Anda di halaman 1dari 40

UCRL-JC- 115690 PREPRINT

Mixing in Explosions

A. L. Kuhl

This paper was prepared for submittal to the Technical Meeting at the Russian Academy of Sciences Moscow, Russia December 1-3,1993

. ..v..,v,...:....,,,,

:!i
.:::.::':::::" ,::::::::::::., .:. .... ,.. ..........

::::::::::::::::::::::::: i:!:!:!:'.:i

Thisis a preprintof apaperintendedforpublicationinajcmrnalorproceedinKs.Since chanses may be made before publication, this preprintis made available with the understandinKthat it will not be cited or reproducedwithout the permission of the author.

.............................................. .... iiiiii_,!!:iiiii:_iiiii!ii!ii!i!iii:i!!

DISTRIBUTION OF THIS DOCUMENT IS UNLIMITED

v d

DISCLAIMER This document was prepared as an account of work sponsored by an agency of the United States Government. Neither the United States Government nor the University of California nor any of their employees, makes any wm-rJmty, express or implied, or assumes my legal liability or respemibility for the Rccuracy, completeness, er usefulness of any information, apparatus, product, or process disclosed, or represents that its use weald not infringe privately owned rights. Referenceherein to any specific commercial produc/s, process, or service by trade name, trademark, mmufacturer, or otherwise, does net necessarily constitute or imply its endorsement, _mmendatiomb or favoring by the United States Government or (he University of Californ_ The views and opinions of authors expressed herein do not necessarily state or reflect lhose of _he United States Government or the University of California,.and shall net be used for advertising or product endorsement purposes.

J o

NI........

_1 _''

-,.42.7,

Mixing in Explosions
A. L. Kuhl Lawrence Livermore National Laboratory El Segundo, California
ip

Abstract

Explosions

always contain embedded turbulent

mixing regions, for example: Described here is one

boundary layers, shear layers, wall jets, and unstable interfaces. an HE-driven blast wave. dimensional mesh. numerical Vorticity

particular example of the latter, namely, the turbulent mixing occurring in the fh'eball of The evolution of the turbulent mixing was studied via twoon the fireball interface by baroclinic effects. The simulations of the convective mixing processes on an adaptive

was generated

interface was unstable, and rapidly evolved into a turbulent mixing layer. Four phases of mixing were observed: (i) a strong blast wave phase; (ii) and implosion phase; (iii) a reshocking phase; and (iv) an asymptotic mixing phase. vorticity decayed due to a cascade process. The flowfield was azimuthally enstrophy turbulence averaged to evaluate the mean and r.m.s, fluctuation profiles across the mixing layer. The This caused the corresponding parameter to increase linearly with time -- in agreement with homogeneous calculations of G.K. Batchelor.

KEY WORDS:

fireball instabilities, baroclinic effects, turbulent mixing layer profiles, scaling, enstrophy, cascade of vorticity.

INTRODUCTION One can say that the science of explosion dynamics publication of similarity solutions to the point-explosion , and L.I. Sedov (1946). spectrum of blast wave problems, ' began with the seminal

problem by G. I. Taylor (1941) by Sedov (1959),

In Russia, this phase plane analysis was used to solve a broad as described in the monographs

Zel'dovich and Raizer (1966), Stanyukovich (1955), Barenblatt (1979), and Korobeinikov (1985). In the United States, A. K. Oppenheim (1972 a,b) also utilized the phase plane method to systematically investigate explosions (e.g., all blast wave solutions bounded by

MASTEB
DISTRIBUTION OF THIS DOCUMENT IS UNLIMITED
I

a strong shock, all blast wave solutions bounded by a strong C. J. detonation, etc.), and thereby discovered whole flew classes of blast wave solutions (Barenblatt et al., 1980). The next phase of research led to investigations of non-self-similar Initially, the research focused on one-dimensional point explosions dimensional Eventually, solutions by Goldstine driven blast waves by Brode (1959). problems the numerical to very complex algorithms blast waves,

which by necessity, required finite difference solutions of the gas dynamics equations. problems, for example, the decay of were extended to twosurfaces. accurate and von Neumann (1955), and Brode (1955), and HELater, such calculations were improved of blast waves from planar to the point where reflection

such as the reflection

flows such as double-Mach However, such numerical

(DMR) could be

obtained (Colella et al., 1986). laminar solutions.

simulations only produced

Obviously, real explosions always contain turbulent mixing regions, for example: boundary layers, slip lines and shear layers, wall jets, and unstable interfaces. recent advances in computational power and improvements now feasible to perform direct numerical simulations of three-dimensional, Due to turbulent in numerical algorithms, it is

mixing. This paper offers an example of such mixing in explosions. The tool used for this purpose simulations of compressible turbulent convective mixing--rather Reynolds-number is a Convective flows. Mixing Model (CMM) for that

It is based on the idealization

than diffusive _'ansportmis by Oppenheim

the dominant physics for highextension of the Zero

flows. It can be viewed as a compressible-flow

Mach Number Model proposed laws of gasdynamics

(1986) and Chorin (1986) for direct algorithm (Colella and Glaz,

simulations of turbulent combustion.

Specifically, CMM solves the inviscid conservation

by means of a high-order Godunov

1985). Adaptive Mesh Refinement (AMR) is used to follow the turbulent regions of the explosion, and calculate the convective mixing processes on the computational mesh. CMM avoids the perils and limitations of turbulent models which are, by

necessity, based on integration of mean-field equations.

As Chorin (1975) has so aptly CMM is based on the the convective here, this

pointed out, the laws of fluid dynamics are nonlinear, so the integration of the averaged equations does not equal the average of the integrated equations. latter approach, namely, we integrate the equationsmby mixing on an adaptive grid_and then average the solution. calculating

As demonstrated

approach provides

a more faithful numerical simulation to a wide spectrum Oppenheim

of fluid-

dynamic mixing problems.

himself was one of the first and most forceful

proponents of this approach for explosion dynamics, hence it is natural that this review should appear in a volume dedicated in his honor. THE PROBLEM

The problem considered here is the turbulent mixing in the fireball, created by the detonation of a spherical high-explosives high velocity, surrounding initially atmosphere. The detonation (HE) charge. The fireball gases expand at a charges are Such density instabilities (1960) Small around 9 km/s, which drives a strong blast wave into the products from solid explosives air (-0.01 g/cm3),

typically very dense (--2.5 g/cm 3) compared to the shock-compressed interfaces (Youngs, instabilities, perturbations are unstable to constant accelerations 1984), impulsive and misaligned accelerations pressure gradients

hence the density ratio across the interface is very large (around 250). such as Rayleigh-Taylor (i.e., baroclinic as in Richtmyer

(1960)-Meshkov effects).

on the interface, arising either from granular irregularities

on the charge

surface or from molecular fluctuations, rapidly grow into a turbulent mixing region--as is quite evident from high-speed photography of explosions (Figure 1). Turbulent mixing in fireballs is of scientific interest for a variety of reasons. First, mixing on a f'k,'eball interface can lead to afterburning for explosives that are not oxygen balanced (e.g., TNT). Afterburning can add considerable energy to the blast wave (as much as 50 percent) that is exceedingly difficult to calculate, since it depends on the On a more general note, turbulent fireballs

turbulent mixing rate which is not known. spherical geometry.

provide an example of nonsteady compressible, turbulent mixing on density interfaces in As will be shown here, such problems contain a wealth of fluid mechanic phenomena that are certainly worthy of detailed scientific investigation. The first finite difference 1959) were one-dimensional, . instabilities calculations interfaces of HE-driven blast waves (e.g., Brode. The growth of and was f'n'st mentioned by Anisimov

hence mixing effects were neglected.

on spherical HE-fireball

Zel'dovich (1977). This was followed by experimental studies for cylindrical explosions by Davydov et al. (1978, 1979). Some of the f'n-st analysis of such mixing layers was reported by Anisimov et al. in 1983. They identified two regimes: (i) a strong blast wave regime where the mixing width was thin compared to the distance between the inner and

outer shock (they called this free Rayleigh-Taylor regime, recent progress on interface instabilities and Inogamov (1993).

turbulence);

and (ii) an asymptotic Much of the

where the mixing width was of the order of the fireball radius.

is summarized in a review article by Anisimov

While others (Marcus et al., 1991; Ruppert, 1991) have studied the growth of instabilities investigates wave. convective on planar density interfaces accelerated by shock waves, this paper instabilities on a spherical density interface imbedded in an HE-driven blast we shall study the dynamics of mixing in f'u'eballs by following This same methodology has been instabilities in high-altitude explosions (Kuhl et al., mixing processes on an adaptive mesh.

In particular,

used to study spherical interface

1993) and implosions (Klein et al., 1993). The formulation of the calculations section energy. presents: visualization is described in the next section. The results

of the mixing process,

analyses of the mixing layer for further work.

growth, mean and r.m.s, profiles across the layer, and evolution of the turbulent kinetic This is followed by conclusions and recommendations

FORMULATION

The problem considered spherical charge of PBX-9404, corresponding

here is the blast wave created by the detonation For convenience,

of a

with an initial charge density of P0 = 1.84 g/cm 3 and a the charge

detonation velocity of WCj = 8.8 krn/s.

radius was assumed to be Ro = 1.0 m; this results in a charge mass of 7,700 kg. The ambient gas was air at sea level conditions: p**= 1.01325 bars; p**= 1.2254 x 10.3 g/cm3; e_,--- 1.94x10 ) erg/g; a**- 334 m/s; u = 0; y = 1.41.

In order to focus on the convective approximations spherically detonation balanced approximation, were made. In particular, (2) symmetric;

mixing effects, a number of idealizing (i.e., an infinite Reynolds number

we assumed that: (1) the explosion is initially

The flow is inviscid

where molecular diffusion effects are neglected); (3) both the air and the equilibrium; (4) the explosive is oxygen so the blast wave has constant energy; (5) buoyancy

products are in local thermodynamic (i.e., no afterburning),

effects are neglected (in order to focus on baroclinic dimensional (2-D).


0

effe:cs); (6) the flow is two-

Based on the preceding conservation laws of gas dynamics: O

assumptions,

the flow is governed

by the inviscid

_ttP+ V. (pu) = 0 a _ttPu + V. (puu) = -Vp 3 x,pE + V .(pEu)=-V.(pu)

(1)

(2)

at

(3)

Where U denotes the velocity and E represents the total energy E = e + 0.5 usu. air, it takes the well-known form:

The

pressure p is related to the density O and internal energy e by the equation of state. For

p=(y,-1)pe

(4)

where the effective gamma _t = ye(O,e) comes from the equilibrium air calculations of e Gilmore (1955). A JWL equation of state was used for the detonation products:

p= A(1-wPo/PR1)e

-R'p*/p+ B(1- WPo/PR2)e -R'p*/p+ (yd - 1)pe

(5)

where according to Dobratz (1974): A = 8.545 Mbars, B. = 0.2049 Mbars, RI= 4.60, R2 = 1.35 and Yd = 1.25 for PBX-9404. In mixed cells, the pressure was defined as a massweighted average of the air and HE pressures. In order to follow the mixing processes, an additional transport equation was used for concentration C:

-_3C =(us V)C=O at

(6)

where C was initialized an C = 1 inside the charge (r < r0) and C = 0 in the air (r > r0).

The above set of equations (1, 2, 3 and 6) were integrated by means of a secondorder Godunov algorithm (Colella and Glaz, 1985). mixing region of the fireball. In order to initialize 1950). the problem, we assumed that the chemical energy (CJ) detonation was Adaptive Mesh Refinement (AMR) by Berger and Colella (1989) was used to follow the convective dynamics in the turbulent

released in the charge by an ideal Chapman-Jouguet

wave (Taylor,

Hence, the flowfield in the charge was assumed to be that of a spherical CJ charge are (Dobratz, 1974)

detonation at the time when the wave reached the edge of the charge. The CJ parameters for a PBX-9404

PCJ = 370 kbars; PCJ = 2.485 g/cm3; ecj = 8.142 x 1010 erg/g; qcJ = 5.543 x 1010 erg/g; ucj = 2.28 krn/s; F = 2.85

To de-singularize charge surface,

the problem, and to qualitatively model the effects of HE gains on the a linear gradient region was added to the outside of the charge to In this region (1.0 < r/R o < 1.10),

transition from the CJ state to the ambient state.

Gaussianly distributed random fluctuations were, added to the density and energy profiles. This also served as a "white noise" seed for instabilities on the f'treball interface

An r-z cylindrical coordinate grid was used. To save computational quarter-space region of the spherical blast wave was actually calculated.

time, only a

Three levels of

AMR grid refinement were used: a base grid, with a mesh size A1 that extended to 100 Ro, was used to capture the shock; a transitional grid A2; and fine grid A3 to follow the mixing. The refinement strategy was based on following the detonation products region (i.e., where C _<1). To check for convergence, both a medium and a fine-grid calculation were performed. The mesh sizes are listed in Table 1. Table 1
IIII

ml

ii

II

CPU (hours) Case 1-medium zoning 2-fine zoning A1/Ro 0.90 0.45 A2/Ro
Ill II

A3/Ro
I IIIII Illll

t = 62 ms 5 30

t = 125 ms 10 -

0.15 0.075

0.0250 0.0125

The calculations were run out to a time of 62 ms or 125 ms (i.e., 2 to 4 times the positive phase duration of the blast wave). This took between 5 and 30 hours of CPU time on a Cray Y-MP computer, depending on grid resolution. next section. RESULTS The results are described in the

A. Flow Visualization The next four figures present a sequence of frames that show the evolution of the blast wave in cross-section. shocks and mixing processes. Figure 2 depicts the evolution of the flow during the strong blast wave phase. It shows that the detonation products rapidly expand, forming the main blast wave shock S and a secondary backward facing shock I that is swept outward by the blast wave flow. The f'u'eball interface is located in the region between these two shocks. The fireball interface is impulsively accelerated when the detonation wave reaches the edge of the charge. decelerated Thus, the interface is flu'st subjected to impulsive accelerations (RM) instabilities. As the fireball expands, it is air, and it is thus subjected to Rayleigh-Taylor that can lead to Richtmyer-Meshkov by the shock-compressed (RT) instabilities. In particular, density contours are used to visualize the

By a time of 2.4 ms, perturbations are clearly visible on the fireball

interface (Figure 2a), which has expanded to about 8 charge radii. Perturbations in this calculation originate from discretization errors resulting from the projection of the spherical interface onto the square r-z mesh. In the physical experiment, perturbations arise either from granular irregularities or from molecular fluctuations. will grow exponentially difference grid. with time. Eventually on the charge surface, on the finite

According to linear stability analysis, such perturbations they will be resolvable

Here we are interested in the late-time phase where there are strong This leads to a similarity spectrum of frequencies, Hence, for our purposes loss of memory of the initial perturbations.

interactions between various modes. and a concomitant


,11

the cause of the initial perturbations is not particularly important.

By a time of 3.9 ms (or about 12 charge radii), the perturbations have evolved into mushroom-shaped structures that are characteristic of RM and RT-driven mixing layers By a time of 9.8 (KH) instabilities interface cause a (Figure 2b). These structures rapidly evolve into the nonlinear regime. nonlinear cascade process (Figure 2d). radial. Also, Kelvin-Helmholtz on the fireball

ms (or about 17.5 charge radii), they have generated new fine-scale structures through a develop due to shear flow along the bubble walls Note that pressure gradients in the blast wave are essentially misalignment mechanism: Perturbations of pressure and density gradients, that creates vorticity by the baroclinic

= -V(l/p) vp
This leads to a self-induced production of vorricity in the mixing region.

(7)

Figure 3 depicts the evolution of the flow during the implosion phase. starts propagating increasing back toward the origin.

At a time

of about 12 ms, the secondary shock I falls into the negative phase of the blast wave, and It implodes at a time of about 24 ms. This draws the inner boundary characteristic of the mixing layer back toward the origin, thus greatly During this phase, the flow exhibits many of features

the mixing width.

of imploding spherical shells (e.g., elongated bubbles with KI-I structures

on the bubble walls, perturbations on the imploding shock shape, etc; see Klein et al., 1993).

The next two figures depict the evolution of the mixing in the f'trebaU during each of the four characteristic internal energy--where phases of the mixing. shock-compressed Colors are used to visualize various aspects of the mixing processes. For example, color bar (Figure 4c) is used for air air is red and the ambient air is blue, while

f'treball gases are yellow. The green and white lines represent vorticity contours (positive and negative, respectively) dilatation contours which help to visualize the mixing, while black lines represent Figure 4a depicts a cross-sectional which make the shocks visible.

view of the flow at the end of the strong blast wave phase (t = 12 ms). Figure 4b depicts the flow field at the time shock I implodes (t = 24 ms). Clearly, the mixing extends all the way to the origin.

The implosion of shock I creates a second explosion shock S' that expands outward from the origin. This re-shocks the mixing layer, and creates additional

vorticity through baroclinic fireball (Figure 5a).

effects.

At a time of 63 ms, shock S' emerges from the of fine-scale mixing is evident. Shock S'

The intensification

continues to expand, and leaves behind a turbulent fh'eball. This leads to an asymptotic mixing phase that is shown in Figure 5b (t = 126 ms). It is characterized by a continued cascade process among the various scales of vonicity, fireball interface. fireball size remains essentially Figure 5b. B. Growth of th, Mixing Laver e In order to study the growth of the mixing region, we found it useful to follow the evolution of certain mean concentration lines. Three lines were selected '
m

and a continued folding of the has ceased, the

Since the mean radial expansion of the explosion constant.

However, the mixing processes continue to

generate acoustic waves within the f'treball which are quite evident as fine black lines in

1. The line R0.9(t) where C = 0.9, which characterizes the inner boundary of the mixing region; 2. The line R0.5(t) where C = 0.5, which represents the center of the mixing region; and 3. The line R0.1(t) where C = 0.1, which characterizes the outer boundary of the mixing region. These are depicted in the wave diagram for an HE-driven blast wave (Figure 6). Four distinct phases are evident: (i) a blast wave phase--where the mixing region is swept stretches the inner reoutward by the shock-induced flow; (ii) an implosion phase--that

boundary of the mixing region back toward the origin; (iii) a re-shocking phasemthat the fireball size remains essentially constant. Let us define the mixing width by the relation: 3(t) / Ro = [R0., (t)- Ro.9 / R0 (t)] (8)

energizes the mixing layer by RM effects; and (iv) an asymptotic mixing phase--where

The evolution of the mixing width is shown in Figure 7. During the blast wave phase, the mixing width grows linearly in time:

_(t) / R0 = 0.396t

(0 < t < 10 ms)

(9)

with a velocity of 396 rn/s. This is consistent fireball interface (i.e., a RM growth). exponent increases to a value of 1.44: / R0 = 0.151t
TM

with an impulsive acceleration

of the

During the implosion

phase, the power-law

(10ms < t < 32 ms)

(10)

as a result of the inward stretching of the interface back toward the origin. Mixing occurs in a strong gradient of velocity (i.e., acceleration is not constant), so the quadratic growth associated with RT mixing is not achieved. is compressed During the re-shock phase, the mixing layer from 8/Ro - 22 to 8/Ro ---16 by the shock S'. During the asymptotic

phase, the mixing width is essentially constant : 8/Ro -- 21.

C. Azimuthal Averaging In order to study the profiles in the mixing region, it is convenient to define an

averaging procedure. In this problem, azimuthal averaging is most appropriate. The flow field variables _, were stored as a function of r, z, and t, that is _(r,z,t). One can define a polar-coordinate system in the r-z plane: R = _Jr2 + z2 (11)

0 = Tan-_(z / r) and convert the flowfield tG the ilew variables, i.e., O(R,0,t). an azimuthal average of the flowfield by integrating over 0 :

(12) It is then possible to define

(R,t) = 2 f,r,_ _(R,O,t)dO


,10

(13)

Given a mean, one can then calculate r.m.s, fluctuations about the mean from the relation:

10

(_')2 =--J0 _/2 [_(R'O't)--_(R't)12dO 2

(14)

and

"

_' (R,t)= __

(15)

D. Averaged Profiles In order to check for similarity, the profiles were scaled with the mixing width 8: r/= (R- Ro.5) / _ (16)

In addition,

the flowfield

was nondimensionalized

by the ambient

atmospheric

conditions, denoted by subscript .o. Figure 8 presents the azimuthally-averaged mean flow profiles across the mixing profiles

layer, at the end of each of the four phases of mixing. The mean concentration of the validity of the scaling procedure. spread over a larger and larger volume. pressure gradient at the fireball that the azimuthal demonstrates

tend to collapse to a single curve, independent of time. This provides indirect verification The mean density profiles peak at the fireball Figure 8c indicates that there is a mean radial Figure 8d interface at a value of ~ 7/9. at early times, and then decay with time as that mass is

interface during the blast wave phase.

averaging procedure is adequate to produce smooth,

well-behaved mean radial velocity profiles. Shock S and I axe indicated on the figure. Of course, the magnitude of the velocity decays with time. The mean azimuthal velocities
m

are typically less than 0.1 u, and they oscillate about a mean value of zero--as one might expect. Figure 8f displays the azimuthally-averaged vorticity profiles which decay with time (Figure 8f), by means of a cascade process. The corresponding . . well-behaved concentration percent. profiles r.m.s, profiles across the mixing layer are presented in Again the around 40

Figure 9. Again note that the azimuthal-averaging of the second moments profiles are approximately

procedure is adequate to give smooth, of the flow variables.

self-similar, with peak fluctuation

Again, density fluctuations

decay with time, as do the pressure fluctuations. smooth, and have a trimodal

The radial velocity fluctuation

profiles are surprisingly 11

character, with peaks at the inner and outer edges of the rrrxing layer and a peak at the center of the layer. Even the azimuthal velocity profiles are relatively smooth and wellbehaved. Hence, we conclude that the azimuthal-averaging independent of time: procedure is adequate to give smooth, well-behaved profiles. The following profiles are approximately self-similar and c, p, p, c', p', p', u', and v'. Profiles exhibiting a marked decay are: the mean radial velocities u (due to blast wave decay), and vorticity _ and co' (as a censequence of a cascade process). MIXING PROPERTIES

In order to check the global properties

of the mixing processes,

the various

components of the kinetic energy were integrated over the mixing volume, in particular, the kinetic energy of the mean flow k (t), the kinetic energy of the fluctuation flow k"(t), and the total kinetic energy K(t), were calculated from the following relations:

3-- 3 J0"" u(R,t) + 2R. f

(17)

, 3.._fR, v' t)2]RZdR k' (t)= 2t?,3, o [u'(R't)2 + (R, d

(18)

._3 fR.
K(/)-'d_j fo _

[u(R,O,t)2+v(g,o,l)2]R2dOdR

(19)

= k(t)+k"(t)

(20)

where a value of R, = 40 Ro was used. The results are presented in Figure 10. It shows that the kinetic energy reaches a peak value of 6.5 x 108 (cm/s) 2 at t _.= ms, or about 6 halfway through the blast wave phase. The mean kinetic energy then decays as the

interface does work on the ambient atmosphere, and eventually approaches zero at late times. This is, of course, consistent with the finite energy of the explosion, which implies that the velocity field of the blast wave must eventually decay to zero. The fluctuating kinetic energy k" reaches a peak value of 3.1 x 107 (era/s) 2 (or about 10 percent of the
UlE

instantaneous k) during the implosion phase. It decays slightly during the re-shock phase, and finally reaches a constant asymptotic value of k"** = 2.2 x 107 (crrds)2; this is

12

consistent with a residual fluctuating velocity of about 66 m/s that continues to mix the fluid in the fireball.
B

The total kinetic energy, which represents the sum of the two approaches k" at late times.

components, tracks k at early times and asymptotically The fluctuating kinetic energy corresponds

to the rotational energy of mixing

induced by the vorticity distribution.

The persistence of fluctuating kinetic energy at late

times may be explained by considering the mechanical energy equation:

d K -u (Vp / p +vu. (Wu


dr (21)

where the overbars present calculations

denote averaging over the computational

volume.

The pressure Since the Hence,

gradients rapidly disappear and can no longer influence the kinetic energy. are inviscid, shear work cannot affect the kinetic energy.

conservation of mechanical energy implies that the rotational kinetic energy approaches a constant for inviscid blast waves. In the present calculations, the asymptotic value of this constant is K**- 2.3 x 107 (cm/s)2 m or about 3.5 percent of its peak value at early times. From a numerical point of view, this persistence of fluctuating kinetic energy demonstrates that Godunov-AMR times. Three-dimensional algorithm is non-dissipative. From a physical point of in

view, additional dissipation is required so that the kinetic energy approaches zero at late (3-D) calculations have shown that vorticity amplification hairpin structures leads to the continued decay of kinetic energy (Bell and Marcus, 1992). Since the velocity or kinetic energy fluctuations are driven by the vorticity field, it is useful to consider the global properties of the vorticity, namely:

===== _R3 _0_. ,,/2 w(R,O,t) 6 f co(t) Jo

R 2dOdR

(22)

I_l
r

6 = _R. SR.f,,/2 leo _-z-mo Jo (R,O,t)i R2aOdR

(23)

==_=co(t)2 ='_-6

f0_ 00['_/2 co2(R'O't)R2 dOdR

(24)

13

where the overbars denote integration

over the mixing volume, and the vorticity was

calculated as a finite difference approximation to the curl of the velocity field.


m m

Figure 11 presents the evolution of the vorticity co(t) averaged over the mixing region 0 <_R _<40 Ro. The calculation Although vorticity is baroclinically mushroom-capped essentially zero 1. To demonstrate shows that the net vorticity remains near zero. pairs that form the one must generated in positive-negative

structures seen in Figures 2-5, the net amount of vorticity remains that vorticity is not zero everywhere,

consider the magnitude of vorticity

la, or I

its energy co2. Figure 11 shows that vorticity Figure 12

reaches a peak value during the implosion phase (t = 22 ms) and then decays.

shows that o92 decays as t-2 at subsequent times. Evidently vorticity in this calculation corresponds to a function that fluctuates about a zero mean, with the magnitude of the fluctuations decaying with time, while the kinetic energy K remains constant. The decay of the vorticity parameter _(t): is best illustrated by considering the enstrophy-like

E(t) ---1/ _0.5_--'_

(25)

whose evolution

is shown in Figure

13.

Indeed, the enstrophy

parameter

in this

calculation increases with time, and may be approximated by the linear relation: E(t) = 0.005 + 0.0345t (t> 10ms) (26)

that is shown as the straight line in this figure. This is similar to our previous results (Kuhl et al., 1993) for interface instabilities in high-altitude explosions, where e(t) = 0.15 + 0.046 t. (27)

The present results have essentially Batchelor (1969): E(t) = 0.134 + 0.0323t.

the same slope as the enstrophy curve derived by

(28)

14

His results were based on 2-D simulations of incompressible, turbulence in a box, as performed by R.W. Bray. Batchelor

isotropic, homogeneous showed that this linear

growth in time was associated with a cascade process for to2. Although the kinetic energy was constant, w2 decayed as 1/t2. He found that the spectrum of t02 had a slope of k"1 in the inertial range 2, and that this spectrum leads to an overall decay in t02. Evidently there is also a cascade process in the present calculations. small scales which grow with time. Additional fine-scale Vorticity starts at

structures are continuously

created by a folding mechanism in 2-D flows, analogous to the re-generation of fine-scale structures by hairpin vorticies in three-dimensional in volume-averaged SUMMARY vorticity--even (3-D) flows. This leads to a net decay though the kinetic energy remains constant.

AND CONCLUSIONS

The fireball interface was unstable, and rapidly evolved into a convective mixing layer. During the blast wave phase the mixing width grew linearly with time by the 8/Ro = 0.396 t, while during the implosion phase it grew with a power-law function of time 8lKo = 0.151 t 1.44 as a result of the inward stretching of the interface implosion shock. At late times it attained an asymptotically-constant The azimuthal-averaging procedure was adequate width of 8/Ro = 21. well-

to produce smooth,

behaved profiles in the mixing layer. The profiles scale with the mixing layer thickness, i.e., according to rl = (R-R0.5)/8; some of the profiles were self-similar (i.e., independent of time). The mean kinetic energy of the blast wave rapidly decayed to zero, but the fluctuating kinetic energy approached a constant (about 0.035 Kmax) at late times. This small residual value represents the rotational energy associated with turbulent mixing, and is driven by the vorticity field. Three-dimensional calculations are required to dissipate this small residual energy. Vorticity is rapidly generated by baroclinic effects, and then evolves through a cascade process by convective mixing. For example, vorticity is created at fine scales that grow to larger and larger scales by vortex merging. same time, fine-scale structures are regenerated whereby vorticity gradients are amplified by velocity gradientstsimilar At the in 2-D by stretching of material lines, to vortex-line

stretching in 3-D flows (see Batchelor, 1969). This leads to a vorticity distribution that decays with time because of a cascade process. These results are in agreement with other

15

2-D simulations

of turbulent mixing for incompressible

flows (Batchelor,

1969) and

compressible shear layers (Vuillermoz and Oran, 1992). The HE-driven blast wave problem considered here provides an opportunity investigating wave breaks a variety of fluid-mechanic out of the charge effects related to interface instabilities spherical geometry. Initially the interface is accelerated impulsively and instabilities for in

as the detonation

grow by the Richtmyer-Meshkov structures resemble those found in shock S'

mechanism during the blast wave phase. During the implosion phase, the mixing layer is
I

drawn back toward the origin, and the mixing imploding spherical shells.

During the re-shocking

phase, the secondary

interacts with the density structures in the f'n'eball thus generating additional fine-scale structures by the baroclinic mechanism. What remains after shock S' leaves the f'n-ebaUis This can late in spherical mixing layers to arbitrarily a vorticity distribution that evolves according to convective mixing dynamics. be used to study the decay of turbulence times.

Turbulence recalculate

is, or course,

a three-dimensional

phenomenon.

We plan to

this problem in 3-D to study the evolution of turbulence in spherical mixing

layers. In addition, more detailed experimental data are needed to check the accuracy of the numerical simulations.

16

ACKNOWLEDGMENTS Work performed under ihe auspices of the U. S. Department of Energy by the Lawrence Livermore National Laboratory under contract No. W-7405-Eng-48. Also sponsored by the Defense Nuclear Agency under DNAIACRO#93-824, and Work Unit 0000I.

17

REFERENCES

Anisimov,

S.I., and Zel'dovich,

Y.B. (1977). Rayleigh-Taylor

instability

of

boundary between detonation products and gas in spherical explosion. Eksp. Teor. Fiz. 3, 1081-1084.

Pis'ma Zh.

Anisimov, S.I., Zel'dovich, Ya.B., Inogamov, M.A. and Ivanov, M.F. (1983). The Taylor instability of contact boundary between expanding detonation products and a surrounding gas. Shock Waves, Explosions and Detonations, J.R. Bowen, J.-C. Leyer, and R.I. Soloukhin, eds., Progress in Astronautics and Aeronautics Series,

87, ( M. Summerfield, ed.), AIAA, Washington, D.C., pp. 218-227. Anisimov, S.I., and Inogamov, N.A. (March 1993). Studies in Physics of High

Density Energy, Landau Institute, Moscow.

Batchelor, G.K., two-dimensional,

(1969). Computation of the energy spectrum in homogeneous turbulence, Phys. of Fluids (Supplement II), pp. I1233 to II 239. Self-Similarity and Intermediate Asymptotics

Barenblatt, G.I. (1979). Similarity,

(N. Stein, translator), M. Van Dyke, ed., Consultants Bureau of Plenum Publishing, New York.

Barenblatt, G.I., Gurguis, R.H., Kamel, M.M., Kuhl, A.L., and Oppenheim, A.K. (1980). Self-similar explosion waves of variable energy at the front. J. Fluid Mech., 99 (4), 841-858. Bell, J.B. and Marcus, D.L. (1992). Vorticity intensification and transition to

turbulence in the 3-D Euler equations, Comm. Math. Phys., 147, 374-394. Berger, M.J. and Colella, P. (1989) Local Adaptive Mesh Refinement for shock

hydrodynamics. J. Comput Phys. 82, 64-84. Brode H.L. (1955). Numerical solutions of spherical blast waves, Applied Physics, 26 (6), 766-775. Journal of

18

Brode, H.L. (1959). Blast wave from a spherical charge, Physics of Fluids, 2 (2), 217-229.

Chorin, A.J. (1975). Lectures on Turbulence Theory, of dishonesty", pp. 21-23). .

Mathematics Lecture Series

5, Publish or Perish, Inc., Boston (vid. esp. Chapter I-3: "An example of the perils

Chorin, A.J. (1986). Vortex methods for the study of turbulent combustion at low math number, Dynamics of Reactive Series, Systems, (Part 1: Flames ed.), and Configurations), Astronautics J.R. Bowen, J.-C. Leyer, and R.I. Soloukhin, and Aeronautics 105 (M. Summerfield, eds., Progress in AIAA,

Washington, D.C., pp. 14-21. Colella, P. and Glaz, H.M. (1985) Efficient solution algorithms for the riemann 59, 264-289. Mach reflection ed.),

problem for real gases, J. Comput. Phys,

Colella, P., Ferguson, R.E., Glaz, H.M., and Kuhl, A.L., (1986). from an HE-driven blast wave, Progress in Astronautics Dynamics and Aeronautics-Series,

of Explosions and Reactive Systems, 106, ( M. Summerfield,

AIAA, Washington, D.C., pp. 388-421. Davydov, A.N., Lebedev, E.F., and Perkov, S.A. (1978). Gasdynamic instability in propagation of cylindical shock waves. Detonation. Critical Phenomena: Chemical Transformations 76-79. in Shock Waves, Chemogolovka, PhysicoOIKhF AN SSSR, pp.

Davydov, investigation

A.N.,

Lebedev,

E.F.,

and Perkov,

S.A. (1979).

Experimental the

of gas-dynamic

instability

in the a plasma flow following

cylindrical shock wave. Preprint N 1-40, Institute of High Temperatures, USSR, p. 26.

Dobratz, B. (1974). Properties of Chemical Explosives and Explosive Simulants, UCRL-51319, Lawrence Livermore Laboratory, Livermore, CA. Gilmore, F.R. (1955). Equilibruim Composition and Thermodynamic Air to 24000 K, RM-1543, Rand Corp., Santa Monica, CA. Properties of

19

Klein, R.I., Bell, J. Pember, R., and Kelleher, T., Numerical Workshop:

simulations

using

Adaptive Mesh Refinement hydrodynamics of a shock-driven mixing layer, 4th Int. Physics of Compressible Turbulent Mixing, 29 March1 April, 1993, Cambridge, England, (in press). Korobeinikov, V.P. (1985). Problems of Point-Blast Theory, (Translated by G.

Adashko), American Institute of Physics, New York. Kuhl, A.L., Bell, J.B., Ferguson, R.E., White, W.W. and McCartor, T.H. (1993). Turbulent Mixing in High-Altitude Explosions, UCRL-JC-11818, Lawrence Physics Livermore National Laboratory, Livermore, CA; also in 4th Int. Workshop: (in press). Marcus, National D.L., Puckett, Laboratory, E.G., Bell, J.B. and Saltzman, Interfaces, UCRL-JC-108720, CA; also Livermore, J. (1991). Numerical Lawrence Livermore Phys. Workshop:

of Compressible Turbulent Mixing, 29 March - 1 April, 1993, Cambridge, England,

Simulation

of Accelerated

in 3rd Int.

Compressivle Turbulent Mixing, Royaumont, France, 1991. Meshkov, E.E. (1960). Instability of the interface of two gases accelerated shock wr,ve, Izv AN SSRE Mekhanika Zhidkosti i Gaza, 4 (15), 151-157. yon Neumann, J., and Goldstine, H. (1955). Blast wave calculation, Communication on Pure and Applied Mathematics, 8, 327-353; reprinted in John yon Neumann Collected Works, Taub, A.H., ed., Vol. VI, Pergamon Press, New York, 1963, pp 386-412. by a

Oppenheim, A.K., Kuhl, A.L., and Kamel, M.M. (1972). A parametric study of selfsimilar blast waves. J. Fluid Mech. $2(4), 657-682. Oppenheim, A.K., Kuhl, A.L., and Kamel, M.M. (1972). On self-similar waves, headed by the Chapman-Jouguet Oppenheim, A.K. (1986). The blast

detonation J. Fluid Mech. 55 (2), 257-270. of combustion fields, and their

beauty

aerothermodynamic

significance, Dynamics of Reactive Systems,

(Part 1: Flames

20

and Configurations), Astronautics

J.R. Bowen, J.-C. Leyer, and R.I. Soloukhin, eds., Progress in Series, 105 (M. Summerfield, ed.), AIAA,

and Aeronautics

Washington, D.C. pp. 3-13. Richtmyer, R.D. (1960). Taylor instability in shock acceleration of compressible fluids, Comm. Pure Appl. Math, 13, 297-319. . Ruppert, V. (1991). Shock-interface interaction: current research on the RichtmyerMeshkov problem, Shock Waves, Vol. 1, K. Takayama editor, Springer-Verlag, Berlin, pp.83-94. Stanyukovich, K.P. (1955). Unsteady Motion of Continuous Media, Gostekhizdat,

Moscow; English translation, edited by M. Holt, Pergamon Press, New York, 1960. Sedov, L.I. (1946). Rasprostraneniya sil'nykh vzryvnykh voln, Prikladnaya

Matematika i Mekhanika, 10,(2), 241-250. Sedov, Friedman L.I. (1959). Similarity and Dimensional Method in Mechanics, M.

translator, M. Holt editor., 4th Printing, Academic

Press, New York,

Chapter IV. Taylor, Sir G.I. (1941) The formation of a blast wave by a very-intense explosion, first published London, in British Report RC-210; revised version in Proc. Royal Soc., A 201 (1), 1950, 175-186.

Taylor, G.I. (1950). The dynamics of the combustion products behind plane and spherical detonation fronts in explosives, Proc. Royal Soc., A 200, 235-247. VuiUermoz, P. and Oran, E.S. (1992). Mixing regimes in a spatially-confined, dimensional, supersonic shear layer, (submitted to Phys. of Fluids). . Youngs, D.L. (1984). "Numerical simulation of turbulent mixing by Rayleightwo-

Taylor instability," Physica, 12D, pp. 32-44. Zerdovich, Temperature Ya. B., and Raizer, Yu. P. (1966). Physics of Shock Waves and HighHydrodynamic Phenomena, 2nd ed., Izdatel'stuo Nauka, Moscow;

21

English translation: Hayes, W.D., and Probstein, R.F., eds., Academic Press, New York, Vols. I and 2, 1967.

22

FOOTNOTES , 1. I.e., zero, to within a few percent of [co[. Nonzero values arise due to asymmetries in the vortex pairings, finite-difference errors in computing V x u, and anomolies arising for the r - z form of the governing equations. Note: The corresponding slope for 3-D turbulent is k "5/3.

2.

23

CAPTIONS Figure I. Shadow photograph showing a cross-sectional view of an HE-driven blast wave at t = 0.2 ms. A 0.45-g hemispherical charge of Nitropenta was mounted on the explosion chamber window, and centrally-detonated by an intense electrical discharge. (Courtesy of Dr. H. Reichenbach of the Ernst Mach Institut, Freiburg, Germany; 1993).

Figure

2. Growth of instabilities on the fireball interface during the blast wave expansion phase (log p contours): (a) t = 0.31 ms (b) t = 1.1 ms; (c) t = 2.4 ms; (d) t = 3.9 ms; (e) t = 5.8 ms; (f) t = 7.8 ms; (g) t = 9.8 ms.

Figure 3. Inward stretching of the mixing region during the implosion phase (log p contours): (a) t -- 12 ms (b) t ,,- 14.5 ms; (c) t ---17 ms; (d) t ---20 ms; (e) t = 22 ms; (t3 t = 23 ms. Figure 4. Cross-sectional view of the mixing in a HE-driven blast wave: (a) at the end of the strong blast wave phase (t = 12 ms); (b) at the end of the implosion phase (t = 24 ms); (c) color scale for air internal energy. Note that yellow denotes the detonation products gases in the fireball, shock-compressed air is red, while the ambient air is blue. The green and white lines represent vorticity contours (positive and negative, respectively) which help visualize the mixing, while the black lines represent dilatation contours which make the shocks visible. Figure 5. Cross-sectional view of the flow field at later times: (a) at the end of the reshocking phase (t = 63 ms); (b) in the asymptotic-mixing phase (t ---126 ms). Color scheme is the same as the previous figure. Figure 6. Figure 7. Wave diagram for a HE-driven blast wave. Growth of the mixing region.

Figure 8. Evolution of the mean-flow profiles in the mixing region: (a) concentration; (b) density; (c) pressure; (d) radial velocity: (e) azimuthal velocity; (f') vorticity. Figure 9. Evolution of the fluctuation profiles in the mixing region: (a) concentration; (b) density; (c) pressure; (d) radial velocity; (e) azimuthal velocity; (f) vorticity.
111

Figure 10. Evolution of the kinetic energy in the region R < 40 Ro: mean component fluctuating component k", and total K = k + k". Figure 11. Evolution of volume-averaged Figure 12. Evolution of volume-averaged vortieity in the fn-eball. _ in the f'trebaU.

k,

Figure 13. Evolution of the enstrophy parameter I_ in the mixing region. Data points denote the present calculation and their fit: E (t) = 0.005 + 0.0345 t.

24

Figure 1.

Shadow photograph showing a cross-sectional view of an HE-driven blast wave at t = 0.2 ms. A 0.45-g hemispherical charge of Niwopenta was mountedon theexplosionhamberwindow, and cenn-ally-<letonated c by an intense electrical discharge.Courtesy ( ofDr. H. Reichenbach ftheErnst o Mach Insfimt, Freiburg, Germany;1993).

Figure 2. Growth of instabilities on the fireball interface during the blast wave expansion phase (log p contours): (a) t = 0.31 ms (b) t = I.I ms; (c) t = 2.4 ms; (d) t = 3.9 ms; (e) t = 5.8 ms; (f) t = 7.8 ms; (g) t = 9.8 ms.

Figure3. Inwardstretching the mixingregionduring of theimplosiori phase(log p contours): t= 12ms (b)t= 14.5 (a) ms; (c) 17ms; (d) 20 ms; (e) 22 t= t= t= ms; (t') 23 ms. t=

"tU.

20.

10.
I

Figure 4. Cross-sectional view of the mixing in a HE-driven blast wave: (a) at the end of the strong blast wave phase (t - 12 ms); (b) at the end of the implosion phase (t - 24 ms); (c) color scale for air internal energy. Note that yellow denotes the detonation products gases in the fireball, shock-compressed air is red, while the ambient air is blue. The green and white lines represent vordcity contours (positive and negative, respectively) which help visualize the mixing, while the black lines representdilatation contours which make the shocks visible.

Figure 5. Cross-sectional view of the flow field at later times: (a) at the end of the reshocking phase (t = 63 ms); (b) in the asymptotic-mixing phase (t = 126 ms). Color scheme is the same as the previousfigure.
e

Figure 6. Wave diagram for a HE-driven blast wave.

ii .+.

uo._ Su_t_ oq_.;o _o.z D

"L_!I!.4

oot

o_

c_,

c_

ol,

tr

i ! ' '_ r _ .tL i-,_ i _i _!'_[+:t !__!.!_4-/,.' ................. I ;-,-!----_-_i,,+,,, _u_wm, Ir>,:__l[-_2_J:m,_F-I ] +^_,u.._:-_ _ l_.-.,:..-t-:: ,--_
......... ..... i" -t'i' '

.L' '1 ,'!" I:,,v_


':i ! t '

--

..... : ,,--

.,: .....
..

"+':.. ._
:..... ,:,....,-,-

-,-l,

,,

, t
--,+.

I ..<_J+,, i
+_ .I 'i'::_"----_

'

+,,

:-::

t+l....

+It_l

i '

+ +

'__+------+

"........

"":" i"+

'+ ++:_--.r-'i"-:t-- --!- - +' ' ._ +++.+,._..;._ ='--"--=++ i_" Z :, __.:..+.., .+-'_.+ ._+---___ +.+I'..- ;"_'+' + ,_ I :' + :_

_I_ ''+
+

.... _+++....
---'--I_'"'" r," i_ : ':+" '+i | --o

;.! ,,i _" .._11, ,I "" ' _+

i_

,--"i"--.-"i-.' ,,, ,+ i i

,,

+
,_

.Li ;'+' i r, .+_., --,.._++'_---.-'-:-" _iIi' ;;,1 !i'.,,i + ---'," ' ,'t+ '! _'

,,,+,,
+'E

_ .:-_...-i I t

I",-l + + ":b++il:, ,+_+!"- I_:+. .;.'..,:+.;+:+- !ii:+*+._ . +, +:.'-+-+

v_

4t_

,4

"l

L '
Jl
i i

"*
0a,

w'--|.I "l,O

i
__
LI t.O i._.

'
1

,,.,
'i,i ,l.| O,O

'_24._/t
I,I . *

'I
d i iiii _ ii _ iii!ii __ i i i i __ iii iii Q i 'q-_ e,

"L
}

-" 0 "" .;
, :
a.e *_,e

Iz 6.,.I

.s,.: b._ _
0

Ib}
!I
*l Ike _.e

iI .............. (e)
,.
e

12_ Z4

i!

! '
1,:

:'"a.o

,,,,

Ij
._.e

.....
o.o _.o-

A
a.:

i O

ql O

"l.O

"l,O

O.O

1,0

1.,:

' " Figure 9. Evolution of the r.rn.s, fluctuation profiles in the mixing region: (a)concentration;density; p_ssm-c: (d) (b) !c.) radialeloci_; v (=)azimu_ velocity; vomc2_. (f)

Figure10. Evolutionof thekineti energyin theregon R _ 40 Ro: meancomponent_, c nucmaungcomponentk, andtotalK = _:+ k".

Figure 13.Evolution of the cnsu'ophy parameterg in the mixing r_&,ion. Data points denote the presentcalculationand theirfit: g (t) --0.005 . 0.0345 t.

o o

W ql

,J
dP

Anda mungkin juga menyukai