XUE JIANQING

136
UNIVERSITY OF CINCINNATI Date:___________________ I, _________________________________________________________, hereby submit this work as part of the requirements for the degree of: in: It is entitled: This work and its defense approved by: Chair: _______________________________ _______________________________ _______________________________ _______________________________ _______________________________

Transcript of XUE JIANQING

Page 1: XUE JIANQING

UNIVERSITY OF CINCINNATI Date:___________________

I, _________________________________________________________, hereby submit this work as part of the requirements for the degree of:

in:

It is entitled:

This work and its defense approved by:

Chair: _______________________________ _______________________________ _______________________________ _______________________________ _______________________________

Page 2: XUE JIANQING

COMPUTATIONAL SIMULATION OF FLOW INSIDE PRESSURE-SWIRL ATOMIZERS

A dissertation submitted to the

Division of Research and Advanced Studies of the University of Cincinnati

in partial fulfillment of the requirements for the degree of

DOCTOR OF PHILOSOPHY (Ph. D.)

in the Department of Mechanical, Industrial and Nuclear Engineering of the College of Engineering

2004

by

Jianqing Xue

B.S., Tianjin University, 1989 M.S., Tianjin University, 1992

Dissertation Committee:

Dr. Milind A. Jog, Chair Dr. San-Mou Jeng Dr. Raj M. Manglik Dr. Rupak K. Banerjee

Page 3: XUE JIANQING

ii

ABSTRACT

Simplex atomizers (pressure-swirl atomizers) are widely used in air-breathing gas

turbine engines as they have good atomization characteristics and are relatively simple

and inexpensive to manufacture. To reduce emissions, it is critical to design fuel

atomizers that can produce spray with a predetermined droplet size distribution at the

desired combustor location (small mean droplet diameters and uniform local air/fuel

ratios). Manufacturing methods are now available where complex atomizer geometries

can be easily obtained. However to use such methods, the influence of atomizer

geometry on its performance must be well understood.

In this dissertation, a two-dimensional axi-symmetric computational fluid

dynamics (CFD) model based on the Arbitrary-Lagrangian-Eulerian (ALE) method to

predict the flow in pressure-swirl atomizers was developed. The Arbitrary-Lagrangian-

Eulerian method was applied so that the free interface between gas and liquid could be

tracked sharply and accurately. The developed code was validated by comparison of

predictions with experimental data for large scale prototype and with semi-empirical

correlations at small scale. The computational predictions agreed well with experimental

data for the film thickness at the exit, spray cone angle, and the pressure drop across the

atomizer as well as velocity field in the swirl chamber.

Using the validated code, a comprehensive parametric study on simplex atomizer

performance was conducted. The geometric parameters of atomizer covered in this study

include: atomizer constant (K), the ratio of length to diameter in swirl chamber (Ls/Ds),

the ratio of length to diameter in orifice (lo/do), the swirl chamber to orifice diameter ratio

(Ds/do), inlet slot angle (β), trumpet angle (θt), trumpet length (lt), and swirl chamber

Page 4: XUE JIANQING

iii

convergent angle (θc). The effects of these geometric parameters on the atomizer

performance were studied for a fixed mass flow rate through the atomizer as well as for a

fixed pressure drop across the atomizer. The atomizer performance was described in

term of dimensionless film thickness at the exit (t*), discharge coefficient (Cd) and spray

cone half angle (θ).

To address applications in pharmaceutical and food processing industry, flow of

non-Newtonian power-law fluids through pressure-swirl atomizers was considered.

Detailed flow patterns inside the atomizer for shear-thinning, Newtonian and shear-

thickening fluids were investigated. A range of power law index from 0.7 to 1.3 was

considered. With a fixed flow rate through the atomizer, the shear-thickening fluids

exhibit higher film thickness at exit, lower spray angle, and higher discharge coefficient

compared to Newtonian fluids. For the range of power law index considered in this study,

the atomizer performance parameters for shear-thinning fluids show small change from

Newtonian fluids. The variation of atomizer performance with the atomizer constant was

delineated for different power-law index.

Page 5: XUE JIANQING

iv

Page 6: XUE JIANQING

v

ACKNOWLEDGEMENTS

I am greatly indebted to my advisor, Professor Milind Jog, who not only served as

my supervisor but also encouraged and challenged me throughout my academic program.

His technical and editorial advice was essential to the completion of this dissertation. I

have learned quite a lot from his extensive knowledge and many brilliant ideas. His all-

aspect help has made the past years a good memory in my life.

I would deeply thank my co-advisor, Professor San-mou Jeng, for hosting me in

his group as a PhD student. His supervision and continuous supports laid a smooth way

for my research work. Through rewarding discussions with him, I have learnt much more

knowledge in CFD and atomization. I thank, with the highest respect, Professor Raj

Manglik and Professor Rupak Banerjee for serving on my dissertation committee and for

being there whenever I needed their help.

I am especially grateful to Dr. Erlendur Steighthorsson for his help on automatic

grid generator creating and much valuable advice on CFD code developing. I learnt quite

much CFD knowledge from him. Special thanks go to Dr. Michael Benjamin and Dr.

Adel Mansour, who gave me much encouragement and help during my three-month co-

op work time in Gas Turbine Spray System department, Parker Hannifin Corporation in

1999. I am greatly thankful to Dr. Dexin Wang, one of my best friends, who not only

gave me much advice on my academic work but helped me much in my life too. I am

deeply grateful to Mr. Tolga Sakman, who gave me much help on the ALE method

understanding at the beginning days I was involved in this project. My sincere thanks

should also go to Dr. Zhanhua Ma, Dr. Jun Cai, Dr. Juntao Zhang and Mr. Ibrahim

Page 7: XUE JIANQING

vi

Ashraf for their technical suggestion on my research work. Discussion with them for

technical issue was very enjoyable.

I would like to thank my wife, Ms. Li Jia, for her understanding and love during

these years. Her support and encouragement was in the end what made this dissertation

possible. My parents receive my deepest gratitude for their ever-loving supports and

understanding during the years of my study and work.

At last, the financial support of Parker Hannifin Corporation and the National

Science Foundation is greatly appreciated.

Page 8: XUE JIANQING

1

TABLE OF CONTENTS

ABSTRACT ..............................................................................................ii

ACKNOWLEDGEMENTS .....................................................................v

TABLE OF CONTENTS .........................................................................1

LIST OF TABLES ....................................................................................4

LIST OF FIGURES ..................................................................................5

NOMENCLATURE ..................................................................................8

CHAPTER 1 INTRODUCTION ……………………………….…......11

1.1 Motivation …………………………………………………………………….....11

1.2 Literature Review ………………………………………….………………........14

1.3 Scope of Dissertation ………………………………………………………........18

CHAPTER 2 COMPUTATIONAL MODEL ………………………....21

2.1 Review of Computational Methods ....................................................................21

2.2 Mathematical Formulation …………………………………………………......23

2.3 Stress Calculation ..............................................................................................24

2.4 Discretization ....................................................................................................26

2.5 Pressure Field Iteration ...........................................................................................28

2.6 Boundary Conditions ………………………………………………………........32

Page 9: XUE JIANQING

2

2.7 Grid Generator ..................................................................................................35

2.8 Code Structure ..................................................................................................36

CHAPTER 3 CODE VALIDATION ……………………………….....37

3.1 Verification of Grid Independence of Results ……………………………........37

3.2 Comparison with Experimental Data and Empirical Correlations .......................37

3.3 Comparison with Correlations at Small Scale ……………………………….....43

3.4 Summary ………………………………………………………………….......49

CHAPTER 4 EFFECTS OF GEOMETRIC PARAMETERS ON

ATOMIZER PERFORMANCE (PART I) ………………………….....50

4.1 Introduction ……………………………………………………………….........50

4.2 Results for Constant Mass Flow Rate through the Atomizer ….……………….52

4.3 Results for Constant Pressure Drop across the Atomizer ………........................61

4.4 Summary …………………………………………………………………….......70

CHAPTER 5 EFFECTS OF GEOMETRIC PARAMETERS ON

ATOMIZER PERFORMANCE (PART II) ……………………….....71

5.1 Introduction ......................................................................................................71

5.2 Effect of Variation in Inlet Slot Angle β ............................................................74

5.3 Effect of Variation in Trumpet Angle θt .............................................................80

5.4 Effect of Variation in Dimensionless Trumpet Length (lt/do) ………………..83

Page 10: XUE JIANQING

3

5.5 Effect of Variation in Swirl Chamber Convergence Angle θc ………………...85

5.6 Summary …………………………………………………………………….......88

CHAPTER 6 NON-NEWTONIAN FLUID FLOW IN

PRESSURE-SWIRL ATOMIZERS ………………………………......90

6.1 Introduction …………………………………………………………………......90

6.2 Mathematical Model ………………………………………………………........91

6.3 Influence of Power-law Index on the Performance of Atomizers with

Constant Flow Rate across the Atomizer ………………………..……………........92

6.4 Influence of Atomizer Constant on the Performance of Atomizers for both

Newtonian and non-Newtonian Fluids ………………………….………….........107

6.5 Summary .........................................................................................................111

CHAPTER 7 CONCLUSIONS AND RECOMMENDATIONS …...113

7.1 Conclusions ……………………………………………………………….....113

7.2 Recommendations for Future Work ……………………………………….......115

BIBLIOGRAPHY ………………………………………………..........116

APPENDIX INVISCID ANALYSIS FOR PRESSURE SWIRL

ATOMIZERS ………………………………………………………....124

Page 11: XUE JIANQING

4

LIST OF TABLES

Table 3.1 Results of different grid density …………………………………………......37

Table 3.2 Cases summary ............................................................................................39

Table 3.3 Comparison results of experimental, CFD and correlation data ………….....42

Page 12: XUE JIANQING

5

LIST OF FIGURES

Figure 1.1 Schematic of a simplex atomizer ……………………………………….......13

Figure 2.1 The assignment of the variables about a cell…………………………..........27

Figure 2.2 The schematic of mass and momentum control volume ………………........27

Figure 2.3 Flowchart of the code ………………………………………………….........36

Figure 3.1 Final grids of different grid density……………………………………........38

Figure 3.2 Comparison of swirl velocity variations at three axial locations………........40

Figure 3.3 Variation of discharge coefficient ……………………………………..........44

Figure 3.4 Variation of dimensionless film thickness …………………………….........46

Figure 3.5 Variation of Spray Angle………………………………………………........48

Figure 4.1 Variation of the dimensionless film thickness at the exit and discharge

coefficient with Ls/Ds under constant flow rate …………………………………..........53

Figure 4.2 Variation of the spray cone half angle and wav/uav with Ls/Ds under

constant flow rate ………………………………………………………………….........54

Figure 4.3 Variation of the dimensionless film thickness at the exit and discharge

coefficient with lo/do under constant flow rate ……………………………………........56

Figure 4.4 Variation of the spray cone half angle and wav/uav with lo/do under

constant flow rate ………………………………………………………………….......57

Figure 4.5 Variation of the dimensionless film thickness at the exit and discharge

coefficient with Ds/do under constant flow rate …………………………………….......59

Figure 4.6 Variation of the spray cone half angle and wav/uav with Ds/do under

constant flow rate …………………………………………………………………........60

Page 13: XUE JIANQING

6

Figure 4.7 Variation of the dimensionless film thickness at the exit and discharge

coefficient with Ls/Ds under constant pressure drop………………………………........62

Figure 4.8 Variation of the spray cone half angle and wav/uav with Ls/Ds under

constant pressure drop …………………………………………………………….........63

Figure 4.9 Variation of the dimensionless film thickness at the exit and discharge

coefficient with lo/do under constant pressure drop ………………………….......65

Figure 4.10 Variation of the spray cone half angle and wav/uav with lo/do under

constant pressure drop ……………………………………………………………........66

Figure 4.11 Variation of the dimensionless film thickness at the exit and discharge

coefficient with Ds/do under constant pressure drop ………………………………........68

Figure 4.12 Variation of the spray cone half angle and wav/uav with Ds/do under

constant pressure drop …………………………………………………………….........69

Figure 5.1 Schematic of a simplex atomizer with trumpet ………………………........72

Figure 5.2 Variation of atomizer performance parameters with inlet slots angle β ........75

Figure 5.3 Comparison of results for lo/do = 1(solid lines) and lo/do =0.5 (dashed

lines) ...........................................................................................................................76

Figure 5.4 Variation of atomizer performance parameters with trumpet angle θt ….......81

Figure 5.5 Variation of atomizer performance parameters with ratio of trumpet

length to orifice diameter lt/do ……………………………………………………........84

Figure 5.6 Variation of atomizer performance parameters with convergence angle

θc …………………………………………………………………………………........86

Figure 5.7 Streamlines for flow inside the atomizer for different θc. Inlet slot

angle β=900 ......................................................................................................................87

Page 14: XUE JIANQING

7

Figure 6.1 Initial grid ………………………………………………………………......93

Figure 6.2 Final grid ....................................................................................................94

Figure 6.3 Streamline and pressure distribution…………………………………….......95

Figure 6.4 Variation of film thickness with power index ………………………..........96

Figure 6.5 Variation of discharge coefficient with power index …………………......97

Figure 6.6 Variation of Spray half angle with power index ……………………..….....98

Figure 6.7 Locations for the velocity inside nozzle study ............................................99

Figure 6.8 Axial velocity profiles at the swirl chamber ………………………….......100

Figure 6.9 Axial velocity profiles at the orifice ……………………………................101

Figure 6.10 Axial velocity profiles at the orifice ..........................................................102

Figure 6.11 Swirl velocity profiles at the swirl chamber ..............................................103

Figure 6.12 Swirl velocity profiles at the orifice ..........................................................104

Figure 6.13 Swirl velocity profiles at the exit …………………………………….......105

Figure 6.14 Variation of film thickness with atomizer constant ………………….......108

Figure 6.15 Variation of discharge coefficient with atomizer constant.........................109

Figure 6.16 Variation of spray cone angle with atomizer constant ...............................110

Page 15: XUE JIANQING

8

NOMENCLATURE

Aa air core area at orifice exit

Ao orifice area

Ap total swirl slot area

At the trumpet end area

Ata air core area at the trumpet end

Cd discharge coefficient, ( ) 5.02 ρPA

m

o ∆

&

Ds spin chamber diameter

do orifice diameter

dt trumpet diameter at atomizer exit

fr

body force

K atomizer constant, Ap/(Dsdo)

K1 Ap/(πrors )

Kt Ap/(πrtrs)

Ls spin chamber length

lo orifice length

lt trumpet length

p static pressure

Q volume flow rate

ri the radial distance from axis to inlet slot

ro orifice radius, do/2

Page 16: XUE JIANQING

9

rs spin chamber radius, Ds/2

rt dt/2

S(t) surface enclosing control volume V(t)

t film thickness at exit

t* dimensionless film thickness, t/(do/2)

U average total velocity at the end of orifice

u axial velocity component

Ur

arbitrary velocity vector for the control volume V(t)

ur velocity vector

ou average axial velocity at the end of orifice

uoa axial velocity at the liquid-air interface at the end of orifice

tu average axial velocity at the end of trumpet

uta axial velocity at the liquid-air interface at the end of trumpet

V(t) control volume

w tangential velocity component

wi average tangential velocity at the inlet

ow average tangential velocity at the end of orifice

woa tangential velocity at the liquid-air interface at the end of orifice

tw average tangential velocity at the end of trumpet

wta tangential velocity at the liquid-air interface at the end of trumpet

X Aa/Ao

Xt Ata/At

∆p pressure drop across the nozzle

Page 17: XUE JIANQING

10

β inlet slot angle

θ spray cone half angle

θc spin chamber convergence angle

θt trumpet angle

θ` difference between spray cone half angle and trumpet angle, θ -θt

ρ density

τ viscous stress

Subscript

i parameter at the inlet

o parameter at the end of orifice

oa parameter at the liquid-air interface at the end of orifice

t parameter at the end of trumpet

ta parameter at the liquid-air interface at the end of trumpet

Page 18: XUE JIANQING

11

CHAPTER 1 INTRODUCTION

1.1 Motivation

Liquid atomization, the process of producing a large number of droplets from

bulk liquid, is used in a variety of engineering applications, in pharmaceutical industries,

process industries, fuel injection in combustion applications, and in agricultural sprays,

among others. A number of spray devices have been developed for this purpose, and

they are generally designated as atomizers or nozzles (Lefebvre, 1989). Among these,

pressure-swirl atomizers or simplex atomizers are commonly used for liquid atomization

due to their simple design, ease of manufacture, and good atomization characteristics.

The applications such as fuel injection systems of the gas turbines engines, crop spraying

in agriculture, the production of powders by spray drying, water sprays for fire

suppression, oil spray for combustors in furnaces and power stations, involve Newtonian

fluids. On the other hand, applications such as atomization of aqueous polymeric

solutions in manufacture of pharmaceutical products, paint sprays, spray drying of food

and detergents, sprays in healthcare products, involve fluids that exhibit non-Newtonian

flow behavior. Figure 1 shows a schematic of simplex atomizer geometry. In a simplex

atomizer, the liquid is forced under high pressure to enter a swirl chamber through inlet

slots at the outer wall. An air core is formed along the centerline due to high swirl

velocity of the fuel. The liquid exits the atomizer through a small orifice with even higher

swirl velocity that forces the liquid to disperse radially outwards to form a hollow-cone.

The thin liquid sheet then becomes unstable and breaks up to form a spray of droplets.

Page 19: XUE JIANQING

12

Owing to the numerous applications of pressure-swirl atomizers, a large number

of studies are available in published literature. Focus of earlier studies has been on the

phenomena taking place after the liquid exits the atomizers, break-up of simple forms of

bulk fluids, on the statistical nature of fuel sprays and on the development of correlations

and operating parameters relevant to spray characteristics. Internal flows in atomizers are

of interest due to their potential effect on the atomization process that takes place external

to the atomizers. Numerous experimental studies have shown that small changes in

geometric features internal to the atomizers can greatly impact the characteristics of the

spray that is produced. However, due to the difficulties in measurements inside a small

scale atomizer and the challenges in modeling of two-phase flow, the flow phenomena

inside the atomizer have not received much attention. The flow inside the atomizer

involves both air and fuel and the interface between the gas and liquid is not known a

priori and must be determined as part of the solution. Furthermore, the flow is turbulent

and contains regions of recirculating flow. Recent advances in accurate tracking of

gas/liquid interface are used in this thesis to simulate the flow in the atomizer. We have

used the Arbitrary-Lagrangian-Eulerian method to numerically simulate the two-phase

flow inside the atomizer and have determined the characteristics of the liquid sheet

emanating from the atomizer. Other students in our research group (Wang, 1999, and

Ma, 2001) have used large-scale prototype atomizer to measure the velocity field inside

the atomizer, film thickness at exit, and the spray cone angle. The computational code

was validated by using their data. The developed computational code was used to

conduct a comprehensive parametric study of the effect of atomizer geometrical

parameters on its performance under both the constant flow rate and constant pressure

Page 20: XUE JIANQING

13

drop conditions for Newtonian and non-Newtonian (power-law) fluids. This is first such

computational study that computationally investigates flow inside simplex atomizers in

detail. The study provides useful insights in the flow phenomena inside an atomizer and

the results provide guidance for simplex atomizer designers.

A A

Ls Ds

do

x, Ux

View A-A

dp

θ, Uθ

r, Ur

Air-coredvortex

Thin, hollow, conical liquid sheet

Inlet Inlet

Lp

swirlchambersection

exitorificesection

swirler section

Θ

lo

Figure 1 Schematic of a simplex atomizer

Page 21: XUE JIANQING

14

1.2 Literature Review

Early studies of simplex atomizers employed analytical and/or experimental

methods to measure and/or predict the external spray characteristics (Rizk and Lefebvre

1983, 1985 and 1986; Suyari and Lefebvre 1986; Dodge and Biaglow 1986; Dodge,

Rhodes and Reitz 1987; Dodge and Schwalb 1989; McDonell and Samuelsen et al 1992;

Yule and Widger 1996; Vanderwege and Hochgreb 1998; Okamoto et al 1998).

Rizk and Lefebvre (1985 and 1986) investigated the internal flow characteristics

of simplex swirl atomizers using approximate analytical treatment of the flow. They

investigated the effects of variation of the individual swirl atomizer geometrical

dimensions on the thickness of annular liquid film at the nozzle exit and the effects of the

variation of the fluid properties on the values of the discharge coefficient, the spray angle

and the liquid film thickness. They developed a general expression for the liquid film

thickness at the exit of the swirl atomizer and stated that the air-core diameter increases

with increasing pressure, decreasing inlet area, increasing swirl chamber diameter,

decreasing swirl chamber length, increasing orifice length, decreasing liquid viscosity

and decreasing liquid density.

Horvay and Leuckel (1985 and 1986) studied the velocity profiles within a

pressure swirl atomizer. The experiment were conducted using three different

convergence configurations (standard, concave, and plain conical) and two different

inlet/swirl chamber configurations (four 20 x 10 mm and four 20 x 5 mm rectangular

inlet slots). The atomizers were manufactured from Plexiglas and have same overall

dimensions: radius of swirl chamber rs = 50 mm, length of swirl chamber ls = 25 mm,

Page 22: XUE JIANQING

15

radius of orifice ro = 10 mm and length of orifice lo = 20 mm. The measurement of the

liquid velocity components within the atomizer were carried out using LDA and a

refractive index matching fluid, which is a mixture of tetraline, turpentine and castor oil.

The seeding particles are small air bubbles. Radial profiles of the axial and tangential

velocities were taken at six different cross-sections through the atomizer.

De Keukelaere (1995) investigated the pressure drop and diameter of air core

from inlet to outlet of swirl atomizer. He found that the air core was almost fully formed

at a pressure of 15 kPa and observed the fluctuations of the air core. This confirmed

Hsieh and Rajamani’s observation.

Herpfer et al (1996 and 1997) developed a nonintrusive instrumentation called

Streaked Particle Imaging Velocimetry and Sizing (SPIVS) probe and used this probe to

measure the size and velocity of droplets in the spray of a Delavan WDB 45 solid-cone

simplex swirl atomizer. They compared the results of SPIVS with that of PDPA and

concluded that the SPIVS diagnostic technique has the ability to accurately and reliably

characterize the liquid droplet properties from a spray even though the SPIVS system

exhibits a tendency towards sampling bias in favor of the larger drop sizes within a spray.

Holtzclaw et al (1997 and 1998) examined the geometrical effects on the internal

flow field in a large-scale simplex fuel nozzle. They measured the tangential and radial

velocity components using PIV techniques and found that the radial velocity was

significantly less than the swirl velocity at any point within the swirl chamber of a

simplex nozzle. They also derived an empirical equation based on the measured swirl

velocity component. Due to the limitations of the imaging technique and post-processing

Page 23: XUE JIANQING

16

software of image analysis, Holtzclaw’s PIV measurements are not very accurate but

provide qualitative results.

Benjamin et al (1998) investigated the effects of various geometric and flow

parameters on the performance of large-scale pressure swirl atomizers using optical

methods. They measured the film thickness, droplet size and spray angle of a series of

large-scale pressure-swirl atomizers. After testing and analyzing a large number of

geometric variations covering a wide range of flow capacities, they developed some

correlations on the discharge coefficient, flow number, velocity coefficient, spray angle

and Sauter mean diameter based on their experimental data.

Ma (2001) studied the internal flow characteristics in the swirl chamber for both

large-scale and medium-scale pressure swirl atomizers. The internal flow field was

measured using a two-color PIV system and refractive index matching fluids method.

The spray cone angle and liquid film thickness of the nozzle were also obtained from the

PIV image. The measurements of the droplet size and velocity distribution were carried

out using PDPA. According the experimental data, he gave the relationship between the

internal flow field and the external spray characteristics. A non-dimensional correlation

between the vortex flow pattern in the swirl chamber and atomizer design variables was

presented. A discussion on flow regime (turbulent or laminar) within the atomizer body

was given.

Apart from the experimental measurement, the approximate analytical methods

are also used to study the performance of the pressure swirl atomizers. Giffen and

Muraszew (1953) carried out an inviscid analysis of the flow in a simplex atomizer.

They showed that the atomizer constant, which is the ratio of inlet area to the product of

Page 24: XUE JIANQING

17

swirl chamber diameter and exit orifice diameter, is the most important geometric

parameter of the atomizer. They derived expressions for liquid film thickness, spray cone

angle, and the discharge coefficient in terms of the atomizer constant. However, in a

practical atomizer, the flow is viscous, and other geometric parameters also influence the

atomizer performance.

Som and Mukherjee (1980) conducted a theoretical investigation on the discharge

coefficient and spray cone angle of a swirl spray nozzle. These two parameters were

theoretically evaluated through the analytical solution of hydrodynamics of flow inside

the nozzle. They found among the nozzle geometries, an increase in the orifice to swirl

chamber diameter ratio, swirl chamber angle or decrease in swirl chamber length to

diameter ratio decreases the discharge coefficient and increases the spray cone angle.

Yule and Chinn (1997) conducted a numerical study by treating the entire

computational domain as single phase, and then guessing the interface by joining grid

points where pressure is found to be atmospheric. Solution was re-calculated by creating

a new grid using the calculated interface and treating the interface as a “with-slip”

boundary. However, the condition of normal stress balance was not applied at the

interface. The velocity and pressure distribution in atomizers were calculated and

discharge coefficient and spray angle were predicted.

Sakman et al (1998) studied the effects of simplex nozzle geometry on its

performance numerically. They concluded that when designing a nozzle, particular

attention should be paid to optimum values of the performance variables like the local

maximum of spray cone half angle at Ds / Do = 4.1 and extreme caution should be taken

when designing a simplex nozzle. The performance variables are dependent on the actual

Page 25: XUE JIANQING

18

physical dimensions, since these may determine whether the flow is laminar or turbulent,

and may change the trends of the performance parameters.

Steinthorsson and Lee (2000) conducted three-dimensional simulations of

internal, free-surface flow in a pressure-swirl atomizer via commercial software

FLUENT. The atomizer used in the simulations is a large-scale atomizer. The Volume

of Fluid (VOF) method was adopted to capture the formation of the air-core and

Reynolds Stress Turbulence model was used to model the effects of turbulence. The

results were compared to experimental data given by Wang et al (1999, 2000). They

concluded that the effect of the discrete inlet slots disappears before the liquid enters the

orifice.

Hansen (2001) simulated the flow in a scaled model of a Danfoss pressure-swirl

atomizer via commercially available CFX-4.3 code. The Volume of Fluid (VOF) method

was used to track the liquid-gas interface. The simulations were performed in a three-

dimensional curvilinear grid representing the swirl chamber of the atomizer and managed

to capture the overall flow characteristics of a pressure-swirl atomizer with the

formations of an air-core. A simulation using the k-ε turbulence model over-predicted

viscosities and failed to predict a stable air-core in the atomizer. Results from LES and

simulations assuming laminar flow were verified against experimental findings from

LDA and pressure measurements.

1.3 Scope of the Dissertation

In Chapter 2, the details of the numerical method are given. The flow inside a

simplex atomizer is turbulent, unsteady and contains re-circulating regions. Additionally,

Page 26: XUE JIANQING

19

the shape and the location of the liquid-air interface is not known and must be determined

as part of the solution. This is the primary difficulty in computational modeling of the

flow in a simplex atomizer. To tackle this problem, the Arbitrary-Eulerian-Lagrangian

method is used to track the liquid/gas interface. In this method, the grid points on the

interface always remain on the interface and move with the interface. The advantage of

this approach is that the gas-liquid interface stays sharp at all times. This is especially

important for the flow in a simplex nozzle to accurately determine the film thickness to

evaluate the mean droplet size. In contrast, commercially available general-purpose flow

codes do not have the capability to track a sharp interface.

In Chapter 3, the code validation is presented. Validation of the computational

code is essential before the code can be used to study the flow phenomena inside an

atomizer. As the flow field inside a small scale atomizer is extremely difficult to

measure, a large-scale prototype nozzle was used to measure the detailed flow field for a

number of geometric configurations by Jeng et al. (1998) and Wang (1999). A detailed

comparison with the experimental data was carried out. The comparison showed that for

the cases with four inlet slots the code is able to predict performance parameters very

well (generally within 5% of the experimental measurements). We note that the

differences are less than the uncertainty in the experimental measurements. Grid

independence verification of the results was carried out as well.

In Chapter 4, effects of changes in three non-dimensional geometric parameters

on the atomizer performance are studied. These geometric parameters are - the length to

diameter ratio of the swirl chamber and that of the orifice, and the swirl chamber to

orifice diameter ratio. The variation in the atomizer performance is obtained by keeping

Page 27: XUE JIANQING

20

the pressure drop across the atomizer or inlet mass flow rate constant and is presented in

terms of the dimensionless liquid film thickness at the exit of the orifice, spray cone half

angle and discharge coefficient.

In Chapter 5, effects on the atomizer performance of varying a number of

geometric parameters that are not considered in available correlations are studied. The

geometric parameters are: inlet angle, smoothing radius between exit orifice and trumpet,

trumpet angle, trumpet length, orifice length, length of the spin chamber, and diameter of

the spin chamber.

In Chapter 6, the numerical code for Newtonian flow described in Chapter 2 is

extended to time-independent purely viscous power-law non-Newtonian flow. A

parametric study is conducted under both constant mass flow rate at the inlet and constant

pressure drop across the atomizer. Detailed flow pattern inside the simplex nozzle is

investigated.

In Chapter 7, the conclusions from this study are listed, the limitations of the

present code are addressed, and suggestions for future work are outlined.

Appendix provides a derivation based on inviscid flow assumption for film

thickness, spray cone angle, and discharge coefficient for nozzles that do not have purely

tangential fuel entry and for nozzles with a trumpet.

Page 28: XUE JIANQING

21

CHAPTER 2 COMPUTATIONAL MODEL

2.1 Review of Computational Methods

Due to the high swirl velocity in pressure swirl atomizers, an air core exists at the

centerline of the atomizer. The position and geometry of the gas/liquid interface is not

known a priori, and must be determined as a part of the solution. Therefore, the main

difficulty in a numerical scheme that models the interaction of two fluid phases is to

accurately track the interface. Modeling of motion of a liquid gas interface is relevant to

many engineering applications and a number of methods have been used for this purpose.

The easiest approach is to treat the entire computational domain as single phase, and then

guess the interface by joining grid points where pressure is found to be atmospheric. Yule

and Chinn (1997) and Datta and Som (2000) have used such approach using

commercially available software. This is an approximate method as the interface

conditions of stress balance are not applied at the interface. As the liquid gas interface

behaves vary differently from a single-phase region, such method is not likely to provide

accurate results. The volume-of-fluid (VOF) method is widely used in flow simulations

with multiple phases (Steinthorsson and Lee, 2000). In this method the interface is

represented by the fraction of cell volume occupies by the liquid. As the grid points do

not exactly lay on the interface, the interface has to be reconstructed using volume

fractions for each cell. In general, the accuracy strongly depends on the fineness of the

grid around the interface and the method can be computationally very expensive. The

shape of the liquid gas interface greatly affects its stability and breakup. The pressure,

surface tension, and aerodynamic forces determine the breakup behavior of a liquid gas

Page 29: XUE JIANQING

22

interface. The curvature of the interface can change the pressure variation along the

interface, surface tension forces, and the aerodynamics forces. As such, for the

application being considered here, accurately tracking a sharp interface is very important.

The VOF method does not provide tracking of the sharp interface and as such may not be

the best method for this application. The gradient method does not define the exact

location of the interface but represents the interface as a continuous gradient over several

cells. Although this method is computationally efficient, the accurate determination of

the location is not possible. The most promising approach for such problems, is a fully

Lagrangian scheme (Hirt et al, 1970). In this scheme each node point moves with its

velocity so that the same cell mass is followed in time. Therefore, in a Lagrangian

approach, node points on a two-phase interface always remain on the interface and the

motion and shape of the interface can be evaluated very accurately. Such method works

very well only for well-behaved flows. For domains that include shear, fluid separation,

and recirculation, such as the problem at hand, a Lagrangian scheme would lead to a

highly distorted grid very quickly. When differential operators are calculated on a highly

distorted mesh, the approximations generally lose accuracy. Furthermore, obtaining

converged solutions for a complex flow on a highly distorted grid would be virtually

impossible. To achieve the Lagrangian accurate tracking of the interface and yet

maintain accuracy and obtain convergent solutions, The Arbitrary-Lagrangian-

Eulerian(ALE) method (Hirt et al, 1974) was adopted in this study. The ALE method

comprises of two computational steps that combine the Lagrangian tracking with

Eulerian re-gridding. In the first step, the computational grid vertices move with the

same velocity as the fluid. Therefore there is no mass exchange among computational

Page 30: XUE JIANQING

23

cells. The continuity and momentum conservation equations are solved using a predictor-

corrector numerical scheme. This is the Lagrangian computational step. In the second

step, a new adaptive grid is generated. The mass and momentum for each new cell is

calculated based on the motion of new grid vertices from their Lagrangian positions.

This is the Eulerian step of the calculation. The Lagrangian step ensures that the points

on the interface remain on the interface and thus the interface is tracked accurately. The

Eulerian step ensures that the grid does not get distorted thereby maintaining solution

accuracy. Among the available methods for interface tracking, the ALE method provides

the most accurate tracking of the interface without numerical smearing.

2.2 Mathematical Formulation

The conservation equations governing the flow are,

Continuity Equation 0. =∇ ur (2.1 a)

Momentum Conservation fPDt

uD rr

ρτρ +∇+−∇= . (2.1 b)

Where ur is velocity, τ is viscous stress term, fr

is a body force, and ρ is

density. Integrating Eqns. 2.1 over a control volume V(t) with a surface S(t) that moves

with an arbitrary velocity, the equations become,

∫∫∫∫∫∫∫∫∫∫

∫∫

=−∇+−−

=−

)()()()(

)(

0.).(

0).(

tAtVtAtV

tA

SdPdVSduUudVudtd

SduU

rrrrrr

rrr

τρρ (2.2)

The Equations 2.2 provides the form in which the equations are used to develop

the code.

Page 31: XUE JIANQING

24

2.3 Stress Calculation

For an axisymmetric flow ( 0=∂∂θ

) with a swirling velocity, the six components

of sheer stress tensor can be simplified as:

rv

err ∂∂

= µτ 2 (2.3)

rv

eµτθθ 2= (2.4)

zu

ezz ∂∂

= µτ 2 (2.5)

)(rw

rw

er −∂∂

= µτ θ (2.6)

zw

ez ∂∂

= µτθ (2.7)

)(ru

zv

ezr ∂∂

+∂∂

= µτ (2.8)

Here u, v, and w denote the axial, radial and tangential (swirl) velocity

components respectively. µ e is the effective viscosity. It is taken to be the sum of laminar

and turbulent components:

tle µµµ += (2.9)

where the subscripts l and t represent laminar and turbulent, respectively. In this

study, the Baldwin-Lomax two layer turbulence model (Baldwin and Lomax, 1978) is

used to predict value of µe.

The Baldwin-Lomax model is based on the renowned Cebeci-Smith model, which

is a two-layer algebraic model for the turbulent viscosity:

Page 32: XUE JIANQING

25

⎩⎨⎧

>≤

=co

cit yyfor

yyfor............

µµ

µ (2.10)

where µi is turbulent viscosity within inner layer

µo is turbulent viscosity within outer layer,

and yc is the distance from a wall to the first point at which µi>µo.

A Prandtl-van Driest formulation is used for the inner layer:

ρτν

κωρµ

w

i

u

uyy

AyD

yDll

=

=

⎭⎬⎫

⎩⎨⎧−

−=

=

=

+

+

+

*

*

2

exp1 (2.11)

where ω is the vorticity, ρ is the density, κ is the von Karman constant, and D is the van

Driest damping function with A+=26.

The outer layer formulation is:

⎟⎟⎠

⎞⎜⎜⎝

⎛=

max

2

maxmaxmax ,min.fu

yCfyCK diffwkcpo ργµ (2.12)

where K=0.0168 is the Clauser constant and Ccp=1.6 is an additional constant. In wakes,

D is set to 1 and the lower portion of this equation is used. Cwk was originally reported as

0.25, but is generally taken to be 1.0. udiff stands for the difference between the highest

and the lowest velocities in the profile. The Klebanoff intermittency function γ is given

by:

Page 33: XUE JIANQING

26

12

max

5.51

⎥⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛+=

yyCKlebγ (2.13)

where CKleb=0.3.

The function f(y) is defined by:

Dyyf ω=)( (2.14)

Noted that fmax = max(f(y)) and ymax = y at fmax

2.4 Discretization

The popular discretizing methods in space consist of Finite Difference Method

(FDM), Finite Volume Method (FVM) and Finite Element Method (FEM). Unlike FDM,

FVM is based on integral formulation. The integral form of conservation laws are

satisfied for each control volume and for the entire domain. Therefore, the finite volume

method is more preferable to the finite difference approach. In code developing, FVM is

much easier to implement compared to FEM. Hence, FVM was chosen here for space

discretization. For time, the explicit time-marching method was selected for simplicity.

The mesh used here consists of a network of quadrilateral cells with vertices

labeled by integral pairs (i, j). Fluid variables are assigned to staggered locations in the

mesh (shown in Figure 2-1). Pressures (p), cell volumes (V), and Masses (M) are all

assigned to cell centers whereas coordinates (z, r) and velocity components (u, v, w) are

all assigned to cell vertices.

Due to the staggered assignment of the fluid variables, the control volume for

mass conservation is different from that for momentum conservation. In particular, the

control volume and control surface for mass continuity are the volume and corresponding

Page 34: XUE JIANQING

27

Figure 2-1 A schematic showing assignment of the variables for a cell

Figure 2.2 Schematic of mass and momentum control volume

Momentum C. V.

Mass C. V.

J+1

J

J-1

J-2

I-3 I-2 I-1 I I+1

i+1, j

i, j+1

i+1, j+1

p, V, M

i, j

z, r, u, v, w

Page 35: XUE JIANQING

28

surface of a cell in the mesh, and control volume and control surface for momentum

conservation are the vertex-centered volume and corresponding surface surrounding a

vertex point (shown in figure 2.2). In contrast to a mass cell which has four faces, the

momentum cell has eight faces, each comparable in size to one-half of a regular cell face.

2.5 Pressure Field Iteration

As the time moves from t (time step N) to t +∆t (time step N+1), both velocity

field and pressure field have to be updated. Because velocity and pressure are coupled,

they can not be computed explicitly. Here a predictor-corrector method is applied to

solve this problem. Specifically, the Conjugate-Gradient method (O’Rourke et. al., 1986)

together with Jacobian matrix preconditioning is used to solve the velocity and pressure

implicitly. A brief discussion of this technique is presented here.

The Conjugate-Gradient method is used to solve the problem

0)( =PRrr

(2.15)

Where Rr

is the vector of residuals in the solution of a set of implicit linear finite-

volume equations and Pr

is the vector of unknowns - pressure for our case - we wish to

solve for. Number of components for both Rr

and Pr

are equal to the number of

computational cells. Assuming that vPr

predicts Pr

with volume residual )( vPRrr

, based on

a Taylor expansion neglecting higher order terms;

( ) 0)()( =−⋅⎟⎟⎠

⎞⎜⎜⎝

⎛+= v

Pv PP

PRPRPR

v

rrr

rrrrr

r∂∂ (2.16)

Define a Jacobian matrix A as,

Page 36: XUE JIANQING

29

vPPRA

rr

r

⎟⎟⎠

⎞⎜⎜⎝

⎛−=

∂∂ (2.17)

The vector vPr

∆ can be defined as:

vv PPPrrr

−=∆ (2.18)

Then,

)( vv PRPArrr

=∆ (2.19)

The problem can now be rephrased as finding the solution of vPr

∆ in Equation

2.19. The convergence criterion is the satisfaction of the Equation 2.15. Note that A has

to be non-singular, and also of rank n where n is the number of components of vPr

in

order to be able to apply the Conjugate-Gradient method. The method is known to be

faster for symmetric A. Assume that A is non-singular at this point.

Define a preconditioning matrix M such that

( ) vv PPRMrrr

∆=)( (2.20)

It is best to take M as close as possible to 1−A , the inverse of A. The iteration

procedure can be explained in four steps:

1) Calculate the current volume residual vRr

on guessed pressure field vPr

at

iteration v ;

2) Calculate

( ))(ˆ vvv PRMPqrrrr

=∆= (2.21)

Which is a guess for the solution to pressure adjustment;

Page 37: XUE JIANQING

30

3) Orthogonalize vqrˆ with respect to 1−vqr , the previous pressure correction

direction vector, the new pressure correction direction vector is obtained,

1ˆ −−= vvvv qqq rrr λ (2.22)

where

( ) ( )

( ) ( )∑∑

−−

=

iiviv

iiviv

v qMAqA

qMAqA

11

1ˆrr

rr

λ (2.23)

4) Find the component 1+∆ vPr

in the direction of vqr ,

vvv qP rrα=∆ +1 (2.24)

Where

( ) ( )

( ) ( )∑∑ −

=

iiviv

iivivB

v qMAqA

qMAPPArr

rrr()(

α (2.25)

Then,

11 ++ ∆+= vvv PPPrrr

(2.26)

Next, return to the first step where 1+= vB PPrr

is the predicted pressure at (N+1)

time step.

To initiate this iteration algorithm for v = 0 at time step (N+1), the pressure fields

of time steps (N) and (N-1) is used to extrapolate linearly. The Conjugate Gradient

method is proved to effectively converge to the solution after L iterations, where L is the

number of computational cells.

Page 38: XUE JIANQING

31

The key to the implementation of this method into the code is to find the non-

singular matrix A and its pre-conditioning Jacobian matrix M. For given cell (i, j), a nine-

point formula relating residuals and pressures can be expressed as,

),,,,,,,,( 1,11,1,1,11,11,1,1,1,, ++++−−−−−−++= jijijijijijijijijiji PPPPPPPPPfRES (2.27)

It is difficult to include all nine points of the pressure influence into matrix A.

Since the pre-conditioning matrix M needs to be as close as possible to A-1, and also

considering the requirement of computational efficiency, only the main diagonal

elements of A have been used to define an approximation to matrix A, and its pre-

conditioning Jacobian matrix M is given as,

⎟⎟⎠

⎞⎜⎜⎝

⎛= lm

lmaM δ1 (2.28)

Where lma is the respective element of matrix A, and lmδ is the delta function:

⎩⎨⎧

≠=

mlml

lm if 0= if 1

δ (2.29)

This simplification greatly reduces the computational complexity, although it is at

the expense of an increase in the number of iterations.

From Equation 2.15, matrix A can be determined to be:

ji

jiji P

RA

,

,, ∂

∂= (2.30)

Coefficients vα and vλ are evaluated by applying A to the vectors vqArˆ , vqAr , and

)( vB PPArr

− ,

Page 39: XUE JIANQING

32

1111

1

)()()()(

ˆ

)ˆ()(ˆ

−−−−

−=∆−−=−=

−=

+−=

vvvvvBvBv

vvvv

vvvv

qAPRPPPAPPAPR

qAqAqA

qPRPRqA

rrrrrrrrrr

rrr

rrrrrr

α

λ (2.31)

The convergence criterion for each cell (i,j) is based on finding an acceptably

small value of )(PRrr

which obeys

Ajiv VPR ,)( ε≤

rr (2.32)

Where ε is a small number in the order of 710− , and AjiV , is some initial cell

volume.

2.6 Boundary Conditions

The governing equations are elliptic in space and require boundary conditions on

all boundaries of the computational domain. There are four kinds of boundary conditions

(BC) encountered in this study: inflow BC, static wall BC, outflow BC and free film BC.

Inflow Boundary Condition

The specified values of the velocity components in axial, radial and tangential

directions are applied here as the inflow boundary condition. The values of the velocity

components are calculated according to the flow rate and the inlet slots geometry. As the

present model is axisymmetric, boundary conditions require an "equivalent"

axisymmetric inlet. The width of this inlet and the velocity components are calculated by

equating the mass flow rate, angular momentum, axial momentum and kinetic energy to

the actual inlet conditions. The equations to calculate the width of the equivalent

axisymmetric inlet and the velocity components vary with nozzle geometry.

Page 40: XUE JIANQING

33

Static Wall Boundary Conditions:

The rigid no-slip boundary condition is used for the static wall. All three

components of velocity are set to zero on the rigid boundaries.

Outflow Boundary Conditions:

Typically outflow boundary condition of zero velocity gradient along a stream

line is used in many simulations. In the problem at hand, as the liquid exits from the

nozzle in form of a hollow cone, the cross-sectional area for the flow increases as the

liquid spreads out. As such the typical outflow boundary condition introduces slight error

in mass conservation. As a reason, in the present model, a slightly different boundary

conditions is used such that the mass flow, the angular momentum and kinetic energy are

conserved between the grid points on the outflow boundary and their interior neighbors.

According to these three rules, the velocity components at the outflow BC are calculated.

Several cases were run using the typical outflow boundary condition and the new

condition that conserves mass and angular momentum. It was seen that the differences in

the results obtained using the two outflow boundary conditions are slightly different.

However, application of the new condition made the code converge faster and made it

somewhat more stable under extreme conditions. As such the new outflow condition is

better and therefore it is used the present model. The equations for this boundary

condition are:

v(nxp,j) = sqrt[(1-rw**2)*w(nx,j)**2/(1+ru(j)**2)+v(nx,j)**2]

u(nxp,j)=ru(j)*v(nxp,j)

Page 41: XUE JIANQING

34

w(nxp,j)=rw*w(nx,j)

where,

rw = r1/r2

r1 = (y(nx,1)+y(nx,nyp))/2.

r2 = (y(nxp,1)+y(nxp,nyp))/2

ru(j) = u(nx,j)/v(nx,j)

The node points nxp and nyp indicate the last points in the computational domain

in the z and r directions, respectively.

Free film BC

At a two-phase interface, the normal stress and the shear stress must be

continuous across the interface. As this code deals with an interface between liquid and

air, and as the liquid viscosity is significantly greater than that of air, the shear stress at

the interface is negligible. Such interface is referred to as a "free surface" and the

condition of zero shear stress is imposed at the interface. The surface tension and

pressure is balanced at the interface. It is noted that in certain geometry, the distortion of

the free film inside the atomizer may be large during first few iterations and may lead to

the code to diverge. To overcome this problem, the motion of points on about half of the

free film inside the atomizer is initially constrained in a way that all the points have the

same vertical displacement. Once the initial transients diminish, this constraint is

removed. All of this is done automatically inside the code based on our experience with

nozzle geometries and the user does not need to control this. Additionally, it should be

emphasized that, for most cases, part of inner free film in the swirl chamber has been

Page 42: XUE JIANQING

35

constrained in order to avoid radial location of any point on the interface becoming zero.

The radial location (r) of any point on the interface going to zero implies that the air core

has collapsed. Should this happen, the boundary condition for balance of surface tension

and pressure is expressed in terms of curvature surface curvature (~1/r) and will become

singular. As such the code would diverge.

2.7 Grid Generator

A pre-processor is developed to generate the initial grid automatically. The entire

calculation domain is divided into 5 blocks. They are: swirl chamber block, convergent

section block, orifice block, trumpet block and free film block. If the geometry has no

trumpet, then there are only 4 blocks. Next, for each block, the boundary is divided into a

set of discrete points. Experience of running the code for a variety of flow condition and

geometries has been used to provide a "reasonable" grid. Finally, the whole domain is

divided based on transfinite Hermite interpolation (Stoer and Bulirsch, 1993). This pre-

processor generates a two-dimensional algebraic grid automatically.

Page 43: XUE JIANQING

36

2.8 Code Structure

Figure 2.3 Flowchart of the code

Pressure Field Solving Iteration

Time Step ∆t Estimation

Velocity Updating to Take Viscous Stress into Account

Velocity Updating due to the Convective Transport

Velocity Updating with the Pressure Acceleration

Mesh Regridding

Start

Initialization

Converged?

End

Yes

No

Page 44: XUE JIANQING

37

CHAPTER 3 CODE VALIDATION

3.1 Verification of Grid Independence of Results

For any computational flow simulation, the validation of grid independence is

always necessary (Wilcox, 1993). The code was run under several flow conditions

with different grid density to assure grid independence of results. Two sets of results

with the same flow condition and geometry are shown here with 80X21 and 159X41

grid, respectively. Figure 3.1 shows the final grids. The results shown in Table 3.1

indicate that the difference between coarse grid and fine grid is very small. As such

an 80X21 or similar grid should be sufficient to get accurate results.

Table 3.1 Results of different grid density.

Grid 1

(80X21) Grid 2 (159X41)

t (mm) 0.5378 0.5322 uav(m/s) 24.20 24.51 wav(m/s) 19.15 19.32 Cd 0.4350 0.4505 θ(degree) 38.35 38.25

3.2 Comparison with Experimental Data and Empirical Correlations

Measurement of flow field inside a small scale nozzle (exit orifice diameter ~

1mm) is extremely difficult. Such measurements are essential to validate the

computational model. Limited data for spray angle and film thickness are available in the

literature for small-scale nozzles. Unfortunately the details of the nozzle geometry are

Page 45: XUE JIANQING

38

Figure 3.1 Final grids of different grid density

(2D

)⏐

29A

ug20

00⏐

cycl

e40

000

0

00

.001

0.00

20.

003

0.00

40.

005

0.0

060

.007

0.0

08X

0

0.0

005

0.00

1

0.0

015

0.00

2

0.0

025

0.00

3

0.0

035

Y

Grid

:80

X21

Rea

l-Siz

eno

zzle

(2D

)⏐

29A

ug20

00⏐

cycl

e5

0000

0

00.

001

0.00

20.

003

0.0

040

.005

0.0

06

0.00

70.

008

X0

0.00

05

0.0

01

0.00

15

0.0

02

0.00

25

0.0

03

0.00

35

Y

Grid

:159

X41

(2D

)⏐

29A

ug20

00⏐

cycl

e5

0000

0

Page 46: XUE JIANQING

39

generally not provided and therefore the available data can not be used for code

validation. To overcome this difficulty, Jeng et al. (1998), Wang (1999) and Ma (2001)

carried out flow measurements on large-scale prototype nozzle. A brief description of

their setup and measurement technique is described below for sake of completeness. I

have used their data to validate the developed computational code. The large-scale

prototype nozzles provided sufficient spatial resolution for flow measurements. The

body of the prototype nozzle was made of optical quality plexiglass to use optical

diagnostic techniques (Jeng et al, 1998). The Particle-Image-Velocimetry (PIV)

technique, CCD camera, and still photography were used for measurements. In general,

the uncertainty in the measurements is about 2% for discharge coefficient, about 2

degrees in spray angle, and less than 15% in film thickness measurements. Three cases

were simulated for which experimental measurements were available. The details of

three cases are given in Table 3.2.

Table 3.2 Cases summary

Case # Ds

(mm)

Ls

(mm)

d0

(mm)

Inlet

slots

l0

(mm)

Ap

(mm2)

θ Flow rate

(gal./min)

1 76 89 21 4 5 406 90 15

2 76 89 21 4 5 406 45 15

3 76 89 21 4 5 406 45 20

Page 47: XUE JIANQING

40

Comparison of the Swirl Velocity

Particle-Image-Velocimetry (PIV) technique was used to measure swirl velocity

variation at different axial locations in the large-scale prototype nozzle. For Case # 2

listed in Table 3.2, comparison of experimental data and the CFD predictions of the swirl

velocity profiles at three different axial locations are shown in Figure 3.2. It is seen that

the swirl velocity is the largest close to the free surface and decreases away from the

gas/liquid interface. The variation predicted by the CFD code is seen to match with the

experimental data extremely well. The slight difference near the gas/liquid interface can

be attributed to the viscous effects of the air which are neglected in the CFD model. For

any CFD code, excellent match with local parameters, such as velocity variation at

different points in the flow field, is difficult to achieve compared to good agreement with

global parameters (such as Cd). As such the excellent agreement between the swirl

velocity variation from experiments and CFD, provides a strong validation of the CFD

model.

Figure 3.2 Comparison of swirl velocity variations at three axial locations

Page 48: XUE JIANQING

41

Comparison of predicted film thickness, spray angle, and discharge coefficient

with experiments and empirical correlations

To further validate the code and to assess its accuracy relative to the available

empirical correlations, comparisons for spray angle, film thickness, and discharge

coefficient were carried out. Several correlations are available in the literature (Lefebvre,

1989).

The correlations listed below are probably the most widely used and as such these

were considered for comparison with CFD predictions. The correlations are listed in two

groups. The first group consists of correlations that are given by Lefebvre and co-workers

(Lefebvre, 1989) and the second group by other researchers (Jones, 1982, Suyari, 1986).

25.0

66.3 ⎟⎟⎠

⎞⎜⎜⎝

⎛∆

=LL

LLo

Pmd

µ& (3.1)

11.0

2

215.062 ⎟⎟

⎞⎜⎜⎝

⎛ ∆= −

L

LoLm

dPK

µρ

θ (3.2)

25.05.0

35.0 ⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛=

o

s

os

pd d

DdD

AC (3.3)

Other correlations:

XX

dD

dDA

o

s

os

p

+−

=⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛1

)1(09.035.0

(3.4)

XX

+−

=11cos2 θ (3.5)

23.052.005.003.002.0

45.0 ⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛=

−−

o

s

os

p

s

s

o

o

L

thLod d

DdD

ADL

dlUd

Cµρ

(3.6)

Page 49: XUE JIANQING

42

Table 3.3 Comparison results of experimental, CFD and correlation data

CFD Lefebvre Other Experimental

Value Error Value Error Value Error

Cases with 4 inlet slots (flow nearly axi-symmetric)

∆p (Pa) 48263 52022 +8%

t (mm) 2.633 2.932 2.434

θ 42.90 41.26 -4% 50.39 +17% 61.22 +43%

CASE 1

Cd 0.2781 0.2678 -4% 0.2435 -12% 0.2446 -12%

∆p (Pa) 46883 58919 +25%

t (mm) 2.68 2.544 -5.1% 2.953 +10% 2.434 -9%

θ 39.35 42.26 +7.4% 50.23 +28% 61.22 +55%

CASE 2

Cd 0.2821 0.2517 -10.8% 0.2435 -14% 0.2446 -13%

∆p (Pa) 89631 105808 +18%

t (mm) 2.64 2.542 +3.7% 2.699 +2% 2.434 -8%

θ 40.35 42.40 +5.1% 53.95 +34% 61.22 +52%

CASE 3

Cd 0.2721 0.2504 -8% 0.2435 -10% 0.2430 -11%

Table 3.3 shows that for all the cases considered here, the CFD code gives a

significantly better prediction for the spray cone angle compared with empirical

correlations. In general, the difference in experimental measurements is about 5%.

Considering that the experimental uncertainty in the measurements is greater than 5%,

this is indeed excellent agreement. In contrast, the correlations deviate from the

experimental values from 20% to 50%. For the free film thickness at the exit and

discharge coefficient, in the cases (case 1, 2 and 3) that have 4 inlet slots, the results from

CFD code are better than those from correlations with deviation from experimental

Page 50: XUE JIANQING

43

values less than 5%. Once again, this is within the experimental uncertainty and shows

excellent agreement.

3.3 Comparison with Correlations at Small Scale

The results obtained from the CFD code for small scale atomizers were compared

with correlations available in the literature. As mentioned earlier, details of the nozzle

geometry are not available where experimental data for spray angle, film thickness and

discharge coefficient are given in the literature. As such a comparison can only be made

using empirical correlations for small-scale nozzles. Predictions for discharge

coefficient, film thickness, and spray angle were obtained with varying the atomizer

constant. Some of the correlations are known to work well to evaluate the effect of

variation of atomizer constant, and the good agreement with CFD results provides

additional confidence in the computational code.

Discharge coefficient

The discharge coefficient is governed partly by the pressure losses incurred in the

nozzle flow passes and also by the extent to which the liquid flowing through the final

discharge orifice makes full use of the available flow area (Lefebvre, 1989). The

discharge coefficient is the ratio of the actual to the maximum theoretical flow rate that is

determined from the measured pressure drop across the atomizer.

LoL

Ld PA

MCρρ /2∆

=&

(3.7)

Page 51: XUE JIANQING

44

Atomizer Constant, K

Dis

char

gerC

oeffi

cien

t,C

d

0 0.2 0.4 0.60.1

0.2

0.3

0.4

0.5

CFDBenjaminJonesRizk

Figure 3.3 Variation of discharge coefficient

Page 52: XUE JIANQING

45

Rizk et al. (1984) derived the following expression:

25.05.0

35.0 ⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛=

o

s

os

pd d

DdD

AC (3.3)

From analysis of experimental data on the discharge coefficient of large-scale

pressure swirl atomizers, Jones (1982) obtained the following correlation:

23.052.005.003.002.0

45.0 ⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛=

−−

o

s

os

p

s

s

o

o

L

thLod d

DdD

ADL

dlUd

Cµρ

(3.6)

Benjamin et al (1998) used the similar equation to Jones’s to correlate his

experimental data and resulted in the following expression:

187.0517.0091.0229.0027.0

466.0 ⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛=

o

s

os

p

s

s

o

o

L

thLod d

DdD

ADL

dlUd

Cµρ

(3.8)

As seen in Figure 3.3, there is fair amount of agreement between the different

correlations; and CFD results match with the predicted values quite well.

Film thickness

At small scale, the measurement of film thickness is fairly difficult to accomplish.

A number of techniques have been reported in the literature. Our large-scale nozzle flow

simulations as well as experiments have shown that the film thickness in the exit orifice

is not constant. It varies with axial location and the variation is not monotonic. Film

thickness measurements tend to capture an average value of thickness over a small but

finite axial distance, whereas the film thickness obtained from the CFD code is at the exit

of the orifice. The correlations for film thickness are of two types, first that directly

provide thickness based on fuel properties, pressure drop, and orifice exit diameter.

Page 53: XUE JIANQING

46

Atomizer Constant, K

Dim

ensi

onle

ssTh

ickn

ess

atE

xit,

t*

0 0.2 0.4 0.60.1

0.2

0.3

0.4

0.5

CFDRizkGiffenBenjamin

Figure 3.4 Variation of dimensionless film thickness

Page 54: XUE JIANQING

47

Rizk et al. (1984) gave the following correlation,

25.0

66.3 ⎥⎦

⎤⎢⎣

⎡∆

=P

mdt

L

LLo

ρµ&

(3.1)

Benjamin et al. (1998) derived their correlation as,

25.0

5.0)(78.3 ⎥

⎤⎢⎣

⎡∆

=P

FNdt

L

Lo

ρµ

(3.9)

Another type of correlation was given by Suyari and Lefebvre (1986) where the

film thickness is obtained by equating discharge coefficient to the discharge coefficient

obtained by Giffen and Muraszev (1953) by inviscid analysis.

XX

dD

dDA

o

S

oS

p

+−

=⎥⎦

⎤⎢⎣

⎡⎥⎦

⎤⎢⎣

⎡1

)1(09.035.0

(3.4)

Suyari and Lefebvre (1986) have compared small scale film thickness

measurements to a number of correlations and found that there is large disagreement

between different correlations. They found that Equation 3.4 provides the best fit for

available data. Figure 3.4 indicates that the CFD predictions match very well predictions

by Equation 3.4. Also, the CFD results are significantly better than other correlations.

This provides further validation of the CFD code.

Spray cone half angle

There is no unanimity for the spray cone angle predictions by using different

correlations. The following three correlations are compared with CFD results.

Rizk’s correlation 1 (Rizk and Lefebvre, 1984)

XX

+−

=11cos2 θ (3.5)

Page 55: XUE JIANQING

48

Atomizer Constant, K

Spr

ayC

one

Hal

fAng

le,θ

0 0.2 0.4 0.6

20

30

40

50

60

70

CFDATAN(wav/uav)Rizk2BenjaminRizk1

Figure 3.5 Variation of Spray Angle

Page 56: XUE JIANQING

49

Rizk’s correlations 2 (Rizk and Lefebvre, 1984)

11.0

2

215.062

⎥⎥⎦

⎢⎢⎣

⎡∆= −

L

LoPdK

µρ

θ (3.2)

Benjamin gave the following correlation (Benjamin et al, 1998),

0647.0

2

2237.075.92

⎥⎥⎦

⎢⎢⎣

⎡∆= −

L

LoPdK

µρ

θ (3.10)

Of these correlations, Benjamin's correlations was developed using data for large

scale prototype nozzle and as such the difference between that and the CFD results is not

surprising. It is interesting that the other two correlations have significant disagreement.

Figure 3.5 shows the CFD results match well with one of Rizk's correlations.

3.4 Summary

The grid independence study was conducted and the result indicates an 81X21 or

similar grid should be sufficient to get accurate simulation for typical simplex nozzle.

A detailed comparison with the experimental data was carried out. The

comparison showed that for the cases with four inlet slots the code is able to predict

performance parameters very well (generally within 5% of the experimental

measurements). We note that the differences are less than the uncertainty in the

experimental measurements.

The results obtained from the CFD code for small scale atomizers were compared

with correlations available in the literature. Predictions for discharge coefficient, film

thickness, and spray angle were obtained with varying the atomizer constant. The good

agreement with CFD results provides additional confidence in the computational code.

Page 57: XUE JIANQING

50

CHAPTER 4 EFFECTS OF GEOMETRIC PARAMETERS ON

ATOMIZER PERFORMANCE (PART I)

4.1 Introduction

Pressure-swirl atomizers used in gas turbine engines and other spray applications

have to operate over a wide range of flow conditions. It is very important to be able to

predict spray performance based on the relevant input parameters into the nozzle. Since

predictive breakup models are not available with the level of accuracy required for the

design process, design engineers still have to struggle with the cut-and-try methodology

in order to fine-tune the nozzle into the required spray properties. Presently, semi-

empirical correlations are used to provide guidance in designing simplex nozzles

(Benjamin et al 1998). However, the semi-empirical correlations do not always predict

the effects of changes in the length-scales on spray properties. Furthermore, not all the

geometric parameters are included in the currently available correlations. The validated

CFD code is used in this study to perform such parametric study to understand the effect

of atomizer geometry on atomizer performance.

For pressure-swirl atomizers, it is now generally accepted that the dimension of

most importance for atomization is the thickness of the liquid sheet as it leaves the final

orifice (Lefebvre 1983). Theory predicts, and experiment confirms, that mean drop size

is roughly proportional to the square root of liquid sheet thickness at the exit (Lefebvre

1989). Discharge coefficient is another important parameter for simplex nozzle, because

Page 58: XUE JIANQING

51

this coefficient not only affects the flow rate of any given nozzle but also can be used to

calculate its velocity coefficient and spray cone angle (Lefebvre 1989). The spray angles

produced by pressure-swirl nozzles are of special importance in their application. One

important reason is that normally any increase in spray cone angle will lead to improved

atomization (Lefebvre 1989). Therefore, the dimensionless film thickness at the exit,

discharge coefficient and spray cone half angle are used here to describe the performance

of pressure-swirl atomizers.

In this chapter, the effects of the length to diameter ratio of the swirl chamber

(Ls/Ds) and of orifice (lo/do), the swirl chamber to orifice diameter ratio (Ds/do) on the

performance of a simplex atomizer are numerically investigated. The definitions of these

geometric parameters are described in Figure 1.1. These three parameters are generally

not included in correlations for film thickness and spray angle. Two flow situations are

important – constant mass flow rate through the nozzle and constant pressure drop across

the nozzle. Here results are presented for both working conditions. A range of Ls/Ds from

0.1 to 1.5, a range of lo/do from 0.2 to 2.0, a range of Ds/do from 2 to 6 have been covered.

The base geometric parameters are Ls/Ds = 0.25, lo/do = 0.5, Ds/do = 4.0, and K = 0.47.

While studying the effects of changes in Ls/Ds, only Ls is varied and all other dimensions

are kept as base geometry. Similarly, lo is varied to change lo/do. While studying the

effects of changes in Ds/do, in order to hold K as constant, both Ds and do are varied such

that Ds*do is kept constant. The atomizer performance is monitored in terms of

dimensionless film thickness at the exit, spray cone half angle and discharge coefficient.

Page 59: XUE JIANQING

52

4.2 Results for Constant Mass Flow Rate through the Atomizer

Effect of variation in Ls/Ds

For given inlet mass flow rate, eight different cases, which Ls/Ds is 0.1, 0.2, 0.25,

0.5, 0.75, 1 and 1.5 respectively, are investigated numerically here. Noted that the

change of Ls/Ds results from the variation of Ls whereas Ds is held constant.

Figure 4.1 gives the variations of the dimensionless film thickness at the exit and

discharge coefficient with changing Ls/Ds. As Ls/Ds increases, both t* and Cd increase

monotonously. Because Ds is held constant, larger Ls/Ds corresponds to longer swirl

chamber length and results in an increasing decay of swirl energy and consequently, a

smaller air core, i.e., a bigger film thickness at the exit. As the inlet flow rate keeps

constant, the thicker the film is, the smaller the pressure drop throughout the atomizer and

consequently the larger Cd is.

Figure 4.2 indicates that with an increase of Ls/Ds, the spray cone half angle

diminishes slightly. Theoretically, as Ls increases, both the swirl energy and axial energy

are deceased at the exit. However, Figure 4.2 shows that wav/uav decreases with increase

in Ls/Ds. That means the decay of swirl energy is larger than that of axial energy

throughout the nozzle. Therefore, a decrease in the spray angle results from a decrease in

the ratio of exit swirl to axial velocity components.

Page 60: XUE JIANQING

53

Ls/Ds

Dim

ensi

onle

ssTh

ickn

ess

atE

xit,

t*

Dis

char

geC

oeffi

cien

t,C

d

0 0.5 1 1.50.3

0.35

0.4

0.3

0.35

0.4

0.45

Cdt*

Figure 4.1 Variation of the dimensionless film thickness at the exit and discharge

coefficient with Ls/Ds under constant flow rate

Page 61: XUE JIANQING

54

Ls/Ds

Spr

ayC

one

Hal

fAng

le,θ

wav

/uav

0 0.5 1 1.530

35

40

0.6

0.7

0.8

0.9

θwav/uav

Figure 4.2 Variation of the spray cone half angle and wav/uav with Ls/Ds under constant

flow rate

Page 62: XUE JIANQING

55

Effect of variation in lo/do

In order to conduct the influence of variation in lo/do on the atomizer performance,

eight different values of lo/do are considered here, which are 0.2, 0.3, 0.4, 0.5, 0.75 1, 1.5

and 2. Similar to the variation of Ls/Ds, only lo is changed to change the value of lo/do

and do is constant for the all eight cases.

Figure 4.3 shows that as lo/do increases from 0.2 to 0.5, t* decreases dramatically;

when lo/do = 0.75, t* is minimum; and as lo/do changes from 0.75 to 2, t* increases

slightly. This behavior of film thickness matches the trends discussed in Benjamin et al

(1998).

Figure 4.3 also indicates that as lo/do increases, Cd decreases with lo/do. This is

similar to the variation of Cd with Ls/Ds and because of the similar reason. It is noted that

the trend is rather steep at lower lo/do, while its gradient decreases at higher lo/do.

The variations of spray cone half angle and wav/uav with changing lo/do are given

in Figure 4.4. With increasing lo/do, θ initially diminishes with a high gradient when lo/do

is less than 0.75; as lo/do is greater than 0.75, the gradient of θ decreasing diminishes

rapidly. The varying tendency of wav/uav with increase in lo/do is similar to that of θ.

Page 63: XUE JIANQING

56

lo/do

Dim

ensi

onle

ssTh

ickn

ess

atE

xit,

t*

Dis

char

geC

oeffi

cien

t,C

d

0 0.5 1 1.5 20.3

0.35

0.4

0.3

0.35

0.4

0.45

Cdt*

Figure 4.3 Variation of the dimensionless film thickness at the exit and discharge

coefficient with lo/do under constant flow rate

Page 64: XUE JIANQING

57

lo/do

Spr

ayC

one

Hal

fAng

le,θ

wav

/uav

0 0.5 1 1.5 230

35

40

0.6

0.7

0.8

0.9

θwav/uav

Figure 4.4 Variation of the spray cone half angle and wav/uav with lo/do under constant

flow rate

Page 65: XUE JIANQING

58

Effect of variation in Ds/do

Eight values of Ds/do are considered 2, 2.5, 3, 3.5, 4, 4.5, 5 and 6. Figures 3.5, 3.6

and 3.7 show that the atomizer performance is strongly influenced by atomizer constant

K. In order to extract the effect of variation in Ds/do independently, K has to keep

constant. This is achieved by keeping Ds*do as constant when varying the Ds/do.

Figure 4.5 illustrates that with increasing Ds/do, the dimensionless film thickness

at the exit diminishes when Ds/do is less than 3; as Ds/do is greater than 3, t* becomes to

increase. Since K (and Ds*do) is constant in this study, to keep Ds*do constant, do has to

be decreased with increasing Ds/do. Actually, computational result indicates that the

absolute film thickness at the exit decreases monotonously with increasing Ds/do.

Figure 4.5 also shows the discharge coefficient variation with changing Ds/do. It

can bee seen that with increasing Ds/do, Cd decreases at smaller value of Ds/do, eventually

becoming almost constant at larger values of Ds/do.

The variation of spray cone half angle and wav/uav with changing Ds/do are given

in Figure 4.6. With increase in Ds/do from 2 to 6, the spray cone half angle increase

initially and decreases at larger values of Ds/do but the maximum relative change is only

about 4%. This indicates that Ds/do has little influence on θ. With increasing in Ds/do,

wav/uav shows the similar behavior to Ds/do.

Page 66: XUE JIANQING

59

Ds/do

Dim

ensi

onle

ssTh

ickn

ess

atE

xit,

t*

Dis

char

geC

oeffi

cien

t,C

d

1 2 3 4 5 60.3

0.35

0.4

0.3

0.35

0.4

0.45

Cdt*

Figure 4.5 Variation of the dimensionless film thickness at the exit and discharge

coefficient with Ds/do under constant flow rate

Page 67: XUE JIANQING

60

Ds/do

Spr

ayC

one

Hal

fAng

le,θ

wav

/uav

1 2 3 4 5 630

35

40

0.6

0.7

0.8

0.9

θwav/uav

Figure 4.6 Variation of the spray cone half angle and wav/uav with Ds/do under constant

flow rate

Page 68: XUE JIANQING

61

4.3 Results for Constant Pressure Drop across the Atomizer

Effect of variation in Ls/Ds

The variations of the dimensionless film thickness at the exit and discharge

coefficient with changing Ls/Ds are shown in Figure 4.7. With the increase in Ls/Ds, both

t* and Cd increase monotonously. Larger Ls/Ds corresponds to longer swirl chamber

length and results in an increasing decay of swirl energy and a smaller air core at the exit.

The same trend was also observed in the experiment (Kutty et al 1978). As the pressure

drop is held constant, the thicker the film is, the larger the mass flow rate and

consequently the larger Cd is.

Figure 4.8 shows that with an increase of Ls/Ds, the spray cone half angle and

wav/uav decrease almost linearly but the diminishment of the spray angle is somewhat

small (only about 9% in the given condition and atomizer). This tendency coincides with

the experimental observation (Kutty et al 1978). Theoretically, this also results from

increasingly decaying swirl energy at the exit for a longer swirl chamber. A decrease in

the ratio of swirl to axial velocity components at the exit results in a decrease in the spray

angle.

Page 69: XUE JIANQING

62

Ls/Ds

Dim

ensi

onle

ssTh

ickn

ess

atE

xit,

t*

Dis

char

geC

oeffi

cien

t,C

d

0 0.5 1 1.50.3

0.35

0.4

0.45

0.3

0.35

0.4

0.45

Cdt*

Figure 4.7 Variation of the dimensionless film thickness at the exit and discharge

coefficient with Ls/Ds under constant pressure drop

Page 70: XUE JIANQING

63

Ls/Ds

Spr

ayC

one

Hal

fAng

le,θ

wav

/uav

0 0.5 1 1.530

35

0.6

0.7

0.8

θwav/uav

Figure 4.8 Variation of the spray cone half angle and wav/uav with Ls/Ds under constant

pressure drop

Page 71: XUE JIANQING

64

Effect of variation in lo/do

Figure 4.9 shows that as lo/do increases from 0.2 to 0.5, t* decreases dramatically;

when lo/do = 0.75, t* is minimum; and as lo/do changes from 0.75 to 2, t* slightly

increases. .

Figure 4.9 also shows that as lo/do increases, Cd decreases with lo/do. This is

similar to the variation of Cd with Ls/Ds and because of the similar reason. Note that the

trend is rather steep at lower lo/do, while its gradient decreases at higher lo/do.

The variations of spray cone half angle and wav/uav with changing lo/do are given

in Figure 4.10. With increasing lo/do, both θ and wav/uav initially have a high gradient

which decreases rapidly at lo/do greater than 0.75.

Page 72: XUE JIANQING

65

lo/do

Dim

ensi

onle

ssTh

ickn

ess

atE

xit,

t*

Dis

char

geC

oeffi

cien

t,C

d

0 0.5 1 1.5 20.3

0.35

0.4

0.3

0.35

0.4

Cdt*

Figure 4.9 Variation of the dimensionless film thickness at the exit and discharge

coefficient with lo/do under constant pressure drop

Page 73: XUE JIANQING

66

lo/do

Spr

ayC

one

Hal

fAng

le,θ

wav

/uav

0 0.5 1 1.5 230

35

40

0.6

0.7

0.8

0.9

θwav/uav

Figure 4.10 Variation of the spray cone half angle and wav/uav with lo/do under constant

pressure drop

Page 74: XUE JIANQING

67

Effect of variation in Ds/do

From Figure 4.11, we can see that with increasing Ds/do, the dimensionless film

thickness at the exit first decreases, then increases for Ds/do greater than 4.5. Note that in

this case K (and Ds*do) is constant. To keep Ds*do constant, do has to be decreased with

increasing Ds/do. It is found that the absolute film thickness at the exit decreases

monotonously with increasing Ds/do. Such behavior has been reported in several

experimental observations (Lefebvre 1989).

The discharge coefficient variation with changing Ds/do is also given in Figure

4.11. Cd decreases with increasing Ds/do, eventually becoming almost constant at larger

values of Ds/do.

The variation of spray cone half angle and wav/uav with changing Ds/do are shown

in Figure 4.12. With increasing Ds/do from 2 to 6, the spray cone half angle decreases by

about 15%.

Page 75: XUE JIANQING

68

Ds/do

Dim

ensi

onle

ssTh

ickn

ess

atE

xit,

t*

Dis

char

geC

oeffi

cien

t,C

d

1 2 3 4 5 60.3

0.35

0.4

0.3

0.35

0.4

0.45

Cdt*

Figure 11 Variation of the dimensionless film thickness at the exit and discharge

coefficient with Ds/do under constant pressure drop

Page 76: XUE JIANQING

69

Ds/do

Spr

ayC

one

Hal

fAng

le,θ

wav

/uav

1 2 3 4 5 630

35

40

0.6

0.7

0.8

θwav/uav

Figure 12 Variation of the spray cone half angle and wav/uav with Ds/do under constant

pressure drop

Page 77: XUE JIANQING

70

4.4 Summary

The effects of changes in the ratio of length to diameter in swirl chamber, the ratio

of length to diameter in orifice and ratio of orifice to swirl chamber diameters were

numerically investigated. The investigation was conducted under both constant inlet flow

rate and constant pressure drop across the atomizer. For both working conditions, the

dimensionless film thickness at the exit increases with increasing the ratio of length to

diameter in swirl chamber, the ratio of length to diameter in orifice when its value is

small and ratio of swirl chamber to orifice diameters when its value is small and with

decreasing the ratio of length to diameter in orifice and the ratio of swirl chamber to

orifice diameters when their values are large; the discharge coefficient increases with

increasing the ratio of length to diameter in swirl chamber and with decreasing the ratio

of length to diameter in orifice and the ratio of swirl chamber to orifice diameters; the

spray cone half angle increases with decreasing the ratio of length to diameter in swirl

chamber and the ratio of length to diameter in orifice. The spray angle has little change

with changing the ratio of swirl chamber to orifice diameters under constant flow rate

whereas it decreases with increase in the ratio of swirl chamber to orifice diameters under

constant pressure drop.

Page 78: XUE JIANQING

71

CHAPTER 5 EFFECTS OF GEOMETRIC PARAMETERS ON

ATOMIZER PERFORMANCE (PART II)

5.1 Introduction

With the advent of new manufacturing techniques, simplex atomizers with

different geometries can be produced easily. However to design such atomizers, the

effect of all geometric parameters on the performance of the atomizer must be well

understood. Newer designs of simplex atomizers may include inlets that are not

perpendicular to the axis and inlet slot angle (angle β in Figure 5.1) may not be 90

degrees. The geometry may include a trumpet at the end of the exit orifice to limit the

spray cone angle. It is noted that, to achieve rapid fuel air mixing, air flow may be

introduced around the fuel injector through a swirler cup. In such an arrangement, it is

desired to impinge the atomized fuel on the wall of a swirler cup to undergo further

atomization due to shearing action of air flow along the swirler cup wall. Therefore, in

such configurations, a particular spray angle is desirable rather than the maximum

possible spray angle. Hence a trumpet is introduced in the atomizer geometry to limit the

cone angle of the spray. Also a particular trumpet length may be needed to inject the

atomized fuel at the right location in the air flow for rapid mixing. The internal

convergence angle (angle θc in Fig. 5.1) may be different from 45 degrees. These

geometrical parameters have not been considered in studies available in literature and the

effect of variation in these parameters on the performance of simplex atomizers has not

Page 79: XUE JIANQING

72

Figure 5.1 Schematic of a simplex atomizer with trumpet

β

θc

θt do

dt

DS

lo

LS

Free Film

Inlet Slot

Interface Liquid

Gas

lt

Page 80: XUE JIANQING

73

been reported in literature. None of the available semi-empirical correlations address

these geometric parameters and as such the present correlations do not provide any

guidance about these parameters to atomizer designers.

The knowledge of flow in a pressure-swirl atomizer and the effect of geometric

parameters on the characteristics of the liquid sheet emanating from the atomizer are

extremely important to design a fuel injector. The characteristics of the liquid sheet

emanating from the atomizer depend on the geometrical parameters of the atomizer

which can be expressed in dimensionless form as atomizer constant (K), ratio of length to

diameter of the swirl chamber (Ls/Ds), that of the orifice (lo/do), ratio of swirl chamber to

orifice diameter (Ds/do), inlet angle (β), trumpet length, trumpet angle (θt), and the

convergence angle (θc). The detailed study of first four parameters listed above was

presented in Chapter 3 and Chapter 4. An atomizer designer must choose a combination

of geometric parameters to achieve the desired spray angle, film thickness, and discharge

coefficient. Therefore, it is important to know how each geometric parameter affects the

atomizer performance. Semi-empirical correlations have been developed by varying the

atomizer constant, Ls/Ds, lo/do, and Ds/do and measuring their effect on atomizer

performance parameters. Also, the influence of K, Ls/Ds, lo/do, and Ds/do on atomizer

performance has been studied computationally (Jeng et al 1998, Sakman et al 2000, Xue

et al 2002). However, no studies are currently available that provide any guidance on the

change in atomizer performance with a change in inlet slot angle, swirl chamber

convergence angle, trumpet angle and trumpet length.

Page 81: XUE JIANQING

74

In this chapter, the effect of four geometric parameters viz., inlet slot angle β,

swirl chamber convergence angle θc, trumpet angle θt, and ratio of trumpet length to

orifice diameter lt/do, on the performance of pressure-swirl atomizers is studied. The

variation in the atomizer performance is presented in terms of the dimensionless liquid

film thickness at the exit of the orifice (t*), spray cone half angle (θ) and discharge

coefficient (Cd). All cases were run keeping mass flow rate constant. When one

parameter was changed, all other geometric parameters were held constant. Also,

analytical solutions are developed based on inviscid approximation to determine the film

thickness, spray cone angle, and discharge coefficient for the atomizer.

5.2 Effect of Variation in Inlet Slot Angle β

To investigate the effect of inlet slot angle, An atomizer without a trumpet is used.

The dimensionless film thickness, spray cone half angle, and discharge coefficient are

plotted in Figures 5.2 and 5.3 with variation in inlet slot angle. The film thickness was

determined from the CFD results at the orifice exit. In the numerical method used here,

the interface between liquid and gas is tracked as a sharp interface and therefore the CFD

results directly provide the film thickness. The discharge coefficient is calculated based

on the pressure drop between in inlet and exit of the atomizer obtained in the numerical

solution.

Page 82: XUE JIANQING

75

Inlet Slots Angle β

Dim

ensi

onle

ssTh

ickn

ess

t*an

dD

isch

arge

Coe

ffici

entC

d

Spr

ayC

one

Hal

fAng

leθ

40 50 60 70 80 90

0.2

0.3

0.4

30

35

40

45

50

55

t* (CFD)t* (Equation 5.3)Cd (CFD)Cd (Equation 5.2)Cd (Equation 5.5)θ (CFD)θ (Equation 5.4)

Figure 5.2 Variation of atomizer performance parameters with inlet slots angle β

Page 83: XUE JIANQING

76

Inlet Slots Angle β

Dim

ensi

onle

ssTh

ickn

ess

t*an

dD

isch

arge

Coe

ffici

entC

d

Spr

ayC

one

Hal

fAng

leθ

40 50 60 70 80 90

0.2

0.3

0.4

30

35

40

45

50

55

t* (lo/do=1)t* (lo/do=0.5)Cd (lo/do=1)Cd (lo/do=0.5)θ (lo/do=1)θ (lo/do=0.5)

Figure 5.3 Comparison of results for lo/do = 1(solid lines) and lo/do =0.5 (dashed lines)

Page 84: XUE JIANQING

77

With a fixed mass flow rate through the atomizer, a change in inlet slot angle

results in a change in the ratio of inlet axial to swirl velocity components. With β = 90o

the fuel has no axial velocity component at the inlet. Hence for the same mass flow rate,

β = 90o gives greater swirl velocity compared to β = 40o case.

Figure 5.2 shows that with an increase of β from 40o to 90o, dimensionless

thickness t* decreases about 25%, discharge coefficient Cd decreases about 35% and

spray cone half angle θ increases by about 15%. The higher swirl velocity that

corresponds to a higher inlet slot angle pushes the liquid to the atomizer walls and the

liquid film thickness in the exit orifice decreases. The spray cone angle is governed by

the ratio of axial to swirl velocity components at the exit. The cone angle is seen to

increase with inlet slot angle. This is due to the larger centrifugal force caused by the

higher swirl velocity component. As the mass flow rate is kept constant, thinner liquid

film in the exit orifice section corresponds to higher axial velocity. A combination of

increased swirl and axial velocity leads to a higher pressure drop across the atomizer and

the discharge coefficient decreases. As evident from Figure 5.2, the changes in spray

cone angle, discharge coefficient, and film thickness are large, indicating that inlet angle

is an important parameter that significantly influences the performance of the atomizer.

It should be noted that empirical correlations available in the literature do not include the

inlet slot angle and are unable to predict its influence on atomizer performance.

Analysis can be carried out using inviscid theory to examine the effect of the inlet

slot angle. We have modified the inviscid analysis of Giffen and Muraszew (1953) to

Page 85: XUE JIANQING

78

include the inlet slot angle. The details are given in the appendix. Only the final

equations are given here.

βπ 22

322 sin)1(

32 XXK −

= (5.1)

5.03

1)1(

⎥⎦

⎤⎢⎣

⎡+−

=XXCd

(5.2)

From the definition of X, we get,

Xt −= 1* (5.3)

⎟⎟⎠

⎞⎜⎜⎝

+=

s

id

rr

XKC

)1(sin)2/(

sinβπ

θ (5.4)

It is interesting to note that Equation 5.2 is similar to one obtained by Giffen and

Muraszew (1953), however the effect of inlet slot angle is apparent in Equation 5.1. The

results Equations 5.2, 5.3, and 5.4 are plotted in Figure 2(a) using dashed lines. For film

thickness, comparing with the CFD results, it is seen that the analytical inviscid solution

provides the correct qualitative variation but predicts lower values. This is to be expected

as the inviscid theory does not account for the viscous effects that would tend to decrease

the velocity and increase film thickness.

Equation 5.2 predicts lower values for discharge coefficient than the

computational results. Giffen and Muraszew (1953) also observed this trend in

comparison with available experimental data. They modified their equation by

introducing a coefficient (A) as shown below and suggested A = 1.17 based on

experimented data.

Page 86: XUE JIANQING

79

0.53(1 )1d

XC AX

⎡ ⎤−= ⎢ ⎥+⎣ ⎦

(5.5)

It is seen in Figure 5.2 that our computational results match well with this

modified equation (Equation 5.5).

It is noted that there is significant difference in the prediction of spray cone angle

using inviscid approximation and numerical results. The numerical results are based on

solution of Navier-Stokes equations which fully account for viscous forces. Due to the

accounting of viscous forces, difference in the predictions of CFD results and those with

inviscid approximation (theoretical results) is to be expected. Furthermore, the spray

angle is governed by the ratio of the average swirl to the average axial velocity at the

atomizer exit. More precisely it is related to the tan inverse of the average swirl to axial

velocity ratio. A change in this ratio will change the spray angle. For the cases

considered here, the decrease in swirl velocity compared to inviscid prediction is larger

compared to the decrease in the axial velocity. Hence the inviscid results over-predict the

spray angle. The experiments (Jeng et al, 1998) also show considerable difference

between experimental measurements and inviscid analysis for spray angle, with inviscid

theory over-predicting the spray angle compared to experimental measurements,

consistent with the CFD results.

It is important to note that the equations derived using inviscid approximation

(Equations 5.1-5.4) do not account for variation in all geometric parameters. For

example, the ratio of length to diameter of the swirl chamber (Ls/Ds), that of the orifice

Page 87: XUE JIANQING

80

(lo/do) influence the spray angle, however these parameters do not appear in Equations

5.1-5.4 at all. Numerical results obtained with a different value of Ls/Ds (or lo/do) will be

different, but the inviscid predictions will remain the same. Therefore, for different value

of Ls/Ds (or lo/do) the difference between numerical results and inviscid approximation

will be different. This is seen in Figure 5.3 where results for spray cone angle, discharge

coefficient, and film thickness at exit, are presented for lo/do = 0.5 and lo/do = 1. It is

noted that the variations in atomizer performance parameters with β, for the two values of

lo/do are essentially parallel. This makes the results very useful in designing simplex

atomizers.

5.3 Effect of Variation in Trumpet Angle θt

Note in this study, the length of the trumpet lt has been held constant as θt is

varied. Figure 5.4 indicates that as θt increases from 10o to 40o, dimensionless thickness

t* decreases about 42%, spray cone half angle θ increases from 22o to 40o but discharge

coefficient Cd remains almost constant. This shows that the trumpet can be very effective

in controlling the spray cone angle and the film thickness without significantly affecting

the pressure drop across the atomizer. As pressure drop across the atomizer remains

relatively steady, the axial velocity at the end of the exit orifice is expected to exhibit

little variation. However, trumpet diameter dt increases with increase in θt so the film

thickness t* decreases when the flow rate through the atomizer is kept constant.

Page 88: XUE JIANQING

81

Trumpet Angle θtDim

ensi

onle

ssTh

ickn

ess

t*an

dD

isch

arge

Coe

ffici

entC

d

Spr

ayC

one

Hal

fAng

leθ

10 20 30 40

0.1

0.2

0.3

0.4

20

25

30

35

40

45

50

t* (CFD)t* (Equation 5.7)Cd (CFD)Cd (Equation 5.2)Cd (Equation 5.5)θ (CFD)θ (Equation 5.8)

Figure 5.4 Variation of atomizer performance parameters with trumpet angle θt

Page 89: XUE JIANQING

82

Considering inviscid flow through the atomizer, Analytical solutions for film

thickness, spray cone half angle, and the discharge coefficient for atomizer geometry with

trumpet have been developed. Detailed analysis and a schematic of the geometry are

provided in the appendix. The final equations are:

22

2

222

21sin1

)1(cos1

ds

i

ttttt

o

Crr

XKXAA

=⎥⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛+

−⎟⎟⎠

⎞⎜⎜⎝

⎛β

θ (5.6)

Where,st

pt

t

tat rr

AK

AA

== ,

to

tt r

rXt θcos)1(*

⎟⎟⎠

⎞⎜⎜⎝

⎛−= (5.7)

'θθθ += t (5.8)

Where, ts

i

t

t

rr

KX

θβθ cossin)1(2

tan '⎟⎟⎠

⎞⎜⎜⎝

⎛−=

Results using Equations 5.6-5.8 are shown in Figure 5.4 with dashed lines. Once

again the trends shown by the analytical solutions are seen to match well with

computational results. Similar to Figure 5.2, the analytical inviscid solutions give lower

values of film thickness and discharge coefficient, and higher values of spray cone half

angle. The results of Equation 5.2 under-predict the discharge coefficient. The modified

Equation 5.5 provides improved match with the numerical results. As stated earlier, the

differences in inviscid solutions and numerical results can be attributed to omission of

viscous forces in the inviscid treatment. It is interesting to note that compared to inviscid

flow through the atomizer, the decrease in average axial velocity due to viscous forces is

Page 90: XUE JIANQING

83

less than the reduction in average swirl velocity resulting in lower values of spray cone

half angle.

5.4 Effect of Variation in Dimensionless Trumpet Length (lt/do)

To study the effect of variation trumpet length, both orifice diameter (do) and

trumpet angle θt are held constant as lt is varied. Figure 5.5 indicates that with an

increase in lt/do, dimensionless thickness t* decreases about 14%, discharge coefficient

shows virtually no change, whereas spray cone half angle decreases about 9% for the

range of dimensionless trumpet length considered here. For constant flow rate across the

atomizer, the pressure drop across the atomizer, and the axial velocity at the end of the

exit orifice are not expected to change significantly with a change in the trumpet length.

As such the discharge coefficient shows very little change with trumpet length.

However, as the trumpet length increases, to keep the flow cross-sectional area relatively

constant, the film thickness decreases. Also, as the trumpet diameter increases, the

principle of conservation of angular momentum dictates that the liquid swirl velocity

would decrease as it moves along the trumpet length. Therefore, the spray cone angle is

seen to decrease with increasing trumpet length. Equations 5.6-5.8 outlined above

provide atomizer performance parameters for inviscid flow through the atomizer.

Although the trumpet length does not directly appear in Equations 5.6-5.8, it is noted that

for a fixed trumpet angle, Xt is a function of the trumpet length. The effects of variation

in trumpet length for inviscid flow are plotted in Figure 5.5 with dashed lines. Similar to

Page 91: XUE JIANQING

84

lt/doDim

ensi

onle

ssTh

ickn

ess

t*an

dD

isch

arge

Coe

ffici

entC

d

Spr

ayC

one

Hal

fAng

leθ

0.3 0.4 0.5 0.6 0.7 0.8 0.9

0.1

0.2

0.3

0.4

30

35

40

45

50

55

60

t* (CFD)t* (Equation 5.7)Cd (CFD)Cd (Equation 5.2)Cd (Equation 5.5)θ (CFD)θ (Equation 5.8)

Figure 5.5 Variation of atomizer performance parameters with ratio of trumpet length to

orifice diameter lt/do

Page 92: XUE JIANQING

85

inviscid flow solutions for variations in inlet slot angle and trumpet angle, the values for

film thickness and discharge coefficient predicted with computational model are higher

than analytical solutions and those of spray angle are lower than analytical solutions. The

results for discharge coefficient using modified Equation 5.5 are significantly closer to

computational results.

5.5 Effect of Variation in Swirl Chamber Convergence Angle θc

In this case, the geometry of the atomizer is one without trumpet and all

geometric parameters have been kept fixed except for θc. As seen from Figure 5.6, for the

atomizer configuration considered in this study, with an increase in θc from 45o to 90o,

dimensionless thickness t* increases about 16%, Cd increases about 34% and spray cone

half angle θ decreases about 9%. Earlier studies (Jeng et al 1998, Xue et al 2002) have

shown that the flow field inside a simplex atomizer has a region of recirculating flow.

The majority of the liquid entering from the inlet slots flows through a region close to the

liquid gas interface and enters the orifice. The location and size of the recirculating flow

affects the axial velocity variations in the swirl chamber and consequently in the exit

orifice. As the convergence angle changes, the re-circulating region in the swirl chamber

can change considerably. This is illustrated in Figures 5.7(a) and 5.7(b) which show flow

streamlines for two atomizer geometries with convergence angle 60o and 90o,

respectively. As the flow structure changes, the velocity variation in the atomizer is

altered and consequently the film thickness at the exit, spray cone half angle, and

Page 93: XUE JIANQING

86

Spin Chamber Convergence Angle θcDim

ensi

onle

ssTh

ickn

ess

t*an

dD

isch

arge

Coe

ffici

entC

d

Spr

ayC

one

Hal

fAng

leθ

40 50 60 70 80 900.2

0.3

0.4

36

38

40

42

t*

Cdθ

Figure 5.6 Variation of atomizer performance parameters with convergence angle θc

Page 94: XUE JIANQING

87

(b) θc = 90o

Exit

AirC

ore

Inlet

(a) θc = 60o

Exit

AirC

ore

Inlet

Figure 5.7 Streamlines for flow inside the atomizer for different θc. Inlet slot angle β=900.

Page 95: XUE JIANQING

88

discharge coefficient change. Although the spray angle is lowest and the film thickness

is largest at convergence angle of 90o, this geometry may be preferred in some cases as it

is easier and less expensive to manufacture compared to the geometry with smaller

convergence angle. It is noted that the effects due to changes in swirl chamber

convergence angle can not be predicted by inviscid analysis.

5.6 Summary

By using a computational model based on the Arbitrary-Lagrangian-Eulerian

method, the flow in simplex atomizers was predicted. The effect of four geometric

parameters on the atomizer performance was investigated. The four geometric

parameters considered were the inlet slot angle, trumpet angle and length, and the swirl

chamber convergence angle. The atomizer performance was shown in terms of variations

of dimensionless film thickness, spray cone half angle, and discharge coefficient. Results

indicated that increase in inlet slot angle results in lower film thickness and discharge

coefficient and higher spray cone angle. The swirl chamber convergence angle has an

opposite effect on performance parameters. With an increase in convergence angle, film

thickness and discharge coefficient increase whereas the spray cone angle decreases. The

trumpet angle has very little influence on discharge coefficient. However, the film

thickness decreases and spray cone angle increases with increasing trumpet angle. The

discharge coefficient is insensitive to trumpet length whereas both the spray cone angle

and the film thickness decrease with increasing trumpet length. The effect of the

Page 96: XUE JIANQING

89

geometric parameters has not been reported in the literature and as such the results

presented here will provide useful guidance in designing simplex atomizers.

Page 97: XUE JIANQING

90

CHAPTER 6 NON-NEWTONIAN FLUID FLOW IN PRESSURE-

SWIRL ATOMIZERS

6.1 Introduction

Although the fluid flow in the pressure-swirl atomizers used in combustors

involves Newtonian fluids, the fluid to be atomized is likely to have non-Newtonian

viscous characteristics in many other atomization applications, such as paint spray,

agricultural spray, pharmaceutical and food processing. Compared to the large number of

publications on the Newtonian fluid sprays from simplex nozzles, only two studies by

Som and co-workers on non-Newtonian fluids are available (Som, 1983, Biswas and

Som, 1986). They determined both theoretically and experimentally the discharge

coefficient and spray cone angle of swirl nozzle using a time-independent purely viscous

power-law non-Newtonian fluid. In their papers, theoretical predictions were made

through an approximate analytical solution of the hydrodynamics of flow inside the

nozzle and experiments were carried out with aqueous solutions of CMC powder of

various concentrations as the working fluid. Their experiments were restricted to shear-

thinning fluids. Their experiment data indicated that at fixed values of generalized

Reynolds number based on tangential velocity at the inlet, ReGi, an increase in the power

index n increases the discharge coefficient and the power index n has a negligible

influence on the spray cone angle. Biswas’s experimental result showed that for fixed

values of ReGi, an increase in the power index n decreases the spray cone angle and the

power index n has a negligible influence on the discharge coefficient. The experiment

were conducted on two atomizer geometries one with pure tangential inlet (Som) and

Page 98: XUE JIANQING

91

with combined axial and tangential entry (Biswas and Som). Note that in their study,

with the decrease in value of n, the condition of constant ReGi requires much higher

tangential velocity at the inlet. That means the mass flow rate varies considerably.

Furthermore, the fluid consistency (K) of the fluids used was significantly higher for low

power law index. These limitations make direct comparison of our predictions with their

results difficult.

We have extended the CFD model developed earlier for Newtonian flow by using

a two-parameter non-Newtonian power law fluid model, the flow of time-independent

purely viscous power-law fluids in pressure-swirl atomizers is investigated.

6.2 Mathematical Model

The generalized Newtonian fluid model is (Bird et al., 2002):

γηητ &−=∇+∇−= + ))(( vv (6.1)

Where,

)(γηη &= (6.2)

+∇+∇= )( vvγ& (6.3)

According to two-parameter power law model:

1−= nmγη & (6.4)

Where n is power index (fluid behavior index) and m is fluid consistency index.

So the shear stress tensor can be expressed as,

nmγτ &−= (6.5)

For n < 1, the fluid is shear-thinning non-Newtonian fluid; n = 1, is Newtonian

fluid; for n > 1, the fluid is shear-thickening non-Newtonian fluid. When Equation 6.5 is

Page 99: XUE JIANQING

92

used to compute the shear stress term in the Navier-Stokes equations, the model can be

used to handle both Newtonian and non-Newtonian fluid flow.

6.3 Influence of power-law index on the performance of atomizers with constant

flow rate across the atomizer

The effect of different power law index on atomizer performance was investigated

keeping the mass flow rate across the atomizer constant. In most applications the

atomizers operate in a flow regime where the Reynolds number has little effect on the

atomizer performance and resulting spray characteristics. To make sure that we consider

an appropriate flow rate for such a regime, the results were obtained with constant inlet

flow rates of 300lbm/h and 400lbm/h. The range for the power index n is from 0.7 to 1.3.

The atomizer performance is shown in terms of variations of dimensionless film

thickness at exit, spray cone half angle, and discharge coefficient.

Figure 6.1 shows the geometry of atomizer used in this study. The final grid is

shown in Figure 6.2. Figure 6.3 gives the simulation result of typical velocity field and

pressure distribution.

Figure 6.4 shows the variation of dimensionless film thickness at the exit with the

power index and indicates that as power index increases, the film thickness increases.

Due to the increase in shear stress with the power index, the air core diameter at the exit

should decrease. The variation of discharge coefficient with the power index is given in

Figure 6.5. The discharge coefficient increases slightly with increase n power index.

The increase in film thickness will decrease the exit velocity and consequently decrease

the pressure drop across the atomizer and increase the discharge coefficient. Figure 6.6

Page 100: XUE JIANQING

93

Radial Direction (r)A

xial

Dire

ctio

n(z

)

Exit

Air

core

Figure 6.1 Initial grid

Page 101: XUE JIANQING

94

Radial direction (r)A

xial

Dire

ctio

n(z

)

Exit

Air

core

Figure 6.2 Final grid

Page 102: XUE JIANQING

95

Figure 6.3 Streamline and pressure distribution

Axialdirection

(z)

Radial Direction (r)

P1.17198E

+061.09096E

+061.00994E

+0692891584789376687268585160483052380844278736176628074419972311870237680.3

ExitA

ircore

Page 103: XUE JIANQING

96

Power index (n)

Dim

ensi

onle

ssTh

ickn

ess

(t* )

0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.40.3

0.31

0.32

0.33

0.34

0.35

0.36

0.37

400 lbm/h300 lbm/h

Figure 6.4 Variation of film thickness with power index

Page 104: XUE JIANQING

97

Power index (n)

Dis

char

geco

effic

ient

(Cd)

0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.40.36

0.37

0.38

400 lbm/h300 lbm/h

Figure 6.5 Variation of discharge coefficient with power index

Page 105: XUE JIANQING

98

Power index (n)

Spr

ayco

nean

gle

(θ)

0.6 0.8 1 1.2 1.430

31

32

33

34

35

400 lb/h300 lb/h

Figure 6.6 Variation of Spray half angle with power index

Page 106: XUE JIANQING

99

Radial Direction (r)A

xial

dire

ctio

n(z

)

Exit

Air

core

AA

B B

CC

Figure 6.7 Locations for the velocity inside nozzle study

Page 107: XUE JIANQING

100

r/rSA

xial

Vel

ocity

(m/s

)0 0.2 0.4 0.6 0.8 1

0

2

4

6

8

n=1.3n=0.7n=1.0

Locaion A-A

Figure 6.8 Axial velocity profiles at the swirl chamber

Page 108: XUE JIANQING

101

r/rSA

xial

Vel

ocity

(m/s

)0.4 0.6 0.8 10

10

20

30

40

50

n=1.3n=0.7n=1.0

Locaion B-B

Figure 6.9 Axial velocity profiles at the orifice

Page 109: XUE JIANQING

102

r/rSA

xial

Vel

ocity

(m/s

)0.6 0.8 1

0

10

20

30

40

50

n=1.3n=0.7n=1.0

Locaion C-C

Figure 6.10 Axial velocity profiles at the exit

Page 110: XUE JIANQING

103

r/rSS

wirl

Vel

ocity

(m/s

)0 0.2 0.4 0.6 0.8 1

0

10

20

30

40

50

n=1.3n=0.7n=1.0

Locaion A-A

Figure 6.11 Swirl velocity profiles at the swirl chamber

Page 111: XUE JIANQING

104

r/rSS

wirl

Vel

ocity

(m/s

)0.4 0.6 0.8 10

10

20

30

40

50

n=1.3n=0.7n=1.0

Locaion B-B

Figure 6.12 Swirl velocity profiles at the orifice

Page 112: XUE JIANQING

105

r/rSS

wirl

Vel

ocity

(m/s

)0.6 0.8 1

0

10

20

30

40

50

n=1.3n=0.7n=1.0

Locaion C-C

Figure 6.13 Swirl velocity profiles at the exit

Page 113: XUE JIANQING

106

indicates the spray cone angle decreases when power index increases. This behavior is

consistent with measurements of Som (1983) and Biswas (1986) of discharge coefficient

and spray cone angle with variation in power index n. However, in their experiment, the

swirl and axial velocity decreased with increase in n whereas here the mass flow rate is

constant.

The axial and swirl velocity profiles at three different locations inside the

atomizer are investigated here. The locations are shown in Figure 6.7, and include a

location in the swirl chamber (A-A), in the orifice (B-B) and at the orifice exit (C-C).

The velocity profiles for three different power index values, viz. n= 0.7, 1 and 1.3, are

given in Figure 6.8-6.13.

Figure 6.8 shows that in the swirl chamber, the axial velocity is larger near the

wall and near the liquid/gas interface. The axial velocities are relatively weak in the

middle. Such velocity variations have also been observed in experiments (Wang et al

1999). Among the three fluids, the shear-thickening fluid (n= 1.3) shows the largest

variation in the velocity magnitude throughout the flow cross-section. Figure 6.9

indicates that at the orifice, the profiles of the axial velocities are quite flat except for the

zone near the wall. The velocity gradients for the n = 0.7 and 1 are greater than that of

the n = 1.3 in the zone near the wall but the shear-thickening fluid has greater maximum

axial velocity. The axial velocity profiles at the exit are shown in Figure 6.10. At the

exit, the locations of the maximum axial velocities are more close to the wall than the

Page 114: XUE JIANQING

107

interface. Comparing to the axial velocities at the orifice, the average values become

larger due to the decrease of the film thickness.

The swirl velocity profiles at the swirl chamber are given in Figure 6.11. All

three profiles show the typical free-vortex swirl flow pattern. The shear-thickening fluid

has slightly smaller velocity magnitude. Figure 6.12 indicates that at the orifice, the

shear-thinning fluid and Newtonian fluid have the combined vortex structure, a solid

vortex in the inner region connected with a free vortex in the outer region and the

velocity variation in the shear-thickening fluid closely resembles a free vortex structure.

The swirl velocity profiles at the exit are shown in Figure 6.13. The flow patterns at the

exit are similar to those at the orifice.

6.4 Effect of atomizer constant on the performance of atomizers for both

Newtonian and non-Newtonian fluids

In this section, the influence of atomizer constant on the performance of atomizers

for both Newtonian and non-Newtonian fluid is investigated. The working condition is

constant pressure drop throughout the atomizer. Three different power index values, viz.

0.8, 1.0 and 1.2, are considered here.

The variation of dimensionless film thickness with atomizer constant is shown in

Figure 6.14. As atomizer constant K changes from 0.2 to 0.6, the film thickness increases

about 50% for n = 0.8 and 1, and about 46% for n = 1.2. Increase in K corresponds to

increase in inlet area and consequently a decrease in swirl velocity. Therefore, with

Page 115: XUE JIANQING

108

Atomizer Constant K

Dim

ensi

onle

ssTh

ickn

ess

t*

0.2 0.3 0.4 0.5 0.60.2

0.25

0.3

0.35

0.4

t* (n = 0.8)t* (n = 1.0)t* (n = 1.2)

Figure 6.14 Variation of film thickness with atomizer constant

Page 116: XUE JIANQING

109

Atomizer Constant K

Dis

char

geC

oeffi

cien

tCd

0.2 0.3 0.4 0.5 0.60.2

0.25

0.3

0.35

0.4

0.45

Cd (n = 0.8)Cd (n = 1.0)Cd (n = 1.2)

Figure 6.15 Variation of discharge coefficient with atomizer constant

Page 117: XUE JIANQING

110

Atomizer Constant K

Spr

ayC

one

Hal

fAng

leθ

(deg

)

0.2 0.3 0.4 0.5 0.632

34

36

38

40

42

44

46

θ (n = 0.8)θ (n = 1.0)θ (n = 1.2)

Figure 6.16 Variation of spray cone angle with atomizer constant

Page 118: XUE JIANQING

111

increasing K, the film thickness increases. We note that for given atomizer constant

value, as power index increases, the film thickness increases.

Figure 6.15 gives variation of discharge coefficient with atomizer constant. It

indicates that the discharge coefficient increases about 60% for n = 0.8 and 1, and about

65% for n = 1.2. With increasing K, the decrease in swirl results in thicker fluid film.

With a constant pressure drop across the atomizer, the thicker fluid film leads to higher

mass flow rate and hence increases in the discharge coefficient. Also note that the

discharge coefficient increases with power index for a given atomizer constant.

The variation of spray cone half angle with atomizer constant is shown in Figure

6.16. As atomizer constant varies from 0.2 to 0.6, the spray angle approximately changes

from 44 to 36 for n = 0.8 and n = 1.0, from 44 to 32 for n = 1.2. The spray angle

decreases with increase in power index for a given atomizer constant. Once again, this is

to be expected as lower swirl with higher K would lead to smaller spray cone angle.

6.5 Summary

Using a two-parameter power law model, the flow of power-law fluids inside

pressure-swirl atomizers was investigated. Under constant mass flow rate through the

atomizer, the dimensionless film thickness at the exit increases and the spray cone angle

decreases with increase in the power index. The discharge coefficient exhibits only a

small influence of power-law index. Detailed flow pattern inside the atomizer for shear-

thinning, Newtonian and shear-thickening fluids were investigated.

Page 119: XUE JIANQING

112

Under constant pressure drop across the atomizer, the dimensionless film

thickness at the exit and the discharge coefficient increase, and the spray cone angle

decreases with increase in the atomizer constant. For the given atomizer constant value,

the dimensionless film thickness at the exit and the discharge coefficient increase, and the

spray cone angle decreases with increase in the power index.

Page 120: XUE JIANQING

113

CHAPTER 7 CONCLUSIONS AND RECOMMENDATIONS

7.1 Conclusions

A 2D CFD code for simulation of flow in pressure-swirl atomizers was

developed. The code can track the free interface between gas and liquid accurately via

ALE method. The code incorporated a method for grid generation at each time step. The

code was validated by comparison of computational predictions with experimental data

for large scale prototype and with semi-empirical correlations at small scale. The CFD

predictions of spray angle were closer to the experimental measurements on large scale

prototype atomizer than those obtained with semi-empirical correlations for all the cases

considered. The CFD prediction agreed well with experimental data of the film thickness

at the exit and the pressure drop across the nozzle. The agreement was found to be

substantially better than the predictions using correlations.

Among the geometric parameters considered here, atomizer constant was found to

be the most dominant parameter. With the atomizer constant (K) increasing from 0.1 to

0.6, dimensionless film thickness at the exit increases by about 0.15; the discharge

coefficient increases by about 0.2; and the spray cone half angle decreases by about 25%.

With other parameters, the dimensionless film thickness and spray cone half angle

variations are not always monotonic and exhibit optimal conditions (small thickness and

large cone angle) at certain parameter values. This may be of interest in designing

atomizers.

Page 121: XUE JIANQING

114

The effect on the performance of the nozzle with variations with Ls/Ds, lo/do, and

Ds/do were considered under constant inlet flow rate as well as constant pressure drop

across the atomizer. The code could predict these results easily, whereas the present

correlations can not provide such guidance for nozzle designers.

There are several parameters that are not considered in the correlations. Four

such parameters were investigated here, viz., inlet slot angle, trumpet angle, trumpet

length and convergent angle. The result illustrated that the film thickness increases with

decreasing inlet angle, trumpet angle and trumpet length and increasing convergent angle;

discharge coefficient increases with decreasing inlet angle and increasing convergent

angle; spray cone angle increases with increasing inlet angle and trumpet angle and

decreasing trumpet length and convergent angle; trumpet angle and trumpet length have

little effect on discharge coefficient.

The code was extended to model the flow of time-independent purely viscous

power-law fluids in pressure-swirl atomizers. A range of power-law index from 0.7 to

1.3 was considered. Detailed flow patterns inside the atomizer for shear-thinning,

Newtonian and shear-thickening fluids were investigated. The effect of variation in

atomizer constant on the atomizer performance was studied for fluids with different

power index values for a fixed pressure drop across the atomizer. The results indicated

that the dimensionless film thickness at the exit and the discharge coefficient increase,

and the spray cone angle decreases with increase in the atomizer constant. For a given

Page 122: XUE JIANQING

115

atomizer constant, the dimensionless film thickness at the exit and the discharge

coefficient increase, and the spray cone angle decreases with increase in the power index.

7.2 Recommendations for future work

For the atomizers with very short swirl chamber and two inlet slots, the flow may

be have three-dimensional features that can not be determined by the present

axisymmetric approach. In order to simulate such flows, a fully three dimensional

treatment may be needed. The present code can be extended to model three dimensional

flows.

In this study, the performance of atomizers was expressed indirectly in term of

dimensionless film thickness at the exit, the spray cone half angle and discharge

coefficient. If a model to simulate the break-up of the liquid sheet can be developed and

is integrated in current code, the spray characteristics, such as mean drop size, drop size

distribution, can be predicted directly from atomizer geometry and input flow parameters.

Such a break-up model, combined with the present code, will represent a major advance

in developing predictive tools for fuel injection.

Page 123: XUE JIANQING

116

BIBLIOGRAPHY

Amsden, A. A., and Ruppel, H. M., “SALE-3D: A Simplified ALE Computer Program

for Calculating Three-Dimensional Fluid Flow”, LA-8905, Los Alamos National

Laboratory, 1981.

Baldwin, B. S., and Lomax, H., “Thin-Layer Approximation and Algebraic Model for

Separated Turbulent Flows,” AIAA paper 78-257, Huntsville, AL, 1978.

Benjamin, M. A., Jeng, S. M., and Jog, M. A., “Comparison of Simplex Atomizer

Correlations with Detailed CFD and Experimental Data,” Proceedings of the 10th Annual

Conference on Liquid Atomization and Spray Systems, Inst. For Liquid Atomization and

Spray Systems-Americas, Irvine, CA, 1997, pp. 45-49.

Benjamin, M. A., Mansour, A., Samant, U. G., Jha, S., Liao, Y., Harris, T. and Jeng, S.

M., “Film Thickness, Droplet Size Measurements and Correlation for Large Pressure-

Swirl Atomizers”, International Gas Turbine & Aeroengine Congress & Exhibition,

Stockholm, Sweden, June2-June5, 1998.

Bird, R. B., Stewart, W. E., and Lightfoot, E. N., Transport Phenomena, Second Edition,

John Wiley Publishing, 2002.

Biswas, G. and Som, S. K., “Coefficient of Discharge and Spray Cone Angle of a

Pressure Nozzle with Combined Axial and Tangential Entry of Power-Law Fluids”,

Applied Scientific Research, 43, pp 3-22, 1986.

Page 124: XUE JIANQING

117

Chung, I., and Presser, C., “Fluid Property Effects on Sheet Disintegration of a Simplex

Pressure-Swirl Atomizer”, Journal of Propulsion and Power, Vol. 17, No.1, 2001.

Datta, A. and Som, S. K., “Numerical Prediction of Air Core Diameter, Coefficient of

Discharge and Spray Cone Angle of a Swirl Spray Pressure Nozzle,” International

Journal of Heat and Fluid Flow, Vol. 21, pp. 412-419, 2000.

De Keuelaere, H. J. K., “The Internal Flow in a Swirl Atomizer Nozzle”, MSc. Thesis,

Department of Mechanical Engineering, University of Manchester Institute of Science

and Technology, 1995.

Deng, Z., “Numerical Investigation of Non-Spherical Droplet Dynamics in Viscous

Convective Flows”, Ph.D. Dissertation, University of Tennessee, Knoxville, 1991.

Dodge, L. G. and Biaglow, J. A., “Effect of Elevated Temperature and Pressure on

Sprays from Simplex Swirl Atomizers”, Journal of Engineering for Gas Turbines and

Power, January 1986, Vol. 108, pp. 209-215.

Dodge, L. G., Rhodes, D. J., Reitz, R. D., “Drop-size measurement techniques for sprays:

comparison of Malvern laser-diffraction and Aerometrics phase/Doppler”, Applied

Optics, Vol. 26, No. 11/1 June 1987, pp.2144-2154.

Dodge, L. G. and Schwalb, J. A., “Fuel Spray Evolution: Comparison of Experiment and

CFD Simulation of Nonevaporating Spray”, Journal of Engineering for Gas Turbines

and Power, January 1989, Vol. 111, pp. 15-23.

Elkotb, M. M., Rafat, N. M., and Hanna, M. A., “The Influence of Swirl Atomizer

Geometry on the Atomization Performance”, ICLAS 78, pp 109-115, 1978.

Page 125: XUE JIANQING

118

Fang, Pingping, “Numerical investigation of laminar forced convection in Newtonian and

non-Newtonian flows in eccentric annuli”, Ph. D. thesis, Department of Mechanical,

Industrial and Nuclear Engineering, University of Cincinnati, 1998.

Giffen, E., and Muraszew, A., The Atomization of Liquid Fuels, John Wiley and Sons Inc.,

1953.

Herpfer, D. C. and Jeng, San-Mou, “Comparisons between SPPPIVS and PDPA Results

for a Simplex Atomizer”, 34th Aerospace Sciences Meeting & Exhibit, January 15-18,

1996, Reno, NV.

Herpfer, D. C. and Jeng, San-Mou, “Planar Measurements of Droplet Velocities and

Sizes within a Simplex Atomizer”, AIAA Journal, Vol. 35, No.1, January 1997.

Hirt, C. W., Amsden, A. A., and Cook, J. L., et al, “A Lagrangian Method for Calculating

the Dynamics of an Incompressible Fluid with Free Surface,” Journal of Computational

Physics, No. 5, 1970, pp. 103-124.

Hirt, C. W., Amsden, A. A., and Cook, J. L., “An Arbitrary Lagrangian-Eulerian

Computing Method for All flow Speeds,” Journal of Computational Physics, Vol. 14, No.

3, 1974, pp. 227-253.

Holtzclaw, D., Sakman, A. T., Jeng, S. M., Jog, M. A. and Benjamin, M. A.,

“Investigation of Flow in a Simplex Fuel Nozzle”, AIAA paper No. 97-2970, 33rd

AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, Seattle, WA, 1997.

Page 126: XUE JIANQING

119

Holtzclaw, J. D., “Parametric study of the internal flow field in a simplex fuel nozzle

using PIV characterization”, MSc. Thesis, Department of Aerospace Engineering and

Engineering Mechanics, University of Cincinnati, 1998.

Horvay, M. And Leuckel, W., “LDA-Measurements of Liquid Swirl Flow in Converging

Swirl Chambers with Tangential Inlets”, Proc. 2nd Int. Symp. LDA Lisbon. 1985, Paper

14.

Horvay, M. And Leuckel, W., “Experimental and Theoretical Investigation of Swirl

Nozzles for Pressure-Jet Atomization”, German Chem. Eng., Vol. 9, 1986, pp 276-283.

Hwang, Y. H., “Solution of the Steady Navier-Stokes Equations by the Pressure

Correction Method in the ALE Grid”, Numerical Heat Transfer, Part B, vol. 23, pp. 237-

255, 1993.

Jeng, S. M., Jog, M. A., Benjamin, M. A., “Computational and Experimental Study of

Liquid Sheet Emanating from Simplex Fuel Nozzle”, AIAA Paper No. 97-0796, AIAA

Journal, Vol. 36, No. 2, pp. 201-207, 1998.

Jones, A. R., “Design Optimization for a Large Pressure-Jet Atomizer for Power Plant,”

Proceedings of the 2nd International conference on Liquid Atomization and Spray

Systems, Madison, WI, 1982, pp.181-185.

Kutty, S. P., Narasimhan, M., and Narayanaswany, K., “Design and Prediction of

Discharge Rate, Cone Angle, and Air Core Diameter of Swirl Chamber Atomizers”,

Proceedings of the 1st International Conference on Liquid Atomization and Spray

Systems, The Fuel Society of Japan, Tokyo, pp 93-100, August, 1978.

Page 127: XUE JIANQING

120

Lefebvre A. H., Gas Turbine Combustion, Hemisphere, Washington, D. C., 1983.

Lefebvre A. H., Atomization and Sprays, Hemisphere, Washington, DC, 1989.

Liao, Y., Sakman, A. T., Jeng, S. M., Jog, M. A., and Benjamin, M. A., “A

Comprehensive Model to Predict Simplex Atomizer Performance,” Journal of

Engineering for Gas Turbines and Power, Vol. 121, No. 2, 1999, pp. 285-294.

Ma, Zhanhua, “Investigation on the Internal Flow Characteristics of Pressure-swirl

Atomizer”, Ph.D. Thesis, Department of Aerospace Engineering and Engineering

Mechanics, University of Cincinnati, 2001.

McDonell, V. G., Samuelsen, G. S., Wang, M. R., Hong, C. H. and Lai, W. H.,

“Interlaboratory Comparison of Phase Doppler Measurements in a Research Simplex

Atomizer Spray” 28th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit,

July 6-8, 1992, Nashville, TN, AIAA 92-3233.

Nonnenmacher, S., and Piesche, M., “Design of Hollow Cone Pressure Swirl Nozzles to

Atomize Newtonian Fluids”, Chemical Engineering Science, pp 4339-4348, 2000.

O’Rourke, P. J. and Amsden, A. A., “Implementation of a Conjugate Residual Iteration in

the KIVA Computer Program”, LA-10849-MS, 1986.

Okamoto, T., Takagi, T., Kaji, T., Shimazaki, K. and Nakanishi, K., “Studies on the

Behavior of Droplets and the Air Flow in a Hollow-Cone Spray”, Transactions of the

ASME, Vol. 120, September 1998.

Page 128: XUE JIANQING

121

Ren, W. M., and Nally, J. F., “Computations of Hollow-Cone Sprays from a Pressure-

Swirl Injector”, SAE Paper 982610, International Fall Fuels and Lubricants Meeting and

Exposition, San Francisco, CA, October 19-22, 1998.

Rizk, N. K. and Lefebvre, A. H., “Spray Characteristics of Simplex Swirl Atomizers”,

Progress in Aeronautics and Astronautics, Vol. 95, 1984, pp.563-580.

Rizk, N. K. and Lefebvre, A. H., “Internal Flow Characteristics of Simplex Swirl

Atomizers, Journal of Propulsion and Power, Vol. 1, No. 3, May-June, 1985, pp 193-

199.

Rizk, N. K. And Lefebvre, A. H., “Influence of Liquid Properties on the Internal Flow

Characteristics of Simplex Swirl Atomizers, Atomization and Spray Technology, Vol. 2,

No. 3, May-June, 1986, pp 219-233.

Sakman, A. T., A Numerical Parametric Study of Simplex Fuel Nozzle Internal Flow and

Performance, M.S. Thesis, University of Cincinnati, 1998.

Sakman, A. T., Jha, S., Jog, M.A, Jeng, S.M. and Benjamin, M. A., “A Numerical

Parametric Study of Simplex Fuel Nozzle Internal Flow and Performance”, 34th

AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, July13-15, 1998,

Cleveland, OH.

Sakman, A. T., Jog, M. A., Jeng, S. M., andBenjamin, M. A., “Parametric Study of

Simplex Fuel Nozzle Internal Flow and Performance,” AIAA Journal, Vol. 38, No. 7,

2000, pp. 1214-1218.

Page 129: XUE JIANQING

122

Som, S. K., “Theoretical and Experimental Studies on the Coefficient of Discharge and

Spray Cone Angle of a Swirl Spray Pressure Nozzle Using a Power-Law Non-Newtonian

Fluid”, Journal of Non-Newtonian Fluid Mechanics, 12, pp 39-68,1983.

Steinthorsson, E., and Lee, D. M., “Numerical Simulations of Internal flow in a Simplex

Atomizer,” 8th International on Liquid Atomization and Spray Systems, Pasadena, CA,

July 2000.

Stoer, J. and Bulirsch, R; translated by Bartels, R., Gautschi, W., and Witzgall, C.,

Introduction to Numerical Analysis, Springer-Verlag, New York, 1993.

Suyari, M. and Lefebvre, A. H., “Film Thickness Measurements in Simplex Swirl

Atomizer”, Journal of Propulsion, Vol. 2, No.6, pp.528-533, 1986.

Vanderwege, B. A. and Hochgreb, S., “The Effect of Fuel Volatility on Sprays from

High-Pressure Swirl Injectors”, 27th Symposium (International) on Combustion, The

Combustion Institute, 1998, pp. 1865-1871.

Wang, D., and Ma, Z et al, “Experimental Study on Large-Scale Simplex Nozzle,” AIAA

paper NO. 99-2401, 35rd AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit,

1999.

Wilcox, D. C., Turbulence Modeling of CFD, DCW Industries, Inc, La Canada, CA, 1993.

Xue, J., Jog, M. A., Jeng, S. M., “Numerical Simulation of the Non-Newtonian Flow

inside a Pressure-Swirl Atomizer Using Time-Independent Power-Law”, ILASS

Americas, 17th Annual Conference on Liquid Atomization and Spray Systems, Arlington,

VA, May 2004

Page 130: XUE JIANQING

123

Xue, J., and Jog, M. A. et al, “Computational Model to Predict Flow in Simplex Fuel

Atomizer,” AIAA paper NO. 2002-3710, 38rd AIAA/ASME/SAE/ASEE Joint Propulsion

Conference & Exhibit, Indianapolis, IN, July, 2002.

Xue, J., and Jog, M. A. et al, “Influence of Geometry on the Performance of Simplex

Nozzles under Constant Pressure Drop,” ILASS Americas, 15th Annual Conference on

Liquid Atomization and Spray Systems, Madison, WI, May 2002.

Xue, J., Jog, M. A. et al “Effect of Geometric Parameters on Simplex Atomizer

Performance”, AIAA Journal, Vol. 42, No. 11, November, 2004

Yuel, A. J. and Widger, I. R., “Swirl Atomizer Operating at High Water Pressure”, Int. J.

Mech. Sci. 1996, Vol. 38, No. 8-9, pp. 981-999.

Yule, A. J., and Chinn, J. J., “Pressure Swirl Atomizer Internal flow and Performance,”

Proceedings of the 10th Annual conference on Liquid Atomization and Spray Systems, Inst.

For Liquid Atomization and Spray Systems-Americas, Irvine, CA, 1997, pp.205-209.

Page 131: XUE JIANQING

124

APPENDIX INVISCID ANALYSIS FOR PRESSURE SWIRL

ATOMIZERS

The analysis closely follows the development by Giffen and Muraszev (1953)

with appropriate additions for the new geometric parameters considered here. First we

consider the effect of inlet slot angle for an atomizer without trumpet. Here we consider

a more general case where the inlet is not necessarily at r = rs (shown in Figure A-1).

Figure A-1 Schematic of pressure swirl atomizer

In pressure swirl atomizer, the liquid flow can be considered as a spiral motion as

a result of an axial flow being superimposed on a free vortex.

For a free vortex:

Gas/Liquid interface

ri

roa ro

β

rs

Page 132: XUE JIANQING

125

== ii rwwr constant (A-1)

where:

pi A

Qw βsin= (A-2)

According to Bernoulli’s equation, the total pressure (injection pressure) in the

liquid flowing through the orifice is:

=++=∆ 22

21

21 wupP ρρ constant (A-3)

Where p is the static pressure.

At the air core, p = 0. So,

22

21

21

oaoa wuP ρρ +=∆ (A-4)

The axial velocity can be calculated as:

)( oaooa AA

Qu−

= (A-5)

According to Eq. (A-1) and Eq. (A-2),

oap

ioa rA

rQw

βsin= (A-6)

Substituting into Eq. (A-3) u and w from Eq. (A-5) & (A-6),

⎥⎥

⎢⎢

⎟⎟⎠

⎞⎜⎜⎝

⎛+⎟⎟

⎞⎜⎜⎝

⎛−

=∆22

sin21

oap

i

oao rAQr

AAQP

βρ (A-7)

From the definition of Cd, we get,

5.02

⎟⎟⎠

⎞⎜⎜⎝

⎛ ∆=

ρPACQ od (A-8)

Substitution of Q from Eq. (A-8) into Eq. (A-7),

Page 133: XUE JIANQING

126

β22

21

22 sin1)1(

11⎟⎟⎠

⎞⎜⎜⎝

⎛+

−=

s

i

d rr

XKXC (A-9)

Where,so

p

o

oa

rrA

KAA

== 1,

Assume that the size of the air core in the orifice will be such as to give maximum

flow. That means Cd is maximum:

β22

2

32

1

2

sin2

)1(0

1

⎟⎟⎠

⎞⎜⎜⎝

⎛−=⇒=

⎟⎟⎠

⎞⎜⎜⎝

s

id

rr

XXK

dXC

d (A-10)

Substitute Eq. (A-10) into Eq. (A-9),

5.03

1)1(

⎥⎦

⎤⎢⎣

⎡+−

=XXCd

(A-11)

14KK

ddA

Kos

p π=⇒=Q (A-12)

Substitute Eq. (A-12) into Eq. (A-10),

βπ 22

2

322 sin)1(

32 ⎟⎟⎠

⎞⎜⎜⎝

⎛−=

s

i

rr

XXK (A-13)

From the definition of X, we get,

Xt −= 1* (A-14)

Following Giffen and Muraszew[12], for the spray cone half angle θ, we have,

Uw

oruw o

o

o == θθ sintan (A-15)

odod AC

QUUACQ =⇒=Q (A-16)

At the exit, the total angular momentum in the orifice,

Page 134: XUE JIANQING

127

∫−

=o

oa

r

rp

oaoio

p

io A

rrrQurA

Qrdrur β

πρβρπ sin

)(2sin)2( (A-17)

The total mass flow in the orifice,

)( oaoo AAu −ρ (A-18)

⇒ mean tangential velocity at the exit,

)(sin)(2

oaop

oaoio AAA

rrQrw

−−

=βπ (A-19)

Substituting Eq. (A-16) & Eq. (A-19) into Eq. (A-15), we get,

⎟⎟⎠

⎞⎜⎜⎝

+=

s

id

rr

XKC

)1(sin)2/(

sinβπ

θ (A-20)

Now we consider an atomizer geometry with trumpet.

Figure A-2 Schematic of orifice with trumpet

Following the foregoing analysis, consider the air core at the trumpet end, we

have,

do

doa

dt

dta

Page 135: XUE JIANQING

128

ttatta AA

Quθcos)( −

= (A-21)

tap

ita rA

rQw

βsin= (A-22)

⎥⎥

⎢⎢

⎟⎟⎠

⎞⎜⎜⎝

⎛+⎟⎟

⎞⎜⎜⎝

⎛−

=∆⇒22

sincos)(2

1

tap

i

ttat rAQr

AAQP

βθ

ρ (A-23)

Substitution of Q from Eq. (A-8) into Eq. (A-23),

22

2

222

21sin1

)1(cos1

ds

i

ttttt

o

Crr

XKXAA

=⎥⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛+

−⎟⎟⎠

⎞⎜⎜⎝

⎛β

θ (A-24)

where,st

pt

t

tat rr

AK

AA

== ,

to

tt r

rXt θcos)1(*

⎟⎟⎠

⎞⎜⎜⎝

⎛−=⇒ (A-25)

For the spray cone half angle θ, approximately we have,

'θθθ += t (A-26)

where t

t

uw

='tanθ (A-27)

the mean velocity along with the trumpet direction at the trumpet end,

ttatt AA

Quθcos)( −

= (A-28)

At the trumpet end, the total angular momentum,

∫−

=t

ta

r

rp

tatitt

p

itt A

rrrQurA

Qrdrur β

θπρβθρπ sin

)(cos2sin)cos2( (A-29)

The total mass flow at the trumpet end,

Page 136: XUE JIANQING

129

)(cos tattt AAu −θρ (A-30)

⇒ mean tangential velocity at the trumpet end,

)(sin)(2

tatp

tatit AAA

rrQrw

−−

=βπ (A-31)

Substituting Eq. (A-28) & Eq. (A-31) into Eq. (A-27), we get,

ts

i

t

t

rr

KX

θβθ cossin)1(2

tan '⎟⎟⎠

⎞⎜⎜⎝

⎛−= (A-32)