Вы находитесь на странице: 1из 7

PHYSICS OF FLUIDS 22, 024103 2010

Viscous linear stability of axisymmetric low-density jets: Parameters


influencing absolute instability
V. Srinivasan,a M. P. Hallberg, and P. J. Strykowski
Department of Mechanical Engineering, University of Minnesota, Minneapolis, Minnesota 55455, USA

Received 20 May 2009; accepted 18 December 2009; published online 19 February 2010
Viscous linear stability calculations are presented for model low-density axisymmetric jet flows.
Absolute growth transitions for the jet column mode are mapped out in a parametric space including
velocity ratio, density ratio, Reynolds number, momentum thickness, and subtle differences between
velocity and density profiles. Strictly speaking, the profiles used in most jet stability studies to date
are only applicable to unity Prandtl numbers and zero pressure gradient flowsthe present work
relaxes this requirement. Results reveal how subtle differences between the velocity and density
profiles generally used in jet stability theory can dramatically alter the absolute growth rate of the
jet column mode in these low-density flows. The results suggest heating/cooling or mass diffusion
at the outer nozzle surface can suppress absolute instability and potentially global instability in
low-density jets. 2010 American Institute of Physics. doi:10.1063/1.3306671
I. INTRODUCTION

Many fluid systems yield nonlinear synchronized structures of inherent originsglobal modesand these are
known to exist in certain jets,1,2 wakes,3,4 and shear layers.5
The precise origin of these dynamics is believed rooted in
the linear stability of the underlying flow. Recently, remarkable agreement between the marginally absolutely unstable
frequency and the nonlinear global frequency6 in a synthetic
wake, as well as the corresponding spatial growth rates,
strongly support the presence of a steep nonlinear global
mode. These results are in full agreement with theoretical
predictions7 showing that model flows can be governed by
fronts acting as wavemakers positioned at the convective-toabsolute transition in the flow; the resulting nonlinear evolution is directly connected to the local linear characteristics.
Similar excursions have been undertaken in the low-density
axisymmetric jet,8 as well as the swirling jet9 where again
the frequencies and spatial growth rates support the presence
of a steep nonlinear global mode with close ties to the
convective-to-absolute transitions in the underlying base
flows.
Clearly, the nonlinear evolution of these self-sustaining
flows is intimately connected to the underlying linear stability and in order to accurately predict the dynamics using
linear tools, the precise details of the flow development are
required. Therefore, in the case of the low-density jet, the
spatial evolution of both the velocity and density profiles are
required, since it is the unique combination of the two profiles at a particular location in the flow that most likely establishes the nonlinear dynamics. To this end, a brief investigation was carried out by Raynal et al.10 in an attempt to
explain the differences observed between the critical density
ratio in a low-density jet created via heating compared to one
created using a pure gas. They showed in the case of an
a

Present address: Department of Mechanical Engineering, University of


California Berkeley.

1070-6631/2010/222/024103/7/$30.00

inviscid planar jet that the precise details of the profile have
a dramatic impact on the absolute growth rates; ultimately
the absolute growth can become negative if the velocity and
density profiles are sufficiently displaced.
Recently, it has been shown that the velocity and density
profiles in a low-density jet are indeed different in certain
situations.11,12 These studies focused on either fully developed flows11 or less developed flows but somewhat further
downstream than might be of interest;12 they nonetheless
suggest the low-density jet velocity and density profiles are
dissimilar in the exhaust plane. Three things are apparent at
the nozzle exit: 1 diffusion will occur no matter how the
density difference is created; 2 heating will only lead to an
increase in diffusion via conduction into the nozzle hardware; and 3 viscosity acts on the velocity field upstream of
the jet exit and this action is independent, at least in the
incompressible case, of the fluid choice insofar as the velocity profile in the nozzle exit plane is known to be a function
of the Reynolds number. Based on these facts it is easy to
envision conditions at the jet exit where the density profile is
much steeper than the velocity profile regardless of the
choice made in creating the density difference. Heating will
only lead to an increase in diffusion where, in the limiting
adiabatic case, it will be identical to binary gas diffusion
assuming identical Prandtl and Schmidt numbers. To help
shed light on the change in stability due to subtle differences
between velocity and density profiles at the nozzle exit, the
following study has been undertaken. It is intended to extend
the work of Raynal et al.10 and includes a comprehensive
investigation into the viscous linear stability of low-density
jet profiles including systematic differences between the velocity and density profiles that might be observed very near
the nozzle exit. The work also independently considers the
role of Reynolds number and external coflow.
The study was motivated in part by the fact that global
instability in jets is generally only observed in relatively
clean laboratory flows or computationally idealized jets as

22, 024103-1

2010 American Institute of Physics

Downloaded 30 Sep 2010 to 202.3.77.11. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

024103-2

Phys. Fluids 22, 024103 2010

Srinivasan, Hallberg, and Strykowski

pointed out by Lesshafft.13 This is in sharp contrast to the


rather ubiquitous nature of global instability observed in the
case of bluff body wakes. It will be shown in what follows
that by independently varying the velocity and density profiles, subtle displacements and/or shape differences between
the velocity and density profiles can dramatically alter the
absolute growth of the jet-column mode leading to an effective control technique much like base-bleed and suction in
wake flows.1416
II. STABILITY THEORY

In what follows, we use spatiotemporal theory to explore


the linear stability of a viscous parallel flow, specifically, a
variable density round jet with coflow. In general, the variation in density in the jet shear layer will depend on the
mechanism used to generate the nonuniform density
profileheating the jet versus the use of dissimilar fluids for
the jet and external streams. Nonhomogeneous jets typical of
laboratory flows have a steep, nearly top-hat density profile
at the nozzle exit, while heating a jet results in a density
variation that is comparatively less steep at the nozzle exit
largely due to the difficultly of achieving the adiabatic
boundary condition in the laboratory. In general, therefore,
the density thickness which we define subsequently is not
equal to the velocity shear layer thickness, which arises due
to boundary layer development on the nozzle walls. Further,
for the same reasons outlined above, the points of inflection
of the density and velocity profiles do not generally occur at
the same radial location.
We elect to model a round jet in coflow by using the
velocity profile of the form

Ur = 1 + R1 1 + R tanh bu

1
r
r

the velocity ratio R = Uc U / Uc + U is a measure of the


strength of the external coflowing stream, and the subscripts
c and correspond to the velocities on the jet centerline and
at large radial distances, respectively. This profile matches
laboratory velocity profiles to good agreement in the nearfield of the jet, and has the advantage of having been extensively investigated in theoretical studies.17,18 The density
profile is chosen to satisfy the BusemanCrocco19 relation
and is given by r = 2 / Fr, where

Fr = 1 + S + 1 Stanh b

1
r
r

with S = c / . For nonunity Prandtl numbers we include a


radial separation between the density and velocity profiles by
a distance , so that r = 2 / Fr . One can define the
extent of the regions where density and velocity variation are
significant by means of the density and momentum thicknesses

u
=
r1/2

Ur U
Ur U
1
dr.
Uc U
Uc U

For a constant density and the prescribed velocity profile, we


get D / u = 8bu, which is the definition we use henceforth. A

FIG. 1. A representative velocity profile as well as shifted and unshifted


density profiles.

density thickness is computed in the same manner resulting


in the dimensionless parameter describing the density profile
shape D / = 8b. Figure 1 shows the velocity profile Ur
for a relatively thick shear layer having D / u = 20. The
scaled density profile r S is plotted for D / = 100 with
and without a shift of = 0.1; note that the subscripts u and
indicate velocity and density profile characteristics, respectively.
We consider the linearized NavierStokes equations for
infinitesimal axisymmetric disturbances and restrict our attention to modes with zero helicity i.e., m = 0. Further, in
keeping with previous studies,17,18 gravity has been neglected i.e., Richardson number approaching zero. From
these assumptions one obtains the OrrSommerfeld equation,
viz.,

c U +

+ i Re1 iv 2
+

1
U
U
2 +
+ U

+ 2 22
r
r

22 3
3 + 4 = 0,
r
r

which is a generalized eigenvalue problem in = rur. The


Reynolds number is defined based on the jet centerline velocity, density, and viscosity. The length scale in Re is the jet
half-width, i.e., the radial distance at which the axial velocity
ratio falls by a factor of 2.
For a given velocity and density profile specified by S,
R, D / u, D / , and and Reynolds number, one can solve
the dispersion relation for axisymmetric disturbances,
namely the OrrSommerfeld eigenvalue problem stated
above, and obtain the phase velocity spectrum c = for
the wave packet. It can be shown20 that the most highly
amplified waves in time say, with r = m occur when the
wavenumber is real, while the most highly amplified waves
in space occur when the frequency is real. Following Bers,21
a test profile is absolutely unstable if the mapping =
gives rise to saddle points o points where d / d = 0 in the

Downloaded 30 Sep 2010 to 202.3.77.11. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

024103-3

Viscous linear stability of axisymmetric low-density jets

Phys. Fluids 22, 024103 2010

complex wavenumber plane, subject to the conditions that


a the imaginary part of is positive; b that m; and
c the saddle point is the result of the pinching of two waves
of the same wavenumber, one traveling upstream and the
other traveling downstream. To obtain the absolute/
convective instability boundary in the parameter space, we
look for saddle points where o,i = 0, and conditions b and
c are satisfied.
To find saddle points of the dispersion relation, we follow the method proposed by Deissler.22 The method consists
of expanding = in a Taylor series about three points
1 , 2 , 3 whose corresponding images 1 , 2 , 3 in the
complex frequency space have been calculated. With the requirement that d / d = 0 in each of these expansions, an
improved guess for the saddle point can be generated. This
procedure is repeated until the change in absolute value of
and is less than 103 in successive iterations.
The numerical technique of Chebyshev collocation is
employed to solve the generalized eigenvalue problem. In
brief, the solution procedure involves the assumption that the
solution can be expanded as the sum of a finite number of
Chebyshev polynomials with undetermined coefficients an.
The polynomials we use in this study are the renormalized
functions TLn, developed specifically for use on a semiinfinite interval; see Boyd.23 The expansion is substituted
into the differential equation and enforced at N collocation
points. The collocation points are given by the mapping from
, to 0 ,
ri =

Rmap1 cost/2
,
sin2t/2 +

with the constants given by Rmap , = 1.9, 0.03. Here, ti


= 2i + 1 / 2N + 2, i = 0 , 1 , 2 , . . . , N. The Chebyshev functions are defined by
TLnr = cosnt,

FIG. 2. a Contours of i in the complex -plane for S = 0.5, Re= 5000,


U / Uc = 0.025, D / u = 25.2, D / = 70, and = 0.1. b Axial velocity disturbance function at the saddle point o.

n = 0,1,2 . . . .

This mapping concentrates points at Rmap / 2 and is functionally identical to that used by Ref. 18 except for the change in
constants from Rmap , = 1.8, 0.036. The transformed differential equation, when enforced at the collocation points,
results in a matrix equation of the form Ax= Bx, where is
the generalized eigenvalue and corresponds to the undetermined coefficients an. The complex eigenvalue problem is
solved using QR factorization, which is implemented in the
intrinsic function ZGGEV mathematical library of LAPACK.
For each calculation, 160 points are used to obtain results
independent of grid size.

III. PARAMETRIC EFFECTS: VISCOSITY, COFLOW,


AND DENSITY FIELD

We are interested in examining the role of finite Reynolds number, ambient coflow, and the density field on the
absolute and convective instability of axisymmetric jets. Experiments of Kyle and Sreenivasan24 and Hallberg et al.25
clearly indicate that suppression of the self-excited oscillations in low-density jets occurs when the momentum

thickness of the separating shear layer is small relative to the


jet diameter. Raynal et al.10 suggested that this may be due in
part to the relative position of the velocity and density discontinuities, which they studied in planar jets. Here we extend their work to axisymmetric jets, but also include the role
of viscosity and ambient coflow.
An example of solving the dispersion relation is shown
in Fig. 2a, where contours of i are shown in the complex
wavenumber plane for a representative parameter set S
= 0.5, U / Uc = 0.025, D / u = 25.2, D / = 70, = 0.1, Re
= 5000. It is seen that at o = 0.9655 1.3640i the dispersion
relation has a saddle point, and that for higher values of i,
the contours move into the upper and lower halves of the
complex wavenumber plane. Examination of the axial disturbance uz corresponding to the saddle-point mode, shown in
Fig. 2b, reveals that the axial fluctuations have a peak in
the shear layer but are nonzero at the jet axis, and hence the
most unstable eigenmode is the jet-column mode identified
by Jendoubi and Strykowski.17

Downloaded 30 Sep 2010 to 202.3.77.11. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

024103-4

Srinivasan, Hallberg, and Strykowski

Phys. Fluids 22, 024103 2010

FIG. 4. Effect of viscosity on the absolute-convective instability boundary


for S = 0.5; as in Fig. 3, the momentum and density thicknesses are linked.

FIG. 3. Absolute-convective inviscid stability boundaries for a family of


low-density jets with linked velocity and density profiles D / u = D / ;
= 0.

A. Effects of ambient coflow for inviscid jets

We first consider the case of an inviscid axisymmetric jet


issuing into an external coflowing stream. Results are presented in Fig. 3 for density ratios ranging from 0.1 to 0.7,
with similar and unshifted velocity and density profiles i.e.,
D / u = D / and set to zero. For a particular choice of
density ratio, the region below the stability curve is absolutely unstable and the region above is convectively unstable.
The calculations show that for each density ratio, there is a
finite amount of coflow that suppresses absolute instability,
with the required coflow level decreasing as the density ratio
rises from 0.1 to 0.7. Note that jets issuing into quiescent
surroundings U = 0, hereinafter referred to as free jets experience a transition from absolute to convective instability
if the shear layer thickness is sufficiently large compared to
the jet diameter, namely by decreasing D / u for axisymmetric disturbances.17 A similar transition is observed for large
values of D / u and a density ratio of 0.7 based on inviscid
theory; this nonmonotonic behavior for modest density ratios
allows a critical density ratio to be defined for free jets at
S = 0.713 and D / u 55 and agrees well with other
calculations.18,26 Experimental values of the critical density
ratio range from approximately 0.6 for axisymmetric jets24 to
0.9 in heated planar jets.27 It should be pointed out however,
in experiments conducted with helium jets issuing into air
S = 0.14, self-excited oscillations have not been observed
for sufficiently large D / u;2427 clearly, a result not observed
in Fig. 3.
B. Reynolds number effects

The frequency scaling as well as the observed boundaries of convective-absolute instability in the ReD / space
found by Hallberg and Strykowski28 indicate that the Reynolds number is an independent parameter. Computationally,
it is easy to isolate the two effects and explore the independent role of Reynolds number on stability properties. We

present the absolute-convective stability boundary for various Reynolds numbers and a fixed density ratio of S = 0.5 in
Fig. 4; the velocity and density profiles have D / u = D / ,
and = 0. The most significant effect of the Reynolds number
is to stabilize the jet for thin shear layers, similar to the
effects observed by Lesshafft and Huerre18 for a free jet. For
sufficiently low Re, absolute instability is significantly suppressed relative to the inviscid results, with a reduction in the
maximum coflow level required to suppress absolute instability for the entire range of D / u.
C. Effects of a steep density profile

As a way to investigate the stability response to situations where Pr 1, we chose to vary the density profile independent of the velocity profile through the addition of a
density profile factor and a shift parameter . We now
study two separate effects: 1 having a steep but constant
density profile, while varying the velocity profile independently; and 2 introducing a displacement between the
density and velocity profiles. Both these effects have been
studied by Raynal et al.10 for planar jets with zero viscosity,
in the context of explaining the differences in the behavior of
nonhomogeneous and heated jets. Raynal et al. found that
the effect of steep density profiles was to make the jet more
unstable, and rejected it as a reason for the disappearance of
the unstable mode at high values of D / u, pointing instead to
the stabilizing influence of . As we have just shown in Fig.
4, this stabilization may partly be due to the effect of viscosity, as well.
Figure 5 shows the effect of steepening the density profile without offset = 0 at fixed S of 0.5, and two values of
the Reynolds number, Re= 1000 and 5000. Solid lines are
drawn for the case of D / = D / u, while dashed lines denote
a fixed D / = 70 and 100. For the choice of profiles employed here, the density gradient leads to greater instability
when steeper than the velocity gradient, while the converse
is true when D / u D / . Also noteworthy, for these profiles, is the shifting toward smaller values of D / u at which
the maximum amplification occurs for fixed D / .
Next we explore the effect of various values of at a
fixed Reynolds number. Figure 6 examines the behavior of
o,i for = 0, 0.075, 0.1, and 0.125 for Re= 5000, S = 0.5, and

Downloaded 30 Sep 2010 to 202.3.77.11. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

024103-5

Phys. Fluids 22, 024103 2010

Viscous linear stability of axisymmetric low-density jets

surements of Kyle and Sreenivasan25 Figs. 18 and 19 in


their paper who observe peak fluctuation intensities near
D / u 35 for S = 0.5, and report that as D / reaches its upper limit, the base-to-peak intensity varies relatively slowly.
So although there is no rigorous justification for the use of
the imposed shift and shape factor in the density profile,
several features appear consistent with laboratory observations. Finally, the inset to Fig. 6 illustrates the effect of an
inward or outward radial shift 0.125 on profiles with
D / = D / u. While subtle differences are evident, qualitatively the behavior is the same with regard to positive and
negative displacements, namely both effects are stabilizing.
IV. DISCUSSION

FIG. 5. Effect of steepening the density profile on the amplification rate i,o.
For a particular density field, the upper and lower curves correspond to
Reynolds numbers of 5000 and 1000, respectively.

U / Uc = 0. In order to study the effects of displacement


alone, we keep the density profile fixed at values of D /
= 70 solid line and 100 dashed line. The effect of adding a
displacement is quite dramatic, especially for thin shear layers. For sufficiently large displacements and/or thin shear
layers, the flow field under consideration is such that the
velocity variation occurs in a region of nearly constant density, while the density variation occurs in a region of nearly
constant velocity. The stability characteristics of the jet are
markedly altered, with a transition from absolute to convective instability occurring for increasing D / u values ranging
from 60 to 120 depending on the value of . The amplification rate increases rapidly at low D / u, reaches a maximum
at D / u 40, and then decreases slowly as shear layer thickness decreases. This behavior compares well with the mea-

FIG. 6. Effect of shifting the density profile on the amplification rate i,o
S = 0.5, Re= 5000, U / Uc = 0. For a given shift, the solid and dashed
curves correspond to D / of 70 and 100, respectively. Inset illustrates role
of positive and negative shift when D / = D / u.

One interesting feature of the above findings is the potential to use these ideas as a means to control low-density
jet self-excitation. It has been fairly well established that
there exists a unique connection between absolute instability
and self-excited global modes. To this end, unique solutions
to control these flows can be sought using theory as a
jumping-off point. Clearly the results laid out above reveal
the potential to use radial differences between velocity and
density profiles to augment or suppress flow self-excitation.
Imagine either through heating or variable species, a selfexcited low-density jet is established. In an experiment it
would be rather straightforward to devise a means to shift
either the density profile or temperature profile radially using
heating/cooling or mass diffusion along the nozzle. Neglecting buoyancy, which may or may not be justified,2931 the
velocity profile at the nozzle exit will be unchanged, while
the density profile can readily be shifted until the absolute
growth rate becomes negative suppressing the self-excited
global mode altogether. This type of control is in the same
spirit as cylinder heating and base-bleed or suction applied to
wake flows.1416
Another motivation for conducting this stability analysis
was to explore potential explanations for the lack of experimental evidence for self-excitation in low-density jets at
large D / u. Kyle and Sreenivasan24 were the first to systematically document the absence of self-excitation in laboratory
low-density jets with thin shear layers; recent work corroborated those findings in the presence of an external stream,25
and as a function of D / u and Reynolds number.28 Furthermore, high-speed exhaust flows will typically satisfy the
nominal conditions for absolute instabilitylow S, high Re,
and large D / uyet do not display global self excitation. It
has been shown above that the independent effects of Reynolds number, steepened density gradients, profile offsets,
and coflow can lead to the stabilization of axisymmetric jets
having thin shear layers, though none taken alone appear to
fully explain the experimental findings.
We begin here by comparing the results of calculations
for coflowing jets with S = 0.5 for two scenarios: an inviscid
jet with BusemanCrocco velocity/density fields in one case
and another with Re= 5000, D / = 70, and = 0.1, see Fig. 7.
Contrary to the inviscid calculations, which do not predict
the free jet to be convectively stable at S = 0.5 in the limit for
thin shear layers, inclusion of the above parameters in the

Downloaded 30 Sep 2010 to 202.3.77.11. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

024103-6

Srinivasan, Hallberg, and Strykowski

Phys. Fluids 22, 024103 2010

FIG. 7. Comparison of the absolute-convective stability boundary for lowdensity jets with coflow. A standard theoretical profile using inviscid theory
is compared to a profile selected to mimic representative laboratory
conditions.

stability calculation results in an absolute-convective stability boundary that mimics the behavior of laboratory jets
whose global modes are suppressed for relatively low levels
of coflow for thin shear layers see in particular Fig. 11 in
Ref. 25.
The experiments of Hallberg and Strykowski28 varied
Reynolds number and D / u independently using nozzle extensions of various lengths, producing curves for the globalmode transition. At a density ratio of 0.5, they observed a
finite region in the ReD / u space, where the jet was observed to be globally unstable. Their results indicated that the
transitional Reynolds number would be facility dependent,
and for sufficiently strong nozzle contractions, the selfexcited oscillations would be absent altogether. While we do
not presume to have identified a universal model profile, we
did explore the stability properties as a function of Reynolds
number and D / u using the profile from Fig. 7.
Figure 8 shows the absolute-convective instability
boundary for jets having S = 0.5, D / = 70, and = 0.1 with
subtle ambient coflow 2% and for a free jet. As observed in
Fig. 7, the influence of even small levels of external flow has
a marked impact on the stability boundary in the ReD / u
space, particularly for thin shear layers. The surprisingly
strong role of external flow on jet stability suggests that
changes in nozzle hardware affecting near-field entrainment
e.g., knife-edge, block nozzle, and confinement will undoubtedly alter the global flow dynamics, bringing into question the ability to identify a truly universal stability curve in
the laboratory setting.
The results of Fig. 8 may help explain some of the experimental difficulties in identifying the underlying governing parameters responsible for low-density self-excited oscillations. We have plotted on Fig. 8 the Reu relationship for
several experimental facilities, labeled 15 on the figure; the
expected 21 -power dependence is borne out as would be expected for initially laminar boundary layers. A comparison of
the uppermost experimental curves curves 1 and 3 and the
stability boundary U / Uc = 0 suggests that conditions for
absolute instability may not be supported for free jets in

FIG. 8. Stability boundaries as a function of Reynolds number and momentum thickness S = 0.5, D / = 70, = 0.1. Shaded region corresponds to
experimentally determined region of global instability for S = 0.5 from
Ref. 28.

those facilities. Interestingly, while Kyle and Sreenivasan24


observed self-excited oscillations curve 1, Fig. 8, Hallberg
and Strykowski28 did not for L / D = 0 curve 3, Fig. 8
while a slightly longer extension section curve 4, Fig. 8 did
yield global oscillations near Re= 3200. The domain over
which global oscillations were observed in the experiments
of Hallberg and Strykowski,28 is indicated by the shaded region in Fig. 8.
At first impression the agreement between theory and
experiment is not remarkable. While the model profile is
perhaps representative of low-density jets, the theoretical
curve for U / Uc = 0 and the shaded region for free jets at S
= 0.5 are not coincident. The theoretical predictions certainly
capture the nontrivial dependence of the jet physics on Reynolds number and D / u, but are confounded by the strong
sensitivity of the stability transitions with low levels of ambient coflow, which in the laboratory will be a function of the
near-field jet entrainment. It is further complicated by the
sensitivity of the jet stability to subtle shifts between the
velocity and density field, as first illustrated by Raynal
et al.10 Generally however, given the nature of experimental
facilitieswith laminar separating boundary layersand the
stability transitions in Fig. 8, in the limit of high enough Re
and D / u, low-density jets could be either absolutely or convectively unstable, depending on the experimental trajectory
taken through the ReD / u plane.
While the exact velocity and density fields at the nozzle
exit remain to be measured, the stability results presented
suggest upstream mass diffusion or heating in the ambient
stream near the nozzle surface can lead to the elimination of
absolute instability and hence global mode suppression. Further it has been speculated through shifting the density

Downloaded 30 Sep 2010 to 202.3.77.11. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

024103-7

profile and changing its shape from that of the velocity profile, some trends observed in experimental low-density jets
can be explained.
ACKNOWLEDGMENTS

The authors would like to acknowledge the support received through Grant No. CTS-0317429 from the National
Science Foundation.
1

Phys. Fluids 22, 024103 2010

Viscous linear stability of axisymmetric low-density jets

K. Sreenivasan, S. Raghu, and D. Kyle, Absolute instability in variable


density round jets, Exp. Fluids 7, 309 1989.
2
P. A. Monkewitz, D. W. Bechert, B. Barsikow, and B. Lehmann, Selfexcited oscillations and mixing in a heated round jet, J. Fluid Mech. 213,
611 1990.
3
M. Provansal, C. Mathis, and L. Boyer, Benardvon Karman instability:
transient and forced regimes, J. Fluid Mech. 182, 1 1987.
4
P. J. Strykowski and K. R. Sreenivasan, On the formation and suppression of vortex shedding at low Reynolds numbers, J. Fluid Mech. 218,
71 1990.
5
P. J. Strykowski and D. L. Niccum, The stability of countercurrent mixing layers in circular jets, J. Fluid Mech. 227, 309 1991.
6
B. Pier and P. Huerre, Nonlinear self-sustained structures and fronts in
spatially developing wake flows, J. Fluid Mech. 435, 145 2001.
7
B. Pier, P. Huerre, and J.-M. Chomaz, Bifurcations to fully nonlinear
synchronized structures in slowly varying media, Physica D 148, 49
2001.
8
L. Lesshafft, P. Huerre, P. Sagaut, and M. Tarracol, Nonlinear global
modes in hot jets, J. Fluid Mech. 554, 393 2006.
9
F. Gallaire, M. Ruith, E. Meiburg, J.-M. Chomaz, and P. Huerre, Spiral
vortex breakdown as a global mode, J. Fluid Mech. 549, 71 2006.
10
L. Raynal, J.-L. Harion, M. Favre-Marinet, and G. Binder, The oscillatory instability of plane variable-density jets, Phys. Fluids 8, 993 1996.
11
R. P. Satti and A. K. Agrawal, Flow structure in the near-field of buoyant
low-density gas jets, Int. J. Heat Fluid Flow 27, 336 2006.
12
S. Boujemaa, M. Amielh, and M. P. Chauve, Velocity and concentration
distributions in globally unstable axisymmetric helium jets, C. R. Mec.
335, 449 2007.
13
L. Lesshafft, Nonlinear global modes and sound generation in hot jets,
Ph.D. thesis, Ecole Polytechnique Palaiseau, 2006.

14

P. W. Bearman, The effect of base bleed on the flow behind a twodimensional model with a blunt trailing edge, Aeronaut. Q. 18, 207
1967.
15
K. Hannemann and H. Oertel, Numerical simulation of the absolutely and
convectively unstable wake, J. Fluid Mech. 199, 55 1989.
16
D. A. Hammond and L. G. Redekopp, Global dynamics of symmetric and
asymmetric wakes, J. Fluid Mech. 331, 231 1997.
17
S. Jendoubi and P. J. Strykowski, Absolute and convective instability of
axisymmetric jets with external flow, Phys. Fluids 6, 3000 1994.
18
L. Lesshafft and P. Huerre, Linear impulse response in hot round jets,
Phys. Fluids 19, 024102 2007.
19
H. Schlicting, Boundary Layer Theory, 7th ed. McGraw-Hill, New York,
1979.
20
M. Gaster, Growth of disturbances in both space and time, Phys. Fluids
11, 723 1968.
21
A. Bers, Space-time evolution of plasma instabilitiesabsolute and convective, in Handbook of Plasma Physics I, edited by M. N. Rosenbluth
and R. Z. Sagdeev North-Holland, Amsterdam, 1983.
22
R. J. Deissler, The convective nature of instability in plane Poiseuille
flow, Phys. Fluids 30, 2303 1987.
23
J. P. Boyd, Orthogonal rational functions on a semi-infinite interval, J.
Comput. Phys. 70, 63 1987.
24
D. Kyle and K. Sreenivasan, The instability and breakdown of a round
variable-density jet, J. Fluid Mech. 249, 619 1993.
25
M. P. Hallberg, V. Srinivasan, P. Gorse, and P. J. Strykowski, Suppression
of global modes in low-density axisymmetric jets using coflow, Phys.
Fluids 19, 014102 2007.
26
P. A. Monkewitz and K. Sohn, Absolute instability in hot jets, AIAA J.
26, 911 1988.
27
M. H. Yu and P. A. Monkewitz, Oscillations in the near field of a heated
two-dimensional jet, J. Fluid Mech. 255, 323 1993.
28
M. P. Hallberg and P. J. Strykowski, On the universality of global modes
in low-density axisymmetric jets, J. Fluid Mech. 569, 493 2006.
29
B. S. Yildirim and A. K. Agrawal, Full-field concentration measurements
of self-excited oscillations in momentum-dominated helium jets, Exp.
Fluids 38, 161 2005.
30
R. P. Satti and A. K. Agrawal, Computational analysis of gravitational
effects in low-density gas jets, AIAA J. 44, 1505 2006.
31
J. W. Nichols, P. J. Schmid, and J. J. Riley, Self-excited oscillations in
variable-density round jets, J. Fluid Mech. 582, 341 2007.

Downloaded 30 Sep 2010 to 202.3.77.11. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Вам также может понравиться