ArticlePDF Available

Shock-Wave Mach-Reflection Slip-Stream Instability: A Secondary Small-Scale Turbulent Mixing Phenomenon

Authors:

Abstract and Figures

Theoretical and experimental research, on the previously unresolved instability occurring along the slip stream of a shock-wave Mach reflection, is presented. Growth rates of the large-scale Kelvin-Helmholtz shear flow instability are used to model the evolution of the slip-stream instability in ideal gas, thus indicating secondary small-scale growth of the Kelvin-Helmholtz instability as the cause for the slip-stream thickening. The model is validated through experiments measuring the instability growth rates for a range of Mach numbers and reflection wedge angles. Good agreement is found for Reynolds numbers of Re 2 x 10(4). This work demonstrates, for the first time, the use of large-scale models of the Kelvin-Helmholtz instability in modeling secondary turbulent mixing in hydrodynamic flows, a methodology which could be further implemented in many important secondary mixing processes.
Content may be subject to copyright.
Shock-Wave Mach-Reflection Slip-Stream Instability:
A Secondary Small-Scale Turbulent Mixing Phenomenon
A. Rikanati
Department of Physics, Ben Gurion University, Beer-Sheva 84015, Israel
O. Sadot
Department of Mechanical Engineering, Ben Gurion University, Beer-Sheva 84015, Israel,
and Department of Physics, Nuclear Research Center, Negev 84190, Israel
G. Ben-Dor
Department of Mechanical Engineering, Ben Gurion University, Beer-Sheva 84015, Israel
D. Shvarts
Department of Physics, Ben Gurion University, Beer-Sheva 84015, Israel,
Department of Mechanical Engineering, Ben Gurion University, Beer-Sheva 84015, Israel,
and Department of Physics, Nuclear Research Center, Negev 84190, Israel
T. Kuribayashi and K. Takayama
Interdisciplinary Shock-Wave Research Center, Institute of Fluid Science, Tohoku University, Sendai, Japan
(Received 15 December 2005; published 3 May 2006)
Theoretical and experimental research, on the previously unresolved instability occurring along the slip
stream of a shock-wave Mach reflection, is presented. Growth rates of the large-scale Kelvin-Helmholtz
shear flow instability are used to model the evolution of the slip-stream instability in ideal gas, thus
indicating secondary small-scale growth of the Kelvin-Helmholtz instability as the cause for the slip-
stream thickening. The model is validated through experiments measuring the instability growth rates for
a range of Mach numbers and reflection wedge angles. Good agreement is found for Reynolds numbers of
Re > 2 10
4
. This work demonstrates, for the first time, the use of large-scale models of the Kelvin-
Helmholtz instability in modeling secondary turbulent mixing in hydrodynamic flows, a methodology
which could be further implemented in many important secondary mixing processes.
DOI: 10.1103/PhysRevLett.96.174503 PACS numbers: 47.20.Lz, 47.20.Ft, 47.40.Nm, 47.63.mc
Understanding secondary turbulent mixing in complex
unstable hydrodynamic flows is of great importance in
achieving gain in laser driven inertial confinement fusion
(ICF) as well as in many astrophysical processes [1–3].
Mach reflections (MRs) are a well known shock-wave
related phenomenon occurring when an oblique shock
wave reflects from a rigid wall (or interacts with a second
shock wave) and are of great importance in many hydro-
dynamic flows [4]. The basic feature of a Mach reflection is
that of the three shocks structure, giving rise to the slip-
stream (SS) instability. In this Letter, we show, for the first
time, that the growth of the SS instability is due to sec-
ondary turbulent mixing. Through modeling this instabil-
ity, as described further on in this Letter, we implement a
new technique for understanding secondary turbulent mix-
ing, a technique which could be further implemented for
many other hydrodynamic flows.
The MR three shocks structure is a phenomenon widely
demonstrated in many experimental works (see [4] for
examples), as well as in the example from the current
work illustrated in Fig. 1 (see figure caption for details).
The three shocks structure appears when the inclination
angle, between the incident shock and the bounding wall, is
smaller than a critical angle defined through the detach-
ment criteria [5], depending on the shock-wave Mach
number and the material equation of state (see [4] and
references therein for further details). In the MR structure,
the SS is of a unique hydrodynamic nature, being a dis-
continuity which is not a shock. The SS separates between
two regions (regions 2 and 3 in Fig. 1) of different densities
(
2
3
) and tangential velocities (v
k
2
v
k
3
) but of equal
pressures (p
2
p
3
) and of zero perpendicular velocities
(v
?
2
v
?
3
0), i.e., a shear flow. As can be seen from
Fig. 1, and in contrast with the sharp nature of the three
shock waves, the SS discontinuity thickness increases
downstream from the triple point. To our knowledge, the
full nature of this experimentally observed phenomenon,
known as the SS instability, is still unresolved, with the
leading assumption for the cause for this instability relating
to the viscosity generated boundary layer effect [4].
As is the case in the SS discontinuity, the Kelvin-
Helmholtz (KH) shear flow instability [6] occurs when
two fluids flow with proximity to each other with a tangen-
tial velocity difference, defined as the shear velocity. Under
this instability, small perturbations on the two fluid inter-
face evolve into a formation of vortices causing the two
PRL 96, 174503 (2006)
PHYSICAL REVIEW LETTERS
week ending
5 MAY 2006
0031-9007=06=96(17)=174503(4) 174503-1 © 2006 The American Physical Society
fluids to turbulently mix. In this Letter, we demonstrate
that the cause for the SS thickening is the KH instability
evolving on the two fluid interface, generating secondary
small-scale turbulent mixing.
The KH instability large-scale behavior has been thor-
oughly investigated through a wide range of experimental,
numerical, and theoretical work resulting in an understand-
ing of its growth rates and characteristics and of the main
mechanisms dominating its evolution (see, for example,
[6,7]). By implementing previously reported KH large-
scale instability growth rates [6,7], we try to model the
spread angle of the SS instability as a function of the MR
flow parameters. In the following paragraphs, a brief de-
scription of the growth rates of the large-scale KH insta-
bility will be presented, followed by a detailed modeling of
the SS instability evolution. The model predictions are then
verified through comprehensive experimental research.
Finally, a short description of viscous effects and the
relation to the Re number of the flow is presented, again
supported by experimental results.
Apart from revealing the nature of this previously un-
resolved phenomenon (the SS instability), the success of
the process described in this Letter demonstrates, for the
first time, that large-scale models of the KH instability can
be implemented in describing secondary turbulent mixing
in hydrodynamic flows. Similar methodologies could be
further implemented to many other phenomena involving
secondary mixing, such as the ICF relevant Rayleigh-
Taylor and Richtmyer-Meshkov instabilities [2,3]. It
should be mentioned that an attempt to numerically de-
scribe secondary mixing phenomena through the solution
of full 2D or 3D Euler equations is expected to be very
difficult, if not impossible. That is due to the very large
number of computational cells required to describe both
the large-scale and the small-scale mixing processes.
As expected from dimensional considerations, the width
of the KH large-scale turbulent mixing zone (TMZ)
evolves with time according to htcvt, where v
is the shear velocity, t represents the evolution time, and
c 0:19 0:01 is a dimensionless constant derived ex-
perimentally [6], numerically, and recently even theoreti-
cally [7]. It should be mentioned that, in most experiments,
the instability growth is measured spatially, i.e., as a func-
tion of the advection distance from the mixing starting
point. The spatial mixing growth rate is easily related to
the temporal growth by assuming the average flow flows
downstream with the fluid average velocity, resulting in
hx2cv
1
v
2
=v
1
v
2
x, where v
1
and v
2
are
the two fluid velocities and c is the previously mentioned
dimensionless constant. To this equation, two corrections
must be introduced. The first, as reported in Ref. [7], is for
cases of fluids of two different densities, and the second is
for high Mach number flows (
v
a
> 1, where a is the
fluid average sound speed). Following these two correc-
tions, the TMZ width hx becomes
hx0:38 0:02
Sv
1
;v
2
1 2f
d
1
=
2
Sv
1
;v
2
xf
HiMach
v
a
;
f
d
1
;
2

1
2
1

2
=
1
p
1

2
=
1
p
;
(1)
where Sv
1
;v
2
v
1
v
2
=v
1
v
2
. f
HiMach
v
a

0:51 tanh2
v
a
1:2 is the high Mach correction
which is based on a parametric fit of the results shown in
Ref. [8]. Finally, based on Eq. (1), the spread angle of the
spatially growing instability is found according to
spread
arctan
hx
2x
: (2)
When trying to implement Eq. (2) for the growth of the SS
instability, one needs to describe v
1
, v
2
,
1
, and
2
of the
two fluids along the SS. These physical properties are the
velocities (in the frame of reference moving with the triple
point) and the densities of regions 2 and 3 in Fig. 1. All are
analytically calculated using the three-shock theory first
suggested by von Neumann [5]. The theory is based on the
traditional shock-wave conservation equations which are
implemented for a single oblique shock. For an ideal gas
(see [9]), the flow parameters behind an oblique shock can
be analytically derived through a set of analytical trans-
lation functions, as a function of the shock inclination
angle and the preshock conditions. By implementing these
translation functions 3 times, once for each shock in the
FIG. 1. A holographic interferometry image of a M 1:9
Mach reflection in air with a wedge angle of
w
30
(see
text for further experimental details). Marked are the following
features: the incident shock (IS), the Mach stem (MS), the
reflected shock (RS), the triple point (TP), the slip stream
(SS), and the triple-point trajectory (TPT) which is also indicated
by a dashed line. Four flow regions are distinguished: the non-
shocked air (0), the shocked air above the reflection (1), the
shocked air after the reflection (2), and the Mach-stem shocked
air, before the SS (3).
PRL 96, 174503 (2006)
PHYSICAL REVIEW LETTERS
week ending
5 MAY 2006
174503-2
three-shock complex, and by demanding the closure rela-
tions of pressure equalization between regions 2 and 3,
P
2
P
3
, and zero perpendicular velocity, v
?
2
v
?
3
0,
the flow parameters can be solved in the entire domain. By
additionally assuming that the Mach stem is perpendicular
to the wedge, one can also find the angle of the triple-point
trajectory. Once the flow parameters are known, the four
physical parameters mentioned above are easily obtained.
It can be shown from the translation function mentioned
earlier that the resulting spread angle depends only on the
incident shock-wave Mach number, the reflecting wedge
angle, and the adiabatic index of the gas , while the initial
density and pressure are of no importance. In Fig. 2, one
can see the resulting SS instability spread angles for MR in
air ( 1:4) as a function of the incident shock-wave
Mach number for reflecting wedge angles of
w
20
; 30
; 40
; 45
.
Notice that, initially, the expected spread angle increases
with the increase in Mach number or wedge angle, until a
value of about 8
, after which the spread angle decreases.
This effect is a result of the Mach number reduction factor
of Eq. (1).
In order to verify the model predictions, complementary
experimental research was conducted using a shock-tube
facility at the Interdisciplinary Shock-Wave Research Cen-
ter of the Institute of Fluid Science, Tohoku University.
The tube allows the generation of Mach 1.1–5 shock waves
passing through a rectangular tube with a 10 cm by 18 cm
cross section. Near the end of the tube, a windowed test
section allows the user to implement optical diagnostics. A
steel wedge with a varying angle was placed in the test
section, and holographic interferometry images were taken
of the MR generated from the shock-wedge interaction.
For further details on the shock tube and the interferometry
technique, see [10]. Experiments were done with ambient
air at an initial pressure of P
0
10:1 kPa, wedge angles of
w
20
; 30
; 40
; 45
, and incident shock-wave Mach
numbers of M
i
1:55; 1:9; 2:3; 2:78. One can see two
examples: in Fig. 1 for
w
30
, M
i
1:9 and in Fig. 3
for a constant wedge angle of
w
40
and for all four M
numbers.
It is evident that, as the gas on two sides of the SS flows
away from the triple point, the SS thickness linearly in-
creases. A density profile is evident through the increasing
spacing between the fringes along the SS. Notice that, as
predicted by the model, the spread angle increases with
increasing Mach number. In order to quantitatively analyze
the model predictions, a detailed comparison of the mea-
sured and predicted spread angles was done for all Mach
numbers. Spread angle measurements were taken between
two straight lines bounding the SS thickening region, as
demonstrated in Fig. 3(b). Error estimates were set accord-
ing to half the thickness of a single fringe. The results are
plotted in Fig. 2 as a function of the incident shock-wave
Mach number for reflecting wedge angles of
w
20
; 30
; 40
; 45
. Additional experiments conducted at a
higher initial pressure of 100 kPa and at an incident Mach
number of M 1:5 are also plotted.
1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
0
1
2
3
4
5
6
7
8
9
10
Mach number
spread angle [deg]
FIG. 2. SS instability spread angle as a function of incident
shock-wave Mach number. Theoretical predictions in thick lines:
w
45
(solid line),
w
40
(dashed-dotted line),
w
30
(upper dashed line), and
w
20
(lower dashed
line). Thin lines around the thick lines mark the error in the
model prediction [see Eq. (1)]. Experimental results and error
bars are also plotted; (), (), (4), and (5) mark experiments
conducted with
w
45
,40
,30
, and 20
, respectively. All
of the experiments were conducted with ambient air at P
0
10:1 kPa, apart from the M 1:5 experiments conducted at
P
0
100 kPa.
FIG. 3. Holographic interferometry images for MRs with a
wedge angle of
w
40
and incident shock-wave Mach num-
bers of M
i
1:55 (a), 1.9 (b), 2.3 (c), and 2.78 (d). Two white
lines in (b) bound the SS, demonstrating the growth angle
measurement technique.
PRL 96, 174503 (2006)
PHYSICAL REVIEW LETTERS
week ending
5 MAY 2006
174503-3
In the figure, good agreement is demonstrated, apart
from the following experiments at P
0
10:1 kPa: all the
Mach 1.55 experiments, the
w
30
experiments at M<
2, and the
w
20
experiments at M<2:4. In order to
understand the cause for the above disagreements, the
discrepancy between the model predictions and the experi-
mental results is plotted in Fig. 4 against the SS shear
velocity Re number. The Re number is calculated accord-
ing to Re
vl
, where v is the shear velocity, is the
average ideal gas kinematic viscosity calculated at re-
gions 2 and 3 according to
air
181:92 0:536T=
[11], and l stands for a typical length scale of 1 cm (chosen
according to the typical size of the Mach stem).
From the figure, it is clear that all of the experiments
conducted at Re > 2 10
4
show good agreement with
theory. Special attention should be pointed out to experi-
ments conducted at M 1:55 with P
0
10 kPa where no
spread angle could be measured, whereas similar M 1:5
experiments but with P
0
100 kPa show very good
agreement. For the
w
45
experiments, for example,
the resulting physical properties of regions 2 and 3 are
2
0:4kg=cm
3
,
3
0:36 kg=cm
3
, v
2
412 m= sec ,
v
3
257 m= sec , and a 460 m= sec for the P
0
10 kPa experiment and
2
3:8kg=cm
3
,
3
3:5kg=cm
3
, v
2
402 m= sec , v
3
257 m= sec , and
a 454 m= sec for the P
0
100 kPa experiment, both
leading to a very similar spread angle according to
Eq. (1). The important difference between the two classes
of experiments are the higher densities obtained at the
P
0
100 kPa experiments, introducing weaker kinematic
viscosities and larger Re numbers (see caption of Fig. 4).
Together with Fig. 4, it is clear that the disagreements
between the model and the experiments are due to viscosity
effects, which are not accounted for in Eq. (1). Viscosity
effects, serving as a secondary mix cutoff stabilizing
mechanism (through a cutoff Re number), are a well
known phenomenon. The critical Re number, measured
at 2 10
4
, is similar to critical Re numbers demonstrated
for several other mixing phenomena, such as secondary
mixing at the Richtmyer-Meshkov and KH instabilities
[2,3].
Summary.—It is shown that, from the properties of a
large-scale hydrodynamic phenomena, i.e., densities, vis-
cosities, velocities, and average sound speed, one can
predict the existence and width of a secondary small-scale
turbulent mixing zone through a simple analytical proce-
dure based on models of large-scale KH instability growth.
Using this technique, the evolution of the MR SS insta-
bility was analytically characterized. Good agreement is
achieved with an experimental evaluation of the SS spread
angle, conducted at Mach numbers of M
i
1:52:78 and
spread angles of
w
20
45
. A critical Re number of
2 10
4
is found to distinguish between turbulent and
laminar flows, thus indicating secondary turbulent mixing
generated through the evolution of the KH instability as the
cause for the SS discontinuity thickening. The success in
modeling secondary mixing using large-scale models of
the KH instability should provide a guideline for future
research in secondary turbulent mixing phenomena.
The financial support of the Interdisciplinary Shock-
Wave Research Center of the Institute of Fluid Science,
Tohoku University, for conducting the experimental phase
of the presented research and helpful discussions with
D. Oron are acknowledged by the authors.
[1] S. P. Regan et al., Phys. Rev. Lett. 89, 085003 (2002).
[2] Y. Zhou, H. F. Robey, and A. C. Buckingham, Phys. Rev. E
67, 056305 (2003).
[3] P. E. Dimotakis, J. Fluid Mech. 409, 69 (2000).
[4] G. Ben-Dor, Shock Wave Reflection Phenomena (Springer-
Verlag, Berlin, 1992).
[5] J. von Neumann, Navy Department, Bureau of Ordinance,
Washington, DC, Explosives Research Report No. 12,
1943.
[6] G. L. Brown and A. Roshko, J. Fluid Mech. 64, 775
(1974).
[7] A. Rikanati, U. Alon, and D. Shvarts, Phys. Fluids 15,
3776 (2003).
[8] S. A. Ragab and J. L. Wu, Phys. Fluids A 1, 957 (1989).
[9] H. Li and G. Ben-Dor, J. Fluid Mech. 341, 101 (1997).
[10] F. Ohtomo, K. Ohtani, and K. Takayama, Shock Waves 14,
379 (2005).
[11] American Institute of Physics Handbook, edited by
Dwight E. Gray (McGraw-Hill, New York, 1972).
0 2 4 6 8 10 12 14 16
x 10
4
−6
−5
−4
−3
−2
−1
0
1
Reynolds Number
θ
measured
θ
predicted
[deg]
FIG. 4. Predicted spread angle subtracted from the measured
spread angle, as a function of the SS Re number. (), (), (4),
and (5) mark experiments conducted at
w
45
,40
,30
, and
20
, respectively. Experiments conducted at an initial pressure of
100 kPa resulted in Re numbers of Re 15:3 10
4
, 9:2 10
4
,
and 4:8 10
4
for
w
40
,30
, and 20
, respectively.
PRL 96, 174503 (2006)
PHYSICAL REVIEW LETTERS
week ending
5 MAY 2006
174503-4
... Rikanati et al. [36,37] measured shear layer growth rates in shock reflection experiments and found that these growth rates agreed with previous measurements for turbulent shear layers only above a Reynolds number of Re 2×10 4 . They associated this critical Reynolds number with the transition to turbulence in Kelvin-Helmholtz instabilities, where the instabilities begin to of Re ≈ 100 at t = 17, when ignition is expected behind the Mach stem, and Re ≈ 800 at t = 120 for ignition behind the transverse shock. ...
... This is two orders of magnitude smaller than the critical Reynolds number necessary for Kelvin-Helmholtz instabilities to contribute to contact surface thickness. This estimation is extended to stoichiometric detonations under atmospheric conditions in table 2, which lists the Reynolds numbers when ignition occurs behind the Mach stem, when Kelvin-Helmholtz instabilities become important [36,37], and at the maximal cell width. The ignition Reynolds numbers were calculated for prereflection incident wave strengths of 70% [8] of the ideal one-dimensional steady state (Chapman-Jouguet) velocity at the end of the detonation cell, using the Shock & Detonation Toolbox [38] for Cantera [34] with the GRI 3.0 mechanism [35]. ...
Preprint
The reflection of a triple-shock configuration was studied numerically in two dimensions using the Navier-Stokes equations. The flow field was initialized using three shock theory, and the reflection of the triple point on a plane of symmetry was studied. The conditions simulated a stoichiometric methane-oxygen detonation cell at low pressure on time scales preceding ignition, when the gas was assumed to be inert. Viscosity was found to play an important role on some shock reflection mechanisms believed to accelerate reaction rates in detonations when time scales are small. A small wall jet was present in the double Mach reflection and increased in size with Reynolds number, eventually forming a small vortex. Kelvin-Helmholtz instabilities were absent and there was no Mach stem bifurcation at Reynolds numbers corresponding to when the Mach stem had travelled distances on the scale of the induction length. Kelvin-Helmholtz instabilities are found to not likely be a source of rapid reactions in detonations at time scales commensurate with the ignition delay behind the Mach stem.
... The perturbation of the shear layer preceding the embedded shock led to CRVR formation. This process involved the slipstream originating from the shock-wave Mach reflection's triple point, its thickening due to small-scale Kelvin-Helmholtz vortices, aspects reported in prior studies [91,92]. ...
Article
Full-text available
Vortex loops are compact toroidal structures wherein fluid rotation forms a closed loop around a fictitious axis, manifest in many natural occurrences. These phenomena result from brief impulses through vents or apertures in fluid systems, such as in caves, volcanic crusts, downbursts, or the descent of liquid droplets. The majority of naturally occurring and laboratory-generated vortex loops, studied for fundamental research on their formation, growth, instability, and dissolution, are classified as incompressible. This categorisation denotes negligible alterations in thermodynamic properties within the vortex loop. However, a distinct category of vortex loops emerges from processes involving artillery, shock tubes, explosions, chemical interactions, and combustion. This class primarily constitutes compressible vortex loops. Their presence in flow fields spans over a century, and they have been observed since the application of open-ended shock tubes to explore phenomena like diffracting shock waves, blast wave interactions with objects, and noise mitigation. The study and comprehension of compressible vortex loops and their interactions have historically relied heavily on optical techniques, lacking comprehensive insight into the intricate flow dynamics. Nevertheless, the advancements in flow visualisation tools and computational capabilities in the 21st century have significantly aided scientists in scrutinising and characterising these vortex loops and their interactions in intricate detail. Unfortunately, a comprehensive review of the literature addressing compressible vortex loops originating from shock tubes, their evolution, and interactions with shockwaves and various objects, including walls, appears lacking. This review article aims to address this gap.
... First, the initial shock deposits ω 0 at the initial position due to the velocity gradient across the shock [see Eq. (11)]. Second, the slipstream downstream to the shock is identified by ω SS , as a shear layer spread out from the triple point [29]. ...
Article
Full-text available
We report the mechanism and modeling for the formation of cavitylike structures on a planar interface subjected to a perturbed shock wave. The cavity is distinguished from bubbles and spikes formed in the standard Richtmyer-Meshkov instability (RMI). The two-dimensional direct numerical simulation is conducted at a range of shock Mach numbers and Atwood numbers. We elucidate the effects of the interfacial vorticity and the shock-induced vorticity on the cavity formation. The interfacial vorticity, which is important in the standard RMI, only has a minor influence on the cavity width in the linear stage. Alternatively, the cavity width is determined by the Mach-stem height when the shock accelerates the interface. A pair of vorticity patches connecting the Mach stem, as a part of the shock-induced vorticity, penetrate the interface to form the cavity via strong shear layers generated by slipstreams during shock propagation. Inspired by this mechanism, we develop a model of the Mach-stem height to estimate the cavity width in the linear stage at various Mach numbers.
... Since the jet is under-expanded, strong expansion waves extend from the lips of the inlet boundary and interact with the mixing layer. This mixing layer is unstable because of the huge velocity difference inside and outside it, which belongs to the Kelvin-Helmholtz instability [50]. The interaction between the mixing layer and expansion waves enlarges this instability and rolls the mixing layer into a series of counter-rotating vortices ( = 0.24, 0.30 ms). ...
Article
Full-text available
Numerical simulations of high-speed compressible flows remain challenging in engineering as the appearance of shock waves poses difficulties for high-resolution schemes as well as explicit large eddy simulation methods. Therefore, in this work, we propose a practical numerical solver for simulations of shock-vortex and shock-turbulence problems with the implicit large eddy simulation approach and newly developed low-dissipative schemes. In order to handle the multi-scale raised in the shock-vortex and shock-turbulence problems, it is required that the flow solver can simultaneously solve the large-scale flow structures such as shock waves with numerical oscillation-free, the resolvable flow structures above the grid scale with low-dissipation, and the under-resolve isotropic sub-grid scale with numerical stability. To this end, a low-dissipative, structure-preserving scheme with rigorously adjusted numerical dissipation is employed in the proposed solver. The structure-preserving property of this scheme can ensure that the large-scale structure is solved without numerical oscillations and the low-dissipative property of this scheme can produce high-resolution results for the resolvable flow scales. Moreover, the sub-grid scale is solved and stabilized by the inherent numerical dissipation in the shock-capturing scheme. The proposed numerical solver is then applied to simulate a wide range of shock-vortex and shock-turbulence interaction problems including supersonic planar jets, transonic flows past a deep cavity and impingement of a supersonic jet on a cone mounted on a flat plate. A comparison is also made with the solver using conventional total variation diminishing schemes. The numerical results of the supersonic planar jet have demonstrated that the proposed solver can resolve the small-scale structure such as Kelvin–Helmholtz instability involved in shock and shear layer with high-resolution. Through the results of transonic flow past a deep cavity and comparisons with the experimental data, it is verified that the proposed solver can reproduce the turbulence statistical data. Furthermore, the proposed solver resolves the complex turbulent fluid mechanical phenomena in the impingement of a supersonic jet on a cone, which demonstrates the proposed solver can robustly handle irregular geometry. The proposed solver also enjoys simplicity without using the explicit sub-grid scale model and involving the complexity of very high-order schemes. Thus, this work provides an accurate and practical numerical solver for shock-vortex and shock-turbulence problems in high-speed flows.
... They also revealed that multiple CRVRs were formed due to strong impulse. The growth rates of the slipstream observed in their study followed the model given by Rikanati et al. 17 Qin et al. 4 performed PIV experiments to investigate the PVR formed in the shock-tube generated impulsive jets at five different shock Mach numbers, that is, M s ¼ 1.28, 1.39, 1.48, 1.56, and 1.59, respectively. The corresponding Reynolds number was Re ¼ 7:52 Â10 5 ; 10:65 Â 10 5 ; 13:42 Â 10 5 ; 13:70 Â 10 5 , and 13:98 Â 10 5 , respectively. ...
... Studies of shock waves have promoted developments in many scientific fields, such as nuclear physics, 1 plasmas, 2 and aerodynamics. 3,4 Shock reflection phenomena, as a remarkable property of shock waves, have extensively been investigated because of their important physical mechanisms. [5][6][7][8][9][10][11][12][13][14][15][16][17] After first observation of shock reflection 5 and classic shock reflection theory, 6 studies of shock reflection have made breakthroughs of various areas, such as the transition of shock reflection, 7-9 unsteady shock reflection, [10][11][12] and asymmetric shock reflection. ...
Article
Mach reflection (MR) is an essential component in the development of the shock theory, as the incident shock curvature is found to have a significant effect on the MR patterns. Curved-shock Mach reflection (CMR) is not yet adequately understood due to the rotational complexity behind curved shocks. Here, CMR in steady, planar/axisymmetric flows is analyzed to supplement the well-studied phenomena caused by oblique-shock Mach reflection (OMR). The solution from the von Neumann's three-shock theory does not fully describe the CMR case. A CMR structure is presented and characterized by an incident shock, reflected shock, Mach stem, and expansion/compression waves over the slipline or occasionally an absence of waves due to pressure equilibrium. On the basis of this CMR structure, an analytical model for predicting the Mach stem in the CMR case is established. The model reduces to the OMR case if the shock curvature is not applicable. Predictions of the Mach stem geometry and shock structure based on the model exhibit better agreement with the numerical results than predictions using previous models. It is found that the circumferential shock curvature plays a key role in the axisymmetric doubly curved CMR case, which results in a different outcome from the planar case.
... The Knudsen number can be finite in many practical applications, and rarefaction effects become important. For example, rarefaction effects play a critical role in the free shear (mixing) layers of slipstream behind Mach stems [1,2], exhaust plumes of satellite nozzles, [3][4][5] and other space applications. The extent of rarefaction effects can vary with the type of flow and the Mach and Knudsen numbers. ...
Article
Gas-kinetic simulations of rarefied and compressible mixing layers are performed to characterize continuum breakdown and the effect on the Kelvin-Helmholtz instability. The unified gas-kinetic scheme (UGKS) is used to perform the simulations at different Mach and Knudsen numbers. The UGKS stress tensor and heat-flux vector fields are compared against those given by the Navier-Stokes-Fourier constitutive equations. The most significant difference is seen in the shear stress and transverse heat flux. The study demonstrates the existence of two distinct continuum breakdown regimes, one at low and the other at high convective Mach numbers. Overall, at low convective Mach numbers, the deviation from continuum stress and heat flux appears to scale exclusively with the micro-macro length scale ratio given by the Knudsen number. On the other hand, at high convective Mach numbers, the deviation depends on the global micro-macro timescale ratio given by the product of Mach and Knudsen numbers. We further demonstrate that, unlike shear stresses and transverse heat flux, the deviations in normal stresses and the streamwise heat flux depend separately on Knudsen and Mach numbers. A local parameter called the gradient Knudsen number is proposed to characterize the rarefaction effects on the local momentum and thermal transport. Noncontinuum aspects of gas-kinetic stress-tensor and heat-flux behavior that Grad's 13-moment equation model reasonably captures are identified.
... The vortices become large when they reach the end wall. The evolution of K-H vortices formed from the triple point is completely different from the K-H vortices formed in a steady [31][32][33] and transient jet [26,34]. ...
Article
Over the last couple of decades, the shock-wave boundary-layer interaction has gained a lot of attention due to its practical importance in many engineering applications. It is a complex problem involving shock-wave bifurcation, boundary-layer separation, the interaction of contact discontinuity with the shock wave, the formation of shocklets, and the evolution of vortical structures having different length scales. It is difficult to capture experimentally the detailed flow field having the shocklets and vortices. Numerical solvers having negligible numerical dissipation are highly essential to predict these structures accurately. Over the years, many researchers have obtained a grid-converged solution for shock-wave boundary-layer interaction at Reynolds numbers of 200 and 1000. The shock-wave boundary-layer interaction at higher Reynolds numbers is not attempted due to the requirement of huge computational resources and challenges associated with convergence. In the investigation, the grid-converged solution is obtained for a Reynolds number of 2500 with a 13th order high-resolution hybrid scheme using 100 cores of a computational cluster equipped with 3.0 GHz Intel Xeon processors incorporating the MPI library for parallelization. The complex flow field is analyzed in detail using wall density, density gradient, vorticity, pressure, and enstrophy plots after validating the solver with the benchmark wall pressure,density, and v-velocity around the primary vortex provided by Zhou et al. (2018), Phys. Fluids, 30, 016102 for a Reynolds number 1000. It is observed that the triple point height and number of vortices in both separated zone and at the shear layer increase with an increase in Reynolds number. The grid-converged data obtained from the present simulation can be used as benchmark data to validate different numerical schemes/solvers at higher Reynolds numbers.
Article
Full-text available
In the starting phases of continuously blowing under-expanded jets, this numerical study investigates the effect of co-flow ([Formula: see text]) (a) on the circulation and evolution of primary vortex ring ( PVR) and (b) on the occurrence of Mach reflection, slipstream generation, and subsequent formation of counter rotating vortex rings ( CRVRs). With increase in co-flow ([Formula: see text]), the PVR circulation gradually decreases. The size of supersonic PVR gradually decreases with increase in co-flow ([Formula: see text]), and at high magnitudes of co-flow ([Formula: see text]), the supersonic PVR attains a circular shape. The strengths of embedded shock ( ES) and vortex-induced shock are found to decrease with increase in co-flow ([Formula: see text]), and at high magnitudes of co-flow ([Formula: see text]), these shocks may even cease to form inside the supersonic PVR. An increase in co-flow ([Formula: see text]) causes the expansion fan to become more and more narrow. This reduces the acceleration of the supersonic flow inside the inviscid core, thereby weakening the incident oblique shock ( IOS), which in turn increases the pressure prevailing downstream of this shock inside the inviscid core. The increase in co-flow ([Formula: see text]) also leads to a simultaneous decrease in the pressures prevailing in front of the downstream marching PVR and Mach disk ( MD) of the inviscid core due to the reduction in the strength of precursor shock. As the magnitudes of pressures prevailing in the upstream and downstream of Mach disk approach each other, hence, MD also weakens. This shows that with the increase in co-flow ([Formula: see text]), there is weakening of the different shocks (i.e., ES, IOS, and MD) involved in Mach reflection. This causes a reduction in the strength of the resulting slipstream, thereby affecting the formation of CRVR patterns.
Article
A discrete-velocity Boltzmann equation (DVBE) with Bhatnagar-Gross-Krook (BGK) approximation is discretized in time and space using a third-order Runge-Kutta (RK3) and fifth-order weighted essentially nonoscillatory (WENO) finite-difference scheme to simulate benchmark inviscid compressible flows. The implementation of the WENO ensures that solutions behave with minimal or no oscillations, narrowing the gap between the exact and the numerical results. Discrete-velocity sets given by Kataoka and Tsutahara [Phys. Rev. E 69, 056702 (2004)] are used. The additional dissipation terms as well as artificial viscosity are incorporated in the formulation to solve the compressible flows at high Mach number. Further, the flows which are subjected initially to a high density ratio are effectively simulated. In this article, one-dimensional benchmarks are simulated at initial Mach number up to 30 and density ratio up to 1000, whereas, the benchmarks in two dimensions are simulated with a Mach number up to 10. The algorithm is assessed by simulating numerous benchmarks, namely, (i) one-dimensional Riemann problem for various shock waves combinations [namely (a) shock-shock waves in the case of different Mach numbers, (b) rarefaction-shock waves for various density ratios, (c) sudden contact shock discontinuity, and (d) shock-rarefaction waves for density ratio 5], (ii) isentropic vortex convection test, (iii) regular shock reflection (RR) for Mach numbers 2.9 and 10, (iv) double Mach reflection (DMR) for inflow Mach numbers as 6 and 10, and (v) forward-facing step for inflow Mach numbers 2 to 5. Further, the effect of change in Mach numbers and wedge angles on the flow structures in the case of DMR are detailed. In the case of a forward-facing step, the variations of flow structure (e.g., the Mach stem height, triple points location, and shock standoff distance) are detailed with respect to Mach number, step height, and specific-heat ratios. Finally, the numerical stability of the proposed formulation is carried out to assess the behavior of the free parameters.
Book
This text describes shock wave reflection phenomena from a phenomenological point of view. Organized in five parts, the book covers an introduction to oblique shock wave reflection and the governing equations of the two- and three-shock theories; shock wave reflection in pseudo-steady flows; and reflection phenomena in steady flows and unsteady flows. With regard to pseudo-steady, steady and unsteady flows, the possible types of specific reflections are described, criteria for their formation and termination are presented and their governing equations are solved analytically and graphically and compared with experimental results. Modification of the governing equations by accounting for viscous and real gas effects are suggested. In addition, unresolved problems are pointed out and ideas for future research are suggested. The fifth part of the book constitutes a detailed source list of most of the scientific papers and reports which have been published in the subject area.
Article
Linear instability waves in supersonic shear layers are analyzed. Both viscous and inviscid disturbances are considered. The basic state is obtained by solving the compressible laminar boundary‐layer equations or is specified by the hyperbolic tangent profile. The effects of the velocity ratio and temperature ratio are determined. The numerical results show that the maximum growth rate depends nonlinearly on the velocity ratio. The results also substantiate the convective Mach number as a compressibility parameter for mixing layers.
Article
The nonlinear growth, of the multimode incompressible Kelvin-Helmholtz shear flow instability at all density ratios is treated by a large-scale statistical-mechanics eddy-pairing model that is based on the behavior of a single eddy and on the two eddy pairing process. From the model, a linear time growth of the mixing zone is obtained and the linear growth coefficient is derived for several density ratios. Furthermore, the asymptotic eddy size distribution and the average eddy life time probability are calculated. Very good agreement with experimental results and full numerical simulations is achieved.
Article
The unsteady inviscid two-dimensional flow field and the wave configurations which result when a supersonic vehicle strikes a planar oblique shock wave were modelled and analytically predicted using some approximations and simplifying assumptions. Based on the two- and three-shock theories together with the geometric shock dynamics theory, both regular (windward) and irregular (leeward) shock-on-shock (S-O-S) interactions were investigated, and the transition criterion between them was suggested. For the case of regular S-O-S interaction, the transmitted shock wave reflects over the vehicle body surface either as a regular (RR) or a Mach reflection (MR) depending on the inclination angle and the strength of the impingement shock wave. A pronounced peak surface pressure jump was found to exist during the transition from RR to MR. A RR[leftrightarrow A: l&r arrow]MR transition criterion when the flow ahead of the shock pattern is not quiescent was proposed. Predictions based on the model developed here are superior to those of approximate theories when compared to the available experimental data and numerical simulations.
Article
The flow fields associated with Mach reflection wave configurations in steady flows are analysed, and an analytical model for predicting the wave configurations is proposed. It is found that provided the flow field is free of far-field downstream influences, the Mach stem heights are solely determined by the set-up geometry for given incoming-flow Mach numbers. It is shown that the point at which the Mach stem height equals zero is exactly at the von Neumann transition. For some parameters, the flow becomes choked before the Mach stem height approaches zero. It is suggested that the existence of a Mach reflection not only depends on the strength and the orientation of the incident shock wave, as prevails in von Neumann's three-shock theory, but also on the set-up geometry to which the Mach reflection wave configuration is attached. The parameter domain, beyond which the flow gets choked and hence a Mach reflection cannot be established, is calculated. Predictions based on the present model are found to agree well both with experimental and numerical results.
Article
The paper reports results of shock tube experiments of the attenuation of shock waves propagating over arrayed baffle plates, which is motivated to simulate shock wave attenuation created accidentally at the acoustic delay line in synchrotron radiation factory upon the rupture of a metal membrane separating the acceleration ring at high vacuum and atmospheric test chambers. Experiments were carried out, by using double exposure holographic interferometry with double path arrangement, in a 100 mm×180 mm shock tube equipped with a test section of 180 mm×1100 mm view field. Two baffle plate arrangements were tested: Oblique and staggered baffle plates; and vertical symmetric ones. Pressures were measured along the shock tube sidewall at individual compartments for shock Mach numbers ranging from 1.2 to 3.0 in air. The results were compared with a numerical simulation. The rate of shock attenuation over these baffle plates was compared for vertical and oblique baffle plates. Shock wave attenuation is more pronounced in the oblique baffle plate arrangements than in the vertical ones.
Article
Plane turbulent mixing between two streams of different gases (especially nitrogen and helium) was studied in a novel apparatus. Spark shadow pictures showed that, for all ratios of densities in the two streams, the mixing layer is dominated by large coherent structures. High-speed movies showed that these convect at nearly constant speed, and increase their size and spacing discontinuously by amalgamation with neighbouring ones. The pictures and measurements of density fluctuations suggest that turbulent mixing and entrainment is a process of entanglement on the scale of the large structures; some statistical properties of the latter are used to obtain an estimate of entrainment rates. Large changes of the density ratio across the mixing layer were found to have a relatively small effect on the spreading angle; it is concluded that the strong effects, which are observed when one stream is supersonic, are due to compressibility effects, not density effects, as has been generally supposed.
Article
The Rayleigh-Taylor instability in its highly nonlinear, turbulent stage causes atomic-scale mixing of the shell material with the fuel in the compressed core of inertial-confinement fusion targets. The density of shell material mixed into the outer core of direct-drive plastic-shell spherical-target implosions on the 60-beam, OMEGA laser system is estimated to be 3.4(+/-1.2) g/cm(3) from time-resolved x-ray spectroscopy, charged-particle spectroscopy, and core x-ray images. The estimated fuel density, 3.6(+/-1) g/cm(3), accounts for only approximately 50% of the neutron-burn-averaged electron density, n(e)=2.2(+/-0.4)x10(24) cm(-3).
Article
A turbulent mixing transitional criterion and a procedure for modeling and predicting the required time interval was developed to achieve transition when the background flow is unsteady rather than stationary. It was emphasized that such a criterion and estimation procedure is essential for analysis, experimental design, and diagnostic development. Focus was on studies of extremely energetic, high pressure, supersonic combustion, hypersonic aerothermodynamic design, and astrophysical stellar and planetary evolution research, among others.
Article
A correct description of turbulent mixing is particularly taxing on our understanding of turbulence; such a description relies on an account of the dynamics spanning the full spectrum of scales. Specifically, to describe the entrainment stage that is responsible for the engulfment of large pockets of irrotational fluid species into the turbulent flow region (Brown & Roshko 1974), the large-scale flow structures need to be correctly described. Secondly, to describe the subsequent kinematic stirring process responsible for the large interfacial surface generation between the mixing species, the intermediate range of scales must be correctly accounted for. These are below the largest in the flow in size, but above the smallest affected by viscosity and molecular diffusivity. Finally, the dynamics at the smallest scales must be captured to describe the molecular mixing process itself. These three phases of turbulent mixing...