ArticlePDF Available

Richtmyer–Meshkov-like instabilities and early-time perturbation growth in laser targets and Z-pinch loads

Authors:

Abstract and Figures

The classical Richtmyer–Meshkov (RM) instability develops when a planar shock wave interacts with a corrugated interface between two different fluids. A larger family of so-called RM-like hydrodynamic interfacial instabilities is discussed. All of these feature a perturbation growth at an interface, which is driven mainly by vorticity, either initially deposited at the interface or supplied by external sources. The inertial confinement fusion relevant physical conditions that give rise to the RM-like instabilities range from the early-time phase of conventional ablative laser acceleration to collisions of plasma shells (like components of nested-wire-arrays, double-gas-puff Z-pinch loads, supernovae ejecta and interstellar gas). In the laser ablation case, numerous additional factors are involved: the mass flow through the front, thermal conduction in the corona, and an external perturbation drive (laser imprint), which leads to a full stabilization of perturbation growth. In contrast with the classical RM case, mass perturbations can exhibit decaying oscillations rather than a linear growth. It is shown how the early-time perturbation behavior could be controlled by tailoring the density profile of a laser target or a Z-pinch load, to diminish the total mass perturbation seed for the Rayleigh–Taylor instability development. © 2000 American Institute of Physics.
Content may be subject to copyright.
RichtmyerMeshkov-like instabilities and early-time perturbation growth
in laser targets and Z-pinch loads*
A. L. Velikovich,J. P. Dahlburg, and A. J. Schmitt
Plasma Physics Division, Naval Research Laboratory, Washington, D.C. 20375
J. H. Gardner and L. Phillips
Laboratory for Computational Physics and Fluid Dynamics, Naval Research Laboratory,
Washington, D.C. 20375
F. L. Cochrana) and Y. K. Chong
Berkeley Research Associates, Incorporated, Springfield, Virginia 22150
G. Dimonte
Lawrence Livermore National Laboratory, Livermore, California 94551
N. Metzler
Physics Department, Nuclear Research Center Negev, P. O. Box 9001, Beer Sheva, Israel, and Science
Applications International Corporation, McLean, Virginia 22150
Received 18 November 1999; accepted 19 January 2000
The classical Richtmyer–Meshkov RMinstability develops when a planar shock wave interacts
with a corrugated interface between two different fluids. A larger family of so-called RM-like
hydrodynamic interfacial instabilities is discussed. All of these feature a perturbation growth at an
interface, which is driven mainly by vorticity, either initially deposited at the interface or supplied
by external sources. The inertial confinement fusion relevant physical conditions that give rise to the
RM-like instabilities range from the early-time phase of conventional ablative laser acceleration to
collisions of plasma shells like components of nested-wire-arrays, double-gas-puff Z-pinch loads,
supernovae ejecta and interstellar gas. In the laser ablation case, numerous additional factors are
involved: the mass flow through the front, thermal conduction in the corona, and an external
perturbation drive laser imprint, which leads to a full stabilization of perturbation growth. In
contrast with the classical RM case, mass perturbations can exhibit decaying oscillations rather than
a linear growth. It is shown how the early-time perturbation behavior could be controlled by
tailoring the density profile of a laser target or a Z-pinch load, to diminish the total mass perturbation
seed for the Rayleigh–Taylor instability development. © 2000 American Institute of Physics.
S1070-664X0093305-6
I. INTRODUCTION
The classical Richtmyer–Meshkov RMinstability1,2
develops when a planar shock wave interacts with a corru-
gated interface between two different fluids. After 45 yr of
studies that followed Richtmyer’s theoretical discovery,1the
main features of this instability at the linear and weakly non-
linear regimes are firmly established, and the experimental
data, numerical simulation results and theoretical predictions
are in good agreement e. g., see Refs. 3–5 and references
therein.
In this paper, we discuss a wider class of instabilities
which, not being caused by the shock-interface refraction,
are still driven by exactly the same physical mechanisms as
the classical RM instability. This class includes but is not
limited tointerfacial instabilities caused by the ‘‘virtual
gravity’’;6,7 instabilities produced in collisions of perturbed
fluid layers8–10 and/or of perturbed shock waves with each
other or material interfaces;11 instabilities excited in simulat-
ing the evolution of an initial discontinuity in a fluid the
perturbed Riemann problem12,13; instabilities of shock-
piston flows, such as those produced when a shock wave is
driven from a rippled surface by a uniform laser beam, or
when a nonuniform laser beam irradiating a planar surface
imprints mass perturbations into the flow.14–22 In the litera-
ture, these instabilities are sometimes referred to simply as
RM,6–10 and sometimes identified as new kinds of
instability.11 For lack of a better term, all these instabilities
will be called RM-like. Our discussion is mostly limited to
the small-amplitude, linear phase of the instability develop-
ment.
One of the most interesting features of the flows that
exhibit RM-like instabilities in contrast, say, with the
Rayleigh–Taylor RTunstable flowsis their proximity to
the boundary separating stable and unstable situations. Con-
sequently, perturbation evolution is sensitive to relatively
small changes in the formulation of the problem, physical
mechanisms accounted for or ignored, initial conditions, etc.,
which can turn a stable situation into unstable, and vice
*Paper GT1 2 Bull. Am. Phys. Soc. 44, 122 1999.
Tutorial speaker.
aPresent address: Los Alamos National Laboratory, Los Alamos, New
Mexico 87545.
PHYSICS OF PLASMAS VOLUME 7, NUMBER 5 MAY 2000
16621070-664X/2000/7(5)/1662/10/$17.00 © 2000 American Institute of Physics
Downloaded 06 Aug 2007 to 128.138.96.216. Redistribution subject to AIP license or copyright, see http://pop.aip.org/pop/copyright.jsp
versa. This is one of the reasons that the studies of such
instabilities are an exciting playground for perfecting simu-
lation techniques. Even more important, however, are the
opportunities for laser target and Z-pinch load design that
such studies might present. Suppression of the RM-like per-
turbation growth at the early stages of laser–target accelera-
tion or a Z-pinch implosion could make a big difference in
the uniformity of imploded or stagnated plasmas.
This paper is structured as follows. In Sec. II we review
the basic features of the classical RM instability that are
relevant for further discussion, and in Sec. III present some
examples illustrating the RM-like instabilities that are dis-
tinct from the classical case of shock–interface interaction.
Section IV reviews the RM-like instabilities in ablatively
driven shock-piston flows. Section V shows how the RM-
like growth could be suppressed using tailored density pro-
files in both Z-pinch loads and laser targets. In Sec. VI, we
conclude with a discussion.
II. THE CLASSICAL RM INSTABILITY
Here we briefly review the basic features of the classical
RM instability in the small-amplitude regime. Let a planar
shock wave be normally incident from fluid 1 upon a contact
interface separating it from a different fluid, 2. The interface
is slightly corrugated. In the small-amplitude theory we deal
with a single-mode sine corrugation
x0exp(iky), where
x0is the preshock corrugation amplitude, and k2
/is
the perturbation wave number. After the interaction, the in-
terface acquires a constant velocity Uin the direction of the
incident shock wave, a perturbed shock wave is transmitted
into fluid 2, and either a perturbed shock wave or a perturbed
centered rarefaction wave is reflected back into fluid 1.
Typical behavior of perturbation amplitudes and growth
rates is illustrated by Fig. 1b. The fluid parameters in the
figure correspond to the ideal gas model23 developed to
simulate interaction of strong radiatively driven shock waves
with a contact interface between a solid beryllium ablator
and a foam tamper in the RM experiments on Nova laser at
Lawrence Livermore National Laboratory LLNL.24,3 Fig-
ure 1 is plotted for a strong incident shock wave, Mach num-
ber M010.8.
The displacement amplitude of the contact interface
x(t) grows with time. Note that for the whole class of the
RM-like instabilities, the growth rate, which we denote by
C(t)(d/dt)
x, has a dimensionality of velocity cm/sin
contrast with the case of RT and most of the other fluid
instabilities, which are characterized by exponential growth
rates expressed in inverse seconds. This is because the RM-
like instabilities exhibit at most linear with time, secular,
rather than exponential perturbation growth. When an insta-
bility of this class develops at a contact interface, then the
growth rate defined above is the same as the perturbation
amplitude of the axial velocity at the interface, C(t)
vx(xxC,t)this is not necessarily the case, e.g., the
local velocity perturbation amplitude at the ablation front
does not coincide with the growth rate. The growth rate C,
after some damped oscillations at late time kUt1, ap-
proaches a constant asymptotic value, ; in other words
linear late-time perturbation growth. The solid line labeled S
shows the displacement amplitude at the shock front
xs;
the dotted line shows relative pressure perturbation
ps/p.
The transmitted shock wave is seen to be superstable, which
means that its perturbations actually decrease in time as
t3/2, except when the transmittedshock wave is very
strong, and they decay as t1/2, see Refs. 25. If a shock wave
is reflected, its perturbations behave similarly.
The reflected centered rarefaction wave is unstable.26,27
The ripple amplitude at its leading edge remains constant in
time, whereas perturbations of its trailing edge grow linearly
in time. For the case shown in Fig. 1, the corresponding
asymptotic growth rate is about four times higher than the
interface growth rate and has a different sign.
The physical factors that drive the classical RM instabil-
ity are illustrated in Fig. 1a. Instant deposition of localized
vorticity at the interface during the shock refraction makes
the fluid move along the contact interface, decreasing the
pressure at the convex side of a bubble darker side, which
then makes the bubble grow. The pressure perturbations
coming from the stable shock frontsaffect the growth, typi-
cally slowing it down even a full cancellation is possible,
leading to zero asymptotic growth rate, so called
freezeout12,28. The asymptotic growth rate is expressed via
an integral of all these pressure perturbations
p(x,t), com-
ing to the contact interface C:
FIG. 1. aThe flow structure at the density interface supports the RM
instability growth; vortical flow along the contact interface Ccreates a pres-
sure gradient directed from the convex side of the bubble, and sonic pressure
perturbations sare exchanged between the interface and the shock front S.
bLinear RM growth for adiabatic exponents of fluids 1 and 2
1
1.8,
21.45, respectively, and preshock Atwood number 0.868. Time
histories of normalized growth rates at the interface (C) and at the trailing
edge of the reflected rarefaction wave (R/4) , shock displacement and shock
pressure perturbation amplitudes solid and dotted lines S, respectively;the
dotted line MB is the asymptotic interfacial growth rate predicted by the MB
prescription Ref. 30. The perturbation amplitudes and growth rates are
shown in units of k
x0and kU
x0, respectively.
1663Phys. Plasmas, Vol. 7, No. 5, May 2000 Richtmyer-Meshkov-like instabilities and early-time...
Downloaded 06 Aug 2007 to 128.138.96.216. Redistribution subject to AIP license or copyright, see http://pop.aip.org/pop/copyright.jsp
⫽⫺ 1
*
x
0
px,tdt
xxC
,1
where
*is the constant postshock density at the either side
of the interface xxC.
The two distinct contributions to the growth rate are
separated in the following exact formula for :
1
*
vy1
*
2
*
vy2
*
1
*
2
*
1
*Fs1
2
*Fs2
1
*
2
*⬅⌫0⌬⌫s,
2
derived by Wouchuk and Nishihara WN.13 Here
1,2
*,
vy1,2
*
are densities and transverse velocities immediately after the
shock–interface interaction, and Fs1,2 are parameters ac-
counting for the contribution of perturbations to the
asymptotic growth rate from reflected and transmitted shock
waves in the case of a reflected rarefaction wave, Fs10).
The first contribution 0comes from the localized vorticity
deposited at the contact interface immediately after the shock
refraction, and has an explicit analytical formula. The second
contribution ⌬⌫swhich, as mentioned above, typically has a
different sign than 0, comes from the interaction of the
perturbed interface with one or two shock waves propagating
from it. Formulas for Fs1,2 are not yet available in a closed
form; the procedure of their approximate evaluation de-
scribed in Ref. 13 yields the same asymptotic growth rates as
the other methods.12,26,29 Interaction with the unstable re-
flected rarefaction wave does not contribute to . The
value of ⌬⌫sin the WN formula 2could be determined via
the time histories of the pressure perturbations at the shock
front.
The RM instability is often compared to and sometimes
is believed to be the same asthe particular case of RT in-
stability, when the gravity acceleration, rather than being
constant which corresponds to the classical exponential
growthor slowly varying, is impulsive: g(t)U
(t), simu-
lating instantaneous shock acceleration. This approach yields
the well-known Richtmyer’s estimate for the asymptotic
growth rate
0
0dtkgtAt
xCtkUAeff
xeff ,3
where the instant t0 corresponds to the shock–interface
interaction, and Aeff and
xeff are some effective values of
the Atwood number and interfacial perturbation amplitude
that both change discontinuously when the shock wave
passes the interface. To actually use Eq. 3for estimates, it
has to be replaced by some prescription. The available op-
tions
kUA*
xC
*R;
kUA*1
2
xC
*
x0MB;
kU1
2A*
xC
*A
x0VMG;
4
have been suggested, respectively, by Richtmyer R,1Meyer
and Blewett MB,30 and Vandenboomgaerde et al.
VMG.31 The prescriptions are not exact expressions for
, but rather heuristic formulas approximating it. In the
weak-shock limit, all the prescriptions yield the same result,
which is consistent with compressible theory. The R pre-
scription sometimes works better for the reflected shock
case, MB—for the reflected rarefaction case, and VMG—for
both cases, if the shock wave is not too strong. None of the
prescriptions are reliable for strong shocks.For the particu-
lar case of a strong shock in Fig. 1, the MB prescription
happens to be a very good approximation of the asymptotic
growth rate, while R and VMG predictions are not 0.11 and
0.38, respectively, normalized as in Fig. 1. For the ex-
ample of air/He shock–interface interaction at M010, none
of the prescriptions are found to work: R, MB and VMG
normalized growth rates are 0.11, 0.28, and 0.32, respec-
tively, whereas the actual asymptotic growth rate equals
0.16.
III. SOME EXAMPLES OF RM-LIKE INSTABILITIES
Comparing 2and 3, we see that the former formula
describes a wider class of instabilities driven by the same
physical mechanisms. To drive this kind of instability, we
have to deposit initial localized vorticity at the contact inter-
face and/or provide sonic interaction of the interface with
one or two shock waves propagating from it. This could be
accomplished not only through classical shock–interface in-
teraction, but also in many other ways. This is how we come
to the generalization of RM-like instabilities. They are de-
fined here as the interfacial instabilities driven by the same
physical factors as the classical RM, but differing from it by
the source of the unstable configuration, and sometimes as
well by some other factors affecting the perturbation evolu-
tion.A very good example is the instability of an impulsively
accelerated fluid interface, observed using the novel experi-
mental techniques that allow the experimentalists to produce
controlled time histories of acceleration acting upon the ob-
served interface. The linear electric motor LEMbuilt at
LLNL6provides variable electromagnetic acceleration of test
vessels. The free-falling tank technique7makes the test ves-
sel fall and then bounce off a fixed spring, producing a sharp
acceleration pulse. Both techniques allow the possibility of
incompressible fluids and the production of a RM-like insta-
bility development in the absence of any shocks. For the case
of a small-amplitude, single-mode initial perturbation, Richt-
myer’s formula 3is an exact result rather than an approxi-
mation; for incompressible fluids, it is well defined, since
neither Anor
xCchange during the acceleration pulse, and
no prescriptions are necessary. This exact linear asymptotic
growth rate coincides with the contribution 0of instantly or
rapidly deposited interface vorticity, as defined in 2. This
case could be regarded as a ‘‘more-than-classical’’ regime of
the RM-like growth, which literally corresponds to the Rich-
tmyer’s model of impulsive acceleration. Figure 2 presents
an example of experimental results obtained on LEM6for
such an acceleration regime. Here, the RM-like growth ex-
cited by a short acceleration pulse is undoubtedly a particular
case of and, understandably, slower thanthe RT growth
supported by continuous acceleration.
1664 Phys. Plasmas, Vol. 7, No. 5, May 2000 Velikovich
et al.
Downloaded 06 Aug 2007 to 128.138.96.216. Redistribution subject to AIP license or copyright, see http://pop.aip.org/pop/copyright.jsp
Now consider a completely different example, illustrated
in Fig. 3. In the initial state, two identical uniform gases
(
5/3), separated by a sine-shaped contact interface, have
identical constant mass velocities, UandUat t0, both
directed from the interface. This ‘‘anticollision’’ produces
two centered rarefaction waves of identical strength that
propagate from the interface.12 If Udoes not exceed the ve-
locity of free gas expansion into vacuum as in the case of
Fig. 3, then the perturbed contact interface continues to exist
at t⬎⫹0, separating the gas layers left behind the trailing
edges of the rarefaction waves. There is no impulsive accel-
eration here: the xvelocity of the interface is exactly zero
both at t⫽⫺0 and at t⫽⫹0. The displacement amplitude
x0does not change in this interval. Similarly, there is no
uncertainty in the Atwood number, it remains exactly zero
throughout the interaction. This situation therefore has no
resemblance to the shock-excited or impulsively driven RT
instability, and formula 3, as well as any of the prescrip-
tions 4, predicts zero perturbation growth rate. However,
due to our discontinuous initial conditions, a known amount
of localized vorticity is deposited at the interface at t⫽⫹0.
According to 2, this is enough to drive a RM-like pertur-
bation growth. Moreover, since there are no shock waves for
the contact interface to interact with, the prediction of the
WN formula 2is available in a closed form, 0
⫽⫺kU
x0. This result13 is shown in Fig. 3 to be in full
agreement with the linear theory;12 so far, it is the only
known example when the asymptotic growth rate of a RM-
like instability in a compressible fluid is exactly expressed by
a closed analytical formula.
Let the same problem be modified for initial conditions
corresponding to a collision rather than anticollision Fig. 4,
inset, with two identical shock waves propagating from the
contact interface after interaction. For the same reasons as in
the previous example, this case has nothing to do with the
impulsively driven RT instability, and formulas 3,4pre-
dict zero growth rate for it. In this case, however, the shock-
interface contribution to the growth rate ⌬⌫sis nonzero. To
verify the prediction of the linear theory for this case, we did
a numerical simulation of this problem with the FAST2D hy-
drocode developed at the Naval Research Laboratory
NRL.32 Figure 4 demonstrates good agreement between the
perturbation growth predicted by the linear theory and the
simulation results. This is also an unmistakably RM-like type
of perturbation growth.
The interface does not even have to exist before the in-
teraction. Remove any half from the initial configurations of
Figs. 3, 4, and impose instead a boundary condition requiring
constant pressure at the perturbed surface no longer an in-
terface. None of the results of Figs. 3, 4 would change. We
see that any shock-piston flow driven by a pressure applied to
a free surface would excite a RM-like instability, no matter
whether it is the surface that is initially perturbed, or whether
the pressure driving it is slightly nonuniform. Alternatively,
the contact interface could be created when two uniform lay-
FIG. 2. aThe acceleration profiles produced in four linear electric motor
runs, varying from approximately constant, to increasing, to decreasing, to
impulsive. bTime evolution of the bubble amplitudes for these four cases.
Most of the observed growth is in the nonlinear stage. The RM-like growth
corresponds to impulsive acceleration .
FIG. 3. Perturbation growth at a contact interface for a symmetrical Rie-
mann problem producing two outgoing rarefaction waves, and U0.6
a0,wherea0is the speed of sound before interaction. Units are the same as
in Fig. 1. Insetscheme of the initial state.
FIG. 4. Same as in Fig. 3 for the case of two outgoing shock waves,
6/5 and U2.9a0.Insetscheme of the initial state.
1665Phys. Plasmas, Vol. 7, No. 5, May 2000 Richtmyer-Meshkov-like instabilities and early-time...
Downloaded 06 Aug 2007 to 128.138.96.216. Redistribution subject to AIP license or copyright, see http://pop.aip.org/pop/copyright.jsp
ers of different fluids collide head-on, depositing initial vor-
ticity at the interface separating them, and/or producing
rippled shock waves. The existing methods of calculating
linear time histories of perturbation growth and asymptotic
growth rates,1,12,13,26 including Eq. 2, are directly appli-
cable to all these cases, provided that the fluidsahead of the
shock and or rarefaction wavespropagating from the inter-
face or surface, are uniform and unperturbed.
If the latter condition is not met, it does not change the
nature of the RM-like growth. However, in this case, no
linear theory is yet available to describe it. The simplest
example is the reshock of the contact interface by the trans-
mitted shock wave that is reflected from the end of a shock
tube or a computational box, and is incident back upon the
interface, carrying the perturbations that are coherent with
those already in the flow. Another example is interaction of a
rippled shock wave with a contact interface.11 In both cases,
the factors that drive the perturbation growth remain the
same as shown in Fig. 1a. Calculation of the time histories
and asymptotic growth rates, where applicable, is more com-
plicated because it has to account for the perturbations enter-
ing the flow via either perturbed shock wave or perturbed
rarefaction wave. This has not been done yet. Probably the
same computational difficulty has so far prevented a theoret-
ical analysis of the classical RM instability in magnetohydro-
dynamics MHD, which for some cases has been studied
numerically in Ref. 33. Here, after interaction, three waves
slow, Alfve
´n and fastpropagate in each direction, and one
has to describe how the perturbations reaching the interface
interact with all of these. This case, when appropriately stud-
ied, would present abundant opportunities for benchmarking
multidimensional MHD hydrocodes: there are 189 non-
equivalent cases of perturbed Riemann problems in MHD,
compared to only five in ordinary gas dynamics two of
which have been illustrated above.
We conclude that the family of the RM-like instabilities
extends well beyond the strict limits of applicability of Rich-
tmyer’s original analogy 3of the perturbed interface with a
pendulum driven by a short pulse of gravity. There might be
no gravity, and the pendulum does not have to exist before it
is kicked. Nevertheless, the analogy with a pendulum with-
out gravity remains useful for all the RM-like instabilities. It
clarifies that the instability has no external energy source to
drive the perturbation growth, only a limited deposit of per-
turbation energy left from the initial conditions or some tran-
sient phase. This would generally imply a linear rather than
exponential perturbation growth. Any physical factor capable
of dissipating this limited energy, or converting it into some
other form, can completely suppress the RM-like instability
development. For instance, even though the MHD RM prob-
lem has not yet been solved, there can be little doubt that in
any configuration with k"B0 at either side of the contact
interface after interaction, its perturbations would oscillate
instead of growing linearly, because bending of magnetic
force lines requires extra energy.
IV. RM-LIKE INSTABILITIES OF ABLATIVELY DRIVEN
SHOCK-PISTON FLOW
A shock-piston flow driven by a uniform pressure ap-
plied to a rippled surface is a good example of a RM-like
unstable flow. The analogy with a pendulum without gravity
suggests that if a constant pressure is applied to a rippled
surface of a half-space filled with a uniform gas, driving a
rippled shock into it, then a linear surface perturbation
growth will follow, the asymptotic growth rate being ex-
pressed by Eq. 2. However, if the pressure is nonuniform
and constant in time, then the constant force driving our
pendulum would supply it with a constant acceleration, pro-
ducing quadratic in time perturbation growth at the interface.
On the other hand, if the external nonuniform pressure acts
only for a limited time, or decays fast enough, then it is
equivalent to a single kick, producing linear growth. Apply-
ing the RM linear theory to these cases, one can see that
perturbation growth indeed occurs in all these cases just as
described above see Fig. 5cand Ref. 20.
The shock-piston flow, with a shock of constant strength
driven by a constant pressure, is a good model for interaction
of the low-intensity foot of the laser beam with a direct-drive
target. During the foot of the laser pulse, the target material
is slightly shock-accelerated and somewhat shock com-
pressed to prepare it for the arrival of the main laser pulse.34
During the shock-transit interval, most of the mass perturba-
tions due to the laser beam nonuniformity and the surface
roughness of the target are imprinted into the flow. This im-
printing that proceeds essentially as a RM-like instability
development, as first identified in Ref. 14, has been studied
by many authors.14–22 Due to thermal smoothing of the laser
perturbations in the developed corona, the effect of the driv-
ing laser beam nonuniformity on the uniformity of ablative
pressure for any given wavelength lasts only for a finite in-
terval of time. Then both causes of the nonuniformity could
be expected to cause linear perturbation growth.
Surprisingly, perturbation growth due to both laser non-
uniformity and surface roughness has been observed to slow
down instead of growing linearly.15–17 It should be empha-
sized that the saturation-like slowing down was observed in
the linear, small-amplitude regime, and has nothing to do
with nonlinear saturation of perturbation growth at large am-
plitudes, like that observed in Fig. 2b. Moreover,
simulations19 have demonstrated that after reaching some
maximum value the mass perturbations exhibit decaying os-
cillations, like those presented in Figs. 5a,5b. Having
observed decaying oscillation of our pendulum, we must
identify two factors: the effective gravity that makes it oscil-
late, and the effective friction that makes the oscillations
decay.
The role of effective friction is played by mass
ablation.18–20 Since the ablative mass leaves the flow region
between the ablation front and the shock front, it carries
away all of its vorticity and pressure perturbations, which
drain the limited energy resource available for perturbation
development, as explained in Sec. III. If the friction was the
only stabilizing effect taken into account, then the pendulum,
after being kicked, would stop at some amplitude, implying
1666 Phys. Plasmas, Vol. 7, No. 5, May 2000 Velikovich
et al.
Downloaded 06 Aug 2007 to 128.138.96.216. Redistribution subject to AIP license or copyright, see http://pop.aip.org/pop/copyright.jsp
saturation of perturbation growth curve 3 in Fig. 5c兲兴.
There is, however, another stabilizing influence, the ‘‘rocket
effect’’35,36 that acts as effective gravity. It is based on the
fact that the ablation front is an isotherm.37 When the front
moves into a higher temperature area, its temperature does
not increase, but the temperature gradient near it does. In
other words,22 since the temperature distribution in the co-
rona is smoother than the ablation front, the tops of the
ripples that come closer to the laser receive more energy than
the bottoms. Then, as first noticed long ago Ref. 38, Section
‘‘Stability’’, the increased temperature gradient speeds up
the mass ablation, producing the dynamic pressure pushing
the front back. This mechanism acts as an effective gravity.
The theory of RM-like growth with the rocket effect term
taken into account was developed in Ref. 21 and demon-
strated oscillations qualitatively similar to those observed in
Fig. 3. It would be highly desirable to carry out a special
experiment to observe the full oscillation cycle in imprinting,
beyond the vicinity of the first peak, to test both the simula-
tion codes and the analytical models.21,22
It should be noted how different the effect of these
physical phenomena on the ablative RT and RM-like insta-
bilities is. In the developed acceleration regime, for relevant
wavelengths, neither the small friction nor the small
‘‘rocket’’ gravity can strongly affect the RT instability of the
flow, where perturbations derive their energy from a power-
ful external energy source, ‘‘gravity’’ ablative acceleration.
On the other hand, each of these mechanisms separately is
sufficient to completely suppress the RM-like growth.
Since the mass perturbations in the target should be
small otherwise laser fusion will not work, the contribution
of the surface roughness could be simulated as described
above, and shown in Fig. 5b, modeling each relevant wave-
length separately and then summing it up. If the behavior of
each Fourier mode is known, then we only need to know the
spectrum of initial surface roughness to predict the total rms
contribution of this source for any instant. This is not the
case, however, for the laser imprint. A laser beam that is well
smoothed with, for example, induced spatial incoherence
ISItechnology39 does not impose constant phase perturba-
tions onto the target. Rather, phases and amplitudes of the
perturbations change very rapidly, so that the perturbations
in the target are essentially noise driven. We can however
apply the elementary theory of noise driven oscillators to see
if it provides any helpful insight into the actual behavior of
ISI perturbations.
Let as assume linear response of the mass perturbation to
the corresponding Fourier component of the driving pertur-
bation
mkt
0
tdtGkt,t
Ikt,5
where Gk(t,t) is the corresponding Green function, and
Ik(t) is the randomFourier component of the relative in-
tensity variation. Ensemble averaging immediately shows
that
mk(t)
0the noise does not drive mass perturba-
tions in any preferential direction, but
mkt
2
1/2
mk,rmst
c
0
t
Gkt,t
2dt
1/2,
6
where the driving noise is assumed to be delta correlated:
Ik(t)
Ik
*(t)
c
(tt). Note that the linear relation
5is by no means guaranteed. Since the laser speckle struc-
FIG. 5. aSimulated growth of mass perturbations in a DT target irradiated
by a KrF laser at 2.81013 W/cm2with single-mode, constant phase 10%
laser intensity perturbation; bsame for initial surface ripple; cfor the
same conditions and 10% laser intensity perturbation: no thermal smooth-
ing, no mass ablation 1; thermal smoothing, no mass ablation 2; thermal
smoothing and mass ablation 3; thermal smoothing, mass ablation and the
’’rocket effect’’ 4.
1667Phys. Plasmas, Vol. 7, No. 5, May 2000 Richtmyer-Meshkov-like instabilities and early-time...
Downloaded 06 Aug 2007 to 128.138.96.216. Redistribution subject to AIP license or copyright, see http://pop.aip.org/pop/copyright.jsp
tures driving the perturbations are not small perturbations
all the smallness comes from the time averaging, higher-
order combinations of
Ik(t) could contribute to
mk. The
linear relation can be only validated in the limit of large
bandwidth, and verified by the inverse-square-root relation
between
mkand
. For a finite bandwidth there are con-
tributions of higher orders, which makes establishing the link
between constant phase and ISI simulations difficult. Let us
assume that the Gaussian full width half maximum band-
width
of the driving noise is high enough, so that 5
holds.
Then we can evaluate the Green function from a series
of computer simulations, where a small-amplitude, constant
phase single-mode perturbation
Ik(t)␧⌰(tt) is ap-
plied subsequently at tn
,n0,1,2,...,producing the re-
sponses
mk(t;t). Then, by definition 5,G(t,t)
⫽⫺(1/)(
/
t)
mk(t,t). Interpolating this function to
substitute it into 6, we can predict the ISI imprint at a given
wavelength produced with the given bandwidth.
Figure 6 shows comparison of this prediction with an
actual ISI simulation performed for a 1
4
m laser light, 1 THz
bandwidth, 31012 W/cm2, on a solid plastic target. The
correspondence is quite reasonable, and demonstrates that
our qualitative results remain valid with the laser beam
smoothing taken into account. Improvement will be needed
to remove the small discrepancy at early time.
It is natural to inquire whether we can reduce the laser
imprint. The pendulum model suggests that we have to in-
crease the friction, or gravity, or both, so that the pendulum
swings less under the same drive. Since the effective friction
scales as kVa, and effective gravity due to the ‘‘rocket ef-
fect’’ as kVa
2, where Vais the ablation velocity, the natural
way of decreasing the imprint is to increase Va, which could
be done by decreasing the density
0of the target. Indeed, as
we have demonstrated in Ref. 20 for plastic targets, the peak
imprint amplitude and the time of growth both scale approxi-
mately as
0
1/2 . The imprint, however, could be further re-
duced, if we add some more gravity independent of the
‘‘rocket effect.’’40 This is discussed in Sec. V.
V. SUPPRESSION OF THE RM-LIKE INSTABILITIES
DUE TO DENSITY TAILORING
We have noticed in Sec. III that a collision of two
slightly nonuniform surfaces excites a RM-like perturbation
growth. This kind of growth has been recently observed in
LLNL experiments simulating interaction of a supernovae
ejecta with circumstellar gas.9In these experiments, how-
ever, a relatively slow RM-like growth rapidly evolved into
much faster RT growth. Once the conditions for the bulk
convective instability p
0 were satisfied in the stag-
nated material driving the interface, exponential perturbation
growth followed. This is quite natural—inverted gravity
would drive our pendulum exponentially from its equilib-
rium.
The opposite however is also true—appropriately di-
rected gravity could quench the RM-like instability develop-
ment in colliding fluids, producing oscillations instead of the
linear growth. This effect was identified in an early simula-
tion of Z-pinch implosions with double shells10 and seems to
contribute to the enhanced stability observed in most of the
Z-pinch implosions with nested shells,41–43 including dy-
namic hohlraum experiments.44 To illustrate how this stabi-
lization mechanism works, consider the experiment where an
argon double-puff load on the 4 MA Double Eagle generator
in Maxwell Physics International was imploded in a long-
pulse regime.42 The simulation illustrated by Fig. 7 was per-
formed with the NRL radiation-MHD code PRISM which
stands for plasma radiating imploding source model, see
Refs. 45, 46for realistic initial density profiles in the col-
liding shells. For tutorial purposes, a single-mode mass per-
turbation was introduced in this run at the instant of colli-
sion, to avoid contamination of the flow by the RT
perturbations developing at the outer boundary of the pinch.
At the instant of impact, the inner shell has a higher density
than the outer.
We observe that the contact interface remains virtually
nonperturbed throughout the implosion, while in the outer
part of the pinch the RT instability actively develops. Near
the interface, however, sufficient mass is located where the
density gradient is positive, and the local gravity—positive,
therefore remaining stable up to the stagnation. After the
stagnation Figs. 7d–7e兲兴, the density peaks at the axis and
the pinch is decelerated local gravity is negative. This leads
to oscillation of the whole structure, shifting the density
peaks by a half-period cf. Figs 7eand 7f兲兴 and straight-
ening the outer pinch boundary.
Specifically designing the density profile of a pinch in
such a way that the magnetic field/plasma interface is always
decelerated, we can in principle fully suppress even the RT
instability of implosion.46,47 Full stabilization, however,
comes at the price of having a shock wave present through-
out the implosion. This might be tolerable for a Z pinch used
as a plasma radiation source, but not for a laser target. With
a laser target, however, we might pursue a less ambitious
goal of reducing the imprint by designing the target to pro-
duce additional, favorable acceleration during the shock tran-
sit time, as discussed in Ref. 40. For this, we should make
the target density increase in the direction where the shock
FIG. 6. Perturbation growth for 30
m obtained in an ISI simulation
solid lineand predicted for this case from the simulations imposing a small
constant phase single-mode perturbation.
1668 Phys. Plasmas, Vol. 7, No. 5, May 2000 Velikovich
et al.
Downloaded 06 Aug 2007 to 128.138.96.216. Redistribution subject to AIP license or copyright, see http://pop.aip.org/pop/copyright.jsp
wave propagates, so that the shock will decelerate, and so
will the ablation front. The ‘‘virtual gravity’’ due to inertia
adds to that caused by the rocket effect, thus making our
pendulum stiffer. We illustrate this by some of the results of
Ref. 40.
We compare the mass perturbations imprinted by a 0.2%
constant phase laser nonuniformity, in four planar plastic tar-
gets that have the same mass, 68.2
m g/cm3. Target 1 fea-
tures a parabolic density profile where density increases in
the direction of the laser beam from 0.05 g/cm3to the solid
plastic density 1.07 g/cm3within 120
m, which is followed
bya20
m solid plastic payload. In target 2, the mass con-
centrated in the profile is uniformly distributed over 120
m,
the payload remaining the same; in target 3, the whole mass
is uniformly distributed over the whole thickness, 140
m.
Finally, target 4 is made of solid plastic, and is thinner than
the other three. Here, we deal with premanufactured density
profiles, although one might consider producing them dy-
namically, by irradiating the target surface by the laser
prepulse, and then releasing the pressure.48In Fig. 8 we
compare the rms mass perturbations imprinted into these tar-
gets by three perturbation wavelengths: 15, 30, and 60
m.
Averaging over the three perturbation wavelengths is seen to
smooth out most of the peaks due the oscillations of the
separate modes. For target 1, some oscillatory behavior is
still visible as a result of substantially increased gravity os-
cillations of the three modes, similar to those see in Fig. 5a,
are shown as thin lines.
FIG. 7. Density maps in a PRISM simulation of a double puff implosion on Double Eagle. Figures aecorrespond to times 190, 220, 230, 242, 250, and
270 ns.
FIG. 8. Total rms perturbation growth due to 15, 30 and 60
m perturba-
tions, driven by constant phase 0.2% laser nonuniformity in four compared
targets thick linesand oscillations of the three modes in the tailored den-
sity target 1.
1669Phys. Plasmas, Vol. 7, No. 5, May 2000 Richtmyer-Meshkov-like instabilities and early-time...
Downloaded 06 Aug 2007 to 128.138.96.216. Redistribution subject to AIP license or copyright, see http://pop.aip.org/pop/copyright.jsp
It is not surprising at all that in all cases some extra
gravity in the right direction helps stabilize a pendulum and
makes its swing less sensitive to external drive perturbations.
Whether laser targets making use of this property could be
manufactured or are consistent with high-gain direct drive
fusion remains to be seen.
VI. CONCLUSIONS
We have demonstrated that the RM instability is just the
best-known member of a family that includes numerous non-
classical RM-like instabilities. The main features of the clas-
sical RM instability at the linear and weakly nonlinear re-
gimes are firmly established in the experiments, simulations
and theory.3–5 This level of understanding has yet to be ex-
tended to cover the whole family. Some of the outstanding
basic issues are listed below.
1For the RM-like instability of colliding fluid layers,
the small-amplitude phase is described by a simple applica-
tion of the analytical techniques developed for the classical
RM instability. The same must be true for a weakly nonlin-
ear theory, but this still has to be shown. Much more detailed
numerical studies are needed. An experimental observation
of this instability with colliding laser- or Z-pinch-driven tar-
gets is very difficult, since uniformity of the densities in both
plasma layers is hard to maintain after collision. A clean
relevant experiment might require massive solid plates col-
liding at a relative velocity on the order of 10 km/s.
2There are numerous RM-like instabilities that in-
volve perturbations coming to the contact interface through
perturbed shock/rarefaction waves instability of a planar in-
terface hit by a rippled shock wave, instabilities involving
reshocked and externally driven unstable interfaces, as well
as the whole host of respective MHD problems.For these
instabilities, the accurate compressible theory has yet to be
developed, even for the small amplitude case. Although a
considerable amount of experimental data and numerical
simulation results is already available, more experiments in-
cluding specially designed shock tube experimentsare nec-
essary. Developing an analytical theory of shock interaction
with an interface at various stages of mixing is an issue of
much interest and importance.
3The RM-like instability of ablative acceleration has
been extensively studied numerically and theoretically. A di-
rect experimental validation of these findings is now becom-
ing a focus of interest. Observation of the whole oscillation
cycle of the mass perturbation amplitude during the shock
transit time, although a serious challenge for the experimen-
talists, nevertheless appears to be feasible.
Progress in the theoretical and numerical studies of the
RM-like instabilities of the ablative shock-piston flow is a
major development toward reliable modeling of the mass
nonuniformity produced in a laser target by interaction with
surface roughness and by nonuniformity of laser beams. The
understanding gained could be helpful in designing imprint-
resistant laser targets. Similarly, better understanding of the
RM-like instability in a collision might help in the design of
multiple-shell loads in Z pinches providing improved radia-
tive performance, particularly for long current pulses.
ACKNOWLEDGMENTS
Stimulating discussions with and help of S. E. Bodner,
R. H. Lehmberg, R. P. Drake, B. Fryxell, H. Sze, R. B.
Baksht, H. Azechi, and J. G. Wouchuk are gratefully ac-
knowledged. This work was supported by the Defense Threat
Reduction Agency, U.S. Department of Energy and the Of-
fice of Naval Research.
1R. D. Richtmyer, Commun. Pure Appl. Math. 13,2971960, ‘‘This paper
was originally published while the author was at the Los Alamos Scientific
Laboratory...asReport LA-1914, July, 1954’’ footnote on p. 297.
2E. E. Meshkov, Fluid Dyn. 45, 101 1969;Izv. Akad. Nauk SSSR,
Mekh. Zhidk. Gaza 45, 151 1969兲兴.
3R. L. Holmes, G. Dimonte, B. Fryxell, M. L. Gittings, J. W. Grove, M.
Schneider, D. H. Sharp, A. L. Velikovich, R. P. Weaver, and Q. Zhang, J.
Fluid Mech. 389,551999.
4G. Dimonte, Phys. Plasmas 6, 2009 1999.
5N. Zabusky, Annu. Rev. Fluid Mech. 31, 495 1999.
6G. Dimonte, J. Morrison, S. Hulsey, D. Nelson, S. Weaver, A. Susoeff, R.
Hawke, M. Schneider, J. Batteau, D. Lee, and J. Ticenhurst, Rev. Sci.
Instrum. 67, 302 1996; M. B. Schneider, G. Dimonte, and B. Remington,
Phys. Rev. Lett. 80, 3507 1998. G. Dimonte, D. Nelson, S. Weaver,
Marilyn Schneider, E. Flower-Maudlin, R. Gore, J. R. Baumgardner, and
M. S. Sahota, J. Rheol. 42, 727 1998; G. Dimonte Phys. Plasmas 6, 2009
1999.
7J. W. Jacobs and J. M. Sheeley, Phys. Fluids 8, 405 1996; J. W. Jacobs,
M. A. Jones, and C. E. Niederhaus, Proeedings of the 5th International
Workshop on Compressible Turbulent Mixing, July 18-21, 1995, Univer-
sity at Stony Brook, NY, edited by R. Young, J. Glimm, and B. Boston
University of Stony Brook Press, Stony Brook, NY, 1995,p.195.
8R. P. Drake, J. J. Carroll, III, K. Estabrook, S. G. Glendinning, B. A.
Remington, R. Wallace, and R. McCray, ApJ. Lett. 500, L157 1998;R.
P. Drake, S. G. Glendinning, Kent Estabrook, B. A. Remington, R. Mc-
Cray, R. J. Wallace, L. J. Suter, T. B. Smith, J. J. Carroll, R. A. London,
and E. Liang, Phys. Rev. Lett. 81, 2068 1998; J. Kane, D. Arnett, B. A.
Remington, S. G. Glendinning, G. Bazan, R. P. Drake, B. A. Fryxell, R.
Teyssier, and K. Moore, Phys. Plasmas 6,20651999; R. P. Drake, J.
Geophys. Res. 104, 14505 1999.
9J. Kane, R. P. Drake, and B. A. Remington, ApJ. 511, 335 1999.
10F. L. Cochran, J. Davis, and A. L. Velikovich, Phys. Plasmas 2, 2765
1995.
11R. Ishizaki, K. Nishihara, H. Sakagami, and Y. Ueshima, Phys. Rev. E 53,
R5592 1996.
12A. L. Velikovich, Phys. Fluids 8, 1666 1996.
13J. G. Wouchuk and K. Nishihara, Phys. Plasmas 3, 3761 1996;4,1028
1997.
14M. H. Emery, J. H. Gardner, R. H. Lehmberg, and S. P. Obenschain, Phys.
Fluids B 3, 2640 1991.
15M. Desselberger, M. W. Jones, J. Edwards, M. Dunne, and O. Willi, Phys.
Rev. Lett. 74, 2961 1995.
16T. Endo, K. Shigemori, H. Azechi, A. Nishiguchi, K. Mima, M. Sato, M.
Nakai, S. Nakaji, M. Miyanaga, S. Matsuoka, A. Ando, K. A. Tanaka, and
S. Nakai, Phys. Rev. Lett. 74, 3608 1995.
17H. Azechi, M. Nakai, K. Shigemori, M. Miyanaga, H. Shiraga, H. Nish-
imura, M. Honda, R. Ishizaki, J. G. Wouchuk, H. Takabe, K. Nishihara,
and K. Mima, Phys. Plasmas 4, 4079 1997.
18R. Ishizaki and K. Nishihara, Phys. Rev. Lett. 78, 1920 1997; Phys. Rev.
E58, 3744 1998.
19K. Nishihara, R. Ishizaki, J. G. Wouchuk, Y. Fukuda, and Y. Shimuta,
Phys. Plasmas 5, 1945 1998.
20R. J. Taylor, A. L. Velikovich, J. P. Dahlburg, and J. H. Gardner, Phys.
Rev. Lett. 79, 1861 1997; A. L. Velikovich, J. P. Dahlburg, and J. H.
Gardner, Phys. Rep. 5, 1491 1998.
21V. N. Goncharov, Phys. Rev. Lett. 82,20911999.
22N. Matsui, K. Mima, M. Honda, and A. Nishiguchi, J. Plasma Phys. 61,43
1999.
23G. Dimonte, C. E. Frerking, M. Schneider, and B. Remington, Phys. Plas-
mas 3, 614 1996.
24G. Dimonte and B. Remington, Phys. Rev. Lett. 70, 1806 1993.
25N. C. Freeman, Proc. Phys. Soc., London, Sect. A 228, 341 1955;P.M.
Zaidel, J. Appl. Math. Mech. 24,3161958.
1670 Phys. Plasmas, Vol. 7, No. 5, May 2000 Velikovich
et al.
Downloaded 06 Aug 2007 to 128.138.96.216. Redistribution subject to AIP license or copyright, see http://pop.aip.org/pop/copyright.jsp
26Y. Yang, Q. Zhang, and D. H. Sharp, Phys. Fluids 6, 1856 1994.
27A. L. Velikovich and L. Phillips, Phys. Fluids 8, 1107 1996.
28K. O. Mikaelian, Phys. Rev. Lett. 71, 2903 1993;6356 1994.
29G. Fraley, Phys. Fluids 29,3761986.
30K. A. Meyer and P. J. Blewett, Phys. Fluids 15,7531972.
31M. Vandenboomgaerde, C. Mu
¨gler, and S. Gauthier, Phys. Rev. E 58,
1874 1998.
32J. P. Boris and D. L. Book, J. Comput. Phys. 11,381973; also see
Solution of the continuity equation by the method of flux-corrected trans-
port, Methods in Computational Physics Academic, New York, 1976,
Vol. 16, pp. 85–129.
33C. C. Wu and P. H. Roberts, Geophys. Res. Lett. 26, 655 1999.
34S. E. Bodner, D. G. Colombant, J. H. Gardner, R. H. Lemberg, S. P.
Obenschain, L. Phillips, A. J. Schmitt, J. D. Sethian, R. L. McCrory, W.
Seka, C. P. Verdon, J. P. Knauer, B. B. Afeyan, and H. Powell, Phys.
Plasmas 5, 1901 1998.
35J. Sanz, Phys. Rev. Lett. 73,27001994.
36A. R. Piriz, J. Sanz, and F. L. Iban
˜
ez, Phys. Plasmas 4, 1117 1997.
37S. E. Bodner, Phys. Rev. Lett. 33, 761 1974.
38J. Nuckolls, L. Wood, A. Thiessen, and G. Zimmerman, Nature London
239,1391972.
39A. V. Deniz, T. Lehecka, R. H. Lehmberg, and S. P. Obenchain, Opt.
Commun. 147, 402 1998.
40N. Metzler, A. L. Velikovich, and J. H. Gardner, Phys. Plasmas 6, 3283
1999.
41R. B. Baksht, A. V. Luchinskii, and A. V. Fedyunin, Sov. Phys. Tech.
Phys. 37, 1118 1992; S. A. Sorokin and S. A. Chaikovsky, J. X-Ray Sci.
Technol. 5, 307 1996; A. G. Rousskikh, R. B. Baksht, A. Yu. Labetsky,
A. V. Shishlov, and A. V. Fedyunin, Plasma Phys. Rep. 25,5271999.
42H. Sze, P. Coleman, B. Failor, J. Levine, Y. Song, E. Waisman, F. Co-
chran, A. Velikovich, and J. Davis, Bull. Am. Phys. Soc. 44,531999.
43C. Deeney, M. R. Douglas, R. B. Spielman, T. J. Nash, D. L. Peterson, P.
L’Eplattenier, G. A. Chandler, J. F. Seamen, and K. W. Struve, Phys. Rev.
Lett. 81, 4883 1998.
44T. J. Nash, M. S. Derzon, G. A. Chandler, R. Leeper, D. Fehl, J. Lash, C.
Ruiz, G. Cooper, J. F. Seaman, J. McGurn, S. Lazier, J. Torres, D. Jobe, T.
Gilliland, M. Hurst, R. Mock, P. Ryan, D. Nielsen, J. Armijo, J. McKen-
ney, R. Hawn, D. Hebron, J. J. MacFarlane, D. Peterson, R. Bowers, W.
Matuska, and D. D. Ryutov, Phys. Plasmas 6, 2023 1999.
45F. L. Cochran and J. Davis, Phys. Fluids B 2, 1238 1990; C. Deeney, P.
D. LePell, F. L. Cochran, M. C. Coulter, K. G. Whitney, and J. Davis,
ibid. 5, 992 1993.
46A. L. Velikovich, F. L. Cochran, and J. Davis, Phys. Rev. Lett. 77, 853
1996; A. L. Velikovich, F. L. Cochran, J. Davis, and Y. Chong, Phys.
Plasmas 5, 3377 1998.
47J. H. Hammer, J. L. Eddleman, P. Springer, M. Tabak, A. Toor, K. Wong,
G. B. Zimmerman, C. Deeney, R. Humphreys, T. J. Nash, T. W. L. San-
ford, R. B. Spielman, and J. S. DeGroot, Phys. Plasmas 3, 2063 1996.
48T. J. B. Collins, J. P. Knauer, and S. Skupsky, Bull. Am. Phys. Soc. 43,
1664 1998.
1671Phys. Plasmas, Vol. 7, No. 5, May 2000 Richtmyer-Meshkov-like instabilities and early-time...
Downloaded 06 Aug 2007 to 128.138.96.216. Redistribution subject to AIP license or copyright, see http://pop.aip.org/pop/copyright.jsp
... As they move into the fluid, they leave density and vorticity perturbation behind them. 14,18,[46][47][48] The interface perturbation shows damped oscillations, and when the shock fronts go away from the interface more than a wavelength of the perturbation, 3 the perturbed contact surface reaches a constant velocity, which we call the asymptotic linear growth rate. 14,18,24,46,48,49 The asymptotic linear growth rate can be exactly expressed in the following form: 14,18,24 ...
... 14,18,[46][47][48] The interface perturbation shows damped oscillations, and when the shock fronts go away from the interface more than a wavelength of the perturbation, 3 the perturbed contact surface reaches a constant velocity, which we call the asymptotic linear growth rate. 14,18,24,46,48,49 The asymptotic linear growth rate can be exactly expressed in the following form: 14,18,24 ...
... The terms F 1 and F 2 in Eq. (1) represent the sonic interaction between the rippled shock fronts and the rippled contact surface for t > 0þ. They can be written as integrals of the pressure perturbations along the shock front trajectories 14,18,48 (we can also regard F 1 and F 2 as averaged measures of the vorticity field left by the rippled shock fronts 14 ). The second term in Eq. (1) comes from the vorticity generation by the deformed shock fronts in the bulk fluid. ...
Article
When a planar shock hits a corrugated interface between two fluids, the Richtmyer–Meshkov instability (RMI) occurs. Vortices are generated in bulk behind the transmitted and reflected shocks in RMI. As the shock intensity becomes larger, the stronger bulk vortices are created. The nonlinear evolution of RMI is investigated within the vortex sheet model (VSM), taking the nonlinear interaction between the interface and the vortices into account. The fluid becomes incompressible as the shocks move away from the interface, and VSM can then be applied. The vorticity and position of the bulk vortices obtained from the compressible linear theory [F. Cobos-Campos and J. G. Wouchuk, Phys. Rev. E93, 053111 (2016)] are applied as initial conditions of the bulk point vortices in VSM. The suppression of RMI due to the bulk vortices is observed in the region such that the corrugation amplitude is less than one-tenth of the wavelength, and the reduction of the growth is quantitatively evaluated and compared with the compressible linear theory. In the nonlinear stage, the interaction between the interface and the bulk vortices strongly affects the interfacial shape and the dynamics of bulk vortices, e.g., the creation of a vortex pair is observed. Strong bulk vortices behind the transmitted shock enhance the growth of spike, supplying flow from spike root to its top and mushroom umbrella in the fully nonlinear stage.
... • Laboratory and numerical experiments on hydrodynamic instabilities and opacity measurements in relation to stellar and supernova physics [8,70,71]. ...
... It should be pointed out that previous work on laboratory astrophysics of core-collapsing supernovae was concentrated more on problems of Rayleigh-Taylor and Richtmyer-Meshkov instabilities in the supernova shocks [8,70,71]. More recent work is done on the simulations growth of instabilities in collapse phase, see [73] and references/citations in the latter paper. ...
Article
Radiation–matter interaction depends mainly on the state of matter (its density, temperature, etc.), and also on the radiation spectrum. The opacity of thick plasma also depends on plasma velocity—the Doppler effect shifts atomic lines. For the cases when there are many bound–bound transitions, i.e., the plenty of lines contribute to the opacity, the latter is enhanced when the plasma expands with a nonuniform velocity field. It is known as “expansion opacity” in the literature. Existing models are discrepant and predict diverse results in some cases. Here, we present a rigorous derivation of the effect and show that the effect is available for experimental study at modern laser facilities. The plasma created by a Cu target irradiated with an ∼ 100 J nanosecond laser pulse is rich in lines and has enough expansion velocity so that its opacity is increased in the spectral range ∼ 10 2 − 10 3 eV by the order of magnitude. The possible experimental measurement of the effect is briefly discussed.
... It is important to notice that the complexity order of the sound wave dynamics within the shock fronts is responsible of the above recurrence expressions, and thus, of the mathematical model of the fully-compressible model for the RMI. That analysis is particularly useful as it can be employed to distinguish dominant phenomena among other RMI-like situations, see Ref. [41]. Simpler examples verse on the stability of an isolated shocks, where an explicit finite solution can be reached [39], or for the dynamic between a steady shock and a simple supporting mechanism [36,42,43], for which a closed recurrence expression is achievable. ...
Article
Full-text available
We present a linear stability analysis of the Richtmyer-Meshkov instability that develops when a shock wave reaches a sinusoidally perturbed premixed flame from behind. In the hydrodynamic regime, when acoustic contributions dominate the flame growth rate, the problem is analytically addressed by the direct integration of the sound wave equations at both sides of the flame, which are bounded by the reflected and transmitted shock waves and the flame front that acts as a contact surface in the hydrodynamic limit. The resolution involves: a hyperbolic change of variables to modify the triangular spatio-temporal domain, a transformation in the Laplace variable, the resolution of the functional equations in the frequency domain, and the final inverse Laplace transform. The latter involves a novel resolution method that is proven beneficial for long-time dynamics. Asymptotic analysis is also carried out to describe the early time and late time hydrodynamic response. The nonuniform flow field resulting from distorted oscillating shocks is characterized by acoustic, rotational, and entropic disturbances, each of which exerts a substantial influence on flame dynamics. These disturbances contribute to the intricate interplay of factors shaping the behavior of the flame in response to the nonuniform flow. In particular, the sensitivity of local flame propagation to temperature disturbances is investigated. This work contributes to a deeper understanding of Richtmyer-Meshkov instability dynamics and offers insights into instability reduction through the modulation of temperature disturbances generated by the transmitted shock.
... "Richtmyer-Meshkovlike" instabilities, including "anti-collisions," where such a mechanism becomes relevant, have been discussed in the context of shock-piston flows relevant to the modeling of laser experiment targets. 42 Local vorticity deposition from curved shocks can also significantly affect the late-time growth of single-mode Richtmyer-Meshkov instability, particularly in the case of an initial light-to-heavy gas transition. 30 The use of vortex methods also provides insight into later stages of the instability development not readily accessible using analytical treatments. ...
Article
Two- and three-dimensional simulation results obtained using a new high-order incompressible, variable-density vorticity–streamfunction (VS) method and data from previous ninth-order weighted essentially nonoscillatory (WENO) shock-capturing simulations [M. Latini and O. Schilling, “A comparison of two- and three-dimensional single-mode reshocked Richtmyer-Meshkov instability growth,” Physica D 401, 132201 (2020)] are used to investigate the nonlinear dynamics of single-mode Richtmyer–Meshkov instability using a model of a Mach 1.3 air(acetone)/SF6 shock tube experiment [J. W. Jacobs and V. V. Krivets, “Experiments on the late-time development of single-mode Richtmyer–Meshkov instability,” Phys. Fluids 17, 034105 (2005)]. A comparison of the density fields from both simulations with the experimental images demonstrates very good agreement in the large-scale structure with both methods but differences in the small-scale structure. The WENO method captures the small-scale disordered structure observed in the experiment, while the VS method partially captures such structure and yields a strong rotating core. The perturbation amplitude growth from the simulations generally agrees well with the experiment. The simulation bubble and spike amplitudes agree well at early times. At later times, the WENO bubble amplitude is smaller than the VS amplitude and vice versa for the spike amplitude. The predictions of nonlinear single-mode instability growth models are shown to agree with the simulation amplitudes at early-to-intermediate times but underpredict the amplitudes at later times in the nonlinear regime. Visualizations of the mass fraction and enstrophy isosurfaces, velocity and vorticity fields, and baroclinic vorticity production and vortex stretching terms from the three-dimensional simulations indicate that, with the exception of the small-scale structure within the rollups, the VS and WENO results are in good agreement.
... The consistent results in RMI and DDM demonstrate that the present multi-mode RMI with a weak incident shock is mainly driven by the synergistic effect of the density 925 A39-11 difference across the perturbed interface and the PBV (see § § 4.2 and 4.3 in Peng et al. 2021), whereas the pressure perturbation from distorted waves only plays a minor role in reducing the growth rate (e.g. Velikovich et al. 2000) and their effect on the PBV distribution is neglected. Therefore, the DDM can serve as a simplified model for the present multi-mode RMI. ...
Article
Full-text available
We investigate the effect of the secondary baroclinic vorticity (SBV) on the energy cascade in the mixing induced by the multi-mode Richtmyer–Meshkov instability (RMI). With the aid of vorticity-based simplified models and the vortex-surface field, we find that the effect of the SBV peaks at a critical time when the vortex reconnection widely occurs in the mixing zone. Before the critical time, spikes and bubbles evolve almost independently, and we demonstrate that the variation of the kinetic energy spectrum induced by the SBV has the $-1$ scaling law at intermediate wavenumbers using the model of vortex rings. This SBV effect causes the slope of the total energy spectrum at intermediate wavenumbers to evolve towards $-3/2$ at the critical time. Subsequently, the SBV effect diminishes and the energy spectrum decays to the $-5/3$ law. Inspired by the vortex dynamics, we develop a model for estimating the mixing width and validate the model using numerical simulations of the multi-mode RMI with various modes of initial perturbations. This model captures the nonlinear growth of the mixing width before the self-similar growth stage.
... where k = 2π λ , ∆u is the velocity change due to the shock, A is the Atwood number, and a 0 is the initial perturbation amplitude. Other work [190] has indicated that some restrictions to this model may apply. However, this model is a good first approximation for the growth in this phase.The vorticity distribution and the circulation have been shown to be [37] ω(x) = −2ȧ 0 sin(kx) and Γ = −4 kȧ 0 . ...
Thesis
Developing a highly accurate numerical framework to study multiphase mixing in high speed flows containing shear layers, shocks, and strong accelerations is critical to many scientific and engineering endeavors. These flows occur across a wide range of scales: from tiny bubbles in human tissue to massive stars collapsing. The lack of understanding of these flows has impeded the success of many engineering applications, our comprehension of astrophysical and planetary formation processes, and the development of biomedical technologies. Controlling mixing between different fluids is central to achieving fusion energy, where mixing is undesirable, and supersonic combustion, where enhanced mixing is important. Iron, found throughout the universe and a necessary component for life, is dispersed through the mixing processes of a dying star. Non-invasive treatments using ultrasound to induce bubble collapse in tissue are being developed to destroy tumors or deliver genes to specific cells. Laboratory experiments of these flows are challenging because the initial conditions and material properties are difficult to control, modern diagnostics are unable to resolve the flow dynamics and conditions, and experiments of these flows are expensive. Numerical simulations can circumvent these difficulties and, therefore, have become a necessary component of any scientific challenge. Advances in the three fields of numerical methods, high performance computing, and multiphase flow modeling are presented: (i) novel numerical methods to capture accurately the multiphase nature of the problem; (ii) modern high performance computing paradigms to resolve the disparate time and length scales of the physical processes; (iii) new insights and models of the dynamics of multiphase flows, including mixing through hydrodynamic instabilities. These studies have direct applications to engineering and biomedical fields such as fuel injection problems, plasma deposition, cancer treatments, and turbomachinery.
Article
Full-text available
This paper characterizes the refraction of a triple-shock configuration at planar fast–slow `gas interfaces. The primary objective is to reveal the wave configurations and elucidate the mechanisms governing circulation deposition and velocity perturbation on the interface caused by triple-shock refraction. The incident triple-shock configuration is generated by diffracting a planar shock around a rigid cylinder, and four interfaces with various Zr(i.e. acoustic impedance ratio across the interface) are considered. An analytical modeldescribing the triple-shock refraction is developed, which accurately predicts both the wave configurations as well as circulation deposition and velocity perturbation. Depending on Zr, three distinct patterns of transmitted waves can be anticipated: a triple-shockconfiguration; a four-shock configuration; a four-wave configuration. The underlying mechanism for the formation of these wave configurations is elucidated through shock polar analysis. A novel physical insight into the contribution of triple-shock refraction to the interface perturbation growth is provided. The results indicate that the reflected shock in an incident triple-shock configuration makes significant negative contribution to both circulation deposition and velocity perturbation. This investigation elucidates the underlying mechanism responsible for the relatively insignificant contribution of baroclinic circulation to the Richtmyer–Meshkov-like instability induced by a non-uniform shock, and provides an explanation for the decrease in growth rate of interface perturbation amplitude with increasing Atwood number.
Article
Full-text available
Impact-induced mixing between bolide and target is fundamental to the geochemical evolution of a growing planet, yet aside from local mixing due to jetting – associated with large angles of incidence between impacting surfaces – mixing during planetary impacts is poorly understood. Here we describe a dynamic instability of the surface between impacting materials, showing that a region of mixing grows between two media having even minimal initial topography. This additional cause of impact-induced mixing is related to Richtmyer-Meshkov instability (RMI), and results from pressure perturbations amplified by shock-wave refraction through the corrugated interface between impactor and target. However, unlike RMI, this new impact-induced instability appears even if the bodies are made of the same material. Hydrocode simulations illustrate the growth of this mixing zone for planetary impacts, and predict results suitable for experimental validation in the laboratory. This form of impact mixing may be relevant to the formation of stony-iron and other meteorites.
Article
A unified analytical approach to study the effects of elasticity, viscosity, and magnetic fields on the Richtmyer–Meshkov (RM) instability by using the impulsively accelerated model is presented. This model clarifies the discontinuity in the oscillation periods and yields the asymptotic decaying rate in elastic solids. It reveals that the complex eigenvalues produce better results compared with the numerical simulations for RM instability in viscous fluids and resolves the standing controversy between the analytical theory and numerical simulations at a vacuum/fluid interface. At last, it easily retrieves the results when the normal or tangential magnetic field is present. Those good agreements, between numerical simulations and theoretical analysis, would enable the model to be valuable in more complex situations such as in the elastic–plastic slabs with or without the presence of magnetic fields, as well as in the nonlinear regime.
Article
Full-text available
Experiments on the implosion of both single and double argon gas puffs were carried out in the IMRI-4 (tfr = 1.1 μs and Im = 350 kA) and GIT-4 (tfr = 0.12 μs and Im = 1.7 MA) current generators. Three different sources of gas-puff preionization were used: (i) UV spark radiation, (ii) a cylindrical magnetron discharge in the crossed E × B fields, and (iii) a planar magnetron discharge in the crossed E × B fields. With the use of an optical streak camera, for all three methods of preionization, conditions are found under which the stratification of the current sheet occurs during the implosion of the gas puff. Experimental data show that the dynamics of the gas-puff implosion depends on the preionization method. The use of magnetron preionization results in suppression of the instabilities excited during the gas-puff implosion and allows the obtaining of stable plasma parameters in the final stage of implosion. It is shown that, in a single gas puff, a plasma column with an initial radius of 30 mm can be uniformly compressed to a stable plasma cylinder with a radius less than 2 mm.
Article
Perturbations of small-amplitude [eta] at the interface between two fluids grow linearly in time after the passage of a shock. According to Richtmyer's prescription, the growth rate is proportional to the Atwood number after the interface has been shocked, [dot [eta]][similar to][ital A][sub after]. The focus is on highly compressible fluids starting with [ital A][sub before][ge]0. By carrying out two-dimensional numerical simulations, several exceptions to Richtmyer's prescription are found, in particular, when [ital A][sub after][le]0. Neither the expected freeze-out when [ital A][sub after]=0 nor the phase reversal when [ital A][sub after][lt]0 is observed. The results, however, are in agreement with Fraley's analysis, which is compared and contrasted with Richtmyer's prescription. Previous calculations and experiments on the Richtmyer--Meshkov instability are analyzed, it is explained why they have not detected the failure of Richtmyer's prescription, and several new numerical and physical experiments are proposed.
Article
Nested-wire arrays on the 20-MA Z-pinch accelerator have produced x-ray powers up to 280+/-40 TW (a 40% increase over a single array) and an x-ray pulse width of 4 ns. The short x-ray pulse widths are associated with the formation of tight (1-mm-diameter), uniform pinches at stagnation. Two-dimensional radiation magnetohydrodynamic calculations suggest that the inner array mitigates the growth of implosion instabilities thereby leading to smaller diameter pinches that radiate at higher power than single-wire arrays.
Article
Nonuniform laser ablation caused by nonuniform laser irradiation and initial target roughness induces ripples of shock and ablation surfaces in laser implosion. Hydrodynamic perturbation growth before the shock breakout is investigated by using an analytical model and two-dimensional simulations. The model agrees well with simulation and experimental results. Areal mass density perturbations and growth rate of the Richtmyer–Meshkov instability are estimated for an ignition target. The thermal smoothing in the ablation layer is also studied for perturbations with a wavelength longer than the layer thickness. A large increase of temperature and density perturbations is shown instead of the smoothing for such a wavelength.
Article
Recent experiments by Meshkov have demonstrated that a shock‐accelerated perturbed surface separating two gases of different densities is unstable for shocks traveling from the lighter gas to the heavier gas and vice versa. Using a two‐dimensional Lagrangian hydrodynamic code, the authors show qualitative agreement with Meshkov's experiments. Experiment, numerical calculations, and linear analysis are compared. Good agreement was obtained between numerical calculations and linear analysis.
Article
Z-pinch implosions driven by the SATURN device [D. D. Bloomquist {ital et} {ital al}., {ital Proceedings} {ital of} {ital the} 6{ital th} {ital Institute} {ital of} {ital Electrical} {ital and} {ital Electronics} {ital Engineers} ({ital IEEE}) {ital Pulsed} {ital Power} {ital Conference}, Arlington, VA, edited by P. J. Turchi and B. H. Bernstein (IEEE, New York, 1987), p. 310] at Sandia National Laboratory are modeled with a two-dimensional radiation magnetohydrodynamic (MHD) code, showing strong growth of the magneto-Rayleigh{endash}Taylor (MRT) instability. Modeling of the linear and nonlinear development of MRT modes predicts growth of bubble-spike structures that increase the time span of stagnation and the resulting x-ray pulse width. Radiation is important in the pinch dynamics, keeping the sheath relatively cool during the run-in and releasing most of the stagnation energy. The calculations give x-ray pulse widths and magnitudes in reasonable agreement with experiments, but predict a radiating region that is too dense and radially localized at stagnation. We also consider peaked initial density profiles with constant imploding sheath velocity that should reduce MRT instability and improve performance. Krypton simulations show an output x-ray power {approx_gt}80 TW for the peaked profile. {copyright} {ital 1996 American Institute of Physics.}
Article
An analytic model for the Richtmyer‐Meshkov instability in the linear regime is presented. Approximate formulas for the interface asymptotic velocity are obtained for weak incident shocks. When a rarefaction is reflected, a comparison with recent experiments is shown. The phase reversal is also discussed. The areal mass density perturbations are studied and they are found to grow asymptotically linearly with time. For strong shocks there is a shift due to the entropy disturbances behind the shock(s). The impulsive model is modified to consider the different accelerations exerted by the reflected and transmitted fronts. It agrees with the compressible solution in the weak shock limit.
Article
We have studied the spatial fluence nonuniformities in single beams from the Nike KrF laser smoothed by the Induced Spatial Incoherence (ISI) technique. This study compares time-integrated CCD camera measurements with a numerical model of ISI speckle used in the NRL FAST2D hydrocode, and with analytic calculations that examine the ensemble-average behavior of that model. The results verify the model and calculations, showing good agreement of both the spatial mode spectra and the total nonuniformity versus pulsewidth and bandwidth.
Article
The [ital K]-shell x ray yields from argon gas puff Z pinches are observed to increase from 3.5[plus minus]1.0 kJ to 13[plus minus]1.0 kJ when inwardly tilted nozzles are used on a 6 TW, 4 MA generator. This increase is associated with the elimination of the zipper effect and the achievement of higher density plasmas, as confirmed by x-ray diagnostics. Two-dimensional (2-D) magnetohydrodynamic modeling of the gas puff implosions indicates that the collapsing shell reaches a higher density and smaller diameter when the axial zipper is eliminated. The calculations show that axial mass flow in the zippering cases result in nonhollow collapses which limit the final pinch radii and densities. The calculations also indicate that narrower nozzle exits contribute to increasing the assembled plasma density.