ArticlePDF Available

Energy transfer in Rayleigh-Taylor instability

Authors:

Abstract and Figures

The spatial structure and energy budget for Rayleigh-Taylor instability are examined using results from a 512 x 512 x 2040 point direct numerical simulation. The outer-scale Reynolds number of the flow follows a rough t(3) power law and reaches a final value of about 5500. Taylor microscales and Reynolds numbers are plotted to characterize anisotropy in the flow and document progress towards the mixing transition. A mixing parameter is defined which characterizes the relative rates of entrainment and mixing in the flow. The spectrum of each term in the kinetic energy equation is plotted, at regular time intervals, as a function of the inhomogeneous direction and the two-dimensional wave number for the homogeneous directions. The energy spectrum manifests the beginning of an inertial range by the latter stages of the simulation. The production and dissipation spectra become increasingly opposite and separate in wave space as the flow evolves. The transfer spectrum depends strongly on the inhomogeneous direction, with the net transfer being from large to small scales. Energy transfer at the bubble/spike fronts is strictly positive. Extensive cancellation occurs between the pressure and advection terms. The dilatation term produces negligible energy transfer, but its overall effect is to move energy from high to low density regions.
Content may be subject to copyright.
Energy transfer in Rayleigh-Taylor instability
Andrew W. Cook and Ye Zhou
Lawrence Livermore National Laboratory, University of California, Livermore, California 94551
Received 1 November 2001; revised manuscript received 8 February 2002; published 30 August 2002
The spatial structure and energy budget for Rayleigh-Taylor instability are examined using results from a
5125122040 point direct numerical simulation. The outer-scale Reynolds number of the flow follows a
rough t3power law and reaches a final value of about 5500. Taylor microscales and Reynolds numbers are
plotted to characterize anisotropy in the flow and document progress towards the mixing transition. A mixing
parameter is defined which characterizes the relative rates of entrainment and mixing in the flow. The spectrum
of each term in the kinetic energy equation is plotted, at regular time intervals, as a function of the inhomo-
geneous direction and the two-dimensional wave number for the homogeneous directions. The energy spectrum
manifests the beginning of an inertial range by the latter stages of the simulation. The production and dissi-
pation spectra become increasingly opposite and separate in wave space as the flow evolves. The transfer
spectrum depends strongly on the inhomogeneous direction, with the net transfer being from large to small
scales. Energy transfer at the bubble/spike fronts is strictly positive. Extensive cancellation occurs between the
pressure and advection terms. The dilatation term produces negligible energy transfer, but its overall effect is
to move energy from high to low density regions.
DOI: 10.1103/PhysRevE.66.026312 PACS numbers: 47.27.Ak, 47.27.Cn
I. INTRODUCTION
Rayleigh-Taylor instability RTIoccurs at the interface
between two fluids of different densities whenever the
heavier fluid is decelerated by the lighter fluid 1–4; i.e., if
density and pressure gradients are in opposite directions,
then vorticity, deposited at the interface through baroclinic
torque, will cause the fluids to interpenetrate and mix.
Richtmyer-Meshkov instability RMIcorresponds to the
case of impulsive acceleration of an interface 5,6, e.g.,
shock passage, and is sometimes considered a special case of
RTI with time-dependent acceleration. RTI presents a seri-
ous design challenge for inertial confinement fusion ICF
capsules, where high density shells are decelerated by low
density fuel. Depending on the acceleration history and the
ratio of shell radius to thickness, RTI may lead to break up of
the shell prior to ignition and/or significant mixing of the
fuel with the plastic ablator 7. RTI also plays a prominent
role in supernovae, where ejecta are decelerated by circum-
stellar matter 8,9. Furthermore, mixing from RTI alters
thermonuclear burn in supernovae in such a manner as to
affect the rates of formation of heavy elements; hence, the
relative abundance of elements in the universe, and the cor-
responding potential for life, are directly related to astro-
physical RTI mixing.
Most RTI research thus far has focused on predicting the
rate of growth of the turbulent mixing zone 10–15. Mixing
zone amplitudes are routinely measured in high-energy laser
experiments conducted at very high Reynolds number 16.
In its early stages, RTI growth is characterized by ‘‘spikes’
of heavy fluid penetrating into light material and ‘‘bubbles’
of light fluid rising into heavy material. In the strongly non-
linear stages, the bubbles and spikes merge to form larger
structures. If the only imposed length scale is from a constant
acceleration, then the mixing layer will grow quadratically in
time; e.g., a growth constant ‘‘
’’ can be defined and mea-
sured 17,18,10–13. However, if long wavelength perturba-
tions compared to domain sizeare present, the scaling
analysis is more complicated and growth may not be qua-
dratic 19.
The range of scales participating in RTI dynamics con-
tinually grows as the flow evolves. Kelvin-Helmholtz insta-
bilities, occurring along the sides of the interpenetrating fin-
gers, along with vortex stretching and bending motions,
serve to broaden the energy spectrum. Eventually the flow
may become fully turbulent, while still remaining highly an-
isotropic. A complete description of the flow field requires
resolution down to the Kolmogorov scale 20,21. Due to
limitations of diagnostics, laser experiments performed thus
far have not yielded much information on the internal struc-
ture of the mixing region. Larger-scale experiments 22–24
have provided some information on mixing zone structure,
but lack the resolution necessary for a close investigation of
the energy budget.
Over the past three decades, direct numerical simulation
DNShas emerged as an accepted surrogate for experiment
when detailed information, not readily measured in the labo-
ratory, is needed. DNS is restricted to low Reynolds number
flows, due to the limited range of wave numbers that can be
supported on a computational mesh. Nevertheless, it has
proven capable of following the three phases of turbulent
mixing identified by Eckart 25, i.e., entrainment, stirring,
and molecular mixing. It also provides a complete, three-
dimensional, time-dependent description of the flow field.
DNS data of RTI flow can be used to test and/or tune models
for the overall growth of the mixing region, and for devel-
oping subgrid-scale SGSmodels for large eddy simulation.
The latter is intimately connected with energy transfer to and
from unresolved scales. The primary goal of this paper is to
gain insight into the energy transfer processes in order to
guide future SGS model development.
The outline of this paper is as follows. In Sec. II, the
PHYSICAL REVIEW E 66, 026312 2002
1063-651X/2002/662/02631212/$20.00 ©2002 The American Physical Society66 026312-1
governing equations and solution technique of the DNS are
described. In Sec. III, flow visualizations are presented,
along with statistical measures of flow structure, such as
growth rates and Reynolds numbers. In Sec. IV, a variable-
density formulation is proposed for energy transfer analyses
previously performed for isotropic 26–31, anisotropic 32,
and wall-bounded 33flows. The analysis is carried out in
Sec. V, where the spectrum of each term in the kinetic energy
equation is computed from the DNS data. Finally, conclu-
sions are given in Sec. VI.
II. DIRECT NUMERICAL SIMULATION
A. Governing equations
The conservation laws governing the flow of two incom-
pressible fluids in a gravitational field with no surface ten-
sion are
⳵␳
Yl
t
⳵␳
ujYl
xj
xj
D
Yl
xj
l1,2,1
⳵␳
ui
t
⳵␳
uiuj
xj
⫽⫺
p
xi
ij
xj
gi,2
where
ij2
Sij1
3
ij
uk
xk
,
Sij1
2
ui
xj
uj
xi
.
Here
is the mixture density, Ylis the mass fraction of
species l,uiis the mass-averaged mixture velocity, pis the
pressure, Dis the Fickian diffusivity,
is the dynamic vis-
cosity, and gi(0,0,g) is the acceleration. The mass frac-
tions satisfy
Y1x,tY2x,t13
and, defining
1and
2to be the constant densities of the
light and heavy fluids, respectively, the specific volume sat-
isfies
1
x,tY1x,t
1
Y2x,t
2.4
Equations 1,3, and 4can be used to derive the follow-
ing divergence relation for miscible fluids 34:
uj
xj
⫽⫺
xj
D
⳵␳
xj
.5
Hence, for incompressible mixing, a convenient equation for
is
⳵␳
tuj
⳵␳
xj
xj
D
⳵␳
xj
.6
B. Solution technique
The equations were solved in nondimensional form, with
length measured in units of box width L, time measured in
units of
L/g, and density measured in units of
1. The
diffusivity was set to D
/
1, with the viscosity being
5124/3 in the units just described. The numerical
scheme for solving the governing equations is described in
detail in 19. In summary, the code computes xand yde-
rivatives spectrally via fast Fourier transform. The zderiva-
tives are computed with an eighth-order compact scheme
35. Periodic boundary conditions are applied in xand y,
with no-slip walls imposed in z. The zgrid spacing is set to
8
13 times the grid spacing in xand y, in order to account for
the difference in resolving power between the spectral and
compact methods. Time advancement is accomplished via a
pressure-projection algorithm with third-order, Adams-
Bashforth-Moulton integration.
The simulation was performed on a computational mesh
with 5125122040 grid points in x,y, and zdirections,
respectively. The bottom 17
32of the domain was initialized
with
11 and the top 15
32portion with
23. The
density interface between the fluids was specified in the same
manner as in 19, i.e., as an error function in zfive grid
points thickwith isotropic perturbations in xand y. The
perturbed interface was initially located at the z0 plane.
Diffusion velocities were initialized as in 19in order to
satisfy the divergence relation 5.
The spectral/compact algorithm was chosen to ensure that
numerical dissipation did not enter into the calculation. No
filtering or artificial diffusion of any kind was applied in the
simulation, i.e., the viscous and diffusive terms in Eqs. 2
FIG. 1. ColorInitial density perturbations on z0 plane.
ANDREW W. COOK AND YE ZHOU PHYSICAL REVIEW E 66, 026312 2002
026312-2
and 6were solely responsible for energy dissipation. Con-
sequently, the simulation eventually became unstable for
Re5500once significant energy reached the Nyquist wave
number. A prominent symptom of under-resolution and con-
sequent aliasing errorsbeyond Re5500 is a curling up of
the energy spectrum at the highest wave numbers. The data
presented herein were selected at times prior to those where
significant aliasing errors occurred.
FIG. 2. ColorSnapshots of density field from DNS of Rayleigh-Taylor instability. Images, proceeding from upper left to lower right,
were taken at t1, 2, 3, 4, 5, and 6. The heavy fluid is red (
3), the light fluid is blue (
1), and the mixed fluid is green (
2).
ENERGY TRANSFER IN RAYLEIGH-TAYLOR INSTABILITY PHYSICAL REVIEW E 66, 026312 2002
026312-3
III. FLOW STRUCTURE
A. Visualization
The initial perturbations are depicted in Fig. 1, which pro-
vides a top-down view of density on the z0 plane at t
0. The flow was seeded with fine scale perturbations to
minimize the influence of the periodic boundary conditions.
Figure 2 displays a time sequence of the density field. The
images, taken at unit time intervals, illustrate the evolution of
the fluid interface defined as the
2 isosurface. The early
evolution is weakly nonlinear and is characterized by the
formation of bubbles of light fluid rising upwards and spikes
of heavy fluid penetrating downward. Later on, the bubbles
and spikes begin to merge and the flow becomes strongly
nonlinear.
The kinetic energy,
uiui/2, on the back, bottom, and
side planes of the flow domain is shown in Fig. 3 for t3, 4,
5, and 6. It grows rapidly and appears to be fairly evenly
distributed across the mixing layer. The gravitational poten-
tial provides the source for kinetic energy production. Figure
4 displays the kinetic energy on the z0 plane at the same
times. Like the density field, the kinetic energy is homoge-
neous and isotropic in xand y. Also, large values of
appear
to become concentrated in localized regions of the flow.
The source of vorticity and consequently energy genera-
tionis the baroclinic torque term in the vorticity equation,
which is nonzero only in the mixing region. The magnitude
of vorticity,
u
, and the magnitude of baroclinic torque,
p
/
,att6, are plotted in Fig. 5. A few spikelike
and bubblelike structures are discernible in the fields; how-
FIG. 3. ColorKinetic energy
on side
boundary planes of DNS domain. Images were
taken at t3, upper left,4upper right,5
lower left, and 6 lower right.
ANDREW W. COOK AND YE ZHOU PHYSICAL REVIEW E 66, 026312 2002
026312-4
ever, the fields are quite chaotic. Early on, when the bubbles
and spikes are growing more or less independently, and
hence are readily identified, vorticity and consequently en-
ergyis primarily generated along the sides of the structures.
However, by late time, the flow has become weakly turbulent
and vorticity generation occurs throughout the mixing zone.
The complexity of the baroclinic torque field increases as
centrifugal forces pull pressure gradients out of alignment
with the gravity vector.
B. Statistics
The penetration lengths of the bubbles and spikes, hb(t)
and hs(t), respectively, are defined as in 19, i.e., by aver-
aging the heavy fluid mole fraction X(
1)/(
2
1)
in xand y, and measuring the distance from z0 for which
X
xy0.99 and
X
xy0.01
具典
xy denoting horizontal aver-
age. The bubble and spike penetrations are plotted in Fig. 6.
The change in slope around t2 occurs as modal growth
overtakes diffusive growth.
The outer-scale Reynolds number is plotted, versus time,
in Fig. 7. The Reynolds number is based on the vertical
extent of the mixing region, hhbhs, and its rate of
growth, h
˙, i.e.,
Re
1
2h
˙h
2
.7
The Reynolds number grows roughly like t3, which is ex-
pected if ht2and h
˙t.Reasons why hmay depart from
quadratic growth are discussed in 19.The terminal outer-
scale Reynolds number of 5500 is about a third of the Rey-
nolds number (1–2)104] suggested by Dimotakis 36as
the critical value for reaching the mixing transition and
achieving fully developed turbulence. It should be noted,
however, that the Reynolds number for mixing transition has
yet to be documented for RTI flow. If data above the mixing
transition could be obtained for RTI flow, it would be inter-
FIG. 4. ColorKinetic energy
on z0 plane at t3, upper left,4upper right,5lower left, and 6 lower right.
ENERGY TRANSFER IN RAYLEIGH-TAYLOR INSTABILITY PHYSICAL REVIEW E 66, 026312 2002
026312-5
esting to see if the ‘‘hot spots’ in kinetic energy would per-
sist or if the energy would become more evenly distributed.
It is possible to define Taylor microscales and Reynolds
numbers for this flow in a manner that accommodates the
anisotropic forcing. A microscale in the idirection can be
defined as 37,38
i
ui
2
xy
ui/
xi2
xy
1/2
no sum on i,8
with statistics computed in the (z0) plane. With statistical
isotropy in the (z0) plane, the xand ymicroscales are very
close and can be averaged to define a single horizontal mi-
croscale,
xyxy
2.9
Figure 8 depicts the temporal growth of the vertical and
horizontal Taylor microscales in the (z0) plane. The ver-
tical and horizontal scales both grow as the bubbles increase
in size, broadening the velocity correlation functions. The
difference between the vertical and horizontal scales gives a
FIG. 5. ColorMagnitude of vorticity leftand magnitude of baroclinic torque rightat t6.
FIG. 6. Amplitude of bubbles (hb) and spikes (hs) in the mixing
region. The total width of the layer is hhbhs.FIG. 7. Outer-scale Reynolds number, based on extent and rate
of growth of mixing region.
ANDREW W. COOK AND YE ZHOU PHYSICAL REVIEW E 66, 026312 2002
026312-6
direct measure of anisotropy in the flow. The ratio z/xy
starts out near unity, increases during the diffusive growth
stage to a maximum value of about 3.7, and then appears to
asymptote to a value around 1.4 in the far nonlinear regime.
Figure 9 depicts the temporal evolution of the horizontal
and vertical Taylor Reynolds numbers on the (z0) plane.
These are defined as
Re,i
xyi
ui
2
xy1/2
no sum on i,10
again, with spatial averages computed in the (z0) plane.
As with the microscales, horizontal isotropy permits a hori-
zontal Taylor Reynolds number to be defined as the average
of Re,xand Re,y, i.e.,
Re,xyRe,xRe,y
2.11
The anisotropy in microscales Fig. 8is also manifest in the
Taylor Reynolds numbers. A Taylor Reynolds number of
roughly 100 is required to cross the mixing transition 36.
Judging from the proximity of the outer-scale Reynolds num-
ber to the critical value of 1000020000, it appears that the
transition criterion would probably apply best to the horizon-
tal Taylor Reynolds number, rather than the vertical.
In order to quantify the degree of mixing within the layer,
a parameter, analogous to the Youngs 13‘molecular mix-
ing fraction,’’ 共⌰兲 is defined. Assuming a passive, equilib-
rium chemical reaction between fluids, the chemical product
is
Xp
X/Xsif XXs
1X/1Xsif XXs,12
FIG. 8. Vertical labeled zand horizontal labeled xyTaylor
microscales on the z0 plane.
FIG. 9. Vertical zlabeland horizontal xy labelTaylor Rey-
nolds numbers on the z0 plane.
FIG. 10. Mixing parameter, indicating the ratio of actual chemi-
cal product for a hypothetical infinite-rate reactionto the product
that would be formed if the fluid inside the mixing region were
completely mixed no xy variation.
FIG. 11. Evolution of two-dimensional energy spectrum Eat
z0.
ENERGY TRANSFER IN RAYLEIGH-TAYLOR INSTABILITY PHYSICAL REVIEW E 66, 026312 2002
026312-7
where Xsis the heavy-fluidmole fraction for a stoichio-
metric mixture, here taken as X1
2. A mixing parameter, ,
is defined as the ratio of mixed to entrained fluid, i.e.,
H
XpX
xydz
HXp
X
xydz,13
where His the height of the flow domain. Thus, 1 indi-
cates fluids are completely mixed within the mixing zone;
whereas, 0 corresponds to fully segregated fluids im-
miscible case. Note that with no perturbations (X
X
xy),
for an increasingly sharp interface, the numerator and de-
nominator go to zero at the same rate; hence, 1 for a
Heaviside function. The mixing parameter is plotted in
Fig. 10 as a function of time. Initially, the layer is diffuse
with small amplitude perturbations; hence, starts out near
unity. As the perturbations grow, they entrain fluid at a rate
proportional to their wavelengths longer wavelengths result
in bigger ‘‘gulps’ of pure fluid. At early times, the rate of
entrainment exceeds the rate of mixing and decreases.
This is a consequence of the fact that, early on, the surface
area across which the fluids can diffuse is relatively small.
However, later on the interface begins to wrinkle due to
baroclinic vorticity mushroom capsand Kelvin-Helmholtz
instabilities in the shearing regions along the mushroom
necks; the interfacial surface area then rapidly increases, the
mixing rate overtakes the entrainment rate, and the curve
reverses direction. The curve appears to asymptote to a value
somewhere around 0.8; however, mixing and entrainment
rates have not come into balance within the time span of the
simulation. Furthermore, this curve is likely to rise after the
mixing transition occurs.
IV. ENERGY BUDGET
In order to extend the methodology of constant-density
energetics to the variable-density case, a new variable is in-
troduced, i.e.,
vi
1/2ui,14
such that, the kinetic energy may be written as
vivi/2.
This variable has been used for similar purposes by various
authors 39–41. The left-hand side of the Navier-Stokes
equation can then be written as
⳵␳
ui
t
⳵␳
uiuj
xj
1/2
vi
t1
2
1/2vi
⳵␳
t
1/2
viuj
xj
1
2
1/2viuj
⳵␳
xj
1/2
vi
t
viuj
xj
1
2vi
uk
xk
,
such that the transport equation for vibecomes
vi
t
1/2gi
viuj
xj
1/2
p
xi
1
2vi
uk
xk
2
1/2
xj
Sij1
3
ij
uk
xk
.15
For simplicity, Eq. 15is rewritten as
vi
tFiNiDi,16
where the buoyant forcing is
Fi
1/2gi,17
the nonlinear quadratic, pressure, and dilatationcontribu-
tion is
Ni⫽⫺
viuj
xj
1/2
p
xi
1
2vi
uk
xk,18
and the viscous diffusion is
Di2
1/2
xj
Sij1
3
ij
uk
xk
.19
FIG. 12. ColorEnergy spectrum Eversus zverticaland
log10(k)horizontalat t3top left,4top right,5lower left,
and 6 lower right. Blue0 and red5104.
ANDREW W. COOK AND YE ZHOU PHYSICAL REVIEW E 66, 026312 2002
026312-8
Now let v
ˆi,F
ˆi,N
ˆi, and D
ˆibe horizontal Fourier transforms
of vi,Fi,Ni, and Di, respectively; e.g., v
ˆi(k,z,t)
Fxy
vi(x,t)
, where k(kx,ky) is the horizontal wave
vector. If v
ˆi
*(k,z,t) denotes the complex conjugate of
v
ˆi(k,z,t), then multiplying the transform of Eq. 16, and its
conjugate, by v
ˆi
*and v
ˆi, respectively, and adding the equa-
tions together yields
v
ˆi
*v
ˆi
tv
ˆi
*F
ˆiv
ˆiF
ˆi
*v
ˆi
*N
ˆiv
ˆiN
ˆi
*v
ˆi
*D
ˆiv
ˆiD
ˆi
*.
20
Integrating Eq. 20over Fourier annuli of radius k
kx
2ky
2leads to the energy budget equation
tEk,z,tk,z,tTk,z,tEk,z,t,21
where the kinetic energy is
E1
2
v
ˆi
*v
ˆid
,22
the production from gravity is
1
2
v
ˆi
*F
ˆiv
ˆiF
ˆi
*d
,23
the nonlinear transfer is
T1
2
v
ˆi
*N
ˆiv
ˆiN
ˆi
*d
,24
and the viscous dissipation is
E1
2
v
ˆi
*D
ˆiv
ˆiD
ˆi
*d
,25
with d
being a differential element of a wave space annu-
lus.
There are three contributions to the nonlinear energy
transfer. The first is from the quadratic term, which is respon-
sible for passive-vector advection, while the second and third
contributions are from pressure and dilatation effects. In or-
der to ascertain the relative importance of each process, the
total nonlinear transfer is subdivided into each individual
component, i.e.,
Tk,z,tTqk,z,tTpk,z,tTdk,z,t,26
where
FIG. 13. ColorProduction 共⌸: left, transfer T: middle, and
dissipation E: rightspectra versus zverticaland log10(k)hori-
zontalat t3topand 4 bottom. Blue⫽⫺2104, green0,
and red2104.
FIG. 14. ColorProduction 共⌸: left, transfer T: middle, and
dissipation E: rightspectra versus zverticaland log10(k)hori-
zontalat t5topand 6 bottom. Blue⫽⫺2104, green0,
and red2104.
ENERGY TRANSFER IN RAYLEIGH-TAYLOR INSTABILITY PHYSICAL REVIEW E 66, 026312 2002
026312-9
Tm1
2
v
ˆi
*N
ˆm,iv
ˆiN
ˆm,i
*d
mq,p,d27
and
Nq,i⫽⫺
viuj
xj,28
Np,i⫽⫺
1/2
p
xi,29
Nd,i1
2vi
uk
xk.30
V. SPECTRA
The time evolution of the two-dimensional energy spec-
trum, E(k,z0,t0,2,4,6), is plotted in Fig. 11. The initial
diffusion velocities result in a nonzero spectrum at t0. The
spectrum increases by several orders of magnitude as kinetic
energy is deposited into the flow. The peak of the spectrum
migrates toward lower wave numbers as bubbles and spikes
merge to form larger structures. The spectrum also fills out at
higher wave numbers as vortex stretching and bending mo-
tions transfer energy to smaller scales. Near the end of the
simulation it appears that an inertial range is just beginning
to form. In 42, Zhou argues that the inertial range for the
RTI flow will follow a 7
4power law. He reasons that the
spectrum will be modified from the classical Kolmogorov
5
3power law for isotropic turbulence as a result of the
external time scale introduced by gravity. Both power laws
are shown on the figure for comparison to the data. The
late-time spectra appear somewhat consistent with Zhou’s
theory; however, the statistical fluctuations in the spectra
there are not many points in the Fourier annuli at lower
wavenumbersare larger than the difference in slope be-
tween k7/4 and k5/3. Furthermore, there is no clear begin-
ning to the dissipation range, which appears to extend well
into the lower wave numbers, thereby steepening the slope of
the spectrum. Due to the closeness of the power laws, much
higher Reynolds number data and improved statistics will be
needed in order to discriminate between the two.
The zdependence of the energy spectrum is graphically
portrayed in Fig. 12 at four different times. The plots are in a
semi-Fourier domain with log10(k) along the abscissa and z
along the ordinate. The bulk of the energy is initially depos-
ited at a moderate wave number, corresponding to the domi-
nant wavelength of the density perturbations. As time
progresses, E(k,z,t) expands in both kand zbut maintains
its maximum value close to z0. The spectrum decreases
near the edges of the mixing zone, becoming negligible out-
side it.
The production 共⌸兲, transfer T, and dissipation Espec-
FIG. 15. ColorQuadratic (Tq: leftand pressure (Tp: right
contributions to transfer spectrum versus zverticaland log10(k)
horizontalat t3topand 4 bottom. Blue⫽⫺5104,
green0, and red5104.
FIG. 16. ColorQuadratic (Tq: leftand pressure (Tp: right
contributions to transfer spectrum versus zverticaland log10(k)
horizontalat t5top, and 6 bottom. Blue⫽⫺5104,
green0, and red5104.
ANDREW W. COOK AND YE ZHOU PHYSICAL REVIEW E 66, 026312 2002
026312-10
tra are plotted in Figs. 13 and 14 on the same zversus
log10(k) domain for the same times. Early on, the peaks of
the production and dissipation spectra are close to one an-
other, but later the production moves toward lower wave
numbers while the peak of the dissipation spectrum stays
roughly fixed. By the end of the simulation, there is some k
separation between the two, i.e., as structures merge within
the mixing layer, energy is deposited at larger scales. The
increasing separation of peaks between the production and
dissipation spectra is a direct result of the increasing Rey-
nolds number for this transitional flow. As time progresses,
both spectra advance in zand k, becoming increasingly large
and opposite.
In contrast to the straightforward nature of and E, the
transfer spectrum behaves in a very complicated manner. It
exhibits an intricate web of positive and negative regions,
interspersed over a wide range in zand k. At higher wave
numbers Tis mostly positive, indicating a net cascade of
energy to smaller scales. It is also positive at the top and
bottom of the mixing zone, suggesting production of energy
at the bubble and spike fronts. Inside the mixing zone, back-
scatter appears approximately equal to forwardscatter. It fur-
ther appears that, at each instant in time, some zlocations
may be undergoing forward energy cascade, while neighbor-
ing regions are simultaneously experiencing inverse cascade.
In order to unravel the irregular patchwork that constitutes
T, each individual component Tq,Tp, and Tdis plotted
separately. In Figs. 15 and 16, Tqand Tpare plotted on the
same semi-Fourier domain and at the same times as before.
The dilatation spectrum, Td, is plotted at all four timesin
Fig. 17. The dilatation spectrum is roughly two orders of
magnitude smaller than Tqand Tp, and hence, makes negli-
gible contribution to T. Interestingly, the quadratic (Tq) and
pressure (Tp) components are near opposites of one another,
their net contribution to Tbeing the result of extensive can-
cellation between the two. The quadratic term is mostly
negative at lower wave numbers inside the mixing zone and
mostly positive at higher wave numbers and near the edges
of the mixing envelope. Regarding the pressure term, nearly
the opposite is true, except at higher wave numbers where
positive and negative regions appear to be roughly equally
distributed. The net positive transfer of energy to the higher
wave numbers usual cascade pictureand to the bubble/
spike fronts is thus a result of quadratic interactions; with
pressure counterbalancing advection, for the most part.
As regards the dilatation, Td, although its influence on the
energetics is likely negligible, it is interesting to observe that
it is strongly polarized, i.e., positive for z0spike region
and negative for z0bubble region. The velocity diver-
gence is related to diffusion through Eq. 5. The net effect of
diffusion is to increase the density of fluid in the lower re-
gion and decrease the density of fluid in the upper region.
This results in a net transfer of energy in the zdirection
due to diffusion of heavy fluid into light fluid. At late times,
vigorous stirring at the center of the mixing zone causes the
positive and negative regions of Tdto overlap.
VI. CONCLUSIONS
We have examined the flow structure and energy budget
for Rayleigh-Taylor instability using the results of a high-
resolution direct numerical simulation. The outer-scale Rey-
nolds number was observed to follow a t3power law and
reached a final value of 5500, the highest Reynolds number
attained in a DNS or RTI flow to date. The curvature of
velocity correlation functions, as manifest in the Taylor mi-
croscales, exhibits strong anisotropy between the vertical and
horizontal directions, with a similar anisotropy observed in
the Taylor Reynolds numbers. This is due to the directed
forcing term in the governing equations. The energy spec-
trum, computed at the center of the mixing zone, appears to
manifest the beginning of an inertial range by the latter
stages of the simulation. Unfortunately, statistical fluctua-
tions in the spectrum make it difficult to establish whether
the inertial range follows a 5
3Kolmogorov power law or
the 7
4power law proposed by Zhou 42.
A formulation of the kinetic energy equation was pro-
posed, which enables straightforward extension of method-
ologies commonly employed for constant-density, isotropic
turbulence. The spectrum of each term in the energy equation
was computed as a function of height, horizontal wave num-
ber, and time. The peak of the energy spectrum migrates to
lower wave numbers as structures merge inside the mixing
layer. The production spectrum also moves to lower wave
numbers as gravity acts on the larger structures. The dissipa-
tion spectrum expands in both zand kwhile its peak stays
roughly fixed in wave number space. The limited Reynolds
number of the DNS appears to inhibit movement of the peak
FIG. 17. ColorDilatation (Td) component of transfer spectrum
versus zverticaland log10(k)horizontalat t3top left,4top
right,5lower left,and6lower right. Blue⫽⫺7106, green
0, and red7106.
ENERGY TRANSFER IN RAYLEIGH-TAYLOR INSTABILITY PHYSICAL REVIEW E 66, 026312 2002
026312-11
dissipation to higher wave numbers.
The transfer spectrum depends strongly on the inhomoge-
neous direction z. While the net energy transfer is from large
to small scales, there is significant inverse cascade over a
wide range of z. Energy transfer at the bubble/spike fronts is
strictly positive. Examination of the individual contributions
to the transfer reveals this to be due to the quadratic advec-
tionterm. Pressure acts to counterbalance advection, such
that the net transfer is substantially smaller than the transfer
from either single component. The dilatation term accounts
for energy transfer via diffusion of unequal-density fluids. It
is very small but serves to move energy from high to low
density regions.
The flow induced by Rayleigh-Taylor instability, as seen
here, has rather different character than that of homogeneous,
isotropic turbulence. The flow is highly anisotropic, even at
small scales, as evidenced by the Taylor microscales and
Reynolds numbers. Initial rates of entrainment and mixing
are determined by the initial conditions. Production rates al-
ways exceed dissipation rates; hence, the kinetic energy
grows rapidly in time. Furthermore, the evolution of the
spectra depends sensitively on initial conditions; since the
interfacial perturbations set the scale at which energy is ini-
tially injected into the flow. This has serious consequences
for large eddy simulations, where the initial perturbations
may be much smaller than the grid scale. In such cases,
subgrid-scale models must be capable of treating not only
backscatter, but also growth of spectra below the grid scale
and migration of the energy peak through the cutoff wave
number. The detailed analysis of the spectra, performed
herein, serves as a step toward developing subgrid-scale
models capable of treating RTI flows in ICF and astrophysics
applications.
ACKNOWLEDGMENTS
This work was performed under the auspices of the U.S.
Department of Energy by the University of California,
Lawrence Livermore National Laboratory, under Contract
No. W-7405-Eng-48. Additionally, we wish to thank Profes-
sor P. E. Dimotakis for many stimulating discussions on this
topic and for providing keen insights into this flow.
1Lord Rayleigh, Scientific papers Cambridge University Press,
Cambridge, England, 1900, Vol. II, p. 200.
2G. I. Taylor, Proc. R. Soc. London, Ser. A 201, 192 1950.
3S. Chandrasekhar, Hydrodynamic and Hydromagnetic Stability
Oxford University Press, Oxford, 1961.
4D. Sharp, Physica D 12,31984.
5R. D. Richtmyer, Commun. Pure Appl. Math. 13, 297 1960.
6E. E. Meshkov, Sov. Fluid Dyn. 4, 101 1969.
7J. Lindl, Internal Confinement Fusion: The Quest for Ignition
and Energy Gain Springer, New York, 1997.
8L. Smarr, J. R. Wilson, R. P. Barton, and R. L. Bowers, Astro-
phys. J. 246, 515 1981.
9W. D. Arnett, J. N. Bahcall, R. T. Kirshner, and S. E. Woosley,
Annu. Rev. Astron. Astrophys. 27, 629 1989.
10K. I. Reed, Physica D 12,451984.
11D. L. Youngs, Physica D 12,321984.
12D. L. Youngs, Physica D 37, 270 1989.
13D. L. Youngs, Laser Part. Beams 12, 725 1994.
14G. Dimonte, Phys. Plasmas 7, 2255 2000.
15G. Dimonte and M. Schneider, Phys. Fluids 12, 304 2000.
16B. A. Remington, R. P. Drake, H. Takabe, and D. Arbett, Phys.
Plasmas 7, 1641 2000.
17V. A. Andronov, S. M. Bakhrakh, E. E. Meshkov, V. N.
Mokhov, V. V. Nikiforov, A. V. Pevnitskii, and A. I. Tolshmya-
kov, Sov. Phys. Dokl. 27,51982.
18N. N. Annuchina, Yu. A. Kucherenko, V. E. Neuvazhaev, V. N.
Ogibina, L. I. Shibarshov, and V. G. Yakovlev, Izv.Akad. Nauk
SSSR, Mekh. Zhidk. Gaza 6, 157 1978.
19A. W. Cook and P. E. Dimotakis, J. Fluid Mech. 443,69
2001.
20G. K. Batchelor, The Theory of Homogeneous Turbulence
Cambridge University Press, Cambridge, UK, 1953.
21A. S. Monin and A. M. Yaglom, Statistical Fluid Mechanics
MIT Press, Cambridge, 1975, Vol. 2.
22P. F. Linden and J. M. Redondo, Phys. FluidsA 3, 1269 1991.
23P. F. Linden, J. Redondo, and D. L. Youngs, J. Fluid Mech.
265,971994.
24S. B. Dalziel, P. F. Linden, and D. L. Youngs, J. Fluid Mech.
399,11999.
25C. Eckart, J. Mar. Res. VII, 265 1948.
26Y. Zhou and C. G. Speziale, Appl. Mech. Rev. 51, 267 1998.
27J. A. Domaradzki and R. S. Rogallo, Phys. Fluids A 2, 413
1990.
28P. K. Yeung and J. G. Brasseur, Phys. Fluids A 3, 884 1991.
29K. Ohkitani and S. Kida, Phys. Fluids A 4, 794 1992.
30Y. Zhou, Phys. Fluids A 5, 1092 1993.
31Y. Zhou, Phys. Fluids A 5, 2511 1993.
32Y. Zhou, P. K. Yeung, and J. G. Brasseur, Phys. Rev. E 53,
1261 1996.
33J. A. Domaradzki, W. Liu, C. Hartel, and L. Kleiser, Phys.
Fluids 6, 1583 1994.
34D. D. Joseph, Eur. J. Mech. B/Fluids 9, 565 1990.
35S. K. Lele, J. Comput. Phys. 103,161992.
36P. E. Dimotakis, J. Fluid Mech. 409,692000.
37H. Tennekes and J. L. Lumley, A First Course in Turbulence
MIT Press, Cambridge, 1972.
38K. K. Nomura and S. E. Elghobashi, Theor. Comput. Fluid
Dyn. 5, 153 1993.
39C.-S. Yih, J. Fluid Mech. 9,681960.
40J. L. Rodriguez Azara and G. Emanuel, Phys. Fluids 31, 1058
1998.
41S. Kida and S. A. Orszag, J. Sci. Comput. 5,851990.
42Y. Zhou, Phys. Fluids 13, 538 2001.
ANDREW W. COOK AND YE ZHOU PHYSICAL REVIEW E 66, 026312 2002
026312-12
... For each concentration profile, the mixing layer width is defined as the distance between the 10% and 90% threshold locations. This is similar to the definition of visual width used in simulation studies of both RMI and RTI (see Cook & Dimotakis 2001;Cook & Zhou 2002;Zhou & Cabot 2019), where the plane-averaged mole fraction or volume fraction profile is used along with a typical threshold cuttoff of 1% and 99%, e.g. ...
... (3.11) Figure 19 shows the temporal variation for both definitions of the outer-scale Reynolds number. The outer-scale Reynolds numbers using the TKE-based definition for ℎ are roughly a factor of 2 larger, mostly due to the TKE-based width being a lot larger than the VF-based width in all cases, with neither definition close to reaching the critical value of Re ℎ ≳ 1-2 × 10 4 for fully developed turbulence (Dimotakis 2000). For both the = −1 and = −2 perturbations the VF-based Reynolds number is approximately constant in time, consistent with the measured values of given in table 4. Dimotakis (2000) showed that for stationary flows, fully developed turbulence is obtained when / ⩾ 1 where = 5 is the Liepmann-Taylor length scale and = 50 is the inner-viscous length scale, with and the Taylor and Kolmogorov length scales respectively. ...
... The outer-scale Reynolds numbers using the TKE-based definition for ℎ are roughly a factor of 2 larger, mostly due to the TKE-based width being a lot larger than the VF-based width in all cases, with neither definition close to reaching the critical value of Re ℎ ≳ 1-2 × 10 4 for fully developed turbulence (Dimotakis 2000). For both the = −1 and = −2 perturbations the VF-based Reynolds number is approximately constant in time, consistent with the measured values of given in table 4. Dimotakis (2000) showed that for stationary flows, fully developed turbulence is obtained when / ⩾ 1 where = 5 is the Liepmann-Taylor length scale and = 50 is the inner-viscous length scale, with and the Taylor and Kolmogorov length scales respectively. These length scales may be related to the outer-scale Reynolds number by from which it can be shown that Re ℎ ⩾ 10 4 for fully developed turbulence. ...
Preprint
Full-text available
The effects of initial conditions on the evolution of the Richtmyer-Meshkov instability (RMI) at early to intermediate times are analysed, using numerical simulations of an idealised version of recent shock tube experiments performed at the University of Arizona. The experimental results are bracketed by performing both implicit large-eddy simulations of the high-Reynolds-number limit as well as direct numerical simulations (DNS) at Reynolds numbers lower than those observed in the experiments. Various measures of the mixing layer width, based on both the plane-averaged turbulent kinetic energy and volume fraction profiles are used to explore the effects of initial conditions on $\theta$ and are compared with the experimental results. The decay rate of the total fluctuating kinetic energy is also used to estimate $\theta$ based on a relationship that assumes self-similar growth of the mixing layer. The estimates for $\theta$ range between 0.44 and 0.52 for each of the broadband perturbations considered and are in good agreement with the experimental results. Overall, the results demonstrate important differences between broadband and narrowband surface perturbations, as well as persistent effects of finite bandwidth on the growth rate of mixing layers evolving from broadband perturbations. Good agreement is obtained with the experiments for the different quantities considered; however, the results also show that care must be taken when using measurements based on the velocity field to infer properties of the concentration field.
... The integrated kinetic energy spectrum and density fluctuation spectrum in the RMI were studied by Schilling, Latini & Don (2007) and Tritschler et al. (2014) using data from numerical simulations. Numerical experiments have explored some of the terms in the transport of kinetic energy in wavenumber space such as in Cook & Zhou (2002) for the Rayleigh-Taylor instability (RTI) case and in Thornber & Zhou (2012) for the RMI case, while a similar experimental study of the combined scalar and kinetic fields has not been performed to date. Energy transfer within a reshocked gas curtain was studied by Zeng et al. (2018) calculating the integrated homogeneous flux of kinetic energy. ...
... The structures seem to correlate, and the values are comparable; however, some of the details are different. Cook & Zhou (2002) calculate the time-varying dissipation, production and transport in non-scaling coordinates for an RTI problem. The production term of the system is very different between the RTI and RMI, so differences are expected in the temporal behaviour; however, transport and dissipation have the same mathematical structure. ...
Article
Full-text available
The Richtmyer–Meshkov instability (Richtmyer, Commun. Pure Appl. Maths , vol. 13, issue 2, 1960, pp. 297–319; Meshkov, Fluid Dyn. , vol. 4, issue 5, 1972, pp. 101–104) of a twice-shocked gas interface is studied using both high spatial resolution single-shot (SS) and lower spatial resolution, time-resolved, high-speed (HS) simultaneous planar laser-induced fluorescence and particle image velocimetry in the Wisconsin Shock Tube Laboratory's vertical shock tube. The initial condition (IC) is a shear layer with broadband diffuse perturbations at the interface between a helium–acetone mixture and argon. This IC is accelerated by a shock of nominal strength Mach number $M = 1.75$ , and then accelerated again by the transmitted shock that reflects off the end wall of the tube. An ensemble of experiments is analysed after reshock while the interface mixing width grows linearly with time. The kinetic and scalar energy spectra and the terms of their evolution equation are calculated and compared between SS and HS experiments. The inertial range scaling of the scalar power spectrum is found to follow Gibson's relation (Gibson, Phys. Fluids , vol. 11, issue 11, 1968, pp. 2316–2327) as a function of Schmidt number when the effective turbulent Schmidt number is used in place of the material Schmidt number that controls equilibrium scaling. Further, the spatially integrated scalar flux follows similar behaviour observed for the kinetic energy in large eddy simulation studies by Zeng et al. ( Phys. Fluids , vol. 30, issue 6, 2018, 064106) while the spatially varying scalar flux exhibits back scatter along the centre of the mixing layer and forward energy transfer in the spike and bubble regions.
... For each concentration profile, the mixing layer width is defined as the distance between the 10 % and 90 % threshold locations. This is similar to the definition of visual width used in simulation studies of both RMI and RTI (see Cook & Dimotakis 2001;Cook & Zhou 2002;Zhou & Cabot 2019), where the plane-averaged mole fraction or volume fraction profile is used along with a typical threshold cuttoff of 1 % and 99 %, e.g. ...
... (3.11) Figure 19 shows the temporal variation for both definitions of the outer-scale Reynolds number. The outer-scale Reynolds numbers using the TKE-based definition for h are roughly a factor of 2 larger, mostly due to the TKE-based width being a lot larger than the VF-based width in all cases, with neither definition close to reaching the critical value of Re h 1-2 × 10 4 for fully developed turbulence (Dimotakis 2000). For both the m = −1 and m = −2 perturbations the VF-based Reynolds number is approximately constant in time, consistent with the measured values of θ given in table 4. ...
Article
Full-text available
The effects of initial conditions on the evolution of the Richtmyer–Meshkov instability (RMI) at early to intermediate times are analysed, using numerical simulations of an idealised version of recent shock tube experiments performed at the University of Arizona (Sewell et al. , J. Fluid Mech. , vol. 917, 2021, A41). The experimental results are bracketed by performing both implicit large-eddy simulations of the high-Reynolds-number limit as well as direct numerical simulations (DNS) at Reynolds numbers lower than those observed in the experiments. Various measures of the mixing layer width $h$ , known to scale as ${\sim }t^\theta$ at late time, based on both the plane-averaged turbulent kinetic energy and volume fraction profiles are used to explore the effects of initial conditions on $\theta$ and are compared with the experimental results. The decay rate $n$ of the total fluctuating kinetic energy is also used to estimate $\theta$ based on a relationship that assumes self-similar growth of the mixing layer. The estimates for $\theta$ range between 0.44 and 0.52 for each of the broadband perturbations considered and are in good agreement with the experimental results. Decomposing the mixing layer width into separate bubble and spike heights $h_b$ and $h_s$ shows that, while the bubbles and spikes initially grow at different rates, their growth rates $\theta _b$ and $\theta _s$ have equalised by the end of the simulations. Overall, the results demonstrate important differences between broadband and narrowband surface perturbations, as well as persistent effects of finite bandwidth on the growth rate of mixing layers evolving from broadband perturbations. Good agreement is obtained with the experiments for the different quantities considered; however, the results also show that care must be taken when using measurements based on the velocity field to infer properties of the concentration field.
... Researchers have extensively studied the kinetic energy transport for turbulence induced by incompressible RTI and buoyancy-driven turbulence 21,26,36,53,61,62 . On the other hand, the analysis of kinetic energy budgets for compressible RTI is still limited except some very recent works 20,63,64 . ...
Article
Full-text available
Rayleigh–Taylor instability, RTI, occurs at the interface separating two fluids subjected to acceleration when the density gradient and the acceleration are in opposite directions. Previous scientific research primarily considered RTI under the incompressible assumption, which may not be valid in many high-energy-density engineering applications and astrophysical phenomena. In this study, the compressibility effects of the background isothermal stratification strength on multi-mode two-dimensional RTI are explored using fully compressible multi-species direct numerical simulations. Cases under three different isothermal Mach numbers – Ma=0.15, 0.3, and 0.45 – are investigated to explore weakly, moderately, and strongly stratified compressible RTI, respectively, at an Atwood number of 0.04. Unlike incompressible RTI, an increase in the flow compressibility through the strength of the background stratification can suppress the RTI growth and can lead to a termination of the RTI mixing layer growth with a highly molecularly mixed state. Our findings suggest that even at the chosen relatively low Atwood number, the variable-density effects can be significantly enhanced due to an increase in the background stratification for the compressible RTI as different spatial profiles become noticeably asymmetric across the mixing layer for the strongly stratified case. In addition, this study compares the chaotic behavior of the cases by studying the transport of the turbulent kinetic energy as well as the vortex dynamics. The Reynolds number dependence of the results is also examined with three different Reynolds numbers, and the findings for the large-scale mixing and flow quantities of interest are shown to be universal in the range of the Reynolds numbers studied.
... power spectral density) is positive. For example, the Fourier transform of an auxiliary field w ≡ √ ρu is performed and then squared such that the spectrum is positive and integrates to ρ|u| 2 /2 [77,78]. This approach, while being mathematically sound on its own, creates difficulties when analyzing the scale-dependent equations governing such KE spectra as shown in [79]. ...
Preprint
Full-text available
We generalize the `filtering spectrum' [1] to probe scales along different directions by spatial coarse-graining. This multi-dimensional filtering spectrum quantifies the spectral content of flows that are not necessarily homogeneous. From multi-dimensional spectral information, we propose a simple metric for shape anisotropy at various scales. The method is applied to simulations of 2D and 3D Rayleigh-Taylor (RT) turbulence, which is inhomogeneous and anisotropic. We show that 3D RT has clear shape anisotropy at large scales with approximately $4:3$ vertical to horizontal aspect ratio, but tends toward isotropy at small scales as expected [2,3,4]. In sharp contrast, we find that RT in 2D simulations, which are still the main modeling framework for many applications, is isotropic at large scales and its shape anisotropy increases at smaller scales where structures tend to be horizontally elongated. While this may be surprising, it is consistent with recent results in [5]; large-scale isotropy in 2D RT is due to the generation of a large-scale overturning circulation via an upscale cascade, while small scale anisotropy is due to the stable stratification resultant from such overturning and the inefficient mixing in 2D.
Article
An expected statistical law of passive scalar turbulence subjected to high Schmidt numbers is studied with the Hessian-based Lagrangian renormalized approximation (HBLRA), a self-consistent closure theory recently developed for passive scalar turbulence. Without relying on any empirical parameters, HBLRA derives the scalar variance spectrum proportional to k−1 in the viscous-convective range predicted by Batchelor. Thanks to improvements in the physical description, HBLRA yields quantitatively better predictions of the scalar variance spectrum and its universal constant in comparison with pioneering closure theories. Our results suggest that the turbulence motion of only a limited scale range may contribute to the scalar flux within the viscous-convective range. Numerical solutions for finite Schmidt numbers (Sc=10–100000) are obtained, where a clear viscous-convective range appears with a well-developed inertial-convective range.
Article
Full-text available
The importance of non-locality of mean scalar transport in two-dimensional Rayleigh–Taylor Instability (RTI) is investigated. The macroscopic forcing method is utilized to measure spatio-temporal moments of the eddy diffusivity kernel representing passive scalar transport in the ensemble averaged fields. Presented in this work are several studies assessing the importance of the higher-order moments of the eddy diffusivity, which contain information about non-locality, in models for RTI. First, it is demonstrated through a comparison of leading-order models that a purely local eddy diffusivity is insufficient to capture the mean field evolution of the mass fraction in RTI. Therefore, higher-order moments of the eddy diffusivity operator are not negligible. Models are then constructed by utilizing the measured higher-order moments. It is demonstrated that an explicit operator based on the Kramers–Moyal expansion of the eddy diffusivity kernel is insufficient. An implicit operator construction that matches the measured moments is shown to offer improvements relative to the local model in a converging fashion.
Article
This work investigates the compressible turbulence induced by Richtmyer–Meshkov (RM) instability using high-resolution Navier–Stokes simulations. Special attention is paid to the characteristics of RM turbulence including the mixing width growth, the turbulent kinetic energy (TKE) decay, the mixing degree, inhomogeneity and anisotropy. Three distinct initial perturbation spectra are designed at the interface to reveal the initial condition imprint on the RM turbulence. Results show that cases with initial large-scale perturbations present a stronger imprint on statistical characteristics and also a quicker growth of mixing width, whereas cases with small-scale perturbations present a faster TKE decay, greater mixing level, higher isotropy and homogeneity. A thorough analysis on the inter-scale energy transfer in RM turbulence is also presented with the coarse-graining approach that exposes the two subfilter-scale (SFS) energy fluxes (i.e. deformation work and baropycnal work). A strong correlation between the nonlinear model of baropycnal work ( Fluids , 4 (2), 2019) and the simulation results is confirmed for the first time, demonstrating its potential in modelling the RM turbulence. Two primary mechanisms of baropycnal work (the straining and baroclinic generation processes) are explored with this nonlinear model. The evolutions of two SFS energy fluxes exhibit distinct behaviours at various filter scales, in different flow regions and under various flow motions (strain and rotation). It is found that all three cases share the common inter-scale energy transfer dynamics, which is important for modelling the RM turbulence.
Article
We generalize the “filtering spectrum” to probe scales along different directions by spatial coarse-graining. This multidimensional filtering spectrum quantifies the spectral content of flows that are not necessarily homogeneous. From multidimensional spectral information, we propose a simple metric for shape anisotropy at various scales. The method is applied to simulations of two-dimensional (2D) and 3D Rayleigh-Taylor (RT) turbulence, which is inhomogeneous and anisotropic. We show that 3D RT has clear shape anisotropy at large scales with approximately 4:3 vertical to horizontal aspect ratio but tends toward isotropy at small scales as expected. In sharp contrast, we find that RT in 2D simulations, which are still the main modeling framework for many applications, is isotropic at large scales and its shape anisotropy increases at smaller scales where structures tend to be horizontally elongated. While this may be surprising, it is consistent with recent results that large-scale isotropy in 2D RT is due to the generation of a large-scale overturning circulation via an upscale cascade, while small-scale anisotropy is due to the stable stratification resultant from such overturning and the inefficient mixing in 2D.
Article
We demonstrate a methodology for diagnosing the multiscale dynamics and energy transfer in complex HED flows with realistic driving and boundary conditions. The approach separates incompressible, compressible, and baropycnal contributions to energy scale-transfer and quantifies the direction of these transfers in (generalized) wavenumber space. We use this to compare the kinetic energy (KE) transfer across scales in simulations of 2D axisymmetric vs fully 3D laser-driven plasma jets. Using the FLASH code, we model a turbulent jet ablated from an aluminum cone target in the configuration outlined by Liao et al. [Phys. Plasmas, 26 032306 (2019)]. We show that, in addition to its well known bias for underestimating hydrodynamic instability growth, 2D modeling suffers from significant spurious energization of the bulk flow by a turbulent upscale cascade. In 2D, this arises as vorticity and strain from instabilities near the jet's leading edge transfer KE upscale, sustaining a coherent circulation that helps propel the axisymmetric jet farther (≈25% by 3.5 ns) and helps keep it collimated. In 3D, the coherent circulation and upscale KE transfer are absent. The methodology presented here may also help with inter-model comparison and validation, including future modeling efforts to alleviate some of the 2D hydrodynamic artifacts highlighted in this study.
Article
Full-text available
A missing factor of 2 in the scaled diffusivity, used to evaluate σ(k) from the Rayleigh-Taylor linear-stability theory for viscous and diffusive fluids of Duff et al. (1962), led to an incorrect plot of this curve in figure 2 of Cook & Dimotakis (2001). The corrected curve is plotted in figure 1, as a dimensionless product, σ(k) τ , where τ = L/(Ag) is the characteristic time (6c). The error led to an incorrect inference regarding the agreement of the relative times for the A, B, and C cases when mixing-zone growth emerges from its diffusion-dominated, h ∝ t 1/2 regime to the subsequent, faster-growth regime, with h the Rayleigh-Taylor mixing-zone height. The relative values of these times (figure 8) are then in qualitative accord with linear theory. This is despite the fact that the initial perturbation amplitudes are moderate (21) and may place even early-time growth outside the applicability of linear theory. Additionally, the 1/3 factor in the Neumann pressure boundary condition (17) is erroneous. The correct boundary condition is, ∂p ∂z = 1 Re ∂ 2 w ∂z 2 − ρ , prior to the arrival of mixed fluid at the end walls, when ∇ρ = 0 and therefore also (15), ∇ · u = − 1 Pe ∇ · 1 ρ ∇ ρ = 0 , at the end walls. In separate simulations with large bubbles/spikes and at a lower Re, the difference in mixing-zone behavior caused by the effectively lower-viscosity boundary condition was confirmed to be negligible. This is as expected in view of the higher Re in the original DNS runs and their termination some time prior to the arrival of mixed fluid at the end walls.
Article
Full-text available
The nonlinear interscale couplings in a turbulent flow are studied through direct numerical simulations of the response of isotropic turbulence to isotropic and anisotropic forcing applied at the large scales. Specifically, forcing is applied to the energy‐containing wave‐number range for about two eddy turnover times to fully developed isotropic turbulence at Taylor‐scale Reynolds number 32 on an 1283 grid. When forced isotropically, the initially isotropic turbulence remains isotropic at all wave numbers. However, anisotropic forcing applied through an array of counter‐rotating rectilinear vortices induces high levels of anisotropy at the small scales. At low wave numbers the force term feeds energy directly into two velocity components in the plane of the forced vortices. In contrast, at high wave numbers the third (spanwise) component receives the most energy, producing small‐scale anisotropy very different from that at the large scales. Detailed analysis shows that the development of small‐scale anisotropy is caused primarily by nonlocal wave‐vector triads with one leg in the forced low‐wave‐number range. This latter result is particularly significant because asymptotic analysis of the Fourier‐transformed Navier–Stokes equations shows that distant triadic interactions coupling the energy‐containing and dissipative scales persist at asymptotically high Reynolds numbers, suggesting that the structural couplings between large and small scales in these moderate Reynolds number simulations would also exist in high Reynolds number forced turbulence. The results therefore imply a departure from the classical hypothesis of statistical independence between large‐ and small‐scale structure and local isotropy.
Article
Full-text available
The fundamental aspects of isotropic turbulence are reviewed in order to gain a better insight into the physical processes of turbulence. After first reviewing the Kolmogorov energy spectrum and the energy cascade, the Kolmogorov hypothesis of local isotropy is discussed in depth. Then, the detailed physical processes involving energy transfer and interacting scales in isotropic turbulence, including triad interactions, are reviewed. The inertial range and self-similarity are also discussed along with the response of the small scales to large-scale anisotropy and the final stages of the decay process. Results from direct and large-eddy simulations of isotropic turbulence -including a discussion of subgrid scale modeling - are then discussed in detail to illustrate these points. The article closes with a review of self-preservation in isotropic turbulence and a discussion of the prospects for future research. It contains 155 references.
Article
Full-text available
This paper describes experiments on the mixing produced by Rayleigh–Taylor instability between two miscible fluids. A layer of brine is placed over a layer of fresh water in a gravitational field, and the ensuing flow is visualized by laser‐induced fluorescence and measurements are made of the concentration fields as the flow develops. It is found that large‐scale disturbances develop which produce intermingling of the fluids, but molecular mixing occurs as a result of small‐scale instabilities which grow on the edges of the larger scale motions. When the system overturns stably stratified fluid is produced, and the mixing efficiency of the process is measured and found to be large compared with other forms of mixing. It is suggested that this increased efficiency is due to the fact that much of the mixing occurs when the system is unstably stratified.
Article
We investigate inertial subrange energy spectra associated with turbulent flows developed by the Rayleigh-Taylor instability (RTI) and Richtmyer-Meshkov instability (RMI). We argue that the extended Kolmogorov-Kraichnan phenomenology originally developed for turbulent flows with an external agency should also be applicable to these instability driven turbulent flows. A prediction of the mixing zone width for the RMI induced turbulent flow is presented using the RMI modified energy spectrum and a two-equation turbulence model. A possible application to subgrid modeling for large-eddy simulation is discussed briefly. This work was performed under the auspices of the U.S. Department of Energy by the University of California Lawrence Livermore National Laboratory under contract No. W-7405-Eng-48.
Article
The world's water resources are coming under increasing stress, a stress that will become critical globally sometime during the next century. This is due to the rapidly rising population demanding more and more water and an increasing level of affluence. The book discusses the background to this issue and the measures to be taken over the next 20-30 years to overcome some of the difficulties that can be foreseen, and the means of avoiding others, such as the hazard of floods. It looks at the water resource and its assessment and management in an integrated fashion. It deals with the requirements of agriculture and of rural and urban societies and to a lesser extent with those of industry and power, against the background of the needs of the natural environment. It presents a number of ways and means of improving the management of national and international affairs involving fresh water. It highlights the importance of fresh water as a major issue for the environment and for development.
Article
It is shown that, when two superposed fluids of different densities are accelerated in a direction perpendicular to their interface, this surface is stable or unstable according to whether the acceleration is directed from the heavier to the lighter fluid or vice versa. The relationship between the rate of development of the instability and the length of wave-like disturbances, the acceleration and the densities is found, and similar calculations are made for the case when a sheet of liquid of uniform depth is accelerated.
Article
An analysis of the data of a direct numerical simulation of a forced incompressible isotropic turbulence at a high Reynolds number (Rλ≊180) is made to investigate the interaction among three Fourier modes of wave numbers that form a triangle. The triad interaction is classified into six types according to the direction of the energy transfer to each Fourier mode: (+,+,−), (+,−,+), (+,−,−), (−,+,+), (−,+,−), (−,−,+), where the + (or −) denotes the energy gain (or loss) of the modes of the largest, the intermediate, and the smallest wave numbers in this order. The last three types of the interaction are very few. In the first type of the interaction, a comparable amount of energy is exchanged typically among three modes of comparable magnitude of wave numbers. In the second and third types, the magnitudes of the larger two wave numbers are comparable and much larger than the smallest one, and a great amount of energy is exchanged between the former two. This behavior of the triad interaction agrees very well with the prediction due to various quasinormal Markovianized closure theories of turbulence, and was observed before for lower Reynolds number turbulence by Domaradzki and Rogallo [Phys. Fluids A 2, 413 (1990)]. The fact that the dominant triad interactions involve different scales of motion suggests that the statistics of the small-scale motions of turbulence may be directly affected by the large-scale motions. Nevertheless, Kolmogorov’s local energy cascade argument may hold at least partially because the energy is exchanged predominantly between modes of two comparable scales in the triad interaction.
Conference Paper
The aim of this talk is to survey Rayleigh-Taylor instability, describing the phenomenology that occurs at a Taylor unstable interface, and reviewing attempts to understand these phenomena quantitatively.