You are on page 1of 101

ON THE EFFECTS OF A LOW-DENSITY WEAK

CO-FLOW ON THE NEAR FIELD OF AN


AXISYMMETRIC JET

by

Michael G. Bonarski

September 2010

A thesis submitted to the Faculty of the Graduate School of the University at


Buffalo, State University of New York in partial fulfillment of the requirements for
the degree of

Master of Science

Department of Mechanical and Aerospace Engineering

Major Professor
David J. Forliti, Ph.D.
ACKNOWLEDGEMENTS
The author wishes to express his gratitude to a number or individuals who were vital to

the success of this project and his graduate studies as a whole. First and foremost, he would like

to thank his advisor, Dr. David Forliti. He has been extremely helpful and supportive, and

provided the insight to pursue this interesting research opportunity. The author would also like

to thank Dr. Paul Desjardin and Dr. Matthew Ringuette for being kind enough to serve as

committee members. He is also very appreciative of the funding received throughout his

graduate studies from the University at Buffalo through the Presidential Fellowship.

Many thanks go out to all the past and present graduate students of the Combustion

Laboratory including, Kareem Ahmed, Zak Carr, Arezoo Hajesfandiari, Dan Jason, Ben Knox,

Rahul Mulinti, Sann Naing, and Joe Richter. Their guidance, and assistance was highly valued,

as was their friendship. Joel Gabrielson and Paul Dack are also acknowledged for volunteering

to help during different phases of this project.

The design, manufacturing, and construction of the coaxial nozzle facility were difficult

and the project may not have materialized as well as it did, had it not been for the expertise of

Ken Peebles. His advice in these areas was second to none and was greatly appreciated. The

author is also grateful for the assistance from Dr. Bahattin Koc of the Industrial Engineering

Department, in manufacturing the nozzles used in this project by means of rapid prototyping.

The author would also like to thank Toni Schumacher and Carole Dentico for all their

administrative help throughout his graduate career.

Lastly, the author wishes to express his most sincere gratitude to all his family, friends,

employers, and all other UB faculty and staff, for their support and motivation throughout his 6+

years at the University at Buffalo, and making it a truly memorable and enriching experience.

- ii -
TABLE OF CONTENTS
ACKNOWLEDGEMENTS ................................................................................................ ii
LIST OF FIGURES ........................................................................................................... vi
LIST OF TABLES ............................................................................................................. ix
NOMENCLATURE ........................................................................................................... x
ABSTRACT...................................................................................................................... xii
CHAPTER 1: INTRODUCTION ....................................................................................... 1
1.1 Motivation ................................................................................................................ 1
1.1.1 Mixing Enhancement ....................................................................................... 1
1.1.2 Turbulence and Mixing Suppression................................................................ 2
1.1.3 Current Study ................................................................................................... 3
CHAPTER 2: LITERATURE REVIEW ............................................................................ 5
2.1 Free Shear Layers ..................................................................................................... 5
2.1.1 Axisymmetric Jets ............................................................................................ 6
2.1.2 Instability.......................................................................................................... 8
2.1.3 Effects of Initial Conditions ........................................................................... 13
2.1.4 Effects of Density ........................................................................................... 16
2.1.5 Effects of Compressibility.............................................................................. 20
CHAPTER 3: FACILITY & INSTRUMENTATION...................................................... 22
3.1 Nozzle Design ........................................................................................................ 22
3.2 Test Facility............................................................................................................ 25
3.3 Test Conditions ...................................................................................................... 25
3.3.1 Initial Conditions ............................................................................................ 27
3.4 Instrumentation & Data Collection Techniques ..................................................... 28
3.4.1 Particle Image Velocimetry ............................................................................ 28
3.4.2 Hot-wire Anemometry ................................................................................... 31
3.4.3 Schlieren Flow Visualization ......................................................................... 32
3.4.4 Flow Metering ................................................................................................ 33
CHAPTER 4: RESULTS .................................................................................................. 35
4.1 Schlieren................................................................................................................. 35

- iii -
4.2 Particle Image Velocimetry.................................................................................... 37
4.2.1 Instantaneous Images ..................................................................................... 37
4.2.2 Velocity Field Statistics ................................................................................. 38
4.3 Hotwire Power Spectra .......................................................................................... 45
CHAPTER 5: CONCLUSIONS ....................................................................................... 48
5.1 Conclusions ............................................................................................................ 48
5.2 Future Work ........................................................................................................... 49
APPENDIX A: FIGURES .......................................................................................... 51
REFERENCES .................................................................................................................. xi

- iv -
-v-
LIST OF FIGURES
Figure 1: Example free shear layer boundaries and evolving velocity profile. ........................... 51
Figure 2: Velocity profiles at 5 downstream locations for an axisymmetric air jet ..................... 51
Figure 3: Vorticity profiles at 5 downstream locations for an axisymmetric air jet .................... 52
Figure 4: Axisymmetric jet structure ........................................................................................... 53
Figure 5: Evolution of vortex growth in the free shear layer. ...................................................... 53
Figure 6: Instantaneous PIV image demonstrating two vortices pairing ..................................... 53
Figure 7: Comparison of an air jet and a helium jet at similar Reynolds numbers...................... 54
Figure 8: Schlieren photograph of a helium jet exhibiting a side jet. .......................................... 55
Figure 9: The scaled contours of the axisymmetric nozzles used in the experiment. .................. 56
Figure 10: Experimental apparatus setup for PIV........................................................................ 57
Figure 11: Cross-sectional view of piping leading up to coaxial nozzle apparatus ..................... 58
Figure 12: Velocity profile of the central air jet at 0.63 mm above the nozzle exit plane........... 59
Figure 13: Velocity fluctuations profile of the central air jet at 0.63 mm above the nozzle exit . 60
Figure 14: Setup used for PIV. .................................................................................................... 61
Figure 15: Sample instantaneous vector fields ............................................................................ 62
Figure 16: Schlieren setup ........................................................................................................... 62
Figure 17: Sample schlieren images ............................................................................................ 63
Figure 18: Sample PIV images for a central air jet with a 50 m/s mean velocity ....................... 66
Figure 19: Sample PIV images for a central air jet with a 17 m/s mean velocity ....................... 68
Figure 20: Streamwise velocity contours. ..................................................................................... 69
Figure 21: Streamwise centerline velocities. ............................................................................... 70
Figure 22: Transverse velocity contours. ..................................................................................... 71
Figure 23: Linear region of the shear layer width........................................................................ 72
Figure 24: Shear layer spreading rate for varying mass flow ratios. ........................................... 72
Figure 25: Streamwise velocity fluctuation contours. ................................................................. 73
Figure 26: Transverse velocity fluctuation contours. .................................................................. 74
Figure 27: Centerline transverse velocity fluctuations. ................................................................ 75
Figure 28: Centerline streamwise velocity fluctuations................................................................ 75
Figure 29: Turbulent kinetic energy production contours. .......................................................... 76
Figure 30: Peak transverse velocity fluctuations. ........................................................................ 77
Figure 31: Peak streamwise velocity fluctuations. ...................................................................... 77

- vi -
Figure 33: Maximum vorticity per streamwise location. ............................................................. 79
Figure 34: Instantaneous vorticity contours ................................................................................. 80
Figure 35: Power spectra at x/D=.45, y/D = .45. ......................................................................... 81
Figure 36: Power spectra at x/D = 2, y/D = .25. .......................................................................... 82

- vii -
- viii -
LIST OF TABLES

Table 1: Annular to Inner Coaxial Jet Parameters ....................................................................... 26


Table 2: PIV Test Conditions ...................................................................................................... 26
Table 3: Measured Shear Layer Properties .................................................................................. 42

- ix -
NOMENCLATURE
a speed of sound

B shear layer width

D jet diameter

f focal length

vortex passage frequency

L nozzle contraction length

Lp potential core length

M annular to inner momentum flux ratio

Mc convective Mach number

PIV particle image velocimetry

volumetric flow rate

r nozzle contraction radius

rQ annular to inner mass flow ratio

rv annular to inner mean velocity ratio

R planar velocity ratio

R0 nozzle inlet radius

Re Reynolds number

RL nozzle outlet radius

R.xx radial location of xx% the local centerline velocity

S annular to inner density ratio

St Strouhal number

u local streamwise velocity component

U mean jet exit velocity

-x-
Uo jet centerline exit velocity

v transverse velocity component

x streamwise distance

x0 virtual origin location

y transverse distance

Greek Symbols

∆t laser pulse separation

ρ density

θ momentum thickness

kinematic viscosity

Notations

( )i relative to inner jet

( )a relative to annular co-flow

( )’ root mean square of the velocity fluctuation component

- xi -
ABSTRACT

The near-field flow characteristics of an axisymmetric air jet issuing into ambient air

were studied, with a focus on the structure and growth of the instabilities and vortices within the

shear layer. Changes in the jet’s stability characteristics and spreading properties were observed

with the application of small levels of a helium co-flow shroud. The addition of the low-density

gas served to introduce density gradients within the shear layer, altering the density-weighted

vorticity profile, and a subsequent suppression of the Kelvin-Helmholtz instability. Density

gradients, visualized by the schlieren technique, indicated an increased delay in the formation of

vortex roll-ups and a greater suppression of the vortex pairing process for increased low-density

co-flow. Power spectra plots, obtained using hot-wire anemometry, suggest that the initial

instability frequency of the air jet remains constant throughout all conditions examined, however,

they are made more stable with increased annular flow. The effects of a more stable jet and

changes to the vortex dynamics brought on by a density gradient in the shear layer were revealed

in mean and instantaneous flow data obtained using particle image velocimetry. Co-flowing

mass flow rates of 2.5% resulted in a potential core length increase of 41% due to a 20%

reduction in the shear layer spreading rate. Reduced mixing and entrainment, and an overall

change of the turbulence profile also resulted from the applied annular flow.

- xii -
CHAPTER 1: INTRODUCTION

1.1 Motivation
The coaxial jet, or co-flowing axisymmetric jet, consists of a central jet surrounded by an

annular fluid flow. Each of the two streams exit a nozzle, where they are initially separated by a

thin splitter plate and undergo mixing and develop to a turbulent state typical of free shear flows.

This flow configuration is relevant in the aerospace industry and has been studied for roughly the

past 60 years. Many efforts have been made in the near-field region of the jet exit (less than

approximately 10-20 diameters downstream), where large coherent vortical structures are

generated due to flow instabilities. These structures facilitate mixing and shear layer growth and

ultimately break down to a fully-developed turbulent state. The round jet near-field region is of

particular interest for applications that would benefit from enhanced mixing (e.g. combustion

chamber design), as well as reduced turbulence intensity (e.g. jet exhaust noise suppression).

The aim of the current study is to investigate the near-field mixing characteristics of an

axisymmetric air jet with the presence of a density variation in the shear layer, generated via an

annular nozzle issuing varying amounts of helium co-flow. This simple configuration is

employed to achieve a density profile similar to one seen in jet flames, the inspiration for this

study, but without any of the associated temperature gradients.

1.1.1 Mixing Enhancement

Air-breathing engines in aerospace applications rely on flow dynamics to create a mixture

of fuel and air necessary for combustion. In an industry where weight is a major factor in

driving costs, it is desirable to minimize the required distance to thoroughly mix the two fluids

rapidly to a homogenous condition, as to permit a shorter and more efficient combustion

chamber design. Since bulk mixing and shear layer growth are directly governed by the

-1-
behavior of the coherent vortex structures in the early stages of shear flow evolution, emphasis

for enhanced mixing (or reduced mixing, for that matter) has been placed on efforts that alter the

expression of jet instabilities. Passive control can achieve this through geometric modifications,

while active control accomplishes this with energy addition to the flow by means of excitation:

acoustics, a flapping ribbon, thermal techniques, or hydrodynamic methods, for example(Ho and

Huerre 1984). Using schlieren flow visualization and pressure measurements, Kedia and

Kurian(Kedia and Kurian 2005) showed that tabbed and lobed nozzles, a form of passive mixing

control, result in supersonic jets with reduced potential core lengths and greater mixing

capabilities. Samimy et al.(Samimy, Kim et al. 2007) demonstrated control over instability

mode expression of a jet, and was able to decrease the potential core length and enhance mixing

with electrical excitations from arc filament plasma actuators, located azimuthally around the

nozzle exit and using less than 1% of the jet flow power. These are but two accounts

demonstrating different methods of control.

1.1.2 Turbulence and Mixing Suppression

Reduced mixing and reduced turbulence via active or passive control is also of interest to

the aerospace industry. Tam(Tam, Viswanathan et al. 2008) identified both fine-scale turbulence

and the large coherent turbulent structures, produced from jet instabilities, as the sources of noise

in axisymmetric jets. It was also concluded that the dominant source of noise is located

immediately downstream of the jet potential core, in agreement with others(Laufer 1976;

Schaffar 1979) who identified it to be confined between one and two potential core lengths

downstream of the nozzle exit. Acoustic theory first put forth by Lighthill(Lighthill 1952)

speculated jet noise changes proportionally to the exit velocity to the eighth power. It was later

shown that noise reduction could also be achieved by introducing an annular jet of the same

-2-
fluid. A regime of annular to central jet velocities below one resulted in reduced noise levels as

observed experimentally by Williams, Ali, and Anderson(Williams, Ali et al. 1969), with the

maximum attenuation near a velocity ratio of one half. Balsa and Gliebe(Balsa and Gliebe 1977)

also demonstrated reduced noise levels for velocity ratios less than one using numerical methods,

citing reduced turbulence as the cause. More recently, other intricate methods, such as water

injection(Krothapalli, Venkatakrishnan et al. 2003) and microjets(Arakeri, Krothapalli et al.

2003), suppressed turbulence by 10-30% in a Mach 0.9 round jet, resulting in near-field noise

reductions of 2 to 6 dB. In both experiments, the injection of secondary flow was small in

comparison to the central jet: mass flow rates of approximately 5% and 1%, respectively.

A relevant mixing suppression application has been developed and patented by Praxair

Inc. for electric arc furnaces in the steel industry. Engineers developed a process and apparatus

called CoJet™ to deliver oxygen at supersonic speeds into molten metal baths. The technique

uses an oxy-fuel flame shroud and special injector nozzle to maintain original oxygen jet

diameter and velocity up to 70 diameters(Mathur and Messina 2001). This type of active

control, which has an overwhelming impact on the original oxygen jet dynamics, is similar to

those being utilized in the present investigation.

1.1.3 Current Study

Similar to some of the more recently developed methods eluded to, the current

investigation attempts to explore an active control technique to alter jet mixing behavior, using

an input that is small in magnitude in relation to the jet. Specifically, the goal of the study is to

explore the impact of a density variation in the shear layer of an axisymmetric jet. This is

achieved by introducing a weak co-flow of helium through an annular nozzle in a coaxial jet

configuration with the assumption that the low-density gas is entrained into the shear layer of the

-3-
jet. Motivation for studying this technique of introducing a density variation comes from the

insight of Day, Reynolds, and Mansour(Day, Reynolds et al. 1998), who showed through

numerical simulations that the behavior of the mixing layer can be described and predicted from

the density-weighted vorticity profile, which is altered from the effects of both compressibility

and heat release. In addition, Rehm and Clemens(Rehm and Clemens 1999) demonstrated a

striking difference in jet structure between reacting and non-reacting planar jets, also indicating

that density variation within the shear layer may influence jet behavior; in this case a result of

heat release. Although density variation in the shear layer, as a means of control is a key point of

the study, direct measurements of density/composition within the flow-field are not made.

Instead, qualitative and quantitative techniques are used to compare jet structure and behavior,

with and without helium co-flow. Mean and instantaneous measurements are made using

particle image velocimetry to extract velocity, vorticity, and turbulence properties of the flow. In

addition, schlieren is used to visualize the flows, and a hot-wire probe is used to measure jet exit

conditions and obtain power spectra measurements.

-4-
CHAPTER 2: LITERATURE REVIEW

2.1 Free Shear Layers


A free shear layer, or mixing layer, is the region in an unbounded fluid flow,

characterized by transverse velocity gradients, high levels of fluctuating velocity, and intense

mixing. It originates from the detachment of a boundary layer from the end of a surface, usually

a sharp splitter plate or trailing edge in the experimental setting, and it spreads approximately

linearly(Brown and Roshko 1974). Although not experimentally realistic, a shear layer can also

originate from an abrupt discontinuity in velocity, in the idealized inviscid condition(Lugt 1983).

Initial and downstream mean velocity profiles are shown in Figure 1 for a generic free shear

layer consisting of a single semi-infinite planar stream with some arbitrary velocity, U, coming

into contact with a stagnant fluid. Progressing downstream, the velocity profile broadens and the

shear layer grows in thickness, resulting in increased mixing between the main flow and ambient

fluid. The broadening of the velocity profile is coupled with the diffusion of vorticity as

governed by the vorticity transport equation (see Figure 2 and Figure 3 for profiles of the

axisymmetric shear layer from the current study). Due to the large velocity gradients at the

separation point, the vorticity will be at a maximum there in the absence of vorticity-generating

mechanisms within the flow. Progression downstream also leads to the evolution of a fully-

developed mixing layer, regardless of the initial conditions. Bradshaw(Bradshaw 1966) found

this to transpire at approximately 1000 initial momentum thicknesses downstream of the shear

layer separation point.

Aside from the simplistic single stream shear layer, emphasis has recently been placed on

the study of free shear layers between multiple interacting streams, as there are many practical

applications where two streams at different conditions are experiencing mixing. Ko et al.(Ko and

-5-
Kwan 1976; Ko and Au 1985) for example, investigated co-flowing shear layers in the

axisymmetric jet, while Strykowski et al.(Strykowski and Niccum 1991; Strykowski and

Wilcoxon 1993; Forliti, Tang et al. 2005) has extensively studied the counter-flowing shear layer

for both planar and axisymmetric geometries.

2.1.1 Axisymmetric Jets

The axisymmetric shear layer is more complex than its planar counterpart, however, the

principles governing its behavior are essentially the same. In most experimental efforts, focus is

placed on the round jet exiting from a contraction nozzle, such that only a small boundary layer

exists near the wall, giving way to a top-hat velocity profile at the trailing edge. Due to this

geometry, the shear layer spreads radially both away from and towards the jet axis in the

downstream direction. Unlike the planar case, the axisymmetric shear layer eventually

converges on itself, shown experimentally to occur at the jet axis between 4 and 5 diameters

downstream(Dimotakis, Miakelye et al. 1983; Gutmark and Ho 1983). This pinches off a region

referred to as the potential core. The conical core of flow is nearly inviscid, maintaining the

initial jet exit velocity, and it is only weakly influenced by vorticity within the shear layer. The

extent of the potential core can be measured by the intersection of the constant and the

hyperbolic decrease in jet centerline velocity(Hinze 1959). Turbulence intensity, and more

specifically, the centerline turbulence intensity, is shown to reach a maximum at a location

between the end of the potential core and approximately 10 diameters downstream due to the

merging of shear layers(Boguslawski and Popiel 1979; Warda, Kassab et al. 1999).

While the shear layer inner boundary intersects at the end of the potential core, the outer

boundary continues to spread indefinitely, influencing a greater region of fluid. Downstream of

the potential core is a development or transition region, followed by a region of fully-developed

-6-
turbulent flow and self-similarity(Hinze 1959) (see Figure 4). The self-similar solutions for both

mean velocity and turbulence quantities correspond to flow induced by a point source of

momentum located at a point on the jet axis, called the virtual origin (x0), which is found to vary

in location between upstream, downstream, or at the jet exit as a result of differing jet exit

conditions(Hussain and Zedan 1978; Revuelta, Sanchez et al. 2002).

2.1.1.1 Coaxial Jets

The structure of the coaxial jet varies from the single round jet by the presence of a

secondary or outer potential core and mixing region from the existence of the annular jet.

Interaction between the central and annular jets occurs within the inner mixing layer of the jet,

while the outer or secondary shear layer combines annular and ambient fluids. This type of

configuration was utilized in the current investigation as a means of introducing helium in close

proximity to the shear layer of an air jet. Because of the flows studied, focus is strictly placed on

coaxial jets with small outer to inner velocity ratios.

The homogenous coaxial jet structure and behavior was extensively studied by

Champagne and Wyganski(Champagne and Wygnanski 1971), and Ko and Kwan(Ko and Kwan

1976; Kwan and Ko 1977). They found the velocity and turbulence profiles far downstream

reach a self-similar state that depends only on virtual origin location and total jet momentum, in

agreement with single free jet results. They also established that a velocity ratio less than one

lengthens the potential core, while a velocity ratio greater than one shortens it, enhancing mixing

between the two streams. Hinze(Hinze 1959) presents an empirical relation established by the

experimental work of W. Forstall and Shapiro(Forstall and Shapiro 1950), correlating the length

of the potential core to velocity ratios over the range of 0.2 to 0.5, which will be referenced to in

Chapter 4. According to the relation, a trend of increasing potential core lengths should be

-7-
observed for increasing velocity ratios, λ, (below one) of annular to inner jet flow. Subsequent

studies have gone on to examine how jet dynamics are affected by the velocity ratio in more

detail and over a broader range, as well as other parameters such as the ratio of exit areas, lip

thickness, density ratio, and momentum flux ratio (taking both density and velocity ratios into

account). One particular focus was placed on determining the effect of coaxial jet parameters

on acoustic noise production. Williams and Ali(Williams, Ali et al. 1969) and Balsa and

Gliebe(Balsa and Gliebe 1977) identified attenuation of jet noise for velocity ratios less than one,

both experimentally and computationally, and cited reduced mean shear and turbulence levels in

the inner mixing region as the cause. Maximum reductions were found for velocity ratios

between 40% and 50%. Dahm et al.(Dahm, Frieler et al. 1992), utilizing flow visualization,

suggested that variation in jet structure and behavior from changes in velocity ratio are due to

changes in the instability characteristics in the near field of the flow. A more general discussion

on the effects of velocity ratio on a free shear layer will follow in section 2.1.3.

2.1.2 Instability

The mechanism governing the interaction of parallel streams of fluid is the Kelvin-

Helmholtz instability. It dictates the behavior of the flow in uniform density shear layers, as well

as the configuration in the current investigation, which is a variable density shear layer. The

stability characteristics of a particular flow are determined when small disturbances are

introduced, naturally or artificially, and disrupt the equilibrium base flow. If those disturbances

are amplified, the flow is termed unstable, and subsequently if the disturbances are suppressed,

or not amplified, it is termed a stable flow (with respect to the type of disturbance introduced).

The interaction of numerous parameters results in the specific stability characteristics of a

flow. A fluid’s viscosity has both a stabilizing effect by means of energy dissipation, while at

-8-
the same time a destabilizing one by means of momentum diffusion(Drazin and Reid 1981).

Nonetheless, the Reynolds number, whose lengthscale is defined according to the type of flow, is

attributed to driving the behavior. Above some critical Reynolds number, a flow is unstable, and

even above this value, the exact behavior of the instability is Reynolds number dependent.

Multiple investigations showed the instability of round jets evolve from a sinuous to pulsatile, or

varicose behavior at a critical Reynolds number (based on jet diameter) near the order of 103 for

both Pouiseuille flow jet exit conditions in water(Crow and Champagne 1971) and a top-hat

velocity profile in air(Becker and Massaro 1968). Therefore, experimental results suggest that at

Reynolds numbers above approximately 103, a round jet may be considered

turbulent(Boguslawski and Popiel 1979). However, other research has demonstrated flaws with

this train of thought. Heeg, Dijkstra, and Zandbergen(Heeg, Dijkstra et al. 1999) for example,

established that differences in pressure gradients resulted in drastically different critical

Reynolds numbers, varying by orders of magnitude. Lower critical Reynolds numbers resulted

from the presence of shear layer inflection points (where the derivative of vorticity equals zero),

explaining why counterflowing geometries are typically a more unstable flow. A wide range of

critical Reynolds numbers were also found for a number of different fluids under the same

Hagen–Poiseuille flow conditions by Novopashin and Muriel(Novopashin and Muriel 2002),

further demonstrating a universal independence of the Reynolds number on stability.

2.1.2.1 Vortex Evolution

At and above moderate Reynolds numbers in free shear flows, the appearance of large

coherent vortical structures or eddies in the shear layer is an inherent property of the Kelvin-

Helmholtz instability and they play an integral role in the transition of an initially laminar jet to

turbulence. The evolution of vortices in a free shear flow was described by Sato(Sato and Okada

-9-
1966) and Freymuth(Freymuth 1966) as consisting of four distinct development regions or

processes and is pictured(Lugt 1983) in Figure 5. The free shear layer originates from the

separation of the boundary layer from the surface of a trailing edge and is followed by

exponential growth. Due to non-linear growth downstream, vortices roll up and this

phenomenon occurs at regular intervals, referred to as the vortex shedding frequency. Becker

and Massaro(Becker and Massaro 1968) determined that the distance from a nozzle exit to the

location of vortex roll-up, or the “wave-breaking” length, is a function of the jet diameter and

Reynolds number. As instability waves form, less dominant fluid stream (or the ambient fluid) is

engulfed into the shear layer by the more dominant one, and a larger interface between the fluids

facilitates greater mixing. This process, termed entrainment, is more generically defined as the

inclusion of fluid from outside the jet boundaries into the main turbulent stream(Falcone and

Cataldo 2003). The final region in the evolution of the shear layer is characterized by vortex

pairing and transitioning to turbulence. A sample instantaneous PIV image is provided in Figure

6, which shows two vortices of equal size rolling around each other and pairing to create a larger

vortex, similar in size to the one seen downstream. In an axisymmetric jet, the final vortex

merging occurs near the end of the potential core(Gutmark and Ho 1983). Winant and

Browand(Winant and Browand 1974) showed that the pairing process between two vortices

results in a vortex structure doubled in both size and spacing (frequency) and occurs repeatedly

until they breakdown to turbulence. A more significant finding is that the pairing process

directly controls the growth of the shear layer. In fact, the growth between vortex pairings is

relatively small, giving rise to the appearance of a step in the shear layer thickness at the site of

vortex mergings. Ho and Huang(Ho and Huang 1982) expounded on this concept when they

- 10 -
demonstrated that a forcing mechanism could be used to significantly increase the shear layer

growth by the coalescence of multiple vortices at once, termed ‘collective interaction’.

2.1.2.2 Vortex Frequency

Besides controlling the growth of the shear layer, vortex structures also play an important

role in jet energy transport. Within the shear layer, vortices contain the largest amount of

turbulent kinetic energy, and transport this energy my means of inertial forces. Dominant vortex

passage frequencies are generally indicated by peak(s) within spectral density plots.

Determination of the vortex frequency within a flow is dependent upon the disturbances

(naturally or artificially produced) initiating the growth of instabilities in the flow. Wave

formation is dominated by a discrete frequency, corresponding to the most amplified (most

unstable) frequency predicted by linear theory. Forcing, or creating artificial disturbances,

commonly achieved by means of a mechanically oscillating ribbon or flap, or acoustic excitation

by an array of speakers, has been shown to influence a flow to express specific frequencies

and/or modes of instability; i.e. axisymmetric, helical(Ho and Huerre 1984). In the absence of

forcing, the peak frequency was demonstrated by Ko and Davies(Ko and Davies 1971) to be a

function of jet exit velocity, as well as axial and radial position in the near field (of an

axisymmetric jet).

Frequency is typically reported in the non-dimensional form of the Strouhal number, and

is given by , where is the preferred or peak frequency, U is the mean jet exit velocity

and L is some characteristic length scale. For a jet, the characteristic length is chosen to describe

the passage frequency for one of two sizes of instabilities, one on the order of the boundary

layer, scaling with the momentum thickness, and the other involving the jet column and scaling

to the jet diameter or width(Crow and Champagne 1971). In axisymmetric jets, Crow and

- 11 -
Champagne(Crow and Champagne 1971) found that the most amplified frequency occurring at

the end of the potential core, or the preferred mode, has a Strouhal number based on jet diameter

of 0.3. Gutmark and Ho(Gutmark and Ho 1983) acknowledged that a wide range of Strouhal

numbers, from 0.24 to 0.64, have been reported in other studies, and attributed this to scatter in

the location of measurement and the range of frequencies that could be encountered as a result

due to jumps in frequency from the vortex pairing process. Ho and Huerre(Ho and Huerre 1984)

note that the range of data could be a result of data processing techniques, noise present in

individual facilities, or varying initial conditions (momentum thicknesses). Instabilities on the

order of the initial boundary layer, referred to as the shear layer or initial instability mode, also

show scatter in reported values, but maintains a Strouhal number of approximately 0.017 for the

axisymmetric jet(Gutmark and Ho 1983).

2.1.2.3 Linear Stability Theory

An analytical/numerical means has been established which analyzes the origins of

instability within a shear flow, and can be a useful means for studying factors influencing their

formation. The basis of this technique, termed linear stability theory, is to determine when a

flow becomes unstable due to the introduction of infinitesimal perturbations. The equations

governing the behavior of the flow are derived from a linearization of the Navier-Stokes

equations, where second order or higher fluctuating quantities are assumed negligible. This

assumption limits the relevance of linear stability theory to the very early stages of shear flow

instability. The resulting linearized equations for the parallel shear layer include the Orr-

Sommerfeld equation for viscous flows and the Rayleigh equation for inviscid (high Reynolds

number) flow. Solving the fourth order differential equation eigenvalue problem requires the

utilization of advanced approaches such as the shooting method, or projection method; a popular

- 12 -
projection method is to use Chebyshev polynomials, for example. Because of the complexity of

the governing equations, early analyses have only been performed on a number of simple

velocity profiles for parallel and axisymmetric shear flows, including Blasius, Pouiselle, and

hyperbolic tangent profiles. However, advancement in techniques and computing has led to the

study of more complex velocity profiles and even variable density shear layers(Colucci 1993;

Srinivasan, Hallberg et al. 2010).

To successfully model a flow using linear stability theory, one must choose between

using spatial or temporal theory, which refers to the disturbances evolving in either space or

time(Huerre and Monkewitz 1985). Disturbances in unstable flows that grow and develop into

large vortex structures form at some preferred frequency depending on the conditions of the

flow. This frequency can be approximated by linear stability theory by mapping wavelength

amplification rates, as Ho and Huerre(Ho and Huerre 1984) have done (in their Figure 2). The

dominant instability forms at only the most amplified wavelength. Comparisons to experimental

results confirm the usefulness of this analytical tool. For example, one can compare the

experimental results of Strykowski and Niccum(Strykowski and Niccum 1992) to the linear

stability theory predictions by Pavithran and Redekopp(Pavithran and Redekopp 1989).

Nonetheless, there is still much ground to cover in gaining a more complete understanding of the

overall interaction between the numerous possible initial conditions and the resulting instability

behavior of a free shear flow.

2.1.3 Effects of Initial Conditions

The spatial developmental behavior of the free shear layer is determined by the initial

conditions. Consequently, the large scatter in jet data between similar studies has been attributed

to the varying initial conditions between jet facilities(Hussain and Zedan 1978; Gutmark and Ho

- 13 -
1983; Ho and Huerre 1984), as previously mentioned. The velocity ratio, density ratio, laminar

or turbulent initial boundary layer, facility disturbances, and the velocity profile and associated

flow properties are believed to be key initial condition parameters influencing the shear layer and

jet development. The influence of the jet Reynolds number and velocity ratio were previously

discussed: the velocity ratio effect was earlier described in terms of jet development, and will be

briefly touched on again with a focus on vortex structures. A thorough discussion on density

ratio will follow this section, as density is the focus driving the current research.

The initial laminar or turbulent state of the boundary layer was found to play a key role in

the jet near-field, specifically in shear layer development. It is suggested that the far-field

behavior is relatively unaffected by this characteristic, as self-similarity and asymptotic

turbulence levels were shown to occur at approximately the same x/D value for both cases(Ho

and Huerre 1984; Antonia and Zhao 2001). The state of the boundary layer is not a clear-cut

classification process, but has been defined based on a collection of factors including the

frequency spectrum, shape factor, and fluctuation intensity magnitude and profile(Hussain and

Zedan 1978). The criteria used in this study to classify the initial boundary layer will be

described later. A turbulent boundary layer, however, is typically achieved via a tripping

mechanism and is found to disrupt vortex formation and pairing during early nonlinear growth,

resulting in smaller spreading rates. This is attributed to turbulence being spread over a wider

range of wavenumbers(Russ and Strykowski 1993; Xu and Antonia 2002). Adversely, the

initially laminar boundary layer leads to regular pairing of coherent, organized structures and

thus, usually a larger shear layer growth rate in the developing region.

- 14 -
The shear layer growth rate is also highly dependent on the ratio of velocities across

streams of fluid. The velocity ratio is given by , where ∆U is the velocity difference

and provides a measure of the shear that drives amplification of instabilities, and Uav is the

average velocity, and provides a measure of the convection velocity of the structures(Ho and

Huerre 1984). Brown and Roshko(Brown and Roshko 1974) made the argument that R related

linearly to the shear layer growth rate, but that dependence is also based on the density ratio

across the streams. For lower values of R, vortex structures are stretched and flow development

is extended(Ho and Huerre 1984). This corresponds to increased potential core lengths in

axisymmetric jets. Dahm et al.(Dahm, Frieler et al. 1992) investigated the mechanisms behind

such trends using a laser-induced-fluorescence technique to visualize a coaxial jet configuration

over a wide range of velocity ratios. He found that vortex size and structure were greatly

affected across varying velocity ratios due to competition between instability modes.

Dominance of either the inner or outer shear layer, the mode of streamwise vortices (e.g.

axisymmetric or helical), and rotational interaction in the shear layer were all found to be

dependent upon the velocity ratio. Rotational interaction within the shear layer was defined by

the velocity profiles and termed shear-layer-like where dominated by vorticity concentrations of

one sign of circulation (e.g. coflowing streams) or wake-like where dominated by vorticity

concentrations of both signs of circulation (e.g. counterflowing streams). Although a smaller

range of velocity ratios were considered, Ko and Au(Ko and Au 1985) also observed competition

between the inner and outer shear layer structures, which led to varying mean jet properties.

Thus, it has been established that the jet shear layer is highly dependent on the exiting conditions

at the trailing edge because of varied vortex dynamics and instability mode competition.

- 15 -
2.1.4 Effects of Density

2.1.4.1 Single Shear Layers

The effects of density on the free shear layer have been studied both in terms of an initial

or freestream condition (i.e. variable density jets) as well as a more complicated entity entailing

compressibility (i.e. high mach number flows) and heat transfer (i.e. combustion and jet flames)

over the entire field of the jet. Brown and Roshko(Brown and Roshko 1974) studied the effects

of the ratio of the density of the freestream fluids on the plane mixing layer and found increased

spreading rates, based on vorticity thickness, when the high speed stream has lower density than

the low speed stream. Correspondingly, axisymmetric jet studies have demonstrated that the

low-density helium jet(Amielh, Djeridane et al. 1996) and heated jet(Russ and Strykowski 1993)

increase the spreading rate, decrease the potential core length, and result in more rapid velocity

decay compared to constant density jets.

More intense mixing and a quicker breakdown to turbulence was observed by the author

in some preliminary schlieren flow visualization in a helium jet compared to that of an air jet at

equal Reynolds numbers (see Figure 7). Observed trends result from changes in the stability

characteristics of the flow. Kyle and Sreenivasan(Kyle and Sreenivasan 1993) cited the

energetic and regular vortex pairing process, occurring closer to the jet exit, as the cause of early

potential core breakdown in low-density jets. Russ and Strykowski(Russ and Strykowski 1993)

also concluded a more organized vortex formation and pairing process taking place at lower jet

densities from narrower spectral distributions of disturbances. Kyle and Sreenivasan found the

vortex passage frequency to be dependent on density as well, as they observed a decrease in the

Strouhal number with jet density. This finding agrees with linear stability theory predictions:

see Trouve(Trouve, Candel et al. 1988) for a heated jet example.

- 16 -
Variation in density can also alter the type of instability within the flow. Kyle and

Sreenivasan(Sreenivasan, Raghu et al. 1989; Kyle and Sreenivasan 1993) observed a shift to a

more intense oscillatory instability, occurring for jet-to-ambient density ratios less than 0.6. This

compares to the linear stability theory analysis of Monkewitz and Sohn(Monkewitz and Sohn

1988), who determined a transition from convective to absolute instability to occur in hot jets for

density ratios less than a critical value of 0.72. Absolutely unstable flows are characterized by

exponential growth of disturbances everywhere in the flow, whereas disturbances are convected

away as they amplify, and leave the flow undisturbed, in the case of convectively unstable

flows(Huerre and Monkewitz 1985). Absolutely unstable flows also exhibit frequency spikes in

the spectral density plot with a dominant peak corresponding to the main disturbance frequency

and numerous harmonic and subharmonic peaks(Sreenivasan, Raghu et al. 1989).

Enhanced spreading and mixing in low-density jets is also facilitated by the formation of

side jets. Side jets result in radial ejection of fluid from the jet and subsequent entrainment.

Sreenivasan, Raghu, and Kyle(Sreenivasan, Raghu et al. 1989) captured photos of these side jets

using laser induced fluorescence, as did the present author using the schlieren flow visualization

(see Figure 8). Monkewitz et al.(Monkewitz, Lehmann et al. 1989) observed anywhere between

2 and 6 of these non-axisymmetric structures and attributed their formation to the Widnall

instability of the primary vortex rings. He also suggested that side jets are similar to, but more

intense than the formation of streamwise vortices, which result from azimuthal instabilities and

are found in unforced homogenous jets; observed by Yule(Yule 1978) and shown by Liepmann

and Gharib(Liepmann and Gharib 1992) to enhance entrainment.

2.1.4.2 Variable Density Coaxial Jets

- 17 -
The study of coaxial jets with large density differences has been limited, and usually

considers only a very specific configuration and range of parameters. Favre-Marinet et

al.(Favre-Marinet, Camano et al. 1999; Favre-Marinet and Schettini 2001) investigated the

regime of large momentum flux ratio coaxial flows with a low velocity, high-density central jet

and high velocity, low-density annular jet configuration, in order to model the conditions of

liquid propellant rocket engine injectors. It was suggested that the dynamics of a variable

density coaxial jet is governed by the momentum flux ratio (which takes the velocity ratio and

the density ratio into account) as it was found that the location of a regime of recirculation and

the potential core length were highly dependent on it. For the case of the single jet in coflow, a

trend of enhanced mixing was observed when the high-velocity stream was lower density. This

same trend was also confirmed for the coaxial jet configuration by Favre-Marinet, as well as by

Gladnick. In addition, Gladnick et al.(Gladnick, Enotiadis et al. 1990) concluded that the near

field mixing of a variable density coaxial jet is dominated by the large structures in the primary

shear layer, which he found to vary drastically for ratios between the co-flow and jet velocity

above, below, and equal to one. Transition to absolute instability has also been observed for the

right combination of velocity and density ratios in coaxial jets(Sreenivasan, Raghu et al. 1989;

Kyle and Sreenivasan 1993). Critical density ratios on the absolute/convective instability

interface for axisymmetric counterflowing streams were experimentally found and mapped for

varying velocity ratios by Strykowski and Niccum(Strykowski and Niccum 1992), with the trend

of decreased density ratios (greater than one) needed to obtain an absolutely unstable flow for

increased velocity ratios.

2.1.4.3 Reacting Jets

- 18 -
Reacting jets exhibit strikingly different behavior than a similar non-reacting jet as a

result of heat release. The heat release results in volumetric expansion in a localized region near

the flame, exhibiting low density and reduced viscosity. Debate ensues as to the specific cause

for the consequential changes in jet behavior. Nonetheless, reacting jets are a more complicated

subject matter because their behavior is no longer solely dominated by their initial conditions;

rather they are highly influenced by the combustion process in or near the free shear layer. For

round jet flames, the combustion process has a stabilizing and laminarizing effect on jet

structure, causing reduced shear layer growth rate, slower decay of the jet centerline velocity,

and reduced entrainment(Rehm and Clemens 1999; Han and Mungal 2001). Savas and

Gollahalli(Savas and Gollahalli 1986) demonstrated that these effects may be due to stark

differences in the behavior of the vortex structures formed in the shear layer. Schlieren

photographs showed that in the case of an attached flame, the formation and roll-up of the

vortices is delayed and pairing less frequent than those in a non-reacting jet, resulting in a slower

growth rate. However, he notes the lifted flame case displays a structure more similar to that of

the non-reacting case, implying that flame location matters. Furi et al.(Furi, Papas et al. 2002)

also concluded that flame location has a significant effect on the Kelvin-Helmholtz instabilities

and attributed it to a redistribution of the density-weighted vorticity profile, ρ du/dy (where

dv/dx is assumed negligible). Day, Reynolds, and Mansour(Day, Reynolds et al. 1998)

performed numerical simulations on a compressible, reacting mixing layer and demonstrated that

the density-weighted vorticity profile is a useful characterization to explain and predict growth

trends of the shear layer. He established that density variation from heat release could result in

multiple peaks in the density weighted vorticity profile, giving rise to the development of inner

and outer instability modes corresponding to the high and low speed side of the shear layer

- 19 -
2.1.5 Effects of Compressibility

A high-speed, compressible flow is a complex entity where large velocity, density, and

pressure gradients are present within the shear layer. The magnitude of the effects of

compressibility are generally described in terms of the convective Mach numbers, Mc, of the

turbulent shear layer, defined by Papamoschou and Roshko(Papamoschou and Roshko 1988) as

Mc1=(U1-Uc)/a1 and

Mc2=(Uc-U2)/a2, where a is the speed of sound and the subscripts 1 and 2 are the two free

streams. The general trend observed both experimentally and through numerical means is

reduced growth rates with increased convective Mach numbers(Papamoschou and Roshko 1988;

Gutmark, Schadow et al. 1995; Freund, Lele et al. 2000; Pantano and Sarkar 2002). Spreading

rates drop steadily until an asymptotic value approximately 20% of the spreading rate of

incompressible shear layers is reached near a convective Mach number of 0.8(Gutmark,

Schadow et al. 1995). Gutmark, Schadow, and Yu(Gutmark, Schadow et al. 1995) point out that

density ratios of 49 produced by Brown and Roshko(Brown and Roshko 1974) resulted in

changes in the growth rate by less than a factor of 2, implying density is not the only significant

parameter in compressible flows. However, just like variable density flows, the compressible

shear layer is greatly influenced by the coherent structures formed by jet instabilities.

Numerous studies find vortex structure and behavior in the compressible shear layer

begin to deviate from what is typical for the incompressible case somewhere in the range of

convective Mach numbers between 0.5 and 0.6(Sandham and Reynolds 1990; Papamoschou

1991; Elliott, Samimy et al. 1995; Gutmark, Schadow et al. 1995). Gutmark(Gutmark, Schadow

et al. 1995) notes that compressibility results in less organized and less coherent structures. In

addition, Papamoschou and Roshko(Papamoschou and Roshko 1988) observed suppression of

- 20 -
vortex pairing within the shear layer and suggest energy loss from supersonic convective Mach

numbers may play a role. Linear stability theory was used to show the dominant instability

modes to transition from a 2-D to 3-D entity above convective Mach numbers of 0.6(Sandham

and Reynolds 1990). The transport of compressible shear layer instabilities also varies in the

convection velocity. Papamoschou(Papamoschou 1991) found the convection velocity of

structures to be close to one of the free streams under high speed conditions, as opposed to an

average of the free streams for the incompressible case. The combination of these differences in

compressible free stream vortex behavior results in a drastically reduced spreading rate.

Determining how the varying conditions of free streams interact still remains the struggle. The

current investigation is an attempt to isolate density from velocity, temperature, and

compressibility effects on shear layer instabilities and growth.

- 21 -
CHAPTER 3: FACILITY & INSTRUMENTATION

3.1 Nozzle Design


The purpose of this experiment was to fashion a means to modify the density profile

within the free shear layer of an axisymmetric jet, without also inducing any significant

temperature gradients, velocity gradients, or compressibility effects in the system. The intent of

this action was to observe coherent vortex structures emanating from the Kelvin-Helmholtz

instability, record their changes in structure and behavior due to density gradients, and determine

their influence on the growth of the shear layer. In order to accomplish this, the nozzle of the jet

was designed to produce a clean boundary layer with certain specific parameters to aid in the

investigation.

The momentum thickness of the boundary layer was minimized (D/θ maximized) such

that the different instability modes scaled with either the jet column or shear layer could be

identified separately. An ideal facility would yield a thin boundary layer with a density

inhomogeneity present at separation. While more complex modifications to an axisymmetric jet

could have been used to better achieve this, it was decided that a coaxial jet was an effective and

simplistic approach to accomplish not ideal, but sufficient conditions. With the addition of a

weak variable density co-flowing jet, the boundary layer of the main jet separates with a uniform

density profile, but it quickly entrains low-density fluid, early in the development of the shear

layer, creating the desired flow conditions slightly downstream of the jet exit.

The nozzles used in this study were designed by the author to accommodate appropriate

flow conditions; a discussion of the more specific considerations made during the design process

will follow. The ultimate goal was the design of coaxial nozzles facilitating an initially laminar

- 22 -
central air jet with a thin exiting laminar boundary layer, surrounded by a weak annular low-

density jet that could be manufactured and run economically. To assure a uniform jet would

issue from the nozzle with only a thin boundary layer, approximations for the boundary layer

development along the nozzle were made using Thwaites’ method(Thwaites 1960) where the

momentum thickness, θ, could be found at any location using the relation, ,

where ν is the kinematic viscosity, u the local mean velocity, and x the streamwise direction with

the origin at the nozzle inlet. The displacement thickness was then estimated by finding the

shape factor, H, using the relation H = 2.61 – 3.75λ + 5.24λ2, where λ is given as

. Boundary layer growth of the main jet flow was minimized with a large inlet-

to-exit area contraction ratio of 75, to produce favorable pressure gradients and result in a small

boundary layer Reynolds numbers based on displacement thickness to maintain a laminar

boundary layer state(Schlichting and Gersten 2003) over a moderate range of jet exit Reynolds

numbers. A compromise on the length of the nozzle contraction was made between a short

design to hinder boundary layer growth and lower manufacturing costs, and a longer design to

avoid flow separation(Mehta and Bradshaw 1979).

After selecting a nozzle contraction length, L, and inlet and exit radii, R0 and RL, a

smooth contraction contour was designed using a 5th order polynomial curve, which was found to

be the optimal geometric choice by Bell and Mehta(Bell and Mehta 1988) for low-speed wind

tunnels. Satisfying all boundary conditions ( r(0)=R0, r(L)=RL, and first and second derivatives

of r with respect to x are zero) leads to the solution of the polynomial coefficients yielding

, where r(ξ) is the contraction radius at the

non-dimensional location ξ = x/L. This chosen design method resulted in the contours depicted

- 23 -
in Figure 9. A final design consideration concerning the inner nozzle was the thickness of the

lip, or trailing edge. Mehta(Mehta 1991) argued that the finite thickness of a splitter plate

between co-flowing streams creates a wake region which affects mixing layer development.

Therefore, lip thickness was designed as thin as possible to minimize any of these effects, but it

was also limited by the fabrication technique. The nozzle was fabricated in ABS plastic by Dr.

Bahattin Koc of the Industrial Engineering Department at the University at Buffalo using a

Stratsys FDM 3000 rapid prototyping machine, which cannot create features smaller than 1.27

millimeters.

The focus of the secondary flow was more on mean quantities of velocity and mass

flowrate than on the velocity profile itself, contributing to a decision to use a moderate

contraction ratio. The purpose of the annular co-flow is simply to provide the presence of a low-

density fluid next to the main jet, and would contribute very little to the total momentum or

vorticity at the interface of the two streams due to velocity differences. To minimize the “co-

flow” or “velocity ratio” effects of the annular jet, a large exit area, approximately 1.8 times

larger than the main flow exit area was designed, such that the velocity of the annular jet was

less than 10 to 15 percent of the central jet for all experimental cases. Coupled with this velocity

requirement was a constraint for the annular mass flow rate to be small in relation to the main jet:

less than 5%. This was imposed for both economic reasons and because other recent

studies(Arakeri, Krothapalli et al. 2003; Krothapalli, Venkatakrishnan et al. 2003; Samimy, Kim

et al. 2007) on the active control of axisymmetric jets have shown significant changes in jet

behavior with minimal inputs.

- 24 -
3.2 Test Facility
A test facility was fabricated and assembled in house to study the effects of a low-density

co-flow on an axisymmetric air jet within the confines of the Combustion Laboratory at The

State University of New York at Buffalo. The facility is pictured in Figure 10 along with the

camera used for taking PIV images. Unistrut metal framing was used to create a stable support

for the test facility in a vertical orientation, with the nozzle exit plane standing at a height of

approximately 50 inches. A vertical orientation was chosen to avoid a situation where Rayleigh–

Taylor instabilities may have been a dominant feature of the flow, due to variable density layers

of fluid perpendicular to the gravitational vector. Air was delivered to the facility from the

building air supply through ½” I.D. braided tubing. Air then passed through 22 inches of 3¼”

I.D. polycarbonate tubing and a total of five 60 x 60 metal screens to condition the flow before

entering the nozzle inlet (see Figure 11). The screens assure uniform flow within the jet and

reduce turbulence levels to approximately 1.5% at the nozzle exit (see section 3.3).

The gas used for the annular co-flow was Praxair UN1046 helium, initially compressed to

2640 psig in a size T Praxair cylinder. A Praxair model # PRX31233 regulator was used to

control the output pressure of the tank. The helium flow was divided up into 4 separate streams

through two sets of tees and entered the annulus between the nozzle walls through four inlets

perpendicular to the nozzles and spaced 90 degrees apart.

3.3 Test Conditions


Results are presented in Chapter 4 for a single axisymmetric air jet with no co-flow as

well as for the same air jet with five cases of increasing amounts of helium co-flow. Properties

of the flow were scaled and non-dimensionalized according to the jet centerline exit velocity, Uo.

Regardless, the conditions of the central air jet were maintained approximately constant for all

- 25 -
cases. The mean velocity of the air jet, determined by dividing the volume flow rate by the

nozzle exit area, was roughly 50 m/s yielding a Reynolds number based on exit diameter, given

as of approximately 36,000.

The selection for the maximum amount of helium co-flow, as well as the incremental

steps in co-flow to be studied were selected after preliminary data was taken with schlieren flow

visualization and PIV, respectively. A schlieren technique (described in section 3.4.3) was

useful in determining the amount of co-flow (determined to be 2.5% helium by mass) necessary

to have a reasonably significant impact on the behavior of the axisymmetric air jet, while being

able to run experiments economically.

Table 1: Annular to Inner Coaxial Jet Parameters


Momentum Flux
Density Ratio Mean Velocity Ratio Mass Flow Ratio
Ratio

The (variable density) coaxial jet may be characterized by a number of different

parameters; four of which are defined in Table 1. Correspondingly, the experimental values for

each test case considered are given in Table 2.

Table 2: PIV Test Conditions

Case Base 1 2 3 4 5

Annular Mean Velocity (m/s) 0.0 1.0 2.0 3.0 4.0 5.1

Velocity Ratio (%) 0.0 2.0 4.1 6.4 8.1 10.2

Mass Flow Ratio (%) 0.0 0.5 1.0 1.5 2.0 2.5

Momentum Flux Ratio (%) 0.0 0.0056 0.0225 0.0510 0.0907 0.1415

- 26 -
In the case of most homogenous coaxial jet studies (e.g. Ko and Au(Ko and Au 1985);

Dahm et al.(Dahm, Frieler et al. 1992); Rehab, Villermaux, and Hopfinger(Rehab, Villermaux et

al. 1997)), the outer to inner velocity ratio is used to characterize different flow regimes.

Although few studies have focused on variable density coaxial jets, the work of Favre-Marinet et

al.(Favre-Marinet, Camano et al. 1999; Favre-Marinet and Schettini 2001) suggests that

momentum flux ratio is a more appropriate parameter used to describe the flow dynamics of that

configuration. Nonetheless, the results from this investigation are mainly given in terms of the

mass flow ratio. Although this term scaled linearly with the velocity ratio, it is more appropriate

to use to describe the conditions that are driving the changes. It is the presence and amount of

density variation within the shear layer, and not the speed of the low-density co-flow that is

responsible for the observed behavior changes, which is best expressed in terms of the mass flow

ratio.

3.3.1 Initial Conditions

A hotwire technique, described in section 3.4.2, was used to characterize the initial

conditions of the central jet, and specifically the exiting boundary layer by probing the plane

0.63 mm above the nozzle exit. It should be noted that measurements were taken above the exit

plane, and therefore the separated boundary layer of the free jet, although minimal, had some

distance from the nozzle exit in which growth could have taken place. The center of the jet was

located after probing with the hotwire and assuming a symmetric velocity profile across the jet.

High precision travel of the hotwire probe was achieved with a one axis ;spatial uncertainty of

.0005” using a 3-axis micrometer positioning stage. The axis of the jet was used as a starting

point to probe along four paths 90 degrees apart. The resulting mean velocity profile, given in

Figure 12 shows overlap at all four angles, indicating symmetry in the exiting flow. The thin

- 27 -
boundary layers due to the large contraction in the nozzle design result in a top-hat velocity

profile. The average boundary layer thickness on the perimeter of the nozzle, defined by 99% of

the freestream velocity, is approximately 0.49 mm or 15% of the nozzle radius. Using the

standard definition, given by , the initial momentum thickness of the jet

was found to be 0.047 mm, giving the ratio D/θ equal to 225. Velocity fluctuation intensity

profiles are plotted in Figure 13 for all four angles probed. The peak fluctuation intensity, found

to be less than 3% of the jet velocity, is located away from the wall and near the center of the

boundary layer thickness indicating an initially laminar boundary layer(Hussain and Zedan

1978).

3.4 Instrumentation & Data Collection Techniques

3.4.1 Particle Image Velocimetry

Particle image velocimetry (PIV) was the primary means of data collection in the present

investigation. Particle image velocimetry is a non-intrusive optical technique used to measure

instantaneous velocity vectors within a flow-field that can be used to extract information on

turbulence and velocity-dependent flow properties. The method employs a high-speed camera to

record pairs of images of tiny particles within a seeded flow, which are illuminated by a laser

light sheet. The technique produces a vector field by tracking the displacements of groups of

localized particles between image pairs separated by a known time. A hotwire technique is a

simple and typical way to sample flows; however, PIV was chosen to be the main means of data

collection in the current study due to multiple benefits over the hotwire technique. A hotwire

technique uses an intrusive probe, which has been shown to affect instability characteristics

when placed at small x/D(Sreenivasan, Raghu et al. 1989). In addition, a single hotwire probe

can only be used to sample a single point over a discrete time period. Additionally, hotwire

- 28 -
signals are dependent on fluid velocity and concentration(Kyle and Sreenivasan 1993), so due to

the inhomogeneity of gases within the shear layer of this study, hotwires could not be used in this

region. Despite these shortcomings, a hotwire probe was used to characterize the initial

conditions of the jet, particularly the boundary layer, due to its spatial resolution.

For the current study, tracer particles were introduced into the two flows using pressure-

regulated Laskin nozzles to create finely atomized droplets of olive oil. Olive oil was selected

due to its nontoxic properties and relatively constant mean diameter droplet size of 1µm under a

wide range of operating conditions(Kahler, Sammler et al. 2002). For the given particle size, a

compromise is made between the tracer adequately following the flow (a trait of smaller

particles) and providing a strong signal-to-noise ratio from the scattering of light (a trait of larger

particles)(Hart 1998). The building compressed air supply drove the Laskin atomizer for the

central air jet, and seeding density was controlled independently of the flow rate with an air

bypass across the atomizer. All of the helium co-flow passed through a second Laskin atomizer,

which was modified to produce an adequate seeding concentration for the low flowrates. A

reliable seeding concentration is suggested to be greater than approximately 15 tracer particles

per interrogation region(Keane and Adrian 1990). Due to entrainment of ambient air into the jet

flow, it was also necessary to seed the ambient environment using a third Laskin atomizer.

Seeding of the ambient was accomplished again using the building pressurized air supply and

was done prior to, and in between experimental runs, so as not to introduce any ambient velocity

gradients in the vicinity of the jet. A schematic of the PIV setup is provided in Figure 14.

A 50 mJ/pulse New-Wave research Solo III Nd:YAG laser, operating at a rate of 15 Hz,

was utilized for the present study. The 532 nm wavelength laser beam was turned into a

vertically planar light sheet by passing the beam through a spherical lens (f = 500 mm) and a

- 29 -
plano-concave cylindrical lens (f = -25.4 mm). Positioning of the laser sent the light sheet

through the coaxial jet axis and illuminated particles in a plane approximately 1 mm thick. The

time separation between laser pulses was determined to allow particles to be displaced

approximately ¼ of the size of the interrogation region at the maximum flow velocities, based on

the recommendation by Keane(Keane and Adrian 1990) to optimize the correlation between

images, and was calculated using the following equation: .

Images were captured using a high-speed IDT X-Stream XS-5 CCD digital camera with a

1280x1024 resolution. Synchronization between the laser and camera system was controlled

with IDT ProVISION-XS software through an HP xw4300 PC Workstation and an IDT

MotionPro X timing hub. The camera was oriented such that the larger viewing area

corresponded with the vertical or streamwise direction. Three camera positions, overlapping by

approximately 20% along the streamwise direction, were needed in order to obtain data up to 10

diameters downstream of the nozzle exit. At each camera position, 812 image pairs were

collected and correlated to acquire mean and turbulent velocity statistics. Data from the three

positions were later merged to produce the plots given in Chapter 4.

Image pairs were correlated using IDT ProVISION-XS software after applying a

brightness and contrast adjustment to improve correlations. Interrogation regions of 32x32

pixels (with a pixel equivalent in size to 33.96µm per side) were utilized with a 50% overlap,

improving the spatial resolution of the flow-field and eliminating correlation errors(Hart 1998),

and resulting in an 80x64 grid of velocity vectors. Evaluation of sample instantaneous vector

fields (Figure 15) reveal numerous spurious vectors close to the nozzle exit plane due to poor

image correlation as a result of light reflection off the nozzle as well as lack of penetration of the

- 30 -
laser sheet on the far side of the nozzle. As a result, reported data excludes information for the

first 0.3 diameters downstream of the exit plane.

3.4.2 Hot-wire Anemometry

A hotwire technique was used to characterize the initial conditions of the central jet, and

specifically the boundary layer by probing the plane 0.63 mm above the nozzle exit. Hotwire

was used instead of PIV for its high sensitivity, quick response time, spatial resolution, and

because of spurious vectors produced from PIV near the nozzle due to laser reflections. A

downside of hot-wire anemometry is that components of the velocity vector are

indistinguishable, and only the magnitude of the velocity can be obtained. It is assumed that near

the nozzle exit where measurements were taken, the transverse velocity component is negligible,

and measurements can be approximated as streamwise components only.

Data acquisition was conducted using a five micron diameter probe with an active sensor

length of approximately 1.2 mm, operating in constant temperature mode with an overheat ratio

of 50%. Signals from the Standard U-Wire Probe A55P01 from Auspex Scientific were acquired

with a DISA type 55M10 CTA standard bridge. A DISA 55D26 Signal Conditioner removed

noise from the signal using its low-pass filter mode. Output voltages from this configuration

were monitored with LABVIEW on a PC using a National Instruments BNC-2111 Data

Acquisition system. The hotwire sensor was initially calibrated with a pitot tube, using a Fluke

111 multimeter to monitor the voltage output from a Setra model 239 pressure transducer.

Power spectra measurements were made via a fast Fourier transformation of the auto-correlation

function, and sampling occured at 20,000 Hz to obtain a frequency range up to 104 Hz.

- 31 -
3.4.3 Schlieren Flow Visualization

Schlieren is a technique used to visualize gradient disturbances of inhomogeneous

transparent media resulting from temperature differences, high-speed flows, or mixing of

dissimilar materials(Settles 2001). The technique exploits variations of the refractive index

within transparent media to visualize density gradients in a spatial direction (i.e. ) with

varying degrees of light intensity. In order to generate the schlieren effect, parallel light rays

must pass through the region of interest and be focused to a point where a knife edge blocks a

portion of the original light. Without the use of a knife-edge, the shadowgraph technique results,

which shows the second derivative of density (i.e. ), and produces less contrasted

images. Schlieren can be done with a nominal point light source and different combinations of

lenses and mirrors. The schlieren setup for the current study is depicted in Figure 16 in a top

view. An Osram 150W xenon arc lamp was used as the light source and situated in an Ealing

Stabilarc 250 Lamphouse. An aperture on the lamphouse was used to focus light onto a 25.4 cm

diameter concave mirror approximately 203 cm away. The mirror collimates the light, which

then passes through the gases dispersing from the coaxial jet arrangement, and is refocused by an

identical concave mirror. At the focal point, a razor blade is oriented in the vertical direction,

partially blocking some of the light source to create the schlieren effect, and indicates density

gradients in the transverse direction of the jet. Light passing the razor blade is focused by a lens

and images are captured with the same IDT X-Stream XS-5 digital camera used for PIV.

Resulting images display high intensity (bright) regions corresponding to positive density

gradients and low intensity (dark) regions corresponding to negative density gradients.

- 32 -
Five total cases were investigated, each involving a central air jet with a mean velocity of

approximately 22 m/s and Reynolds number of 16,000. The base case with no co-flow, required

the addition of 0.5% helium by mass to introduce a slight density variation from the ambient air

and produce a great enough contrast for the schlieren system, while not significantly altering the

momentum or instability characteristics of the jet. The subsequent cases consisted of increasing

the ratio of helium co-flow to main air flow by increments of 0.74% by mass up to a maximum

of 2.96%.

3.4.4 Flow Metering

Gas flow rates entering the experimental setup were monitored using Dwyer Series RM

Rate-Master® Flowmeters at the locations depicted in Figure 14. The air supply for the center

jet passed through a model RMC-104-SSV flowmeter. A model RMB-51 flowmeter was used to

monitor the helium co-flow for cases 1-3, as defined in Table 1, and a model RMB-52-SSV for

cases 4 and 5. Westward 60 psi max pressure gages were installed near the rotameter exits to

monitor the pressure and account for gas density variation using the ideal gas law, assuming a

deviation from standard temperature was negligible. Flow rates could then be corrected using

the following: .

- 33 -
- 34 -
CHAPTER 4: RESULTS

4.1 Schlieren
A high intensity light source and a series of lenses and mirrors were set up as described in

section 3.4.3 to produce a schlieren effect, and images were captured using a high-speed camera.

The knife-edge used to produce the schlieren effect was oriented vertically, parallel to the axis of

the jet, in order to highlight transverse density gradients. Qualitative comparisons in jet structure

and behavior are conducted by examining images from each trial; images representative of

typical behavior are provided in Figure 17. These 2-D images can be considered an idealized

cross-sectional view of the jet, however there are some features resulting from the 3-

dimensionality of the jet structure.

The images taken for the co-flowing jets show high intensity bright and dark horizontal

strips near the exit plane on either side of the central jet, and correspond to positive and negative

density gradients, respectively. These strips visualize the significant reduction and then increase

of density in moving radially from the jet axis: going from the air jet, to the helium co-flow, and

finally to the ambient air. The regions on either side of the jet axis are symmetric in shape, but

opposite in intensity due to opposing signs in the spatial direction of the density variation. In all

cases, the symmetry of the jet behavior is further visualized by examining the instability

structures, especially in Figure 17(e). Symmetry of the vortex structures and the presence of

faint transverse bands in the jet core, a line of sight integration effect, confirm that axisymmetric

modes of instability are being expressed (and not helical or azimuthal modes). Although the

expressed mode of instability is consistent throughout the cases, the vortex behavior and

interaction is significantly different, comparing the base case to the 2.96% co-flow case. In the

base case, a definite growth in size of the vortex structures as they progress downstream is

- 35 -
clearly seen, and is due to the pairing process described by Winant and Browand(Winant and

Browand 1974). This causes growth of the shear layer, leading to a relatively quick merging and

degradation of the jet to a turbulent state. The addition of increased low-density co-flow is

observed to result in a transition to turbulence farther downstream, which also corresponds to the

approximate end of the large density gradients (seen at the interface of the central and annular

jets). However, changes in density gradient go hand-in-hand with the spreading of the shear

layer and transition to turbulence, analogous to the influence of velocity gradients on vorticity

and vice versa. Transition to a turbulent state appears to take place in image (e) even in the

presence of a relatively thick intense band (i.e. a large density gradient). This suggests that a

turbulent state is still able to transpire in the presence of a large density gradient, and without

significant vortex pairing.

Images of increased co-flow also show an increased number of rolled-up instabilities

present, which appear to have a smaller wavelength and transverse height. Although the vortices

appear to show different frequencies, they are not necessarily representative of the initial

instability frequency and this issue will be addressed in more detail with hotwire spectral plots

presented in Chapter 4.3. For the higher co-flow cases, numerous vortices in a row appear to be

the same in size, indicating that the mechanism inducing vortex pairing has been damped and

amalgamation is not taking place. Thus shear layer growth has been inhibited by the presence of

the low-density gas.

The smaller structures seen in the high co-flow cases don’t penetrate as far into the core

of the central jet, are less capable of inducing entrainment, and leave a more coherent column of

fluid. The helium appears to have an overall stabilizing effect on the jet, resulting in a more

laminar, unsteady behavior near the exit. These trends were also observed by means of schlieren

- 36 -
by Savas and Gollahalli(Savas and Gollahalli 1986), who achieved a similar density profile due

to heat release. The wave-breaking length, the distance from the exit plane to vortex roll-up

formation, is also observed to increase with the addition of a low-density annular flow. The

magnitude of this effect is apparently density driven and correlates well with the trends observed

by Kyle and Sreenivasan(Kyle and Sreenivasan 1993), who found wave-breaking lengths of jet

instabilities decrease as the ambient fluid was made denser.

4.2 Particle Image Velocimetry


A total of six flow cases with the same central jet properties but varying amounts of

helium co-flow were examined with particle image velocimetry, as defined in Table 2. Mean

flow properties were obtained using 812 image pairs at each of three camera locations in order to

capture data extending up to ten diameters downstream of the nozzle exit plane and 1.6 diameters

radially from the jet axis. Both mean and instantaneous PIV data are presented in the following

sections. All figures containing six contour plots correspond to the six cases aforementioned,

with the base case (no co-flow) on top and increasing co-flow from top to bottom.

4.2.1 Instantaneous Images

Sample instantaneous PIV images are provided which highlight differing seeding

densities between the three experimental fluid domains by adjusting image brightness and

contrast. Although not as informative as the schlieren images, the PIV images illustrate similar

near field behavior and dynamics. Images are referenced to in Figure 18 and Figure 19 for cases

corresponding to the central air jet mean velocity of 50 m/s and 18 m/s, respectively. The no co-

flow and low co-flow cases (approximately (a) through (c)) demonstrate large structures that

penetrate deep into the core of the air jet, pinching off the column of fluid, and causing a quick

transition to turbulent behavior. Again, the latter cases of relatively higher co-flow velocities

- 37 -
(approximately 1.5% co-flow by mass and above) exhibit smaller, more frequent vortices and a

more coherent jet column

Although absent from the schlieren cases, but typical of coaxial jet behavior, vortex

formation and instability growth is present in the secondary shear layer of the PIV images.

Nonlinear growth and the rolling-up of vortex structures begin to emerge at a mass flow ratio

around 2.0% and above. The provided images only show one or two vortex roll-ups each,

however they are multiple times larger than the instabilities of the inner shear layer, possessing a

longer wavelength. The formation of these secondary structures is not related to the initial

objective of this research, which was to limit the influence of coaxial jet behavior. However,

these additional structures facilitate the entrainment of ambient air into the annular column of

helium, reducing its ability to shroud the air jet in a pure low-density gas. Despite this, it is

assumed the outer structures have minimal effects on the overall behavior of the central jet

instabilities and spreading dynamics. Thus, the coaxial configuration is idealized as a single jet

configuration and increased co-flow can be thought of as maintaining a significant density

gradient over a greater streamwise distance of the central jet. However, progressing to even

higher co-flow rates may result in a totally different near field behavior all together, as was

found for homogenous coaxial jets by Dahm, Frieler, and Tryggvason (Dahm, Frieler et al. 1992)

and Ko and Kwan(Ko and Au 1985).

4.2.2 Velocity Field Statistics

4.2.2.1 Mean Streamwise Velocity and Potential Core Length

Mean data is presented to establish average behavior of the jet as a whole, followed by

instantaneous data to shed light on some of the underlying phenomena responsible for the

observed mean trends. Contours of the non-dimensional streamwise velocity, u, are first

- 38 -
presented in Figure 20. The most obvious observation is the presence of high velocities

maintained farther downstream as more co-flow is introduced around the central jet. The extent

of the contour line corresponding to 95% of the exiting velocity moves from 4.5 diameters

downstream with no co-flow to approximately 6.5 diameters downstream at the maximum co-

flow condition. The potential core length, Lp, was approximated for each case where the

centerline velocity was 98% of the exiting velocity, matching closely to the start of velocity

decay. A visual of this location is provided by Figure 21, a plot of the streamwise centerline

velocity distribution. Although the decay of centerline velocity occurs with nearly identical

slopes, the start of the decay (end of the potential core) moves downstream in somewhat regular

increments with the mass flow ratio. Potential core length values are reported in Table 3 and

range from 4.0D to 5.6D. Fitting a linear regression to the data to corresponding velocity ratios

yields a squared correlation coefficient of 0.984 and the relation: ,

where is again the velocity ratio. The empirical values reported by Forstall and

Shapiro(Forstall and Shapiro 1950) are 12 and 4 for the slope and intercept, respectively, but

come from a slightly higher velocity ratio range of 0.2 to 0.5. Nonetheless, the potential core

length from the current study is found to be comparatively more dependent and sensitive to

changes in the velocity ratio. Thus, the non-unity density ratio of the current study confirms

instability growth is density ratio dependent

Again looking at Figure 20, but this time at the low velocity (5%) contours, the width of

the jet and/or shear layer location can be approximated. For increasing co-flow, the intersection

of the 0.05u/Uo contour to the radial location of 1.6r/D occurs farther downstream, moving from

approximately 8.5 to 10.5 diameters downstream. This does not necessarily mean the spreading

rate of the shear layer has decreased, as the location of the virtual origin may also be affected by

- 39 -
varying co-flow ratios. Nonetheless, added co-flow results in a more compact, coherent jet in the

initial region less than ten diameters downstream. The spreading rate of the shear layer and

virtual origin are important jet characteristics for processes dependent on mixing, such as

combustion, because they greatly influence the gas exchange, or entrainment, taking place at the

jet boundary.

4.2.2.2 Mean Transverse Velocity and Entrainment

The shear layer growth characteristics are dependent upon vortex formation and the

pairing process. Consequently, the variation in wave-breaking length and pairing characteristics

seen in the instantaneous schlieren and PIV images has a distinct impact on the mixing of the

present air jet. Figure 22 illustrates the transverse velocity of the jet under all co-flow

conditions. The contours appear noisy because the transverse velocity is a mere fraction (less

than 5% ) of the streamwise exit velocity and closer to the order of magnitude to the resolution of

the velocity measurements. Despite this, trends are still able to be found by comparing the

different cases of mass flow ratio. For the single jet, where wave roll-up occurs almost

immediately, ambient fluid is drawn radially inward at maximum transverse velocities very close

to the jet exit. Likewise, jet fluid near the shear layer is immediately moving radially outward

after exiting the nozzle. The peak magnitudes seen immediately following the exit indicate

entrainment is taking place, corresponding to significant vortex activity. Progressing

downstream, the velocity of ambient fluid decreases, while transverse velocities within the jet

remain at nearly the same magnitude until the shear layer pinches off the end of the potential

core. As increased helium co-flow is added to the system, two trends are observed. First, the

downstream location at which maximum transverse velocities and significant entrainment start to

occur, is noted to shift from about x/D = 1 to x/D = 2. This again corresponds to the increased

- 40 -
wave-breaking length found to occur with the addition of co-flow. The second trend is an

overall decrease in maximum transverse velocities for increased co-flow. This characteristic

likely relates to the suppression of vortex pairing. Liepmann(Liepmann and Gharib 1992) found

that the instantaneous entrainment field, and overall mixing of a jet is highly dependent on the

formation of vortices, particularly streamwise instabilities. He plotted the inward volume flux to

demonstrate the dependence of entrainment on such structures. While the immediate near-field

of the jet under different conditions shows varied behaviors, farther downstream the transverse

velocity fields are similar in appearance. Towards the end of the downstream range considered,

the jet dynamics are no longer dominated by organized vortex structure dynamics, but instead by

fully developed turbulence.

4.2.2.3 Shear Layer Spreading Rate

Hussain and Zedan(Hussain and Zedan 1978) cite that shear layer width, momentum

thickness, and vorticity thickness are all indicators of the relative size of the local shear layer and

they grow linearly in the self-preserving region. As is typical in other studies(Hussain and Zedan

1978; Arakeri, Krothapalli et al. 2003), the shear layer width, B, is used to determine the local

thickness of the shear layer, given by the relation B = R0.95 – R0.10, where R0.95 and R0.10 are the

radial locations where the velocity is 95% and 10% of the local centerline velocity. The linear

region of the shear layer width is plotted in Figure 23 up to the approximate end of the potential

core. For each case, a least squares linear regression was applied and all were found to have a

high correlation, with the worst case having a squared correlation coefficient value of 0.9964.

The slope of the regression fit corresponds to dB/dx, or the shear layer growth rate, and was

found to be 0.2008 for the axisymmetric jet with no co-flow. This growth rate falls within the

range of 0.17 to 0.23, found in previous studies for an axisymmetric jet(Hussain and Zedan

- 41 -
1978). The addition of a low-density co-flow results in a reduction of the shear layer spreading

rate as evidenced by the reported values in Table 3. This data is plotted in Figure 24 and shows

an approximately linear trend between mass flow ratios and spreading rates. For the range of

flow rates examined the shear layer spreading rate decreases by approximately 0.0078 per each

0.5% of helium co-flow, although this trend is not hypothesized to continue for much larger

quantities of co-flow. This behavior of decreased spreading correlates with the trend of vortex

pairing suppression for increased co-flow, observed in schlieren and PIV images.

Table 3: Measured Shear Layer Properties

Mass Flow Ratio (%) 0.0 0.5 1.0 1.5 2.0 2.5
2
Correlation Coefficient Squared (R ),
0.9988 0.9964 0.9979 0.9973 0.9983 0.9986
Linear Regression of Shear Layer Width
Spreading Rate (dB/dx) 0.2008 0.1925 0.1857 0.1764 0.1714 0.1611

Virtual Origin (x0/D) 0.086 0.235 0.569 0.672 0.771 0.678

Potential Core Length (Lp /D) 4.00 4.44 4.78 5.17 5.46 5.64

4.2.2.4 Virtual Origin

The virtual origin is the theoretical streamwise location of a point source of momentum

from which the shear layer originates and self-similar properties scale to. The location of the

virtual origin for each case was approximated by the intersection of the jet axis and a linear

extension of shear layer width data and is provided in Table 3. While reported virtual origin

values in literature range from upstream to downstream locations relative to the jet exit plane for

the axisymmetric jet, values are typically small in value. For this jet configuration and initial

conditions, the virtual origin is always found downstream of the jet exit, with the normal

axisymmetric jet very close to the jet exit plane. The introduction of helium has the overall

effect of shifting the origin farther downstream for increased mass flow rates. This contradicts

- 42 -
the homogenous coaxial jet results of Ko and Kwan(Ko and Kwan 1976), who found larger shifts

of the virtual origin downstream for decreased co-flow values (for velocity ratios also less than

one). In addition, Arakeri(Arakeri, Krothapalli et al. 2003) reported a significant upstream shift

in location for an axisymmetric jet with the addition of homogenous microjets at the jet exit. The

apparent uncharacteristic shift of the virtual origin with the addition of co-flow is attributed to

the density variation. The virtual origin location data, in connection with previously presented

results, suggests an initial suppression in the linear growth of Kelvin-Helmholtz instabilities

from the annular flow. The shift in virtual origin and stability characteristics may stem from

reduced amplification rates as a result of a different density weighted vorticity profile directly

following the jet exit.

4.2.2.5 Turbulence Statistics

Jet mixing is also facilitated by the chaotic fluid motion associated with velocity

fluctuations. For the single axisymmetric jet, the peak streamwise and transverse fluctuating

velocity values are seen within the shear layers (top contour plot in Figure 25 & Figure 26) and

initially within the separating boundary layer (Figure 13). Along the centerline of the jet,

maximum values are reached after the merging of the shear layer, after the end of the potential

core, as demonstrated in Figure 27and Figure 28. The application of a density gradient achieved

with a helium co-flow results in delayed behavior and a shift in peak velocity fluctuations from

the shear layer to downstream of the potential core, and more along the jet axis. This contradicts

the behavior typically demonstrated by homogenous coaxial jets. Ko and Kwan(Ko and Kwan

1976) found greater shifts in the maximum turbulence intensity towards the axis of the central jet

for decreasing co-flow velocity due to a more dominant inner jet drawing the annular jet more

- 43 -
towards the axis. In this study, the shrouding gas is low density and acts to suppress vortex

formation, mixing, and entrainment resulting in a more laminar and inviscid development.

Contour plots in Figure 29 present turbulent kinetic energy production data. Instead of a

gradual shift or trend in behavior, there seems to be a complete jump or change in going from

1.5% to 2% co-flow by mass. In cases 1 through 4, maximum amounts of turbulent kinetic

energy production takes place within the shear layer between one and four diameters

downstream. However, at 2.0% and 2.5% co-flow, production within the shear layer seems to

have been retarded, and instead there is high magnitude production centered on the jet axis

beginning at about eight diameters downstream. Although the very initial region of the

axisymmetric jet has been laminarized by helium co-flow, it would appear that the breakdown to

turbulence became a stronger, more turbulent, more chaotic effect. The reason for this dramatic

shift in behavior is not completely clear. Overall peak velocity fluctuations are seen to be

reduced by up to 50% within the first 2 diameters with only 2% helium co-flow by mass, but

they quickly become and remain similar in magnitude up to 6 x/D. At approximately 8-10

diameters downstream of the nozzle exit, levels for the same 2% helium co-flow are about 20%

greater than the single jet (Figure 30 and Figure 31), accounting partially for the increased

turbulent energy production.

4.2.2.6 Mean and Instantaneous Vorticity

The formation of discrete eddies and vortex roll-ups result from the Kelvin-Helmholtz

instability mechanism and concentrations of vorticity. Vorticity is defined as the curl of the

velocity vector field, and in this instance it is noted that the most dominant component is made

up from gradients of the axial velocity in the radial direction, which is also the most significant

in quasi-parallel shear flows (see Figure 2 and Figure 3). Examination of the mean vorticity

- 44 -
field, given in Figure 32, shows thin, high magnitude bands of vorticity near the jet exit, which

broaden and weaken in the axial direction. Comparison of the cases in Figure 32 and Figure 33

suggests that the higher co-flow cases maintain high levels of vorticity over greater distances,

suggesting that diffusion of vorticity is retarded (in agreement with the reduced growth rates).

Because of the relationship between vorticity and velocity, it comes as no surprise that

the mean vorticity profiles (Figure 32) are elongated to match those of velocity (Figure 20). In

order to gain a better understanding of the flow dynamics, one must look to instantaneous data.

From images, it was previously determined that the wave-breaking length is greater in the

presence of helium co-flow. This is reflected in instantaneous vorticity data in Figure 34, which

shows a coherent band of vorticity being pinched off into discrete pockets farther downstream

for the cases of more helium. The high intensity pockets of vorticity correspond to instabilities

in the shear layer that have grown to an unstable state and have rolled up into coherent vortex

structures. The late appearance of these pockets in the presence of a low density co-flow are an

effect of the corresponding delayed vortex formation and retarded diffusion processes. The

earlier appearance of pockets in the absence of a shear layer density gradient, in turn, results in

the quicker transport and diffusion of energy and mass, and breakdown to a turbulent state.

4.3 Hotwire Power Spectra


Power spectra plots show the distribution of energy of a signal over the frequency domain

at a given location and over a discrete period of time. The convection of energy associated with

turbulent fluctuations within a shear layer is greatly facilitated by the vortex structures that form

from the initial instabilities emanating from the Kelvin-Helmholtz instability mechanism.

Consequently, energy peak(s) within a narrow frequency band in a power spectra plot usually

indicate the most dominantly expressed instability frequency. Hot-wire measurements were

- 45 -
made to survey the power/frequency plane and examine how shear layer instabilities are affected

by a low-density co-flow, as a way to quantitatively compliment the PIV data and schlieren flow

visualization studies. Co-flow mass flow ratios were increased in increments of 0.5%, and the

central jet mean velocity was 18m/s. Lower velocities were utilized such that preferred

frequency modes were within the frequency range sampled (10,000 Hz).

Hot-wire probe locations were selected to be within the potential core such that density

fluctuations would not be sensed by the instrument. As a result, both locations yielded a large

amount of noise in the signals, which may have be due to positioning the probe too far from the

shear layer to feel adequate pressure fluctuations from instabilities. Peaks at low frequencies (on

the order of 10 Hz) at both locations are assumed to be a characteristic of the lab and jet facility.

The first probe location was located at x/D = 0.45 and y/D = 0.45 in order to examine the

initial instability mode with varying amounts of helium co-flow. All cases examined (0.0% to

3.0% co-flow) showed a large broadband peak around 3800 Hz in Figure 35. While this is not

the only peak, the others are assumed to be noise, since calculation of the Strouhal number based

on the initial momentum thickness yields a value of approximately 0.017, matching the value

predicted by Michalke(Michalke 1965) and found experimentally by others(Gutmark and Ho

1983). In addition, the amplitude of this peak was reduced in each case of increasing co-flow,

coinciding with stabilization and mode suppression. The first case of co-flow (0.5%) yielded a

peak decrease of about one half decade, but the reduction in amplitude slowed with increasing

co-flow, as the largest amount (3.0%) yielded a peak approximately 1.5 decades smaller than the

single air jet. Hypothesizing from this spectral plot leads to the conjecture that co-flow has a

stabilizing effect on the initial shear layer instabilities, yet it does not alter the frequency of the

most unstable mode. If this is the case, consideration should be made as to the location of

- 46 -
instability formation compared to the distance required for adequate density gradient formation

from the entrainment and diffusion of helium into the shear layer.

Power spectra were also monitored at downstream locations. The power spectrum at x/D

= 2, and y/D = .25 is presented as Figure 36. This location was contained within the potential

cone and again demonstrated large broadband peaks, with a few dismissed noise peaks. At this

location, the single air jet is dominated by a large spike at approximately 950 Hz with a smaller

harmonic peak of nearly twice the frequency, around 1860 Hz. At 1.0% helium co-flow, the

amplitude of the dominant peak diminished in magnitude and shifted to a higher frequency at

1660 Hz. Upon further increase in co-flow, the amplitude of the peak at 1660 Hz diminished,

and the development of and shift to another peak at a frequency of 3800 Hz is observed. This

frequency matches that of what is hypothesized to be the initial instability previous suggested

and indicates pairing has yet to take place at this location for the higher co-flow rates, matching

the observations in photographs of schlieren and PIV for the particular conditions. Besides

inhibiting vortex pairing in the shear layer, the reduced peak amplitudes of the energy spectrum

also suggest that the low-density co-flow has a stabilizing effect on the jet.

- 47 -
CHAPTER 5: CONCLUSIONS

5.1 Conclusions
The present investigation utilized hot-wire anemometry, particle image velocimetry, and

schlieren flow visualization to study the effects of a low density co-flowing gas on an

axisymmetric air jet in a coaxial jet configuration. Five cases of varying amounts of helium co-

flow up to a maximum mass flow ratio of 2.5% were considered and compared to the case of a

single axisymmetric air jet issuing into ambient air. For all cases, the air jet exited a large

contraction ratio nozzle with a top-hat velocity profile and laminar boundary layer, with a mean

velocity of approximately 50m/s and a Reynolds number of 36,000. Data collection took place

in the jet near-field, up to ten diameters downstream, where transition from a laminar to turbulent

state took place.

Shrouding the turbulent air jet with small flowrates of a low-density gas resulted in

stabilization and laminarization the jet. Over the range of conditions examined, the shear layer

spreading rate was seen to decrease linearly to a maximum of approximately 20% resulting in a

potential core length increase of 41%. The creation of such drastic changes in the entrainment

and mixing in the near-field were a result of stark differences in the vortex structure dynamics.

Instantaneous schlieren and PIV images reveal the growth and roll-up of coherent vortices from

initial instabilities occur over a greater streamwise distance. In addition, the pairing process is

greatly inhibited, resulting in a more coherent jet with a longer transition to turbulence. Hotwire

data indicates the possibility that initial instabilities form at the same frequency, and that co-flow

results in a more stable condition. These results provide further insight into the role of density

gradients in vortex behavior, shear layer development, and mixing processes.

- 48 -
5.2 Future Work
The scope of this project consisted of a single coaxial jet configuration: a central air jet

surrounded by an annular helium jet. Flowrates of the helium co-flow were varied to examine

how the presence of a low-density gas within the shear layer of the air jet altered its properties.

The behavior of the jet with helium co-flow has been documented and could now be used to

predict at other flowrates. Nonetheless, it would be useful, as a continuation of this experiment,

to detail the characteristics of the same air jet for differing co-flow densities and velocities in an

attempt to find a more universal set of parameters for which to describe the behavior of this

generic axisymmetric, weak co-flow configuration. Included in this would be to introduce a

high-density co-flow, for which it is hypothesized that the jet would become more unstable and

transition to turbulence quicker, similar to a low-density jet(Sreenivasan, Raghu et al. 1989).

Utilizing both flow visualization such as schlieren, as well as PIV for mean and instantaneous

flow field measurements is a suggested approach.

Another aspect of the current work that would be important to examine further is the

density field. Mapping out the mean density field for all co-flow conditions would provide more

insight into the mechanism behind the observed results. This initial study is important in its

findings about the effects of introducing a weak low-density co-flow, however, it is not complete

in answering the specifics of exactly how or why the observed trends are taking place. It is not

perfectly clear what the density and density weighted vorticity profiles are. It would be useful to

know how far the helium penetrates into the shear layer radially at any downstream position, or

how far the region of pure helium is maintained downstream for the different co-flows? One of

two techniques is suggested to be used to measure density/concentration for this additional study.

- 49 -
One technique, described by Favre-Marinet and Schettini(Favre-Marinet and Schettini 2001),

utilizes a special hot-wire probe apparatus called an aspirating probe, in which the probe is

insensitive to velocity, and density can instead be sampled. A better approach to obtaining

density measurements would be to use a Rayleigh scattering technique. This technique is a non-

intrusive optical method, which can extract velocity, density, and temperature measurements of a

flow, and is based on frequency shifts in light scattered by gas molecules due to the Doppler

effect(Mielke, Elam et al. 2007).

- 50 -
APPENDIX A: FIGURES

Figure 1: Example free shear layer boundaries and evolving velocity profile.

Figure 2: Velocity profiles at 5 downstream locations for an axisymmetric air jet, Re = 36,000.

- 51 -
Figure 3: Vorticity profiles at 5 downstream locations for an axisymmetric air jet, Re = 36,000.

- 52 -
Figure 4: Axisymmetric jet structure.

Figure 5: Evolution of vortex growth in the free shear layer defined by four regions: (1) Boundary layer
separation (2) Exponential growth of flow disturbances (3) Vortex roll-up (4) Vortex pairing and
degradation to turbulent state.

Figure 6: Instantaneous PIV image demonstrating two vortices pairing to create a larger vortex; one equal in
size to the vortex seen on the right hand side of the image.

- 53 -
(a)

(b)

Figure 7: Comparison of (a) an air jet with 3% helium by mass and (b) a helium jet at similar Reynolds
numbers (Re = 2200).

- 54 -
Figure 8: Schlieren photograph of a helium jet exhibiting a side jet.

- 55 -
Figure 9: The scaled contours of the axisymmetric nozzles used in the experiment.

- 56 -
Figure 10: Experimental apparatus setup for PIV.

- 57 -
Figure 11: Cross-sectional view of piping leading up to coaxial nozzle apparatus. *Screens are denoted by
dashed lines.

- 58 -
Figure 12: Velocity profile of the central air jet at 0.63 mm above the nozzle exit plane.

- 59 -
Figure 13: Velocity fluctuations profile of the central air jet at 0.63 mm above the nozzle exit plane.

- 60 -
Figure 14: Setup used for PIV.

- 61 -
(a)

(b)

Figure 15: Sample instantaneous vector fields combined for three camera positions extending 10 diameters
downstream for the following cases: (a) No co-flow (b) 2.5% co-flow

Figure 16: Schlieren setup with the following components: (A) Xenon Arc Lamp (B) Concave Mirror
(C) Coaxial Nozzles (D) Knife Edge (E) Lens (F) High Speed Camera

- 62 -
(a)

(b)

(c)

(d)

(e)

Figure 17: Sample schlieren images extending 13 diameters downstream for an axisymmetric air jet with a
mean velocity of 22m/s and a Reynolds number of 16,000; (a) No co-flow, but 0.5% helium by mass in main

- 63 -
flow for adequate density gradients (b) 0.74% co-flow by mass (c) 1.48% co-flow by mass (d) 2.22% co-flow
by mass (e) 2.96% co-flow by mass

- 64 -
(a)

(b)

(c)

(d)

- 65 -
(e)

(f)

Figure 18: Sample PIV images with modified contrast and brightness extending 10 diameters downstream
for a central air jet with a 50 m/s mean velocity for the following mass flow ratios as defined in Table 2: (a)
0.0% (b) 0.5% (c) 1.0% (d) 1.5% (e) 2.0% (f) 2.5%

- 66 -
(a)

(b)

(c)

(d)

- 67 -
(e)

(f)

(g)

Figure 19: Sample PIV images with modified contrast and brightness extending 10 diameters downstream
for a central air jet with a 17 m/s mean velocity for the following mass flow ratios; (a) 0.0% (b) 0.9% (c)
1.3% (d) 1.9% (e) 2.7% (f) 3.3% (g) 3.8%

- 68 -
Figure 20: Streamwise velocity contours.

- 69 -
Figure 21: Streamwise centerline velocities.

- 70 -
Figure 22: Transverse velocity contours.

- 71 -
Figure 23: Linear region of the shear layer width.

Figure 24: Shear layer spreading rate for varying mass flow ratios.

- 72 -
Figure 25: Streamwise velocity fluctuation contours.

- 73 -
Figure 26: Transverse velocity fluctuation contours.

- 74 -
Figure 27: Centerline transverse velocity fluctuations.

Figure 28: Centerline streamwise velocity fluctuations.

- 75 -
Figure 29: Turbulent kinetic energy production contours.

- 76 -
Figure 30: Peak transverse velocity fluctuations.

Figure 31: Peak streamwise velocity fluctuations.

- 77 -
Figure 32: Vorticity contours

- 78 -
Figure 33: Maximum vorticity per streamwise location.

- 79 -
(a)

(b)

(c)

(d)

Figure 34: Instantaneous vorticity contours extending 7 diameters downstream for the following co-flow
cases (by mass): (a) 0% (b) 0.5% (c) 1.5% (d) 2.5%

- 80 -
Figure 35: Power spectra at x/D=.45, y/D = .45 for different co-flow rates.

- 81 -
Figure 36: Power spectra at x/D = 2, y/D = .25 for different co-flow rates.

- 82 -
- 83 -
REFERENCES
Amielh, M., T. Djeridane, et al. (1996). "Velocity near-field of variable density turbulent jets."
Int. J. Heat Mass Transfer 39(10): 2149-2164.

Antonia, R. A. and Q. Zhao (2001). "Effect of initial conditions on a circular jet." Experiments in
Fluids 31(3): 319-323.

Arakeri, V. H., A. Krothapalli, et al. (2003). "On the use of microjets to suppress turbulence in a
Mach 0.9 axisymmetric jet." Journal of Fluid Mechanics 490: 75-98.

Balsa, T. F. and P. R. Gliebe (1977). "Aerodynamics and Noise of Coaxial Jets." AIAA Journal
15(11): 1550-1558.

Becker, H. A. and T. A. Massaro (1968). "Vortex Evolution in a Round Jet." Journal of Fluid
Mechanics 31: 435-448.

Bell, J. H. and R. D. Mehta (1988). Contraction design for small low-speed wind tunnels, Joint
Institute for Aeronautics and Acoustics.

Boguslawski, L. and C. O. Popiel (1979). "Flow Structure of the Free Round Turbulent Jet in the
Initial Region." Journal of Fluid Mechanics 90(Feb): 531-539.

Bradshaw, P. (1966). "The effect of initial conditions on the development of a free shear layer."
Journal of Fluid Mechanics 26(2): 225-236.

Brown, G. L. and A. Roshko (1974). "On density effects and large structure in turbulent mixing
layers." Journal of Fluid Mechanics Digital Archive 64(04): 775-816.

Champagne, F. H. and I. J. Wygnanski (1971). "An Experimental Investigation of Coaxial


Turbulent Jets." International Journal of Heat and Mass Transfer 14(9): 1445-1464.

Colucci, P. J. (1993). Linear Stability Analysis of Density Stratified Parallel Shear Flows.
Mechanical and Aerospace Engineering, SUNY Buffalo. M.S.

Crow, S. C. and F. H. Champagne (1971). "Orderly Structure in Jet Turbulence." Journal of


Fluid Mechanics 48(Aug16): 547-591.

Dahm, W. J. A., C. E. Frieler, et al. (1992). "Vortex Structure and Dynamics in the near-Field of
a Coaxial Jet." Journal of Fluid Mechanics 241: 371-402.

Day, M. J., W. C. Reynolds, et al. (1998). "The structure of the compressible reacting mixing
layer: Insights from linear stability analysis." Phys. Fluids 10(4): 993-1007.

Dimotakis, P. E., R. C. Miakelye, et al. (1983). "Structure and Dynamics of Round Turbulent
Jets." Physics of Fluids 26(11): 3185-3192.

- xi -
Drazin, P. G. and W. H. Reid (1981). Hydrodynamic stability. Cambridge [Cambridgeshire] ;
New York, Cambridge University Press.

Elliott, G. S., M. Samimy, et al. (1995). "The characteristics and evolution of large-scale
structures in compressible mixing layers." Physics of Fluids 7(4): 864.

Falcone, A. M. and J. C. Cataldo (2003). "Entrainment Velocity in an Axisymmetric Turbulent


Jet." J. Fluids Eng. 125(4): 620-627.

Favre-Marinet, M., E. B. Camano, et al. (1999). "Near-field of coaxial jets with large density
differences." Experiments in Fluids 26(1-2): 97-106.

Favre-Marinet, M. and E. B. C. Schettini (2001). "The density field of coaxial jets with large
velocity ratio and large density differences." International Journal of Heat and Mass Transfer
44(10): 1913-1924.

Forliti, D. J., B. A. Tang, et al. (2005). "An experimental investigation of planar countercurrent
turbulent shear layers." Journal of Fluid Mechanics 530: 241-264.

Forstall, W. and A. Shapiro (1950). "Momentum and mass transfer in coaxial gas jets." Trans.
A.S.M.E. J. appl. Mech. 10: 399-408.

Freund, J. B., S. K. Lele, et al. (2000). "Compressibility effects in a turbulent annular mixing
layer. Part 1. Turbulence and growth rate." J. Fluid Mech. 421: 229-267.

Freymuth, P. (1966). "On Transition in a Separated Laminar Boundary Layer." Journal of Fluid
Mechanics 25: 683-704.

Furi, M., P. Papas, et al. (2002). "The effect of flame position on the Kelvin-Helmholtz
instability in non-premixed jet flames." Proceedings of the Combustion Institute 29: 1653-1661.

Gladnick, P. G., A. C. Enotiadis, et al. (1990). "Near-Field Characteristics of a Turbulent


Coflowing Jet." AIAA Journal 28(8): 1405-1414.

Gutmark, E. and C. M. Ho (1983). "Preferred Modes and the Spreading Rates of Jets." Physics of
Fluids 26(10): 2932-2938.

Gutmark, E. J., K. C. Schadow, et al. (1995). "Mixing Enhancement in Supersonic Free Shear
Flows." Annu. Rev. Fluid Mech. 27: 375-417.

Han, D. and M. G. Mungal (2001). "Direct Measurement of Entrainment in


Reacting/Nonreacting Turbulent Jets." Combustion and Flame 124: 370-386.

Hart, D. P. (1998). The elimination of correlation errors in PIV processing. 9th Int. Symp. on
Applications of Laser Techniques to Fluid Mechanics. Lisbon, Portugal.

- xii -
Heeg, R. S., D. Dijkstra, et al. (1999). "The stability of Falkner-Skan flows with several
inflection
points." Math. Phys. 50: 82-93.

Hinze, J. O. (1959). Turbulence: An Introduction to Its Mechanism and Theory. New York,
McGraw-Hill Book Company, Inc.

Ho, C. and P. Huerre (1984). "Perturbed Free Shear Layers." Ann. Rev. Fluid Mech. 16: 365-
424.

Ho, C. M. and L. S. Huang (1982). "Subharmonics and Vortex Merging in Mixing Layers."
Journal of Fluid Mechanics 119(Jun): 443-473.

Huerre, P. and P. A. Monkewitz (1985). "Absolute and Convective Instabilities in Free Shear
Layers." Journal of Fluid Mechanics 159(Oct): 151-168.

Hussain, A. K. M. F. and M. F. Zedan (1978). "Effects of Initial Condition on Axisymmetric


Free Shear-Layer - Effects of Initial Momentum Thickness." Physics of Fluids 21(7): 1100-1112.

Kahler, C. J., B. Sammler, et al. (2002). "Generation and control of tracer particles for optical
flow investigations in air." Experiments in Fluids 33: 736-742.

Keane, R. and R. Adrian (1990). "Optimization of particle image velocimeters. Part1: Double
pulsed systems." Meas. Sci. Technol. 1: 1202-1215.

Kedia, K. S. and J. Kurian (2005). Supersonic Freejets from Shaped Nozzles. 32nd National
Conference on Fluid Mechanics & Fluid Power. Osmanabad, Maharashtra.

Ko, N. W. M. and H. Au (1985). "Coaxial Jets of Different Mean Velocity Ratios." Journal of
Sound and Vibration 100(2): 211-232.

Ko, N. W. M. and P. O. A. L. Davies (1971). "Near Field within Potential Cone of Subsonic
Cold Jets." Journal of Fluid Mechanics 50(Nov15): 49-&.

Ko, N. W. M. and A. S. H. Kwan (1976). "Initial Region of Subsonic Coaxial Jets." Journal of
Fluid Mechanics 73(Jan27): 305-332.

Krothapalli, A., L. Venkatakrishnan, et al. (2003). "Turbulence and noise suppression of a high-
speed jet by water injection." Journal of Fluid Mechanics 491: 131-159.

Kwan, A. S. H. and N. W. M. Ko (1977). "Initial Region of Subsonic Coaxial Jets .2." Journal of
Fluid Mechanics 82(Sep7): 273-287.

Kyle, D. M. and K. R. Sreenivasan (1993). "The Instability and Breakdown of a Round Variable-
Density Jet." Journal of Fluid Mechanics 249: 619-664.

- xiii -
Laufer, J., Schlinker, R, Kaplan, R.E. (1976). "Experiments on supersonic jet noise." AIAA
Journal 14(4): 489-497.

Liepmann, D. and M. Gharib (1992). "The Role of Streamwise Vorticity in the near-Field
Entrainment of Round Jets." Journal of Fluid Mechanics 245: 643-668.

Lighthill, M. J. (1952). "On Sound Generated Aerodynamically. I. General Theory." Proc. R.


Soc. Lond. A 211(1107): 564-587.

Lugt, H. J. (1983). Vortex flow in nature and technology. New York, Wiley.

Mathur, P. and C. Messina (2001). "Praxair CoJet™ Technology – Principles and Actual Results
from Recent Installations." AISE Steel Technology (USA). 78(5): 21-25.

Mehta, R. D. (1991). "Effect of velocity ratio on plane mixing layer development: Influence of
the splitter plate wake." Experiments in Fluids 10: 194-204.

Mehta, R. D. and P. Bradshaw (1979). "Design rules for small low speed wind tunnels." Aero. J.
83: 443-449.

Michalke, A. (1965). "On spatially growing disturbances in an inviscid shear layer." Journal of
Fluid Mechanics 23(3): 521-544.

Mielke, A. F., K. A. Elam, et al. (2007). Development of a Rayleigh Scattering Diagnostic for
Time-Resolved Gas Flow Velocity, Temperature, and Density Measurements in Aerodynamic
Test Facilities. 22nd International Congress on Instrumentation in Aerospace Simulation
Facilities. Pacific Grove, CA.

Monkewitz, P. A., B. Lehmann, et al. (1989). "The Spreading of Self-Excited Hot Jets by Side
Jets." Physics of Fluids a-Fluid Dynamics 1(3): 446-448.

Monkewitz, P. A. and K. D. Sohn (1988). "Absolute Instability in Hot Jets." AIAA Journal
26(8): 911-916.

Novopashin, A. and A. Muriel (2002). "Is the Critical Reynolds Number Universal?" Journal of
Experimental and Theoretical Physics 95(2): 262-265.

Pantano, C. and S. Sarkar (2002). "A study of compressibility effects in the high-speed turbulent
shear layer using direct simulation." J. Fluid Mech. 451: 329-371.

Papamoschou, D. (1991). "Structure of the Compressible Turbulent Shear Layer." AIAA 29(5):
680-.

Papamoschou, D. and A. Roshko (1988). "The compressible turbulent shear layer: an


experimental study." J. Fluid Mech. 197: 453-477.

- xiv -
Pavithran, S. and L. G. Redekopp (1989). "The Absolute-Convective Transition in Subsonic
Mixing Layers." Physics of Fluids a-Fluid Dynamics 1(10): 1736-1739.

Rehab, H., E. Villermaux, et al. (1997). "Flow regimes of large-velocity-ratio coaxial jets."
Journal of Fluid Mechanics 345: 357-381.

Rehm, J. E. and N. T. Clemens (1999). "The Large-Scale Turbulent Structure of Nonpremixed


Planar Jet Flames." Combustion and Flame 116: 615-626.

Revuelta, A., L. Sanchez, et al. (2002). "The virtual origin as a first-order correction for the far-
field description of laminar jets." Physics of Fluids 14(6): 1821-1824.

Russ, S. and P. J. Strykowski (1993). "Turbulent structure and entrainment in heated jets: The
effect of initial conditions." Phys. Fluids 5(12): 3216-3225.

Samimy, M., J. H. Kim, et al. (2007). "Active Control of a Mach 0.9 Jet for Noise Mitigation
Using Plasma Actuators." AIAA Journal 45(4): 890-901.

Sandham, N. D. and W. C. Reynolds (1990). "Compressible mixing layer: linear theory and
direct simulation." AIAA 28(4): 618-624.

Sato, H. and O. Okada (1966). "Stability and Transition of an Axisymmetric Wake." Journal of
Fluid Mechanics 26: 237-253.

Savas, O. and S. R. Gollahalli (1986). "Flow Structure in Near-Nozzle Region of Gas Jet
Flames." AIAA Journal 24(7): 1137-1140.

Schaffar, M. (1979). "Direct measurements of the correlation between axial in-jet velocity
fluctuations and far field noise near the axis of a cold jet." Journal of Sound and Vibration 64(1):
73-83.

Schlichting, H. and K. Gersten (2003). Boundary Layer Theory, Springer-Verlag.

Settles, G. S. (2001). Schlieren and Shadowgraph Techniques: Visualizing Phenomena in


Transparent Media, Springer-Verlag.

Sreenivasan, K. R., S. Raghu, et al. (1989). "Absolute Instability in Variable Density Round
Jets." Experiments in Fluids 7(5): 309-317.

Srinivasan, V., M. P. Hallberg, et al. (2010). "Viscous linear stability of axisymmetric low-
density jets: Parameters influencing absolute instability." Physics of Fluids 22.

Strykowski, P. J. and D. L. Niccum (1991). "The stability of countercurrent mixing layers in


circular jets." Journal of Fluid Mechanics 227: 309-343.

- xv -
Strykowski, P. J. and D. L. Niccum (1992). "The Influence of Velocity and Density Ratio on the
Dynamics of Spatially Developing Mixing Layers." Physics of Fluids a-Fluid Dynamics 4(4):
770-781.

Strykowski, P. J. and R. K. Wilcoxon (1993). "Mixing Enhancement Due to Global Oscillations


in Jets with Annular Counterflow." AIAA Journal 31(3): 564-570.

Tam, C. K. W., K. Viswanathan, et al. (2008). "The sources of jet noise: experimental evidence."
Journal of Fluid Mechanics 615: 253-292.

Thwaites, B. (1960). Imcompressible aerodynamics: an account of the theory and observation of


the steady flow of incompressible fluid past aerofoils, wings, and other bodies, Oxford,
Clarendon Press.

Trouve, A., S. M. Candel, et al. (1988). Linear Stability of the Inlet Jet in a Ramjet Dump
Combustor. AIAA 26th Aerospace Sciences Meeting. Reno, NV.

Warda, H. A., S. Z. Kassab, et al. (1999). "An experimental investigation of the near-field region
of free turbulent round central and annular jets." Flow Measurement and Instrumentation 10(1):
1-14.

Williams, T. J., M. R. M. Ali, et al. (1969). "Noise and Flow Characteristics of Coaxial Jets."
Journal of Mechanical Engineering Science 11(2): 133-142.

Winant, C. D. and F. K. Browand (1974). "Vortex Pairing - Mechanism of Turbulen Mixing


Layer Growth at Moderate Reynolds-Number." Journal of Fluid Mechanics 63: 237-255.

Xu, G. and R. A. Antonia (2002). "Effect of different initial conditions on a turbulent round free
jet." Experiments in Fluids 33(5): 677-683.

Yule, A. J. (1978). "Large-scale structure in the mixing layer of a round jet." Journal of Fluid
Mechanics 89(3): 413-432.

- xvi -

You might also like