Sie sind auf Seite 1von 39

Annu. Rev. Fluid Mech. 2000.

32:309345
Copyright q 2000 by Annual Reviews. All rights reserved

SHOCK WAVETURBULENCE INTERACTIONS


Yiannis Andreopoulos, Juan H. Agui, and
George Briassulis
Experimental Aerodynamics and Fluid Mechanics Laboratory, Department of
Mechanical and Aerospace Engineering, The City College of the City University of New
York, New York, New York 10031; e-mail: andre@me-mail.engr.ccny.cuny.edu

Key Words compressible flows, flow structure


Abstract The idealized interactions of shock waves with homogeneous and iso-
tropic turbulence, homogeneous sheared turbulence, turbulent jets, shear layers, tur-
bulent wake flows, and two-dimensional boundary layers have been reviewed. The
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

interaction between a shock wave and turbulence is mutual. A shock wave exhibits
substantial unsteadiness and deformation as a result of the interaction, whereas the
characteristic velocity, timescales and length scales of turbulence change considerably.
The outcomes of the interaction depend on the strength, orientation, location, and
shape of the shock wave, as well as the flow geometry and boundary conditions. The
state of turbulence and the compressibility of the incoming flow are two additional
parameters that also affect the interaction.

1. INTRODUCTION
The interaction of shock waves with turbulent flows is of great practical impor-
tance in engineering applications. A fundamental understanding of the physics
involved in this interaction is necessary for the development of future supersonic-
and hypersonic-transport aircraft, and advances in combustion processes as well
as high-speed rotor flows. In such flows the interaction between the shock wave
and turbulent flow is mutual, and the coupling between them is very strong.
Complex linear and nonlinear mechanisms are involved, that can cause consid-
erable changes in the structure of turbulence and its statistical properties and alter
the dynamics of the shock-wave motion.
Amplification of velocity fluctuations and substantial changes in length scales
are the most important outcomes of interactions of shock waves with turbulence.
This indicates that such interactions may greatly affect mixing. The use of shock
waves, for instance, has been proposed by Budzinski et al (1992) as a means of
enhancing the mixing of fuel with oxidant in ramjets. Turbulence amplification
through shock wave interactions is a direct effect of the Rankine-Hugoniot rela-
tions. However, this type of amplification should be decoupled from other effects
that also contribute to turbulence amplification such as destabilizing streamline

00664189/00/01150309$12.00 309
310 ANDREOPOULOS n AGUI n BRIASSULIS

curvature, flow separation, dilatational effects, or longitudinal pressure gradients


that may be present in flow before or after the interaction with the shock.
The outcomes of the shock-turbulence interaction depend on (a) the charac-
teristics of the interacting shock-wave-like strength, relative orientation to the
incoming flow, and location and shape, (b) the state of turbulence of the incoming
flow as it is characterized by the fluctuation levels of velocity, density, pressure,
and entropy and length scales, (c) the level of compressibility of the incoming
flow, and (d) the flow geometry and boundary conditions.
Basic understanding of the physics of such complex interactions has been
obtained through investigations of conveniently selected and reasonably simpli-
fied flow configurations. The flows to be considered here include shear free flows,
shear layers, and wall-bounded flows. Most of the work in this review is confined
to the following cases: (a) homogeneous and isotropic turbulence interactions
with shock waves (see Figure 1a), (b) constant shear homogeneous turbulence
interactions with shock waves (see Figure 1b), (c) circular jet flows interacting
with shock waves (see Figure 1c), (d) plane shear layers interacting with oblique
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

shock waves (see Figure 1d), (e) wake flows interacting with oblique or normal
shock waves (see Figure 1e), and ( f ) boundary layer interactions with oblique
shock waves (see Figure 1f). This classification of the interactions also reflects
the level of increasing complexity from the first to the last flow configuration.
We restrict our review to nominally two-dimensional interactions, albeit keep-
ing in mind that all of these interactions are three-dimensional in nature because
turbulence in all its complexity is characterized by instantaneous flow variables
that exhibit a variation in time and space. Shock waveboundary layer interactions
have been reviewed in the past by Green (1970), Adamson & Messiter (1980),
and Settles & Dodson (1991). Compressibility effects of turbulence, including
some shock wave interactions, have been recently reviewed by Lele (1994). Com-
pressibility effects in turbulent boundary layers and shear layers in the absence
of shock interactions have been reviewed by Spina et al (1994) and Gutmark et
al (1995), respectively. Birch & Eggers (1972) and Bradshaw (1977) reviewed
compressibility effects in shear layers, whereas Bradshaw (1996) provided a more
recent review of physical and modeling aspects of turbulence in compressible
flows. The reader is also referred to Morkovins (1962, 1992) reviews of com-
pressibility effects on boundary layers and shear layers. Interactions of shock
waves with boundary layers in three-dimensional geometries have been reviewed
by Settles & Dolling (1992) and Dolling (1993). The textbook by Smits & Dus-
sauge (1996) also provides an excellent account of compressibility effects in
turbulent shear layers, with chapter 10 devoted to boundary layershock wave
interactions.
Kovazsnay (1953) suggested the decomposition of turbulent fluctuations into
the vorticity, acoustic, and entropy modes. He performed a small perturbation
analysis, valid for weak fluctuations of density, pressure, and entropy, to show
how acoustic waves may accompany vorticity fluctuations and how entropy is
changed in regions of different shear. In Kovazsnays decomposition the three
SHOCK-TURBULENCE INTERACTIONS 311
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Figure 1 Schematic of flow configuration interactions with shock waves. (a) Homoge-
neous and isotropic turbulence, (b) constant shear homogeneous turbulence, (c) turbulent
jet, (d) free shear layer, (e) turbulent wake, and (f) two-dimensional turbulent boundary
layer.

modes of fluctuations are dynamically decoupled from each other to a first-order


approximation. However, interaction of the modes can take on a second-order
approximation of fluctuations, in which mode couplings arise as one mode can
be generated from the interaction of the other two modes (Chu & Kovazsnay
1958). An interaction of the modes can take place when a shock wave is present
in the flow if conversion of one mode into either of the other two is possible. In
312 ANDREOPOULOS n AGUI n BRIASSULIS

the present work, particular emphasis is given to interactions during which the
vorticity mode dominates the incoming flow.
The fundamental concept behind interactions of shock waves with turbulence
is the transfer of vorticity through shock waves. Longitudinal vorticity, for
instance, is expected to be transferred unchanged through a normal planar shock,
whereas the other components of vorticity parallel to the shock surface should
change substantially.
Discussion of the following unanswered questions is attempted in this review:

1. How much of the amplification of turbulence in interactions with shock waves


is caused entirely by the Rankine-Hugoniot conditions?
2. Why are small-size eddies amplified more than large eddies?
3. Are the length scales of the incoming turbulence reduced or amplified through
such interactions?
4. Is the dissipation rate of turbulent kinetic energy also reduced?
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

5. Why are vorticity fluctuations amplified more than velocity fluctuations?


Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

The ability of existing theoretical models to predict the flow after interactions
with shock waves is also considered: Linear interaction analysis (LIA), rapid
distortion theory (RDT), large eddy simulations, direct numerical simulations
(DNS), and Reynolds-averaged Navier-Stokes give different results even in the
simplest of flow.

2. BACKGROUND INFORMATION

The shock wave, from a simplistic point of view, can be considered as a steep
pressure gradient. Information from experiments and simulation of low-speed
flows with such pressure gradients indicate that rapid-distortion concepts hold
and, in the limit of extremely sharp gradients, the Reynolds stresses and turbulent
intensities are frozen, because there is insufficient residence time in the gra-
dient for the turbulence to alter these quantities at all (Hunt 1973, 1978; Hunt &
Carruthers 1990).
The first attempt to consider theoretically the passing of a turbulent field
through a shock wave is attributed to Ribner (1953), who decomposed the incident
disturbance shear wave into acoustic, entropy, and vorticity waves. In his LIA,
Ribner formulated the interaction of a plane vorticity wave with a shock wave as
a boundary-value problem. Figure 2 shows a schematic of the incident plane
sinusoidal shear wave crossing the shock. The wave number vector is refracted
across the shock owing to the changes in thermodynamic properties and therefore
emerges at a different angle from the incident. The condition that the phases
should remain unchanged across the shock yields that the frequency, as well as
the component of the wave number vector parallel to the shock front, are the
same upstream and downstream of the shock. Ribners analysis obtained the first
SHOCK-TURBULENCE INTERACTIONS 313

Figure 2 Schematic of the incident shear wave crossing the shock.


by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

evidence of turbulence enhancement through interactions with shock waves. His


predictions were verified by Sekundov (1974) and Dosanjh & Weeks (1964).
Several analytical and numerical studies of this phenomenon by Morkovin (1960,
1962), Zang et al (1982), Anyiwo & Bushnell (1982), Rotman (1991), and Lee
et al (1991, 1993, 1994) show very similar turbulence enhancement.
Goldstein (1978) and Goldstein & Durbin (1980) formulated a decomposition
of the disturbances into a vortical and an acoustic or potential part. This
formulation was used by Durbin & Zeman (1992) to study numerically the
response of homogeneous turbulence to bulk compression. Helmholtzs decom-
position into a vortical part, which is rotational, and a dilatational part,
which is irrotational, has been also used extensively. Finally Lee et al (1992,
1993, 1994) and Hannappel & Friedrich (1995) demonstrated a very successful
DNS of a homogeneous and isotropic turbulence interacting with a plane shock
wave.
The time-dependent, three-dimensional equations governing the interactions
that describe the conservation of mass, momentum, and total energy are

Dq ]Ui
`q 4 0, (1)
Dt ]xi

DUi ]p ]sij
q 4 1 ` , (2)
Dt ]xi ]xj

](qEt) ](qEtUi) ](pUi) ](sijUi) ]qi


` 4 1 ` 1 . (3)
]t ]xi ]xi ]xj ]xi

Et is the total energy per unit mass, defined as Et 4 e ` 12 UiUi, qi is the rate of
heat added by conduction, and e is the internal energy, sij is the stress tensor,
314 ANDREOPOULOS n AGUI n BRIASSULIS

equal to 2lSij ` kdijSkk, where k is the second coefficient of viscosity that is


related to the bulk viscosity lb through k 4 lb 1 2/3 l.

2.1 Dissipation Rate of Kinetic Energy


The transport equation for the instantaneous kinetic energy 12 Ui Ui in compress-
ible flows is

D112 U U 2i i
]p ]s
q 4 1Ui ` Ui ij . (4)
Dt ]xi ]xj

Equation 4 can be manipulated to yield

1
1
D UiUi
2 2 ](1pUj ` sijUi)
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

q 4 ` pSkk 1 sijSij, (5)


Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Dt ]xj

where the last term on the right-hand side contains the dissipation rate of kinetic
energy E converted into thermal/internal energy. Sij is the strain tensor, and the
term pSkk represents the work done by pressure forces during compression or
expansion of the flow. Both terms, the dissipation rate E 4 sij Sij and the pressure
work term, also appear with opposite sign in the transport equation for internal
energy. Since the dissipation rate is always positive or zero at any given point in
space and time, the pressure-dilatation term can, in principle, be positive or
negative.
The dissipation rate is given by

]Ui ]Uk ]Ui


E 4 sij 4 sijSij 4 2lSijSij ` k dij . (6)
]xj ]xk ]xj

After invoking Stokess hypothesis, which suggests that the bulk viscosity is
negligible, lb ' 0, Equation 6 becomes

]Uk ]Um
E 4 2lSijSij 1 2
o
l . (7)
3
]xk ]xm

The second term in the right-hand side of Equation 7 represents the additional
contribution of compressibility to the dissipation rate of kinetic energy. This term
vanishes for incompressible flows.
Because ]Uk/]xk ]Um/]xm 4 (]Uk/]xk)2, this term is always positive, and the
negative sign in front of this term in Equation 7 may erroneously suggest that
compressibility reduces dissipation. This is incorrect because the term SijSij also
contains contributions from dilatation effects, which can be revealed if one con-
siders that
SHOCK-TURBULENCE INTERACTIONS 315

1 ]Ui ]Uj
SijSij 4 XX ` ,
2 k k ]xj ]xi
where Xk Xk is the enstrophy rate. The second term in the right-hand side rep-
resents the inhomogeneous contribution for incompressible flows. For compress-
ible flows, terms related to dilatation can be extracted, and the dissipation rate
then becomes

E 4 l XkXk ` 4
o
3 lS 2kk ` 2l 3]U ]U
]x ]x
i

j
1 S 4.
i

i
2
kk (8)

The second term on the right-hand side of Equation 8 describes the direct effects
of compressibility, that is, dilatation on the dissipation rate. It is obviously zero
for incompressible flows.
The first two terms on the right-hand side of Equation 8 are quadratic with
positive coefficients and positive signs, and they are, therefore, always positive
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

or zero. The last term on the right-hand side indicates the contributions to the
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

dissipation rate by the purely nonhomogeneous part of the flow. Its time-averaged
contribution disappears in homogeneous flows. This term, in principle, can result
in negative values, and thus it can reduce the dissipation rate. This does not violate
the second law of thermodynamics as long as the total dissipation remains positive
or zero at any point in space and time. It should be noted that the dissipation term
appears as a source term in the transport equation for entropy s, which, for ideal
gases, is
Ds ]qi
qT 4 ` E. (9)
Dt ]xi
It has been customary in the past [see, for instance, Zeman (1990)] to decompose
E into a solenoidal part Es, which is the traditional incompressible dissipation and
the dilatational part Ed. In this case

E 4 Es ` Ed, Es 4 l XkXk ` 2l 3]U ]U


i

]x ]x
j
1 S 4,
j
j 2
kk Ed 4 4/3 lS 2kk.

3. VORTICITY TRANSFER THROUGH SHOCK WAVES

A better understanding of the nature of the interaction of turbulent structures and


vortex motions of turbulent flows with shock waves requires information on
important quantities such as vorticity, rate-of-strain tensor and its matrix invari-
ants, and dissipation of turbulent kinetic energy. These flow quantities, although
computationally as well as experimentally very demanding to obtain, if resolved
to proper scales are well suited for describing physical phenomena in vortical
flows. One of the fundamental questions is how vorticity is transferred through
316 ANDREOPOULOS n AGUI n BRIASSULIS

shock waves. Ribner (1953) considered the case of a vorticity wave that is trans-
mitted through the shock governed by Snells law. Hayes (1957) derived an equa-
tion for the vorticity jump across a surface of discontinuity in unsteady flows by
considering the thermodynamic relations upstream and downstream of it. The
vorticity changes are given as follows:
r
dXn 4 0, (10)

(D r
v ` vsDtr
r
dXt 4 r 3
n 2 t (qvr )d
1
q 12
1 t t
( qvn)
n)d( q)
. 4 (11)

Here the subscripts n and t represent components in the directions normal and
tangential to the surface of discontinuity, the vector rn being the unit vector normal
to the surface of discontinuity. The vector r v represents the relative normal veloc-
ity component with respect to the surface of discontinuity that is r vr 4 r vn 1
r
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

v s. The operator Dt is the tangential part of the total derivative operator or the
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

material derivative for an observer moving with and along the shock surface with
velocity r nvs ` r v t, where Dtrv t 4 [drv t /dt]t ` r
v t tr
v t and dr x t /dt 4
r r r r
] v t /]t ` vs ] v t /]n and Dt n 4 1tvs 1 v t K, where K is the curvature ten-
sor K 4 1 r n.
The first relation (Equation 10) indicates that the surface normal-vorticity com-
ponent is not affected by the shock. This is a direct outcome of the Rankine-
Hugoniot relations, which show that velocity components parallel to the shock
surface remain unchanged through it. Equation 11 shows that the vorticity change
across the shock is determined by the density jump and by the gradients, tangential
to the surface of discontinuity, of the incoming-flow-velocity.
Although Hayess theory is very significant, in many cases it is not useful
because it provides information in reference to a local system of coordinates
attached to the shock front, which may change position and shape with time
during the interaction. The position and shape of the shock wave must emerge as
part of the solution of the mathematical system of equations and is not known a
priori. Nevertheless, this theory has shown that vorticity parallel to the shock
front will be affected by the interaction with the shock.
The transport equation of vorticity is very useful in understanding the mech-
anisms associated with compressible fluid motions (in tensor notation):
1 ]q ]p ]
DXi
Dt
4 SikXk 1 XiSkk ` eiqg 2
q ]xq ]xg
` eiqg
]xq 1 q1 ]s]x 2.
gj

j
(12)

It describes four dynamically significant processes for the vorticity component


Xi, namely that of stretching or compression and tilting by the strain Sik, vorticity
generation through dilatation, baroclinic generation through the interaction of
pressure and density gradients, and viscous effects expressed by the viscous stress
term. The viscous term may also describe reconnection of vortex lines at very
SHOCK-TURBULENCE INTERACTIONS 317

small scales owing to viscosity. If the viscous term can be ignored because its
magnitude, very often, is small, then the change of vorticity of a fluid element in
a Lagrangian frame of reference can be entirely attributed to vortex stretching
and/or tilting, to dilatational effects, and to baroclinic torque.
If Equation 12 is multiplied by Xi, then the transport equation for the enstrophy
1
2 Xi Xi can be obtained:
D(1/2XiXi) 1 ]q ]p
4 SikXkXi 1 XiXiSkk ` eiqg 2 Xi
Dt q ]xq ]xg

] 1 ]sgj 1 ]Xj ]sgj


` eiqg
]xq 1
Xi
q ]xj 2
1 eiqg
q ]xq ]xj
. (13)

The physical mechanisms associated with the time-dependent changes of enstro-


phy of a fluid element in a Lagrangian frame are similar to those responsible for
the generation or destruction of vorticity. Enstrophy is a very significant quantity
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

in fluid dynamics because it is related not only to the solenoidal dissipation as


Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

was mentioned in Section 2, but also to the invariants of the strain-of-rate matrix
Sij. In addition, enstrophy is a source term in the transport equation of dilatation
Skk:
D(Skk) 1 1 ]q ]p 1 ]2p ] 1 ]skq
4 1SikSki ` XkXk ` 2 1 ` . (14)
Dt 2 q ]xk ]xk q ]xk]xk ]xk q ]xq
This transport equation relates the change of dilatation along a particle path, which
can be caused by the straining action of the dissipative motions (SikSik), as well
as by the rotational energy of the spinning motions as it is expressed by the
enstrophy 12XkXk. Pressure and density gradients, as well as viscous diffusion,
can also affect dilatation. It should be noted that transport Equation 14 reduces
to the well-known Poisson equation
1 ]2p 1
4 1SikSki ` XkXk, (15)
q ]xk]xk 2
for incompressible flows of constant density (Skk 4 0).

3.1 A Demonstration
The transport equations of vorticity and enstrophy can be used to gain further
insight into the processes involved in the interactions of shock waves with tur-
bulence, by looking at the instantaneous signals of the various quantities involved.
To demonstrate how vorticity is being transmitted through shock waves, several
experiments were carried out at City College of New York by Agui (1998), in
which a subminiature multiwire probe was used to measure vorticity and other
velocity gradients in an interaction of a grid-generated turbulence with a traveling
normal shock. This probe, a modification of the probe used by Honkan & Andreo-
poulos (1997), is capable of measuring time-dependent velocity gradients with a
318 ANDREOPOULOS n AGUI n BRIASSULIS

resolution of ;1020 Kolmogorov viscous scales in these high-speed flows. Fig-


ure 3 shows signals of longitudinal velocity component U1, lateral velocity com-
ponent U3, and longitudinal and lateral vorticity components X1 and X2,
respectively, dilatation 1Skk 4 1/qDq/Dt, dilatational dissipation Ed 4 SkkSkk,
and solenoidal dissipation/enstrophy Es 4 XiXi. Each signal has been normalized
by its rms value upstream of the interaction so that signals with relatively large
fluctuations are reduced and signals with relatively small fluctuations are
enhanced. Thus all signals have been brought to about the same level. This nor-
malization also helps to observe whether the signal is amplified after the inter-
action because the rms value of this portion of the signals is the gain G, which
is defined in terms of a representative variable Q as GQ 4 Qrms,d /Qrms,u, where
the superscripts u and d denote upstream and downstream of the interactions,
respectively.
Each of the signals, with the exception of that of U1, has been displaced by
multiples of 10 rms units to provide better visual aid. The shock wave location
is evident on the longitudinal velocity signal, where its value drops substantially.
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

An inspection of the level of fluctuations, after the passage of the shock and the
actual computation of their rms values, indicates that some signals are amplified
and some are not. The longitudinal vorticity X1 and lateral velocity U3 signals
are only slightly affected by the interaction. The computed data also show a 2
5% reduction in the rms values, which practically indicates that within the exper-
imental uncertainty, no changes occur in the transmission of X1 and U3 through
the shock, as is expected from the Rankine-Hugoniot relations. Longitudinal
velocity fluctuations and lateral vorticity fluctuations X2 are substantially ampli-
fied through the interaction, with gains of ;1.4. Fluctuations of dilatational dis-
sipation SkkSkk and solenoidal dissipation also show gains of the order of ;1.2.
These two dissipation signals exhibit strong intermittent behaviors that are char-
acterized by bursts of high-amplitude events. These, sometimes, reach values of
up to 8 rms units after which less violent time periods occur. It should be men-
tioned that fluctuations of dilatation Skk are very small in this experiment, only
10% of the fluctuations of vorticity. These values are typical in flows with low
fluctuations of Mach number. Consequently, baroclinic vorticity generation is
negligible because pressure fluctuations are also small. Thus, the only source of
vorticity change, in addition to viscous diffusion, is through the stretching or
compression and tilting term SikXk. A closer look at the components of this vector
indicates that the level of fluctuations of the terms S11X1, S22X2, and S33X3, which
are the terms indicating vortex stretching or compression, decreases by ;30% on
average. Clearly, compression by the shock wave reduces the level of vorticity
fluctuations associated with the so-called mechanism of vortex line compression
on all vorticity components. However, tilting of vorticity components by the
action of the shear Sik increases through the interaction by various amounts on
each of the three vorticity components so as to compensate for the deficit caused
by the vortex line compression and to further increase in the case of the two
components parallel to the shock front. Thus, the complete stretching term
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

319

Figure 3 Time-dependent signals in grid turbulence-shock interactions at Ms 4 1.2.


320 ANDREOPOULOS n AGUI n BRIASSULIS

increases after the shock. It should be noted that the DNS of Lee et al (1992)
agree with this finding.

4. SHOCK INTERACTIONS WITH HOMOGENEOUS


AND ISOTROPIC TURBULENCE

The interaction of an isotropic and homogeneous turbulent flow with a strong


axisymmetric disturbance, such as a normal shock, is the best paradigm case of
a flow where the geometry is reasonably simplified and basic physics of such
interactions can be investigated. Traditionally, this flow has been used as a test
case where a turbulence model of the large eddy simulations or Reynolds-
averaged Navier-Stokes class can be evaluated. The absence of turbulence pro-
duction and the simplified flow geometry can expose the models strengths and
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

weaknesses.
Experimental realization of a homogeneous and isotropic flow interacting with
a normal shock in the laboratory is a formidable task. There are two major dif-
ficulties associated with this: setting up a compressible and isotropic turbulent
flow is the first one, and the generation of a normal shock interacting with flow
is the second. These two problems may not be independent from each other. As
a result of these difficulties, two different categories of experiments have been
performed. The experimental arrangement in shock tubes offers the possibility of
unsteady shock interactions with isotropic turbulence of various length scales and
intensity. This category includes the experiments of Hesselink & Sturtevant
(1988), Keller & Merzkirch (1990), and Honkan & Andreopoulos (1992). In the
last experiments the induced flow behind the incident shock wave passes through
a turbulence-generating grid and then interacts with the shock reflected off the
end wall. A unique facility has been developed at City College of New York in
which the flow Mach number can, to a certain extent, be controlled independently
from the shock wave strength. This has been achieved by replacing the end wall
of the shock tube with a porous wall of variable porosity. The large size of this
facility allows measurements of turbulence with high spatial and temporal reso-
lution (see Briassulis & Andreopoulos 1996; G Briassulis, JH Agui and J Andreo-
poulos, submitted for publication; Agui 1998). Thus, even shock interactions with
incompressible flows can be generated.
Configuring a homogeneous and isotropic turbulence interacting with a normal
shock in a supersonic wind tunnel appears to be more difficult than in a shock
tube. A turbulence-generating grid or other device is usually placed in the flow
upstream of the converging-diverging nozzle. The flow then interacts with a sta-
tionary shock produced by a suitable shock generator located farther downstream
in the working section. The problem with such configurations is that the flow an-
isotropy is substantially increased through the nozzle.
SHOCK-TURBULENCE INTERACTIONS 321

Although DNS of homogeneous and isotropic turbulence interacting with the


shock waves are restricted to low Reynolds numbers, they have been very useful
in providing physical insights of the mechanisms involved. These computations
can provide results that sometimes are very difficult to obtain experimentally, or
they can be used as diagnostic tools. Lee et al (1992), for instance, computed
results containing pressure fluctuations inside the flow or entropy fluctuations
through shocks, which currently seem impossible to obtain experimentally
because of lack of appropriate experimental techniques for measuring these
quantities.
There are three different experiments of homogeneous isotropic turbulence
interactions with shocks all carried out in supersonic wind tunnel arrangements
in France. Debieve & Lacharme (1985) carried out an oblique shock interaction
with turbulence generated inside the plenum chamber and developed in a flow
with a Mach number of 2.3. A threefold amplification of turbulence was measured
close to the shock, which was attributed to shock oscillations and distortions
caused by the incoming turbulence. Farther downstream, a residual gain of tur-
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

bulence fluctuations of 1.25 was measured.


In the work of Blin (1993), described by Jacquin et al (1993), the grid used to
generate turbulence was also acting as a sonic throat to accelerate the flow to
Mach number 1.4. The position of the shock was controlled by a second throat
downstream and by the suction of the wall boundary layer. The tabulated data
provided by O Leuchter (private communication, 1997) indicated a considerable
deceleration of the incoming flow before the shock. The longitudinal mean veloc-
ity from 469 m/s at x 4 0.03 m becomes 380 m/s at x 4 0.32 m. This deceleration
is most probably caused by Mach waves present in the flow or by pressure waves,
as they are called by Jacquin et al, which emanate from the grid. As a result, the
turbulence level of the incoming flow may be higher than it ought to be because
it is known that supersonic flow deceleration through gradual compression is
associated with turbulence augmentation. This may also explain the lack of tur-
bulence amplification through the shock found in this work. Jacquin et al, how-
ever, attribute this to the role of pressure fluctuations in the energy exchange
between kinetic energy and potential energy in pressure fluctuations. In addition,
the use of laser Doppler anemometry to obtain the data may have contributed to
this because it is known that laser Doppler anemometry in compressible flows
overestimates turbulence intensities. The probe volume of the laser Doppler ane-
mometry was 10 mm 2 5 mm, and it is considered rather large to resolve accu-
rately the turbulence of the flow field. In terms of Kolmogorov length scales, the
probe volume appears to be 500 2 250. Thus, the existence of Mach waves
inside the flow has contaminated the flow, and instrumentation problems may
have biased the results, which did not show the expected amplification of tur-
bulence after the interaction with the shock.
A multinozzle turbulence generator was used in the Mach 3 experiments of
Alem (1995) published by Barre et al (1996). A normal shock was formed by the
interaction of two oblique shock waves of opposite directions. The flow after the
322 ANDREOPOULOS n AGUI n BRIASSULIS

interaction is highly accelerated because of the two shear layers/slip lines in the
boundaries of the useful flow region. The velocity is 200 m/s at x 4 10 mm after
the shock and increases to 250 m/s at x 4 20 mm, which results in an acceleration
of ]u/]x 4 5400 s11.
This level of acceleration is very strong and is expected to reduce turbulence
intensities after the interaction. Thus, the amplification levels found in this work
are probably contaminated by the additional effects of acceleration in the subsonic
flow downstream of the interaction. Another weakness of this data set is that the
level of turbulence intensity at the location of the shock is extremely low, '0.4%.

4.1 Turbulence Amplification Through the Interaction


Amplification of turbulence is one of the major features of shock-turbulence inter-
actions. LIA predicts amplification of turbulence as long as fluctuations of pres-
sure, velocity, and temperature upstream of the shock are small so that the shock
front is not substantially distorted and the Rankine-Hugoniot conditions can be
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

linearized. In interactions in which strong distortions of the shock front are evident
because of higher turbulence intensity, the use of LIA is fully justified. DNS data
in interactions with weak shocks (Lee et al 1993, Hannappel & Friedrich 1995)
and Euler simulations (Rotman 1991) also predict amplification of turbulence.
Almost all experiments confirm qualitatively this analytical and computational
result. The experiments of Jacquin et al (1993) report no significant enhancement
of turbulent kinetic energy. However, as was mentioned earlier, the deceleration
in the flow upstream of the interaction because of the existence of Mach waves
may have contributed toward overestimating the level of turbulence upstream of
the interaction.
In the recent experimental work of Zwart et al (1997), a clear increase in
turbulent intensities is reported. These authors, in their study of the decay behav-
ior of compressible gridgenerated turbulence, encountered shock waves in the
range of flow Mach numbers of 0.7 , M , 1.0. They attributed the increased
turbulent intensities in this regime to shock wave unsteadiness.
Typically, the amplification of turbulent fluctuations depends on the shock
wave strength, the state of turbulence of the incoming flow before the interaction,
and its level of compressibility. Figures 4 and 5 show the amplification of velocity
and vorticity fluctuations, respectively, as a function of the shock strength Ms.
The experimental data have been obtained by Agui (1998). Two different grids
were used in these experiments at two different Mach numbers each. The data
shown in Figure 4 clearly suggest an amplification of longitudinal velocity fluc-
tuations, whereas the fluctuations of the other two components are slightly
decreased or remain unchanged through the interactions. The amplification data
of vorticity, shown in Figure 5, indicate a substantial enhancement of vorticity
fluctuations in the direction parallel to the shock front with no amplification or
small attenuation of longitudinal vorticity fluctuations. The amplification data of
the lateral component of vorticity fluctuations shown in Figure 5 represent the
SHOCK-TURBULENCE INTERACTIONS 323
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Figure 4 Amplification of velocity fluctuations in grid turbulenceshock interactions.


Experimental data of Agui (1998), LIA data of Lee et al (1993), RDT data of Jacquin et
al (1993), and DNS data of Lee et al (1993).

average of the two components parallel to the shock front. The results of the LIA
of Lee et al (1993) as well as their DNS data are also plotted in Figure 5. LIA
shows that lateral vorticity fluctuations increase with the shock strength after the
interaction with the shock, and the experimental data tend to confirm this.
The DNS data of Hannappel & Friedrich (1995) show that the amplification
of the vorticity fluctuations in the lateral direction increases with compressibility
in the upstream flow, whereas the amplification of longitudinal velocity fluctua-
tions is reduced by the same effect. They also found that longitudinal vorticity
fluctuations are slightly reduced through the shock, as it has been found in the
experiments of Agui (1998).
If the damping effects of pressure are ignored, the RDT of Jacquin et al (1993)
leads to the following simple relations for the amplifications of longitudinal veloc-
ity and lateral vorticity fluctuations:

31 `3 2C 4,
2
G u2 4 C 2 , G x2 4

where C is the density ratio C 4 qd /qu. The results of the RDT are also plotted
324 ANDREOPOULOS n AGUI n BRIASSULIS
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Figure 5 Amplification of vorticity fluctuations in grid turbulenceshock interactions.


Experimental data of Agui (1998), LIA data of Lee et al (1993), RDT data of Jacquin et
al (1993), and DNS data of Lee et al (1993).

in Figures 4 and 5. The density ratio in the experiments by Agui (1998) was
between 1.25 and 1.7.
The early work of Honkan & Andreopoulos (1992) showed that amplification
of urms, Gu, also depends on the turbulence intensity and length scale of the incom-
ing isotropic turbulence. The work of Briassulis and Andreopoulos (1996a,b) and
Agui (1998) also confirm this finding. The amplification of turbulent kinetic
energy across shocks seems to decrease with increasing Mt. Both experimental
data from the above-mentioned studies and the DNS data of Lee et al (1993) and
Hannappel & Friedrich (1995) agree with this finding. Thus, the outcome of the
shock turbulence interaction depends also on the compressibility level of the
incoming flow.
The effect of the shock strength on the velocity fluctuations is shown in Figure
6, where the data of Briassulis (1996) are plotted. For the 1 2 1 (cells/in2) grid,
it appears that the amplification of turbulence fluctuations defined as Gu 4 urms,d /
urms,u increases with downstream distance for a given flow case and interaction.
Thus, turbulence amplification depends on the evolution of the flow downstream.
As the Mflow increases, Gu also increases. For finer grids the effects of shock
interaction are felt differently. For the 2 2 2 grid, for instance, the data show
SHOCK-TURBULENCE INTERACTIONS 325
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Figure 6 Amplification of velocity fluctuations in interactions of grid-generated turbu-


lence with a normal shock. Experimental data of Briassulis (1996).

that, for a practically incompressible upstream flow interacting with a rather weak
shock, amplification of turbulence occurs at x/M . 35. The amplification is
greater when Mflow increases to 0.436. However, when compressibility effects in
the upstream flow start to become important, no amplification takes place (Gu is
;1).
Some more dramatic effects of compressibility are illustrated in Figure 7, in
which the amplification Gu is plotted for a finer grid with mesh size 5 2 5. The
interaction of a weak shock with a practically incompressible turbulent flow pro-
duces the highest amplification of velocity fluctuations, with Gu reaching a value
close to 2. As the Mflow increases, Gu decreases, and at Mflow 4 0.576 a slight
attenuation occurs at downstream distances. It is, therefore, plausible to conclude
that for high shock strength (high Mach number), compressibility effects control
the velocity fluctuations, which are generated by fine grids, and no amplification
of turbulent kinetic energy is observed.
Hannappel & Friedrich (1995) have also shown in their DNS work that com-
pressibility effects in the upstream reduce turbulence amplification significantly.
The LIA of Ribner (1953), which was initially developed for an incompressible
isotropic turbulent field, predicts amplification of turbulence fluctuations.
Recently, Mahesh et al (1997) have shown that LIA as well as DNS may show a
complete suppression of amplification of kinetic energy if the upstream correla-
tion between velocity and temperature fluctuations is positive. It is therefore pos-
326 ANDREOPOULOS n AGUI n BRIASSULIS
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Figure 7 Amplification or attenuation of velocity fluctuations in grid-generated turbu-


lence with a normal shock. Experimental data of Briassulis (1996).

sible that, for very fine grids and high Mach number flows in which the dissipation
rate of turbulent kinetic energy is high, entropy or pressure fluctuations may be
responsible for completely suppressing turbulence amplification.
Briassulis (1996) found amplification of velocity fluctuations after the inter-
action in all cases involving turbulence produced by coarse grids. This amplifi-
cation increases with shock strength and flow Mach number. For fine grids,
amplification was found in all interactions with low Mflow, whereas at higher Mflow,
amplification was reduced, or no amplification of turbulence was evident. These
results indicate that the outcome of the interaction depends strongly on the
upstream turbulence of the flow.

4.2 Turbulence Length Scales Through the Interaction


As mentioned by Lele (1994), there is some disagreement between experimental
results and DNS data as far as how the various length scales of turbulence are
affected by the interaction with the shock wave. The DNS data of Lee et al (1993)
and Hannappel & Friedrich (1995) indicate that all characteristic length scales,
namely longitudinal- and lateral-velocity, integral-length scales, longitudinal-
velocity microscales, and dissipation-length scales, as well as integral-length
scales and microscales of density fluctuations, decrease through shock interac-
tions. The experimental data of Keller & Merzkirch (1990) show that the density
microscale increases across the shock. Hannappel & Friedrich (1995) also show
that the reduction in the Taylor microscale parallel to the shock front is weaker
SHOCK-TURBULENCE INTERACTIONS 327

by a factor of 2 when the compressibility level of the incoming turbulence is high.


On the other hand, the dissipation length scale in the experiment of Honkan &
Andreopoulos (1992) was also found to increase after the interaction. DNS results
of Lee et al (1994) have indicated a small increase of dissipative-length scales
through weak shock interactions. Thus, there is no agreement among various
researchers on how shock interactions affect the length scales. Intuitively, one
should expect that compression should reduce length scales.
To resolve the disagreement between experiments and DNS, Briassulis (1996)
carried out detailed space-time correlation measurements by using a rake of six
parallel wires and three temperature wires separated by 1 mm. To estimate the
length scales in the longitudinal n1 direction and normal n2, the cross correlation
coefficients rij (nk) 4 ui (x)uj (x ` nk)/!u2i (x)!u2j (x ` nk) were evaluated by two-
point measurement in the n2 direction and from autocorrelations in the n1 direction
after invoking Taylors hypothesis.
The data show that the integral length scale L11(n1) increases with downstream
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

nondimensional distance x/M for all investigated cases in the flow before the
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

interaction. It is also shown that L11 for Mflow 4 0.475 is higher than for
Mflow 4 0.36. However, if the flow Mach number increases to Mflow 4 0.6 and,
therefore, stronger compressibility effects are present, then the integral-length
scale drops.
After the interaction with the shock wave, the distribution of L11(n1) is more
complicated. All the scales are reduced considerably. However, the reduction of
the larger scales is greater. This is shown in Figure 8, where the attenuation ratio
GL11 4 L11,d (n1)/L11,u(n1) is plotted. At large x/M, where the initial scales were
the largest, the reduction is dramatic. Thus, once again, it is found that amplifi-
cation or attenuation is not the same for all initial length and velocity scales. It
is interesting to observe that the stronger the shock strength, the greater the atten-
uation of the longitudinal-length scales.
The two-point correlation r11(n2) in the lateral direction n2 of the longitudinal-
velocity fluctuations is shown in Figure 9. Not all the curves cross the zero line,
and, therefore, it is very difficult to integrate them to obtain the classically defined
length scale in the lateral direction. However, the slopes of these curves are indic-
ative of their trend. It is rather obvious that the length scales before interaction
are reduced with increasing flow Mach number. This behavior is very similar to
that of L11(n1). After the interaction, however, the length scale L11(n2) increases
in the first two cases and decreases in the strongest interaction.
To investigate the effect of initial conditions on this correlation at the highest
flow and shock Mach number, where the lateral scales are shown to reduce (Figure
9), various grids were used. The data shown in Figure 10 indicate that the cor-
relation increases substantially in the finest grid, 8 2 8 with the lowest Rek, after
the interaction for the range of length scales investigated. However, the coarser
2 2 2 grid with the highest Rek 4 737 shows the greatest attenuation in the
lateral integral scale of turbulence after the interaction.
328 ANDREOPOULOS n AGUI n BRIASSULIS
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Figure 8 Ratio of the longitudinal integral-length scales for various experiments in grid-
generated turbulence interactions with a normal shock. Experimental data of Briassulis
(1996).

Figure 9 Spatial correlation in the lateral direction for three different flow cases in grid-
generated turbulence interactions with a normal shock. Experimental data of Briassulis
(1996).
SHOCK-TURBULENCE INTERACTIONS 329
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Figure 10 Spatial correlation in the lateral direction for three different downstream loca-
tions in grid-generated turbulence interactions with a normal shock. Experimental data of
Briassulis (1996).

Thus, integral-length scales in the longitudinal direction were reduced after


the interaction in all investigated flow cases. The corresponding length scales in
the normal direction increased at low Mach numbers and decreased during
stronger interactions. It appears that, in the weakest of the present interactions,
the eddies are compressed in the longitudinal direction drastically, whereas their
extent in the normal direction remains relatively the same. As the shock strength
increases, the lateral length scale increases, whereas the longitudinal-length scale
decreases. At the strongest interaction of the present cases, the eddies are com-
pressed in both directions. However, even at the highest Mach number case, the
issue is more complicated because amplification of the lateral scales has been
observed in fine grids. Thus, the outcome of the interaction strongly depends on
the initial conditions.
The dissipative length scale was found to increase in several cases, and the
disagreement with the DNS remains an issue, although it may be attributed to the
difference in Rek. The Rek of the DNS was between 12 and 22, which is consid-
erably lower than the values between 200 and 750 achieved in the experiments
of Briassulis (1996) and may be the cause of this disagreement between experi-
ments and DNS. The difference between experiments and DNS is also noticeable
if the Reynolds number
330 ANDREOPOULOS n AGUI n BRIASSULIS

(quiui)2
ReT 4
le
is considered. ReT in the experiments is ;4000, whereas in DNS it is only 750.
As a result, the timescale of turbulence uiui/e and its ratio to the timescale imposed
by the shock 1/S11, uiuiS11/e, are quite different between DNS and experiments.
The results of the experiments of Briassulis (1996) clearly show that most of
the changes, either attenuation or amplification of quantities involved, occur for
large x/M distances in which the length scales of the incoming flows are high and
turbulence intensities are low. Thus, large eddies with low-velocity fluctuations
are affected the most by the interaction with the shock.

5. HOMOGENEOUS SHEARED TURBULENCE


THROUGH SHOCKS
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

The only work available in this type of interactions is that of Mahesh et al (1997),
who used RDT to demonstrate that the amplification of turbulent kinetic energy
depends on anisotropy of the Reynolds stress tensor of the incoming flow and the
ratio of shearing rate to that of compression. The dependence of turbulence ampli-
fication on the initially anisotropy is reduced with the obliqueness of the shock
wave. A change in the sign of the initial shear stress has been observed for
sufficiently strong interactions. This type of idealized interaction has yet to be
configured experimentally.

6. TURBULENT JET FLOWS AND SHOCK WAVE


INTERACTIONS

Interactions of jets with shock waves are considerably more complex than the
previously considered homogeneous flow interactions because inhomogeneity
and anisotropy are additional parameters to consider, which may change the flow
physics completely. The variable shear, for instance, in the shear layer emanating
from the jet is an additional parameter that is responsible for the production of
large amounts of turbulent kinetic energy before the interaction. One obvious
question is how this shear is affected by the interaction with the shock wave.
Jacquin & Geffroy (1997) have shown that this shear is decreased through the
interaction.
Parameters such as the convective Mach numbers Mc, the velocity ratio UJ /
Ua, and density ratio qJ /qa have been considered in this type of interaction in
addition to the shock strength, compressibility level, and velocity and length scale
of the incoming turbulence. A schematic of the flow interaction and the param-
eters involved is shown in Figure 11. Baroclinic generation of vorticity may take
place due to the nonalignment of the density and pressure gradients. Density
SHOCK-TURBULENCE INTERACTIONS 331
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Figure 11 Schematic of the turbulent jetshock interaction.

gradients can be generated by heating the jet flow, as in the experiments of Jacquin
& Geffroy (1997), or by using different gases, as in the case of Cetegen & Her-
manson (1995), Hermanson & Cetegen (1998) or Hermening (1999). The pressure
gradient across the shock interacts with the density gradient across the interface
between the jet and the ambient fluid to generate additional vorticity.
Two types of configurations have been used in the past to study experimentally
shock-jet interactions: steady interaction in supersonic wind tunnels and unsteady
interaction with traveling shock waves in shock tube. Jacquin & Geffroy (1997)
designed a facility with a jet flow exiting at the throat of a nozzle. The jet and
the sonic coflowing stream were further expanded to M 4 1.6. Density ratios
from 0.33 to 1 were achieved through heating of the jet flows. The velocity ratio
ranged from 0.58 to 1 and the convective Mach number was #0.42. Turbulence
amplification was measured in the quasihomogeneous region of the jet, whereas
turbulence kinetic energy was decreased in the mixing layer region of the flows.
According to Jacquin and Geffroy, this is, most probably, caused by baroclinic
vorticity effects resulting from a reduction in the near shear by a factor of 2.0.
No substantial effects of jet temperature, that is, density ratio on amplification or
reduction of turbulent kinetic energy, were identified in experiments. Their flow
visualization indicated that the shock front is not distorted by the shock. Thus,
baroclinic vorticity generation or destruction can take place in the absence of
shock front deformations. Substantial shock front deformations, Mach reflections,
and additional shocks can be developed when an overexpanded jet interacts with
a normal shock (see Jacquin et al 1991). Large amplification of turbulence has
been found in this case.
332 ANDREOPOULOS n AGUI n BRIASSULIS

Considerable generation of baroclinic vorticity is reported by Hermanson &


Cetegen (1998) in their experiments on nonuniform-density jets interacting with
traveling shock waves. Density ratios of 0.0595 (helium jet) through 1.31 (CO2
jet) were achieved in these shock tube experiments. Figure 12 shows an image
of the flow containing a vortical structure produced behind the traveling shock
wave for a helium jet interaction obtained at x/d 4 50 (from JC Hermanson &
BM Cetegen, private communication, 1999). Planar Mie scattering of planar laser
light was used to visualize the flow, which was seeded with mineral oil. The
structure contains vorticity with a sign corresponding to that of jet-like vortices
formed by jets with velocities considerably higher than the coflowing fluid veloc-
ity behind the traveling shock wave. The formation of this structure is attributed
by Hermanson & Cetegen to baroclinic generation of vorticity. There is some
apparent inconsistency between the experiments of Jacquin & Geffroy (1997),
who found that baroclinic effects reduce vorticity after the interaction with the
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Figure 12 Helium jetshock interaction in a shock tube (JC Hermanson & BM Cetegen,
private communication, 1999).
SHOCK-TURBULENCE INTERACTIONS 333

shock, and the experiments of Hermanson & Cetegen (1998), who found that the
same effects enhance vorticity after the interaction. This disagreement may be
resolved if the baroclinic term of the vorticity transport Equation 12

1 ]q ]p
BXi 4 eiqg
q2 ]xq ]xg
4 px
1
q 12
is considered. BXi reaches maxima or minima when the two gradients p and
(1/q) are at 908 or 2708 to each other. If only vorticity X3 is considered, then

1 ]q ]p ]q ]p 1 ]q ]p
BX3 4 3
q ]x1 ]x2
2
1
]x2 ]x1
1 24
q ]x2 ]x1
.

The first term can be neglected because pressure is expected to change very little
in the normal direction owing to the thin shear layer approximation. In the exper-
iments of Jacquin & Geffroy (1997), ]p/]x1 is positive across the shock and ]q/
]x2 is positive in the upper half of the flow (see also Figure 11) and negative in
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

the lower half of the flow. This shows that BX is negative in the upper half, where
the initial vorticity X3 is positive, and positive in the lower half of the flow, where
X3 is negative. Therefore, for stationary shocks such as in the experiments of
Jacquin & Geffroy, baroclinic effects tend to suppress vorticity; that is, the bar-
oclinic torque tends to bring the vorticity toward zero values. In the experiments
by Hermanson & Cetegen, the shock wave travels in the same direction as the jet
flow, and it introduces a negative pressure gradient ]p/]x1 into the flow (for a
stationary observer traveling with the shock), and, because the density gradient
is the same as in the previous case, all contributions from BX tend to enhance
vorticity X3. In that respect, there is no disagreement between the two experi-
ments. Hermanson & Cetegen found no such vortical structures in their experi-
ments with air or CO2 jets.
No distinct vortex structures were observed in the experiments by Hermening
(1999), who used helium, air, and krypton gases to generate baroclinic vorticity
of various signs because helium is lighter than air and krypton is heavier than air.
Two jet configurations were used in these experiments. In the first one, the jet
fluid flowed in the same direction as the traveling shock wave. In the second
configuration the shock traveled in the opposite direction to that of the jet flow.
Figure 13 shows several images of the interaction of a shock wave moving in the
same direction as the air jet flow obtained by Hermening. The images were
obtained from different experiments with the same bulk flow parameters and with
the shock wave located at different positions from the jet exit. The jet flow had
a Mach number Mj 4 0.9, and the shock wave was traveling with a Mach number
of Ms 4 1.33. Talcum powder with particles ,1 lm in diameter was used to
seed the flow, and a Nd:YAG laser was used to visualize the flow. The first image
(Figure 13a) shows the jet flow before the interaction. The jet flows downstream,
while it spreads laterally and its velocity decreases. The complexity of the inter-
action is shown in Figure 13b and c. Figure 13b shows the flow during the inter-
334
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Figure 13 (ac) Images of the interaction of a MS 4 1.34 shock wave with a MJ 4 0.9 coflowing jet. Flows from left to right. Images are from
Hermening (1999). (d) Schematic of the interaction.
SHOCK-TURBULENCE INTERACTIONS 335

action with the shock wave located at 5 diameters downstream of the jet exit;
traveling from left to right. Figure 13c shows the flow considerably after passage
of the shock, which is located outside the viewing area. There are three regions
of interest in the flow during the interaction as shown in Figure 13b. The first
region is characterized by a stenosis of the jet cross section observed at a distance
between 2 and 3 diameters downstream of the exit. This region is followed by a
region closer to the exit, with a jet stream of apparently larger diameter than the
previous one, and it is proceeded by a third region immediately behind the shock
wave, where spreading of the flow can be observed. This structure of the jet can
be explained if one considers the time-dependent motion of the shock from the
time it is located at the jet exit until the time it has left the near-field region. The
shock wave is characterized by an induced flow with higher pressure and higher
density behind it than ahead of it. This changes the velocity, pressure, and density
of the ambient fluid, which is now put into motion. Thus, the imposition of higher
ambient pressure compresses the jet fluid to a higher density, which reduces its
cross section and causes the apparent stenosis. At the same time, the higher ambi-
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

ent pressure requires a reduction of the exit Mach number Mj because the pressure
ratio of the ambient pressure to the jet stagnation pressure, pa /poJ, has increased.
Thus, the jet velocity UJ is reduced because the stagnation pressure of the jet is
maintained as constant by a plenum pressure controller. The new velocity of the
jet is sightly lower than the new ambient velocity Ua, which results in no spreading
of the jet into the ambient coflowing fluid and the formation of a wakelike
vortex between the readjusted jet and the region with the stenosis, as shown in
Figure 13c.
The downstream region, however, exhibits a different behavior, which indi-
cates no stenosis of the jet flow after the shock. Before the interaction with the
shock, this region is characterized by a decaying mean velocity, high shear, and
increased spreading of the jet, which results in the production of substantial levels
of turbulence. It appears that, after the interaction, considerable mixing has taken
place, and the turbulence most probably has increased. Figure 13d depicts a sche-
matic illustration of the interaction of the jet flow with the shock wave, which
includes all three regions of the flow discussed above.
Images of the counterflow configuration of a jet-shock interaction, obtained
by Hermening (1999), are shown in Figures 14a and b. The air jet flows from
right to left, whereas the shock travels in the opposite direction. The jet and shock
Mach numbers are the same as in the previous configuration. The most striking
feature of this interaction is the substantial level of spreading of the jet flow
observed after the shock, which is accompanied by high levels of turbulence.
Flow reversal is also expected at some distance from the jet exit.
The experiments of Hermening (1999) with helium and krypton indicated
some similar behavior. The initial spreading rate of the helium jet before the
interaction was higher than that of the air jet. As a result, substantially more
mixing was observed after the interaction with the shock than in the case of air.
The spreading rate of the krypton jet, which was initially smaller than that of the
336 ANDREOPOULOS n AGUI n BRIASSULIS

Figure 14 Airjet interaction with a traveling shock wave in a counterflow configuration.


Jet flows from right to left, while shock wave travels from left to right; MJ 4 0.9, MS 4
1.33. Images are from Hermening (1999).
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

air jet. The shock interaction in the coflowing configuration increased the mixing
compared with the initial jet. However, substantially more mixing was observed
in the counterflow configuration, which supports the notion explained above that
baroclinic effects, in this case of a heavier-than-ambient jet, enhance the vorticity
of the jet. In this case the interaction of the density and pressure gradients brings
about additional vorticity into the flow, which enhances the mixing.
The mixing enhancement shown in the experiments of a helium jet interacting
with a traveling shock wave has been also verified by the calculations of Obata
& Hermanson (1998), who used a Reynolds-averaged Navier-Stokesbased
model to calculate the turbulent jet before the interaction and a TVD scheme to
simulate the passage of the shock.

7. PLANE SHEAR LAYER INTERACTIONS WITH SHOCK


WAVES

Interactions of plane shear layers with shock waves have been configured in
supersonic wind tunnels where a boundary layer flow separates over a backward-
facing step formed by a cavity in the tunnel wall. A ramp, located at some distance
downstream, is used to generate an oblique shock. This configuration has been
used in the experiments by Settles et al (1982), Hayakawa et al (1984), and
Samimy et al (1986). The separated shear layer develops over the cavity as a
compressible shear flow under the influence of a recirculating zone beneath it and
subsequently undergoes compression by the shock system and then reattaches to
the ramp. The phenomena appear complicated within the reattachment region,
showing a great degree of unsteadiness, which is characterized by large fluctua-
tions of wall pressure and mass flux in the flow above. The location of the shock
SHOCK-TURBULENCE INTERACTIONS 337

wave is also time dependent, which further complicates our understanding. All
measurements of turbulence reported in the papers mentioned above indicate sub-
stantial amplification of turbulence intensities that depend on the Mach number.
It is not clear, however, how these results are affected by the shock oscillation
and the unsteady phenomena associated with the reattachment region.
Standard Reynolds-averaged Navier-Stokes methods also have been used to
calculate the flow (see Horstman et al 1982, Visbal & Knight 1984) with moderate
success in predicting the mean flow behavior. Turbulence is considerably over-
estimated in these calculations.

8. TURBULENT WAKESHOCK INTERACTIONS

This type of interaction is very often found inside compressors or turbines of


turbomachinery, where a shock may be formed and very often may travel down-
stream. The shock can subsequently interact with wakes of individual blades of
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

stators or rotors and alter their structure. As a result of this interaction, blade
wakes carrying more turbulence impinge on the surface of blades of downstream
stages and can cause additional fatigue of the blades. No previous experimental
or analytical work could be identified on these types of interactions.

9. BOUNDARY LAYERSHOCK WAVE INTERACTIONS

The majority of the previous work in turbulenceshock interactions belongs in


this category, and it mainly includes interactions of boundary layers with oblique
shock waves. Usually, an oblique shock wave is generated either at the opposite
wall to that of the boundary layer by a sharp edge or protrusion (see Rose &
Johnson 1975) or on the same wall of the boundary-layer flow by a compression
turning, which can be a sharp turning as for a compression corner or a gradual
turning by a smooth change in curvature. In the first configuration the incident
shock emanating from the opposite side of the wall is also reflected over the
boundary-layer wall, and therefore the flow interacts with two oblique shocks. In
the second configuration the flow after the interaction is further compressed by
the new boundary imposed by the ramp slope, whereas concave curvature of the
streamlines has a destabilizing effect on turbulence. Thus, turbulence can not only
be affected by the shock interaction but also by the spatiotemporal oscillation of
the shock system, the destabilizing effects of the concave streamline curvature,
and the continuing downstream compression. All these effects contribute to an
increasing level of turbulence. In addition, unsteady flow separation and flow
reversal take place at the corner region, which at large turning angles can affect
the shock system upstream through a feedback mechanism.
One of the major characteristics of the turbulent boundary-layer interaction
with an oblique shock over a compression ramp is the unsteady motion of the
338 ANDREOPOULOS n AGUI n BRIASSULIS

shock system, which has been observed to occur over a significant distance in
the longitudinal and lateral directions. The shock system itself may at times con-
sist of multiple shocks, a K shock, or a single shock. In addition, shocks are
deformed by the interaction and therefore no longer remain planar. Figure 15
shows a sequence of two images obtained by a double-pulse planar laser scattering
technique at the University of Texas at Austin by NT Clemens & DS Dolling
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Figure 15 A hypersonic compression ramp interaction is visualized using planar laser


scattering imaging of a condensed alcohol fog (NT Clemens & DS Dolling, private com-
munication, 1999). The ramp angle is 288 and the field of view is 75 mm in the streamwise
direction by 36 mm in the transverse direction. The upstream boundary layer is turbulent,
and the free stream Mach number and velocity are 4.95 and 765 m/s, respectively. These
two images are of a double-pulse sequence with a time delay between pulses of 30 ls. In
these images the free stream is rendered gray, boundary layer fluid is darker gray, very
low velocity fluid is black (owing to evaporation of the particles), and the separation shock
is the thin white interface, as labeled. Note that these images are somewhat difficult to
interpret owing to the potential for evaporation and condensation of the alcohol seed. For
reference, arrows point to the same structure in both images.
SHOCK-TURBULENCE INTERACTIONS 339

(private communication, 1999). The images are from an M 4 5 interaction over


a 288C compression ramp. They very clearly show a substantial deformation of
the shock front, which changes significantly from the first to the second image
as a result of the mutual interaction with the incoming turbulence. A turbulent
structure going through the system is also visible. The outer part of the eddy
seems to go through the shock without substantial changes. The shock itself,
however, seems to have moved upward, although its extent toward the wall
remains strong.
The first evidence of the grossly unsteady nature of the shock wave system
came from measurements of the time-dependent pressure at the wall beneath the
region of interaction. Kistler (1959) provided one of the first measurements of
pressure fluctuations in a supersonic boundary layer approaching a forward-facing
step. Coe (1969), Dolling & Murphy (1982), Dolling & Or (1983), and Muck et
al (1985) measured the fluctuating wall pressure in compression corners of dif-
ferent geometries. These studies, as well as the studies of Smits & Muck (1987),
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

Muck et al (1988), and Selig et al (1989), clearly demonstrated that the instan-
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

taneous structure of the shock wave system is quite different from that of the
time-averaged system. Two possible mechanisms have been proposed to explain
the shock system unsteadiness: the first one is the turbulence of the incoming
boundary layer (Plotkin 1975, Andreopoulos & Muck 1987), and the second is
that the separated shear layer amplifies the low frequencies of the contracting/
expanding bubble motion, which are felt upstream through the large subsonic
region of the separated zone.
The experimental work of Andreopoulos & Muck (1987) suggested that the
frequency of the shock wave system unsteadiness may scale on the bursting fre-
quency of the incoming boundary layer. This finding supports the notion that the
incoming turbulence plays a dominant role in triggering the shock unsteadiness.
Since this work, several other experiments have provided support for one or the
other of the two proposed mechanisms. Erengil & Dolling (1991) found that some
of the low-frequency oscillations of the shock are caused by the contraction and
expansion cycle of the separated zone. Unalmis & Dolling (1994) also showed
that the large-scale structures of the incoming boundary layer also contribute to
low-frequency oscillations of the shock. The latest work of Beresh et al (1999)
concludes that the small-scale motion of the shock results from the individual
turbulent fluctuations, whereas the large-scale motion is caused by the large-scale
structures of the boundary layer. Thus, this evidence supports the notion that the
small and large scales of the incoming boundary layer play a dominant role in
the shock system unsteadiness.
Spanwise variation of several time-averaged quantities, such as the intermit-
tency factor or correlation coefficient, for instance, in the region of the interaction
points to the supposition that Taylor-Gortler vortices formed by the concave
streamline curvature may exist and contribute to increased complexity of the
phenomena involved (see Ardonceau 1984).
340 ANDREOPOULOS n AGUI n BRIASSULIS

10. FINAL REMARKS


The information that is available now on the interaction of turbulence with shock
waves appears to be limited. The phenomenology associated with the interactions
is not very extensive, and the understanding of their physical attributes is not
very thorough. Experiments or simulations for instance, of shock interactions with
wake flows, do not exist, and experimental data in shock interactions with homog-
enous sheared turbulence are not available for comparison with existing DNS
data. However, the understanding of the physics has been advanced in the last
years owing to the synergy among theory, numerical simulations, and experi-
ments. Despite this progress, a lot of questions remain unanswered, and more
work is needed to illuminate the physics of the interactions. Ribners (1953)
theory, for instance, which predicts turbulent amplification through shock waves,
has, in principle, been verified by DNS and several experiments. There is new
evidence, however, that suggests that under some conditions of the upstream
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

compressible turbulence the interaction may not lead to increased fluctuations of


Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

velocity. Thus, the final word is not yet out. In addition, several of the questions
posed in the introduction remain unanswered.
The major impediment to understanding the physical aspects of the interactions
is the difficulty in simulating the interactions numerically or experimentally. For
instance, it is very challenging to generate normal shock waves or homogenous
flows inside wind tunnels. Similar difficulties are attributed to DNS applied to
flows with homogeneous or periodic boundary conditions and covering a rela-
tively low range of Reynolds numbers. In addition, the lack of availability of
proper experimental or numerical tools with adequate temporal and spatial reso-
lution remains a severe limitation. The new generation of nonintrusive flow visu-
alization and quantitative information techniques holds a lot of promise (see Miles
& Lempert 1997) with the prospect of further increasing our physical understand-
ing, although extensive development and verification of these techniques are still
needed. This review of available information revealed that the interaction between
a shock wave and turbulence is mutual and very complex: Turbulence is affected
by the shock wave, and the shock wave is affected by the turbulence. Compress-
ibility of the incoming flow and the characteristics of turbulence in the incoming
flow, such as velocity and length scales, are some additional parameters that seem
to affect the outcome of the interaction.

ACKNOWLEDGMENTS

The support from NASA Langley Research Center, NASA Lewis Research Cen-
ter, National Science Foundation, U.S. Army TACOM/ARDEC, and Air Force
Office of Scientific Research in carrying out some of the experiments described
in this review and in writing up this article is greatly appreciated. The authors
also thank J Hermanson and B Cetegen for providing the images of Figure 12
SHOCK-TURBULENCE INTERACTIONS 341

and N Clemens and D Dolling for providing the images of Figure 15. Useful
comments provided by D Knight are also appreciated.

Visit the Annual Reviews home page at www.AnnualReviews.org.

LITERATURE CITED
Adamson TC, Messiter AF. 1980. Analysis of Blin E. 1993. Etude experimentale de
two dimensional interactions between linteraction entre turbulence libre et une
shock waves and boundary layers. Annu. onde de choc. These doc. Univ. Paris, Paris
Rev. Fluid Mech. 12:10338 Bradshaw P. 1974. The effect of mean com-
Agui JH. 1998. Shock wave interactions with pression or dilatation on the turbulence
turbulence and vortices. PhD thesis. City structure of supersonic boundary layers. J.
Univ. New York Fluid Mech. 63:44964
Alem D. 1995. Analyse experimentale dune Bradshaw P. 1977. Compressible turbulent
turbulence homogene en ecoulement super- shear layers. Annu. Rev. Fluid Mech. 9:33
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

sonoc soumise a un choc droit. These doc. 54


Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Univ. de Poitiers, Poitiers, Fr. Bradshaw P. 1996. Turbulence in compressible


Andreopoulos J, Muck KC. 1987. Some new flowPhysics and modeling. Presented at
aspects of the shock-wave boundary layer 1st, Annu. Meet. Jpn. Soc. Fluid Mech.,
interaction in compression ramp corner. J. Tokyo
Fluid Mech. 180:40528 Briassulis G, Honkan A, Andreopoulos J, Wat-
Anyiwo JC, Bushnell DM. 1982. Turbulence kins BC. 1995, Applications of hot-wire
amplification in shock-wave boundary anemometry in shock tube flows. Exp. Flu-
layer interaction. AIAA J. 20:89399 ids 19:2937
Ardonceau PL. 1984. The structure of turbu- Briassulis GK. 1996. Unsteady nonlinear
lence in a supersonic shock-wave/boundary interactions of turbulence with shock
layer interaction. AIAA J. 22:125462 waves. PhD thesis. City Coll. City Univ.
Barre S, Allem D, Bonnet JP. 1996. Experi- New York, NY
mental study of normal shock/homoge- Briassulis GK, Andreopoulos J. 1996. High
neous turbulence interaction. AIAA J. resolution measurements of isotropic tur-
34:96874 bulence interacting with shock waves. Pre-
Beresh SJ, Clemens NT, Dolling DS. 1999. sented at AIAA, Aerospace Sci. Meet.,
The relationship between upstream turbu- 34th, Reno, NV, 96-0042
lent boundary layer velocity fluctuations Buckingham AC. 1990, Interactive shock
and separation shock unsteadiness. AIAA structure response to turbulence. AIAA Pap.
Pap. 990295 90-1642
Birch SF, Eggers JM. 1972. A critical review Budwig R, Zwart PJ, Nguyen V, Bushnell DM,
on the experimental data for developed free Johnson CB, et al. 1969. Comparison of
turbulent shear layers. NASA Spec. Publ. prediction methods and studies of relaxa-
SP-321, pp. 1140. Washington, DC tion in hypersonic turbulent nozzle-
Blaisdell GA, Mansour NN, Reynolds WC. wall boundary layers. NASA Tech. Note
1991. Numerical simulation of compressi- D-5433
ble homogeneous turbulence. Rep. TF-50, Budzinski JM, Zukoski EE, Marble FE. 1992.
Thermosci. Div., Mech. Eng., Stanford Rayleigh scattering measurements of shock
Univ., Stanford, CA enhanced mixing. AIAA/SAE/ASME/ASEE,
342 ANDREOPOULOS n AGUI n BRIASSULIS

Jt. Propulsion Conf., 28th, Nashville, TN, Dolling DS, Murphy M. 1982. Wall pressure
92-3546 fluctuations in a supersonic separated com-
Cetegen BM, Hermanson JC. 1995. Mixing pression ramp flow field. AIAA and ASME.
characteristics of compressible vortex rings Joint Thermophys. Fluids, Plasma Heat
interacting with normal shock waves. Com- Transfer Conf., 3rd, St Louis, MO, 82-0986
bust. Flame 100:23240 Dolling DS, Or CT. 1983. Unsteadiness of the
Chang CT. 1957. Interaction of a plane shock shock wave structure in attached and sepa-
and oblique plane disturbances with special rated compression ramp flow fields. Pre-
reference to entropy waves. J. Aeronaut. sented at Am. Inst. Aeronaut. Astron., Fluid
Sci. 24:67582 Plasma Dyn. Conf., 16th, Danners, MA, 83-
Chen JH. 1993. The effect of compressibility 1715
on conserved scalar entrainment in a plane Dosanjh DS, Weeks TM. 1964. Interaction of
free shear layer. In Turbulent Shear Flows a starting vortex as well as Karman vortex
8, ed. F Durst, R Friedrich, BE Launder, streets with traveling shock wave. Presented
FW Schmidt, U Schumann, JH Whitelaw, at Am. Inst. Aeronaut. Astron., Annu.
pp. 297311. Berlin: Springer-Verlag Meet., 1st, Washington, DC, 64-425
Cheng RK, Ng TT. 1985. Conditional Rey- Durbin PA, Zeman O. 1992. Rapid distortion
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

theory for homogeneous compressed tur-


Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

nolds stress in a strongly heated turbulent


boundary layer with premixed combustion. bulence with application to modeling. J.
Phys. Fluids 28:47388 Fluid Mech. 242:34970
Chu BT, Kovasznay LSG. 1958. Non-linear Dussauge JP, Gaviglio J. 1987. The rapid
interactions in a viscous heat-conducting expansion of a supersonic turbulent flow:
compressible flow. J. Fluid Mech. 3:494 role of bulk dilatation. J. Fluid Mech.
174:81112
514
Elena M, Lacharme JP. 1988. Experimental
Coe CF. 1969. Surface-pressure fluctuations
study of a supersonic turbulent boundary
associated with aerodynamic noise. NASA
layer using a laser Doppler anemometer. J.
Spec. Publ. SP 207. Washington, DC
Mech. Theor. Appl. 7:17590
Coleman GN, Mansour NN. 1993. Simulation
Erengil ME, Dolling DS. 1991. Correlation of
and modeling of homogeneous compressi-
separation shock motion with pressure fluc-
ble turbulence under isotropic mean com-
tuations in the incoming boundary layer.
pression. In Turbulent Shear Flows 8, ed. F
AIAA J. 29(11):186877
Durst, R Friedrich, BE Launder, FW Goldstein ME. 1978. Unsteady vortical and
Schmidt, U Schumann, JH Whitelaw, pp. entropic distortions of potential flows round
26982. Berlin: Springer-Verlag. 8th ed. arbitrary obstacles. J. Fluid. Mech. 89:433
Coleman GN, Mansour NN. 1991. Modeling 68
the rapid spherical compression of isotropic Goldstein ME, Durbin PA. 1980. The effect of
turbulence. Phys. Fluids A 3:225559 finite turbulence spatial scale on the ampli-
Debieve JF. 1986. Problemes de distorsion fication of turbulence by a contracting
rapide en ecoulement compressible. Proc. stream. J. Fluid Mech. 98:473508
Colloq. ONERA/DRET Ecoulements Tur- Green JE. 1970. Interaction between shock
bulents Compressibles, Poitiers-Marseille waves and turbulent boundary layers. Prog.
Debieve JF, Lacharme JP. 1985. A shock wave Aeronaut. Sci. 11:235340
free turbulence interactions. Presented at Gutmark EJ, Schadow KC, Yu KH. 1995. Mix-
Int. Union Theor. Appl. Mech. Conf., Paris ing enhancement in supersonic free shear
Dolling DS. 1993. Fluctuating loads in shock flows. Annu. Rev. Fluid Mech. 27:375417
wave/turbulent boundary layer interaction: Hannappel R, Friedrich R. 1995. Direct numer-
tutorial and update. AIAA Pap. 93-0284 ical simulation of a Mach 2 shock interact-
SHOCK-TURBULENCE INTERACTIONS 343

ing with isotropic turbulence. Appl. Sci. action. Proc. Turbulent Shear Flows, 8,
Res. 54:20521 Tech. Univ., Munich, Ger. pp. 1-2-11-2-6
Hayakawa K, Smits AJ, Bogdonoff SM. 1984. Jacquin L, Blin E, Geffroy P. 1993. An exper-
Turbulence measurements in a compressi- imental study on free turbulence/shock
ble reattaching shear layer. AIAA J. 22:889 wave interaction. In Turbulent Shear Flows
95 8, ed. F Durst, R Friedrich, BE Launder,
Hayes W. 1957. The vorticity jump across a FW Schmidt, U Schumann, JH Whitelaw,
gas dynamic discontinuity. J. Fluid Mech. pp. 22948, Berlin: Springer-Verlag
2:595600 Jacquin L, Cambon C, Blin E. 1992. Turbu-
Hermanson JC, Cetegen BM. 1998. Mixing lence amplification by a shock wave and
enhancement of non-uniform density tur- rapid distortion theory. Phys. Fluids A
bulent jets interacting with normal shock 5:253950
waves. Presented at Symp. Combust., Com- Jacquin L, Geffroy P. 1997. Amplification and
bust. Inst., 27th, Univ. Colo., Boulder reduction of turbulence in a heated jet/
Hermening K. 1999. Experimental investiga- shock interaction. Journal L1217. Free
tion of shock wave interaction with jets. Shear Flow Conf.
Diploma thesis. Tech. Univ. Berlin, Berlin, Keller J, Merzkirch W. 1990. Interaction of a
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

Ger.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

normal shock with a compressible turbulent


Hesselink L, Sturtevant B. 1988. Propagation flow. Exp. Fluids 8:24149
of weak shocks through random medium. J. Kerrebrock JL. 1956. The interaction of flow
Fluid Mech. 196:51353 discontinuities with small disturbances in a
Honkan A, Andreopoulos J. 1992. Rapid com- compressible fluid. PhD thesis. Calif. Inst.
pression of grid-generated turbulence by a Technol., Pasadena
moving shock wave. Phys. Fluids A
Kistler AL. 1959. Fluctuation measurements in
4:256272
a supersonic turbulent boundary layer.
Honkan A, Andreopoulos J. 1997. Vorticity
Phys. Fluids 2:29096
strain-rate and dissipation characteristics in
Kovasznay LSG. 1950. The hot-wire anemom-
the near wall region of turbulent boundary
eter in supersonic flow. J. Aeronaut. Sci.
layers. J. Fluid Mech. 350:2996
17:56584
Honkan A, Watkins CB, Andreopoulos J.
Kovasznay LSG. 1953. Turbulence in super-
1994. Experimental study of interactions of
sonic flow. J. Aeronaut. Sci. 20:65774
shock wave with free stream turbulence. J.
Lee L, Lele SK, Moin P. 1993. Direct numer-
Fluids Eng. 116:76369
Hortsman CC, Settles GS, Williams DR, Bog- ical simulation of isotropic turbulence inter-
donoff SM. 1982. A reattaching free shear acting with a weak shock wave. J. Fluid
layer in compressible turbulent flow. AIAA Mech. 251:53362
J. 20:7985 Lee L, Lele SK, Moin P. 1994. Interaction of
Hunt JCR. 1973. A theory of turbulent flow isotropic turbulence with a strong shock
round two-dimensional bluff bodies. J. wave. AIAA Pap. 94-0311. Dept. Mech.
Fluid Mech. 61:62536 Eng., Stanford Univ., CA
Hunt JCR. 1978. A review of the theory of rap- Lee S, Moin P, Lele SK. 1992. Interaction of
idly distorted turbulent flow and its appli- isotropic turbulence with a shock wave.
cations. Fluid Dyn. Trans. 9:12152 Rep. TF-52. Dept. Mech. Eng., Stanford
Hunt JCR, Carruthers DJ. 1990. Rapid distor- Univ., CA
tion theory and the problems of turbu- Lele SK. 1994. Compressibility effects on tur-
lence. J. Fluid Mech. 212:497532 bulence. Annu. Rev. Fluid Mech. 26:21154
Jacquin L, Blin E, Geffroy P. 1991. Experi- Mahesh KA, Lele SK, Moin P. 1997. The influ-
ments of free turbulence/shock wave inter- ence of entropy fluctuations on the inter-
344 ANDREOPOULOS n AGUI n BRIASSULIS

action of turbulence with a shock wave. J. Rose WC, Johnson DA. 1975. Turbulence in a
Fluid Mech. 334:35379 shock-wave boundary-layer interaction.
McKenzie JF, Westphal KO. 1968. Interaction AIAA J. 13:88489
of linear waves with oblique shock waves. Rotman D. 1991. Shock wave effects on a tur-
Phys. Fluids 11:235062 bulent flow. Phys. Fluids A 3:1792806
Miles RB, Lempert WR. 1997. Quantitative Samimy M, Addy AL. 1985. A study of com-
flow visualization in unseeded flows. Annu. pressible turbulent reattaching free shear
Rev. Fluid Mech. 29:285326 layers. Presented at Am. Inst. Aeronaut.
Morkovin MV. 1960. Note on assessment of Astron., Fluid Dyn. Plasmadyn. Laser
flow disturbances at a blunt body traveling Conf., 18th, Cincinnati, OH 85-1646
at supersonic speeds owing to flow distur- Samimy M, Petrie HL, Addy AL. 1986. A
bances in free stream. J. Appl. Mech. study of compressible turbulent reattaching
27:22329 free shear layers. AIAA J. 24:26167
Morkovin MV. 1962. Effects of compressibil- Sekundov AN. 1974. Supersonic flow turbu-
ity on turbulent flows. In Mechanique de la lence and interaction with a shock wave.
Turbulence, ed. A Favre, pp. 36780. Paris: Izv. Akad. Nauk SSR Mekh. Zhidk. Gaza,
CNRS MarchApril, pp. 816
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

Morkovin MV. 1992. Mach number effects on Selig MS, Andreopoulos J, Muck KC, Dus-
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

free and wall turbulent structures in light of sauge JP, Smits AJ. 1989. Turbulence struc-
instability flow interactions. In Studies in ture in shock wave/turbulent boundary
Turbulence, ed. TB Gatski, S Sarkar, CG layer interaction. AIAA J. 27:86269
Speziale, pp. 26984. Berlin: Springer- Settles GS, Dodson LJ. 1991. Hypersonic
Verlag shock/boundary layer interaction database.
Muck KC, Andreopoulos J, Dussuage JP. 1988. NASA Conf. Rep. CR 177577. Penn. State
Unsteady nature of shock wave/turbulent Univ., Univ. Park, PA
boundary layer interaction. AIAA J. Settles GS, Dolling DS. 1992. Swept shock
26:17997 wave boundary layer interaction. AIAA
Muck KC, Dussuage JP, Bogdonoff SM. 1985. Prog. Ser. Vol. Tactical Missile Aerodyn.
Structure of the wall pressure fluctuations 141:50574
in a shock-induced separated turbulent Settles GS, Fitzpatrick TJ, Bogdonoff SM.
flow. AIAA Pap. 85-0179 1979. Detailed study of attached and sepa-
Obata S, Hermanson JC. 1998. Shear layer rated compression corner flow fields in high
mixing enhancement by shock passing. Pre- Reynolds numbers supersonic flow. AIAA J.
sented at 36th Aircraft Symp., Jpn. Soc. 17:57985
Aeronaut. Space Sci., Yokosuka Settles GS, Williams DR, Baca BK, Bogdonoff
Plotkin KJ. 1975. Shock wave oscillation drive SM. 1982. Reattachment of a compressible
by turbulent boundary layer fluctuations. turbulent free shear layer. AIAA J. 20:60
AIAA J. 13:103640 67
Ribner HS. 1953. Convection of a pattern of Smits AJ, Dussuage JP. 1996. Turbulent Shear
vorticity through a shock wave. NACA Rep. Layers in Supersonic Flow. New York: AIP
2864. Natl. Advis. Comm. Aeronaut., Press. 358 pp.
Cleveland, OH Smits AJ, Muck KC. 1987. Experimental study
Ribner HS. 1969. Acoustic energy flux froze of three shock wave/boundary layer inter-
shock-turbulence interaction. J. Fluid actions. J. Fluid Mech. 182:291314
Mech. 35:299310 Spina EF, Smits AJ, Robinson SK. 1994. The
Ribner HS. 1986. Spectra of noise and ampli- physics of supersonic turbulent boundary
fied turbulence emanating from shock- layers. Annu. Rev. Fluid Mech. 26:287
turbulence interaction. AIAA J. 25:43642 319
SHOCK-TURBULENCE INTERACTIONS 345

Trolier JW, Duffy RE. 1985. Turbulent mea- Zang TA, Hussaini MY, Bushnell DM. 1982.
surements in shock-induced flows. Proc. Numerical computations of turbulence
Am. Inst. Aeronaut. Astron., Fluid Dyn. amplification in shock wave interactions.
Conf., 25th, Colorado Springs, CO AIAA Pap. 82-0293. Am. Inst. Aeronaut.
23:117278 Astron., Aerospace Sci. Meet., 20th,
Unalmis OH, Dolling DS. 1994. Decay of wall Orlando, FL
pressure field and structure of a mach 5 adi- Zeman O. 1990. Dilatation dissipation: the
abatic turbulent boundary layer. AIAA Pap. concept and application in modeling com-
94-2363 pressible mixing layers. Phys. Fluid A
Visbal M, Knight D. 1984. The Baldwin- 2:17888
Lomax turbulence model for two- Zwart PJ, Budwig R, Tavoularis S. 1997. Grid
dimensional shock-wave/boundary layer turbulence in compressible flow. Exp. Flu-
interactions. AIAA J. 22:92128 ids 23:52022
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org
Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org
by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.
Annual Review of Fluid Mechanics
Volume 32, 2000

CONTENTS
Scale-Invariance and Turbulence Models for Large-Eddy Simulation,
Charles Meneveau, Joseph Katz 1

Hydrodynamics of Fishlike Swimming, M. S. Triantafyllou, G. S.


Triantafyllou, D. K. P. Yue 33

Mixing and Segregation of Granular Materials, J. M. Ottino, D. V.


Khakhar 55

Fluid Mechanics in the Driven Cavity, P. N. Shankar, M. D. Deshpande 93


by Stanford University - Main Campus - Green Library on 06/10/12. For personal use only.

Active Control of Sound, N. Peake, D. G. Crighton 137


Annu. Rev. Fluid Mech. 2000.32:309-345. Downloaded from www.annualreviews.org

Laboratory Studies of Orographic Effects in Rotating and Stratified


Flows, Don L. Boyer, Peter A. Davies 165
Passive Scalars in Turbulent Flows, Z. Warhaft 203

Capillary Effects on Surface Waves, Marc Perlin, William W. Schultz 241


Liquid Jet Instability and Atomization in a Coaxial Gas Stream, J. C.
Lasheras, E. J. Hopfinger 275
Shock Wave and Turbulence Interactions, Yiannis Andreopoulos, Juan H.
Agui, George Briassulis 309

Flows in Stenotic Vessels, S. A. Berger, L-D. Jou 347

Homogeneous Dynamos in Planetary Cores and in the Laboratory, F. H.


Busse 383

Magnetohydrodynamics in Rapidly Rotating spherical Systems, Keke


Zhang, Gerald Schubert 409

Sonoluminescence: How Bubbles Turn Sound into Light, S. J. Putterman,


K. R. Weninger 445

The Dynamics of Lava Flows, R. W. Griffiths 477

Turbulence in Plant Canopies, John Finnigan 519

Vapor Explosions, Georges Berthoud 573

Fluid Motions in the Presence of Strong Stable Stratification, James J.


Riley, Marie-Pascale Lelong 613

The Motion of High-Reynolds-Number Bubbles in Inhomogeneous Flows,


J. Magnaudet, I. Eames 659

Recent Developments in Rayleigh-Benard Convection, Eberhard


Bodenschatz, Werner Pesch, Guenter Ahlers 709

Flows Induced by Temperature Fields in a Rarefied Gas and their Ghost


Effect on the Behavior of a Gas in the Continuum Limit, Yoshio Sone 779

Das könnte Ihnen auch gefallen