Anda di halaman 1dari 58

UCRL-JC-111404 PREPRINT

JUl 191893

I
A Baroclinic Model of Turbulent Dusty Flows
9

A. L. Kuhl

This paper was prepared for submittal to the DNA Numerical Methods Symposium Stanford Research Institute April 28-30,1992 Menlo Park, California

April

1992

... ..!:! . :.,::.:.: : .. '": :C_::'..,. .,..,..,.,..... ,i:i::'i.::::: ::: ::- : "':

.,_

Thisis apreprint ofa paperintended forpublication tna journal orproceedings. Since changes may be made before publication, this preprint is made available with the understanding thatit will not be cited or reproducedwithout the permission of the author.

: :

DISCLAIMER This document was prepared an a.ccount as ofwork sponsored an agencyof by theUnitedStatesovernment. Neither G theUnitedStatesovernmentnor the G Unive_ity ofCalifornia noranyoftheirmployees,akes any _warranty, e m express or implied, r assumes any Iegal o liability responsibility or tortheaccuracy, completeness, orusefulness ofany information, apparatus, product,r process o disclosed, orrepresents its that usewould not infringe privately ownecl, right.s. Referenceerei/_ h toany specific commercial product,rocess, p orservice trade by name, trademark,anuf/icturer, m orotherwise, doesnotnecessarily constitute or imply itsendorsement, recommendation, favoring the"United or by St.ates Government or theUniversity ofC:a. lifornia. viewsand opinions fauthors The o expressed herein notnecessarily or reflect do state thoseoftheUnitedStates Government ortheUniversity ofCalifornia, shall otbeusedfor and n advertising orproduct endorsement purposes.

'

ml,l_!

A BAROCLINIC

OF TURBULENT DUSTY A.L. Kuhl Lawrence Livermore National Laboratory E1 Segundo, California ABSTRACT

MODEL

FLOWS*

The problem explosions ,f layer height such surfaces; mass becomes While models, calculates others

considered

here is the numerical

simulation

of the turbulent

dusty flow induced

by

over soil surfaces.

Some of the unresolved

issues are: (i) how much dust is scoured layer; (3) what is the dusty layer profiles;

from

(2) where does the dust go in the boundary versus time; (4) what are the dusty boundary entrained

boundary cloud? closure actually

(v) how much of the dust

into the dust stem; and (6) where does the dust go in the buoyant models based on mean-flow equations and turbulence variations that

have used numerical

here we propose the turbulent

a Baroclinic

Model for flows with large density of dust on an adaptive

mixing and transport

grid. The model is based on shock fluidization of the dust and (5) the lead to the the no-slip condition);

the following idealizations: (4) the dust-air turbulent an inviscid Godunov turbulent mixing algorithm mixture

(1) a loose dust bed; (2) an instantaneous is treated as a continuum

layer; (3) the dust and air are in local equilibrium is dominated

(so air viscosity enforces vorticity.

dense fluid with zero viscosity; which are solved by means (AMR)

by baroclinically-generated laws for the mixture, Adaptive

These assumptions

set of conservation mixing processes

of a high-order

for gasdynamics. on the grid.

Mesh Refinement

is used to capture

One of the unique characteristics by an inviscid, shock and bed; mixing interactions the fireball. layers. turbulent baroclinic with strong These shear The rotational mixing distribution. and

of these flows is that mixing occurs because 0) -_p _p. For example, unstable occurs via the density structures gradients such as the fluidized l,_yer entrain occurs because bed,

vorticity

is produced is created by layer,

mechanism:

vorticity

the thermal

layers are hydrodynamically in the boundary then generation of the dust vorticity structures),

and roll up into turbulent dust velocity from the fluidized field induced by pressure

transport Additional

the vorticity gradients gradients A number teractions conclusion the fluidized

of large-scale forces.

(e.g., during the positive (induced by the rotational

and negative

phase of the blast wave), because of local pressure and because these of buoyancy baroclinic

of examples

will be presented layers and dust

to illustrate beds,

effects including

shock

inThe

with dense-gas of these studies

and dusty

wall jets of airblast

precursors.

is that dusty boundary

layers grow because as proven

of mass entrainment

from

bed (and not because

of viscous wall drag)

by the Mass Integral

Equation.

*Work National Nuclear

performed Laboratory Agency

under

the auspices Contract

of the U.S. Dept. number #92-824

of Energy

by the

Lawrence

Livermore

under

W-7405-ENG-48.

Also sponsored

by the Defense

under DNA IACRO

and Work Unit 00001.

1. INTRODUCTION The problem considered here is the numerical simulation of the turbulent dusty flow induced by explosions over soil surfaces. Figure 1.1 presents a photograph of Event PRISCILLA, which provides an excellent example of the type of dusty flow to be modeled here. Some of the unresolved technical issues associated with such flows are: (i) How much dust is scoured from soil surfaces? (ii) Where does the dust go in the boundary layer? (iii) What is the dusty boundary layer height versus time? (iv) What are the dusty boundary layer profiles? (v) How much dust mass becomes entrained into the dust stem? (vi) Where does the dust go in the buoyant cloud? We have tried to answer such questions by means of direct numerical calculations of the turbulent mixing in such flows, as illustrated in Figure 1.2.
!

One of the unique characteristics of these flows is that mixing occurs because of inviscid, baroclinic effects. This is perhaps best illustrated by considering the Vorticity Transport Equation: 0 _w

+ (u. V)w

= (w" V)u + w(V' I II

u) - V(1/p)

x rP+ III

V x (B/p) IV

+ v V2w V

which is derived by taking the curl of the momentum equation. The terms on the left-hand side represent the time rate-of-change of the vorticity vector w and the convection of vorticity by the velocity field u, The vorticity transport is balanced by a number of source terms (and a diffusion term) on the right-hand side: I II III " IV V vorticity vorticity stretching dilatation along the streamlines (due to compressibility by baroclinic (a three-dimensional effects), gradients interacting with effect),

vorticity creation density gradients),

effects (pressure

vorticity creation by buoyancy forces B (gravitational sity gradients), and diffusion of vorticity (due to molecular transport

forces accelerating

den-

effects).

Baroclinic effects predominate during the blast wave phase of the explosion. For example, vorticity is created by shock interactions with strong density gradients such as the fluidized bed, the thermal layer, and the fireball. These shear layers are

hydrodynamically unstable and roll up into turbulent mixing layers. The rotational structures in the boundary layer entrain dust from the fluidized bed; turbulent mixing and transport of the dust then occurs via the velocity field induced by the vorticity distribution. Additional vorticity generation occurs because of largescale pressure gradients (e.g., during the positive and negative phase of the blast wave) and because of local pressure gradients (induced by the rotational structures). Buoyancy effects dominate the cloud rise phase. Molecular diffusion effects are negligible for large-scale explosions because the Reynolds numbers are so large: Re = 10s to 101 (and Term V scales as 1Re). While others have used numerical models based on mean-flow equations and turbulence closure models, here we propose a Baroclinic Model for flows with large density variations that actually calculates the turbulent mixing and transport of dust on an adaptive grid. A number of examples will be presented to illustrate these baroclinic effects--including shock interactions with dense-gas layers and dust beds, and dusty wall jets of airblast precursors. The conclusion of these studies is that dusty boundary layers grow because of mass entrainment from the fluidized bed (and not because of viscous wall drag) as will be proven by the Mass Integral Equation.

Figure 1.1. Photographof thePRISCILLAEvent (37-KT burstat 700 ft over FrenchmansFlat drylake bed).

ti

2. BAROCLINIC

MODEL

The BaIoclinic Model is based on the following idealizations: (1) a loose dust bed; (2) an instantaneous shock fluidization of the dust layer; (3) the dust and air are in local equilibrium (so air viscosity enforces the no-slip condition); (4) the dust-air mixture is treated as a continuum dense fluid with zero viscosity; (5) the turbulent mixing is dominated by baroclinically-generated vorticity. These assumptions lead to an inviscid set of conservation laws for the mixture,

0
p -I- V" (pu) -- 0 (1) (2) (3)

0
ONpu + V(puu) = - VP

0
ONpE + V" (pEu) - - V .(pu) where u denotes the velocity and E represents the total energy: E - e + 0.5 u. u. The pressure p was related to the density p and internal energy e by the perfect gas equation of state" p-- (")'-)pe 1 (4) where V represents the effective V of the mixture. For simplicity a value of 7 - 1.4 was used in the calculations. This value is an adequate model for diatomic gases such as air. For the dust-air mixture, it also takes into account the kinetic pressure created particle-particle interactions of the dust which occur in the fluidized bed. If one employs a multi-fluid formulation, more complete equations of state are possible, however, this was considered to be beyond the scope of the present investigations. In the above equations, p actually represents Pd may be calculated from the relation: the mixture density. The dust density

pd= pc
This requires an additional
"

(5)
C, namely:

transport
0

equation for the dust concentration

0-7 + (u. v)c = 0 c


-

(6)

The above equations were integrated numerically by means of a high-order Godunov scheme. 15 Adaptive Mesh Refinement (AMR) is used to capture the turbulent mixing processes on the grid. In addition, we have a two-phase model to study nonequilibrium effects in dusty flows. The equations are presented in Table 2.1. One can see that there are separate conservation laws for the gas and particulate phase, which are coupled by drag and heat transfer effects.

"

I> II i I
II

II

i--

II

II

,-_

_, _

,i_

+
%

+ + I E _ _ _ f_ II

I _ II

_ .._ "_ II II

+ _ _

+ _

I II

II II

II _ II

,, E
II

+ _
II

II _

_ II r_

I II

7 II

II ._ _' =I II II

_II

_ _

II II -_, ._

,,_

,e

_ +
_,.

_I::> + + _, '_ _. _. "_

""
I II II II _a II

II

II

II

II

II

II

II

II

II

II

,._

_
.:" _ _, + ,., o II II II II II II ,.-_" ,._ II II r.,,,) + E "_ _ _" I II

i
I I '_.

r_ ..-.- I:> i::> _ ''_ "-" _" " I II II II II

II

_I II

,,.t _ II

'_ II
....

II

III

o
r_ M r._

_, o
_ r._

....

7
I i

3. PLANAR

SHOCK INTERACTIONS DENSE-GAS WALL LAYER

WITH

. -

Shock interactions with density discontinuities provide a rich source of fluid-dynamic phenomena. For example, normal shock interactions with planar density interfaces lead to instabilities and turbulent mixing at the interface, as studied by Richtmyer (1960) and Meshkov (1960). Planar shock interactions with inclined density interraces lead to a spectrum of shock refraction effects as shown in the shock tube experiments of Fattah et al. (1976), and to the rollup of the interface as demonstrated by the calculations of Zabusky et al. (1991). Planar shock interactions with low density layers located along the wall create shock precursor effects. This leads to a turbulent wall jet as shown in the shock tube experiments of Reichenbach and Kuhl (1987). In these examples, the shock interaction with the interface creates vorticity by a baroclinic mechanism: Ap A(1/p). This shear layer is hydrodynamically unstable, and rapidly evolves into a turbulent mixing region. Considered here is a new class of such problems, which can be used to study turbulent mixing in boundary layers that are dominated by density effectsin contrast with classical boundary layers that are dominated by viscous effects. In this respect, the current problem can be considered as a boundary layer version of the classic free shear layer experiments by Brown and Roshko (1974). The problem consists of the interaction of a planar shock wave with a Freon layer located along the floor of the test section of a shock tube. Shock interactions with the density interface create a wall shear layer that rapidly rolls up into a turbulent boundary layer. Described here are shock tube experiments and direct numerical simulations of this problem. 3.1 Problem Description

Experiments were performed in the Ernst-Mach-Institut shock tube. Figure 3.1 shows a schematic of the test section which had an effective rectangular crosssectional area of 7.5 cm by 4 cm. The wall layer was created by means of a fixture that was inserted into the shock tube. It contained a plenum for the injected gas, covered by a 65 cre-long porous ceramic plate (Filtrokelit) that spanned the 4 cm width of the shock tube. The plenum was filled with Freon-12 at a pre.csure of ApF _ 0.5 bars and run for about one minute -- to achieve essentially 100 percent Freon concentration in the plenum, and to cleanse the pores of the ceramic plate of residual air. Next, the Freon valve VR was closed and the shock tube was purged of contaminated gases. Then the Freon valve was opened and the plenum was pressurized to /kpF -- 1.5 bars for a time AtF (typically 2 s) and the shock tube diaphragm was broken mechanically. To increase the reproducibility of the experiments, all filling procedures were computer controlled.

This procedure created a thin layer (,,_ 0.3 cm) of pure Freon-12 on the floor of the test section. Above this, the Freon concentration decreased gradually in a thicker (,,_ 1.5 cre) diffusion layer. This was caused by residual air pollution in the pores of the ceramic plate, and by bi-molecular diffusion processes. The resulting layer was a 2 cm-thick by 65 cre-long laminar layer of Freon whose concentration profile was independent of x. The layer could be controlled by the Freon plenum pressure _PF and duration AtF; best results were achieved with ApF -- 1.5 bars and AtF -- 2 s. High-speed photography was used as the primary flowfield diagnostic -- to make visible the turbulent mixing processes occurring in the wall layer. It consisted of 24 sequential frames of shadow-schlieren photographs that were recorded by EMI's Cranz-Schardin camera. In addition, Mach-Zhender interferometry was used to evaluate the preshock density profile of the Freon layer. Figure 3.2 shows a schematic of the computational domain that was used for numerical simulations of the experiments. A rectangular x-y cartesian grid of 150 cells by 600 cells was employed. The mesh spacing was uniform with/kx = Ay = 0.05 cre. The grid was initialized with a quiescent air atmosphere: pl = 1.223 10-3 g/cre 3, Pl = 1 atm, ul = vi = 0. The Freon layer was then modeled by a Tanh(y) profile which approached the Freon density on the wall PF = 4ps. The left boundary of the grid was fed with constant conditions corresponding to the state behind a Mx = 1.38 shock wave: p2 -" 1.65 Pl,/92 -_ 2.06 Pl, U2 = 1.86 104 cm/s. Outflow boundary conditions were used on the right boundary. The roof and floor were treated as inviscid (slip) walls. Both the air and the Freon were modeled as an ideal gas with -_- 1.4.

3.2 Experiments

"

Figure 3.3 presents a sequence of photographs showing the interaction of a MI 1.38 air shock with a Freon layer. The preshock structure of the layer is visible in the first frame. The white band along the floor consists of pure Freon, about 0.3 cm thick. The black band above it is the aforementioned diffusion layer, where the Freon concentration gradually decreases to zero. The complete layer is about 2 cm thick in this experiment.

The sound speed in the layer was smaller than that in the air above it, hence the shock propagation was retarded in the layer. This caused the shock front to be curved near the floor. The oblique shock compressed the layer and reflected from the floor as a regular-reflection shock structure. The reflected shock reflected off the layer interface as a rarefaction wave which stopped the compression of the layer. The rarefaction wave reflected back and forth within the layer, causing periodic accretions and depletions of density. These effects are clearer in the numerical simulation presented in the next section. By about 10 cm behind the shock, the wave dynamics effects had damped out, and mixing processes began to dominate. Figure 3.3 shows that the flow was very unstable, and for At >_ 0.35 ms it rapidly evolved into a turbulent mixing layer even in these small-scale experiments. Although the flow was no doubt threedimensional, one can occasionally identify large-scale structures (e.g.,/kt = 0.85 ms and 0.95 ms) that entrain Freon from the wall layer. These structures created density striations that point up and to the right at an angle of about 30 deg from the floor. Similar striation effects are found in the numerical simulations. One can also observe fine-scale structures, especially near the top of the layer. 3.3 Numerical Simulations

A numerical simulation of the preceding experiment was also performed, using the Baroclinic Model. Figure 3.4 presents contour plots that show the initial shock interaction with the wall layer. The pressure contours show the refraction of the shock front in the layer, the reflection at the floor, and the formation of periodic expansion and compression waves in the layer. The internal energy contours show the periodic compaction (z = 61 cm and 71 cm) and expansion (x = 66 cre) of the surface. The vorticity contours show that vorticity is generated at the point where the shock interacts with the layer interface. The Freon concentration contours show that the interface begins to rollup almost immediately behind the shock front. At distances greater than about 10 cm behind the shock (x < 61 cre) the rotational structures pair and merge into larger-scale structures. Figure 3.5 depicts flowfield contours at /kt - 0.5 ms after shock passage at x -70 cm. They show an intense mixing with a spectrum of length scales. As is characteristic of variable density flows, the vorticity contours show that counter-sign vorticity (solid lines) are baroclinically generated in the braid region between rotational structures. Peak values are about -0.5 times the peak value of the vorticity of the main flow. The counter-sign vorticity, of course, increases the complexity of the mixing. The tendency in these two-dimensional calculations is to form multiple merging of vortices. At later times, this tendency leads to predominantly large-

l0

scale structures. In three-dimensional flow, however, vortex stretching and tangling will maintain the fine-scales within the large structures m as is evident in the photographs of Figure 3.3. Future calculations should include the three-dimensional effects. The flowfield was sampled at station x - 70 cm, and time-averaged profiles in the wall layer were evaluated. The mean flow profiles are presented in Figure 3.6. Flow variables were nond':mensionalized by the freestream values, denoted by subscript oo. The profiles are qualitatively similar to our previous simulations of the dusty boundary layer behind a planar shock. The streamwise velocity profile indicates that the layer remained about 2 cm thick during this time period -- thus eliminating the need for self-number, similarly stretching the grid to account for growth of the layer. Density effects cause the velocity to decrease to _ - 0.2 Uoo at the floor; if viscous effects were included in the laminar sublayer, then the velocity would go to zero at y = 0. In accordance with the negative displacement effect of shock-induced boundary layers, the mean transverse velocities are negative in the layer and reach a peak value of _ = -0.035 Uoo. The mean density reaches a peak value of about = 3.8 pco because of the Freon near the floor; hence, this mixing layer is strongly influenced by density effects. The mean pressure is essentially constant throughout the layer. Figure 3.7 presents the corresponding fluctuating flow profiles. Again, these profiles

are qualitatively similar to our dusty boundary layer simulations. The streamwise velocity fluctuations peak at a value of u _ = 0.31 Uoo, which is typical of boundary layers. The transverse velocity fluctuations peak at a value of v _ - 0.26 Uoo: this is probably too large by a factor of two m due to the two-dimensional flow approximation. The Reynolds stress u_v _ is positive, indicating that mixing is feeding fluctuating kinetic energy back into the mean flow; peak values reach u_v _ = 0.0075 U2. Density fluctuations are of order 100 percent, due to the Freon. Fluctuations in dynamic pressure and stagnation pressure are also large (e.g., 50 percent) near the floor. 3.4 Summary Shock interactions with a dense-gas layer create a shear layer on the wall by an inviscid (i.e., baroclinic) mechanism. The wall shear layer is unstable, and rapidly rolls up into a three-dimensional, turbulent mixing layer with a variety of mixing scales. One of the most interesting features of this problem is that it provides a method for studying high-Reynolds number, turbulent wall layers that are dominated by baroclinically-generated vorticity and density effects _ in contrast with classical boundary layers that are dominated by viscous effects. This problem also

|_

11

provides a gasdynamic simulation of turbulent dusty boundary layers if the dust particle diameters are very small. Measurements of both the mean and fluctuating flow profiles of this mixing layer will be performed to check the present numerical predictions. 3.5 References Brown, G.L. and Roshko, A. (1974) "On density effects and large structures turbulent mixing layers," J. Fluid Mech., 64:775-816. Abd-el-Fattah, A.M., Henderson, L.F., and Lozzi, A. (1976) "Precursor at a slow-fast gas interface," J. Fluid Mech. 76(1):157-176. in

shock wave--

Meshkov, E.E. (1960) "Instability of the interface of two gases accelerated shock v.:e.ve_"[zv, AN SSSR Mekhanika Zhidkosti i Gaza 4(5):151-157. Richtmyer, R.D. (1960) "Taylor instability in shock acceleration fluids," Comm. Pure Appl. Math, 13:297-319.

by a

of compressible

Reichenbach, H. and Kuhl, A.L. (1987) "Techniques for creating precursors in shock tubes, 16rh Int. Syrup. on Shock Tubes and Waves (ed. H. Gronig), VCH press, Weinheim, Germany, 847-853. Zabusky, N. et al. (1991) "Vorticity deposition, evolution and mixing for shocked, density stratified interfaces and bubbles," 18rh Int. Syrup. on Shock Waves (in press).

0cm

45cm

65 cm

\\,_'k\\\\\\\_\\\\\\\\\\\\\\\\\\\ I MI
A A

\\\\\\\\_ \\\\\\ \\\\\\\ Optical window


/_ ///A /// 7.5 cm

___.'_

' "_\\'_\\\'_\\\\\\\\\\'c\\\\\\\\\\ll ii' '_ "\"u'\\_


I,,_.._,, \ . Porous plate I_ -"-'"\ (Filtrokclit) Plenum

If
_],

_v_
FREON-12

Figure 3.1.Schematic of the test scction.

'

I I Z

_O_l,

:=

\N ___\\\\\\\\\\\\\\\\\\'__\\\\\\\\\\\\_,'_,\\\\\\5

Figure .2. 3 Schematic ofthe computational domain.

13

At=O
t

O, 5rn., 3 I

,s

O,_5ms
i

O,55ms
, i

Figure 3.3,

Shadow-schlieren photography of the interaction of a MI = 1.38 shock wave with a Fr_on wall layer (PI = I atm). i4

4_

"e

0,75ms
,.! -. : .

O,BSms

Figure 3.3. Concluded.

0.075

,_,

"o'o.o_ \_2<,d _:(*_:',f/i !! :


0 __ "'"_-"!' '' ")/- "" :"-' I: _ 0.075-'-"--" "E" ' - "__' e, ' \ _rl Co) '. _. ..... _ ,_ PRESSURE

,<_> :",,,,.. >-...,, _ : _Ill] _C


-,_ "_ (/ < /
,I

0.O75 F 0.050 )

(c) OME ZONE C C =0

y (m)
m

0.025 I 0 0.075 0.050

/ (d) VORTICrrY

y (m) 0.025 0 0.51


1

0.56

0.61

0.66

0.71

0.76 x (m)

Figure 3.4. Contourplots showingtheinitial shockinteraction withthewa.II ._ayer: (a) internal energy; (b)pttssur_; (c)Fmon concentration; (d)vo_.city. 0.075

o.o5oIcY (m) 0.025

\J i """ ' .- ...._... , " "_'" ' _ :"-" 2'-_-..''-_'_'_:" Y- ',//_.-(


) " ' ,, x _ "

oo: ] (.)__._oY

0.050 y (m) 0.025


-.

0.075

Co) E/qSrl_ D
:' .._. , --'

o_

,i, oo:I =o
0.075, 0.050

I
(d)VORTIClTY
0.80 0.85x (m)

y (m)

0.025 0 0.60 0.65 0.70 0.75

Figure 3.5. Contour plots showing intermediate-dme mixing in the wall layer (At = 0.5 ms): (a) inramal energy; Co) density; (c) entropy; (d) vorticity.

!6

Figure 3.6. Mean flow profiles in the wall layer (x = 70 cre): (a) srreamwise veloci_; Co)transverse velocity; (c) density; (d) pressure; (e) dynamic pressure; (f) stagnation point. !7

Figure3.7. Fluctuating-flow profiles in the walllayer (x - 70 cna):(a)streamwisevelocity; Co)transversevelocity; (c) Reynoldsstress;(d) density; (e) dynamicpressure; (f) stagnation pressrre.
18

4. TURBULENT

DUSTY BOUNDARY BEHIND A SHOCK

LAYER

"

Dusty boundary layers axe a common feature of explosions over soil or ground surfaces. Direct theoretical predictions of such flow fields are extremely difficult because of the two-phase, turbulent nature of the flow. In addition, wall boundary conditions (e.g., dust scouring) are particularly intractable. Hence, many of the initial studies (e.g., Ausherman, 1973) were empirically based. Hartenbaum (1974) measured the mean boundary layer profiles for steady flow over a dust bed, and deduced an empirical dust scouring rate for flow velocities of 30 to 120 m/s. More recently, Batt et al. (1988) have measured the velocity and density profiles in the turbulent boundary layer behind a shock propagating along a dust bed. However, such nonsteady experiments can give only instantaneous point values at various distances behind the shock; mean-flow profiles and R.,k_.S. fluctuations cannot be obtained from such data. Mirels (1984) has published analytic solutions for this problem, but he assumed the boundary layer profiles. Others have performed finite difference simulations; however, they used an empirical mass injection boundary condition on the wall and a tc-e model of the turbulence (Denison, 1988; Wolf and Strawn, 1982). Since these models were developed for clean flow, and since the dust dramatically modifies the boundary layer, the adequacy of this approach is questionable. Described here is a numerical calculation based on the Baroclinic Model for the turbulent boundary layer formed by a planar shock propagating along a dusty wall. It differs from previous studies in two respects: first, it follows the dynamic evolution of the unstable flow without resorting to turbulence modeling; second, the entrainment of dust into the turbulent boundary layer is included as part of the computational flow field, thus eliminating the need for a "dust scouring" model. The formulation of the calculation is described in the next section; this is followed by the results and a discussion that offers suggestions for accurate simulations of such dusty, turbulent flows. 4.1 Formulation

Figure 4.1 depicts a planar shock propagating with constant velocity Ws along a dusty wall. In stationary coordinates (Fig. 4.la), a boundary layer is formed behind the shock because of the no-slip condition on the wall. In addition, air percolates into the dust layer, forming a fluidized mixture of dust and air. For simplicity, the mixture was treated as equilibrium dense fluid of 50 mg/cc, as inferred from experiments (Batt et al. 1988).The specific case considered was that of an M8 -- 1.7

19

shock. As in previous analytic studies (Mirels, 1956), the analysis was performed in shock-fixed coordinates, (Fig. 4.1b). In this coordinate system, the velocities vary from the value at the edge of the boundary layer, ue = Ws - u2, to a maximum value on the wall, Uw = Ws. The computational grid consisted of 500 fine cells in the x direction and 60 fine cells in the y direction; a few coarse zones were used above and to the right of the fine mesh to remove any effects of the computational boundaries. The grid was initialized with the appropriate states from Table 1 and the profiles depicted in Figure 4.1c. The shear layer on the wall was approximated by a Tanh(y) function with proper asymptotes of u/al = 1.7 at y = 0 and u/ai = 0.77 at y = c. The density profile consisted of a three-cell-thick fluidized bed (53 mg/cc) with shocked air above it. Note that the inflection point (IP) of the shear layer was located in the air, three cells above the fluidized bed (FB). The left-hand boundary of the grid was then driven by these same profiles (Fig. 4.1c) with sinusoidal perturbations on the velocity field. Their frequencies corresponded to the frequency of maximum amplification rate from linear stability analysis and its first nine subharmonics. The maximum perturbation amplitude was 1 percent. A more complete description of this calculational approach may be found in Kuhl et al. (1988). The dynamic evolution of the flow field was calculated by means of the Baroclinic Model. The calculation was run for 6,000 cycles (6 hours) on a Cray-XMP to create a sufficient database for statistical analysis of the flow. 4.2 Results Figure 4.2 depicts the location of the material interface of the fluidized bed. Figure 4.3 presents the density (p), internal energy (e), vorticity (w), and pressure (p) contours near the end of the fine-zoned grid. These figures are useful for ftow visualization in the boundary layer region. The shear layer in the air rolled up into large rotational structures that entrained material from the fluidized bed. Dense material from the fluidized bed formed long striations because of the flow field of the rotational structures of the turbulent boundary layer. Thus, the dust scouring occurred naturally in the calculation without any modeling. Figure 4.4 depicts the boundary layer thickness 6 (i.e., the height of the 99rh percentile point of the mean velocity profile) as a function of distance behind the shock. The dusty boundary layer grows linearly with distance (6 - .024x) as a result of entrainment and merging of large-scale vortices. Note that this is in contrast to the clean boundary layer on a flat plate, which grows as the 0.8 power of distance from the leading edge (6 a x'S). The instantaneous flow field at the

"

2O

top of the fluidized bed (y = yo) is presented in Figure 4.5. The flow is clearly very chaotic; hence, useful representations of the flow require some type of averaging. To that end, the calculated environment was sampled for all y-cells at stations x = 400, 600, 800, and 950 for each computational cycle. The mean-flow environment was then calculated from these station histories by integrating over the last 5,000 cycles. . The calculated mean density profiles are depicted in Figure 4.6a. The profiles become thicker with increasing distance from the shock because of turbulent convection. Near the wall all profiles converge to a density of about fi/poo = 6(,,_ 17mg/cc). This convergence point defines the calculated height of the fluidized bed: yo = 3.5 cells. Only the flow field above yo was used in the boundary layer profiles; the solution below yo was then considered to be part of the computational modeling of the dust scouring. Note that these calculated profiles are qualitatively similar to the measured dust density profiles shown in Figure 4.6b. The mean-flow velocity (fi/U_), specific volume (_/Ac), and dynamic pressure (4/Q_) profiles are depicted in Figure 4.7 (subscript c_ denotes the freestream conditions above the boundary layer). Note that by using the boundary layer scaling r/BL "-- (y- yo)/_, the calculated profiles collapse to a similarity profile, independent of distance behind the shock. Note also that the calculated profiles are consistent with the LDV and X-ray measurements (shaded regions in Fig. 4.7) over a soil bed of 10-pm-diameter loose dust (Batt et al. 1988). Near the wM1, densities approach five to six times the freestream value. This causes the velocities in the wall region to be very small (fi/U_ < 0.1). Hence, the velocity profile in the wall region seems to be dominated by inertia effects of the dust, and not by fluid viscosity. Note that this situation differs considerably from that of a turbulent boundary layer on a flat plate, where fi/Uc _- 0.5 in the law-of-the-wall region, and viscous effects play an important role. Using a control volume analysis of the fine-grid region (with a bottom boundary at yo) and the mean-flow profiles, the nondimensional mass scouring rate was evaluated as: rho/pooU_ = 0.036. The root-mean-square (R.M.S.) fluctuations about the mean were also evaluated. These fluctuating-flow profiles are presented in Figure 4.8. The streamwise (u'/Uc_) and transverse (v'/Uoo) velocity fluctuations peak at about 30 and 13 percent, respectively, and extend many boundary layer thicknesses from the wall. Thus, dust will be transported well above the mean boundary layer height. The turbulent Reynolds stress (u-_v_/U_) peaked at about -0.003, similar to other boundary layers. Density fluctuations (p_/p_) peaked at about 8ix times the freestream value because of turbulent entrainment of dense

" .

21

material from the fluidized bed. The static (p'/Po,,) and dynamic (q'/Qoo) fluctuations peaked at about 17 and 42 percent, respectively.

pressure

The local fluctuation-intensity profiles are presented in Figure 4.9. In contrast with clean boundary layers, turbulent velocity fluctuation_ were much larger than their mean values (e.g., u'/fi .'_:> and v'/9 >> 10) because fi is so small in the wall region. 1 Hence, the fluctuations are the most important component of the flow. This seems to pose severe challenges for both experimentalists and turbulence modelers. A high-order Godunov scheme has also been developed for nonequilibrium gas-dust mixtures. This solves the conservation of mass, momentum and energy for each phase, with drag and heat transfer interactions between phases. A nonequilibrium calculation of the dusty boundary layer behind a shock was run assuming very fine (0.1#rn diameter) dust particles. The nonequilibrium mixture results (NE) are depicted by dashed curves in Figures 4.7 and 4.8. Both the mean and the R.M.S. profiles are essentially identical to the dense gas results. Surprisingly then, the dense gas approximation is quite an accurate model for this two-phase flow problem. 4.3 Summary Dust lofting behind a shock can be viewed as a two-step process--the formation of a fluidized bed, followed by entrainment of the dense material from the bed by the rotational structures of the turbulent boundary layer. The principal effect of the dust is to change the velocity of the flow. In the wall region, mean-flow velocities are small because dust densities are large. Also, turbulent velocity fluctuations in the wall region are much larger than the mean values due to this same effect. The present calculations focus on the two most important physical processes in the turbulent boundary layer, namely, the nonsteady velocity field of the large rotational structures, and the inertia effects of the dust. An accurate numerical simulation of these processes allows one to capture the first-order physical effects. For example, the velocity and density profiles, the dust scouring rate, and the boundary layer growth agree with the available data. 3 By inference then, nonequilibrium effects, fluid viscosity and three-dimensionality must have a secondary effect on the mean flow. Nevertheless, experimental data are needed to check the accuracy of the calculated R.M.S. fluctuations. This calculation vividly demonstrates the advantage of working in shock-fixed coordinates where the smooth or laminar solution is steady. By performing a nonsteady calculation in these coordinates, one can not only capture the rotational structures of the turbulent flow, but also record time histories at fixed distances behind the

" -

22

shock. These can be used to determine both the mean and fluctuating flow profiles without resorting to turbulence modeling--the main limitation being the 2-D flow approximation. This approach represents a significant advance over previous approaches that must assume turbulence properties which are, of course, Unknown for this dusty boundary layer flow. 4.4 References Ausherman, Haxtenbaum, D., (1973) Initial Dust Lofting: Shock Tube Experiments, DNA 3162F.

B., (!974) Lofting of Particles by a High-Speed

Wind, DNA 2737.

Batt, R.G., Kulkarny, V.A., Behrens, H.W., Rungaldier, H., (1988) "Shock-induced Boundary Layer Dust Lofting," Shock Tubes and Waves, ed., H. Gronig, VCH, Weinheim, Germany, pp. 209-215. Mirels, H., (1984) Blowing Model for Turbulent AIAA, 22(11), pp. 1582-1589. Denison, M.R., A Two-Layer Boundary-Layer Dust Ingestion,

Model of Dust Lofting (in press). Layers in Surface Burst Flowfields (in press). Wave Moving

Wolf, C.J., Strawn,

R.C., Boundary

Mirels, H., (1956) Boundary Layer Behind a Shock or Thin Expansion into a Stationary Fluid NACA TN 3712.

Kuhl, A.L., Chien, K.-Y., Ferguson, R.E., Glaz, H.M., and Colella, P., (1988) Inviscid Dynamics of Unstable Shear Layers, RDA-TR-161604-006, R & D Associates, Marina del Rey, CA.

23

IL

W=

(b)

Ws-U ,,,,,,,
"
w

(9 _gbiLz

"
FLUIDIZED BED (FBI Ii,

., "

_ ,.......
FB ld) I OUTFLOW B.C.

20
>"

_' --Ue--

olyl
lP

lC)

C)
INFLOW B.C.

10- P(Y) 63U

_.

Tanh(y) ,-"i

UolY), Poly )

.......

"

o ---rl w p(mg/cc) "" 5o'1oo 5' lo'

_/i _/.4Z/////I/I/I/i/I/A_,,,,......,,,,_/HH/'/////III///m" \ WALL SLIP WALL B.C.

Figure 4.1. Dusty bounda_ layer behind a shock: (a) stationary coordinates, (b) shock-fixed coordinates, (c) inflow profiles, (d) grid.
40, Y 20 r i 0_) 40

t
t

FB iii i 1O0 X -1 200

Y 20 (ai 0 800 900


X

1000

40_ Y 20_

1 _
4..0

oL _ (FB_- i
200 300 X Y 20 40 f 0| 400 40 Y 20 0 600 _FB ,i 500x -

4o b)
Y 20_,,__ 0 _',!_-__'_.

-_

_]t 600

800 Y 40 | tc} 20L "-'.-_ 800 0

900X

1000

"

..... ' 900 "_'_:_"_" x

_,jLHL__ 1000 " "

700
X

800

y 20 40 .... 0"--' 800

_ .,_ ---": -'-_.i..-'_.,_._, . 900


X

.. 40t (di 0 800 , _,_i 900


X

--"ii,_- , 1000

..

..-

-t 1000

Figure 4.2. Material interface plots at t = 8202 showing entrainment of dust from the fluidized bed (FB).

Figure 4.3. Flowfield contours at t = 8202: (a) density, (b) internal energy, (c) vorticity, (d) overpressure.

24

'Jl

0.6

.........

I ........

lb. t

@oG

.,/

0.4

6 20-%% /

0 p/S._ _.._ ' 0 200 Figure 4.4. 50 L -

I 400
X

I 600

"

i' 800

" !. _000

-0.4L_ -0.6_, , , i , l

1 4'l
i

=1 ", -.!
"

_.o.,I ,! : ; : , i_l .! I
" :l '. 3o_ ', t a

Boundary layer thicincss 8, momentum o thicirJlCSS O, arid c_ vs distance behind shock. , 0.60 t (ai : : , :i , : !

500

looo
i

ii .t ;i

(c)_

i
] } l-,ig ' ! 1

,. I

,f I , ,

,;i

._

10-_ ..... : i S_ II 0
" + ,

L/_ im i' i t, _L_ :,;" ,4_i t_.,<'* i ' ii I lUl " l

', !'_ i!l / i

,, t ,It , t, ii ! , ' I

"

I .,1 :' ,'1,; :

soo

_ooo

o.o _;' o
0

5oo

II l
'

_ooo

Figure. 4.5. Instantaneous flowficld (t _:8202) at the top of the t_uidizext bed (y = yo): (a) density, (b) and (c) velocity and dynamic pressure (stationary coordinates).
(ai '

t
I

,
"'

(b)

2o _ _
>

r_x: ,.oo o 600

; --:

I 2o[_,=_o

1__D<

_
1

800

-E

15

k_

s;=_yo.. ."""_;t:_.../,,,__ x=o-1 o'l , ,,, ,_',_


2 4 6 8 10 20 PIPoo

s oL
1

,,,

4 PdlP

,,,,i

8 10

20

;:igure 4.6. Density profiles behind a Ms = 1.7 shock: (a) calculated, (b) measured (Batt et al., 1988).

'II

' 'I_'

'

0.8

-:.!:_':

0.6

8
0.4 _ 0.2 -

_'__
{3X = 400 O 600 A 0 800 950 I C'---O.1 17BL i (bl 0.8 i J _ _ = J i j 0.2 , = a 0.4 _ _ _ , i 0.6 0.8

0 ., 0.01 1

0.6 8 0.4 {3 X = 400 ,.'i'_S "'':'' -:_'_-_'_;*'=.:." -

0.2

o
0 (C)

coo
950 . F.

;":;":':_:"'

I I =

0.01

0.1 T/BL

0.2

0.4

0.6

0.8

1 -

0.8

'

IO"

8 o

0.6-

0.4 0.20 0.01

{3X = 400 O 600 Z_ 800 0 950 t 1 l i I s 111 0.1 _BL i 0.2 i i 0.4 I i I I I 0.6 0.8 1

Figure 4.7. Mean-flow boundary layer: (a) velocity, (b) specific volume, (c) dynamic pressure.Shadedregionsdenote databandsof Batt ct al. (1988). Dashed curve denotes NE mixture results

at x = 950.

fir

26

O.3

0% !_4_: B X = 400

la)

O.2 i

'

(e)

o_ _r_
_ "_
0.1 0 0.2
v

_ _oo
_o "a.

_-_

Ib)

0.3

,' _.

oo

"

I:;
-6

I,o. ,
i d:u
, i i

1.0

2.0

3.0

4.0

100 v'/; ld)

" 6

-V'/_ 10 U'/O 8 4

2
m

o'lB

0 . 0 1.0 2.0 'qBL 0.1 Figure 4.8. R.M.S. fluctuating-flow profiles: (a) streamwise, (b) transverse, (C) Reynolds stress, (cD density, (e) pressure, (f) dynamic pressure. Dashed curve shows NE mixture results at x = 950. o.ol o. 1 T/BL
At't

3.0

4.0

_.o

Figure .,.:,.

Loca] fluctuating-intensity profiles at x = 950.

5. DUSTY

WALL

JET

IN A PRECURSOR

Airblast precursors are a shock refraction effect accompanying explosions with a large thermal radiation output. Thermal radiation from the fireball is absorbed by the ground and creates a heated layer of gasma thermal layer (TL)mahead of the blast wave. The shock front refracts into this high-sound-speed layer, forming the precursor shock structure characteristic of such explosions (Glasstone, 1962). The major fluid-mechanic feature of this flow is a supersonic wall jet consisting of a free shear layer (FSL) and a wall boundary layer (BL). The jet is unstable and rapidly becomes turbulent. Airblast precursor flows have been studied experimentally under a variety of conditions. These include small-scale laboratory experiments (Reichenbach and Kuhl, 1988), 1000-1b HE tests of the DIAMOND ARC series (Reisler et al., 1988), and the ANFO surface-burst field test (Lutton, 1988). Precursor experiments have also been conducted in the AFWL 6-ft shock tube (Newell, 1987). The main drawback of all these nonsteady experiments is that the data are always taken in laboratoryfixed coordinates, and thus there is no possibility of time-averaging the nonsteady data to produce accurate mean-flow profiles of the turbulent wall jet and boundary layer. A number of researchers have performed numerical simulations of airblast precursors. Typically, these numerical calculations were based on the time-averaged Navier-Stokes equations, turbulent transport was then modeled; e.g., A mixing length model (Rosenblatt cursor solution (Mirels, 1988) et al., 1989), based on an analytic pre-

A two-equation turbulence model, such as the kor the k -w model (Traci and Su, 1988) A six-equation turbulence model (Donaldson

_ model (Barthel,

1985)

et al., 1982)

Such simulations .

attempt

to calculate the evolution of the mean flow. As mentioned

above, however, the available data are all instantaneous point measurements of the turbulent flow and cannot be used to directly check the mean-flow calculations. In addition, these turbulence models were all developed for steady, incompressible

28

clean flow, whereas the airblast precursor flow field is inherently a nonsteady, compressible dusty flow. Hence, the accuracy of such turbulence models is questionable for airblast precursor flows. In the past we used an alternative approach; we performed a direct calculation of the turbulent mixing by following the dynamic evolution of the large-scale turbulent eddies on the computational grid. This approach was used to parametrically investigate precursed flows (Glowacki et al., 1986), to design precursor simulation experiments in large shock tubes (Kuhl et al., 1985) and on large HE tests (Kuhl et al., 1987), and to simulate the turbulent, dusty boundary layer behind a shock (Kuhl et al., 1989). This section describes an extension of such direct calculations of the turbulent mixing. The problem considered is the interaction of a planar, constant-velocity shock with a constant-property thermal layer over a loose soil bed. The calculated turbulent environment was stored along similarity lines; then the solution was timeaveraged to produce the mean and r.m.s, profiles of the turbulent wall jet and dusty boundary layer flow. 5.1 Formulation The problem considered here is the precursor flow field created by a planar, constantvelocity shock interacting with a constant-property thermal layer, TL (see Fig. 5.1). Dust scouring effects were incorporated in the flow by inclusion of a loose fluidized bed, FB, along the wall below the thermal layer. A 2-D Cartesian grid was used for the computational mesh. This consisted of a slowly growing fine-mesh region (100 < i <:600 with initial Ax = 0.1 m; 1 < j < 100 with initial Ay = 0.1 m) that followed the precursor shock structure, a stretchedmesh region (1 < i < 100 with Ax = variable) to capture the flow behind the precursor, and a coarse-mesh region above the precursor (100300 < j < 140 with Ay -- variable) to simulate the unconfined case. The mesh was initialized " with ambient conditions in the air (subscript 1)"

'

"

pl = 1.01325 x 106dy/cm2;pl el = 1.96 109erg/g;

= 1.293 x 10-3g/cm3; ul = 0

al = 3.31 104cm/s; FB):

"

and a three-cell

thick fluidized dust bed (subscript

aFB/al {PFB/Pl

-- 0.161; ufs = =1;PFB/P1=38"67;

eFB/e1=O'0259}

0 < y < ( lOm<-x<-c_ 0.3m )

29

and a ten-cell

thick; constant-property

thermal layer (subscript

TL):

aTc/a_

= 3.18; UTL = 0

0.3m < y < 1.3m value. (state 2)

Note that the thermal

layer somld speed is about three times the ambient

The left boundary of the mesh (x = 0, y,t) was driven by the conditions behind an MI = 1.7 shock: p2/Pl = 3.2; p2/pl = 2.2; e2 el = 1.46

a2/al

= 1.21; u2/al = 0.94; v2/al

Wall drag was neglected at the bottom of the fluidized bed; hence, an inviscid snp boundary condition (v = 0, Ou/Oy = O, Op/Oy = 0) was used at the bottom boundary. An outflow condition was used at the upper boundary of the mesh. The calculation was run for 6,000 computational cycles to accumulate enough data for a good statistical analysis of the turbulent flow. This required about ten CPU hours on a Cray XMP computer. The results are described in the next section. 5.2 Precursor Flow Visualization

Figure 5.1 presents internal energy contours at various times near the beginning of the calculation (t = 100 to 300 ms). They show the initial formation of the precursor shock structure (shocks P-T-I and triple point TP) and the wall jet. The wall jet is very unstable and becomes turbulent almost immediately. At a time of about 300 ms, a large rotational structure breaks off the rear of the jet. This structure was the result of large-scale vortex merging at x __ 133 m and t = 220 ms. Figure 5.2 presents a series of snapshots of the flow that shows the evolution of the turbulent wall jet from t = 400 ms to 1000 ms. From t = 300 ms to 800 ms the wall jet seems to propagate in a quasi-steady mode, with all features growing linearly with time. Then, at about t = 800 ms the turbulent mixing again penetrates all the way to the wall; large-scale merging occurs and this piece of the flow is shed from the rear of the jet at about t = 900 ms. This is accompanied by an almost total collapse of the jet. At the time of about t = 1000 ms, a new precursor, P', forms and re-energizes the jet. The process then repeats itself with a period of about 500 ms. This quasi-periodicity has been confirmed by a separate calculation using a one-zone-thick thermal layer where the calculated quasi-period was 50 ms. Figure 5.3 gives a more detailed view of the flow field in the quasi-steady regime (t = 600 ms). The wall jet is supersonic; turbulent mixing on the free shear layer

"

"

3O

"

(FSL) creates weak shocklets above and below the shear layer, as depicted in the internal energy and pressure contours of Figures 5.3a, b. The velocity field from the vortex structures on the free shear layer trips the wall shear layer (WSL) and causes it to roll up (Fig. 5.3c). Pairing of the boundary-layer vortex structures with the free-shear layer vortex structures causes the boundary layer to separate locally from the wall at many locations along the wall jet. This phenomenon is similar to the strong interactions found in low-speed wall jets (Bajura and Catalano, 1975). This effect is visible in the entropy contours (Fig. 5.3d), which show that material from above the free shear layer mixes all the way to the wall, and dust from the fluidized bed mixes to the top of the free shear layer. Note that conventional turbulence raodels are not designed to simulate such strong interactions between two shear layers containing vorticity of opposite signs. 5.3 Wall Jet Profiles

,i

Figure 5.4 presents the time-averaged mean-flow profiles of the wall jet during the quasi-steady phase of propagation (300 ms to 700 ms). The profiles are nondimensionalized by the peak values and scaled by the wall jet similarity variables:

= - xsp(t)]/[xj(t)- xsp(t)] ,j = (y- yo)/(y,- yo)


where r/j equals 1 at the centerline of the free shear layer and 0 at the top of the fluidized bed. The wall jet scaling seems to collapse the _/Um velocity profiles quite well, especially in the free shear layer region (0.5 < r/j < 1.5). At the tip of the wall jet (_ -- 1) the boundary layer and free shear layer are decoupled by a constant-velocity plateau. Near the front vortex (_ _ 0.8), one can observe a separating boundary layer profile with an inflection point. The inflection point is closer to the wall in the stagnation point region (_ = 0). Near the wall, the mean vertical velocities are positive, because the turbulent fluctuations scour (entrain) dust from the fluidized bed; the mean density values are quite large (ft/pe "" 10) for the same reason. The mean pressures are constant in the wall jet (except near = 0), indicating an isobaric mixing process, similar to other turbulent flows. Figure 5.5 presents the root-mean-square (r.m.s.) fluctuating-flow profiles of the wall jet during the quasi-steady phase of the propagation (300 ms to 700 ms). The wall jet scaling also seems to collapse these profiles. The streamwise velocity fluctuations peak in the boundary layer (r/g " 0.2) and near the center of the free shear layer (r/j = 1.0). The transverse velocity fluctuations peak near r/j = 0.6, and are non-zero at the bottom of the boundary layer (r/j = 0) due to turbulent fluctuations from the fluidized bed. The Reynolds stress peaks near the wall and near r/j = 1. The density fluctuations are very large (p'/pe " 6) near the wall because

31

of turbulent scouring of dust from the fluidized bed. The pressure fluctuations relatively constant in the jet. 5.4 Discussion

are

"

A shear layer is formed when the precursor shock interacts with the thermal layer. A shear layer is also created at the top of the fluidized dust bed on the wall. The free shear layer is unstable and rolls up into large rotational structures on top of the wall jet. The wall shear layer is also unstable; the velocity field from the turbulent structures on the free shear layer trips the wall shear layer, which then also rolls up. There is a strong interaction between the two; for example, gas from above the free shear layer mixes all the way to the wall, and dust from the fluidized bed mixes to the top of the free shear layer. Boundary layer separation occurs at many locations along the wall jet. Turbulence models are not formulated to model such nonlinear interactions between shear layers; hence the turbulent mixing in the precursor wall jet should be calculated directly--as was done here. The precursor shock structure and wall jet did not continue to grow indefinitely with time. Turbulent mixing eventually destroyed the coherence of the wall jet and limited its propagation to about 500 thermal layer thicknesses. The flow was quasiperiodic with a duration of about 500 ms in the present calculation. During that time, a quasi-steady phase existed where the precursor shock structure and wall jet features grew self-similarly. Similarity coordinates _ and 77were used to sample the flow field during the quasisteady phase of propagation. Time-averaging in these coordinates was used to find both the mean and r.m.s, profiles of the wall jet and boundary layer flow without resorting to turbulence modeling. The calculations showed that there is no single boundary layer velocity profiles in the precursor wall jet; instead, the profiles depend on _ because of gradients in the mean flow and because of increasing dust mass effects at increasing distances behind the precursor. Behind the wall jet region, the profiles are similar to the DG3 calculation of a dusty boundary layer behind a planar shock (Kuhl et al., 1989). In closing, we recommend a direct calculation of the turbulent mixing for such problems--because the dust mass modifies the turbulence, and because of the complexity of the nonlinear interactions between the free shear layer and the wall boundary layer. In other words, one should time-average the solution, not the equations. In the future, this approach should be used to calculate the turbulent mixing in nonsteady blast reflection problems.

"

32

5.5 References

Bajura, R.A. and Catalano, M.R. (1975), Transition Wall Jet, J. Fluid Mech., 70, pp. 773-799.

in a Two-dimensional

Plane

"

Baxthel, J. (1985), $-D Hydrocode Computations Using a k- _ Turbulence Model: Model Description and Test Calculations, SSS-R-85-7115, S-Cubed, La Jolla, California. Colella, P. and Glaz, H.M. (1985), "Efficient Solution Algorithms for the Riemann Problem for Real Gases," J. Comp. Phys., 59(2), pp. 264-289. Donaldson, C. et al. (1982), Second-Order Closure Model: Comparison with a Number of Complex Turbulent Flows, ARAP 469, Aeronautical Research Associates of Princeton, Princeton, New Jersey. Gilmore. F. (1955), Equilibrium Composition and Thermodynamic Properties Air to 2_,000 K, RM-1543, Rand Corp., Santa Monica, California. of

"

"

Glasstone, S. (1962), The Effects of Nuclear Weapons, U.S. Atomic Energy Commission. Glowacki, W.J., Kuhl, A.L., Glaz, H.M., and Ferguson, R.E. (1986), "Shock Wave Interaction with High Sound Speed Layers," Proc. 15rh Int. Sym. on Shock Waves and Shock Tubes, eds. D. Bershader and R. Hanson, Stanford University Press, Stanford, California, pp. 187-194. Kuhl, A.L., Glowacki, W.J., Glaz, H.M., and Colella, P. (1985), "Simulation of Airblast Precursors in Large Shock Tube," Proc. 9rh Int. Sym. on Military Applications of Blast Simulation, Oxford, England. Kuhl, A.L., Glowacki, W.J., Glaz, H.M., Ferguson, R.E., Collins, P., and ColeUa, P. (1987), "Simulation of Airblast Precursors on Surface Burst HE Tests," Proc. l Oth Int. Sym. on Military Applications of Blast Simulation, Bad Reichenhall, Germany. Kuhl, A.L., Chien, K.-Y., Ferguson, R.E., Collins, J.P., Glaz, H.M., and Colella, P. (1989), "Simulation of a Turbulent, Dusty Boundary Layer Behind a Shock," (in preparation). Lutton, T. (ed.) (1988), Proc. MISTY PICTURE Symposium, Defense Nuclear Agency, Kirtland Air Force Base, New Mexico. POR-7187-5,

'

33

Mirels, H. (1988), "Interaction of Moving Shock with Thin Thermal Layer," Proc. 16rh Int. Sym. on Shock Tubes and Waves, cd. H. Gr6nig, VCH, Weinheim, Germany, pp. 177-183. Newell, R. (1987), "Nonideal Airblast Simulation in Shock Tubes," Proc. lOth Int. Sym. on Military Applications of Blas_ Simulation, Bad Reichenhall, Germany, pp. 302-321. Reichenbach, H. and Kuhl, A.L. (1988), "Simulation of Precursors in Shock Tubes," Proc. 16rh Int. Sym. on Shock Tubes and Waves, cd. H. GrSnig, VCH, Weirtheim, Germany, pp. 847-853. Reisler, R., et al. (1988), DIAMOND ARC87-Blast Phenomenology Results from HOB: HE Tests with a Helium Layer, DNA-TR-87-99-V2-AP-A-G, Defense Nuclear Agency, Washington, D.C. Rosenblatt, M. et al. (1989), Nonideal-Surface tions in Support of HML. 1. CRT-3770F, Chatsworth, California. Sedov, L.I. (1959), Similarity Press, New York. and Dimensional Airblast and Vehicle Loads PredicCalifornia Research & Technology,

Methods

in Mechanics,

Academic

Traci, R.M. and Su, F.Y. (1988), Turbulent, Two-Phase Boundary Layer Analyses, FP1-R88-06-05, Fluid California.

Flow Modeling for Dusty Physics Ind, Encinitas,

34
p ,

40.

sz.

t = I00 ms xs = 86.30m

16. o.
20. ilO.

TP (_7_o o " "

TL
J _,, aO. |00. |_0.

45.

FigureS.1.

Internal energy of a precursor

contours wall jet,

showing t = i00

the formation to 300 ms.

I
i,

35

M.

'

'

'

'

'

4s. -

t = 220 ms xs = 179.09 m

:m.

49.

I
XS

t = 26(I ms
l 207.09 m

37.
E

TP

zs.

13. %

O, cue

lot.

128.

154.

lot.

207.

234.

x Cm) Figure 5.1. 36 (Concluded).

r_ .. 4,
E >,

_ TP

Xs = 400 ms m t 295.10

'8

O. tsl.

._, lO0. _. 2m. 267. _F/.

W.

'

'

t - 500 ms x s - 358.38 "' _ TP P

19.

._

Oo

'

' '

'

'

t 600 ms
r_. Q -i I Xs

426,46

a.

"_s 411.

>_

( TP T

2_,

O. 2'30. 216. 321. 35"/. 3B3. q2B.

x (m)
FigureS.2. Evolution of the turbulent 37 wall jet from t = 400 to i000 ms.

Tp

t : 700 ms xs = 495.73 m
r

T
24.

38

(a)
_) t : 600 ms

-'E4:. I >,,
22.

XsP: 426.46 m

WALLJET
O. 2'JO. 286. 321. 3S'/. 393. J,28.

4:1.

-..

_-

c_> _,_ _\_,_


\

22.

O,

'

'

SP '

'

'

'

'

'

'

"'

'

Ct:)
E

22. >_

FSL

I
O.

BL

WSL

j
i

x (m)
43. ' i ' I , l " , l m

(d)

O.

-0.5

0.2

0.4

0.6

0.8

1.0

1.2

Figure5.3.

Contours t = 600 (c)

showing ms: (a)

the flow-field internal energy, (dl entropy,

details at (b) pressure,

vorticSty,

I.

II

-m

gO

'

= 0.8

=
= _=Ov

(a)
.2 ta 4 _ 6 *

40

= ,i ' .8

(c)

.8 1.0 # 1.2 m _

30 , -.2

o.6 :=E
. I_ 0.4

o
BL Y*

l,=.
20 _0

" 0 2
,.

I0
lr"

'

FB

FSL

,_

_I
m 0 I "
e-r

FB

--

0.]2

o
FB

Cb) ,,= == "= =

]. 2-"'lumnumnumnimmmnnmmnmnmmmnini._/_.
amffiffiimmffiffimffiffi

"

0.08
=,

1.0. 0.8GI

0.04
t

j_ o
" -0.04

o.6.
O.4.

.z

"

-0.

0.2-

-0.12 ,il -0.2 0

_@ 0.4 0.8 Tj 1.2 1.6 2.0

......... 4

10.2 0

.8 Tj I2

l6

2,,0

Figuro5.4. Mean-flow profiles of the wall jet (0 _ 1 2): ,_, =_=_,,w==e velocity, tD) transverse velocity, (c) density, (d) pressure, 4O

" 0.I 16-

(d)
_=Ov .2 u

" FB

"%
ii I

,4

0 0.30_

12

.6

A
;
0.24 p ,

(b)
l '_ 'i .'p, _1 .& I
Ni

"I -= _
FB 4
ql/"ip

1.0 :

12 m

0.18

=E
_ _ o 0-

0.12
"
eB

"
ii

0.8i
_

I"

FB

e ()

0.06 . FB
II t I 0 10.2 0 IliliiII

"
II

0.4
la

,, "

,"
0

0.4 0.8
R.MoS. (0 _ _

1.2 1.6

2.0

10.2 0

0.4 0.8

1.2 1.6 2.0

nj
Figure5.5.

nj
fluctuating-flow profiles of the wall jet 1.2): (a) streamwise velocity, (b) transverse 41

++
i

rl

velocity,

(c) shear

stress,

(d) density,

(e) pressure.

6. DISCUSSION This section explores the mechanisms of the growth of the wall layer in the context of boundary layer theory. A complete derivation of the equations can be found in the Appendix. We start by defining the mass thickness 6,,, which is related to the boundary layer thickness 6 according to:

.
" Here, Im represents boundary layer:

=z.,6

(6.1)

the integral of the mass and mass-flux profiles taken over the = --

+']0'
U2

(h - 1)dr/+

/0'

(1 - hf)drl

(6.2)

where u2 denotes the gas velocity behind the shock. If the density and velocity profiles in the wall layer are self-similar (i.e., if they are independent of x so that h = _/p_ = h(r/) and f -- fL/U_ = f(r/), respectively), then the mass integral Im equals a constant; such was the case for the normal shock calculation where Im _ 1.85. Next, consider the Mass Integral Equation = dno - voolu2 (6.3)

d [In 6"= d---( 6/R,]

which may be derived from a control volume analysis of the mass flux in the boundary layer. In the above, rho represents the rate that mass is being entrained into the bottom of the boundary layer at 77- 0 due to turbulent mixing: rho = Po Vo/p2u2 " (6.4)

Assuming that the velocity and density profiles are self-similar, then 6" = Ira6' and Equation (6.3) becomes:

6'=(mo-vlu )lS.,

(6.5)

Thus the Mass Integral Equation states that the fundamental reason that the dusty boundary grows is because of turbulent mass entrainment from the fluidized bed (i.e., because of rho). This is true regardless of momentum considerations. In a similar manner, one can define a momentum 60 = Io 6 thickness 60 of the boundary layer: (6.6)

where = _
IL2

h(1 - f)do

hf(1

- f)drl

(6.7)

Again, using a control volume analysis of the momentum flux in the boundary layer, one can derive the Momentum Integral Equation (for the case where there are no pressure gradients in the freestream): d [Io 5/R_I - rho + cii2 . (6.8)

where c I denotes the local wall drag. This states that the momentum thickness grows because of mass entrainment and wall drag. Typically, rho >> cl/2 (e.g., rho "_ 0.04 and cii2 < 0.001), so again one finds that the boundary layer growth is caused by mass entrainment. Solving Equation (6.5) for the mass entrainment rate, one finds (6.9) relation (6.13):

rho = I,-, 6' + vo_/u2 The boundary layer slope 8' may be evaluated 6'= #a_ a-1 from the empirical

= 0.022_ -2/5 (for a planar shock) ' and eliminated from Equation (6.9), yielding:

(6.10)

,:no = I,n _a_ _-1 + voo/u2

(6.11) calculation case gives: (6.12)

t
it
{4

Evaluating

the constants

in the above for the planar-shock rho = 0.041 _-2/5 _ 0.01

!t
H

li

This relation shows that the mass entrainment behind the shock front. The mass entrainment

rate decays as _-2/5 of the distance rate for a dusty boundary layer in

an ANFO surface burst is presented in Figures 6.1 and 6.2. The dusty boundary layer growth for a variety of problems is presented in Table 6.1. We found that the boundary layer grew as a power function 5/R, = a__ (6.13)

(6.14)
of the distance behind the shock Rs. The exponent was case dependent (3/5 _<_ _< 5/6) but waz similar to exponent of 4/5 for clean turbulent boundary layers on flat plates.

,
t

43

Table CASE Clean Flat Plate is SQUARE " "

6.1.

Boundary

Layer Growth BOUNDARY LAYER GROWTH /_

6/)(. = 0.37Re_ 1/5


5 _-_ X 4/5

4/5

WAVE SHOCK REFLECTIONS

22 5/Rs = 0.037 _a/5 (0.1 < _ < 0.7) 5/Rs = 0.0157_ 3/5 (0.1 < _ < 0.6) = 0.0147_ a/5 (0.1 < _ < 0.3) 5/R_ = f_(_) __ 0.0213_ a/5 (0.1 < _ < 0.25) 5/Rj = f2(_) 5/6 1 __ 0.0325_ 5/6 5Rs = 0.024_ 3/5 3/5 3/5 3/5

Normal Shock- Case 1 (MI = 1.7, _,_ = 0) RR- Ca_e 2 (Mt = 2, Ow = 60 )

SMR- Case 3
(Mi = 2, Ow = 27 ) DMRCase 4 (MI - 10, _w = 30) Precursor Case '7 fluidized bed 13 (MI = 1.7, PTL/Pl = 0.1, PFB/Pl = 50) Normal Shock, infinitely-long (MI = 1.7, PFB/Pl = 50) SHOCK (MI = 1.7) TUBE

5/R,

EXPERIMENTS 6,= 0.0325(Ax)5/6 5/6

Normal shockoverloose soil bed9

[x] =
5,= tangentslopethickness Normal shockalonga cleanwall 9 (Mz = 1.7) DECAYING ANFO B LAST WAVES 5,= 0.00983(Ax) '9z 0.93

Surface Burstoverloosedustbed < 80, PFB/P,


= 50)

(20 < Apz(psi) Point Explosion dust bed

5Rs

= 0.0256_ 5/6

5/6

Surface Burst over loose _<8000, PFB/Pl = 50) 5/Rs = 0.086_ 5/6 5/6

"

(1000 <_Api(psi)

HOB=50 ft/KT 1/a over deep snow (3_<Apl(kb) <_I0,PFB/Pl -- 200) a RR Region MR Region 5 = z where fi = 0.99 U= 5t = tangent slope thickness 44

5_/_R_ = 0.0886_/6 s 5v/_Rs = 0.110_ 5/6

5/6 5/6

'

'

'

/ /
/

jc

0.03

0
/

0.02 _

0.01 _

/ /

I /

0.6 Fi_e 6.1.


i

0.7

0.8
X

0.9

1.0

No_cnsio_ mass en_r_nt _ MOv_s_ _s_ce x an ANFO sm'face burstexplosion (Mo = xP-_o/P2u2). 0._ , , , ,
i

O 0.4 -

.
mo

0.3m

0.2 -

O\x

\
\

!
\
\ \

0.10 x i I I i I t

_
\-

0.6 Fi___ 6.2. _

0.7

0.8
X

0.9

1.0 x m an

mass sco_ng .ra_ rha versus die.ce

45 ANTO sm'fcc burst (mo = p-_olP,j.],).

7. SUMMARY

AND

CONCLUSIONS

The preceding illustrations have shown that turbulent dusty flows are dominated by baroclinic effects. For example, vorticity is created by pressure interactions with density gradients. The dust scouring is a function of the local vorticity. Also, turbulent mixing is driven by the velocity field associated with the vorticity distribution. * The Baroclinic Model demonstrates that dusty boundary layers grow by entrainment of mass from the fluidized bed -- as proven by the Mass and Momentum Integral Equations. Wall drag effects are unimportant for turbulent boundary layers with large dust loadings. The role of air viscosity is reduced to keeping the dust-air mixture in local equilibrium, while the mixture viscosity is essentially zero. The remaining issues are: (1) convergence of the numerical solution; (2) threedimensional effects on the turbulent mixing; (3) improved constitutive relations for the dust-air mixture; (4) improved zoning in the boundary layer (via implicit Godunov schemes; and (5) nonequilibrium effects. These issues will be addressed in future calculations.

45

APPENDIX MASS

AND MOMENTUM INTEGRAL EQUATIONS FOR DUSTY BOUNDARY LAYERS of the turbulent coordinates dusty boundary layer created by a

Described " .

here is an analysis

shock wave that is propagating in Figure A-1. In stationary with a constant subscript velocity

along a loose dust bed. (Fig. A-la), ahead and behind

The problem

is depicted by

the shock wave S propagates the shock are denoted

Wj; states

1 and 2, respectively. Densities

Flow interactions

with the fluidized bed FB, create of dust from entrainment of

a velocity deficit (shown as the shaded regions D_ and Db) in the mean streamwise velocity profiles. increase near the wall due to entrainment layer grows because of turbulent the fluidized bed. The boundary dust.

To analyze the flow, we define the following similarity coordinates. that the shock front propagates as a linear function of time: R,(t) and that the shock-induced lines of = W,t

First,

assume

(Al) layer is constant along

flow field above the boundary

x = r/R,

= 1- _ y=z/R,

(A2) (A3)

These similarity lines propagate

with a wave velocity

'
The streamwise

w =wo
velocity in these similarity coordinates becomes:

(A4)

'
This transformation of the boundary

= W - u = xW0- u
modifies the velocity profiles as depicted in Figure A-lb. this case, the velocity begins with the freestream

(A5)
In

value fioo = xW, - uoo at the edge

layer, and increases to a maximum value of fiw = xW_ on the wall.

A-1

Figure -I. Schematic A ofthe turbulent boundary dusty layer induced by a shock propagating a loose (S) along dustyed(FB): b (a) stationary coordinates; Co)' similarity c(xmlinatcs (_. y). A-2

Described here is a control volume analysis of the mass and momentum balance for such turbulent boundary layers, assuming that the mean velocity and density profiles, f -- u/uoo and h- p/poo, are known. Mass Integral Equation

Let Fi represent the mass flux across surface i of the control volume in the similarity coordinates of Figure Al-b. Then the streamwise fluxes across surfaces a and b are given by: .F',. = Fb "" Similarly, p,.fia dy = (sh/ns) pbfibdy --- (6b/Rs) pafiadr/ + Poofioo(Yc - 6a/Rs) Pbfibdr + pccfioo(y_ - 6b/Rs) (A6) (A7)

_0y

_01

/o

/o 1

the mass fluxes through

the bottom and top of the control volume are:

Po= polo a_

(AS)
of

Since the flow is steady in these similarity coordinates, then the conservation mass requires that the sum of the fluxes is equal to zero:

i or

_o- _ +,_o- _ = 0
Solving the above equation
i

(A_0)
(All)
we find the mass

for the streamwise

flux yields:

(_, - Fb)/Az = -poVo+ poovoo


Taking the limit as Ax approaches conservation law: d zero and using d_ -- -dx,

,, where

d-'-_ _''

= povo - poovoo

(Al2)

= (6/n_,)

(pfi - poofic)dr/ A-3


,'.

(Al3)

The latter represents the surplus mass flux (relative to the freestream values) created by the wall boundary layer: The mass conservation law (Eq. Al2) can be nondimensionalized by the mass flux p2u2, yielding: d , where rho = poVo/p2u2 " " H(x) = Poo/P2 F(x)=uoo/u2 (Al5) (Al6) (Al7) [HF 6",Rs] = rho - poovoo/p2u2 (14)

In the above, rho represents the nondimensional mass entrainment rate, and H(x) and F(x) denote the nondimensional flow field above the boundary layer which can be a function of x. In addition, 6., represents the mass thickness of the boundary layer: 6., = 6 where h(x, rl) = p/po _ /(x,_7) = fi/uoo = xWs/uoo , /_(x)=xW_/uoo-foo f(x,71) = u/u_ - f (Al9) (A20) (A21) (A22) (hi]co)&7 (ALS)

. ,

Next, we convert

the above integrand h/-/oo

to lab-fixed velocity profiles (f = u/uoo): - f](h -1) xW_,/uoo + 1 + (1hf) (A23)

= h[xW,/u_ _ W,

u2 F(z)

The expression

for the boundary

layer mass thickness 5.,=I.,5

then simplifies to: (A24)

where u2 F(x)

/o I

(h - 1)drl +

/o'

(1 - hf)drl

(A25)

A-4

Using the above relations, d

the Mass Integral Equation

becomes: (A26) or

d_ [HF I,n 51R8] = rho - HF vooluoo This equation may be integrated to determine boundary layer thickness as a function of _: " 5m(_)/R_ = I.., 5(_)/R, =

the growth of the mass thickness

(rho - HF voo/uoo) d_IHF

(A27)

' o

Thus, mass conservation in the boundary proves that the fundamental cause of dusty boundary layer growth is mass entrainment from the fluidized bed. Note that this is true independent of momentum considerations (e.g., for zero wall drag). If the freestream conditions are independent of x (e.g., in the normal shock case), then the above relations reduce to a particularly simple form: d

d--_ 5/n,] = rho- vooluoo [In


and

(A2S)

6m(_)/R,= sm_(_)/R, =
Momentum Integral Equation

(,hoL' _/u_)a_

(A29)

Now let Pi represent the momentum flux across surface i of the control volume in similarity coordinates. Then the streamwise fluxes across surfaces a and b are given by:

..

Z
_'b = ,
I

/o

pba_ dy +

/o

pbdy (A31)

= (SblR_) Similarly, are: the momentum

I'

-2 (yoo - 5,1Rs) + pbyoo Pbfi_dT1+ PooUoo the bottom

fluxes through

and top of the control volume (A32) (A33)

Po = (poVo_W + ro)Ax -_oo= po_v_fiooAx A-5

Since the flow is steady in these similarity coordinates, then mentum requires that the sum of the fluxes equals zero:

conservation

of mo-

or

F_ - Fb + Fo - Foo = 0 Solving the above equation


lP

(A34)

for the streamwise

flux yields: - ph)lAx (A35)

(F. -/_'b)/Ax

-- --PoVofi_ - ro + p=voofioo - y_(p. zero and using d_ = -dz,

Taking the limit as Az approaches conservation law:

we find the momentum d

d _ = poVo5W + 7"o p_vooftoo - Yoo_ d_ where

(A36)

= (6/n_)
But from the mass conservation

(o_, 2

p_Uoo) d_7 "_

(A37)

law (Eq. A12) we recall that d

poovoo-poVo-"_.ff'm @

This can be used to eliminate

p_v_

from Equation

A36, yielding (A38)

1,

d .--_ _'o = poVoUoo+ rota where

d yoo --;-; p ag

= (6/n.)
A-6

pa(a - fioo) _ d

(A39)

The latter represents the surplus momentum flux (relative to the freestream values) created by the boundary layer. The momentum conservation law (Eq. A38) can be nondimensionalized by the momentum flux p2u_, yielding: d-'_[HF26/Rsl where " Cf=ro/(O.hpccu G(x) -'P_/P2 2) (A41) (A42) = Fgno + HF 2 Cf/2P2 y_ -_ G (A40)

,
o

v_/p_ = (_ -1)/9
In the above, C! represents the local wall drag coefficient which the fluidized bed exerts on the boundary layer, and G(x) denotes the nondimensional pressure above the boundary layer which can be a function of x. In addition, 6o represents the momentum thickness of the boundary layer:

60=6
Next, we convert the above integrand hf(f - f_)=h

hi(i- ]_) d,7


(xW,/uoof) h(1 (l-f) f)hf(1 - f)

(A43)

to lab-fixed velocity profiles (f = u/u_)"

u2 The expression for the boundary

F(x)

(A44)

layer momentum ,5o= Io,5

thickness

then simplifies to: (A45)

where u2 F(x) Using the above relations,


|

h(1 - f)drl .-

hf(1 - f)drl becomes: --_

(A46)

the Momentum

Integral Equation _

d"-_[HF2 Io ,5/R_,] = Fdno + HF 2 Cf/2 This equation may be formally integrated thickness as a function of _:

to determine

the growth in the momentum

_o(_)/a, = Xo 6/R,

"

= -

lo'(F,ho+ HF2 C_/2)d(/HF_


P_ p2u_ HF 2 fo_(6/R,) da d_ "_ (A4S)
A-7

Thus, momentum conservation in the boundary layer demonstrates mentum thickness grows because of three effects: mass entrainment, exterior pressure gradients.

that the mowall drag and

If the freestream flow is independent of _ such as in the normal shock case (where H - F -- G - 1), then the above relations reduce to a particularly simple form: , , and d

d_ [Xo6/R,]

= 'no + C/12

(A49)

6o(_)/R,- zo6/R, =

(,ho+ c_/2)a_

(A50)

A-8

Anda mungkin juga menyukai