Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/348891897

Reynolds-Averaged Numerical Simulations of Conical Shock-


Wave/Boundary-Layer Interactions

Article  in  AIAA Journal · January 2021


DOI: 10.2514/1.J059582

CITATIONS READS

0 147

3 authors:

Feng-Yuan Zuo Antonio Memmolo


Xi'an Jiaotong University Sapienza University of Rome
15 PUBLICATIONS   73 CITATIONS    6 PUBLICATIONS   37 CITATIONS   

SEE PROFILE SEE PROFILE

Sergio Pirozzoli
Sapienza University of Rome
165 PUBLICATIONS   3,965 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

multiphase flows View project

wall turbulence View project

All content following this page was uploaded by Feng-Yuan Zuo on 30 January 2021.

The user has requested enhancement of the downloaded file.


AIAA JOURNAL

Reynolds-Averaged Numerical Simulations of Conical


Shock-Wave/Boundary-Layer Interactions

Feng-Yuan Zuo∗
State Key Laboratory for Strength and Vibration of Mechanical Structures,
Xi’an Jiaotong University, 710049 Xi’an, People’s Republic of China
Antonio Memmolo†
CINECA-Interuniversity Consortium, 40033 Bologna, Italy
and
Sergio Pirozzoli‡
University of Rome “La Sapienza”, 00184 Rome, Italy
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

https://doi.org/10.2514/1.J059582
We carry out a parametric study of conical shock-wave/turbulent boundary-layer interactions by means of
numerical simulation of the Reynolds-averaged Navier–Stokes (RANS) equations, with the eventual goal of
establishing the predictive capabilities of standard turbulence models. Preliminary assessment of several linear-
eddy-viscosity models for the case of planar interactions shows that the k − ε model and its variants as well as the
Spalart–Allmaras model yield accurate prediction of the typical interaction length scale. Numerical simulations of
conical interactions at high Reynolds numbers under attached and separated flow conditions generally support good
capability of RANS to predict the gross flow features, including conditions of incipiently and fully separated flow.
Comparison with direct numerical simulation data at low Reynolds numbers suggests that caution should be made
because certain turbulence models may yield unrealistic incoming velocity profiles. With this caveat, we again find
that RANS models yield satisfactory prediction of the three-dimensional flow organization and associated length
scales. The N-wave mean wall pressure signature is quantitatively predicted, and accurate information about the
fluctuating pressure variance can be obtained from RANS data in a postprocessing stage by extension of correlations
developed for low-speed separation bubbles.

Nomenclature ω = rate of dissipation per unit turbulence kinetic energy


Cf = wall shear-stress coefficient
E = total energy per unit mass Subscripts
H = total enthalpy per unit mass in = inflow parameters
h = height or enthalpy per unit mass r = reference station
k = turbulence kinetic energy rms = root mean square
L = interaction length or cone longitudinal length rs = position of the reflected-shock foot
M = Mach number t = turbulent quantity
N = grid points ∞ = freestream parameters
p = pressure
Re = Reynolds number
Superscripts
t = time
ui = velocity vector 00 = fluctuation from Favre-averaged mean value
uvd = the Van Driest transformed mean streamwise velocity  = nondimensionalization by the inner scales of boundary
uτ = friction velocity layer
x, y, z = Cartesian coordinates
δ0 = boundary-layer thickness at the reference location
ε = turbulence dissipation I. Introduction
θ = momentum boundary-layer thickness at the reference

μ =
location or semiangle of wedge and cone
viscosity
S HOCK-WAVE/BOUNDARY-LAYER interactions (SBLIs) have
been the subject of extensive research over the past decades
[1]. SBLIs occur in a variety of flows of practical interest, such as
ρ = density supersonic intakes [2,3], overexpanded nozzles [4,5], and fluidic
σ ij = viscous stress tensor control methods [6,7]. Typically, these interactions negatively impact
τij = Reynolds stress tensor the aerodynamic performance, yielding thicker boundary layers
τw = wall shear stress and flow separation, resulting in large total pressure loss and flow
distortion. Furthermore, large pressure fluctuations and thermal loads
Received 19 March 2020; revision received 15 October 2020; accepted for are found in the interaction region, which may threaten the structural
publication 21 October 2020; published online 29 January 2021. Copyright integrity of aircraft. A thorough understanding of such complex
© 2021 by the American Institute of Aeronautics and Astronautics, Inc. All flow physics remains a necessary prerequisite for accurate flowfield
rights reserved. All requests for copying and permission to reprint should be prediction, efficient design of high-speed aerospace vehicles, as
submitted to CCC at www.copyright.com; employ the eISSN 1533-385X to well as advances in combustion processes to improve the design of
initiate your request. See also AIAA Rights and Permissions www.aiaa.org/ engineering systems [1].
randp.
*Assistant Professor, School of Aerospace Engineering, 28 Xianning West SBLIs are observed whenever a shock wave sweeps across the
Road; fengyuan@xjtu.edu.cn (Corresponding Author). boundary layer developing on a wall surface. Planar SBLIs are

High Performance Computing Department, via Magnanelli 6/3. classified as “compression corner” interactions or “impinging shock”
‡ interactions. As Pirozzoli et al. [8] summarized, the overall features
Professor, Dipartimento di Ingegneria Meccanica e Aerospaziale, Via
Eudossiana 18. for the two types of interactions are similar:
Article in Advance / 1
2 Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI

1) The turbulent boundary layer develops under an adverse pres- in RANS turbulence closures based on the Boussinesq hypothesis. In
sure gradient (APG). particular, it was found that the assumed proportionality between
2) The wall pressure rise occurs upstream of the nominal inter- turbulent shear stress and the related mean velocity gradient applies
action point due to an upstream influence mechanism. also when three-dimensional effects are important, and the primary
3) For strong interactions, the flow separates and pressure has a shear stress is predicted with reasonable accuracy.
plateau in the separation zone. In the present paper, we report results of RANS turbulence models
4) The shock system and the separation bubble experience a three- for two CSBLI configurations, whereby a conical shock wave is
dimensional motion characterized by low-frequency unsteadiness. made to interact with a flat-plate boundary layer. The general flow
5) Turbulence and mixing are enhanced across the interacting topology and the three-dimensional separation bubbles shape are
shock, and they undergo a relaxation process downstream of the presented, including a parametric study of the effect of the shock
interaction. strength. The eventual goal of the study is to establish the reliability of
SBLIs have received a lot of attention in the past since they occur simple linear-eddy-viscosity RANS models for the prediction of
almost at any time in a supersonic stream, and important progress has complex shock/boundary-layer interaction cases. The paper is organ-
been achieved both on the side of experimental investigations [9–12] ized as follows: the turbulence models adopted in this paper are
and high-fidelity numerical simulations [13–17]. Morgan et al. [18] described in Sec. II, and their performance is preliminarily assessed
have developed a large-eddy simulation (LES) database covering a in Sec. III for the case of two-dimensional SBLIs at high Reynolds
range of shock strengths and Reynolds numbers, and they compared numbers. A parametric study of CSBLIs along with comparison with
it with Reynolds-averaged Navier–Stokes (RANS) results. Based on experimental and DNS data, at high and low Reynolds numbers,
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

modeling errors associated with turbulence closures using eddy- respectively, is presented in Sec. IV. Concluding remarks are pro-
viscosity k − ω and stress transport, they concluded that although vided in Sec. V.
stress-transport models offer improved predictions, inadequacies in
modeling the turbulence transport terms and isotropic treatment of
dissipation limit their predictive accuracy. Despite the large numbers
of measurements and numerical simulations of SBLIs that have II. Physical Model and Numerical Method
been conducted, there is still much about the physical dynamics of A. RANS Equations
SBLIs remaining unclear, especially concerning the low-frequency In the Reynolds-averaged Navier–Stokes equations, the instanta-
unsteadiness associated with the separation shock, whose frequen- neous variables are decomposed into their mean and fluctuating
cies are about two orders of magnitude larger than those in the components. In Cartesian coordinates, the equations are cast in
upstream boundary layer and whose underlying cause is still the conservation form as follows:
subject of debate [19]. Conical shock-wave/boundary-layer inter-
actions (CSBLIs) have received comparatively less attention with ∂ρ ∂ρu~ j 
respect to the planar case, although axisymmetric surfaces and coni-   0;
∂t ∂xj
cal shock waves are far more common in practice.
In recent hypersonic vehicles, in order to maximize aerodynamic ∂ρ u~i  ∂ρu~ i u~ j  ∂p ∂
 −  σ~ − τij ; i  1; 2; 3;
performance, both the external flow on the fuselage (such as the SR- ∂t ∂xj ∂xi ∂xj ij
72 hypersonic vehicle and the HTV, Hypersonic Technology Vehicle)
and the internal flow in the high-speed intake [20,21] are nearly ∂ρ E
^ ∂ρ H^ u~ j  ∂ h i ∂
  σ~ ij − τij u~ i − q~  Qj  (1)
axisymmetric, and CSBLIs occur [22]. Historically, only limited ∂t ∂xj ∂xj ∂xj j
CSBLI research efforts have been made. Kussoy et al. [23] conducted
an experiment with a conical shock impinging on an axisymmetric where ρ, ui , p, E, H, σ ij , and qi are the density, velocity vector,
wall surface at a freestream Mach number of 2.2. Mean flow quan- pressure, total energy per unit mass, total enthalpy per unit mass,
tities were obtained across the interaction region; but, due to the viscous stress tensor, and heat flux vector. The overbar and the tilde
limited experimental techniques available at that time, the separation denote, respectively, the Reynolds ensemble-averaging operator and
flow topology of this CSBLI problem is not clear. Gai and Teh [24]
investigated the interaction of the shock wave produced by a conical the density-weighted (Favre) averages, f~  ρf∕ρ; fluctuations with
shock with a planar turbulent boundary layer at freestream Mach respect to Reynolds and Favre averages are denoted with a single
number of two. Varying the cone opening angles from 14 to 30 deg, or double prime, respectively. Additional definitions are given as
both attached and separated conditions were realized. The results follows:
showed that the incident shock induces a significant pressure gradient  
1 ∂u~ k
in the spanwise direction, which leads to strong crossflow with a σ~ ij ≃ 2μ Sij −
~ δ ; τij  ρ ug00 00
i uj ;
horseshoe-shaped signature identified through the surface oil-flow 3 ∂xk ij
technique, whose size and strength gradually reduce away from the ∂T~
symmetry plane. Notably, the pressure rise for incipient separation q~ i ≃ −κ ; Qi  ρ hg 0 0u 0 0;
∂xi i
was found to be less than for the two-dimensional case. Hale [25]
considered the impingement of a conical shock wave on a plane 1
E~  E^  k; E^  e~  u~ k u~ k ;
turbulent boundary layer at Mach number 2.05 and gathered infor- 2
mation by means of surface oil flow, pressure-sensitive paint, and 1
particle image velocimetry techniques. The data suggested that the H~  H^  k; H^  h~  u~ k u~ k ;
2
interaction causes locally two-dimensional separation near the 1 g 0
k  uk uk ;0 0 0 h  cp T (2)
centerline and three-dimensional separation away from this region,
with fluid propagating away from the centerline. Significant span- 2
wise and streamwise expansions were observed right downstream of
where S~ij is the strain-rate tensor, τij is the Reynolds stress tensor,
the interaction leading edge, unlike in equivalent two-dimensional
wedge-generated SBLIs. The results further gave hints for instanta- k  τkk ∕2ρ is the turbulence kinetic energy, and Qi the turbulent heat
neous boundary-layer separation, and they showed that the interac- flux vector.
tion tends to suppress large-scale vortical structures in the incoming With these choice, the same relations that hold for the instanta-
boundary layer. neous variables also hold for the averaged ones; hence, p  ρ RT~ and
Zuo et al. [26] carried out a high-fidelity direct numerical simu- e~  cv T.
~ The molecular viscosity μ is determined from Sutherland’s
lation (DNS) of CSBLIs at M∞  2.05 and Reθ ≈ 630 based on the law as a function of T,~ and the thermal conductivity κ is related to μ
upstream boundary-layer momentum thickness. The DNS database through κ  cp μ∕Pr (Pr  0.72, constant). A simple eddy-viscos-
was exploited to verify the validity of fundamental assumptions made ity assumption is here considered to model τij and Qj :
Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI 3

    2 3 2 3 2 3
1 1 ∂u~ i ∂u~ j 1 ∂u~ k ρ ρ u~ 0
τij − δij τkk  −2ρνt  − δij ; 6 7 6 7 6 7
3 2 ∂xj ∂xi 3 ∂xk 6 ρ u~ 7 6 ρ u~ u~ p i^ 7 6 τtot 7
6 7 6 7 6 xi 7
∂T~ 6 7 6 7 6 τyi 7
Qi  −κ t W  6 ρ v~ 7; F  6 ρ u~ v~ p j^ 7; G6 tot
7
∂xi
(3) 6 7 6 7 6 7
6 ρ w~ 7 6 ρ u~ w~ p k^ 7 6 τtot 7
4 5 4 5 4 zi 5
ρ E~ ρ u~ E~ p u~ τij u~ j − q
tot tot

where νt denotes the eddy viscosity, and the turbulent thermal


conductivity κ t is determined from κ t  cp ρ νt ∕Prt (Prt  0.85). (8)
Several linear-eddy-viscosity models have been tested in this study,
Here, W is the vector of the conserved quantities, F is the vector of the
which are briefly described in the Appendix for completeness,
ij  σ
inviscid fluxes, G is the vector of the viscous fluxes, τtot ~ ij − τij ,
namely, the Spalart–Allmaras (S-A) model, the k − ε model (also
in its realizable variant as Rk − ε), the k − ω model, and the shear- and qtot  q~  Q. Vectors are reported in bold characters, and V l is
stress-transport (SST) model. the lth control volume. At low Mach numbers, Eq. (7) becomes
numerically very stiff. To alleviate numerical stiffness, precondition-
ing is used to modify the system of equations by switching from the
1. Near-Wall Treatment
conserved quantities W to the primitive quantities V  p;  u; ~T
~ T
In the S-A model, the wall is accommodated by setting the eddy through a modified Jacobian [29]. The preconditioning is, however,
viscosity to zero at the wall. All the other models rely on a zero wall only active in a narrow near-wall region. The inviscid flux Ff ⋅ Af is
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

boundary condition for k. In the k − ω models, ω in the first off-wall


evaluated using Roe’s flux-difference splitting adapted to the pre-
cell is prescribed according to [27] conditioned system [30]

6μ 1 1
Ff ⋅ Af ≃ F^ f  FR  FL  ⋅ Af − ΓjAG jV R − V L 
ωw  10 (4) (9)
0.075ρΔd2 2 2

where subscript f denotes a generic face surrounding the control


where Δd is the cell center distance from the wall. In the standard and volume V l , and the fluxes FR  FV R  and FL  FV L  are evalu-
realizable k − ε models, two layers are considered: 1) a fully turbu- ated using the reconstructed primitive variables vectors V R and V L on
lent region far away from the wall, which obeys the baseline k − ε the “right” and “left” sides of the face. For the non-preconditioned
model equations; and 2) a viscous region, in which the one-equation system, Eq. (9) reduces to Roe’s flux-difference splitting [31] when
model of Wolfshtein [28] is used. The two regions are discriminated Roe-averaged values are inserted in ΓAΓ ; in the present formulation,
based on a turbulent Reynolds number based on the distance from the arithmetic averaging of states V R and V L is used. Such a method may
nearest wall d thus be regarded as a second-order central difference plus an added
artificial dissipation. The required left and right states are evaluated by
p interpolating the cell center values with a second-order upwind scheme
ρ d k using a Taylor series expansion [32], which requires the knowledge of
Red  (5)
μ the gradient in each cell, computed with a least-squares cell-based
approach. In the reconstruction phase, a minmod (minimum modulus)
slope limiter is used to prevent spurious oscillations that would occur
with the boundary being set at Red  200. Additional details are across shocks, discontinuities, or near-rapid changes in the flowfield.
provided in the Appendix. Gradients appearing in the viscous fluxes are instead evaluated with the
divergence theorem, and then they are corrected to contain all imme-
2. Freestream Conditions diate neighbors cells [33] for enhanced stability.
For all cases, we assume a freestream eddy viscosity of μt∞ ≈ The spatially discretized equations are advanced in time by means
10μ∞ . Previous tests have shown the inlet value of the freestream of a backward Euler implicit scheme combined with a Newton-type
eddy viscosity to be rather unimportant; hence, we decided to use the linearization of the fluxes to produce the linearized system in delta
standard value used in general-purpose commercial software. For the form [33]. Overall, second-order accuracy is achieved. It should
S-A model, this amounts to ν~ ∞  12ν∞ . For all other models, we be remarked that the aforementioned formulation is valid (and
further assume moderate freestream turbulence intensity: conservative) for steady-state solutions.

r C. Numerical Boundary Conditions


2
I k ∕u  5% (6) The computational domain used for the numerical simulations
3 ∞ ∞ of CSBLIs is sketched in Fig. 1. At the inflow boundary, all the
flow variables are set to their freestream value, whereas they are
from which the freestream value of k is inferred. Freestream values of
ε∞ and ω∞ are then determined from Eq. (18) or Eq. (19) with
α  1, respectively.

B. Numerical Discretization
A finite volume solver is used to determine approximate solutions
of Eqs. (1), where the method of lines is used so that spatial and
temporal terms are discretized independently. Spatial discretization is
carried out by integrating the equations about each control volume,
with the resulting equations taking the form

Z I

W dV l  F − G ⋅ dA  0 (7)
∂t Vl ∂V l

Fig. 1 Sketch of boundary conditions for numerical simulation of


where CSBLI.
4 Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI

extrapolated at the outflow boundary. The same inflow conditions are wedge axis and the bottom surface is set to h  10 mm to avoid
used for all turbulence models, with freestream parameters selected interference of the trailing-edge expansion with the interaction zone.
such that the boundary-layer state at a reference station upstream of The global flow parameters of the simulation reproduce the exper-
the interaction zone matches reference experimental or DNS data. All imental ones, with a freestream Mach number of M∞  2.3 and a
solid walls are assumed to be adiabatic and no slip. Although the Reynolds number of Reδ0 ≈ 55;000, based on the boundary-layer
assumption of adiabatic wall behavior is necessarily an approxima- thickness at the upstream reference position (xr  29 mm). A range
tion, it is likely to be appropriate in the case of reference experiments of wedge angles is considered from θ  6 to 12 deg. The computa-
in which the residence time is long. At the upper and spanwise tional mesh includes 750 × 250 cells in the streamwise and wall-
boundaries, a symmetry condition is exploited, mimicking the case normal directions, respectively. Grid points are refined in the
of a free-slip wall. This condition causes reflection of waves back into wall-normal direction according to a hyperbolic algorithm to ensure
the domain, and care is taken to ensure they do not contaminate the near-wall Δy ≤ 1.0. Grid points are also clustered in the interaction
interaction region for any of the flow cases hereafter reported. zones to accurately resolve the separation bubble. The same mesh is
used for all turbulence models.
The number of grid points has been selected after a grid refinement
III. Two-Dimensional Wedge-Induced SBLI study, for the wedge-angle θ  6 deg case, run with the Rk − ε
The various turbulence models have been preliminarily applied to model. Figure 3a shows the friction coefficient for three different
study the case of planar SBLIs at high Reynolds numbers using the meshes. The baseline mesh has a number of points of N x × N y 
same computational mesh. In all cases, the distance of the first off-wall 750 × 250, shown as a black dashed line. The number of points is
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

cell center is Δy ≤ 1. The case of a planar oblique shock impinging then doubled in the streamwise direction (1500 × 250; red solid line)
on a flat plate (OSBLI: oblique shock wave/boundary layer interac- and in the wall-normal direction (750 × 500; blue dotted–dashed
tion), sketched in Fig. 2, has been extensively investigated through line), respectively. As shown by the figure, convergence is already
experiments [9,10] and numerical simulations [8,14,34]. If the pressure achieved on the baseline mesh (750 × 250), which is then adopted for
jump across the impinging planar shock is sufficiently large, the the following study. The effect of the distance of the first cell center
associated adverse pressure gradient yields separation of the boundary from the wall has also been checked; and the results are shown in
layer, resulting in a turbulent separation bubble. The deflection of the Fig. 3b, again in terms of skin friction. Three values of Δy have been
flow away from the wall generates compression waves, which merge to tested, namely, Δy ≃ 0.94 (red solid line), Δy ≃ 0.47 (blue
form the separation shock in the outer flow. Outside the bubble, an dashed–dotted–dotted line), and Δy ≃ 0.093 (orange dashed line).
expansion fan adjusts the flow direction, followed by compression Again, the three distributions are nearly identical; hence, the largest
waves near reattachment. Downstream of the interaction zone, the Δy has then been retained for the following study.
boundary layer undergoes relaxation to an equilibrium state. To estab- The computed boundary-layer properties at the reference station
lish the suitability of the various turbulence models for the description are given in Table 1: again recalling that the computational domain
of the interaction, we first compare the simulation results with a well- and the inflow conditions are the same for all turbulence models.
documented experimental test case [9]. Hence, because of different streamwise evolution of the computed
The selected computational domain has a streamwise length of boundary layer, the flow properties at the reference station have some
Lx  150 mm and a height of Ly  20 mm. The leading edge of the scatter with respect to the target experiment, and the Reynolds
shock generator is fixed at x  30 mm, and the distance between the number is especially low for the k − ω and SST models. To verify

Fig. 2 Sketch of oblique planar shock/turbulent boundary-layer interaction (OSBLI) setup.

a) b)
Fig. 3 Mesh refinement study for OSBLI: friction coefficient with different mesh resolution with wedge angle of θ  6 deg using the Rk − ε model:
a) different number of grid points, and b) different near-wall resolutions with Nx × Ny  750 × 250.
Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI 5

Table 1 OSBLI: boundary-layer properties at the reference station


(xr  29 mm)

Model M∞ Cf Reδr  u∞ δ0 ∕ν∞ Reθr  u∞ θr ∕ν∞


S-A 2.3 2.32 × 10−3 55,000 6,900
k-ε 2.3 2.28 × 10−3 58,000 7,250
k-ω 2.3 2.17 × 10−3 49,000 6,500
SST 2.3 2.27 × 10−3 49,000 6,550
SST-HRa 2.3 2.23 × 10−3 53,000 6,990
Rk-ε 2.3 2.30 × 10−3 55,000 7,150
Experiment
2.3 2.10 × 10−3 55,000 6,900
[35]

a
SST-HR refers to tests carried out with the SST model by increasing the inlet Reynolds
number.

whether this difference has any impact on the results shown hereafter,
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

an additional set of RANS has been carried out with the SST model by Fig. 5 OSBLI: variation of the interaction length scale with the imposed
increasing the inflow unit Reynolds number, which is referred to as shock strength compared with experimental data [9,36].
SST-HR. The van Driest transformed mean streamwise velocity
Z u~  1∕2
ρ experiments quite well throughout the range of flow parameters.
uvd  du~ (10) On the other hand, the SST and k − ω models largely overpredict
0 ρ w
the separation length scale: also in the case of weak interactions.
This behavior does not appear to be connected with mismatch in
is shown in Fig. 4 in a semilogarithmic plot [35]. In the figure, we also
the incoming boundary-layer conditions because increasing the
show the universal logarithmic law of the wall:
Reynolds number in the SST-HR case has minor effects. Similar
shortcomings of certain turbulence models were reported by Morgan
1
u  C  log y (11) et al. [18], who traced the origin to near-wall effects. Similarly good
κ predictions are offered by the S-A, k − ε, and Rk − ε in the case of
weak interactions (say, Δp∕2τwr ≤ 40); whereas differences emerge
with C ≈ 5.0 and κ ≈ 0.41. In agreement with many previous RANS for stronger interactions, with the general trend for S-A to predict
studies (e.g., Pirozzoli et al. [8]), we find that all RANS models yield larger separation bubbles, and for k − ε to predict smaller bubbles,
a reasonable representation of the mean velocity profile under equi- with the Rk − ε model in between. In general, given the experimental
librium conditions, with a linear range for y ≤ 5 and a narrow region data scatter and uncertainty due to three-dimensional effects, no
(30 ≤ y ≤ 200) with nearly logarithmic variation. model can be judged to be clearly superior to other ones, although
Figure 5 shows the variation of the computed interaction length some slight preference can be here attributed to the Rk − ε model.
scale, namely, the distance between the apparent origin of the sepa- In two dimensions, the separation region is defined by the exist-
ration shock and the nominal impingement point, with the shock ence of a bubble containing closed streamlines circling around a
intensity, quantified in terms of the ratio Δp∕2τwr , where Δp is the common point and bounded by a separation streamline [38], starting
pressure rise across the incident shock, and τwr is the wall shear stress at the separation point and ending at the reattachment point. Figure 6
at the upstream reference station [36]. As a note of caution, we point shows the computed streamlines in the interaction zone for a wedge
out that in the case of strong interactions, three-dimensional effects angle of θ  8 deg. For all turbulence models, the recirculation
are likely to arise in experiments, which are not accounted for in zone is quite shallow, as expected for turbulent separation bubbles.
numerical simulations [37]. Nevertheless, some models (including However, the size and shape are quite different for different turbu-
S-A, k − ε, and Rk − ε) seem to capture the trends measured in lence models. As previously noted, the streamwise extent is correctly
predicted by the Rk − ε model, although the thickness appears to be
underestimated.

IV. RANS of CSBLI


Having explored the predictive capabilities of RANS in the case of
oblique shock interactions, we now challenge it for the analysis of
three-dimensional CSBLIs. In the forthcoming sections, we sepa-
rately compare the performance of RANS with available experiments
at high Reynolds numbers, and with DNS data at (necessarily) much
lower Reynolds numbers.

A. High-Reynolds-Number CSBLI: Comparison with Experiments


Based on the outcome of the previous section, the Rk − ε and the
S-A turbulence models are here considered for parametric investiga-
tion of CSBLIs at high Reynolds numbers. We consider as a reference
the experimental study of Gai and Teh [24], in which the freestream
Mach number is 2.0 and the Reynolds number of the incoming
boundary layer is Reδ0  55;500. The computational domain is
Fig. 4 OSBLI: van Driest transformed mean velocity profiles at the sketched in Fig. 7. The domain is chosen with a streamwise length
upstream reference station. Dashed line denotes a compound of u  of Lx  300 mm, a height of Ly  60 mm, and a width of Lz 
y and u  5.0  1∕0.41 log y . Circles denote experimental (Exp) 120 mm. The leading edge of the cone is set at xc  80 mm, and the
data [35]. distance between the cone axis and the flat surface is set as
6 Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI

a) b)
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

c) d)

e)
Fig. 6 OSBLI: computed streamlines for θ  8 deg: a) SST model, b) k − ω model, c) S-A model, d) k − ε model, and e) Rk − ε model. Red dashed line
marks u  0 isoline, and blue dashed–dotted line denotes experimental results [11]. Interaction length L is selected as relevant longitudinal length scale.
Dimensionless longitudinal coordinate is defined as x  x − xrs ∕L, where xrs is mean position of reflected-shock foot.

h  30 mm. The opening semicone angle is changed with different The computational domain is discretized with a block-structured
values to control the conical shock intensity, namely, 14, 16, 18, 20, mesh as shown in Fig. 8, which includes approximately eight million
25, and 30 deg. In all cases, the cone has a base diameter of 20 mm, grid points. Quadrilateral mesh blocks are first generated near the
and it is mounted at 0 deg incidence. cone, (see Fig. 8a), which are then extruded outward into an hexa-
hedral mesh. The grid points are clustered toward the bottom wall in
such a way that the first point has a distance of Δy ≈ 1.37 for all
cases. A uniform distribution of grid points is used in the spanwise
direction, with Δz ranging from 50 to 200, depending on the
Reynolds number of the considered test case. Boundary conditions
are the same as described in Sec. II.C. Convergence of the solution
was checked by monitoring the residuals of the mass, momentum,
and energy equations; and simulations were stopped once the varia-
tion over 1000 iterations was less than 0.02%.
The Mach number contours in the symmetry plane (Fig. 9) show
the main features of the interaction. As expected, the shock generated
by the cone is found to be nearly straight and strengthens as θ is
increased. As a result of its impingement, the underlying boundary
layer is seen to thicken significantly; and a zone of reduced momen-
Fig. 7 Sketch of computational domain for CSBLI analysis, where xc is tum (sometimes separated) flow forms at the wall. Merging of the
the x coordinate of the cone leading edge. The green surface depicts the trailing-edge expansion fan is observed for this flow configuration
conical shock, whose wall trace is marked in red. in the rear part of the interaction zone. Additional features are clearly
Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI 7

Fig. 8 Sketch of mesh used for CSBLI: a) multiblock partitioning, b) mesh in x-y symmetry plane, and c) meshes near wall surface. Note that the large
part of the upstream domain has been cut off for clarity.
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

Fig. 9 High-Reynolds-number CSBLI: Mach number contours in symmetry plane for various cone angles, obtained with Rk − ε turbulence model:
a) θ  14 deg, b) θ  16 deg, c) θ  18 deg, d) θ  20 deg, e) θ  25 deg, and f) θ  30 deg. Contours shown for 0.2 ≤ M ≤ 2.4, from blue to red.
Blank zone represents shock generator.

visible in this representation, including the reflected shock and the underestimation of the upstream influence. The friction coefficient
recompression shock in the wake of the cone. Additional numerical (Figs. 11d–11f) shows some noticeable difference, in that the S-A
shock reflections are found at the upper boundary, whose effect at the model predicts slightly higher friction upstream of the interaction, as
wall is only felt past x ≈ 180 mm, which as a consequence should be well as a larger separation bubble.
discarded when comparing with experimental data. Measurements of the separation bubble length (L, defined as the
The wall pressure contours shown in Fig. 10 yield a complemen- distance between separation and reattachment points in the symmetry
tary view of CSBLIs. The interacting conical shock has a distinctive plane) were made by Gai and Teh [24], which we are comparing to
hyperbolic trace at the wall; past which, large pressure amplification RANS in Fig. 12. Agreement between the two sets of data is quite
is observed. Owing to flow three-dimensionality, the pressure remarkable. In particular, both RANS and experiments suggest that
increase is larger at the symmetry plane, and it tends to vanish away incipient separation should occur at θ ≈ 17 deg. The Rk − ε model
from it. Shock reflections are observed at the outer spanwise boun- yields the best prediction, with some discrepancy only for cases with
daries, which are a result of the numerical symmetry conditions. A the strongest shocks. The S-A model also yields good results, but it
strong expansion is observed past impingement of the trailing-edge shows consistent overprediction of experimental data by about 10%
expansion fan, which results in lower-pressure values than in the for θc ≳ 20 deg.
freestream. Additional compressions and expansions in similar Additional insight into the topology of the flow-reversal zones is
alternating fashion are found further downstream as a result of gained from the analysis of the limit stream traces in the wall plane,
impingement of the wake recompression shock and the additional shown in Fig. 13. No vortex or separation bubble is observed in the
waves shown in the symmetry plane visualizations. Overall, the symmetry plane for θ < 18 deg. At θ  18 deg, a tiny near-wall
flow organization is very similar to that found in experimental recirculation arises. In the separated cases, the bigger the semicone
visualizations [24,25]. angle, the stronger the conical shock strength, and the larger the
A quantitative comparison is shown in Figs. 11a–11c, where we separation bubble becomes. It should be noted that the region of
compare wall pressure in the symmetry plane with the experimental reversed flow, which may be defined as the region between the wall
data of Gai and Teh [24]. The focus here is on the primary interaction and the stream trace entering the saddle point, is quite shallow. In
zone, which in all cases features an upward trend, followed by an separated cases, the wall-surface limiting streamlines exhibit a pat-
expansion. Notably, RANS well predicts both the amplitude and tern similar to that resulting from separation induced by an obstacle
length scale of the bump, and it captures the two-stage pressure rise [39] or vertical cylinder [40], and the flow features a saddle point
that is observed in the weak shock case (Fig. 11a). The agreement is associated with three-dimensional flow separation. This suggests that
also remarkable for stronger interactions, however with possible near the surface the conical SBLI behaves as if there is an obstacle in
8 Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

Fig. 10 High-Reynolds-number CSBLI: wall pressure contours for various cone angles, obtained with Rk − ε turbulence model. a) θ  14 deg,
b) θ  16 deg, c) θ  18 deg, d) θ  20 deg, e) θ  25 deg, and f) θ  30 deg. Contours shown for 0.5 ≤ p∕p∞ ≤ 2.0, from blue to red.

a) b) c)

d) e) f)
Fig. 11 High-Reynolds-number CSBLI: centerplane distributions of a–c) wall pressure (blue symbols denoting experimental data of Gai and Teh [24])
and d–f) friction coefficient.
Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI 9

the previous section. Although RANS turbulence models are not


specifically targeted for marginally turbulent flow, we believe that
comparing with DNS can provide additional insight into their pre-
dictive capabilities by also allowing us to access flow properties that
are not easily measurable in experiments. The DNS dataset of Zuo
et al. is considered for the comparison, which mainly differs from the
present RANS setup in terms of the far-field boundary conditions,
which were of nonreflecting type in that study [26]; whereas here we
assume free-slip walls. This translates into numerical wave reflec-
tions that are not present in the DNS. However, the focus of com-
parison here is the shock impingement region, which is not affected
by differences in the boundary conditions. The incoming Mach
number is M∞  2.05, the Reynolds number upstream of the inter-
action zone is Reδ0 ≈ 7000, and the cone angle is θ  25 deg.
Lengths are here made nondimensional with respect to the boundary-
layer thickness at the inflow δin. The total number of grid points is about
Fig. 12 High-Reynolds-number CSBLI: bubble length (normalized by 8.5 million. The velocity profile upstream of the interaction (see
upstream boundary-layer thickness) vs cone angle. The blue circles Fig. 14) brings to light serious difficulties of the Rk − ε model in
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

denote experimental data by Gai and Teh [24].


predicting a realistic distribution, presumably owing to the modest
Reynolds number under examination. Hence, for the purposes of
comparing with DNS, here we only consider the S-A model, which
the flow. Passing through the shock, there is a region of separated
is the only one among those tested to yield a reasonable incoming
flow that follows the shape of the conical shock, which is delimited by velocity distribution, but which suffers from other issues, as shown in
the coalescence (separation) and divergence (reattachment) lines. A the previous sections.
horseshoe pattern of separation whose strength and size are reduced The main features of the flow are compared in Fig. 15. Similar to
away from the symmetry plane is identified. Downstream of the the high-Reynolds number case, the main interacting shock is gen-
divergence line, the reattached flow continues to diverge away from erated as straight at the cone surface, and it subsequently bends
the centerline. backward upon merging with the expansion fan that arises at the
cone trailing edge. The density field predicted by the S-A model
B. Low-Reynolds-Number CSBLI: Comparison with DNS generally matches the DNS results, with the main exception of severe
RANS is here compared with DNS data, which are carried out at underestimation of the separation bubble length past the cone that,
much lower Reynolds number Re than the experiments reported in however, is outside our primary interest.

Fig. 13 High-Reynolds-number CSBLI: limiting streamlines at wall under different cone angles in interaction zone, obtained with Rk − ε turbulence
model: a) θ  14 deg, b) θ  16 deg, c) θ  18 deg, d) θ  20 deg, e) θ  25 deg, and f) θ  30 deg.
10 Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI

DNS, and the S-A model; whereas the Rk − ε model shows shorter
upstream influence and a higher peak as a result of the incorrect
incoming boundary-layer structure.
In response to the imposed pressure gradient, the friction coeffi-
cient (see Fig. 17) exhibits strongly nonmonotonic variation.
Upstream of the interaction zone, the friction coefficient is well
predicted by the S-A model; whereas again, the Rk − ε model shows
major differences owing to incorrect boundary-layer development.
We should mention that, owing to the recycling/rescaling procedure
[26], the boundary layer of the DNS dataset becomes properly
developed only for x∕δin ≳ 15. The S-A model also works quite well
in predicting the typical length scale of the separated flow zone,
whereas the friction overshoot resulting from subsequent flow expan-
sion is a bit overpredicted. The complex behavior observed down-
stream and differences among models are difficult to judge because of
differences in the numerical setup between RANS and DNS.
The structural modifications of boundary-layer turbulence
upon interaction with the conical shock are compared in Fig. 18.
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

Fig. 14 Low-Reynolds-number CSBLI: van Driest transformed incom- For that purpose, contours of the density-scaled velocity correlation
ing velocity distributions obtained with Rk − ε model (dashed–dotted ρ ∕ρw ug0 0 v 0 0 in the symmetry plane are reported upon normalization
line) and S-A model (solid line) compared with DNS data [26] (symbols). by the friction velocity at the upstream reference station uτref . The
Dashed line denotes a compound of u  y and u  5.0  1∕
0.41 log y .
figure supports general amplification of all turbulence intensities
across the APG zones. The density-scaled velocity correlation of
the S-A turbulence model shows a good agreement with DNS results.
The mean wall pressure p w across the interaction zone is compared An interesting question is whether and to what extent RANS can
with DNS in Fig. 16. This figure shows overall favorable comparison also be used to extract information about certain unsteady flow
with correct alternation of compression and expansion zones. A more features such as the wall pressure fluctuations, which are especially
quantitative comparison (see Fig. 16b) shows good agreement in the important in applications. To extract useful information regarding the
prediction of the primary pressure peak between the experiment, unsteady pressure field from RANS simulations, we exploit the

Fig. 15 Low-Reynolds-number CSBLI: mean density in symmetry plane as obtained from DNS (left) and from S-A model (right). Contour levels shown
within range of 0.2 < ρ∕ρ∞ < 2.0. Isoline u  0 is marked in red.

a) b)
Fig. 16 Low-Reynolds-number CSBLI: a) mean pressure contours in wall plane, and b) profile along symmetry axis. Contour levels in Fig. 16a shown
within the range of 0.6 < p∕p∞ < 1.8. Diamond symbols in Fig. 16b denote experimental data [25].
Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI 11

of the peak pressure position, whose amplitude is overestimated by


about 1 dB, which is acceptable for most practical purposes.

V. Conclusions
Two-dimensional and three-dimensional cone-induced shock/
boundary-layer interactions at both high and low Reynolds numbers
have been numerically studied by RANS using various linear-eddy-
viscosity turbulence models. Extensive tests carried out for the case
of two-dimensional interactions show that RANS is generally
capable of qualitatively predicting the flow organization in the inter-
action zone. However, quantitative comparison with experimental
data [9] shows that only certain models correctly predict the typical
length scale of the interaction and its growth with the incident shock
strength. Among the various models under scrutiny, the two-equation
Rk − ε turbulence model showed the most favorable behavior.
Numerical experiments of three-dimensional interactions have
been carried out to match existing experimental data at high Reynolds
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

Fig. 17 Low-Reynolds-number CSBLI: wall friction coefficient along


the symmetry line. The dashed horizontal line indicates the zero value. numbers [24], as well as DNS data at lower Reynolds numbers [26].
In the former case, the Rk − ε model has been tested in a range of cone
angles corresponding to various shocks strengths. Quantitative com-
parison with experiments is extremely good, showing that the wall
analogy between turbulent shear stress and the rms pressure fluctua- pressure footprint is well characterized, and incipient separation
tions, which was first proposed by Simpson et al. [41] in the context occurs for a cone opening angle of θ  17 deg at a freestream Mach
of attached and separated incompressible turbulent boundary layers, number of 2.0. Some difficulties for conventional turbulence models
and it was extended to SBLIs by Pirozzoli et al. [8]. Those authors were noticed when comparing with DNS at lower Reynolds numbers
found a good collapse of the wall pressure distribution when normal- Re, at which it is difficult to achieve a fully developed turbulent
ized by the corresponding values of the maximum turbulent shear boundary-layer state. This difficulty is particularly evident with the
stress in the wall-normal direction. In particular, they showed that Rk − ε model, whereas the S-A model was found to be less sensitive.
Accounting for these shortcomings, RANS is still found to yield
prms ≈ α ⋅ maxτxy (12) accurate prediction of the mean wall pressure signature and of the
y
typical interaction length scale. Based on a priori checks made with
DNS data, this is due to the good accuracy of the Boussinesq closure,
with α ≈ 2.5. This analogy makes it possible to exploit the distribution even in the case of non-uniaxial flow. Limited (but useful) informa-
of turbulent shear stress obtained from a RANS simulation to infer the tion about flow unsteadiness can be also gained from steady RANS
rms pressure level at the wall. Comparison between DNS data and by suitable additional modeling closures. For instance, information
RANS prediction is shown in Fig. 19, where the rms pressure is on the wall pressure variance can be obtained using an existing
reported on a decibel scale. In fact, the predicted fluctuating pressure correlation with the maximum turbulent shear stress along the
maps well to the DNS data, with stronger fluctuations in the region of wall-normal direction. As previously shown for the case of two-
flow retardation and relief in the reacceleration zone. Quantitative dimensional SBLIs, it is found that for three-dimensional inter-
agreement is also quite good, with special reference to the prediction actions, the correlation yields the prediction of peak and local values

a)

10
b)
Fig. 18 Low-Reynolds-number CSBLI: Maps of density-scaled velocity correlation (−ρ∕ρw u 0g
0 00
v ) in the symmetry plane, scaled by the friction velocity
at the upstream reference station. Figure 18a shows DNS, and Fig. 18b shows S-A model.
12 Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

a)

b)
Fig. 19 Low-Reynolds-number CSBLI: a) rms pressure fluctuations in wall plane [left, DNS; right, S-A model + Eq. (12) with α  2.5]; and b) profiles
along symmetry axis [solid line, DNS; dashed line, theoretical equation (12) with α  2.5 applied to RANS data; and dashed–dotted line, theoretical
equation (12) with α  2.5 applied to DNS data].

of the fluctuating pressure that are also accurate to within 1 dB in where d is the distance from the nearest wall, and Ω
~ ij is the mean rate
the presence of separated flow, which is quite remarkable. Whereas of rotation tensor:
high-fidelity numerical methods such as LES or DNS are needed to
 
probe additional unsteady flow features, such as the long-debated 1 ∂ui ∂uj
low-frequency breathing motion of the separation bubble [18], it is Ωij  −
2 ∂xj ∂xi
found that RANS is still a good candidate to predict the salient
features of complex shock/boundary-layer interactions with reason-
able accuracy. The damping function in the destruction term is given by
 
1  c6w3 1∕6 ν~
fw  g 6 ; g  r  cw2 r6 − r; r
Appendix A: Turbulence Models g  c6w3 kSkκ
~ 2v d2
(A3)
A.1. Spalart–Allmaras Model
In the Spalart–Allmaras one-equation model [42], an additional The eddy viscosity is evaluated as νt  ν~ fν1 , with fν1 
transport equation for the pseudoeddy viscosity ν~ is solved: χ 3 ∕χ 3  c3ν1 . The model constants are here set to cb1  0.1355,
cb2  0.622, σ  2∕3, cν1  7.1, cw1  3.2059, cw2  0.3,
   
∂ρ ν~ ∂ρ ν~ u~ j cw3  2, and κ v  0.4187.
 ~ ρ ν~  1 ∂ μ  ρ ν~  ∂~ν  cb2 ρ ∂~ν ∂~ν
 cb1 kSk
∂t ∂xj σ ∂xj ∂xj ∂xj ∂xj
 2
ν~ A.2. k − ε Model
− cw1 fw ρ (A1)
d The two-equation k − ε turbulence model is based on model trans-
port equations for the turbulence kinetic energy k and its dissipation
where the first and last terms at the right-hand side represent, respec- rate ε. The transport equations are
tively, the production and destruction terms. The term kSk ~ appearing
  
in the production term is defined as ∂ ∂ ∂u~ ∂ μ ∂k
ρk  ρu~ j k  −τij i  μ  t
q ∂t ∂xj ∂xj ∂xj σ k ∂xj
kSk
~  2Ω ~ ij Ω~ ij  ν~ fv2 ; fv2  1 − χ
; χ
ρ ν~
(A2)
κ 2v d2 1  χfv1 μ − ρ ε1  2M2t  (A4)
Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI 13

  
∂ ∂ ε ∂u~ ∂ μt ∂ε with the freestream independence of the k − ε model. The eddy
ρε  ρu~ j ε  −c1ε τij i  μ  viscosity is determined from
∂t ∂xj k ∂xj ∂xj σ ε ∂xj
ε2 ρ k 1
− c2ε ρ (A5) μt  (A16)
k ω max1∕α ; ΩF2 ∕a1 ω
In these equations, the first and last terms at the right-hand side and the transport equations are similar to the parent models:
account
p  for production and dissipation, respectively; and Mt 
2k∕a~ is the turbulent Mach number, with a~ as the sound speed.   
∂ ∂ ∂u~ ∂ μ ∂k
The eddy viscosity is evaluated as ρk  ρu~ j k  −τij i  μ  t − ρ β kω
∂t ∂xj ∂xj ∂xj σ k ∂xj
k2 (A17)
μt  ρ cμ (A6)
ε   
∂ ∂ ρ ∂u~ ∂ μ ∂ω
The model constants are here set to c1ε  1.44, c2ε  1.92, ρω  ρu~ j ω  −α τij i  μ  t
∂t ∂xj μt ∂xj ∂xj σ ω ∂xj
cμ  0.09, σ k  1.0, and σ ε  1.3.
− ρ βω2  Dω (A18)

k − ω Model
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

A.3. where α is evaluated from Eq. (A10) by replacing the (constant) α∞


The standard k − ω model [43] is based on model transport equa- with
tions for k and ω  ε∕k, The eddy viscosity is obtained from
βi;1 κ 2v
k α∞  F1 α∞;1  1 − F1 α∞;2 ; α∞;1   − p ;
μt  α ρ 
(A7) β∞ σ ω;1 β∞
ω
βi;2 κ 2v
α∞;2  − p (A19)
α
where is the low-Reynolds-number damping term. Also, k and ω β∞ σ ω;2 β∞
are obtained by solving
   and
∂ ∂ ∂u~ ∂ μ ∂k
ρk  ρu~ j k  −τij i  μ  t
∂t ∂xj ∂xj ∂xj σ k ∂xj βi  F1 βi;1  1 − F1 βi;2 (A20)

− ρ β fβ kω (A8)
The cross-diffusion term Dω is defined as
  
∂ ∂ ω ∂u~ ∂ μt ∂ω 1 ∂k ∂ω
ρω  ρu~ j ω  −α τij i  μ  Dω  21 − F1 ρσ ω;2 (A21)
∂t ∂xj k ∂xj ∂xj σ ω ∂xj ω ∂xj ∂xj
− ρ βfβ ω2 (A9)
The blending functions F1 and F2 are given by
with
F1  tanhΦ41 ; F2  tanhΦ22  (A22)
   
α0  Ret ∕Rk α α0  Ret ∕Rω
α  α∞ ; α ∞ (A10)   p  
1  Ret ∕Rk α 1  Ret ∕Rω k 500μ 4ρk
Φ1  min max ; 2 ; ;
8 0.09ωd ρ d ω σ ω;2 D
ωd
2

>
< 1  680χ k
2  p 
if χ k > 0 1  70χ ω k 500μ
f β  1  400χ 2k ; fβ  (A11) Φ2  max 2 ; 2 (A23)
>
:1 1  80χ ω 0.09ωd ρ d ω
if χ k ≤ 0
  q
 ~ ~ ~  1 1 ∂k ∂ω
1 ∂k ∂ω  Ωij Ωjk Ski  D
ω  max ρ
2 ; 10−20 ; Ω 2Ω ~ ij Ω~ ij (A24)
χk  3 ; χ ω   (A12) σ ω;2 ω ∂xj ∂xj
ω ∂xj ∂xj β∞ ω3 
1 1
σk  ; σω 
β  βi 1  ζ  FMt ; β  βi − βi ζ  FMt  (A13) F1 ∕σ k;1  1 − F1 ∕σ k;2 F1 ∕σ ω;1  1 − F1 ∕σ ω;2
(A25)
 
4∕15  Ret ∕Rβ 4 ρ k
βi  β∞ ; Ret  (A14) where Ωij is the mean rotation rate, d is the distance from the nearest
1  Ret ∕Rβ 4 μ ω
wall, and Dω is the positive portion of the cross-diffusion term. The
model constants are here set to σ k;1  1.176, σ ω;1  2, σ k;2  1,
FMt   M2t − M2t0 HMt − Mt0 ; M2t 
2k
(A15) σ ω;2  1.168, a1  0.31, βi;1  0.075, and βi;2  0.0828. The addi-
a~ 2 tional model constants have the same values as in the standard
k − ω model.
Here, Hx is the Heaviside step function; and the model constants
are set to α∞  1, α∞  0.52, α0  βi ∕3, α0  1∕9, β∞  0.09,
βi  0.072, Rβ  8, Rk  6, Rω  2.95, ζ  1.5, Mt0  0.25,
σ k  2, and σ ω  2. A.5. Realizable k − ε Model
The realizable k − ε model was introduced by Shih et al. [44] to
guarantee positivity of the normal Reynolds stresses and Schwarz
A.4. SST Model inequality for the turbulent shear stress [44]. The modeled transport
The shear-stress-transport model of Menter [27] effectively blends equation for k is the same as for the standard model [Eq. (A4)], except
the near-wall robust and accurate formulation of the k − ω model for the model constants, whereas the equation for ε is
14 Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI

  
∂ ∂ ∂ μ ∂ε Hence, in the fully turbulent layer, where λε ≃ 1, Eq. (A5) or
ρε  ρu~ j ε  ρ c1 Sε  μ  t Eq. (A26) is recovered; whereas in the viscous layer, where λε ≃ 0,
∂t ∂xj ∂xj σ ε ∂xj
we have
ε2
− ρ c2 p (A26)
k  ν ε dε k3∕2
 −ε (A38)
dt lε
where
  q which reduces to Eq. (A31) under steady conditions.
η k The constants are here set to [47]
c1  max 0.43; ; ηS ; S 2S~ij S~ij (A27)
η5 ε
cl  κc−3∕4
μ ; Aμ  70; Aε  2cl (A39)
and a variable cμ is for the evaluation of μt [Eq. (A6)]:
q Acknowledgments
1
cμ  ; U  S~ij S~ij  Ω ~ ij Ω ~ ij (A28)
A0  As kU ∕ε F-Y. Zuo would like to acknowledge the support of National Natural
Science Foundation of China (no. 12002261), the National Postdoc-
p toral Program for Innovative Talents, China (no. BX20200267), the
A0  4.04; As  6 cos ϕ (A29)
Young Talent fund of University Association for Science and Tech-
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

nology in Shaanxi, China (no. 20200501), the Fundamental Research


1 p S~ij S~jk S~ki Funds for the Central Universities of China (no. xzy012020096), and
ϕ  cos−1  6W; W  q (A30)
3 the China Postdoctoral Science Foundation (no. 2020M673411). We
S~ij S~ij
acknowledge that the results reported in this paper have been achieved
using the PRACE Research Infrastructure resource MARCONI based
The model constants are here set to c2  1.92, σ k  1.0, and at CINECA in Casalecchio di Reno in Italy.
σ ε  1.2.
References
[1] Dolling, D. S., “Fifty Years of Shock-Wave/Boundary-Layer Interaction
A.6. Near-Wall Treatment of k − ε Models Research: What Next?” AIAA Journal, Vol. 39, No. 8, 2001, pp. 1517–
For the k − ε models, the flow is divided into two layers: 1) a fully 1531.
turbulent region far away from the wall, which obeys the k − ε model https://doi.org/10.2514/2.1476
equations; and 2) a viscous region, where ε and μt are both specified [2] Krishnan, L., Sandham, N. D., and Steelant, J., “Shock-Wave/
algebraically [28]: Boundary-Layer Interactions in a Model Scramjet Intake,” AIAA Jour-
nal, Vol. 47, No. 7, 2009, pp. 1680–1691.
https://doi.org/10.2514/1.41107
k3∕2 [3] Loth, E., Titchener, N., Babinsky, H., and Povinelli, L., “Canonical
ε (A31)
lε Normal Shock Wave/Boundary-Layer Interaction Flows Relevant to
External Compression Inlets,” AIAA Journal, Vol. 51, No. 9, 2013,
p pp. 2208–2217.
μt;2l  ρ cμ klμ (A32) https://doi.org/10.2514/1.J052175
[4] Keanini, R. G., and Brown, A. M., “Scale Analysis and Experimental
where the damping effects are contained in Observations of Shock-Induced Turbulent Boundary Layer Separation
in Nozzles,” European Journal of Mechanics—B/Fluids, Vol. 26, No. 4,
h i 2007, pp. 494–510.
lε  dcl 1 − exp−Red ∕Aε  (A33) https://doi.org/10.1016/j.euromechflu.2006.10.002
[5] Verma, S., Chidambaranathan, M., and Hadjadj, A., “Analysis of Shock
s Unsteadiness in a Supersonic Over-Expanded Planar Nozzle,” Euro-
h i ρ pean Journal of Mechanics—B/Fluids, Vol. 68, March 2018, pp. 55–65.
w
lμ  dcl 1 − exp−Red ∕Aμ  (A34)
ρ https://doi.org/10.1016/j.euromechflu.2017.11.005
[6] Gefroh, D., Loth, E., Dutton, C., and McIlwain, S., “Control of an
Oblique Shock/Boundary-Layer Interaction with Aeroelastic Meso-
Equation (A34) accounts for the effect of compressibility through the flaps,” AIAA Journal, Vol. 40, No. 12, 2002, pp. 2456–2466.
density ratio, where ρ w is the wall density [45]. https://doi.org/10.2514/2.1589
The blending between the two layers is smoothly and implicitly [7] Verma, S. B., and Manisankar, C., “Shockwave/Boundary-Layer Inter-
achieved for steady-state solutions as proposed in Ref. [46], intro- action Control on a Compression Ramp Using Steady Microjets,” AIAA
ducing the blending function λε Red : Journal, Vol. 50, No. 12, 2012, pp. 2753–2764.
https://doi.org/10.2514/1.J051577
   [8] Pirozzoli, S., Beer, A., Bernardini, M., and Grasso, F., “Computational
1 Red − 200
λε  1  tanh (A35) Analysis of Impinging Shock-Wave Boundary Layer Interaction Under
2 20∕atanh0.98 Conditions of Incipient Separation,” Shock Waves, Vol. 19, No. 6, 2009,
pp. 487–497.
which is close to zero for Red < 200 (i.e., in the near-wall region) and https://doi.org/10.1007/s00193-009-0215-9
one otherwise. The blended eddy viscosity is computed simply as the [9] Dupont, P., Haddad, C., and Debieve, J. F., “Space and Time Organiza-
convex combination of the two: tion in a Shock-Induced Separated Boundary Layer,” Journal of Fluid
Mechanics, Vol. 559, July 2006, pp. 255–277.
https://doi.org/10.1017/S0022112006000267
μt;sth  λε μt  1 − λε μt;2l (A36)
[10] Humble, R. A., Elsinga, G. E., Scarano, F., and van Oudheusden, B. W.,
“Three-Dimensional Instantaneous Structure of a Shock Wave/
where μt is the one of the fully turbulent layer. The blended dissipa- Turbulent Boundary Layer Interaction,” Journal of Fluid Mechanics,
tion is computed in a similar fashion but acting on the differential Vol. 622, March 2009, pp. 33–62.
equations [Eq. (A5) or Eq. (A26)] by multiplying all the terms but the https://doi.org/10.1017/S0022112008005090
temporal derivative by λε, as well as by adding the term [11] Souverein, L. J., Dupont, P., Debieve, J. F., Dussauge, J. P.,
van Oudheusden, B. W., and Scarano, F., “Effect of Interaction Strength
 3∕2  on Unsteadiness in Turbulent Shock-Wave-Induced Separations,” AIAA
k
1 − λε  −ε (A37) Journal, Vol. 48, No. 7, 2010, pp. 1480–1493.
lε https://doi.org/10.2514/1.J050093
Article in Advance / ZUO, MEMMOLO, AND PIROZZOLI 15

[12] Eagle, W. E., and Driscoll, J. F., “Shock Wave-Boundary Layer Inter- pp. 2050–2057.
actions in Rectangular Inlets: Three-Dimensional Separation Topology https://doi.org/10.2514/3.12946
and Critical Points,” Journal of Fluid Mechanics, Vol. 756, Oct. 2014, [30] van Leer, B., Lee, W.-T, and Roe, P. L., “Characteristic Time-Stepping or
pp. 328–353. Local Preconditioning of the Euler Equations,” AIAA Paper 1991-1552,
https://doi.org/10.1017/jfm.2014.382 1991.
[13] Pirozzoli, S., Bernardini, M., and Grasso, F., “Direct numerical Simu- [31] Roe, P. L., “Characteristic-Based Schemes for the Euler Equations,”
lation of Transonic Shock/Boundary Layer Interaction Under Condi- Annual Review of Fluid Mechanics, Vol. 18, No. 1, 1986, pp. 337–
tions of Incipient Separation,” Journal of Fluid Mechanics, Vol. 657, 365.
Aug. 2010, pp. 361–393. https://doi.org/10.1146/annurev.fl.18.010186.002005
https://doi.org/10.1017/S0022112010001710 [32] Barth, T., and Jespersen, D., “The Design and Application of Upwind
[14] Pirozzoli, S., and Bernardini, M., “Direct Numerical Simulation Data- Schemes on Unstructured Meshes,” 27th Aerospace Sciences Meeting,
base for Impinging Shock Wave/Turbulent Boundary-Layer Interac- AIAA Paper 1989-0366, 1989.
tion,” AIAA Journal, Vol. 49, No. 6, 2011, pp. 1307–1312. https://doi.org/10.2514/6.1989-366
https://doi.org/10.2514/1.J050901 [33] Weiss, J., Maruszewski, J., Smith, W., Weiss, J., and Maruszewski, J.,
[15] Wang, B., Sandham, N. D., Hu, Z. W., and Liu, W. D., “Numerical Study “Implicit Solution of the Navier–Stokes Equations on Unstructured
of Oblique Shock-Wave/Boundary-Layer Interaction Considering Meshes,” 13th Computational Fluid Dynamics Conference, AIAA
Sidewall Effects,” Journal of Fluid Mechanics, Vol. 767, March 2015, Paper 1989-2103, 1997.
pp. 526–561. https://doi.org/10.2514/6.1997-2103
https://doi.org/10.1017/jfm.2015.58 [34] Aubard, G., Gloerfelt, X., and Robinet, J. C., “Large-Eddy Simulation of
[16] Hadjadj, A., “Large-Eddy Simulation of Shock/Boundary-Layer Inter- Broadband Unsteadiness in a Shock/Boundary-Layer Interaction,”
Downloaded by XI'AN JIAOTONG UNIVERSITY on January 29, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J059582

action,” AIAA Journal, Vol. 50, No. 12, 2012, pp. 2919–2927. AIAA Journal, Vol. 51, No. 10, 2013, pp. 2395–2409.
https://doi.org/10.2514/1.J051786 https://doi.org/10.2514/1.J052249
[17] Adams, N. A., “Direct Simulation of the Turbulent Boundary Layer [35] Dupont, P., Piponniau, S., Sidorenko, A., and Debieve, J. F., “Inves-
Along a Compression Ramp at M  3 and Reθ  1685,” Journal of tigation by Particle Image Velocimetry Measurements of oblique Shock
Fluid Mechanics, Vol. 420, Oct. 2000, pp. 47–83. Reflection with Separation,” AIAA Journal, Vol. 46, No. 6, 2008,
https://doi.org/10.1017/S0022112000001257 pp. 1365–1370.
[18] Morgan, B., Duraisamy, K., Nguyen, N., Kawai, S., and Lele, S. K., https://doi.org/10.2514/1.30154
“Flow Physics and Rans Modelling of Oblique Shock/Turbulent Boun- [36] Laurent, H., “Turbulence d’une Interaction Onde de choc/Couche
dary Layer Interaction,” Journal of Fluid Mechanics, Vol. 729, Aug. Limite sur une Paroi Plane Adiabatique ou Chauffée,” Ph.D. Thesis,
2013, pp. 231–284. Aix-Marseille 2, 1996.
https://doi.org/10.1017/jfm.2013.301 [37] Dupont, P., Piponniau, S., and Dussauge, J. P., “Compressible Mixing
[19] Souverein, L. J., Bakker, P. G., and Dupont, P., “A Scaling Analysis for Layer in Shock-Induced Separation,” Journal of Fluid Mechanics,
Turbulent Shock-Wave/Boundary-Layer Interactions,” Journal of Fluid Vol. 863, March 2019, pp. 620–643.
Mechanics, Vol. 714, Jan. 2013, pp. 505–535. https://doi.org/10.1017/jfm.2018.987
https://doi.org/10.1017/jfm.2012.495 [38] Babinsky, H., and Harvey, J. K., Shock Wave-Boundary-Layer Inter-
[20] Zuo, F., Huang, G., and Xia, C., “Investigation of Internal-Waverider- actions, Vol. 32, Cambridge Univ. Press, New York, 2011.
Inlet Flow Pattern Integrated with Variable-Geometry for TBCC,” Aero- [39] Délery, J. M., “Robert Legendre and Henry Werlé: Toward the Eluci-
space Science and Technology, Vol. 59, Dec. 2016, pp. 69–77. dation of Three-Dimensional Separation,” Annual Review of Fluid
https://doi.org/10.1016/j.ast.2016.10.009 Mechanics, Vol. 33, No. 1, 2001, pp. 129–154.
[21] Huang, G., Zuo, F., and Qiao, W., “Design Method of Internal Waverider https://doi.org/10.1146/annurev.fluid.33.1.129
Inlet Under Non-Uniform Upstream for Inlet/Forebody Integration,” [40] Ozawa, H., and Laurence, S. J., “Experimental Investigation of the
Aerospace Science and Technology, Vol. 74, March 2018, pp. 160–172. Shock-Induced Flow over a Wall-Mounted Cylinder,” Journal of Fluid
https://doi.org/10.1016/j.ast.2018.01.012 Mechanics, Vol. 849, Aug. 2018, pp. 1009–1042.
[22] Zuo, F., and Huang, G., “Numerical Investigation of Bleeding Control https://doi.org/10.1017/jfm.2018.433
Method on Section-Controllable Wavecatcher Intakes,” Acta Astronau- [41] Simpson, R. L., Ghodbane, M., and McGrath, B. E., “Surface Pressure
tica, Vol. 151, Oct. 2018, pp. 572–584. Fluctuations in a Separating Turbulent Boundary Layer,” Journal of
https://doi.org/10.1016/j.actaastro.2018.06.059 Fluid Mechanics, Vol. 177, April 1987, pp. 167–186.
[23] Kussoy, M. I., Viegas, J. R., and Horstman, C. C., “Investigation of a https://doi.org/10.1017/S0022112087000909
Three-Dimensional Shock Wave Separated Turbulent Boundary Layer,” [42] Spalart, P. R., and Allmaras, S. R., “A One-Equation Turbulence Model
AIAA Journal, Vol. 18, No. 12, 1980, pp. 1477–1484. for Aerodynamic Flows,” 30th Aerospace Sciences Meeting and
https://doi.org/10.2514/3.50907 Exhibit, AIAA Paper 1992-0439, 1992.
[24] Gai, S. L., and Teh, S. L., “Interaction Between a Conical Shock Wave [43] Wilcox, D., Turbulence Modeling for CFD, Vol. 2, DCW Industries,
and a Plane Turbulent Boundary Layer,” AIAA Journal, Vol. 38, No. 5, La Cañada, CA, 1998.
2000, pp. 804–811. [44] Shih, T. H., Liou, W. W., Shabbir, A., Zhigang, Y., and Jiang, Z., “A
https://doi.org/10.2514/2.1060 New k-Epsilon Eddy Viscosity Model for High Reynolds Number
[25] Hale, J., “Interaction Between a Conical Shock Wave and a Plane Turbulent Flows,” Computers and Fluids, Vol. 24, No. 3, 1995,
Compressible Turbulent Boundary Layer at Mach 2.05,” M.S. Thesis, pp. 227–238.
Univ. of Illinois at Urbana-Champaign, Urbana, Illinois, 2015. https://doi.org/10.1016/0045-7930(94)00032-T
[26] Zuo, F.-Y., Memmolo, A., Huang, G.-P., and Pirozzoli, S., “Direct [45] Guezengar, D., Francescatto, J., Guillard, H., and Dussauge, J. P.,
Numerical Simulation of Conical Shock Wave-Turbulent Boundary “Variations on a k-ε Turbulence Model for Supersonic Boundary Layer
Layer Interaction,” Journal of Fluid Mechanics, Vol. 877, Oct. 2019, Computations,” European Journal of Mechanics—B/Fluids, Vol. 180,
pp. 167–195. No. 4, 1999, pp. 713–738.
https://doi.org/10.1017/jfm.2019.558 https://doi.org/10.1016/S0997-7546(99)00111-9
[27] Menter, F. R., “Two-Equation Eddy-Viscosity Turbulence Models for [46] Jongen, T., and Marx, Y. P., “Design of an Unconditionally Stable,
Engineering Applications,” AIAA Journal, Vol. 32, No. 8, 1994, Positive Scheme for the k-ϵ and Two-Layer Turbulence Models,” Com-
pp. 1598–1605. puters and Fluids, Vol. 26, No. 5, 1997, pp. 469–487.
https://doi.org/10.2514/3.12149 https://doi.org/10.1016/S0045-7930(97)00003-0
[28] Wolfshtein, M., “The Velocity and Temperature Distribution in One- [47] Chen, H., and Patel, V., “Near-Wall Turbulence Models for Complex
Dimensional Flow with Turbulence Augmentation and Pressure Gra- Flows Including Separation,” AIAA Journal, Vol. 26, No. 6, 1988,
dient,” International Journal of Heat and Mass Transfer, Vol. 12, No. 3, pp. 641–648.
1969, pp. 301–318. https://doi.org/10.2514/3.9948
https://doi.org/10.1016/0017-9310(69)90012-X
[29] Weiss, J. M., and Smith, W. A., “Preconditioning Applied to Variable P. G. Tucker
and Constant Density Flows,” AIAA Journal, Vol. 33, No. 11, 1995, Associate Editor

View publication stats

You might also like