Anda di halaman 1dari 11

Shock Waves (2009) 19:487–497

DOI 10.1007/s00193-009-0215-9

ORIGINAL ARTICLE

Computational analysis of impinging shock-wave


boundary layer interaction under conditions
of incipient separation
Sergio Pirozzoli · Alexandre Beer ·
Matteo Bernardini · Francesco Grasso

Received: 13 February 2009 / Revised: 27 May 2009 / Accepted: 15 June 2009 / Published online: 8 July 2009
© Springer-Verlag 2009

Abstract The interaction of an oblique shock wave with a 1 Introduction


turbulent boundary layer under conditions of incipient sepa-
ration is analyzed by means of large-eddy simulation (LES) The interaction of shock waves with turbulence has been the
and Reynolds-averaged Navier–Stokes (RANS) turbulence subject of extensive research over the last decades. Shock
models, with the objective to explore their predictive capabil- wave/turbulent boundary layer interactions (referred to as
ities, in particular with respect to the unsteady features of the SBLI in the following) occur in a variety of flows of prac-
interaction. Consistent with earlier direct numerical simula- tical interest, such as flow over control surfaces, supersonic
tions, we have found that the flow dynamics in the interaction inlets, and scramjet combustion chambers. For aircraft
zone is characterized by strong intermittency associated with flying in the transonic and the supersonic regime, the
the formation of scattered spots of flow reversal near the nom- presence of unsteady flow separation may cause severe
inal position of the reflected shock. Comparison with exper- vibrations due to the onset of large scale, low-frequency
imental results (at much larger Reynolds number) show that oscillations. Very large fluctuating pressure loads may occur
the qualitative features of the interaction are predicted rea- [1,3], with characteristic frequencies close to the resonant
sonably well by both LES and RANS models. RANS models frequencies of aircraft structural components. The localized
supplemented with a semi-empirical closure are also found to heating rates typical of shock/boundary layer interactions can
provide reasonable estimate of the fluctuating pressure loads also threaten the structural integrity of aircraft. The under-
at the wall. standing of such complex physical phenomena is extremely
important for the development of improved engineering
Keywords Shock waves · Boundary layers · models.
Large eddy simulation The shock wave interacting with a (turbulent) boundary
layer may be either produced at the surface where the bound-
PACS 47.27.em · 47.27.ep · 47.27.N- · 47.40.Nm ary layer is developing (as when a control surface is acti-
vated), or impinge on the boundary layer, being generated
Communicated by A. Hadjadj. from an external source. Accordingly, shock wave/turbu-
lent boundary layer interactions are classified [19] as ‘com-
S. Pirozzoli (B) · A. Beer · M. Bernardini · F. Grasso pression corner’ interactions or ‘incident shock’ interactions.
Dipartimento di Meccanica e Aeronautica,
The overall picture of the phenomenon for the two types of
Università di Roma “La Sapienza”,
Via Eudossiana 18, 00184 Rome, Italy interactions is similar: (i) the boundary layer develops under
e-mail: sergio.pirozzoli@uniroma1.it adverse pressure gradient; (ii) the wall pressure rise occurs
A. Beer upstream of the nominal interaction point due to an upstream
e-mail: alexandre.beer@uniroma1.it influence mechanism; (iii) for strong interactions the flow
M. Bernardini separates and pressure has a plateau in the separation zone;
e-mail: matteo.bernardini@uniroma1.it (iv) the shock system and the separation bubble experience
F. Grasso a three-dimensional motion characterized by low-frequency
e-mail: francesco.grasso@uniroma1.it unsteadiness; and (v) turbulence and mixing are enhanced

123
488 S. Pirozzoli et al.

across the interacting shock, and they undergo a relaxation carried out within the framework of the European sixth
process downstream of the interaction. framework program UFAST (Unsteady effects in shock-
The interaction of a shock wave impinging on a flat plate wave-induced separation), with the main objective to vali-
turbulent boundary layer in the presence of mean flow sepa- date the use of LES and RANS models for the prediction of
ration has been experimentally investigated by Dussauge and SBLI, in particular with respect to the unsteady features.
coworkers (see [4] and references therein). The experiments
indicate the dominance of low-frequencies in the interaction
zone, which are related to the formation of a mixing layer
2 Computational strategy
near the mean separation point. With regard to the turbulence
statistical properties, the experimental data show a different
Upon suitable simplifications, the filtered- and the RANS
amplification of the tangential and normal Reynolds stresses,
equations have the same form (although the meaning of the
whose maxima are found to occur away from the wall and
variables is different), which allows to use similar strate-
to have strong association with the coherent structures of the
gies for numerical discretization [8]. In Cartesian coordinates
mixing layer.
(x1 = x corresponding to the streamwise direction, x2 = y
Experiments of SBLI under conditions of incipient flow
to the wall-normal direction, and x3 = z to the spanwise
separation have recently been performed at the Technical
direction), the equations are cast in conservation form as
University of Delft (TUD) [20–22] by means of dual PIV.
follows
Those authors have shown that, under the selected flow con-
ditions, a large region of slow-moving motion forms near the ∂ρ ∂(ρ u j)
+ = 0,
wall, with the occurrence of regions of instantaneous flow ∂t ∂x j
reversal, associated with substantial increase of the velocity u i ) ∂(ρ 
∂(ρ  ui u j) ∂p
fluctuations and Reynolds stress. The TUD experiments have + =−
∂t ∂x j ∂ xi
also confirmed anisotropy between the streamwise and the
∂  
lateral fluctuation component (the peak value of the former + 
σi j −τi j , i = 1, 2, 3,
being a factor of four larger than the latter), and a much faster ∂x j
return to equilibrium of the streamwise velocity fluctuations.  
∂(ρ E) ∂(ρ E + p) uj ∂  
+ = (
σi j − τi j ) ũ i
Garnier et al. [7] carried out the first large-eddy simulation ∂t ∂x j ∂x j
(LES) of an oblique shock wave impinging upon a turbulent ∂  
boundary layer at conditions that mimic the experimental − q̃ j + Q j , (1)
∂x j
conditions of Dussauge et al. [4]. Those authors found gen-
erally good agreement between the computed mean global where ρ, u i , p, E, σi j and qi are the density, the velocity vec-
quantities, such as skin friction and displacement thickness, tor, the pressure, the total energy, the viscous stress tensor and
and experimental results. Mean and fluctuating longitudinal the heat flux vector, with
velocities were found to be in satisfactory agreement with
∂T
experimental data, while the Reynolds shear stress was some- σi j = 2 µ Si∗j , qi = −κ ,
what underestimated. ∂ xi
  (2)
SBLI at conditions close to the ones considered by 1 ∂u i ∂u j 1 ∂u k
Si∗j = + − δi j .
Dussauge et al. [4] boundary layer has been investigated by 2 ∂x j ∂ xi 3 ∂ xk
means of direct numerical simulation (DNS) by Pirozzoli and
Grasso [13]. That study confirmed the experimental finding The molecular viscosity µ is evaluated using Sutherland’s
that turbulence amplification is primarily associated with the law, and the thermal conductivity κ is related to µ through
formation of a mixing layer, resulting in a different amplifi- κ = c p µ/Pr (Pr = 0.72). The overbar denotes the spa-
cation of turbulent kinetic energy and Reynolds shear stress. tial filtering operator for LES, and the Reynolds ensemble
The same authors [14] also proposed a simple analogy to averaging operator in the case of RANS models. The tilde
estimate the acoustic pressure load at the wall from the max- is used to denote density-weighted (Favre) averages,  f =
imum of the Reynolds shear stress across the boundary layer ρ f /ρ; fluctuations with respect to Reynolds and Favre aver-
at a given location. ages are denoted with a single or double prime, respectively.
In the present paper, we report results of LES and The unresolved terms in the momentum and energy equa-
Reynolds-averaged Navier–Stokes (RANS) turbulence mod- tions are to be interpreted as either the effect of the sub-
els under conditions similar to the ones used in the TUD grid scales of motion onto the resolved ones in LES, with
experiments. However, due to the limitations of today’s τi j = ρ ( u i u j − ui uj ), Q j =  uj , or as the effect
Tuj − T
available computational resources, LES was performed at a of turbulent fluctuations on the mean flow in RANS models,
(substantially) reduced Reynolds number. The study has been with τi j = ρ u     
i u j, Q j = T u j.

123
Computational analysis of impinging shock-wave boundary layer interaction 489

2.1 Turbulence modeling The model constants are here set to cb1 = 0.1355, cb2 =
0.622, cw1 = 3.24, σ = 2/3, and the effect of the presence
A simple eddy-viscosity assumption is made for both LES of the solid wall is accounted for through the destruction term
and RANS to model τi j and Q j (last term at r.h.s. of (6)), and through the damping functions
f w , f v1 (for a full account of the model, see [23]).
1
τi j − δi j τkk = −2 ρ νt 
Si∗j , (3)
3

∂T 2.2 Boundary conditions
Q j = −κt , (4)
∂ xi
Isothermal, no-slip boundary conditions are enforced at the
where νt denotes the eddy viscosity, and the turbulent thermal bottom boundary (with Tw = Taw ), and extrapolation of all
conductivity κt is determined as κt = c p ρ νt /Prt (Prt = flow variables is performed at the outlet. The shock is arti-
0.60). ficially generated by enforcing a jump in the flow variables
so as to satisfy the Rankine–Hugoniot relations at the upper
2.1.1 LES model boundary. Periodicity is enforced in the spanwise direction
to exploit homogeneity of the flow.
The mixed-time-scale subgrid-scale (SGS) model of Inagaki Large eddy simulation of turbulent flows requires the pre-
et al. [9] is selected for LES to model the effect of the SGS scription of three-dimensional, unsteady inflow boundary
motions onto the resolved ones, as it guarantees the correct conditions to get fast transition to a fully turbulent state. In the
asymptotic behavior to the eddy viscosity at solid walls with- present work we have implemented the synthetic turbulence
out using ad hoc wall damping functions. The turbulent vis- approach proposed by Sandham et al. [17], whereby time-
cosity is determined from dependent perturbations are introduced at the inlet (super-
CMTS  posed onto a mean turbulent boundary layer profile), which
νt = −1
 kes , (5) mimic coherent boundary layer structures. The synthetic
1 + (R CT )
√ velocity disturbances in the streamwise and wall-normal
where R = kes /(  S ∗ ), and kes is the SGS turbulent directions are specified as follows
kinetic energy, kes = ( u − 
u )2 . The hat symbol indicates
the test filter, derived from the trapezoidal rule [16], and  ρw 4

u (x, y, z, t) = u∞ a j A1 j (y) F j (x, t) G j (z),
is the filter width. To avoid the uncertainties associated with ρ(y)
j=0
filtering in the (inhomogeneous) wall-normal direction [10],

filtering is only performed in the wall-parallel planes, and ρw
4
consistently the filter width is defined as  = (x z)1/2 . v  (x, y, z, t) = u∞ b j A2 j (y) F j (x, t) G j (z),
ρ(y)
As suggested by Touber and Sandham [24], the model con- j=0
stants are set to CMTS = 0.03, CT = 10. (8)

with
2.1.2 RANS model
A1 j (y) = (y/y j ) e−y/y j ,
The Spalart–Allmaras one-equation RANS model is used to
determine νt [23]. The standard formulation of the model A2 j (y) = (y/y j )2 e−(y/y j ) ,
2

involves the solution of an additional transport equation for (9)


the pseudo-eddy-viscosity ν̃,  
F j (x, t) = sin ω j (x/u c j − t) ,
∂ρ ν̃ ∂ρ ν̃
uj  
+ = cb1  S̃ ρ ν̃ G j (z) = cos 2π z/λz j + φ j .
∂t ∂x j

 
1 ∂ ∂ ν̃ The mode j = 0, is associated with inner layer streaks
+ (µ + ρ ν̃) and streamwise vortices with spanwise spacing λ+
σ ∂x j ∂x j z = 120,
 2 and length λ+ x = 520, whereas the modes j = 1, . . . , 4,
∂ ν̃ ∂ ν̃ ν̃
+ cb2 ρ − cw1 f w ρ , correspond to large eddies scaling in outer units. The ampli-
∂x j ∂x j y tudes a j , b j , and the length- and time- scales of the various
(6) modes, selected so as to provide fast transition to fully tur-
bulent state with correct statistics, are provided in Table 1. In
that is related to the eddy viscosity through
order to suppress symmetries, divergence-free random veloc-
χ3 ν̃ ity fluctuations with maximum amplitude u  /u ∞ = 4% have
νt = ν̃ f v1 , f v1 = , χ= . (7)
χ 3 + cv1
3 ν been also added within the boundary layer [11]. The spanwise

123
490 S. Pirozzoli et al.

Table 1 Parameters for


synthetic inlet forcing of LES j yj aj bj ωj uc j λz j φj
(as from (8))
0 12.0δv 1.20 −0.25 0.12u τ /δv 10u τ 120δv 0.00
1 0.25δ0 0.32 −0.06 1.2u ∞ /δ0 0.9u ∞ L z /3 5.01
δ0 and δv 0 = νw /u τ 0 are, 2 0.35δ0 0.20 −0.05 0.6u ∞ /δ0 0.9u ∞ L z /4 4.00
respectively, the boundary layer 3 0.5δ0 0.08 −0.04 0.4u ∞ /δ0 0.9u ∞ L z /5 3.70
thickness and the viscous length 4 0.6δ0 0.04 −0.03 0.2u ∞ /δ0 0.9u ∞ L z /6 0.99
scale at the inlet station

Table 2 Boundary layer


properties at the reference EXP LES RANS-LR RANS-HR
station
Reθ r 50000 2678 2833 44954
Cfr 1.49E-3 2.55E-3 2.59E-3 1.71E-3
δi∗ r /δr 0.116 0.162 0.148 0.108
Hi r 1.25 1.39 1.40 1.26
L/δr 1.92 2.85 2.51 1.67
pi denotes the inviscid L/δi∗ r 16.6 17.6 17.0 15.4
pressure jump across the
incoming shock pi /(2 τwr ) 58.2 33.8 31.1 47.9

velocity component is finally determined assuming that the


inlet velocity field is solenoidal.
The synthetic inlet formulation is found to yield a fully
turbulent state at a distance of approximately 15δ0 from the
inlet (δ0 being the boundary layer thickness at the inlet sta-
tion), with proper distribution of turbulence statistics up to
second order.

2.3 Numerical discretization

The computational strategy to solve the governing equa-


tions relies on a finite-difference approach that has been
extensively validated in previous works both for isotropic
Fig. 1 Sketch of the flow configuration for SBLI analysis
decaying compressible turbulence and for wall bounded tur-
bulent supersonic flows [12,13]. The convective fluxes are
discretized by means of a hybrid seventh-order WENO/cen- experiment is M = 1.7, the shock generator angle is θ = 6◦ ,
tral scheme, with a switch based on the Ducros sensor. Vis- and the Reynolds number of the incoming boundary layer is
cous fluxes (cast in Laplacian form) are approximated with Reθ ≈ 50000. Such large Reynolds number is far from the
second-order central differences. Time advancement is per- current capabilities of LES, and a reduced Reynolds number
formed by means of the classical four-stage, fourth-order (see Table 2) is considered instead. RANS models give access
explicit Runge–Kutta algorithm. The transport equation for to large Reynolds numbers with little computational efforts,
ν̃ in the Spalart–Allmaras model is integrated by means of and therefore two RANS calculations were conducted, one
a standard second-order TVD scheme with Van Leer limiter at conditions similar to the LES (referred to as RANS-LR),
to ensure positivity of the transported variable. and one at conditions similar to the experiment (referred to as
RANS-HR). Note that the Reynolds numbers do not match
2.4 Description of test cases exactly due to the fact that the boundary layer properties are
measured right upstream of the interaction zone, where flow
The selected flow conditions correspond to the SBLI properties cannot be precisely controlled.
experimentally investigated by Souverein et al. [22] (exper- As shown in Fig. 1, the computational domain for both
imental data are referred to in the following with the abbre- LES and RANS simulations is rectangular, and restricted
viation EXP). The free-stream Mach number in the TUD to the boundary layer development, the interaction and the

123
Computational analysis of impinging shock-wave boundary layer interaction 491

relaxation regions. The size of the computation box is L x × temperature field in the x–y plane, reported in Fig. 2a for
L y × L z = 36δ0 × 7.4δ0 × 4.1δ0 . The computational domain LES, reveals the existence of complex organized motion in
for LES is discretized with a grid consisting of 448 × 151 × the outer part of the boundary layer, which is characterized by
141 points, and the grid resolution in wall units is + x × the occurrence of turbulent bulges inclined at an acute angle
+ yw ×  + = 43.5 × 1.76 × 15.54, with 90 grid points inside
z with respect to the wall. As observed experimentally both in
the boundary layer. The grid nodes are uniformly distributed subsonic and supersonic turbulent boundary layers [19] these
in the streamwise and spanwise directions and stretched in structures are separated from the surrounding essentially irro-
the wall-normal direction using a hyperbolic sine function tational fluid by sharp interfaces that have a three-dimen-
up to the edge of the computational domain. To guarantee sional character. The flow pattern observed in the RANS
a fair comparison of LES and RANS simulations, the com- simulation is much simpler, as all turbulent fluctuations are
putational grid for the RANS-LR simulation is taken to be filtered out, resulting in a steady flow.
a two-dimensional slice of the grid used for LES. The com- Figure 3 depicts the instantaneous streamwise velocity
putational grid for the RANS-HR simulation has a similar distribution is wall-parallel planes at various distances from
resolution in terms of wall units, and therefore points are the wall (zones of instantaneous flow reversal are marked
more clustered toward the wall. A series of preliminary cal- with solid lines). The figure clearly shows the occurrence of
culations has shown that the RANS solutions presented in the elongated streaky patterns of alternating high- and low-speed
following are properly grid-converged. Different turbulence in the very near-wall region upstream of the interaction zone,
models, including the k–ω [26] and the Fares–Schröder [5] whose characteristic spacing in the spanwise direction is of
model, were also considered, which gave very similar results the order of 100 wall units, and whose length is of the order of
to Spalart–Allmaras. 1,000 wall units. Inside the interaction zone, scattered spots
The statistical properties of the flow are obtained from of flow reversal are observed, which are prevalently found
LES by averaging both in time and in the spanwise direc- upstream of the nominal impingement point. Moving away
tion. The first part of the computation (lasting approximately from the wall, the structures become less elongated, and more
200 δ0 /u ∞ non-dimensional time units) was necessary to get nearly isotropic, and the probability of instantaneous flow
a statistically steady state. Time averaging has then been per- reversal decreases. The analysis of the flow animations has
formed over additional 200 time units, which allows good further shown that excursions of the instantaneous separation
convergence of flow statistics up to second order. In the fol- line from a straight shape are associated to zones of high/low
lowing, the (inviscid) impingement point of the incoming u upstream of the shock [6].
oblique shock-wave xi (placed at a distance 28.78 δ0 from
the inlet) is used as origin for the streamwise coordinates,
3.2 Statistical properties
and lengths are made non-dimensional with respect to the
99% boundary layer thickness right upstream of the interac-
To assess the validity of LES and RANS simulations upstream
tion (δr ), conventionally measured at the streamwise location
of the interaction zone, in Fig. 4 we report the distribution of
(x − x0 )/δr = −4. Flow properties taken at such reference
the Van Driest-transformed mean streamwise velocity at the
station are hereafter denoted with the subscript r .
reference station in a semi-logarithmic plot. In the figure we
Several integral parameters are reported in Table 2 to
also report the universal logarithmic wall law
characterize the state of the boundary layer upstream of the
interaction, including the skin friction coefficient C f , the 1
‘incompressible’ displacement thickness δi∗ , and the ‘incom- u+
vd = C + log y + , (10)
k
pressible’ shape factor, Hi . The size of the interaction zone
is characterized in terms of the interaction lengthscale L, where, for supersonic boundary layers C ≈ 5.2, k ≈ 0.41
defined as the distance of the nominal shock impingement [15]. Figure 4 shows the occurrence in both LES and
point from the apparent origin of the reflected shock, L = RANS simulations of a (more or less extended) layer with
xi − xo (see Fig. 1). logarithmic dependence of velocity upon the wall distance.
In agreement with many previous LES studies (e.g.
Touber and Sandham [24]) the log-layer constant is found to
3 Results be of the order of 6. As expected, LES and RANS-LR sim-
ulations are far from the experimental data (which are only
3.1 Instantaneous properties reliable starting at about y + ≈ 200) being the Reynolds num-
bers vastly different. The structure of the incoming boundary
A qualitative understanding of the flow organization can layer is rather well reproduced by the RANS-HR simulation,
be gained from the analysis of instantaneous slices of the except for a shift in the log-law constant, which is found to
flow field, as reported in Figs. 2 and 3. The instantaneous be C ≈ 6.2 in the TUD experiment.

123
492 S. Pirozzoli et al.

Fig. 2 Iso-contours of
temperature (T /T∞ ) in x–y (a)
4
plane a LES, b RANS-LR
3 1 1.1 1.2 1.3 1.4 1.5

0
-12 -10 -8 -6 -4 -2 0 2 4

(b)
4

3 1 1.1 1.2 1.3 1.4 1.5

0
-12 -10 -8 -6 -4 -2 0 2 4

Fig. 3 Iso-contours of
streamwise velocity (u/u ∞ ) for
-.10 -.05 -.00 .05 .10 .15
LES in x–z planes at a y + = 2,
b y + = 20, c y + = 100 (the
iso-line u = 0 is shown in black) 3

0
-15 -14 -13 -12 -11 -10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5

-.1 -.0 .1 .2 .3 .4 .5 .6

0
-15 -14 -13 -12 -11 -10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5

.0 .1 .2 .3 .4 .5 .6 .7 .8 .9

0
-15 -14 -13 -12 -11 -10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5

The distributions of the skin friction, of the ‘incom- et al. [22] did not allow direct measurements of wall prop-
pressible’ shape factor and of the mean wall pressure are erties. Thus, the skin friction and shape factor data shown in
shown in Fig. 5. Note that the PIV study of Souverein Fig. 5 are based on extrapolation of the available data inside

123
Computational analysis of impinging shock-wave boundary layer interaction 493

35 0.003

30 0.0025

0.002
(a)
25
0.0015
20
0.001
15 0.0005

10 0

-0.0005
5
-0.001
0 -5 -4 -3 -2 -1 0 1 2 3 4 5
0 1 2 3 4
10 10 10 10 10
2.5
(b)
Fig. 4 Distribution of Van Driest-transformed mean streamwise
velocity upstream of interaction in inner scaling. Symbols denote data 2
from Souverein et al. [22] (open circle dual-PIV, open square boundary
layer zoom), solid line LES, dashed line RANS-LR, dashed dotted line
RANS-HR, dotted line law-of-the-wall 1.5

the boundary layer. An estimate of the wall pressure from the


PIV data was also made based on the approach to integrate the 1
-5 -4 -3 -2 -1 0 1 2 3 4 5
momentum equations introduced by van Oudheusden [25].
As expected, only the RANS-HR simulation is capable to 2
correctly predict the boundary layer structure both upstream (c) p3 /p ∞
1.8
and downstream of the interaction zone. In particular, it very
well estimates the global evolution of the shape factor, and 1.6

also predicts the correct trend of the skin friction coefficient 1.4
past the interaction. However, as also seen from Table 2, the
RANS-HR simulation significantly underpredicts the extent 1.2

of the interaction zone. 1


Both LES and RANS-LR simulations are quite far from
0.8
the experimental data, due to the vast disparity in the -5 -4 -3 -2 -1 0 1 2 3 4 5
Reynolds number, and both exhibit mean flow separation
(x − xi )/δ r
(which is more evident in the RANS simulation). In this
respect, note that the height of the mean separation bub-
ble found in numerical simulations amounts to but a few Fig. 5 Distribution of mean flow properties at the wall a skin fric-
tion, b ‘incompressible’ shape factor, c wall pressure ( p3 is the nom-
wall units, which would imply that mean flow separation (if inal pressure downstream the reflected shock). Open circle Souverein
any) would not be detected in the experiment due to lack et al. [22], solid line LES, dashed line RANS-LR, dashed dotted line
of sufficient resolution near the wall. We also note that the RANS-HR
RANS-LR simulation well agrees with LES in terms of the
prediction of the boundary layer shape across the interac- friction velocity and by the reference density profile, are
tion, and exhibits recovery of boundary layer equilibrium on compared in Figs. 6 and 7. In the selected representation, the
a similar length-scale. However, it ovepredicts skin friction structure of the velocity fields is quite similar, even though
upstream of the interaction, and slightly underpredicts it in the length-scales are somewhat different. In all cases, sig-
the relaxation zone. nificant thickening of the boundary layer is found past the
The wall pressure distribution shows significant similari- interaction zone, and strong wall-normal motion is observed
ties between experiment and LES, especially in the recovery near the foot of the impinging shock. However, note that
zone, whereas the upstream influence mechanisms is greatly the low-Reynolds-number simulations predict very gradual
over-estimated. The occurrence of an inflection point in the growth of the boundary layer even upstream of the reflected
experimental pressure distribution is likely to be due to prob- shock foot, and the latter is found to diffract and spread sig-
lems in integrating the pressure gradient equation across the nificantly above the boundary layer. In the TUD experimen-
impinging shock foot. tal data the boundary layer is seen to be undisturbed up to
The distributions of the mean velocity components (x − x0 )/δr ≈ −2, and the reflected shock is found to be
from LES and RANS simulations, scaled by the reference relatively straight. Such interaction pattern is very similar to

123
494 S. Pirozzoli et al.

Fig. 6 Map of normalized


mean defect velocity 2 2
 1 4 7 10 13 16 1 4 7 10 13 16
ρ r /ρ r w (u ∞ − u)/u τ r in the
interaction zone 1.5 1.5

1 1

0.5 0.5

0 0
-3 -2 -1 0 -3 -2 -1 0

2 2
1 4 7 10 13 16 1 4 7 10 13 16

1.5 1.5

1 1

0.5 0.5

0 0
-3 -2 -1 0 -3 -2 -1 0

Fig. 7 Map of normalized


mean wall-normal velocity 2 2

ρ r /ρ r w v/u τ r in the
-3.5 -2 -0.5 1 2.5 -3.5 -2 -0.5 1 2.5

interaction zone 1.5 1.5

1 1

0.5 0.5

0 0
-3 -2 -1 0 -3 -2 -1 0

2 2
-3.5 -2 -0.5 1 2.5 -3.5 -2 -0.5 1 2.5

1.5 1.5

1 1

0.5 0.5

0 0
-3 -2 -1 0 -3 -2 -1 0

the one found in the RANS-HR simulation, even though the tions of Pirozzoli and Grasso [13], who pointed out that the
overall scale of the interaction there is much smaller. Note generation of turbulent stress is strictly related to the unsteady
that, as pointed out by Delery and Marvin [2], and seen from shedding of coherent vortical structures associated with the
Table 2, the results would be more similar if reported in terms occurrence of inflection points in the instantaneous velocity
of the reference displacement thickness. profiles. The same scenario is also observed in the experi-
The distributions of the normalized turbulent shear stress ment, where, however, strong stresses are only found past
are reported in Fig. 8. For clarity of interpretation we point out the foot of the impinging shock, perhaps due to the above
that the Reynolds stresses reported for LES do not account for mentioned difficulties in measuring very-near-wall veloci-
the contribution of the subgrid scales, whereas in the RANS ties. The peak shearing stress is found at about the same
simulations they are reconstructed from (3). In both LES and distance of about 0.18δr in the experiment and in the
RANS simulations , strong shearing stresses are produced RANS-HR simulation, whereas in the LES and RANS-LR
near the apparent origin of the reflected shock, which are simulations the shear stress peak lies at a distance of about
subsequently transported away from the wall in the down- 0.3 δr from the wall, with substantial underestimation of the
stream direction. Such pattern is consistent with the observa- peak value.

123
Computational analysis of impinging shock-wave boundary layer interaction 495

Fig. 8 Map of normalized


Reynolds shear stress 2 2
-1 0 1 2 3 4 -1 0 1 2 3 4
−ρ r u v  /(ρ
r w u τ r ) in the
2
1.5 1.5
interaction zone
1 1

0.5 0.5

0 0
-3 -2 -1 0 -3 -2 -1 0

2 2
-1 0 1 2 3 4 -1 0 1 2 3 4

1.5 1.5

1 1

0.5 0.5

0 0
-3 -2 -1 0 -3 -2 -1 0

Fig. 9 Probability of 0.2 0.2


instantaneous velocity reversal 50 50
45 45
in the interaction zone in percent 0.15 40
35 0.15
40
35
30 30
scale 25
20
25
20
15 15
10 10
0.1 5
0.1 5

0.05 0.05

0 0
-3 -2 -1 0 -3 -2 -1 0

3.3 Unsteady properties 162


2
150 154 158 162 166 170 174 178

The analysis of the unsteady properties of the flow field is


5
useful in that it can provide insight for the prediction of the
6
16

168
fluctuating pressure loads occurring in the interaction zone.
4
1 150 6 16
LES is a natural candidate to get such information, but, as we 17
164
shall see, some reasonable estimate of the incurred unsteady 156
5 1 70
pressure loads can also be gained from RANS simulations.
0

164
16

168
We preliminarily observe that, although no mean flow 166
0
reversal is observed, the adverse pressure gradient is suffi- -3 -2 -1 0 1
ciently strong to locally cause scattered spots of reversed flow
in the interaction zone. The distribution of the intermittency
of local flow reversal, defined as the statistical frequency of Fig. 10 Map of r.m.s. pressure fluctuations from LES in the interaction
points where u < 0, is shown in Fig. 9. The figure indi- zone (dB scale)
cates that a shallow region with non-zero probability of flow
reversal is found in the LES for −2.5 ≤ (x − xi )/δr ≤ 0, levels upstream of the interaction zone are relatively uniform
with maximum probability of approximately 60% inside the across the boundary layer, and attain values of O(155) dB.
separation bubble. Such pattern differs somewhat from the Noise levels become much larger in the interaction zone,
experimental data, which show a much taller zone (extending with a maximum of about 177 db in the proximity of the
up to y/δr ≈ 0.15), which is less extended in the streamwise mean position of the incoming shock foot, and which is likely
direction. related to its unsteady movement. Large values of the fluctu-
The distribution of the r.m.s. pressure fluctuations esti- ating pressure of approximately 166 dB are also found away
mated from LES, reported in Fig. 10, shows that the noise from the wall in the region where the mixing layer develops.

123
496 S. Pirozzoli et al.

Fig. 11 Pre-multiplied pressure 10


2
2
10
spectra across the interaction 1 (a) (b)
10 1
zone obtained from LES at 10
(x − x0 )/δr = −3.9 (a),

f · P SD
0
10 0
10
−2.4 (b), 0.16 (c), 3.96 (d) -1
10 -1
10
-2
10 -2
10
-3
10 -3
-3 -2 -1 0 1
10 -3 -2 -1 0 1
10 10 10 10 10 10 10 10 10 10

2 2
10 10
1 (c) 1 (d)
10 10
0 0
10 10
-1 -1
10 10
-2 -2
10 10
-3 -3
10 -3 -2 -1 0 1
10
-3 -2 -1 0 1
10 10 10 10 10 10 10 10 10 10

Additional detail into the unsteady dynamics of the 170


interaction can be gained from the analysis of the weighted
power spectral densities (PSD) of the pressure time series,
which are reported in Fig. 11 at selected streamwise stations, 165
corresponding to points located upstream of the interaction,
near the apparent origin of the reflected shock, near the nom-
inal position of the impinging shock, and past the interaction. 160
The spectra upstream of the interaction show the typical
behavior found in canonical boundary layers, with energy
concentrated around St ≈ 1, which is the typical frequency 155
-5 -4 -3 -2 -1 0 1 2 3 4 5
for large eddies of size of the order of the boundary layer
thickness. The energy spectra across the interaction zone
exhibit significant strengthening of the lowest-frequency Fig. 12 Distribution of r.m.s. wall-pressure fluctuations. Solid line
modes, which can be an indication of the occurrence of low- LES, dashed line theory ((11) with α = 2.75) applied to RANS data
frequency motions associated with the appearance of glob-
ally unstable modes [24]. In addition, the spectral peak is
found to be less sharp and move to lower frequencies, which by the corresponding values of the maximum turbulent shear
may be due either to the thickening of the boundary layer, or stress in the wall-normal direction. In particular, they showed
to the onset of frequencies associated with the shedding of that
vortical structures in the mixing layer.
The distribution of the rms pressure fluctuations at the wall 
pw ≈ α max τx y , (11)
is reported in Fig. 12. The figure shows substantial amplifi- y

cation of noise near the shock foot by as much as 10 dB


compared to the levels upstream of the interaction, and a sud- with 1.8 ≤ α ≤ 4.3. This analogy makes it possible to exploit
den relaxation right downstream of the nominal interaction the distribution of turbulent shear stress obtained from the
point. RANS simulation to infer the r.m.s. pressure level at the wall.
To extract useful information on the unsteady pressure The data obtained from the analogy (11) (with α = 2.75 to
field from RANS simulations, we exploit the analogy match the upstream r.m.s. pressure amplitude) applied to the
between the behavior of turbulent shear stress and the r.m.s. RANS data are show in Fig. 12 as dashed lines. Very good
pressure fluctuations, first proposed by Simpson et al. [18], agreement of the semi-empirical predictions with the LES
in the context of attached and separated incompressible data is found. In particular, differences of the order of 1 dB
turbulent boundary layers, and confirmed for SBLI by are observed in the peak of the r.m.s. pressure, which is some-
Pirozzoli and Grasso [14]. Those authors found good what shifted downstream compared to LES, and differences
collapse of the wall pressure distribution when normalized of approximately 2 dB are found in the recovery zone.

123
Computational analysis of impinging shock-wave boundary layer interaction 497

4 Conclusions 6. Ganapathisubramani, B., Clemens, N.T., Dolling, D.S.: Effects of


upstream boundary layer on the unsteadiness of shock-induced
separation. J. Fluid Mech. 585, 369–394 (2007)
The interaction of a shock wave with a turbulent boundary 7. Garnier, E., Sagaut, P., Deville, M.: Large-eddy simulation
layer under conditions of incipient separation has been stud- of shock/boundary-layer interaction. AIAA J. 40(10), 1935–
ied by means of both LES and RANS simulations. LES gives 1944 (2002)
full insight into the unsteady, three-dimensional flow behav- 8. Hanjalic, K.: Will RANS survive LES? A view of perspectives.
J. Fluids Eng. 127, 831–839 (2005)
ior, and shows the occurrence of coherent structures very sim- 9. Inagaki, M., Kondoh, T., Nagano, Y.: A mixed-time-scale sgs
ilar to the ones observed in full DNS. The interaction zone model with fixed model-parameters for practical LES. J. Fluids
is shown to be dominated by strongly unsteady dynamics Eng. 127(1), 1–13 (2005)
characterized by the occurrence of spattered spots of instan- 10. Lesieur, M., Métais, O.: New trends in large-eddy simulations of
turbulence. Annu. Rev. Fluid Mech. 28, 45–82 (1996)
taneously reversed flow to which is associated the shedding
11. Li, Q., Coleman, G.N. : DNS of an oblique shock wave impinging
of vortical structures, which are the main responsible for tur- upon a turbulent boundary layer. In: Geurts, B., Friedrich, R.,
bulence amplification across the interaction zone. Overall, Metais, O. (eds.) Direct and large-eddy simulations, ercoftac
both LES and RANS simulations lead to qualitatively cor- series, pp. 387–396. Kluwer, Dordrecht (2003)
12. Pirozzoli, S., Grasso, F.: Direct numerical simulation of isotropic
rect prediction of the main features of the flow field. How-
compressible turbulence: influence of compressibility on dynamics
ever, the interaction lengthscale sensitively depends upon the and structures. Phys. Fluids 16(12), 4386–4407 (2004)
Reynolds number, and it is underestimated by RANS simu- 13. Pirozzoli, S., Grasso, F.: Direct numerical simulation of imping-
lations performed at the same Reynolds number as the ref- ing shock-wave/turbulent boundary layer interaction at M =
2.25. Phys. Fluids 18(6), 1–17 (2006)
erence experiment, whereas it is overestimated by LES and 14. Pirozzoli, S., Grasso, F.: Self-sustained oscillations in shock
RANS simulations at reduced Reynolds number. The analy- wave/turbulent boundary layer interaction. AIAA Paper 2006-2406
sis of the unsteady pressure field in the interaction zone has (2006)
shown the occurrence of very intense acoustic loads at the 15. Pirozzoli, S., Grasso, F., Gatski, T.B.: Direct numerical simulation
and analysis of a spatially evolving supersonic turbulent boundary
wall, reaching up to 168 dB near the foot of the impinging layer at M = 2.25. Phys. Fluids 16(3), 530–545 (2004)
shock. Using a simple analogy between the r.m.s. pressure 16. Sagaut, P.: Large-eddy simulation for incompressible flows: an
fluctuations and the maximum turbulent shear stress across introduction to large-eddy simulations. Springer, Berlin (2001)
the boundary layer, we have shown that some useful informa- 17. Sandham, N.D., Yao, Y.F., Lawal, A.A.: Large-eddy simula-
tion of transonic turbulent flow over a bump. Int. J. Heat Fluid
tion on the unsteady pressure field can also be extracted from Flow 24(4), 584–595 (2003)
RANS simulations, which leads to maximum error on the 18. Simpson, R.L., Ghodbane, M., McGrath, B.E.: Surface pressure
overall sound amplitude of the order of approximately 2 dB. fluctuations in a separating turbulent boundary layer. J. Fluid
Mech. 177, 167–186 (1987)
Acknowledgments The present research has been carried out within 19. Smits, A.J., Dussauge, J.P.: Turbulent shear layers in supersonic
the UFAST project (Unsteady effects in shock-wave-induced separa- flow. American Institute of Physics, New York (2006)
tion), supported by the European Union within the 6th Framework Pro- 20. Souverein, L., Dupont, P., Debiève, J.F., Dussauge, J.P., van
gram. We are indebted to Louis Souverein and Bas van Oudheusden Oudheusden, B.W., Scarano, F.: Unsteadiness characterization in a
of the Technical University of Delft for providing the experimental shock wave turbulent boundary layer interaction through dual-PIV.
data reported in the paper and for helpful discussions. The authors also AIAA Paper 2008-4169 (2008)
acknowledge the CASPUR computing consortium of the University of 21. Souverein, L.J., van Oudheusden, B.W., Scarano, F., Dupont, P.:
Rome “La Sapienza” for providing the computational resources to per- Unsteadiness characterization in a shock wave turbulent boundary
form the numerical simulations under the 2009 Standard Standard HPC layer interaction through dual-PIV. AIAA Paper 2008-4169 (2008)
Grant “Direct numerical simulation of boundary layers at high Mach 22. Souverein, L.J., van Oudheusden, B.W., Scarano, F., Dupont, P.:
numbers”. Application of a dual-plane particle image velocimetry (dual-PIV)
technique for the unsteadiness characterization of a shock wave tur-
bulent boundary layer interaction. Meas. Sci. Technol. 20, 074,003
(16 pp) (2009)
References 23. Spalart, P., Allmaras, S.: A one-equation turbulence model for aero-
dynamic flows. Rech. Aérospatiale 1, 5–21 (1994)
1. Brusniak, L., Dolling, D.S.: Engineering estimates of fluctuating 24. Touber, E., Sandham, N.D.: Large-eddy simulation of low-
loads in shock wave/turbulent boundary layer interactions. AIAA frequency unsteadiness in a turbulent shock-induced separation
J. 34, 2554–2561 (1996) bubble. Theor. Comput. Fluid Dyn. (2009) (accepted)
2. Delery, J., Marvin, J.G.: Shock-wave boundary layer interactions. 25. van Oudheusden, B.W.: Principles and application of velocime-
AGARDograph 280 (1986) try-based planar pressure imaging in compressible flows with
3. Dolling, D.S.: Fifty years of shock-wave/boundary-layer interac- shocks. Exp. Fluids 45, 657–674 (2008)
tion research: what next? AIAA J. 39, 1517–1530 (2001) 26. Wilcox, D.C.: Turbulence modeling for CFD, 3rd edn. DCW Indus-
4. Dussauge, J.P., Dupont, P., Debiève, J.F.: Unsteadiness in shock tries, La Cañada, CA (2006)
wave boundary layer interactions with separation. Aerosol Sci.
Tech. 10, 85–91 (2006)
5. Fares, E., Schröder, W.: A general one-equation turbulence model
for free shear and wall-bounded flows. Flow Turbul. Com-
bust. 73, 187–215 (2004)

123

Anda mungkin juga menyukai