Sie sind auf Seite 1von 12

Probing Real Gas and Leading-Edge Bluntness

Effects on Shock Wave Boundary-Layer


Interaction at Hypersonic Speeds
Siddesh Desai 1; Shuvayan Brahmachary 2; Hrishikesh Gadgil 3; and Vinayak Kulkarni 4
Downloaded from ascelibrary.org by Nottingham Trent University on 08/09/19. Copyright ASCE. For personal use only; all rights reserved.

Abstract: The present investigations are centered on understanding the discrepancies in shock wave boundary-layer interaction (SWBLI)
for perfect and real gas laminar flows. In view of this, the in-house-developed computational fluid dynamics (CFD) solvers are integrated
with a gradient-based optimization algorithm to predict the critical radii of SWBLI in the case of perfect and real gas flows. The
developed high-fidelity approach has been observed to be useful in the precise estimation of critical radii of bluntness. Further, studies
for SWBLI revealed that real gas effects reduce the extent of separation in comparison with the perfect gas flow and also necessitate lower
magnitudes of critical radii. It has been noted that a reduced requirement of a high entropy layer thickness and upstream overpressure
region demonstrate a need for a lower value of inversion and equivalent radii for real gas flow conditions. Therefore, a larger estimate of
the equivalent radius of SWBLI, obtained for perfect gas flow conditions, or any radius larger than that would definitely provide the
necessary separation control for real gas flows. DOI: 10.1061/(ASCE)AS.1943-5525.0001085. © 2019 American Society of Civil
Engineers.
Author keywords: Shock wave boundary-layer interaction (SWBLI); Reacting flows; Critical radii; Separation control; High-fidelity
framework.

Introduction harm to various engineering systems or subsystems (Hataue 1989;


Piotrowicz and Flaszynski 2016; Kaczyński and Szwaba 2016)
Supersonic or hypersonic flows face a major impediment in shock in the event of such interactions. Such control would not only
wave boundary-layer interaction (SWBLI). Such interaction is enhance the performance of the associated components but also
evident in laminar as well as turbulent flows. A spacecraft or air- ensure the safety of flight. Therefore, it is necessary to devise
craft component that either encounters sudden turning due to ramp- techniques or methods to control SWBLIs.
like disturbances or receives a shock impingement will be affected Techniques such as injection (Inger and Zee 1978; Pasquariello
by SWBLI. The former type of interaction is known as ramp- et al. 2014) and leading-edge bluntness (Townsend 1966; Holden
induced SWBLI (R-SWBLI) and the latter as impingement-based 1971; Neuenhahn and Olivier 2009; John and Kulkarni 2014a)
SWBLI (I-SWBLI). In the presence of any of these SWBLI types, were considered in literature to control the SWBLI. Among those,
control surfaces or engine parts encounter adverse effects such as a leading-edge-bluntness-based passive separation control method
flow separation, boundary-layer transition, higher surface heating, has been noted to have an anomalous trait since the necessary sep-
or vortex shedding. Hence, it is highly desirable, first of all, to aration control is not possible for all leading-edge radii. It has been
understand the basics of such interactions (Marini 2001; John found that the initial increment in the leading-edge radius, provided
et al. 2014; Glushneva et al. 2015), predict the interaction param- with the intention of separation control, actually enlarges the length
eters (Chapman et al. 1957; Needham and Stollery 1966; Katzer of the separation bubble size. As a consequence of this, the extent
1989; John and Kulkarni 2014b), and then control or lessen the of separation reaches its peak at a specific (inversion) radius. Fur-
ther increase of the leading-edge bluntness decreases the separation
bubble size. But there still needs to be a critical of bluntness value
1 (equivalent radius) so as to experience the separation bubble as that
Research Scholar, Dept. of Mechanical Engineering, Indian Institute of
Technology Guwahati, Guwahati, Assam 781039, India. Email: siddesh@
of a sharp leading-edge configuration. Thus, there are two special
iitg.ac.in or critical radii, an inversion radius and an equivalent radius, and
2
Research Scholar, Dept. of Mechanical Engineering, Indian Institute their values should be identified in advance. Prediction of the
of Technology Guwahati, Guwahati, Assam 781039, India. Email: inversion radius is essential since it corresponds to the maximum
b.shuvayan@iitg.ac.in separation bubble size. It is recommended to know the equivalent
3
Assistant Professor, Dept. of Aerospace Engineering, Indian Institute radius since any radius of magnitude higher than this would only be
of Technology Bombay, Mumbai, Maharashtra 400076, India. Email: considered for separation control.
gadgil@aero.iitb.ac.in In view of the aforementioned requirements, present studies fo-
4
Associate Professor, Dept. of Mechanical Engineering, Indian Institute cus on the development of a high-fidelity methodology for estimat-
of Technology Guwahati, Guwahati, Assam 781039, India (corresponding
ing the critical radii of R-SWBLI for perfect gas and high-enthalpy
author). Email: vinayak@iitg.ac.in; vinayak@iitg.ernet.in
Note. This manuscript was submitted on June 15, 2018; approved on flows. Such detailed investigations are essential since some of the
May 28, 2019; published online on August 8, 2019. Discussion period open findings reported in the literature either deal only with the inversion
until January 8, 2020; separate discussions must be submitted for individual radius (Holden 1971; Neuenhahn and Olivier 2009) or examine
papers. This paper is part of the Journal of Aerospace Engineering, both radii but without considering the standard optimization
© ASCE, ISSN 0893-1321. techniques (John and Kulkarni 2014a). As an outcome of those

© ASCE 04019089-1 J. Aerosp. Eng.

J. Aerosp. Eng., 2019, 32(6): 04019089


2 3 2 3 2 3
studies, a wide range of radii would be recommended between ρ ρu ρv
which it can be expected to obtain a critical radius. Hence, it is 6 7 6 7 6 7
6 ρu 7 6 ρu2 þ p 7 6 ρuv 7
better to integrate the computational fluid dynamics (CFD) solver 6 7 6 7 6 7
6 7 6 7 6 7
with an optimization procedure to arrive at the precise value of 6 ρv 7 6 ρuv 7 6 ρv2 þ p 7
6 7 6 7 6 7
critical radii or to narrow their range of prediction. Further, there 6 7 6 7 6 7
6 ρE 7 6 ðρE þ pÞu 7 6 ðρE þ pÞv 7
have been suggestions about the prediction of critical radii for a U¼6 7
6 C 7;
6
EI ¼ 6 7;
7
6
FI ¼ 6 7;
7
given freestream condition and ramp angle (John and Kulkarni 6 17 6 uC1 7 6 vC1 7
6 7 6 7 6 7
2014a). As per this suggestion, the case in which the blunt 6 · 7 6 7 6 7
6 7 6 · 7 6 · 7
leading-edge value equals the inversion radius, the thickness 6 7 6 7 6 7
6 · 7 6 · 7 6 · 7
of the entropy layer becomes the same as that of the boundary 4 5 4 5 4 5
layer around the upstream influence location. Also, the upstream C4 uC4 vC4
overpressure region or favorable pressure gradient reaches the 2 3
upstream influence location when the leading-edge radius be- 0
Downloaded from ascelibrary.org by Nottingham Trent University on 08/09/19. Copyright ASCE. For personal use only; all rights reserved.

6 7
comes equal to the equivalent radius. With this information, it 6 τ xx 7
6 7
is still difficult to predict the critical radii for a given hypersonic 6 7
6 τ 7
flow and plate-ramp configuration, since this would require in- 6 xy
PN 7
6 7
formation on the entire flow field. Further, these observations are 6 uτ xx þ vτ xy − qx − i¼1 hi Ci ūi 7
6
Ev ¼ 6 7
made specifically in connection with a perfect gas flow only and
−C1 ū1 7
one needs to consider high-enthalpy flow condition. Thus, one of 6 7
6 7
the contributions of the present investigations is to cross-check 6 7
6 · 7
the validity of indication about the flow field around critical radii 6 7
6 · 7
for a high-enthalpy flow. This process also puts forth the obser- 4 5
vation of a real gas effect on SWBLI. The effect of enthalpy has −C4 ū4
been studied in the literature (John and Kulkarni 2017), but real 2 3 2 3
0 0
gas effects have not. Therefore, a methodology that is suitable 6 7 6 7
for low- and high-enthalpy flows is not only developed to obtain 6 τ xy 7 607
6 7 6 7
estimates of critical radii of SWBLI but also to verify the resem- 6 7 6 7
6 τ 7 607
blance in flow features around critical radii in either case. In 6 yy
PN 7 6 7
6 7 6 7
view of these objectives, a high-temperature real gas flow solver, 6 uτ xy þ vτ yy − qy − i¼1 hi Ci v̄i 7 607
developed in house and integrated with an optimization pro- Fv ¼ 6
6
7;
7 S ¼ −6
6S 7
7
6 −C1 v̄1 7 6 17
cedure (CFD-O), is used to formalize the philosophy of critical 6 7 6 7
6 7 6 · 7
radius prediction for R-SWBLI. Details of the optimization 6 · 7 6 7
6 7 6 7
procedure, solver, and results are presented in the following 6 · 7 6 · 7
4 5 4 5
sections.
−C4 v̄4 S4

Computational Strategy for CFD-O The mixture pressure is evaluated using the sum of partial
pressures of individual species:

Finite-Volume Formulation XN
Ci
p ¼ Ru T ð2Þ
Estimation of the critical radii of bluntness for SWBLI for per- i¼1
MW i
fect and real gas flow situations is a major motivation of the cur-
rent investigations. To demonstrate the high-enthalpy effects in where p = pressure; Ru = universal gas constant; T = temperature;
SWBLI, two different versions of a laminar flow solver [Unstruc- N = number of species; and Ci and MW i = mass concentration and
tured Solver for Hypersonic Aerothermodynamic Simulations molecular weight of ith species, respectively. The temperature in
(USHAS)] developed in house are employed herein. The first the equation is determined by applying the Newton–Raphson
version of the solver is based on the perfect gas assumption, method iteratively to the internal energy:
which means the specific heat and specific heat ratio of gas remain
XN Z 
constant with temperature and solves only Navier–Stokes equa- Ci T
tions (John and Kulkarni 2013, 2014a; John et al. 2016; John and ρe ¼ Cpi dT þ h0fi − p ð3Þ
i¼1
MW i TR
Kulkarni 2017). Another version considers fully coupled Navier–
Stokes and species continuity equations to deal with real gas where ρ = density; e = energy; T R = reference temperature; and Cpi
effects (Desai et al. 2016, 2017; Agarwal et al. 2017). High- and h0fi = specific heat at constant pressure and heat of formation of
enthalpy simulations using a similar methodology are reported species i. Here, species continuity equations provide the production
in the literature (Lee 1999; Ess and Allen 2005). An axisymmetric rate of various species. The required reaction kinetics data are in-
form of the governing equations, as shown in Eq. (1), is used to corporated into the solver, as mentioned by Dunn and Kang (1973).
develop the present solver: This solver also accounts for the temperature dependence of vari-
ous properties like specific heat, thermal conductivity, and viscosity
for each species, as elaborated by Gordon and Mcbride (1994). In
∂U ∂EI ∂FI ∂Ev ∂Fv this numerical study with a perfect gas flow solver or real gas
þ þ þS¼ þ ð1Þ
∂t ∂x ∂y ∂x ∂y flow solver, convective fluxes are obtained using an advection up-
stream splitting method (AUSM) (Liou and Steffen 1993) scheme,
whereas viscous fluxes are calculated using face values of flow
where quantities and their first derivatives, as detailed by Blazek

© ASCE 04019089-2 J. Aerosp. Eng.

J. Aerosp. Eng., 2019, 32(6): 04019089


2 3
(2001). Further, Venkatakrishnan’s limiter (1993) is equipped to 0 0 0 0 0 0 0 0
ensure a second-order monotone solution. For the present study, 6 7
6 0 0 0 0 0 0 0 0 7
a CFL of 0.1 is used. 6 7
6 7
6 0 0 0 0 0 0 0 0 7
6 7
6 0 0 0 0 0 0 0 0 7
Chemical Kinetics 6 7
6 ∂S ∂S1 ∂S1 ∂S1 ∂S1 ∂S1 ∂S1 ∂S1 7
6 1 7
The process of estimation of the production rate for any species is ∂S 66
7
∂C4 7
explained in this section. Here, N R reversible reactions considering ¼ 6 ∂ρ ∂ðρuÞ ∂ðρvÞ ∂ðρEÞ ∂C1 ∂C2 ∂C3
7
∂U 6 ∂S ∂S2 ∂S2 ∂S2 ∂S2 ∂S2 ∂S2 ∂S2 7
N species are expressed as 6 2 7
6 7
6 ∂ρ ∂ðρuÞ ∂ðρvÞ ∂ðρEÞ ∂C1 ∂C2 ∂C3 ∂C4 7
X
N X
N 6 7
6 ∂S3 ∂S3 ∂S3 ∂S3 ∂S3 ∂S3 ∂S3 ∂S3 7
ν ij0 nj ⇌ ν ij0 0 nj ð4Þ 6 7
6 ∂ρ ∂ðρuÞ ∂ðρvÞ ∂ðρEÞ ∂C1 ∂C2 ∂C3 ∂C4 7
j¼1 j¼1 6 7
6 7
Downloaded from ascelibrary.org by Nottingham Trent University on 08/09/19. Copyright ASCE. For personal use only; all rights reserved.

4 ∂S4 ∂S4 ∂S4 ∂S4 ∂S4 ∂S4 ∂S4 ∂S4 5


where i ¼ 1; 2; : : : ; NR ; ν ij0 and ν ij0 0 are the stoichiometric coeffi-
cients for species j appearing as a reactant in the ith forward and ∂ρ ∂ðρuÞ ∂ðρvÞ ∂ðρEÞ ∂C1 ∂C2 ∂C3 ∂C4
C
backward reactions, respectively; and nj ¼ MWj j is the molar con- ð8Þ
centration for species j. The Arrhenius rate expression is employed
for the rate constant of a reaction i as The elements in the Jacobian matrix can be evaluated via the
−Ei
formula
k i ¼ A i T m i e Ru T ð5Þ     N
∂Sj  XNR
K fi Efi Y v0
00 0
¼ Wj ðvij − vij Þ mfi þ nl il
Here, Ei represents the activation energy, while Ai and mi are ∂T Ci;i¼1 : : : ;N−1 i¼1
T R u T l¼1
constants. A 5-species (N 2 , O2 , N, O, NO) model involving 11  Y 
N
elementary reactions is used for all the simulations. The rate of K E v00
− bi mbi þ bi n il ð9Þ
change of molar concentration of species j is then given by T Ru T l¼1 l
summing up the changes due to all reaction steps:
 
0 1 ∂Sj  u ∂Sj 
XNR Y Y ¼ − ð10Þ
∂ρuρ;ρE;Ci;i¼1 : : : ;N−1 ρCv ∂T Ci;i¼1 : : : ;N−1
N 0
N 00
ν ν
Sj ¼ MW j ðν ij0 0 − ν ij0 Þ@kfi nl il − kbi nl il A ð6Þ
i¼1 l¼1 l¼1  
∂Sj  1 ∂Sj 
¼
where kfi and kbi are the forward and backward rate constants ∂ρEρ;ρE;Ci;i¼1 : : : ;N−1 ρCv ∂T Ci;i¼1 : : : ;N−1
for reaction i. The reactions and the corresponding rate constants
XNR   0
are considered as mentioned by Dunn and Kang (1973) and Desai ∂Sj v K fk Y
N
v0
jρ;ρE;Ci;i¼1 : : : ;N−1 ¼ W j ðvki0 0 − vki0 Þ ki nl kl
et al. (2016). Here, treatment of chemical source terms is done ∂Ci k¼1
Ci l¼1
implicitly. Reaction rates remain as small as the flow time scale 
for high-enthalpy flows. Hence, very little time is required for v 0 0K Y N
v00 e − eN ∂Sj
− ki bk nl kl − i j
explicit simulations to simulate the aerothermodynamics of a flow Ci l¼1 ρCv ∂T Ci;i¼1 : : : ;N−1
field efficiently. Therefore, only the source term implicitization is
adopted to accelerate the computations (Bussing and Murman ð11Þ
1988). Further, discretization of Eq. (1) is done using a finite-
volume method, which leads to Eq. (7):
Z Solver Validation
Δt X
n f
SΔt
U nþ1 ¼ Un − ððEI þ FI Þ − ðEv þ Fv ÞÞ · ds þ ∂S
A real gas flow solver is first used for simulating the low-enthalpy
V i¼1 Si I − Δt ∂U freestream hypersonic flow over a plate-ramp configuration with
ð7Þ different leading-edge radii as considered by John and Kulkarni
(2014a) in their perfect gas flow simulation. A typical computa-
∂S tional domain and the associated boundary conditions are shown
where I = identity matrix; and ∂U = Jacobian matrix. Jacobian
∂S in Fig. 1. The considered freestream conditions employed for this
matrix ∂U can be evaluated using the expression
part of the study are Mach number 6.0, static temperature 131.7 K,
static pressure 199.4345 Pa, and wall temperature 300 K. Finally, a
mesh with a total number of cells close to 40,000 is considered for
Supersonic Outlet

simulations, where the grid remains clustered in the crucial areas


Supersonic Inlet
like the leading edge, ramp plate junction, and so forth. In elaborate
mesh independence studies, it has also been noticed that the first
Rb cell height from the wall boundary should be around 3 × 10−5 m.
Thus, the obtained skin friction coefficient and pressure coefficient
plots for a leading-edge radius of 0.3 mm are shown in Fig. 2. It is
evident here that the real gas effects are not prominent, and there-
fore the results reported for perfect gas flow show an encouraging
Symmetry Adiabatic / Isothermal wall match with the present real gas flow. Skin friction variation can be
used to predict the separation length. This prediction of separation
Fig. 1. Schematic diagram of computational domain and associated
length for sharp as well as blunt leading-edge cases are the same in
boundary conditions.
magnitude for perfect and real gas flows (Table 1). Thus, from this

© ASCE 04019089-3 J. Aerosp. Eng.

J. Aerosp. Eng., 2019, 32(6): 04019089


−3
x 10
14 8

Real gas solver Real gas solver


12 7
John and Kulkarni, 2014 John and Kulkarni, 2014

Skin friction Coefficient (C )


f
10 6

8 5

P/ (P )
6 4

4 3

2 2
Downloaded from ascelibrary.org by Nottingham Trent University on 08/09/19. Copyright ASCE. For personal use only; all rights reserved.

0 1

−2 0
0 0.02 0.04 0.06 0.08 0.1 0.12 0 0.02 0.04 0.06 0.08 0.1 0.12
(a) x (m) (b) x (m)

Fig. 2. Comparison of (a) skin friction coefficient Cf ; and (b) surface pressure distribution P=P∞ with numerical results of John and Kulkarni (2014a).

Table 1. Details of result for validation of present real gas solver into 50,400 quadrilateral cells. The flow is simulated using a com-
Separation Reattachment Plateau
putational approach similar to that described earlier. In this case, the
postshock temperature rise is enough to cause specific heat varia-
Case Point (mm) Point (mm) Pressure (Pa) tion; however, it is insufficient to activate dissociation reactions.
0.3 mm The obtained plots of surface pressure and heat flux distribution
Perfect gas 38.10 67.18 505.27 are plotted in Fig. 3. This figure clearly shows that the results ob-
Real gas 37.60 67.50 504.07 tained from the present simulation agree well with results reported
0.5 mm in the literature. Thus, the solver is validated for the high-enthalpy
Perfect gas 38.28 68.84 527.41 condition as well. Further, a detailed validation of the solver for
Real gas 38.00 69.20 526.60
reacting flow problems, including an unsteady shock tube test case
and steady flow past a sphere, are given in Desai et al. (2016, 2017).

test of total temperature (1,080 K) and the isothermal wall boun-


dary condition (300 K), it is clear that the absence of chemical re- Integration of CFD Solver and Gradient Method
actions and the least variation of specific heat make no difference in A standard optimization procedure yields significant leverage by
SWBLI for perfect and real gas flows. maximizing or minimizing certain objective functions or generating
After successful validation in low-enthalpy conditions feasible solutions for hard combinatorial optimization problems;
(0.786 MJ=kg), the solver is further tested in the high-enthalpy however, it is also plagued by limitations. One limitation is the
case (1.98 MJ=kg) of flow past a cylinder of radius 0.0381 m. computational time involved in the process from initial guess to
For testing the solver in high enthalpy case, the free stream con- final optimal solution, especially when the optimization technique
ditions are Mach number 6.47, static temperature 241.5 K, static employed relies on CFD for the function evaluations (Thévenin and
pressure 627.5 Pa, and wall temperature 278 K, as mentioned in Janiga 2008). This is especially an issue when evolutionary algo-
the literature (Wieting 1987). The computational domain is divided rithms, for example, genetic algorithms (GAs) (Deb et al. 2000),

1.4 1.4
Real gas flow solver Real gas flow solver
1.2 Experimental−Wieting, 1987 1.2 Experimental−Wieting, 1987

1 1

0.8 0.8
0

P/P0
Q/Q

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 20 40 60 80 0 20 40 60 80
(a) θ (b) θ

Fig. 3. Comparison of (a) surface heat flux Q=Q0 ; and (b) surface pressure distribution P=P0 with experimental results of Wieting (1987).

© ASCE 04019089-4 J. Aerosp. Eng.

J. Aerosp. Eng., 2019, 32(6): 04019089


Initial Geometry xk = 1 Initial design point
k=0
R0b

Function evaluation f(x k) & f(xk+ε) ; ε = 0.001


Generate / Update grid

k=k+1 Met convergence criteria? Stop Function evaluation (CFD)


YES k
f(Rb )

NO
Gradient evaluation
Determine λk k=k+1
Downloaded from ascelibrary.org by Nottingham Trent University on 08/09/19. Copyright ASCE. For personal use only; all rights reserved.

Gradient method
k+1 k κ k
xk+1 = xk ± λkgrad. f(xk) R b =R b ± λ grad. f(Rb )

Fig. 4. Flowchart describing genetic algorithm. Convergence met?


NO YES
LR=LR , max or Stop
b b
-03
LR
b = 0 -L R b≤ 10

shuffled frog-leaping algorithm (SFLA) (Eusuff and Lansey 2007),


and simulated annealing (SA) (Kirkpatrick et al. 1983), are used Fig. 5. Schematic behind algorithm used to couple in-house CFD
owing to their inherent requirement of a large number of function solvers and gradient method technique.
evaluations until convergence. This was conclusively shown in a
study by Brahmachary et al. (2017), who compared the conver-
gence rate of GA, SFLA, and SA against gradient-based methods
objective function for the prediction of Rb;inv is LRb ¼ LRb ;max ,
(GMs) for a typical unimodal shape optimization problem. It was
where LRb is the separation length of the bubble formed at the
further shown that the convergence rate of GMs can be significantly
ramp foot, with subscript Rb signifying the corresponding radius
improved by altering the adaptive step function. Therefore, a
of the blunt leading edge. In the case of the equivalent radius
greedy algorithm, such as steepest descent or steepest ascent, that
Rb;eq , the separation bubble size is the same as that of the sharp
relies on the gradient of the cost function is employed for the
leading edge. Hence, the objective function in that case is such that
present investigations. This section presents a discussion of the
various segments that are an integral part of such an optimization jLRb ¼0 − LRb j <¼ 10−03 . Further, it is known that the inversion ra-
framework. In particular, it highlights the objective function for dius is smaller than the equivalent radius. Therefore, the separation
the present study, the optimization algorithm employed, and the details of the sharp-leading-edge case are used to start the optimi-
flow solver used to evaluate the cost function. zation procedure for predicting the inversion radius in either cases.
A GM is chosen here for integration with in-house CFD solvers. But any radius greater than the inversion radius can be thought of as
A flow chart depicting the algorithm behind the optimization meth- an initial guess for the prediction of the equivalent radius in both
ods is shown in Fig. 4. For the GM-based optimization technique, cases. Fig. 5 shows a schematic of the algorithm considered to
first an initial guess is made and then its sensitivities are obtained couple the CFD code with the GM procedure. It must be noted that
using finite forward difference approximation. Each optimization the flow solver takes the most computational time within a design
cycle consists of two function evaluations, one each for the gradient cycle, whereas the optimization steps themselves are negligible
estimation, ∇fðxk Þ, and the perturbed decision variable, xkþ1 . The compared to the cost function evaluations. For the present case,
perturbed decision variable is obtained as shown in Fig. 4, where λk inviscid simulations take around 4 h to attain steady-state solutions,
is the adaptive step function at a particular optimization iteration, k, whereas the high-enthalpy laminar flows take close to 30 h.
used to control the rate of convergence, and is defined as
 k−1
Aλ ; iffðxÞk < fðxÞk−1 Results and Discussion
λ ¼
k
Bλk−1 ; otherwise
Freestream conditions used to study the real gas effects in
where A and B are constants whose values depend on the test case R-SWBLI are given in Table 2. These conditions are given as inlet
at hand, with the inequality sign reversed for steepest ascent. These boundary conditions for both perfect gas and real gas flow solvers
constants are used to accelerate the convergence toward a final and integrate with the aforementioned optimization procedure.
optimal solution, and for the present study their values equal 1.75 An adiabatic wall boundary condition is employed specifically
and 0.5, respectively. to enhance the high-temperature effects for these freestream con-
The major objective of this study is to predict the critical radii of ditions at a total temperature of 3,000 K. As an outcome of this,
bluntness for R-SWBLI using perfect gas and real gas flow solvers
to understand the real gas effect. The study is conducted by evalu-
ating two critical radii Rb , and hence Rb forms the decision variable Table 2. Freestream conditions for present study
for this test case with bounds [0,3] mm. Basic sharp leading-edge Parameter Freestream condition
geometry, for which leading-edge-bluntness-based separation con-
trol is to be investigated, is the same as discussed in the previous M∞ 6
section. One of the critical radii, the inversion radius, Rb;inv , cor- T∞ 365.85 (K)
P∞ 406.63 (Pa)
responds to the maximum extent of separation. Therefore, the first

© ASCE 04019089-5 J. Aerosp. Eng.

J. Aerosp. Eng., 2019, 32(6): 04019089


Table 3. Grid specifications the separation length also appears to be constant for different grids.
Serial number Δymin (m) No. of control volumes Hence, Grid 3 was chosen for further study.
Starting with the case of a sharp leading edge as an initial guess,
Grid 1 1 × 10−4 10,800
optimization studies are performed for predicting the inversion ra-
Grid 2 5 × 10−5 28,800
Grid 3 3 × 10−5 32,040
dius. Results obtained at intermediate states are used to analyze the
Grid 4 2 × 10−5 52,190 real gas effects. Fig. 8 shows the separating streamline and sepa-
ration bubble and clearly depicts the higher separation length for
perfect gas flow conditions. Variations in specific heat and the pos-
sibility of chemical reactions lessen the maximum temperature in
for the sharp-leading-edge case itself, 0.0381 mm (approximately the domain for the real gas flow case. The viscosity of the fluid is
0.04 mm) and 0.0308 mm (approximately 0.03 mm) separation also less, which in turn leads to a smaller boundary-layer thickness.
bubble sizes are noted for perfect and real gas flow simulations, Therefore, the extent of separation is expected to be less with a real
respectively. It is crucial to make sure that the solutions obtained gas effect (Grasso et al. 2001; Deepak et al. 2010). The same facts
Downloaded from ascelibrary.org by Nottingham Trent University on 08/09/19. Copyright ASCE. For personal use only; all rights reserved.

are indeed grid independent, especially for derived quantities like are illustrated for various blunt-leading-edge cases in Fig. 9. One of
wall distribution of the skin-friction coefficient. With that motive, a the major objectives of this study is to highlight and quantify this
thorough grid-independence study is conducted using four different effect for R-SWBLI. Therefore, the skin-friction coefficient was
grids with varying control volumes and first cell thickness normal obtained as an outcome of CFD simulations for each optimization
to the wall boundary. Table 3 shows the grid specifications used for cycle and is shown in Fig. 10 for two sample cases, along with the
this study. Figs. 6 and 7 show the surface distribution of coefficient reference sharp-leading-edge case. From Fig. 10, the separation
pressure (Cp ) and skin-friction coefficient (Cf ) for perfect and real bubble size is estimated for different radii. It is evident from the
gas flow solvers, respectively. It can be observed that both Cp and figure that the prediction of the separation bubble size is higher
Cf distributions are invariant with grid spacing. More importantly, with perfect gas flow simulations. As mentioned earlier, the lower

0.25 0.04
Grid 1
Grid 2
Grid 1 0.035 Grid 3
Grid 2 Grid 4
0.2 Grid 3
Skin friction coefficient (Cf )

Grid 4
0.03
Pressure coefficient (Cp )

−3
x 10
1
0.025
0.15
0.02 0

0.1 0.015 −1

0.01
−2
0.04 0.06 0.08
0.05 0.005
Zoomed view
0
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0 0.02 0.04 0.06 0.08 0.1 0.12
(a) x (m) (b) x (m)

Fig. 6. Grid-independent results of perfect gas solver (Rb ¼ 0.0 mm): (a) pressure distribution; and (b) skin friction distribution.

0.25 0.04
Grid 1
Grid 2
Grid 1 0.035 Grid 3
Grid 2 Grid 4
0.2 −3
Skin friction coefficient (Cf )

Grid 3 0.03 x 10
1
Pressure coefficient (Cp )

Grid 4

0.025
0
0.15
0.02
−1

0.1 0.015
−2
0.04 0.06 0.08
0.01

0.05 0.005
Zoomed view

0
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0 0.02 0.04 0.06 0.08 0.1 0.12
(a) x (m) (b) x (m)

Fig. 7. Grid-independent results of real gas solver (Rb ¼ 0.0 mm): (a) pressure distribution; and (b) skin friction distribution.

© ASCE 04019089-6 J. Aerosp. Eng.

J. Aerosp. Eng., 2019, 32(6): 04019089


−3
x 10

12
Perfect gas flow (R b−0.0 mm)
Real gas flow (R b−0.0 mm)
10
Perfect gas flow (R b−0.726 mm)
Real gas flow (R −0.726 mm)

Skin friction coefficient (Cf )


8 b
Perfect gas flow (R b−1.7 mm)
6 Real gas flow (R −1.7 mm)
b

2
Downloaded from ascelibrary.org by Nottingham Trent University on 08/09/19. Copyright ASCE. For personal use only; all rights reserved.

−2

Fig. 8. Temperature contour for leading-edge radius, Rb ¼ 0.626 mm,


for perfect gas and real gas flow simulations. −4
0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.11
x (m)

Fig. 10. Skin-friction coefficient for sample leading-edge radii.


increment of viscosity due to the rise of specific heat and endother-
mic chemical reactions are taken into account for a lower separation
length in real gas flow simulations. Variation of the surface pressure
coefficient Cp distribution along a wall surface is shown in Fig. 11.
Observations like the shorter separation length, lesser upstream in- A detailed flow-field analysis at the respective critical radii can
fluence, lower plateau pressure, and so forth in the case of real gas help in understanding the difference between the results obtained
flow can also be made in connection with pressure variation. from both solvers. Initially, entropy-layer and boundary-layer inter-
The separation length obtained for each intermediate radius is action is considered to examine the inversion radius. Here as well, a
plotted as a result of the complete optimization procedure shown in high entropy layer (HEL) is defined using the streamline that passes
Fig. 12. A notable observation here is the lower value of the inver- through the edge of the curved bow shock. The height of this
sion radius with real gas effects. The optimization procedure is then streamline at the leading edge is taken to be twice that of the
continued for the equivalent radius with both solvers. The thus ob- leading-edge radius (John and Kulkarni 2014a). The thickness
tained separation bubble size for intermediate and final radii is also of HEL is 1.68 mm for the perfect gas flow case at the leading edge,
given in Fig. 12. As per the definition of equivalent radius, it was while that for the reacting gas flow is 1.562 mm at the same loca-
necessary to match the separation bubble size in the sharp-leading- tion. This difference is mainly due to lower shock stand-off distance
edge case. At this reference state itself, the real gas solver showed a in the case of real gas flow. A decrement of the shock stand-off
lower separation than the perfect gas flow solver, so the extent of distance for reacting flow in comparison with the perfect gas flow
separation at the equivalent radius from the real gas flow solver is was reported earlier (Desai et al. 2016) as well. Velocity variation
lower than the perfect gas flow solver. Therefore, the radius at normal to the wall at a reference location for different radii is plot-
which the equivalent separation is noticed is also lower in the case ted in Figs. 13 and 14 for both cases. This figure includes variation
of real gas flow simulations. for inviscid and viscous simulations to estimate the boundary-layer

Viscous perfect gas flow (R −0.0 mm)


b
Viscous real gas flow (R −0.0 mm)
b
Viscous perfect gas flow (R −0.726 mm)
b
Viscous real gas flow (R b−0.726 mm)
Viscous perfect gas flow (Rb−1.7 mm)
Viscous real gas flow (R −1.7 mm)
b

0.012 0.012 0.012

0.01 0.01 0.01

0.008 0.008 0.008


Y (m)

0.006 0.006 0.006

0.004 0.004 0.004

0.002 0.002 0.002

0 0 0
0 0.5 1 1.5 2 0 2 4 6 8 2 4 6 8
2
/ T/T Viscosity (m /s) x 10−5

Fig. 9. Variation of different properties in boundary layer 42 mm downstream of leading edge for various leading-edge radii cases.

© ASCE 04019089-7 J. Aerosp. Eng.

J. Aerosp. Eng., 2019, 32(6): 04019089


0.3
thickness for flow with a streamwise pressure gradient. The asymp-
Perfect gas flow (R b−0.0 mm)
totic match between the inviscid and viscous velocity profiles gives
Real gas flow (R −0.0 mm) the thickness of the boundary layer in this figure. The edge of the
b
0.25 Perfect gas flow (R b−0.726 mm) HEL at the corresponding location is also shown in the same figure.
Real gas flow (R −0.726 mm) Therefore, Figs. 13 and 14 clearly show that the boundary-layer
Pressure coefficient (Cp )

b
Perfect gas flow (R −1.7 mm) thickness is lower than the HEL at the location upstream of the
0.2 b
Real gas flow (R −1.7 mm) upstream influence location. However, for a leading-edge bluntness
b
radius less than the inversion radius (Rb ¼ 0.626 mm), the boun-
0.15
dary layer remains thicker than the HEL, as shown in Figs. 15 and
16. Further, the thickness of the boundary layer approaches the
thickness of the HEL around the upstream influence location for
0.1 the inversion radius. The matching of the boundary-layer and
HEL thickness at the reference upstream influence station has been
Downloaded from ascelibrary.org by Nottingham Trent University on 08/09/19. Copyright ASCE. For personal use only; all rights reserved.

reported as an indication of the leading-edge radius as the inversion


0.05
radius in the case of perfect gas flow (John and Kulkarni 2014a).
But the present study reasserts this truth for real gas flow as well.
Therefore, the indication of the match of the HEL and boundary-
0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.11 layer thicknesses is noted to be a characteristic trait of the inversion
x (m)
radius. It must be remarked here that the inversion radius is slightly
Fig. 11. Surface pressure distribution for sample leading-edge radii. smaller for real gas flow than perfect gas flow. The lower shock
stand-off distance in the case of real gas flow is accounted for
in this observation.
80 For radii lower than the inversion radius, the entropy layer is
Real gas solver
thinner (Figs. 13 and 14); thus, it leads to a smaller HEL thickness
Perfect gas solver than the thickness of the boundary layer at the SWBLI station.
70 But beyond the inversion radius, the HEL remains thicker than
Rb,inv = 0.84 mm the boundary layer around and upstream of the SWBLI (Figs. 17
and 18). This type of thicker HEL above the boundary layer re-
Separation length, L Rb (mm)

60 mains a source of streamwise vorticity as per Crocco’s theorem.


Further, every elemental increment in leading-edge bluntness leads
to a thicker HEL. For any radius between the inversion and equiv-
50 Rb,inv = 0.781 mm alent radii, a thick HEL does exist but its support is insufficient for
reducing separation since the surface pressure variation has a small
overpressure region. A blast wave theory (BWT)–based prediction
40 of wall pressure for different radii is shown in Figs. 19(a and b).
Rb,eq. = 1.95 mm It is evident in this figure that the size of the overpressure region or
favorable pressure gradient increases with increases in the leading-
30
edge bluntness. The difference in wall pressure variation is consid-
Rb,eq.= 1.85 mm ered negligible for both equivalent radii (1.85 and 1.95 mm).
But the favorable pressure gradient for both radii reaches the up-
20 stream influence location corresponding to the sharp-leading-edge
0.0 0.5 1.0 1.5 2.0 2.5 3.0 case. Hence, a favorable pressure gradient provides the critical
Leading edge radius, Rb (mm) support of a thick HEL, that equates the separation length to that
of the sharpŁ leading-edge case. In addition, a small difference in
Fig. 12. Separation bubble length for different leading-edge radius. the reference separation sizes is expected to yield a marginal

Viscous perfect gas flow (R b−0.84 mm) High entropy layer edge
Inviscid perfect gas flow (R b−0.84 mm)
Boundary layer profile 0.015 0.015

0.01 0.01 0.01


M
Y (m)

0.005 0.005 0.005

0 0 0
−0.5 0 0.5 1 −0.5 0 0.5 1 −0.5 0 0.5 1
U/U U/U U/U
Ramp
Blunt leading edge
x=0.012 m x=0.022 m x=0.042 m
Separation bubble

Fig. 13. Variation of boundary-layer and entropy-layer thickness at three different locations for perfect gas flow with inversion radius (upstream
influence location 0.022 m).

© ASCE 04019089-8 J. Aerosp. Eng.

J. Aerosp. Eng., 2019, 32(6): 04019089


Viscous real gas flow (Rb−0.781 mm) High entropy layer edge
Inviscid real gas flow (R b−0.781 mm)
Boundary layer profile 0.015 0.015

0.01 0.01 0.01


M∞

Y (m)
0.005 0.005 0.005

0 0 0
−0.5 0 0.5 1 −0.5 0 0.5 1 −0.5 0 0.5 1
U/U∞ U/U∞ U/U∞ Ramp
Blunt leading edge
x=0.012 m x=0.022 m x=0.042 m
Separation bubble
Downloaded from ascelibrary.org by Nottingham Trent University on 08/09/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 14. Variation of boundary-layer and entropy-layer thickness at three different locations for real gas flow with inversion radius (upstream
influence location 0.022 m).

Viscous perfect gas flow (R b−0.626 mm)


High entropy layer edge
Inviscid perfect gas flow (R b−0.626 mm)
Boundary layer profile 0.015 0.015

0.01 0.01
M∞ 0.01
Y (m)

0.005 0.005 0.005

0 0 0
−0.5 0 0.5 1 −0.5 0 0.5 1 −0.5 0 0.5 1
U/U U/U U/U Ramp
∞ ∞ ∞
Blunt leading edge
x= 0.012 m x=0.022 m x=0.042 m
Separation bubble

Fig. 15. Variation of boundary-layer and entropy-layer thickness at three different locations for perfect gas flow with Rb ¼ 0.626 mm leading-edge
radius (upstream influence location 0.022 m).

Viscous real gas flow (R b−0.626 mm)


High entropy layer edge
Inviscid real gas flow (R b−0.626 mm)
Boundary layer profile 0.015 0.015

0.01 0.01
M∞ 0.01
Y (m)

0.005 0.005 0.005

0 0 0
−0.5 0 0.5 1 −0.5 0 0.5 1 −0.5 0 0.5 1
U/U∞ U/U U/U∞ Ramp

Blunt leading edge
x=0.012 m x= 0.022 m x= 0.042 m
Separation bubble

Fig. 16. Variation of boundary-layer and entropy-layer thickness at three different locations for real gas flow with Rb ¼ 0.626 mm leading-edge
radius (upstream influence location 0.022 m).

difference in the corresponding equivalent radius. But the need for a critical radii through an optimization process. Further, an optimi-
favorable pressure gradient at the reference upstream influence zation strategy was developed to integrate a gradient-based optimi-
location remains a valid basis for choosing leading-edge bluntness zation technique and CFD solvers developed in-house for perfect
as the equivalent radius for perfect and real gas flows. and real gas flows. As an outcome of this integration, inversion and
equivalent radii are found to be, respectively, 0.84 and 1.95 mm for
perfect gas flow and 0.781 and 1.95 mm for real gas flow. Real gas
Conclusions effects are shown to be responsible for this decrement in the sep-
aration size and thus lower leading-edge bluntness for separation
This article investigated shock wave boundary layer interaction control. Further, it is proposed that the inversion radius should have
control using leading-edge bluntness for the prediction of two a thicker high entropy layer around the upstream influence location

© ASCE 04019089-9 J. Aerosp. Eng.

J. Aerosp. Eng., 2019, 32(6): 04019089


Viscous perfect gas flow (Rb−1.95 mm) High entropy layer edge
Inviscid perfect gas flow (R b−1.95 mm)

Boundary layer profile


0.015 0.015

M∞ 0.01 0.01 0.01

Y (m)
0.005 0.005 0.005

0 0 0
−0.5 0 0.5 1 −0.5 0 0.5 1 −0.5 0 0.5 1
U/U∞ U/U∞ U/U Ramp
Blunt leading edge ∞
x=0.012 m x=0.022 m x=0.042 m
Downloaded from ascelibrary.org by Nottingham Trent University on 08/09/19. Copyright ASCE. For personal use only; all rights reserved.

Separation bubble

Fig. 17. Variation of boundary-layer and entropy-layer thickness at three different locations for perfect gas flow with equivalent radius (upstream
influence location 0.022 m).

Viscous real gas flow (R b−1.85 mm) High entropy layer edge
Inviscid real gas flow (R b−1.85 mm)
Boundary layer profile
0.015 0.015

M 0.01 0.01
∞ 0.01
Y (m)

0.005 0.005 0.005

0 0 0
−0.5 0 0.5 1 −0.5 0 0.5 1 −0.5 0 0.5 1
U/U U/U U/U Ramp
Blunt leading edge ∞ ∞ ∞
x=0.012 m x=0.022 m x=0. 042 m
Separation bubble

Fig. 18. Variation of boundary-layer and entropy-layer thickness at three different locations for real gas flow with equivalent radius (upstream
influence location 0.022 m).

Blast wave theory (R −0.84 mm) Blast wave theory (R −0.781 mm)
b b
0.35 Blast wave theory (R −1.7 mm)
b 0.35 Blast wave theory (R −1.6 mm)
b
Blast wave theory (Rb−1.85 mm) Blast wave theory (Rb −1.7 mm)
Pressure coefficient (Cp )

0.3 0.3
Pressure coefficient (Cp )

Blast wave theory (R −1.89 mm) Blast wave theory (R −1.85 mm)
b b
0.25 Blast wave theory (R −1.95 mm) 0.25 Real gas flow (R −0.0 mm)
b b
Perfect gas flow (R −0.0 mm)
0.2 b 0.2

0.15 0.15
Upstream influence location Upstream influence location
0.1 0.1

0.05 0.05

0 0

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0 0.005 0.01 0.015 0.02 0.025 0.03 0.035
(a) x (m) (b) x (m)

Fig. 19. Comparison of surface pressure distribution from blast wave theory with (a) perfect gas flow; and (b) real gas flow.

than the local boundary layer for perfect and real gas flow. An- region lead to a need for a lower value of critical radii of the blunt-
other valuable remark concers the equivalent radius; it is worth ness of real gas flow conditions. The present investigations also
noting that a streamwise favorable pressure gradient in conjunc- revealed that the special gas dynamic features at the critical radii
tion with a thick HEL is highly desirable to reduce the separation are an independent type of gas flow. Further, the critical radii
length absolutely for separation control. It was noted that the real predicted for perfect gas flow would have the same functionality
gas effects reduce the extent of separation owing to SWBLI in in a real gas flow situation as well. Thus, only perfect gas sim-
comparison with perfect gas flow. Thus, the reduced requirements ulations are sufficient to predict the critical radii to ensure sepa-
of a high entropy-layer thickness and upstream overpressure ration control.

© ASCE 04019089-10 J. Aerosp. Eng.

J. Aerosp. Eng., 2019, 32(6): 04019089


Notation Rep. No. 1356. Moffett Field, CA: National Advisory Committee
for Aeronautics, Ames Aeronautical Laboratory.
The following symbols are used in this paper: Deb, K., S. Agrawal, A. Pratap, and T. Meyarivan. 2000. “A fast elitist
Cd = drag coefficient; non-dominated sorting genetic algorithm for multi-objective optimiza-
tion: NSGA-I.” In Parallel problem solving from nature PPSN VI,
Ci = mass concentration of species i (kg=m3 );
edited by M. Schoenauer, E. Lutton, K. Deb, G. Rudolph, X. Yao,
Cp = pressure coefficient; J. J. Merelo, and H.-P. Schwefel, 849–858. Berlin: Springer.
E = total energy (J=kg); Deepak, N. R., S. L. Gai, and A. J. Neely. 2010. Aerothermodynamics of
Ei = activation energy; hypersonic shock wave boundary layer interactions. In Proc., 17th
EI = convective flux vector in x direction; Australian Fluid Mechanics Conf. Auckland, New Zealand: Univ. of
Ev = viscous flux vector in x direction; Auckland.
Desai, S., V. Kulkarni, and H. Gadgil. 2016. “Delusive influence of non-
FI = convective flux vector in y direction; dimensional numbers in canonical hypersonic non-equilibrium flows.”
Fv = viscous flux vector in y direction; J. Aerosp. Eng. 29 (5): 1–10. https://doi.org/10.1061/(ASCE)AS.1943
Downloaded from ascelibrary.org by Nottingham Trent University on 08/09/19. Copyright ASCE. For personal use only; all rights reserved.

hi = specific enthalpy of species i (J=ðkgÞ; -5525.0000620.


k = optimization cycle counter; Desai, S., V. Kulkarni, H. Gadgil, and B. John. 2017. “Aerothermodynamic
kbi = backward rate constants; considerations for energy deposition based drag reduction technique.”
Appl. Ther. Eng. 122 (Jul): 451–460. https://doi.org/10.1016/j
kfi = forward rate constants;
.applthermaleng.2017.04.114.
LRb = separation length (m); Dunn, M. G., and S. W. Kang. 1973. Theoretical and experimental studies
M ∞ = freestream Mach number; of reentry plasmas. NASA CR, No. 2232. Washington, DC: National
N = number of species; Aeronautics and Space Administration.
nj = molar concentration for species j; Ess, P., and R. C. B. Allen. 2005. “Parallel computation of two-dimensional
p = pressure (N=m2 ); laminar inert and chemically reactive multi-species gas flows.” Int. J.
Numer. Methods Heat Fluid Flow 15 (3): 228–256. https://doi.org/10
qx = heat flux in x direction (W=m2 ); .1108/09615530510583865.
qy = heat flux in y direction (W=m2 ); Eusuff, M., and K. Lansey. 2007. “Shuffled frog-leaping algorithm: A
Rb = leading-edge radius (mm); memetic meta-heuristic for discrete optimization.” Eng. Optim.
S = source term vector; 38 (2): 129–154. https://doi.org/10.1080/03052150500384759.
Si = species production rate (kg=m3 s); Glushneva, A., V. A. S. Saveliev, E. E. Son, and D. V. Tereshonok. 2015.
“Investigation of shock wave-boundary layer instability on the heated
Sj = rate of change of molar concentration of species j;
ramp surface.” In Vol. 653 of Proc., Journal of Physics: Conf. Series,
T ∞ = freestream temperature (K); 012069. Bristol, UK: IOP Publishing.
t = time (s); Gordon, S., and B. J. McBride. 1994. Computer program for calculation of
U = conserved variable vector; complex chemical equilibrium composition and applications. NASA,
u = velocity in x direction (m=s); No 1331. Cleveland: National Aeronautics and Space Administration.
ūi = diffusion velocity in x direction (m=s); Grasso, F., M. Marini, G. Ranuzzi, S. Cuttica, and B. Chanetz. 2001.
“Shock-wave/turbulent boundary layer interactions in non-equilibrium
v = velocity in y direction (m=s); flows.” AIAA J. 39 (11): 2131–2140. https://doi.org/10.2514/2.1209.
v̄i = diffusion velocity in y direction (m=s); Hataue, I. 1989. “Computational study of the shock-wave/boundary layer
ν ij0 = stoichiometric coefficients; interaction in a duct.” Fluid Dyn. Res. 5 (3): 217–234. https://doi.org/10
ν ij0 0 = stoichiometric coefficients; .1016/0169-5983(89)90023-3.
Y i = mass fraction of species i; Holden, M. S. 1971. “Boundary-layer displacement and leading-edge
bluntness effects on attached and separated laminar boundary layers
Δymin = first cell thickness (m);
in a compression corner. Part II: Experimental study.” AIAA J. 9 (1):
λk = adaptive learning rate; 84–93. https://doi.org/10.2514/3.6127.
ρ = density (kg=m3 ); and Inger, G., and R. S. Zee. 1978. “Transonic shock wave/turbulent-boundary-
τ xx , τ xy , τ yy = shear stress components (N=m2 ). layer interaction with suction or blowing.” J. Aircr. 15 (11): 750–754.
https://doi.org/10.2514/3.58442.
John, B., and V. Kulkarni. 2013. “Investigation of energy deposition
References technique for drag reduction at hypersonic speeds.” Appl. Mech. Mater.
90: 222–227. https://doi.org/10.4028/www.scientific.net/AMM.367
Agarwal, S., N. Sahoo, K. J. Irimpan, V. Menezes, and S. Desai. 2017. .222.
“Comparative performance assessments of surface junction probes John, B., and V. Kulkarni. 2014a. “Effect of leading edge bluntness on the
for stagnation heat flux estimation in a hypersonic shock tunnel.” interaction of ramp induced shock wave with laminar boundary layer at
Int. J. Heat Mass Transfer. 114 (Nov): 748–757. https://doi.org/10 hypersonic speed.” Comput. Fluids 96 (Jun): 177–190. https://doi.org
.1016/j.ijheatmasstransfer.2017.06.109. /10.1016/j.compfluid.2014.03.004.
Blazek, J. 2001. Computational fluid dynamics: Principles and applica- John, B., and V. Kulkarni. 2014b. “Numerical assessment of correlations
tions. Kidlington, UK: Elsevier Science. for shock wave boundary layer interaction.” Comput. Fluids 90 (Feb):
Brahmachary, S., G. Natarajan, V. Kulkarni, N. Sahoo, and S. R. Nanda. 42–50. https://doi.org/10.1016/j.compfluid.2013.11.011.
2017. Application of greedy and heuristic algorithm based optimisation John, B., and V. Kulkarni. 2017. “Alterations in critical radii of bluntness of
methods towards aerodynamic shape optimization. In Vol. 1 of Proc., shock wave boundary layer interaction.” J. Aerosp. Eng. 30 (5): 1–8.
7th Int. Conf. on Soft Computing for Problem Solving, 937–948. https://doi.org/10.1061/(ASCE)AS.1943-5525.0000728.
Bhubaneswar, India: Indian Institute of Technology Bhubaneswar. John, B., V. Kulkarni, and G. Natarajan. 2014. “Shock wave boundary layer
Bussing, T., and R. A. E. M. Murman. 1988. “Finite volume method for the interactions in hypersonic flows.” Int. J. Heat Mass Transfer 70 (Mar):
calculation of compressible chemically reacting flows.” AIAA J. 26 (9): 81–90. https://doi.org/10.1016/j.ijheatmasstransfer.2013.10.072.
1070–1078. https://doi.org/10.2514/3.10013. John, B., S. Surendranath, G. Natarajan, and V. Kulkarni. 2016. “Analysis
Chapman, D. R., D. M. Kuehn, and H. K. Larson. 1957. Investigation of of dimensionality effect on shock wave boundary layer interaction in
separated flows in supersonic and subsonic streams with emphasis laminar hypersonic flows.” Int. J. Heat Fluid Flow 62 (Dec):
on the effect of transition. NACA Technical Notes 3869. NACA 375–385. https://doi.org/10.1016/j.ijheatfluidflow.2016.09.009.

© ASCE 04019089-11 J. Aerosp. Eng.

J. Aerosp. Eng., 2019, 32(6): 04019089


Kaczyński, P., and R. Szwaba. 2016. “Influence of cooling intensity on wave/boundary layer interactions.” In Vol. 1 of Proc., Shock Waves:
shock wave boundary layer interaction region in turbine cascade.” In 26th Int. Symp. on Shock Waves, 683–688. Berlin: Springer.
Vol. 760 of Proc., Journal of Physics: Conference Series, 012007. Pasquariello, V., M. Grilli, S. Hickel, and N. A. Adams. 2014. “Large-eddy
Bristol, UK: IOP Publishing. simulation of passive shock-wave/boundary-layer interaction control.”
Katzer, E. 1989. “On the length scales of laminar shock/boundary-layer Int. J. Heat Fluid Flow 49 (Oct): 116–127. https://doi.org/10.1016/j
interaction.” J. Fluid Mech. 206: 477–496. https://doi.org/10.1017 .ijheatfluidflow.2014.04.005.
/S0022112089002375. Piotrowicz, M., and P. Flaszynski. 2016. “Numerical investigations of
Kirkpatrick, S., C. D. Gelatt, and M. P. Vecchi. 1983. “Optimization by shock wave interaction with laminar boundary layer on compressor pro-
simulated annealing.” Science 220 (4598): 671–680. https://doi.org/10 file.” J. Phys. Conf. Ser. 760 (1): 012023. https://doi.org/10.1088/1742
.1126/science.220.4598.671. -6596/760/1/012023.
Lee, M. G. 1999. “Numerical prediction of enhanced heat flux due to
Thévenin, D., and G. Janiga. 2008. Optimization and computational fluid
shock-on-shock interaction in hypersonic non-equilibrium flow.” Int.
dynamics. Berlin: Springer.
J. Numer. Methods Heat Fluid Flow 9 (2): 114–135. https://doi.org/10
Townsend, J. C. 1966. Effects of leading edge bluntness and ramp deflec-
.1108/09615539910255992.
Downloaded from ascelibrary.org by Nottingham Trent University on 08/09/19. Copyright ASCE. For personal use only; all rights reserved.

Liou, M., and S. C. J. Steffen. 1993. “A new flux splitting scheme.” J. Com- tion angle on laminar boundary layer separation in hypersonic flow.
put. Phys. 107 (1): 23–39. https://doi.org/10.1006/jcph.1993.1122. NASA TN D-3290. Springfield, VA: National Aeronautics And Space
Marini, M. 2001. “Analysis of hypersonic compression ramp laminar flows Administration.
under sharp leading edge conditions.” Aerosp. Sci. Technol. 5 (4): Venkatakrishnan, V. 1993. “On the accuracy of limiters and convergence to
257–271. https://doi.org/10.1016/S1270-9638(01)01109-9. steady state solutions.” In Proc., 31st Aerospace Sciences Meeting,
Needham, D. A., and J. L. Stollery. 1966. “Boundary-layer separation in Aerospace Sciences Meetings. Reno, NV: American Institute of
hypersonic flow.” In Proc., 3rd and 4th AIAA Aerospace Sciences Meeting, Aeronautics and Astronautics.
455. Reston, VA: American Institute of Aeronautics and Astronautics. Wieting, A. R. 1987. Experimental study of shock wave interference
Neuenhahn, T., and H. Olivier. 2009. “Numerical study of wall tempera- heating on a cylindrical leading edge. NASA TM A00484. Honolulu:
ture and entropy layer effects on transitional double wedge shock American Institute of Aeronautics and Astronautics.

© ASCE 04019089-12 J. Aerosp. Eng.

J. Aerosp. Eng., 2019, 32(6): 04019089

Das könnte Ihnen auch gefallen