You are on page 1of 11

Absolute and convective instability of axisymmetric jets with external flow

S. Jendoubi and P. J. Strykowski

Citation: Phys. Fluids 6, 3000 (1994); doi: 10.1063/1.868126


View online: http://dx.doi.org/10.1063/1.868126
View Table of Contents: http://pof.aip.org/resource/1/PHFLE6/v6/i9
Published by the AIP Publishing LLC.

Additional information on Phys. Fluids


Journal Homepage: http://pof.aip.org/
Journal Information: http://pof.aip.org/about/about_the_journal
Top downloads: http://pof.aip.org/features/most_downloaded
Information for Authors: http://pof.aip.org/authors

Downloaded 15 Aug 2013 to 150.216.68.200. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
Absolute and convective instability of axisymmetric jets with external flow
s. Jendoubi and P. J. Strykowski
Department of Mechanical Engineering, University of Minnesota, Minneapolis, Minnesota 55455
(Received 27 September 1993; accepted 9 May 1994)
The stability of an axisymmetric jet was examined in the presence of external co-flow and
counterflow. Spatio-temporal theory was used to distinguish regions of absolute and convective
instability in a parameter space including the velocity ratio, density ratio, Mach number, and the
shear layer thickness. The absolute-convective transition was identified for two distinct
axisymmetric modes. One of these modes became absolutely unstable in the presence of ambient
co-flow while the other mode required external counterflow to admit an absolutely unstable solution.
In general, the former mode was most unstable in low density jets, while the latter became more
unstable as the jet density was increased relative to the surrounding fluid. In the range of parameters
studied, both modes became increasingly unstable with decreasing jet density and for lower Mach
numbers. The results of the spatiotemporal theory are also compared to globally unstable modes
identified in laboratory jets.

I. INTRODUCTION direction opposite to the high-speed stream and had a mag-


nitude at least 13.5% of the high-speed stream.
Linear stability theory is a relatively mature science Pavithran and Redekopp6 extended the results of Huerre
which addresses the amplification of small disturbances su- and Monkewitz5 to include the effects of density ratio and
perimposed on a known velocity field. Historically the theory Mach number on the location of the absolute-convective sta-
examined whether monochromatic waves would grow tem- bility boundary in plane mixing layers. In qualitative agree-
porally for a presumed real wave number, or whether spatial ment with the general findings of Michalke 9 obtained using
amplification of waves having real frequency was more ap- spatial theory, Pavithran and Redekopp demonstrated that the
propriate. In the last quarter century the simultaneous ampli- mixing layer became more unstable when the density of the
fication of waves in space and time-spatiotemporal high-speed fluid was reduced relative to the density of the
theory-was considered in a variety of free-shear and wall- low-speed fluid; they also observed that increasing the Mach
bounded flows disturbed by an impulse. Gaster! wrote a note number of the high speed stream had a stabilizing effect. An
on the relationship between the frequency and amplification important conclusion of their work was that plane mixing
rates of temporally unstable flows and those of spatially un- layers admit absolutely unstable solutions only in the pres-
stable flows, identifying a unique transformation between ence of counterflow. This was true over the entire range of
them for the case of small amplification rates. Betchov and density profiles and Mach numbers examined in their study.
Criminale2 studied the spatiotemporal stability of the plane The spatio-temporal stability of the circular jet--or axi-
inviscid jet and found singularities in the dispersion relation symmetric mixing layer-has not received the attention
for the combined mode, which Gaster3 later explained, and given to the plane mixing layer despite the practical motiva-
further demonstrated that the most highly amplified waves in tions for the circular configuration. The most complete study
time arise when the waves are traveling with real group ve- was undertaken by Monkewitz and Sohn lO (also see Sohnl!)
locity and have real wave numbers. He also pointed out that where the absolute-convective stability boundary was exam-
the most rapid spatial growth occurs when the waves are ined as a function of the jet Mach number, the density ratio
traveling with real group velocity and have real frequency. between the jet and surrounding fluids, and the thickness of
More recently, spatiotemporal theory has been used to the shear layer relative to the jet diameter. The stability of the
identify the complex-plane trajectories of modes having zero axisymmetric mixing layer is fundamentalIy different from
group velocity,4-6 since these singUlarities in the dispersion that of the plane mixing layer owing to the boundary condi-
relation may lead to localized time-amplifying modes and tions which must be satisfied on the jet centerline. Monke-
self-excitation. 7,8 FolIowing the general formulation of witz and Sohn lO discovered, for instance, that eigenmodes
Bers, 7 regions of absolutely unstable flow can be identified became absolutely unstable in low-density round jets issuing
when the mode having zero group velocity exists in the up- into a quiescent fluid, a situation which admits a convec-
per half of the complex frequency plane. Correspondingly, tively unstable solution under analogous conditions in a
when this mode passes into the lower half of the complex plane mixing layer. 6
plane the flow is convectively unstable. Huerre and The principal goal of the present study was to identify
Monkewitz5 used a family of hyperbolic-tangent functions to regions of absolute instability for axisymmetric jets with ex-
study the effect of coftow and counterflow on the absolutely ternal coftow and counterflow. This study has been motivated
and convectively unstable nature of plane mixing layers at by experimental work conducted during the last decade indi-
constant density. They found that absolutely unstable flow cating the connection between the local concepts of absolute
could be supported only in the presence of ambient counter- and convective instability and the global stability of spatially
flow, specifically when the low-speed stream traveled in the developing laboratory flows. Pioneering work in this area

3000 Phys. Fluids 6 (9), September 1994 1070-6631/94/6(9)/3000/10/$6.00 © 1994 American Institute of Physics

Downloaded 15 Aug 2013 to 150.216.68.200. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
cous and the ambient fluid is assumed to extend to infinity in
the radial direction. Only axisymmetric modes were consid-
ered in this study.
Let u, v, and w denote the velocities in the x, r, and </>
Axisymmetric direction, respectively, and let p, p, and T denote the fluid
Nozzle
density, pressure, and temperature, respectively. The fluid
motion can be divided into two parts: a mean flow motion
which is assumed to be laminar and a disturbed fluid motion
superimposed on the mean flow. Let q be any quantity (u, v,
w, p, T, or p) to be calculated using Reynolds decomposi-
tion, then

q=Q+q', (1)

HG. 1. Velocity profile of an axisymmetric jet exhausting into an external where the quantities Q and q' refer to the mean and dis-
stream. turbed quantities, respectively. The mean flow is assumed to
be locally parallel to the jet axial direction (x direction) and
described by
was carried out by Mathis et al. 12 and Provansal et al. 13 in
the context of the classic "vortex shedding" phenomenon in U= U(r), V=O, W=O, P=p(r),
the neighborhood of the critical Reynolds number. They suc-
cessfully demonstrated that the onset of vortex shedding was T=T(r), and ji=ji(r).
due to a global instability which could be described by a
Hopf bifurcation. Koch,4 Huerre and MOnkewitz,5 Monke- Note that all fluid quantities presented above are nondimen-
witz and Nguyen,14 Chomas et ai.,15 and others, later pro- sional quantities. All lengths are made nondimensional by
vided the theoretical framework describing the connection the scale rlt2 which is the radius at which the velocity U* is
between the local stability concepts and global flow response equal to the mean velocity U! = (U: + U!)12. All the ve-
observed in the laboratory. locities are made nondimensional by the mean velocity
The analysis of Monkewitz and Sohn lO indicating the U! . The temperature and density are made nondimensional
presence of absolute instability in low-density round jets, by the ambient fluid temperature T'! and the ambient fluid
together with the accompanying experimental observations density ji!, respectively. However, the pressure is made
of jet global self-excitation by Sreenivasan et al., 16 Monke- nondimensional by ji!U!2. The superscript (*) refers to di-
witz et al.,17 Kyle and Sreenvisan,18 provide further evidence mensional quantities.
of the importance of the linear stability concepts applied to The velocity profile considered in this paper has the fol-
spatially developing flows. These experimental studies also lowing hyperbolic tangent shape:
exposed the possibility of using global self-excitation as a
means of efficient flow control. Recent studies in our
(2)
laboratoryl9,20 have explored the connection between spatio-
temporal theory and global flow control by using ambient
counterflow to achieve mixing enhancement in axisymmetric where b is a measure of the momentum thickness and R is
jets. Our experimental work has been guided by the analyses the velocity ratio defined by
of Huerre and Monkewitr> and Pavithran and Redekopp6 in
the context of countercurrent plane mixing layers. The aim of
the present investigation was to extend this theoretical (3)
framework by considering the effect of counterflow on the
absolute-convective transition in axisymmetric jets.
The ambient velocity U is allowed to vary between - Uc
00

and + Uc' thus the velocity ratio will only take positive val-
II. MATHEMATICAL MODEL ues. For ambient counterflowing fluids, R is greater than
unity (R > 1.0) and for co-flowing fluids, R is less than unity
A. Equations of motion
(R<1.0). However, for no external flow (Uoo=O) R is equal
The geometric configuration of the jet to be investigated to unity. Note also that the centerline velocity and the ambi-
is shown in Fig. 1. The axial, radial, and angular directions ent fluid velocity can be expressed in terms of the velocity
of the flow are given by x, r, and </>, respectively. Fluid of ratio R as Uc = 1 + Rand U 1- R. Figure 1 shows a typi-
00=

centerline temperature T c and centerline velocity U c is sur- cal velocity profile for the case of 33% counterflow corre-
rounded by ambient fluid having a temperature and velocity sponding to R=2 and b=12.S.
given by Too and U 00, respectively, where the ambient flow The temperature distribution is taken to satisfy the
direction will be taken as either coflowing or counterflowing Busemann-Crocco relation21 which yields the following ex-
to the jet fluid. Both streams will be considered as nonvis- pression for the density distribution:

Phys. Fluids, Vol. 6, No.9, September 1994 S. Jendoubi and P. J. Strykowski 3001

Downloaded 15 Aug 2013 to 150.216.68.200. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
S
per) (4)
1+[(1-S)/2R] (U-I-R)+[(y-l)/2] [(M)/(1+R)]2(R2-1+2U-U 2 ) '

where y is the specific heat ratio and M is the centerline jet By introducing Eq. (7), the equations of motions of
Mach number defined by M = Uc/ ~ yR g T Cl where R g is the small disturbances can be combined together to yield a sec-
gas constant and S = pJPoo is the centerline to ambient den- ond order ordinary differential equation in p only, thus
sity ratio. For heated subsonic jets S < 1.0 and for cold sub-
sonic jets S>1.0. It should also be understood from Eq. (4)
(8)
that U is equal to U(r) given by Eq. (2).
It is necessary to obtain a relationship between the mo-
mentum thickness and the parameter b. Here the momentum where
thickness () is defined by
(9)
- () __ J'm -p U(r)-U
- _ _
oo (
1- U(r)-
_ _U"') dr, (5)
rl/2 0 S Uc-Uoo Uc-U""
In situations where
and for constant density, and using Eq. (2), the momentum
thickness can be explicitly written as a function of the pa- 2 dU 1
---~-
rameter b as U-c dr r'
D
(j = 8b, (6) which is the case for a very small r, or for a very large r
when dU/dr tends to vanish, Eq. (8) has the following as-
where D = 2r 112 is a characteristic diameter of the jet. Be- ymptotic solution:
cause of its simplicity, Eq. (6) will be used for the definition
(10)
of D / () in all the results presented in this paper.
where eland c 2 are arbitrary constants to be determined
B. Linear stability theory from the boundary conditions, and 1m and Km are the modi-
fied Bessel functions of order m.
Following the analysis of Michalke,9 Reynolds decom- It is possible to combine the equations of motions to
position is used together with the mean flow assumptions to obtain a first order differential equation in only one variable
obtain the equations of motion of small disturbances, i.e., [rather than a second order differential equation (8)] by de-
neglecting nonlinear terms. The derivation of these equations fining a new variable X, which was first introduced by
assumes that the flow is laminar and a single component gas, Michalke,9 as follows:
that the fluid is nonviscous and obeys the ideal gas law, that
there are no body forces, no buoyancy effects, and that the . p
heat transfer due conduction is negligible. x= -la A (11)
V
Although only the stability of the axisymmetric mode
will be presented in this paper, the general form of three- thus
dimensional disturbances will be used to derive the linear
stability equations. It is assumed that the disturbed quantities dx ~-
~ = - a-p(U-c)
admit the following exponential form: dr
q' = q' (r, ify,x ,f) = q(r )ei(ax+m</>-wt), (7)
+x [ ---
1 (0 A +(m/r)2 x--
2
dU) 1]
+-. (12)
where a= a r + i a i and w= w r + i W i are the complex wave
U-c a 2p dr r
number and the complex frequency, respectively, and m is
the azimuthal mode number (m =0 represents the axisym- Since the pressure p and the velocity component ;; are
metric mode which will be considered in this study). The real bounded everywhere in the domain of calculation, the
wave number a r is related to the wavelength A by a r =2 7T/A, 0

boundary conditions for Eq. (12) are


and the angular frequency Wr is equal to 7T times the Strouhal
number based on the mean velocity and the diameter of the Sa 2 (1+R-c) Im(Acr)
jet. The quantity ai is a measure of spatial amplification, x---+ xc= - as r-+O, (13a)
where Wi is a measure of disturbance growth in time. The
complex phase velocity c is defined to be the ratio of W and a 2 (l-R-c) KmCAoor)
a, namely, c = w/ cx= c r + ic i ' where c r is the actual phase x-+x", = - (13b)
speed and c i is a measure of disturbance growth in time. The
quantity q(r) is introduced in Eq. (7) to distinguish it from where again the subscript c refers to values at the centerline
the total quantity q=q(r,o/,x,t) given by Eq. (1). of the jet and the subscript 00 refers to values of the ambient

3002 Phys. Fluids, Vol. 6, No.9, September 1994 S. Jendoubi and P. J. Strykowski

Downloaded 15 Aug 2013 to 150.216.68.200. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
fluid at large radial distances. The prime in the above equa- velocity is defined as vg=dw/da. We will denote by d and
tion refers to a derivative with respect to the argument. The wr the wave number and frequency, respectively, where the
quantities Ac and Ax> are given by group velocity is real and given by

'M)
Ac=a (1-L+R
2
(1+R-c)2
) 1/2
, (14a)
dw(a r ) x
da t

Aoo= a( 1-~ (1 :Rf(1-R-C)2 r /2


• (14b)
and similarly indicate by aO and WO the wave number and
frequency when the group velocity is zero, so that we can
write dw(aO)/da=O.
The amplification rate of the wave motion is then given
For given velocity and density distributions U(r) and per)
and for a given wave number a, Eq. (12) along with the by
boundary conditions given by Eqs. (13a) and (13b) represent
an eigenvalue problem. Each eigenvalue c (or w) is associ- (15)
ated with an eigenfunction X. This eigenValue problem will
be solved using the shooting method. In general, the wave packet is confined between two rays of
For a prescribed profile shape (defined by Die, S, R, M) x/t=constant where the amplification rate 0-=0; the wave
and a given wave number a and a guessed value of c (or w), motion is amplified within the wedge and damped outside
the shooting method typically starts the integration of Eq. the wedge. If the wave of zero group velocity is bounded by
(2) at one boundary using a variable step fourth-order the wedge the flow is said to be absolutely unstable, other-
Runge-Kutta method and marches towards the other bound- wise it is convectively unstable. For absolutely unstable
ary, where a comparison is made with the expected boundary flows, waves from the initial impulse disturbance will even-
condition. If the computed value of X by integration and the tually grow in time for all spatial positions x. However, for
boundary value of X calculated by Eg. (13) match, the com- convectively unstable flows the disturbance can grow and
putation stops. If they do not match a new c (or w) value will propagate away from its origin, so that an observer posi-
be guessed and the calculation is repeated until agreement is tioned at a fixed location in the frame of reference of the
achieved within the desired precision. However, this proce- initial impulse wiII see either a continuous decaying distur-
dure is not the most efficient. Mattingly22 suggested an inte- bance in time, or an initially amplifying one followed by
gration procedure beginning at both boundaries and compar- subsequent decay. If the flow is completely stable, the distur-
ing solutions at an intermediate location in the domain; for bance will decay everywhere with time.
example, Monkewitz and SohnlO selected the position of Bers7 showed that absolute instability occurs when
r=1.0 as a point of comparison. If the computed value of X O<W;:S;;Wi(am ) otherwise the flow is convectively unstable,
at r=1.0 obtained by integrating Eq. (12) from r=O matches where am is the wave number of the most amplified wave in
the computed value of X at r = 1.0 obtained by integrating time. In addition, the saddle point must satisfy the pinching
Eg. (12) from large r [where dUldr is small enough to be requirement, namely, the saddle point must be formed by the
able to use the asymptotic solution; Eg. (10)] the computa- coalescence of an upstream au and a downstream al
tion stops. However, if they do not match an improved branch. lO Thus the problem of locating the transition from
guessed value of c (or w) will be taken using the Newton- absolute to convective instability requires the location of
Ralphson method and the calculation is repeated until both saddle point (aO,wO) and checking if the pinching require-
values will match within the desired tolerance (four signifi- ment is fulfilled.
cant digits were taken in this study).

III. RESULTS
C. Criteria for absolute and convective instability
A representative map of lines of constant Wi in the com-
Many authors (among them Bers,? Briggs,23 Akhierzer plex a plane is shown in Fig. 2, for a jet at M=O with a
and Polovin,24 and Huerre 25) developed criteria for establish- centerline to ambient density ratio of S=2.S06 and a velocity
ing whether wave growth in an unstable system would ratio of R =2. We note that for real a (ai=O), the complex
evolve locally or whether the wave would be convected frequency has positive imaginary part Wi for 0:S;;a:S;;24.16,
away from its origin during amplification. These concepts thus the flow is temporally unstable. Pure spatial instability
lead to the definition of absolutely and convectively unstable exists over a range of frequencies w since the wave number
systems as will be discussed below following the criteria of a is complex with a negative imaginary part ai for the
Bers. 7 branches of wi=O. Thus the wave motion will grow in space
Here we are interested in the evolution of an impulse- and time with an amplification rate given by Eq. (15). The
type disturbance on the stability of axisymmetric jets, as the maximum growth rate of 0-= 8.33 occurs at a~ = am = 11.12
effect of any other disturbance can be obtained through a and af =0 where w r( am) = 1S.70 and Wi( an.} =8.33, which
convolution with the solution of the impulse disturbance. 26 is in agreement with the results of Gaste~ since aw;lacxr=O.
Two kinds of waves are of particular interest here: the one In fact, by starting at the origin and moving along the real
with real group velocity having the maximum amplification axis to the right, Wi will increase from zero to its maximum
rate and the one with zero group velocity, where the group value at am and will decrease after that as a r is made larger.

Phys. Fluids, Vol. 6, No.9, September 1994 S. Jendoubi and p, J. Strykowski 3003

Downloaded 15 Aug 2013 to 150.216.68.200. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
o
...........
-- I OJ i =833

~------~~-----~~
..... ...- - -

0.8
..;, ,
\
:,.., \,
CD

-4 t(r)i :. , \
a..I
-8
ICl
0.6
\':.; "Q)
\
\
\@,,
"
-12
:. /l,,' I " •

004 : ' :,~ ,'"


~:': "', i -
:.; '\",/:

-20
: i'
.: .. "

0.2 ~.:
.....
-24
08

09
.....
"i~8.33
,. ......
~

-28
0 5
" . I'
10
........t....
15 20 25
a.r r

FIG. 3. Radial pressure perturbation eigenfunctions for R =2, S =2.806,


FIG. 2. Complex a plane for R=2, S=2.806, D/O=100, and M=O. D/O=100, and M=O.

Figure 2 also shows the location of nine saddle points the jet centerline, analogous to the eigenfunction distribu-
numbered from 1 to 9. The corresponding values of the tions present in plane mixing layers where the disturbance
saddle point coordinates are shown in Table I. Four of these amplitude must vanish at large cross-stream distances. Dis-
saddle points have w;>O (points 6 to 9), however, they do tinguished from Mode I, the eigenfunction distribution cor-
not meet the pinching requirement since they are formed by responding to saddle point 2 displays a maximum pressure
the coalescence of two downstream al modes. Also point 8 on the jet centerline and decays monotonically in the radial
and point 9 have frequencies with an imaginary part greater direction. Hence, the disturbance can communicate across
than Wi( am). Consequently, points 6 through 9 cannot lead the jet column and cannot be supported as a solution in the
to absolute instability. Points 2 through 5 meet the pinching stability of the plane mixing layer. The eigenfunction at
requirement but have negative imaginary parts indicating the saddle point 2 will be identified in this paper as Mode II.
presence of convective instability. Point 1, however, meets Modes 3 and higher will not be considered further in this
the pinching requirement and has w;=O indicating a bound- paper.
ary value between regions of convective and absolute insta- The disturbance velocities associated with Modes I and
bility. II are presented in Fig. 4, and reflect the general observations
The physical significance of saddle points 1 through 5 seen in the pressure distributions of Fig. 3. Velocity fluctua-
can be gained by examining the magnitude of the pressure tions for Mode I disturbances vanish away from the shear
disturbance per) as a function of the radial distance r; these layer and display a characteristic bimodal feature similar to
eigenfunction distributions are presented in Fig. 3, where the that observed in spatially developing shear layers?7,28 Notice
pressure magnitude is normalized by the maximum value of again that the disturbance energy for Mode I is essentially
the pressure. Simple eigenfunction distributions exist only zero on the jet centerline. Mode II fluctuations also peak in
for saddle points 1 and 2, in contrast to the higher order the jet shear layer, but maintain a significant amplitude along
modes 3 through 5 which indicate the presence of oscilla-
tions in the radial direction. Saddle point 1 displays peak
pressure near the interface of the counterflowing streams
(here t/,,= -O.33U c as shown in the profile of Fig. 1) and
decays away from the shear layer. For this mode-referred to
0.8
henceforth as Mode I-the disturbance pressure is not felt on
IMrlI
fimax.
0.6
TABLE I. Saddle point coordinates (R =2, S=2.806, and D/O=100).

Point (tlr 0); ar ai


004

17.500 0.000 8.517 -10.737


2 06.597 -1.739 0.808 -2.197 0.2 -
3 12.620 -4.035 1.267 -5.546
-3.lB6
- -Q)-
4 17.509 1.542 '8.965
_ _ ...1
5 20.764 -10408 1.656 -12.419 0
6 22.056 2.547 1.625 -15.799 0 0.5 1.5 2 2.5
7 21.458 7.215 1.550 ~19.082 r
i! 19.358 11.960 1.467 -22.290
9 16.275 160447 1.384 -250451 FIG. 4. Radial velocity perturbation eigenfunctions for R=2, S=2.806, DIB
=100, and M=O.

3004 Phys. Fluids, Vol. 6, No.9, September 1994 S. Jendoubi and P. J. Strykowski

Downloaded 15 Aug 2013 to 150.216.68.200. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
5 unstable for parameter combinations above the curves and
Ell Monkewitz et al.
convectively unstable below. Principal grid lines have been
• Strykowskl and Niccwn
included in the figure at R = 1 and S = 1 to identify important
4
half-planes. For conditions where R> 1 ambient counterflow
I
R I is required, and when R<1 the streams are co-flowing. The
3 line of S=1 distinguishes between heavy jets (S>I) and
light jets (S < 1) compared to the surrounding fluid.
Concentrating first on Mode I disturbances it is clear that
2 counterflow is required at all density ratios to achieve abso~
lute instability. This results was first observed in plane mix-
_.... ing layer stability by Pavithran and Redekopp.6 Their
.~.\.-.-.-.-.-. absolute-convective boundary computed for M =0 is pro-
Monkewitz &: Sohn vided in Fig. 5 and agrees very favorably with our calcula-
0 tions indicating the similarity in the eigenfunctions between
0.1 1 10
S plane mixing layers and the Mode I disturbance of an axi-
symmetric mixing layer. The slight disparity in the curves at
large S can be attributed to the finite transverse curvature in
FIG. 5. Absolute-convective instability boundaries for Mode I and II distur-
the present problem where D I e= 120. Calculations to be pre-
bances for D/8=120 and M=O. Theoretical results of Pavithran and Rede-
kopp (Ref. 6) and Monkewitz and Sohn (Ref. 10) are provided for compari- sented below indicate that our boundary approaches that of
son. Locations of the onset of global self-excitation are provided from the Pavithran and Redekopp as Die is increased further. For both
studies of Monkewitz et ai. (Ref. 17) and Strykowski and Niccum (Refs. 19 plane mixing layers and Mode I disturbances in the axisym-
and 29).
metric mixing layer the flow becomes more unstable as the
jet density is reduced relative to the surrounding fluid; Le., as
the jet density is reduced, the magnitude of the counterflow-
the jet centerline indicating that this mode is unique to the
axisymmetry of the problem. Furthermore, we might expect ing velocity necessary to reach absolute instability is also
that a Mode II instability will lead to the formation of struc- reduced. For instance, only 1.33% counterflow
(U oo = -O.0133U c ) is required to achieve absolute instabil-
tures having scales of the order of the jet diameter. It is also
likely that changes in the mean velocity field in the shear ity at S =0.1, but more than 35% is necessary at S =3.0.
layer as well as on the jet centerline will alter the stability of The Mode II stability boundary in Fig. 5 reflects the
the Mode II disturbance. Finally, if we refer back to Fig. 2 general trend of Mode I, namely, that the flow becomes in-
we observe that Mode I is on the boundary between absolute creasingly stable as the jet density is elevated relative to the
and convective instability where Mode II is only convec- surrounding fluid. However, as first demonstrated by Monke-
tively unstable. Hence, for the conditions examined here witz and Sohn,lO the low density jet supports an absolutely
(where R=2, S=2.806, Dle=100, and M=O) we define unstable solution in the absence of external flow, Le., for
Mode I to be more unstable than Mode II since the former R = 1. This behavior is observed in Fig. 5 for Mode II distur-
mode is more likely to become absolutely unstable with bances at density ratios S<0.692, indicating that Mode II is
slight variations in the parameter space. We will demonstrate the eigenfunction identified by Monkewitz and Sohn. (The
with the results presented below that Mode II can also be- slight deviations in the Mode II stability boundary and that
come absolutely unstable (Le., making w;
positive for saddle identified by Monkewitz and Sohn are due to differences in
Die, as addressed below, as well as the precise shape of the
point 2) by variations of R, S, Die, or M.
In the sections to follow we will examine the stability of velocity profile used in the analysis.) As the jet density ratio
Mode I and Mode II disturbances in a parameter space in- is reduced below S =0.692 the absolute-convective transition
volving velocity ratio R, density ratio S, shear layer thick- occurs in a region of ambient co-flow. Consequently, at
ness Die, and Mach number M. We view R as the principal S=O.1 the transition would occur in the presence of a co-
dependent parameter in the problem, and hence will evaluate f!owing stream having a magnitude of approximately 27% of
the velocity ratio at which Modes I and II undergo the tran- Uc ·
sition from convective to absolute instability as a function of One significant feature of the stability boundaries in Fig.
S, Die, and M. 5 occurs near a density ratio of unity. Here, as the density
ratio is increased, the most unstable mode switches from
Mode II to Mode I. This crossover point occurs essentially at
A. Effect of the density ratio
S = 1 for the conditions in Fig. 5, but in general is a function
We begin our examination of the effect of jet density of both Die and M as will be shown in the results to follow.
ratio on the flow stability by restricting our study to incom- The intersection of mode boundaries, however, does explain
pressible situations (M =0) for a representative velocity pro- an observation made by Monkewitz and Sohn.lO By extrapo-
file having D/e=120. The principal density effects can be lating their results obtained for a round jet at S <0.72 to a
elucidated by tracking the absolute-convective stability density ratio of S = 1, they predicted that the critical velocity
boundaries of Mode I and II disturbances in the S-R plane ratio was in agreement with the plane mixing layer calcula-
as shown in Fig. 5. The solid and dashed curves correspond tions of Huerre and Monkewitz,5 implicitly assuming that the
to Modes I and II, respectively, where the flow is absolutely modes were identical. It can be seen from Fig. 5 that this

Phys. FluIds, Vol. 6, No.9, September 1994 S. Jendoubi and P. J. Strykowski 3005

Downloaded 15 Aug 2013 to 150.216.68.200. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
agreement was entirely coincidental, owing to the choice of
M =0 and the relatively large value of D / e used by Monke-
witz and Sohn.
We also compared the absolute-convective boundaries in
the S - R plane with the onset of global instability observed
in laboratory jets. Monkewitz et al. 17 examined the nonlinear
saturation amplitude in low-density round jets (S<l) issuing
into a quiescent environment (R = 1), and predicted the onset
of global self-excitation at two distinct modes having density
ratios of 0.63 and 0.73 in remarkable agreement with the
theoretical value of the Mode II boundary at S =0.692 shown
in Fig. 5. These authors point out that this agreement is to
some extent fortuitous given the expectation that global self-
excitation will occur only in the presence of a finite domain
of absolutely unstable flow, which would be unlikely in the 10
immediate vicinity of the absolute-convective boundary.tS
Experiments performed by Strykowski and Niccum 29 consid-
ered the global response of heavy jets (where S~l) by ex-
hausting mixtures of SF6 and air into a region downstream of
the jet exit where counterflow was used to establish velocity
ratios R ~ 1. The solid symbols and error bars indicate the
best estimates of the onset of global instability in heavy jets
and show reasonable agreement with the Mode I boundary of
the spatiotemporal theory. The apparent switching observed
in the experiments between Mode II and Mode I disturbances
for low-density and high-density jets, respectively, is consis-
tent with the expectation that the most unstable mode should
t!merge at any particular location on the S - R plane.
The branch point frequency and wave number for Mode
I and II disturbances are presented in Fig. 6 in the S-R 10
plane. The real frequency of the Mode I disturbance is higher
than the corresponding values at Mode II over the entire
range of density ratios examined. At the location in the S - R
plane where the two modes coincide, namely at S-l, the
frequency w;
takes values of 12.2 and 4.07 for Modes I and
II, respectively. These frequencies translate into Strouhal
numbers based on shear layer momentum thickness and cen-
terline velocity Sto of 0.014 (Mode I) and 0.0047 (Mode 11).
The global mode frequency of St o=0.015 measured hy
Strykowski and Niccum 19 at S=1 and R=1.32 in Fig. 5
agrees favorably with the Mode I disturbance frequency.
The global flow response in favor of the Mode I distur-
hance at the intersection of Mode I and II (at S = 1) is un-
doubtedly related to the streamwise distance over which par-
allel flow conditions could be maintained in the experiments.
Measurements hy Strykowski and Niccum indicate that 10
counterflow was established only in the first one or two jet
diameters; downstream of this location the velocity ratio ap-
proached unity. Calculations of the disturbance wavelength '1\ FIG. 6. Branch point frequency and wave number for absolute-convective
from Fig. 6, given by '1\=21T/a;, indicate that the Mode I and instability boundary at D/B=120 and M=O.
Mode II wavelengths at S=l are 0.42 and 3.8 jet diameters,
respectively. The likelihood that a Mode I or Mode II distur-
bance would emerge in a laboratory flow depends on will lead to rapid changes in mode selection. To address this
whether the parallel flow assumption is approximately satis- issue we consider in the next section how the convective-
fied over streamwise distances of the order of '1\. These con- absolute transition is affected by spreading of the shear layer
ditions were clearly not maintained for Mode II in the ex- as described by the parameter D / e.
periments of Strykowski and Niccum, leading to a natural Finally, we compared the frequency selection of the glo-
flow selection of Mode I. The selection process will also bal modes observed by Monkewitz et al. 17 in low-density
depend on whether the stream wise variations in profile shape jets at S=0.63 and 0.73 in Fig. 5 (R=lJ. The branch point

3006 Phys. Fluids, Vol. 6, No.9, September 1994 S. Jendoubi and P. J. Strykowski

Downloaded 15 Aug 2013 to 150.216.68.200. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
5 160
-~. Model - - Present calculations
~ -Modell • Monkewitz and Sohn

4
120
R
3
ole
80
Absolute
2 Instability

40

Convective
Instability
..L-.-~ __
0
o ~~-....L _ _~_ .. _~_ ......... ~ .....• _.... _.
0.1 0.3 0.5 0.7
0.1 1 10
S
S

FIG. 8. Absolute-convective instability boundary of Mode II disturbances at


FIG. 7. Absolute-convective instability boundaries for Mode I and II distur- R=l and M=O. Solid points indicate theoretical values from Monkewitz
bances as a function of DIB at M=O. and Sohn (Ref. 10).

frequency calculated at R = 1 and S =0.692 from Fig. 6 is


layer thickness is decreased relative to the jet half-width the
w;=2.92 or a Strouhal number based on jet diameter and
centerline velocity of StD =0.46. Monkewitz et al. reported Mode I stability boundary approaches the plane mixing layer
frequencies of StD =0.30 and 0.45 for the global modes iden- results of Pavithran and Redekopp6 shown previously in Fig.
tified at S =0.73 and 0.63, respectively. The agreement be- 5.
tween the calculated Mode II frequency and the self-excited In the range of transverse curvature 40",;;;DIO",;;;160 the
global mode observed at 0.63 suggests that the two may be Mode II boundary also displays a relative insensitivity to
analogous. This argument is supported by the experimental shear layer thickness for low density ratios. However, at den-
findings of Kyle and Sreenivasan. 18 They described the onset sity ratios above S=0.7 an increasing shear layer thickness
of an "oscillating mode" in low-density jets for S"';;;0.61 and has a destabilizing effect, a trend opposite to that observed in
having a frequency of StD ""'0.45. The likelihood of global Mode I at high density ratios. When the density ratio is re-
self-excitation is also greater at S=0.63 than at 0.73 given duced below S=0.7, the Mode II disturbance becomes
the necessary condition for global instability of a finite re- slightly more unstable as shear layer thickness decreases
gion of absolutely unstable flow. IS relative to the jet half-width; this trend will be explored more
completely in connection with Fig. 8 presented below. The
crossover point in the S -R plane where the most unstable
B. Effect of the shear layer thickness
eigenfunction switches between Modes I and II is clearly a
As a jet develops downstream of a nozzle the mixing function of D I 8 as seen in Fig. 7, although this dependence is
layer will spread transverse to the jet axis while the jet half- weak except when the shear layer becomes very thick. The
width will remain essentially constant in the region prior to density ratios at the crossover points for DI8=40, 80, 120,
the termination of the potential core. Consequently, the initial and 160 are S = 1.541, 1.123, 1.033, and 0.993, respectively.
streamwise evolution of the jet can be modeled quite simply For practical reasons the jet is often studied as it ex-
through variations in the transverse curvature parameter D I 8. hausts into quiescent surroundings corresponding to R = l.
Here we will examine how the absolute-convective stability As Monkewitz and Sohn lO demonstrated in a jet at M =0, an
boundary is influenced by D I 0 as a function of density ratio absolutely unstable mode is supported at sufficiently low
and velocity ratio. We again restrict our discussion to plane density ratios in the absence of an external stream, where the
wave disturbances of type Mode I and Mode II as described value of S depended on the shear layer thickness relative to
above. the jet diameter. We explored the R = 1 cut in the S - R plane
Regions of absolute and convective instability are iden- at zero Mach number for Mode II disturbances where we
tified in the S -R plane for Mode I and II disturbances in a expect regions of absolute instability at low values of S. The
jet at zero Mach number in Fig. 7. All stability boundaries absolute-convective stability boundary in the D I O-S plane
display a monotonically increasing behavior with density ra- is presented in Fig. 8. Consistent with the calculations pre-
tio as observed in Fig. 5 at D/8=120. Examining first the sented in Fig. 7, the critical density ratio is virtually indepen-
behavior of Mode I disturbances. it can be observed that the dent of DIO for values of 40",;;;DI8",;;;160. This trend is
transverse curvature Df(J has essentially no effect on jet sta- abruptly reversed for thick shear layers having values of DI8
bility in low density flow (i.e., for S<1) over the range of less than approximately 40; the highest density ratio at which
DIO considered here. The effect of DI8 becomes distinctly absolute instability exists occurred at S=O.71 for DI8=52.
more pronounced at density ratios greater than unity, with The stability boundary points computed by Monkewitz and
thin shear layers (e.g., DI8=160) displaying greater instabil- Sohn are also provided in Fig. 8 indicating the satisfactory
ity than thick ones (e.g., DI8=40). Furthermore, as the shear agreement between the independent calculations; the slight

Phys. Fluids, Vol. 6, No.9, September 1994 S. Jendoubi and P. J. Strykowski 3007

Downloaded 15 Aug 2013 to 150.216.68.200. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
5
--Model Monkewitz & Sohn
- -Modell
0.7F=~";;;:::/-'_
4

R 0.6

3 S
M 0.5
0-
0.4
0.6'-1
2
/
/
0.3
.--

0
0.1 1 10
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
S
M

FIG. 9. Absolute-convective instability boundaries for Mode I and II distur-


bances as a function of Mach number at D/O=160. FIG. 10. Absolute-convective instability boundaries of Mode II disturbances
at R = 1. Theoretical computations of Monkewitz and Sohn (Ref. 10) for an
axisymmetric vortex sheet are included.
discrepancies in locating the boundary are presumably re-
lated to subtle differences in the mean velocity and density
fields used in the computations. vectively unstable flow (above) are presented in Fig. 10. The
region of absolute instability is relatively insensitive to trans-
C. Effect of the Mach number verse curvature for D I e~40 extending up to Mach numbers
near 0.7 for the lowest density ratios examined. The results
Monkewitz and Sohn lO predicted that the region of ab- of Monkewitz and Sohn lO are also included in Fig. 10 for a
solute instability would vanish at moderately high subsonic vortex sheet (D I e-+oo) at R = 1. Their calculations are in rea-
Mach number even at very low density ratios for jets with sonable agreement with the present ones for the thinnest
thin shear layers and without external flow. We address this shear layer case of Dle=160. In the limit of Die tending to
issue in more detail here by considering the combined effects infinity we obtain a critical density ratio of 0.665 for our
of S, R, Die, and M. To gain insight into the independent profile at zero Mach number which is very close to the value
effect of Mach number, we first consider the Mode I and II of S=0.66 computed by Monkewitz and Sohn.
stability boundaries in the S-R plane for a fixed transverse In the presence of thick shear layers the region of abso-
curvature of D I e= 160. The absolute-convective boundaries lute instability is reduced appreciably, being confined to
were computed for Mach numbers of 0, 0.4, 0.6, and 0.8 and small subsonic Mach numbers in jets with very low density.
are presented in Fig. 9. Both Modes I and II become increas- Monkewitz and Sohn lO demonstrated that helical modes also
ingly stable as the Mach number is increased, a trend which become absolutely unstable in the range of Mach numbers
can be observed at all density ratios. Similar behavior was and density ratios considered here, however for the vortex
documented by Pavithran and Redekopp6 over the same sheet they considered the axisymmetric mode was found to
Mach number range in plane mixing layers. be the most unstable. However, a similar conclusion should
A striking feature of the stability boundary of Mode II not be blindly extended to jets with thick shear layers. Batch-
disturbances is the elimination of a region of absolute insta- elor and Gill 30 found, using temporal theory, that helical
bility for jets with ambient co-flow. The R = 1 cut in the S-R modes become more unstable than axisymmetric one for jets
plane first intersects the Mode II boundary at S=O.1 for with Gaussian-type profiles. The present calculation need to
M=0.718. In other words, at the lowest density ratio exam- be extended to include helical modes to address this issue
ined here of S=O.I, the flow admits an absolutely unstable which may be relevant, for instance, in plasma jets where
solution in the absence of external flow only for Mach num- exit profiles shapes have been shown to be nearly parabolic
bers less than M =0.72. Note that the computations at at a density ratio S-0.03 for moderate subsonic Mach
M =0.6 for Mode II disturbances were terminated at numbers?1
S=0.760 since the real wave number of this mode ap-
proached zero; similar nonphysical solutions were identified
IV. SUMMARY
by Monkewitz and Sohn lO in this range of parameters. No
physical solutions could be found for Mode II at M =0.8 in The spatiotemporal stability of axisymmetric jets with
the range of density ratios S~O.l. external flow was investigated. Axisymmetric instability
A closer examination of Mode II stability properties was Modes I and II were identified, both of which admitted ab-
made to highlight the combined effects of Mach number and solutely unstable solutions in certain regions of a parameter
Die for the important configuration when the jet exhausts in space involving velocity ratio, density ratio, transverse cur-
the absence of external flow (R = 1). Stability boundaries vature, and Mach number. Mode I disturbances displayed
separating absolutely unstable flow (below curve) from con- maximum pressure fluctuations in the jet shear layer,

3008 Phys. Fluids, Vol. 6, No.9, September 1994 S. Jendoubi and P. J. Strykowski

Downloaded 15 Aug 2013 to 150.216.68.200. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
whereas Mode II disturbances were found to have maximum 11 K. Sohn, "Absolute instability in hot jets," M.S. thesis. University of
pressure perturbation amplitudes on the jet centerline. In California at Los Angeles, 1986.
general, both modes became more unstable with decreasing 12C. Mathis, M. Provansal, and L Boyer, "The Benard-von Karman insta-
bility: An experimental study near the threshold," J. Phys. Lett. 45, 483
density ratio S, increasing velocity ratio R, and decreasing (1984).
Mach number M. The transverse curvature D / ewas found to 13M. Provansal, C. Mathis, and L. Boyer, "Benard-von Karman instability:
either increase or decrease the jet stability depending on the Transient and forced regimes," J. Fluid Mech. 182, 1 (1987).
mode type and the density ratio. Finally, the onset of global 14p. A. Monkewitz and L N. Nguyen, "Absolute instability in the near wake

self-excitation identified in laboratory jets agreed well with of two-dimensional bluff bodies," J. Fluids Struct. 1, 165 (1987).
Mode I disturbances in high density jets (S > 1) and with 15J. M. Chomaz, P. Huerre, and L G. Redekopp, "Bifurcations to local and
global modes in spatially-developing flows," Phys. Rev. Lett. 60, 25
Mode II disturbances in low density jets (S<l) indicating a (1988).
cross over between the most amplified mode at a density 16K. R. Sreenivasan, S. Raghu, and D. Kyle, "Absolute instability in vari-
ratio near unity as predicted by the theory. able density round jets," Exp. Fluids 7. 309 (1989).
17p. A. Monkewitz, D. W. Bechert, B. Barsikow, and B. Lehmann, "Self-

ACKNOWLEDGMENTS excited oscillations and mixing in a heated round jet," J. Fluid Mech. 213,
611 (1990).
The authors would like to thank P. A. Monkewitz and L. IRD. Kyle and K. R. Sreenivasan, "The instability and breakdown of a round
G. Redekopp for valuable discussions concerning this work. variable-density jet," J. Fluid Mech. 249, 619 (1993).
Financial support was provided by the National Science 19p. J. Strykowski and D. L Niccum, "The stability of countercurrent mix-
ing layers in circular jets," J. Fluid Mech. 227, 309 (1991).
Foundation under Grant No. CTS-9U6532. A computing 20p. J. Strykowski and R. K. Wilcoxon, "Mixing enhancement due to global
grant from the Minnesota Supercomputer Institute is also oscillations in jets with annular counterflow," AIM J. 31, 564 (1993).
acknowledged. 21 H. Schlichting, Boundary Layer Theory, 7th ed. (McGraw-Hill, New York,
1978).
1M. Gaster, "A Note on the relation between temporally increasing and 22G. E. Mattingly, "The stability of a two-dimensional incompressible
spatially-increasing disturbances in hydrodynamic stability," J. Fluid wake," Ph.D. thesis, University of Princeton, 1968.
Mech. 14,222 (1962). 23R. J. Briggs, "Electron-stream interaction with plasmas," Research Mono-
lR. Bctchov and W. O. Criminale, "Spatial instability of inviscid jet and graph 29 (MIT, Cambridge, MA, 1964).
wake," Phys. Fluids 9, 359 (1966). 24 A. 1. Akhierzer and R. V. Polovin, "Criteria for wave growth," Sov. Phys.
3M. Gaster, "Growth of disturbances in both space and time," Phys. Fluids Uspek. 14, 278 (1971).
11, 723 (1968). 25p. Huerre, "Spatio-temporal instabilities in closed and open flows," in
4W. Koch, "Local instability characteristics and frequency determination of Instability alld NOllequilibrium Structures, edited by E. Tirapegui and D.
self-excited wake flows," J. Sound Vib. 99, 53 (1985). Villarrod (Reidel, New York, 1987), pp. 141-177.
sp. Huerre and P. A. Monkewitz, "Absolute and convective instabilities in 2I>R. V. Churchill, Operational Mathematics, 3rd ed. (McGraw-Hili, New
free shear layers," J. Fluid Mech. 159, 151 (1985). York, 1972).
6S. Pavithran and L. G. Redekopp, "The absolute-convective transition in 27p. Freymuth, "On transition in a separated laminar boundary layer," J.
subsonic mixing layers," Phys. Fluids A 1, 1736 (1989).
Fluid Mech. 25, 683 (1966).
7 A. Bers, "Space-time evolution of plasma instabilities--absolute and con-
28 A. Michalke. "On the spatially growing disturbances in an inviscid shear
vective," in Halldbook of Plasma Physics I, edited by M. N. Rosenbluth
and R. Z. Sagdeev (North-Holland, Amsterdam, 1983). layer," J. Fluid Mech. 23,521 (1965).
20p. J. Strykowski and D. L Niccum, "The influence of velocity and density
"P. Huerre and P. A. Monkewitz, "Local and global instabilities in spatially
developing flows," Annu. Rev. Fluid Mech. 22, 473 (1990). ratio on the dynamics of spatially developing mixing layers," Phys. Fluids
9 A. Michalke, "Instabilitiit eines kompressiblen runden Freistrahls under A 4.770 (1992).
Berucksichtigung des Einflusses der StraWgrenzschichtdicke," Z. Flug- JUG. K. Batchelor and A. E. Gill, "Analysis of the stability of axisymmetric
wiss. 19 (8/9), 319 (1971) [Translation NASA TM 75190, Dec. 1977]. jets," J. Fluid Mech. 14, 529 (1962).
Hlp. A. Monkewitz and K. D. Sohn, "Absolute instability in hot jets," AlAA ,\1 R. A. Spores, "Analysis of the flow structure of a turbulent thermal plasma
J. 26. 911 (1988). jet," Ph.D. thesis, University of Minnesota, 1989.

Phys. Fluids, Vol. 6, No.9, September 1994 S. Jendoubi and P. J. Strykowski 3009

Downloaded 15 Aug 2013 to 150.216.68.200. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions

You might also like