientific Research under Grant FA9550-07-1-0228
plays the Mach number contours and flow streamlines. The shock managed by Dr. John Schmisseur. The assistance of Michael
wave is evident in the abrupt decrease in Mach number ahead of Holden (CUBRC) and Klaus Hannemann (DLR) is gratefully appre-
the cylinder. A very thin boundary layer forms on the cylinder ciated. Dimitris Drikakis wishes to thank A. Mosedale, S. Tissera
surface as indicated by the contours of the magnitude of the and V. Titarev for their support with regard to the numerical
velocity (Fig. 14b). The translational–rotational temperature T simulations. Andrea Lani wishes to thank Dr. Marco Panesi
shown in Fig. 14c displays a peak value within the shock layer. (former collaborator at the VKI), whose role has been fundamen-
The vibrational temperature Tv (Fig. 14d) is the same for N2, O2 tal in the implementation of thermal non-equilibrium models
and NO in this model. The difference T v T is shown in Fig. 14e. used by our flow solvers.
Downstream of the shock the translational–rotational and vibra-
tional temperatures nearly equilibrate; however, the vibrational
temperature does not adapt to the rapid expansion around the References
cylinder corner and differences T v T as large as 2500 K occur at
the shoulder of the cylinder. Significant dissociation of N2 occurs [1] Knight D, Degrez G. Shock wave boundary layer interactions in high Mach
downstream of the shock and the concentration of N remains number flows: a critical survey of current numerical prediction capabilities.
AGARD Report 319, AGARD; 1998. p. 1-1–1-35.
relatively high (Fig. 14f). [2] Knight D. RTO WG 10: Test cases for CFD validation of hypersonic flight. AIAA
Paper 2002-0433. American Institute of Aeronautics and Astronautics; 2002.
5.2. HEG III [3] Candler G. Hypersonic nozzle analysis using an excluded volume equation of
state. AIAA Paper 2005-5202. American Institute of Aeronautics and Astro-
nautics; 2005.
5.2.1. Assessment of accuracy [4] Karl S, Martinez-Schramm J, Hannemann K. High enthalpy cylinder flow in
The computed and experimental surface pressure are pre- HEG: a basis for CFD validation. AIAA Paper 2003-4252. American Institute of
Aeronautics and Astronautics; 2003.
sented in Fig. 15. All five simulations are in good agreement with [5] Vincenti W, Kruger C. Introduction to physical gas dynamics. Malabar, FL:
experiment. The predictions by Nompelis (Fig. 15c) for both non- Krieger Publishing Company; 1965.
catalytic (g ¼ 0) and catalytic (g ¼ 1) are virtually identical. The [6] Millikan R, White D. Systematics of vibrational relaxation. Journal of
Chemical Physics 1953;39:3209–13.
computed and experimental heat transfer are shown in Fig. 16. [7] Park C. Review of chemical-kinetic problems of future NASA mission: Earth
The peak heat transfer is lower than for HEG I as expected due to entries. Journal of Thermophysics and Heat Transfer 1993;7:385–98.
the lower freestream total enthalpy. All five simulations are [8] Drikakis D, Tsangaris S. An implicit characteristic flux averaging scheme for
the Euler equations for real gases. International Journal for Numerical
predicted by the experimental heat transfer profile within the
Methods in Fluids 1991;12:711–26.
experimental uncertainty. In particular, the effects of catalycity in [9] Drikakis D. Advances in turbulent flow computations using high-resolution
the computations by Nompelis (Fig. 16c) are significantly less methods. Progress in Aerospace Sciences 2003;39:405–24.
than in HEG I. [10] Drikakis D, Hahn M, Mosedale A, Thornber B. Large eddy simulation using
high resolution and high order methods. Philosophical Transactions of the
Royal Society A 2009;367:2985–97.
[11] Drikakis D, Tsangaris S. On the accuracy and efficiency of CFD methods in real
6. Conclusions gas hypersonic. International Journal for Numerical Methods in Fluids
1993;16:759–75.
[12] Tissera S, Titarev V, Drikakis D. Chemically reacting flows around a double-
The capability of CFD software to accurately predict hyperso- cone including ablation effects. AIAA Paper 2010-1285. American Institute of
nic shock wave laminar boundary layer interactions is examined. Aeronautics and Astronautics; 2010.
A series of independent computations were performed by [13] Tissera S, Drikakis D, Birch T. Computational fluid dynamics methods for
hypersonic flow around blunted-cone-cylinder-flare. Journal of Spacecraft
researchers in the US and Europe for two generic configurations and Rockets 2010;47:563–70.
(a double cone in N2 and a cylinder in air with significant [14] GASP version 4.2 user’s manual. Blacksburg, VA: Aerosoft, Inc.; 2004.
freestream atomic O) and compared with experimental data for [15] Morrison J. Flux-difference split scheme for turbulent transport equations.
AIAA Paper 1990-5251. American Institute of Aeronautics and Astronautics;
surface pressure and heat transfer. For the higher enthalpy double 1990.
cone configuration (Run 42), the computations and experiment [16] Liou M. A sequel to AUSM: AUSMþ . Journal of Computational Physics
display close agreement. Moreover, both the experiment and all 1996;129:364–82.
[17] Gaitonde D. An assessment of CFD for prediction of 2-D and 3-D high speed
computations indicate that the flowfield is steady for this case.
flows. AIAA Paper 2010-1284. American Institute of Aeronautics and Astro-
Surprisingly, all simulations predict a dramatically unsteady flow nautics; 2010.
for the low enthalpy double cone configuration (Run 40) in direct [18] MacCormack R, Candler G. The solution of the Navier–Stokes equations using
contrast to the experiment where the flowfield was observed to Gauss–Seidel line relaxation. Computers and Fluids 1989;17:135–50.
[19] Candler G, MacCormack R. The computation of hypersonic ionized flows in
achieve steady state. Consequently, an additional experiment chemical and thermal nonequilibrium. Journal of Thermophysics and Heat
(Run 80) was performed at nearly the same enthalpy but the Transfer 1991;5:266–73.
26 D. Knight et al. / Progress in Aerospace Sciences 48–49 (2012) 8–26

[20] Druguet M-C, Candler G, Nompelis I. Effect of numerics on Navier–Stokes [33] Dunn M, Kang S-W. Theoretical and experimental studies of reentry plasmas.
computations of hypersonic double-cone flows. AIAA Journal 2005;43: Contractor Report 2232. NASA; 1973.
616–23. [34] Grossman B, Cinnella P. Flux-split algorithms for flows with non-equilibrium
[21] Wright M, Bose D, Candler G. A data-parallel line relaxation method for the chemistry and vibrational relaxation. Journal of Computational Physics
Navier–Stokes equations. AIAA Journal 1998;36:1603–9. 1990;88:131–68.
[22] Nompelis I, Candler G. Numerical investigation of double-cone flow experi- [35] Chase M, Davies C, Downey J, Frurip D, McDonald R, Syverud A. JANNAF
ments with high-enthalpy effects. AIAA Paper 2010-1283. American Institute thermochemical tables. 3rd ed. American Chemical Society and American
of Aeronautics and Astronautics; 2010. Institute of Physics for the National Bureau of Standards; 1986.
[23] Schwamborn D, Gerhold T, Heinrich R. The DLR TAU-Code: recent applica- [36] Panesi M. Physical models for nonequilibrium plasma flow simulations at
tions in research and industry. In: Wesseling P, Onate E, Périaux J, editors. high speed re-entry conditions. PhD thesis. Italy: Pisa University; 2009.
ECCOMAS CFD 2006. The Netherlands: TU Delft; 2006. [37] Blottner F, Johnson M, Ellis M. Chemically reacting viscous flow program for
[24] Reimann B, Hannemann K. Numerical investigation of double-cone and multi-component gas mixtures. Technical Report SC-RR-70-754. Albuquer-
cylinder experiments in high enthalpy flows using the DLR TAU code. AIAA que, New Mexico: Sandia National Laboratories; 1971.
Paper 2010-1282. American Institute of Aeronautics and Astronautics; 2010. [38] Bottin B. Aerothermodynamic model of an inductively-coupled plasma wind
[25] Holden M. Private communication; 2009. tunnel. PhD thesis. Université de Liége; 1999.
[26] Courant R, Friedrichs KO. Supersonic flow and shock waves. 3rd ed. Springer- [39] White F. Viscous fluid flow. New York: McGraw Hill; 1974.
Verlag; 1976. [40] Barbante P. Accurate and efficient modeling of high temperature none-
[27] Gnoffo P, Gupta R, Shinn J. Conservation equations and physical models for quilibrium air flows. PhD thesis. Von Karman Institute; 2001.
hypersonic air flows in thermal and chemical nonequilibrium. Technical [41] Nompelis I. Computational study of hypersonic double-cone experiments for
Paper 2867. NASA; 1989. code validation. PhD thesis. University of Minnesota; 2004.
[28] Gupta R, Yos J, Thompson R, Lee K. A review of reaction rates and [42] Wilke C. A viscosity equation for gas mixtures. Journal of Chemical Physics
thermodynamic and transport properties for an 11-species air model for 1950;18:517–9.
chemical and thermal nonequilibrium calculations to 30 000 K. Reference [43] Magin T, Degrez G. Transport properties for partially ionized unmagnetized
Report 1232. NASA; 1990. plasmas. Physical Review E 2004;70.
[29] Park C. On convergence of computation of chemically reacting flows. AIAA [44] Magin T, Degrez G. Transport algorithms for partially ionized unmagnetized
Paper 1985-0247. American Institute of Aeronautics and Astronautics; 1985. plasmas. Journal of Computational Physics 2004;198:424–49.
[30] Park C. Assessment of two-temperature kinetic model for dissociating and [45] Hirschfelder J, Curtiss C, Bird R. Molecular theory of gases and liquids. 2nd ed.
weakly ionizing nitrogen. AIAA Paper 1986-1347. American Institute of New York: John Wiley & Sons; 1963.
Aeronautics and Astronautics; 1985. [46] Herning F, Zipperer L. Beitrag zur Berechnung der Zähigkeit technischer
[31] Park C. Two temperature interpretation of dissociation rate date for N2 and Gasgemische aus den Zähigkeitswerten der Einzelbestandteile. Gas- und
O2. AIAA Paper 1988-0458. American Institute of Aeronautics and Astro- Wasserfach 1936;79.
nautics; 1988. [47] Barbante P, Degrez G, Sarma G. Computation of non-equilibrium high-
[32] Park C. Assessment of two-temperature kinetic model for ionizing air. Journal temperature axisymmetric boundary layer flows. Journal of Thermophysics
of Thermophysics and Heat Transfer 1989;3:233–44. and Heat Transfer 2002;16:490–7.

You might also like