Sunteți pe pagina 1din 4

week ending

PRL 106, 184503 (2011) PHYSICAL REVIEW LETTERS 6 MAY 2011

Vortex Formation in a Shock-Accelerated Gas Induced by Particle Seeding


Peter Vorobieff, Michael Anderson, Joseph Conroy, Ross White, and C. Randall Truman
Department of Mechanical Engineering, The University of New Mexico, New Mexico 87131, USA

Sanjay Kumar
Department of Engineering, University of Texas - Brownsville, Texas 78520, USA
(Received 4 November 2010; revised manuscript received 3 February 2011; published 4 May 2011)
An instability forms in gas of constant density (air) with an initial nonuniform seeding of small particles
or droplets as a planar shock wave passes through the two-phase medium. The seeding nonuniformity is
produced by vertical injection of a slow-moving jet of air premixed with glycol droplets or smoke particles
into the test section of a shock tube, with the plane of the shock parallel to the axis of the jet. After the
shock passage, two counterrotating vortices form in the plane normal to that axis. The physical
mechanism of the instability we observe is peculiar to multiphase flow, where the shock acceleration
causes the second (embedded) phase to move with respect to the embedding medium. With sufficient
seeding concentration, this leads to entrainment of the embedding phase that acquires a relative velocity
dependent on the initial seeding, resulting in vortex formation in the flow.

DOI: 10.1103/PhysRevLett.106.184503 PACS numbers: 47.20.k, 07.35.+k, 47.40.Nm

Richtmyer-Meshkov instability (RMI) [1,2] develops on being measurably lower than that of the unseeded gas. This
an initially perturbed interface between two gases of differ- entrainment of gas by multiple particles or droplets would
ent densities subjected to an impulsive acceleration be driven by viscosity, which dominates the flow on very
(shock). Misalignment between the pressure gradient on small scales, somewhat akin to the recently observed en-
the shock front and density gradients in the medium leads trainment of water by swarms of swimming plankton [12].
to vorticity deposition. After a short stage of linear pertur- Here we describe a shock-driven general instability of a
bation growth described by Richtmyer [1], the flow driven gas-droplet or gas-particle mixture with initially nonuni-
by shock-deposited vorticity enters a stage of nonlinear form droplet concentration. While the morphology that
instability growth and eventually transitions to turbulence. emerges is superficially similar to that seen in many RMI
RMI plays an important role in transient processes in experiments [13,14], the physical mechanism responsible
Earths magnetosphere [3], astrophysical phenomena, for the vortex formation is quite different (Fig. 1).
e.g., evolution of supernova remnants [4,5], high energy The data presented in this Letter were acquired using a
density physics, specifically inertial confinement fusion horizontal shock tube with a 75  75 mm cross section.
[6], and supersonic combustion [7]. The driver section at the end of the shock tube is pressu-
Realistically, the shock often propagates through a mul- rized with helium. A planar shock wave is released into the
tiphase medium comprised of an embedding gaseous phase driven section by puncturing the diaphragm initially con-
and embedded droplets or particles. For example, in astro- taining helium in the driver section. The shock wave propa-
physical processes, dusty plasma [8] is ubiquitous. While gation is measured with two pressure transducers separated
the relative velocity of the embedded phase can be zero by a 2.59 m distance along the driven section. A horizontal
prior to the shock acceleration, the high density difference cross section of the flow is illuminated by a sequence of
between the phases leads to the particles and droplets four laser pulses (each approximately 5 ns in duration)
lagging behind the gas after the shock passage. This produced by frequency-doubled Nd:YAG lasers. Images
lag has to be taken into account in experiments where of the flow are captured with a multiple-CCD (charge-
particles are used to trace the flow [9]. In first approxima- coupled device) intensified IMACON 200 digital camera.
tion, the momentum exchange between a small particle and The initial conditions are produced by vertically inject-
the embedding gas transpires via Stokes drag. A real ing a slow ( 6 cm=s) jet of air premixed with micron-
shock-particle interaction is much more complex due to sized dipropylene glycol droplets or submicron particles of
compressibility effects, resulting in Mach number depen- smoke into the optically transparent test section of the
dence of the drag force [10]. Transient forces in compress- shock tube, filled with quiescent unseeded air. This injec-
ible flow can be much greater than the quasisteady drag tion through a 6.35 mm diameter round annular nozzle
[11], leading to greater momentum exchange between the forms a quasi-two-dimensional cylindrical column of
droplets (or particles) and the surrounding gas flow. seeded air, surrounded with a coflow of unseeded air.
A sufficient particle concentration may result in the post- This configuration is known to minimize three-
shock equilibrium local velocity of the gas-particle mixture dimensional initial perturbations [15]. Figure 2 shows

0031-9007=11=106(18)=184503(4) 184503-1 2011 American Physical Society


week ending
PRL 106, 184503 (2011) PHYSICAL REVIEW LETTERS 6 MAY 2011

(submicron sized) form a line parallel to the shock front,


and larger droplets (average size 4 m) trail behind,
forming a tail-like structure, because smaller droplets
achieve momentum equilibrium with the embedding
flow faster than the larger ones. This momentum ex-
change leads to the air-droplet mixture mean velocity
in the section of the flow containing particles being lower
than that of the surrounding particle-free air. Shear be-
tween the areas with lower and higher velocities in turn
leads to formation of counterrotating vortices on both
FIG. 1. Schematic comparison of RMI [(a)(c), top row] and
sides of the column which are apparent in the seventh
seeding-induced instability [(d)(f), bottom row]. RMI is driven
by misalignment between the pressure gradient on the shock and subsequent images in Fig. 2.
[rp, (a), (d)] and density gradient on the interface between two The morphology that develops at later times as the result
gases [r, (a)]. This misalignment leads to vorticity deposition of this shock-driven interaction is superficially similar to
on the interface (b), and to subsequent vortex roll-up (c). In the that of a cylinder of heavy gas initially embedded in lighter
seeded flow (d), there is no macroscopic density interface. gas and accelerated by a planar shock [16], where two
Immediately after shock interaction, seeding particles lag behind counterrotating vortices form due to RMI. However, what
the gas accelerated to piston velocity, acquiring relative veloc- happens in the two cases on the microscopic scale is quite
ities ui (e) and interacting with the surrounding gas. As the different. As there is no gaseous density interface, there is
particle and gas velocities reach equilibrium, the average veloc- no baroclinic vorticity deposition. On the level of indivi-
ity of the medium is lower than that of the unseeded gas, leading dual droplets (or particles), details of the interaction can be
to shear and vortex roll-up (f).
different. For example, droplets might deform as shock
propagates through them, or even break up. However,
flow evolution in the laser-illuminated cross section of the seeding the injected flow with droplets or with particles
droplet-seeded column, as it and the surrounding air are of smoke (Fig. 3) produces similar macroscopic morpho-
accelerated by a planar shock front moving at 570 m=s logies. Figure 3 compares the evolution of the flow with a
(Mach number M 1:66  0:02, with small variations smoke- or droplet-seeded air column with that where the
from experiment to experiment). The shock passage accel- injected gas is sulfur hexafluoride SF6 , which is about
erates the air to a piston velocity of about 303 m=s. 5 times more dense than air. With SF6 injection, an initially
The glycol droplets are about 850 times more dense than cylindrical, diffuse interface forms between the heavy gas
air before shock compression and occupy a small ( 5%) and the lighter gas (air), resulting in a density gradient
volume fraction of the injected column. being added to the initial conditions, producing RMI upon
The initial conditions of the experiment are nearly two shock passage. The images in Fig. 3 were acquired at the
dimensional, with little variation in the direction normal same downstream location, at three Mach numbers for
to the plane of view. In the earlier images of the each case (RMI and two cases of seeding-induced flow).
sequence of Fig. 2 (third to fifth), the smallest droplets In the case of RMI, we still use droplet seeding, as many
earlier experiments also did, for the purpose of flow visual-
ization. Strictly speaking, this Letter demonstrates that the
presence of these droplets may alter the flow, however, as
the following discussion will show, in the presence of a
shock-accelerated density interface, the flow is dominated
by vortices developing from baroclinic vorticity deposi-
tion, and the vortex roll-up due to droplets alone is com-
paratively much weaker. The counterrotating vortex pairs
are evident in all the cases. However, for the case of RMI,
another feature gains prominence as the Mach number
increases, namely, a central spike [17,18] that emerges
due to shock focusing in the heavy gas. This feature is
not apparent in the images of the smoke- and droplet-
seeded column. RMI images also manifest small-scale
structures that form due to secondary instabilities and
FIG. 2 (color). Mosaic of six experimental image sequences
showing the evolution of a shock-accelerated column of glycol
lead to the flow transitioning to turbulence. In the case of
droplets in air. Shock direction is left to right. Mach numbers and RMI, the flow behavior at small scales associated with
exposure timings are marked in the figure, with time t 0 these features is known to become statistically consistent
corresponding to the shock wave reaching the center of the with models of fully developed turbulence [16].
initial conditions (IC). Numbers show downstream distance The small-scale structures in the right column of Fig. 3
from the IC center in mm. arise both due to shear (Kelvin-Helmholtz instability) and
184503-2
week ending
PRL 106, 184503 (2011) PHYSICAL REVIEW LETTERS 6 MAY 2011

the shock (in our case, the piston velocity with the negative
sign), a0 is the preshock amplitude of the initial interface
perturbation, k its characteristic wave number, and A is the
Atwood number comprised of the densities of the gases on
the upstream (1 ) and downstream (2 ) sides of the
interface with respect to the direction of the shock: A
2  1 =2 1 . In the case of cylindrical initial con-
ditions with a characteristic radius r0 , k =r0 and
a0 r0 . Realistically, the vimp growth rate represents the
initial slope to the curve describing the actual interfacial
growth, which is nonlinear. For the case of RMI on an
air-SF6 interface, A  0:67. For our experiment with drop-
lets, we measured the average density of the droplet-seeded
FIG. 3 (color). Evolution of a smoke-seeded air column (top air (1:26  0:01 kg=m3 ) and used that to compute the ef-
row), droplet-seeded air column (middle row) and droplet-SF6 fective Atwood number A  0:03. The consistency of the
column (bottom row) at the same downstream position for smoke fluctuates from experiment to experiment, within the
different Mach numbers. Original images are illuminated with range of effective Atwood numbers A  0:02  0:01.
a green laser, false color is added to distinguish between lower The top plot of Fig. 5 shows a comparison of the
(red), intermediate (green) and high (blue) Mach numbers. The instability growth for RMI and the instability of the
shock direction is from left to right, the image extent in the droplet-seeded air column as the function of downstream
streamwise (horizontal) direction is from 152 mm to 181 mm distance from the initial conditions. The data were acquired
downstream. Mach numbers, timings, and dimensionless times  in about 200 experimental runs (resulting in 572 amplitude
(see following text for definition) are labeled.
measurements), conducted in the range of Mach numbers
from 1.22 to 2.02. In all cases, the amplitude initially
to the secondary baroclinic instability induced by pressure- decreases after the shock compression. Phase inversion
density misalignment in the cores of the vortices com- effects on the downstream side of the gas cylinder are
prising the counterrotating pair. While shear-driven known to play a role in this decrease for RMI.
secondary structures may be present in the late-time Subsequently the perturbation amplitude begins to grow,
(t > 1 ms) images of the droplet-column instability, we much faster for the case of RMI, as the plot shows. It is also
did not see evidence of transition to turbulence for that apparent that, for each case, the instability amplitude as the
case. Another notable feature of Fig. 3 is the rather weak
dependence of the size of the counterrotating vortex pairs
on the Mach number.
In many earlier experiments, tracer particles or droplets
were assumed to act as a passive scalar. While this study
makes it clear that such an assumption is generally not
valid, in the presence of RMI, its much faster growth can
dominate the postshock flow (Fig. 4).
Despite the obvious differences in the formation of the
instability of impulsively accelerated two-phase flow we
describe here and in the formation of RMI in gas with a
density interface, it is tempting to assess to which extent the
two shock-driven instabilities can be related. For RMI, the
initial perturbation growth rate according to Richtmyers
theory is vimp UAka0 , where U is the difference
between the mean velocity of the interface before and after

FIG. 5. Streamwise perturbation amplitude w as the function


FIG. 4 (color online). Early-time evolution of an air-droplet of downstream distance in dimensional (top) and dimensionless
column (top) and SF6 -droplet column (bottom) for M  1:66. (bottom) coordinates. Dashed line in the bottom plot represents a
Interval between frames: 25 ms, horizontal image extent: 24 mm, single-parameter fit to the experimental data for the SF6 -droplet
shock direction: left to right. column at M  1:66 using the modified Jacobs correlation [9].
184503-3
week ending
PRL 106, 184503 (2011) PHYSICAL REVIEW LETTERS 6 MAY 2011

function of downstream distance indeed depends rather measurements for the air-droplet mixture. However, flow
weakly on the Mach number. features present in RMI are absent in the case of the insta-
For RMI, the instability amplitude growth curves pro- bility of the droplet-seeded column, namely, the spike due
duced from the same initial conditions at different Mach to shock focusing and the secondary baroclinic instability in
numbers can be collapsed by plotting them in appropriate the cores of the counterrotating vortices. Vice versa, the
dimensionless coordinates [19]. The dimensionless time  trailing tail of more massive particles characterizing the
relates to time after shock acceleration t as  2kAjUjt. evolution of the air-droplet column is not prominent in
For the same initial conditions and A, plotting the RMI evolution. Experiments at later times (and farther
RMI amplitude as the function of downstream distance downstream distances) will elucidate these differences.
x jUjt should also cause the growth curves for differ- Do we observe a new instability in the narrow sense of
ent Mach numbers to collapse with appropriate scaling of the word? Although the vortices in shocked multiphase
the vertical axis. This notion is supported by examining the flow form due to a combination of known physical mecha-
values of  for the test cases of the Mach and Atwood nisms, the resulting flow is similar to RMI, but evolving in
number study presented in Fig. 3 and noting how the an entirely different medium. This instability (perhaps in a
evolution of the vortex structures is related to those values. more general sense) can occur in a wide variety of impul-
In the bottom plot of Fig. 5, the horizontal coordinate is sively accelerated two-phase flows (e.g., in shocked cosmic
rescaled in terms of dimensionless time , and the vertical dusty plasma). Moreover, it can be one of a class of similar
coordinate is nondimensionalized by w0 the average instabilitiesfor example, nonuniform droplet seeding
minimum width of the perturbed column after the shock of the top layer of a fluid or gas in a gravity field would
acceleration for each subrange of Mach numbers (M produce an analog of Rayleigh-Taylor instability.
1:22  0:01, M 1:66  0:02, M 2:02  0:01)to ac- We thank Professor S. Balachandar (U. of Florida) for
count for compression and phase inversion effects. Note helpful suggestions. This research is funded by the US
that this nondimensionalization should make the initial National Nuclear Security Agency (NNSA) through
slopes of the instability growth according to Richtmyers DOE Grant No. DE-PS52-08NA28920 and by the US
theory (after the compression and phase inversion effects Defense Threat Reduction Agency (DTRA) under Awards
took place) equal for all M and A. No. HDTRA1-07-1-0036 and No. HDTRA1-08-1-0053.
Notably, nondimensionalization produces a plot with
fairly consistent growth trends both for RMI and for the
instability of the two-phase droplet-seeded column. This
may suggest that the long-term behavior of the shocked [1] R. D. Richtmyer, Commun. Pure Appl. Math. 13, 297
two-phase medium is consistent with that of continuous (1960).
medium with the same average density. This result is less [2] E. E. Meshkov, Fluid Dyn. 4, 101 (1969).
surprising when two contributing factors are taken into [3] C. C. Wu, J. Geophys. Res. 105, 7533 (2000).
consideration. First, both in the case of RMI and in the [4] R. A. Chevalier, J. M. Blondin, and R. T. Emmering,
case of seeded and shocked flow, the dominant vortex Astrophys. J. 392, 118 (1992).
structure is a pair of counterrotating vortex columns, and [5] J. Kane, R. P. Drake, and B. A. Remington, Astrophys. J.
it is the roll-up of the shocked material into these columns 511, 335 (1999).
that determines the shape of the nonlinear growth curve. [6] V. N. Goncharov, Phys. Rev. Lett. 82, 2091 (1999).
Second, the effective Atwood number we use implicitly [7] J. Yang, T. Kubota, and E. E. Zukoski, AIAA J. 31, 854
(1993).
accounts for the momentum transfer between the embed-
[8] D. A. Mendis and M. Rosenberg, Annu. Rev. Astron.
ding and embedded phases. Astrophys. 32, 419 (1994).
The dashed curve in the bottom plot is produced by using [9] P. M. Rightley, P. Vorobieff, and R. F. Benjamin, Phys.
flow parameters measured during the experiments: the wave Fluids 9, 1770 (1997).
number k known from the initial conditions, the postshock [10] M. K. Parmar, A. Haselbacher, and S. Balachandar, AIAA
minimal streamwise width of the perturbed column w0 , and J. 48, 1273 (2010).
the corresponding time t0 . Their values are inserted into a [11] M. K. Parmar, A. Haselbacher, and S. Balachandar, AIAA
semianalytical equation [9] developed from an earlier Paper 2009-1124, 2009.
formula [20] known to faithfully describe nonlinear [12] K. Katija and J. O. Dabiri, Nature (London) 460, 624
vortex growth after shock acceleration, although with a (2009).
somewhat different initial geometry [9,19]: wt [13] J. W. Jacobs, Phys. Fluids A 5, 2239 (1993).
[14] C. Tomkins et al., J. Visualization 5, 273 (2002).
2k1 sinh1 k2 t  t0 sinhkw0 =2. Here we can re-
[15] B. J. Balakumar et al., Phys. Fluids 20, 124103 (2008).
place t with x=jUj. With this formula, we curve-fit our [16] P. Vorobieff et al., Phys. Rev. E 68, 065301 (2003).
experimental data for RMI at M  1:66 using the shock- [17] S. Kumar et al., Phys. Fluids 17, 082107 (2005).
deposited circulation (that we do not measure directly) as [18] A. Palekar, P. Vorobieff, and C. R. Truman, Progr.
the sole fit parameter. After the fit curve is rescaled Comput. Fluid Dynam. Int. J. 7, 427 (2007).
consistently with the experimental data, it agrees reason- [19] G. C. Orlicz et al., Phys. Fluids 21, 064102 (2009).
ably well with both the measurements for SF6 and the [20] J. W. Jacobs et al., J. Fluid Mech. 295, 23 (1995).
184503-4

S-ar putea să vă placă și