Вы находитесь на странице: 1из 212

A Comparative Study Between Circular and Elliptical

Nozzle Holes on Natural Gas Combustion and Soot


Formation in a Direct Injection Engine

by

Charles Habbaky

A thesis submitted in conformity with the requirements


for the degree of Master of Applied Science
Department of Mechanical and Industrial Engineering
University of Toronto

c 2012 by Charles Habbaky


Copyright

Abstract
A Comparative Study Between Circular and Elliptical Nozzle Holes on Natural Gas
Combustion and Soot Formation in a Direct Injection Engine
Charles Habbaky
Master of Applied Science
Department of Mechanical and Industrial Engineering
University of Toronto
2012
The effects of changing nozzle hole patterns and hole geometry in a direct injection natural gas optically accessible engine was investigated. Six nozzles were studied having a
1 hole, 3 hole, and 9 hole pattern; each having either elliptical or circular hole geometries. Combustion images were taken with a high speed camera and the nozzles were
compared on the basis of their ignition delay time, rate of heat release, net heat release,
fuel utilization, gross indicated thermal efficiency, and particulate emissions. The best
performance in all categories was achieved by the 9 hole nozzles which was largely attributed to better fuel mixing as a result of its hole distribution. The elliptical hole
geometry exhibited characteristics of improved mixing mainly through reduced ignition
delay time and reduced elemental carbon emissions.

ii

Acknowledgements
First and foremost I would like to thank Professor James S. Wallace for giving me the
opportunity to work on this project. He was always there for me to offer his guidance and
support as well as provide his expertise in times of need which has taught me immensely.
It was truly an honour to have had the chance to work with him. These two years have
been one of the most influential times of life, both personally and professionally, and they
will always be remembered.
I would also like to thank everyone who has supported me during these two arduous
yet enormously constructive years. My labmates Silvio, Mark, Robert, and Phil were
always there to brainstorm or have some great conversations. Its the people I see every
day that made it such a great two years. Osmond Sargeant and Terry Zak are among
them. I can say without exaggerating that without their years of expertise this project
would not have been possible in the time I had. I can simply describe them as great people
to work with. Without a doubt, special thanks must go to everyone in the machine shop:
Ryan, Jeff, Gord, Fred, and Tai, who were always more than willing to help me and have
taught me so much over the course of my time.
Finally, none of this would be possible without the unconditional support of my family
and my darling Josephine, whos continuous love and understanding have kept me going
and allowed me to proudly be where I am today.

iii

Contents

Abstract

ii

Acknowledgments

iii

List of Tables

ix

List of Figures

Nomenclature

xvi

1 Introduction

1.1

The Drive Towards Alternative Fuels . . . . . . . . . . . . . . . . . . . .

1.2

Alternative Fuels for Transportation

. . . . . . . . . . . . . . . . . . . .

1.3

Natural Gas as an Alternative Fuel for Transportation . . . . . . . . . .

1.3.1

Attraction of Natural Gas . . . . . . . . . . . . . . . . . . . . . .

1.3.2

NGV Engines . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.3.3

Natural Gas Fuel Properties and Emissions . . . . . . . . . . . . .

Motivation and Objective . . . . . . . . . . . . . . . . . . . . . . . . . .

1.4

2 The Process of Injection and Ignition in a DING Engine


2.1

11

Direct Injection of Natural Gas . . . . . . . . . . . . . . . . . . . . . . .

11

2.1.1

Gas Jet Development and Structure . . . . . . . . . . . . . . . . .

12

2.1.2

Jet Penetration and Mixing . . . . . . . . . . . . . . . . . . . . .

14

iv

2.1.3
2.2

2.3

2.4

Jet Interactions and Confinement . . . . . . . . . . . . . . . . . .

17

Auto-ignition of Natural Gas . . . . . . . . . . . . . . . . . . . . . . . . .

19

2.2.1

The Effect of Temperature and Pressure . . . . . . . . . . . . . .

20

2.2.2

The Composition Effects of Natural Gas . . . . . . . . . . . . . .

21

2.2.3

Ignition Assist Methods and Hot Surface Ignition . . . . . . . . .

24

The Effect of Nozzle Hole Geometry on Natural Gas Combustion . . . .

26

2.3.1

Elliptical Jet Structure . . . . . . . . . . . . . . . . . . . . . . . .

26

2.3.2

Elliptical Nozzle Fuel Injectors . . . . . . . . . . . . . . . . . . . .

27

Soot Formation from Direct Injection Natural Gas . . . . . . . . . . . . .

29

2.4.1

Principles of Soot Formation . . . . . . . . . . . . . . . . . . . . .

30

2.4.2

Soot Formation in a DING Engine . . . . . . . . . . . . . . . . .

32

3 Experimental Apparatus
3.1

36

The CFR Engine and ERDL Combustion Bomb . . . . . . . . . . . . . .

36

3.1.1

Combustion Bomb Heating . . . . . . . . . . . . . . . . . . . . . .

39

3.1.2

Intake Air Heating . . . . . . . . . . . . . . . . . . . . . . . . . .

39

3.1.3

Intake Air Supercharging . . . . . . . . . . . . . . . . . . . . . . .

40

3.1.4

Engine Oil System . . . . . . . . . . . . . . . . . . . . . . . . . .

40

3.1.5

Engine Coolant System . . . . . . . . . . . . . . . . . . . . . . . .

40

Hot Surface Ignition System . . . . . . . . . . . . . . . . . . . . . . . . .

41

3.2.1

Glow Plug Position and Orientation . . . . . . . . . . . . . . . . .

41

3.2.2

Glow Plug Surface Temperature . . . . . . . . . . . . . . . . . . .

42

3.2.3

Glow Plug Shielding . . . . . . . . . . . . . . . . . . . . . . . . .

42

Natural Gas Injector . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

43

3.3.1

Nozzle Hole Geometry and Configuration . . . . . . . . . . . . . .

43

3.4

Fast Flame Ionization Detector . . . . . . . . . . . . . . . . . . . . . . .

45

3.5

Isokinetic Particulate Sampling System . . . . . . . . . . . . . . . . . . .

46

3.5.1

46

3.2

3.3

Sampling Probe Design . . . . . . . . . . . . . . . . . . . . . . . .


v

3.5.2
3.6

Particulate Filter System . . . . . . . . . . . . . . . . . . . . . . .

47

Data and Image Acquisition Systems . . . . . . . . . . . . . . . . . . . .

48

3.6.1

DAQ Measurements and Hardware . . . . . . . . . . . . . . . . .

48

3.6.2

Image Acquisition System . . . . . . . . . . . . . . . . . . . . . .

49

3.6.3

Timing Circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . .

50

4 Validation of Experimental Apparatus


4.1

4.2

4.3

52

Motored Engine Conditions . . . . . . . . . . . . . . . . . . . . . . . . .

52

4.1.1

Combustion Bomb Pressure . . . . . . . . . . . . . . . . . . . . .

53

4.1.2

Combustion Bomb Temperature . . . . . . . . . . . . . . . . . . .

55

4.1.3

Intake Flow Rate . . . . . . . . . . . . . . . . . . . . . . . . . . .

58

4.1.4

Gas Exchange and Blowby Losses . . . . . . . . . . . . . . . . . .

59

4.1.5

Particulate Sampling Flow . . . . . . . . . . . . . . . . . . . . . .

61

Natural Gas Injector Performance . . . . . . . . . . . . . . . . . . . . . .

62

4.2.1

Injector Mass Flow Testing . . . . . . . . . . . . . . . . . . . . . .

63

4.2.2

Injected Gas Plume and Chamber Bulk Gas Interactions . . . . .

66

FFID Calibration and Performance . . . . . . . . . . . . . . . . . . . . .

68

5 Combustion Data Processing and Heat Release Calculations

70

5.1

Determining Ignition Delay . . . . . . . . . . . . . . . . . . . . . . . . .

70

5.2

Heat Release Calculations . . . . . . . . . . . . . . . . . . . . . . . . . .

71

5.3

Modelling In-Cylinder Conditions . . . . . . . . . . . . . . . . . . . . . .

73

5.3.1

In-Cylinder Mass Model . . . . . . . . . . . . . . . . . . . . . . .

74

5.3.2

In-Cylinder Temperatures . . . . . . . . . . . . . . . . . . . . . .

75

5.3.3

Specific Heat of the Working Fluid . . . . . . . . . . . . . . . . .

75

5.3.4

Signal Derivative . . . . . . . . . . . . . . . . . . . . . . . . . . .

76

5.4

Combustion Efficiency Calculations . . . . . . . . . . . . . . . . . . . . .

76

5.5

Particulate Data Processing . . . . . . . . . . . . . . . . . . . . . . . . .

77

vi

6 Experimental Results and Discussion

79

6.1

Experimental Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . .

79

6.2

Test Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

80

6.3

Results Preamble . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

81

6.4

Ignition Delay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

83

6.4.1

Ignition Delay vs Nozzle Hole Pattern . . . . . . . . . . . . . . . .

84

6.4.2

Ignition Delay vs Nozzle Hole Geometry . . . . . . . . . . . . . .

87

Combustion Image Data . . . . . . . . . . . . . . . . . . . . . . . . . . .

88

6.5.1

Images from the 1 Hole Nozzle Pattern . . . . . . . . . . . . . . .

89

6.5.2

Images from the 3 Hole Nozzle Pattern . . . . . . . . . . . . . . .

89

6.5.3

Images from the 9 Hole Nozzle Pattern . . . . . . . . . . . . . . .

90

ROHR and Net Heat Release: Curve Profiles . . . . . . . . . . . . . . . .

98

6.6.1

Curves from the 1 Hole Nozzle Pattern . . . . . . . . . . . . . . .

98

6.6.2

Curves from the 3 Hole Nozzle Pattern . . . . . . . . . . . . . . .

98

6.6.3

Curves from the 9 Hole Nozzle Pattern . . . . . . . . . . . . . . .

99

6.5

6.6

6.7

6.8

6.9

ROHR and Net Heat Release: Nozzle Performance . . . . . . . . . . . . . 105


6.7.1

Heat Release vs Nozzle Hole Pattern . . . . . . . . . . . . . . . . 105

6.7.2

Heat Release vs Nozzle Hole Geometry . . . . . . . . . . . . . . . 109

Engine Efficiency Calculations . . . . . . . . . . . . . . . . . . . . . . . . 110


6.8.1

Efficiency vs Nozzle Hole Pattern . . . . . . . . . . . . . . . . . . 111

6.8.2

Efficiency vs Nozzle Hole Geometry . . . . . . . . . . . . . . . . . 114

Particulate Emissions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115


6.9.1

Organic Carbon vs Elemental Carbon . . . . . . . . . . . . . . . . 115

6.9.2

PM Emissions vs Nozzle Hole Pattern . . . . . . . . . . . . . . . . 119

6.9.3

PM Emissions vs Nozzle Hole Geometry . . . . . . . . . . . . . . 120

7 Conclusions and Recommendations


7.1

123

Conclusions of Experimental Results . . . . . . . . . . . . . . . . . . . . 124


vii

7.2

Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

References

131

A Equipment Calibration Curves

140

A.1 AST Pressure Transducer . . . . . . . . . . . . . . . . . . . . . . . . . . 140


A.2 Kistler Pressure Transducer . . . . . . . . . . . . . . . . . . . . . . . . . 140
A.3 GM Manifold Absolute Pressure Sensor . . . . . . . . . . . . . . . . . . . 142
A.4 TSI 2013 Laminar Flowmeter . . . . . . . . . . . . . . . . . . . . . . . . 142
A.5 Matheson Rotameters

. . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

B Complete Set of Data Figures

145

C MATLAB Script Description

171

D Drawings

174

E Thermocouple Assembly Instructions

185

E.1 Preliminary Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185


E.2 Assembly Instructions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

viii

List of Tables
4.1

Operating conditions of motored tests . . . . . . . . . . . . . . . . . . . .

4.2

Nozzle hole area and % difference between the circular and elliptical geometry 65

6.1

Operating conditions of fired tests . . . . . . . . . . . . . . . . . . . . . .

80

6.2

Experimental Test Matrix . . . . . . . . . . . . . . . . . . . . . . . . . .

82

ix

53

List of Figures
1.1

Comparison of estimated heavy-duty emission targets from 2008 . . . . .

1.2

Number of NGVs vs natural gas refuelling stations . . . . . . . . . . . .

2.1

Development of a self-similar quasi-static transient jet displaying the twozone model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

12

2.2

Dispersion model of a transient jet . . . . . . . . . . . . . . . . . . . . .

13

2.3

Principal characteristics of the expansion process for an under-expanded jet 15

2.4

Schematic of direct injection combustion . . . . . . . . . . . . . . . . . .

16

2.5

Simulated in-cylinder plume interactions . . . . . . . . . . . . . . . . . .

18

2.6

Temperature and pressure map for natural gas auto-ignition under 2ms .

21

2.7

Pressure Delay vs Temperature for various fuels . . . . . . . . . . . . . .

22

2.8

Ignition delay time and ignition temperature of various additives with


methane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

23

Injected gas plume deflection due to glow plug shield . . . . . . . . . . .

26

2.10 Axis switching phenomenon from elliptical jets . . . . . . . . . . . . . . .

28

2.11 Normalized ROHR curves vs CAD for elliptical and round nozzle holes .

29

2.12 Soot derived from pyrolysis of benzene, ethanol, and acetylene . . . . . .

30

2.13 Soot particle burn-up rate in a diesel combustion environment . . . . . .

33

2.9

2.14 Soot from pilot diesel for a given set of engine parameters: overall equivalence ratio (), %EGR, pilot flow rate (low=0.15kg/hr, high=0.25kg/hr)
x

34

3.1

3-D view of CFR cylinder and ERDL combustion bomb . . . . . . . . . .

38

3.2

Exploded view of ERDL combustion bomb and its components . . . . . .

38

3.3

Glow plug position relative to injector nozzle . . . . . . . . . . . . . . . .

41

3.4

Enlarged image of injector nozzle holes . . . . . . . . . . . . . . . . . . .

44

3.5

Injector nozzle and hole patterns . . . . . . . . . . . . . . . . . . . . . .

45

3.6

Isokinetic sampling system schematic . . . . . . . . . . . . . . . . . . . .

47

4.1

Pressure trace of combustion bomb under motored conditions averaged


over the entire bomb wall temperature range . . . . . . . . . . . . . . . .

4.2

54

Temperature trace of combustion bomb under motored conditions averaged over the entire bomb wall temperature range . . . . . . . . . . . . .

55

4.3

Peak, TDC, and BDC temperature correlations . . . . . . . . . . . . . .

56

4.4

Motored intake manifold absolute pressure trace . . . . . . . . . . . . . .

58

4.5

Motored intake air flow rate . . . . . . . . . . . . . . . . . . . . . . . . .

59

4.6

Motored P-v trace with polytropic compression line . . . . . . . . . . . .

60

4.7

Motored mass fraction in combustion bomb . . . . . . . . . . . . . . . .

61

4.8

Mass of Fuel per Injection vs Injection Duration . . . . . . . . . . . . . .

64

4.9

FFID Calibration Curves . . . . . . . . . . . . . . . . . . . . . . . . . . .

69

5.1

Motored, fired, and differential pressure traces for a typical fired cycle . .

71

5.2

Injector current and pressure differential with mean and 3 lines during
ignition event . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

72

6.1

Ignition Delay vs Mass of Fuel Injected . . . . . . . . . . . . . . . . . . .

84

6.2

Ignition Delay vs Mass of Fuel Injected per Hole . . . . . . . . . . . . . .

85

6.3

Ignition Delay vs Mass of Fuel Injected per Area . . . . . . . . . . . . . .

85

6.4

Set of combustion images for 10 consecutive fired cycles for a 1 hole nozzle,
2 CAD duration (scaled to 5%-95% intensity range) . . . . . . . . . . . .
xi

92

6.5

Typical image histogram for a 1 hole nozzle, 2 CAD duration showing the
number of pixels above 50% intensity . . . . . . . . . . . . . . . . . . . .

6.6

Set of combustion images for 10 consecutive fired cycles for a 3 hole nozzle,
3 CAD duration, 10 consecutive fired cycles . . . . . . . . . . . . . . . .

6.7

95

Set of combustion images for 10 consecutive fired cycles for a 9 hole nozzle,
3 CAD duration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

6.9

94

Typical image histogram for a 3 hole nozzle, 3 CAD duration showing the
number of pixels above 50% intensity . . . . . . . . . . . . . . . . . . . .

6.8

93

96

Typical image histogram for a 9 hole nozzle, 3 CAD duration showing the
number of pixels above 50% intensity . . . . . . . . . . . . . . . . . . . .

97

6.10 Typical ROHR and Net Heat curves for a 3 hole nozzle, 3 CAD duration

100

6.11 Typical ROHR and Net Heat curves for a 9 hole nozzle, 1.4 CAD duration 103
6.12 Typical ROHR and Net Heat curves for a 9 hole nozzle, 3 CAD duration

104

6.13 Max ROHR vs Mass of Fuel Injected . . . . . . . . . . . . . . . . . . . . 106


6.14 Max ROHR vs Mass of Fuel Injected per Hole . . . . . . . . . . . . . . . 106
6.15 Max ROHR vs Mass of Fuel Injected per Area . . . . . . . . . . . . . . . 107
6.16 Net Heat vs Mass of Fuel Injected . . . . . . . . . . . . . . . . . . . . . . 107
6.17 Net Heat vs Mass of Fuel Injected per Hole . . . . . . . . . . . . . . . . . 108
6.18 Net Heat vs Mass of Fuel Injected per Area . . . . . . . . . . . . . . . . 108
6.19 Fuel Utilization vs Mass of Fuel Injected . . . . . . . . . . . . . . . . . . 112
6.20 Fuel Utilization vs Mass of Fuel Injected per Area . . . . . . . . . . . . . 112
6.21 Indicated Thermal Efficiency vs Mass of Fuel Injected . . . . . . . . . . . 113
6.22 Indicated Thermal Efficiency vs Mass of Fuel Injected per Area . . . . . 113
6.23 Distribution of Organic and Elemental Carbon . . . . . . . . . . . . . . . 117
6.24 Elemental Carbon Emissions vs Mass of Fuel Injected . . . . . . . . . . . 118
6.25 Organic Carbon Emissions vs Mass of Fuel Injected . . . . . . . . . . . . 118
6.26 Specific PM Emissions vs Mass of Fuel Injected . . . . . . . . . . . . . . 122
xii

6.27 Mass Collected on Filter vs Mass of Fuel Injected . . . . . . . . . . . . . 122


A.1 AST4700 pressure transducer calibration curve . . . . . . . . . . . . . . . 141
A.2 Kistler piezoelectric pressure transducer calibration curve . . . . . . . . . 141
A.3 GM MAP sensor calibration curve . . . . . . . . . . . . . . . . . . . . . . 142
A.4 TSI 2013 flowmeter calibration curve . . . . . . . . . . . . . . . . . . . . 143
A.5 Matheson rotameter calibration curve, Tube 6 with stainless steel float . 144
A.6 Matheson rotameter calibration curve, Tube 602 with glass float . . . . . 144
B.1 Ignition Delay vs Mass of Fuel Injected . . . . . . . . . . . . . . . . . . . 146
B.2 Ignition Delay vs Mass of Fuel Injected per Hole . . . . . . . . . . . . . . 146
B.3 Ignition Delay vs Mass of Fuel Injected per Area . . . . . . . . . . . . . . 147
B.4 Set of combustion images for 10 consecutive fired cycles for a 1 hole nozzle,
2 CAD duration (scaled to 5%-95% intensity range) . . . . . . . . . . . . 148
B.5 Typical image histogram for a 1 hole nozzle, 2 CAD duration showing the
number of pixels above 50% intensity . . . . . . . . . . . . . . . . . . . . 149
B.6 Set of combustion images for 10 consecutive fired cycles for a 3 hole nozzle,
3 CAD duration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
B.7 Typical image histogram for a 3 hole nozzle, 3 CAD duration showing the
number of pixels above 50% intensity . . . . . . . . . . . . . . . . . . . . 151
B.8 Set of combustion images for 10 consecutive fired cycles for a 9 hole nozzle,
3 CAD duration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
B.9 Typical image histogram for a 9 hole nozzle, 3 CAD duration showing the
number of pixels above 50% intensity . . . . . . . . . . . . . . . . . . . . 153
B.10 Typical ROHR and Net Heat curves for a 1 hole nozzle, 3 CAD duration

154

B.11 Typical ROHR and Net Heat curves for a 3 hole nozzle, 3 CAD duration

155

B.12 Typical ROHR and Net Heat curves for a 9 hole nozzle, 1.4 CAD duration 156
B.13 Typical ROHR and Net Heat curves for a 9 hole nozzle, 3 CAD duration
xiii

157

B.14 Max ROHR vs Mass of Fuel Injected . . . . . . . . . . . . . . . . . . . . 158


B.15 Max ROHR vs Mass of Fuel Injected per Hole . . . . . . . . . . . . . . . 158
B.16 Max ROHR vs Mass of Fuel Injected per Area . . . . . . . . . . . . . . . 159
B.17 Net Heat vs Mass of Fuel Injected . . . . . . . . . . . . . . . . . . . . . . 159
B.18 Net Heat vs Mass of Fuel Injected per Hole . . . . . . . . . . . . . . . . . 160
B.19 Net Heat vs Mass of Fuel Injected per Area . . . . . . . . . . . . . . . . 160
B.20 Fuel Utilization vs Mass of Fuel Injected . . . . . . . . . . . . . . . . . . 161
B.21 Fuel Utilization vs Mass of Fuel Injected per Hole . . . . . . . . . . . . . 161
B.22 Fuel Utilization vs Mass of Fuel Injected per Area . . . . . . . . . . . . . 162
B.23 Indicated Thermal Efficiency vs Mass of Fuel Injected . . . . . . . . . . . 162
B.24 Indicated Thermal Efficiency vs Mass of Fuel Injected per Hole . . . . . . 163
B.25 Indicated Thermal Efficiency vs Mass of Fuel Injected per Area . . . . . 163
B.26 Distribution of Organic and Elemental Carbon . . . . . . . . . . . . . . . 164
B.27 Elemental Carbon Emissions vs Mass of Fuel Injected . . . . . . . . . . . 164
B.28 Elemental Carbon Emissions vs Mass of Fuel Injected per Hole . . . . . . 165
B.29 Elemental Carbon Emissions vs Mass of Fuel Injected per Area . . . . . . 165
B.30 Organic Carbon Emissions vs Mass of Fuel Injected . . . . . . . . . . . . 166
B.31 Organic Carbon Emissions vs Mass of Fuel Injected per Hole . . . . . . . 166
B.32 Organic Carbon Emissions vs Mass of Fuel Injected per Area . . . . . . . 167
B.33 Specific PM Emissions vs Mass of Fuel Injected . . . . . . . . . . . . . . 167
B.34 Specific PM Emissions vs Mass of Fuel Injected per Hole . . . . . . . . . 168
B.35 Specific PM Emissions vs Mass of Fuel Injected per Area . . . . . . . . . 168
B.36 Mass Collected on Filter vs Mass of Fuel Injected . . . . . . . . . . . . . 169
B.37 Mass Collected on Filter vs Mass of Fuel Injected per Hole . . . . . . . . 169
B.38 Mass Collected on Filter vs Mass of Fuel Injected per Area . . . . . . . . 170
D.1 Injector nozzle with 3 hole elliptical geometry, Page 1 . . . . . . . . . . . 175
D.2 Injector nozzle with 3 hole elliptical geometry, Page 2 . . . . . . . . . . . 176
xiv

D.3 Injector nozzle with 3 hole elliptical geometry, Page 3 . . . . . . . . . . . 177


D.4 Injector nozzle with 9 hole elliptical geometry, Page 1 . . . . . . . . . . . 178
D.5 Injector nozzle with 9 hole elliptical geometry, Page 2 . . . . . . . . . . . 179
D.6 Injector nozzle with 9 hole elliptical geometry, Page 3 . . . . . . . . . . . 180
D.7 Simplified nozzle seat, Page 1 . . . . . . . . . . . . . . . . . . . . . . . . 181
D.8 Simplified nozzle seat, Page 2 . . . . . . . . . . . . . . . . . . . . . . . . 182
D.9 9E nozzle seat prototype . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
D.10 Glow plug shield simulator . . . . . . . . . . . . . . . . . . . . . . . . . . 184

xv

Nomenclature
List of Symbols
M

Jet Momentum Flowrate

m
P M Mass Rate of Production of Particulate Matter
F uel% Fuel Utilization Efficiency
GiT E Gross Indicated Thermal Efficiency
m
PM
P

Specific Particulate Matter Emissions

dcv
dT

Rate of Change of Specific Heat with Temperature

dmb
d

Rate of Change of Mass of Blowby Gas with Crank Angle

dm
d

Rate of Change of Cylinder Mass with Crank Angle

dP
d

Rate of Change of Cylinder Pressure with Crank Angle

dQnet
d

Net Rate of Heat Release

dT
d

Rate of Change of Bomb Temperature with Crank Angle

dU
d

Rate of Change of Internal Energy with Crank Angle

dV
d

Rate of Change of Cylinder Volume with Crank Angle

Jet Penetration Constant


xvi

Ratio of Specific Heat

Angular Momentum of the Chamber Bulk Gas

Relative Air/Fuel Ratio

Angular Velocity of the Chamber Bulk Gas

Fuel/Air Equivalence Ratio

ch

Chamber Gas Density

Binary Diffusion Coefficient

cp

Specific heat at Constant Pressure

cv

Specific heat at Constant Volume

Db

Diameter of the Combustion Chamber

Ib

Moment of Inertia of the Chamber Bulk Gas

mb

Mass of the Chamber Bulk Gas

mf

Mass of Fuel Injected

Nb

Rotational Speed of the Chamber Bulk Gas

P0

Reservoir Pressure

pf

Linear Momentum of the Injected Fuel

Pch

Chamber Pressure

QN et Net Heat Released


R

Gas Constant

rj

Nozzle Hole Radius


xvii

tD

Characteristic Laminar Diffusion Time

Tf

Flame Temperature

Tn

Soot Nucleation Temperature

TBDC Bomb Temperature at BDC


Tpeak Peak Bomb Temperature
TT DC Bomb Temperature at TDC
Twall Bomb Wall Temperature
vf

Velocity of Fuel Injected

Zt

Jet Penetration Distance

List of Abbreviations
ABF Adjustable Burst Fire
AT DC After Top Dead Centre
BC

Black Carbon

BDC Bottom Dead Centre


BT DC Before Top Dead Centre
CDM Crank Degree Marker
CF R Cooperative Fuels Research
CI

Compression Ignition

CN G Compressed Natural Gas


DAQ Data Acquisition
xviii

DI

Direct Injection

DIN G Direct Injection Natural Gas


EDM Electric Discharge Machining
EGR Exhaust Gas Recirculation
ERDL Engine Research and Development Laboratory
EV C Exhaust Valve Closing
EV O Exhaust Valve Opening
F CV Fuel Cell Vehicle
F F ID Fast Flame Ionization Detector
GikW h Gross Indicated Kilowatt-Hour
HD

Heavy-Duty

ICCD Intensified Charge Coupled Device


IV C Intake Valve Closing
IV O Intake Valve Opening
LHV Lower Heating Value
M BT Maximum Brake Torque
NG

Natural Gas

N GV Natural Gas Vehicle


N M HC Non-Methane Hydrocarbons
N M OG Non-Methane Organic Gases
xix

N Ox Oxides of Nitrogen (e.g.N O, N O2 )


OEM Original Equipment Manufacturer
P AH Polycyclic Aromatic Hydrocarbon
PM

Particulate Matter

ROHR Rate of Heat Release


RON Research Octane Number
SI

Spark Ignition

sP M Specific Particulate Matter Emissions


T DC Top Dead Centre
T HC Total Hydrocarbons
V W DA Variable Window Data Acquisition

xx

Chapter 1
Introduction

1.1

The Drive Towards Alternative Fuels

As societys technological innovations are growing in scale and complexity, so too does the
demand for the energy to power those innovations. Of the three largest sources of energy,
approximately 38% of the worlds power involves petroleum, 24% makes use of coal, and
24% utilizes natural gas [1]. In todays global world, energy independence and security
is of great importance to ensure economic success as it is an invaluable resource with
growing demand. This is also reflected in the rise of crude oil prices that have increased
substantially in recent years. A key aspect would then be to utilize a local natural
resource that is abundant. Natural gas in particular shows to have great potential as
the industry is experiencing large growth in existing and developing markets, with North
America already producing a significant portion of the global supply [2]. This, coupled
with the push towards a sustainable green future and stricter emission regulations, is
motivation for the use of alternative energy sources and the appeal of natural gas.
Globally, one of the most important uses of energy is in the transportation sector.
The focus of this paper will henceforth be focused on examining natural gas for use as a
vehicular fuel. This leads to the other central driving force for alternative fuels, which
1

Chapter 1. Introduction

Figure 1.1: Comparison of estimated heavy-duty emission targets from 2008 [5]

are the emissions. More specifically this pertains to the degrading air quality and health
hazards which are prevalent in large population dense cities. Exhaust regulations for
heavy-duty (HD) diesel engines began in the 1960s and have progressively grown more
stringent with more aggressive goals that are concurrent with available technologies [3].
The main concerns of emissions in regards to vehicles are that of Particular Matter
(PM), Carbon Monoxide (CO), Oxides of Nitrogen (NOx ), and reactive Non-Methane
Hydrocarbons (NMHC)1 , which are prevalent in current engine exhaust. Particulate
Matter is linked to cause respiratory ailments as well as contribute to smog. NOx and
some NMHC are highly photoreactive and cause ground level ozone, which also lead
to respiratory illnesses. Carbon monoxide and NMHC are toxic or can cause adverse
health effects so exposure to such compounds is strictly regulated. Carbon Dioxide
(CO2 ) is another cause for concern contributing to greenhouse gas emissions. Although
at the moment it is not regulated, proposed incremental reduction of CO2 emissions in
diesel HD fleets is set for implementation in the United States2 in the near future [4]
with regulations on NOx , NMHC, and PM emissions also reducing. As exemplified in
Figure 1.1, emissions standards are aggressively being tightened on a seemingly regular
basis and the main challenge of the vehicle industry is to meet those standards.
1
2

Also known as Non-Methane Organic Gases NMOG.


Canadas emissions regulations are aligned with that of the US EPA federal standards.

Chapter 1. Introduction

Alternative fuels in the transport sector are often first implemented in the heavy-duty
(HD) market, which is largely due to cost and infrastructure. As the vast majority of
alternative fuels are still in their infancy, applying them to HD transport fleets are more
easily achieved due to the smaller volumes compared to passenger cars and the lower
cost to implement due to the localized refuelling and maintenance stations. It is also
common practice for governments to heavily subsidize the application of these fuels to
the public sector vehicles, such as buses and taxis. This provides an incentive and an
economical means of implementation for these small scale road tests. Alternative fuels at
this point are not self-sufficient largely due to their immaturity, which incorporates their
technologies as well as their infrastructure; and government subsidies and regulations are
paramount to their growth [6].

1.2

Alternative Fuels for Transportation

Several alternative fuels have been tested and explored for vehicular use with the goal
of reducing harmful emissions. Their consideration involves several factors. For one
their combustion performance, in terms of emissions as well as efficiency, is their central
driving factor. Furthermore, their adaptability to current engines and their storage is also
crucial: modular fuel, retrofitting an engine, or having a complete engine redesign would
dictate the ease in which they can be applied in the market and at what cost. Lastly, the
degree to which the infrastructure would have to change is also important for widespread
use. This relates to the network of suppliers, OEMs, automotive manufacturers, as well
as fuelling stations and fuel supply lines [6, 7].
Various potential fuels have been researched with the above factors in mind and will
be discussed on that basis. Ethanol, for example, is commonly derived from sugar cane
or corn and its supply can fluctuate heavily. Methanol can be derived from natural gas
and organic materials but a central drawback for alcohols are cold-start performance

Chapter 1. Introduction

due to their high heat of vaporization. For this reason they are likely to be used as
an additive to gasoline [8]. Reformulated gasoline and gasoline-methanol blends of 15%
methanol (M15) and 85% methanol (M85), for example, show some improvements in
emissions over gasoline; though could be comparable in some cases depending on its
operation [9, 10].
Fuel cell vehicles (FCV) can have higher efficiency and very low emissions depending
on the fuel on which they operate. In regards to that, FCV require highly refined, pure
fuels to operate and materials are very specialized and expensive to produce [7]. Needless
to say, this technology is still developing and will require large infrastructure changes to
employ.
Natural gas (NG) was shown to have consistently lower emissions but has much lower
energy density, so on-board storage and drivable distance is an issue. All of these fuels
hold potential for further development and implementation but, for reasons that will be
discussed shortly, this study will focus on the use of natural gas as a vehicle fuel.

1.3

Natural Gas as an Alternative Fuel for Transportation

1.3.1

Attraction of Natural Gas

Natural gas is a vastly abundant resource in North America and is already widely used
across the world for a variety of services, making it a favourable contender as an alternative to gasoline and diesel. Although the creation of NG refuelling stations is essential,
the extraction and distribution infrastructure is already largely in place with suppliers
and pipelines in place to deliver NG for residential, commercial, and power generating
usage. Yeh [6] examined the major nations with developing natural gas vehicle (NGV)
markets and found that government subsidies were essential to their survival due to the

Chapter 1. Introduction

Figure 1.2: Number of NGVs vs natural gas refuelling stations [6]


immaturity of the technology. Furthermore, successful markets tended to gravitate towards 1 refuelling station per 1000 NGV (see Figure 1.2). In North America, it is common
for residential homes to have NG pipelines making it favourable for the implementation
of NGVs with the benefit of home refuelling.
Applying NG engines to existing HD fleets is ideal from a logistics stand point and can
recover savings due to lower operating and fuel costs [11]. Also, NG combustion engines
have been researched for quite some time, and advancements have been made in terms of
retrofitting spark ignition (SI) engines. Hekkert et al. [6] examined the well-to-wheel
process in automotive fuel chains of several alternative fuels and found that, although
FCVs offer great potential for improved efficiency and reduced emissions, the technology
and implementation should be a long term goal and NGVs offer the best short-to-medium
term solution as a transition technology. Thus, CNG is a strong contender over other
fuels as an alternative to gasoline and diesel.

1.3.2

NGV Engines

Natural gas vehicle engines that have been successfully applied to road vehicles are variations of conventional spark ignition engines. The octane rating of natural gas at 130
allows for the engines to run at higher compression ratios of 11 : 1 to 15 : 1. Also,

Chapter 1. Introduction

the flammability limits with respect to conventional fuels is also larger, allowing to run
leaner combustion that can lead to lower NOx and CO and higher thermal efficiency [12].
Generally, NG engines have lower harmful emissions than their counterparts, however
to what extent varies greatly on the engine, operating conditions, and the retrofitting.
Furthermore, great advances in control technologies in recent years have significantly reduced gasoline and diesel emissions such that the use of any alternative fuel would require
a dedicated engine specifically designed and optimized for the fuel in question [13].
The choices for NG engines are currently limited to Stoichiometric Otto Gas Engines
and Lean Burn Otto Gas Engines. The former typically relies on a multipoint gas injection upstream of the intake port to provide a stoichiometric mixture of fuel and air
in the cylinder where an oxygen sensor in the exhaust regulates the fuel injected. They
operate at a compression ratio of 11 : 1 to 11.5 : 1 but the power output is 5 15% lower
than conventional diesel engines and show a 15 20% worse fuel consumption [12].
The Lean Burn Otto Gas Engines operate with relative air/fuel ratios in the range
of = 1.1 1.6 and compression ratios of 11 : 1 to 11.5 : 1, where the lean limits are
dictated by flame stability. They do show considerable improvement in thermal efficiency
and fuel consumption over their stoichiometric counterparts however fuel consumption
still remains 15% and 30% behind diesel engines, for steady state and transient operation,
respectively [12].
Another engine concept under development is direct injection natural gas (DING)
systems. Like a diesel system, the fuel would be injected directly in the cylinder late in the
compression stroke. The charge of air remains constant while the amount of fuel is varied
for the desired power output. This throttleless process offers reduced pumping work
and higher fuel conversion efficiencies due to lean burning [14]. Natural gas, however,
is adverse to autoignition under diesel conditions which is reflected in its high octane
rating. Consequently, it requires some form of ignition assistance to properly ignite in
the required time frame. A spark plug is not capable of performing under these conditions

Chapter 1. Introduction

due to flame stability issues as a result of missing an ignitable mixture leading to high
cyclic variability [15, 16]. Alternatively, the use of a dual-fuel system where a small
quantity of diesel fuel is injected prior to the natural gas to pilot-ignite the mixture is
also under investigation. Using just enough diesel fuel to ignite the natural gas, the
emissions are reduced with a small penalty in efficiency [17,18]. A third method, and the
one used in this study, is hot surface ignition, by which the natural gas is directly injected
in the combustion chamber with a glow plug acting as the ignition source. The glow plug
in this case remains powered during the entire cycle and the gas then auto-ignites once
it comes into contact with the surface. Hence, the rationale behind a DING engine is to
take advantage of the high volumetric and thermal efficiencies, operating at compression
ratios of 16 : 1 to 18 : 1, and benefit from the low emissions of natural gas.

1.3.3

Natural Gas Fuel Properties and Emissions

The composition of natural gas varies by region and over time but typically the majority
of NG consists primarily of 90 95% methane, around 2% ethane, trace amounts of
propane, butane, carbon dioxide, nitrogen, and other minute gases. The relative purity
of NG and the high hydrogen to carbon ratio lends itself to improved emissions. There
is no sulphur that can cause catalyst poisoning or SO2 formation leading to acid rain, a
great reduction in NMHC, less carbon dioxide, carbon monoxide, oxides of nitrogen, and
soot. It also has a wider flammability range and lower adiabatic flame temperature than
the conventional fuels, as mentioned earlier, allowing for leaner and cooler combustion
that would lead to the reduction in NOx and CO.
To reiterate, it has been widely reported that emissions of NG offer the potential
for substantial reduction in harmful emissions compared with conventional fuels but the
degree of success by which they are reduced varies significantly. Reductions in carbon
dioxide emissions can range from 6.5% 22% [9, 19], that of carbon monoxide ranges
from 75% 94% [13, 19, 20], NOx from 25% 76% [13, 1922], particulate matter from

Chapter 1. Introduction

80% 99% [13, 19].


For the case of total unburned hydrocarbons (THC) they have generally been higher
for NGV compared to gasoline and diesel. The composition, however, is largely unreactive methane that is usually around 95% of the THC [10, 19]. The remaining NMHCs
have shown significant reduction; particularly in the toxic emissions, such as benzene,
formaldehyde, and acetaldehyde. Typically, methane is also a very stable compound
that has a very low conversion efficiency in catalysts. The high temperatures needed
for methane to have reasonable conversion efficiencies are difficult to achieve due to the
inherently low temperatures as a result of lean combustion. Overall, there is a net benefit
in emissions in the use of NG over conventional fuels and, as further developments arise,
its performance will improve.

1.4

Motivation and Objective

Over the years the University of Torontos Engine Research and Development Laboratory (ERDL) has performed extensive research in the area of direct injection natural gas
engines. In particular this was accomplished with the use of glow plugs as a form of
ignition assistance in a single cylinder engine coupled with an optically accessible combustion bomb [23]. Early work helped establish the visualization of the injected gas and
a model of the jet was formed [24]. Later the study continued with an evaluation of
ignition delay times using a single circular hole nozzle with the goal of optimizing glow
plug placement and operation [25]. This was continued by examining the effects of the
addition of a second nozzle hole and led to the design of an improved glow plug shield
to reduce ignition delay times [26]. This gave insight into the path at which the flame
propagates from one gas plume to another, noting the interactions that took place.
Furthermore, a computer simulation was performed to characterize the behaviour of
the injected gas and formed good agreement with previous experiments while improving

Chapter 1. Introduction

the understanding of the in-cylinder flows in the ERDL bomb [27]. After this, a separate
venue was explored by attempting to improve the mixing and combustion characteristics
of the injected natural gas with the air in the cylinder through the use of elliptical
nozzle holes [28]. The rationale was to take advantage of the axis switching phenomenon
inherent in an elliptical jet to entrain more air and thus provide more rapid mixing. The
results yielded similar but distinct behaviour between the elliptical and circular nozzle
holes used.
The objective of this thesis is to expand upon the previous works in the ERDL group.
More specifically, this study will examine the effects of changing the nozzle hole pattern
(i.e. 1/3/9 hole nozzles) and hole geometry (i.e. Elliptical/Circular) on the following:
Ignition Delay Time.
Combustion Images and Histograms.
Rate of Heat Release and Net Heat Curve Profiles.
Rate of Heat Release and Net Heat Peak Values.
Fuel Utilization and Gross Indicated Thermal Efficiencies.
Particulate emissions: Organic/Elemental carbon content and gravimetric analysis.

Chapter 1. Introduction

10

The following thesis is broken down into several chapters, each one sequentially presenting information for an extensive description of the rationale behind this study, the
supporting theory, the experimental apparatus and its validation, the methodology in
the experiments calculations, and a discussion on the results with concluding remarks.
It is presented in the following order:
Chapter 1 Introduction
An introduction to this study.
Chapter 2 The Process of Injection and Ignition in a DING Engine
A comprehensive literature survey describing the relevant theory pertaining to this
work.
Chapter 3 Experimental Apparatus
The description of the experimental apparatus used in this study.
Chapter 4 Validation of Experimental Apparatus
The description and methodological validation of key measurement equipment.
Chapter 5 Combustion Data Processing and Heat Release Calculations
An explanation of the main set of combustion calculations used to process the data.
Chapter 6 Experimental Results and Discussion
The experimental procedure, results, and a discussion presenting them.
Chapter 7 Conclusions and Recommendations
A conclusive summary on the findings of this study with recommendations on future
work.

Chapter 2

The Process of Injection and


Ignition in a DING Engine

2.1

Direct Injection of Natural Gas

The injection process during engine operation is inherently a transient process in which
many factors affect its successful operation. The injection process of a gaseous fuel has
many distinct characteristics that differ from its liquid counterpart which must be taken
into consideration. Adhering to these differences, the gas jet structure and development
during injection consequently affects the mixing and chamber penetration. The resulting
ignition, or lack thereof, depends on the speed at which penetration and mixing occur
in conjunction with the ignition source. Studies of quiescent unconfined mixtures give
insight into gas jet development but ultimately need to take into account other interactions including that of multiple gas plumes, gas confinement, chamber gas motion and
nozzle hole geometry.
11

Chapter 2. The Process of Injection and Ignition in a DING Engine

12

Figure 2.1: Development of a self-similar quasi-static transient jet displaying the two
zone model [29]

2.1.1

Gas Jet Development and Structure

The development of a gas plume injected in neutral surroundings can be described by a


two-zone model comprised of a starting plume and a spherical vortex ring [29]. Looking
at the injection of a single plume in a quiescent mixture, when it is first injected, its initial
radius is approximately the same size as that of the nozzle inlet. As the fluid is being
injected, it forms a conical plume that expands outward from the source, the head of
which is the vortex. The vortex begins to grow mostly from the entrained mass entering
from the trailing plume and entrains little from the surroundings [30]. The jet region
exhibits a quasi-static self-similar shape as it develops (see Figure 2.1). The momentum
of the vortex changes as it increases in size and distance from the jet origin due to the
increasing viscous forces of the surrounding fluid. This behaviour remained apparent in
several studies that involved laminar [31] and turbulent jets [32] as well as methane jets
that displayed self-similar concentration profiles [33].
Another model for transient gas jets is described by Hyun et al. [34], which can be
categorized as a four-zone model. Referring to a turbulent transient flow, this draws out
in more detail potential entrainment and dilution regions of the jet with the surrounding

Chapter 2. The Process of Injection and Ignition in a DING Engine

13

gases. In the initial stages, similar to the two-zone model, a conical head starts to form at
the tip while still maintaining a somewhat uniform body near the injector head. During
this process there is a large velocity gradient between the axis of the incoming jet and
the quiescent surroundings. This, however, leads to large shear forces of the fluid at
the boundary leading to the formation of several small vortices at the periphery. As the
injection continues, vortices begin to multiply and grow in size while the jet becomes fully
turbulent further downstream of the nozzle. Near the head, the momentum of the gas
decreases due to viscous forces and velocity is quickly decreased. The structure of the jet
is largely due to the instability caused by the shear stress, even in the initial stages. As
a result, a coalescence of vortices takes place in which small vortices exist within a larger
vortex. This phenomenon contributes to the fast growth and multiplication of vortices
within the jet.
In essence, Hyun et al. identified four
regions (see Figure 2.2): the potential core
near the point of injection that is nearly
the size of the nozzle and whose velocity is
almost completely axial; the main jet region at the centre of the plume which is
dominated by the axial velocity and low
shear that almost reaches the jet tip; the
mixing flow region where the coalescence
and multiplication of vortices are turbuFigure 2.2: Dispersion model of a transient

lently mixing with the quiescent surround-

jet [34]

ings; and the dilution region where the jet

has lost most of its momentum and is being pushed aside by the forthcoming gas.
The cases discussed thus far describe flow of jets that were injected at subsonic velocities, hence the gases were treated as incompressible fluids. For use in a DING engine,

Chapter 2. The Process of Injection and Ignition in a DING Engine

14

natural gas is injected at pressures high enough to induce choked flow resulting in sonic
exit velocities. The rationale behind it is to be able to control the amount of fuel being
delivered by controlling the inlet conditions independent of the conditions downstream of
the injector. Choked flow is defined by the critical pressure ratio given by Equation 2.1:
Pch
=
P0

2
+1

 (1)

(2.1)

where is the ratio of specific heat of the gas 1 , Pch is the exit chamber pressure, and
P0 is the injector pressure. An injector pressure that surpasses this point will result in
choked flow where the gas is said to be underexpanded. Ewan and Moodie [35] describe
the principal region at the nozzle exit for an underexpanded jet (see Figure 2.3). At
the tip before exiting the nozzle, the gas reaches sonic speeds (M = 1). Shortly after
exiting, the gas expands, leading to supersonic (M > 1) speeds just past the nozzle.
This in turn sets off a series of barrel shocks that create a normal shock (Mach disc).
Furthermore, this Mach disc propagates itself forward and outward creating reflected
shocks in the subsonic surroundings. The barrel shock is said to take place from 1 to 3
nozzle diameters from the nozzle hole, depending on the pressure ratio. Once choked,
the velocity of the gas has reached its maximum but an increase in pressure ratio leads
to increased gas penetration [36]. Moreover, studies have shown that underexpanded
jets maintain their self-similar characteristic being unaffected by the initial supersonic
expansion [37, 38].

2.1.2

Jet Penetration and Mixing

Two key factors for successful combustion are the penetration of injected natural gas into
the the chamber and the mixing with the surrounding air to form an ignitable mixture.
Direct injection combustion begins as premixed burning near the region of = 1 then
continues largely as a diffusion burning process (see Figure 2.4). Ouellette and Hill [38]
1

Typically for natural gas = 1.31 giving a critical pressure ratio of 0.544.

Chapter 2. The Process of Injection and Ignition in a DING Engine

15

Figure 2.3: Principal characteristics of the expansion process for an under-expanded


jet [35]
describe a relationship for the penetration of transient turbulent jets as:
! 41
1
M
Zt =
t2
ch

(2.2)

where M is the jet momentum flowrate, ch is the chamber gas density, t is the injection
time, and is constant. They have shown this correlation to have have very good
agreement with several studies with a value of = 3.0 0.1. They also found this to
correlate well for underexpanded gas injections.
In the case of a liquid spray injection, as is the case with diesel, the fuel is atomized in the chamber as small droplets which then vaporize and mix with the air in the
chamber to form a flammable mixture. One would think that injecting fuel already in
its gaseous form would speed up the mixing process, leading to faster combustion rates,
however the opposite is true. As Abraham et al. [39] explained, an atomized liquid spray
provides quicker and more thorough mixing. In the case of a spray, the smaller droplets
quickly transfers its momentum to the air in the chamber while the larger droplets, with
more momentum, can penetrate deeper in the chamber and vaporize further down. This
transfer of momentum increases the local turbulence allowing for quicker mixing to reach
the flammability limits. In essence, the liquid droplets act as a vehicle that carries the
fuel for it to vaporize across the chamber.

Chapter 2. The Process of Injection and Ignition in a DING Engine

16

Figure 2.4: Schematic of direct injection combustion [14]

Gaseous fuels, in comparison, form plumes from the nozzle holes that form rich cores
and have a smaller flammable region. The jet created from an underexpanded injection
is also much more turbulent than the surrounding chamber gas even with added swirl
effects. Consequently, the mixing and penetration of a gas was shown to be heavily
dominated by its own turbulence. Increasing the swirl, squish ratio, and turbulence in
the chamber had little effects on mixing [39, 40]. Chamber turbulence comes into play
once the head velocity of the jet becomes of the same order of magnitude as that of the
chamber.Hence, once it reaches that point, swirl can have an affect on bending the axis
of the jet.
The findings from Jennings and Jeske [41] also agreed that mixing is largely dependent
on the injected gas own turbulent energy. It was also found that the fuel rich cores of the
plumes are proportional to the size of the nozzle holes of the injector and that chamber
penetration increases with hole size. Despite this, they concluded that mixing could
be improved by the use of multiple smaller holes in the nozzle while maintaining the
same effective area. The reason for this being that smaller nozzle holes produce smaller
rich cores improving the mixing with the surrounding air. There are some limitations

Chapter 2. The Process of Injection and Ignition in a DING Engine

17

to this due to interactions, which will be discussed in Section 2.1.3. In essence, what
is of importance to the combustion process is the rate and completeness at which air
is entrained by the injected fuel to form a flammable mixture. It is also worth noting
that simulations with the onset of combustion, as opposed to simply mixing, showed
reduced differences between the liquid and gaseous injectors but further emphasizes the
importance of optimizing all conditions, not only the injector, for a given fuel [42].

2.1.3

Jet Interactions and Confinement

The description of jet structures mentioned above described for the most part a free
jet entering a quiescent mixture. In practice however, the injected gas is confined to
the combustion chamber and can experience interactions with adjacent gas plumes or
the enclosing walls. Ouellette and Hill [36] observed strong jet interactions taking place
when confined. They concluded that the air that the jets entrain as they exit the nozzle
create a low pressure zone below the gas sheet since the surrounding chamber air cannot
fill the void fast enough. This causes the plumes to collapse downward below the nozzle,
greatly reducing the radial penetration. The severity of the collapse depended on the
pressure ratio between the injected gas and chamber as well as the poppet valve angle.
At a shallow poppet angle of 10 , jet penetration followed the predicted values closely.
At high poppet angles of 20 and 30 , the severity of the jet collapse was much more
apparent. Increasing the pressure ratio reduced this effect and increased jet penetration.
Essentially, if a sufficient volume of air surrounds the jets, then the change in pressure
will be minimized, which in turn reduces this collapsing effect.
Furthermore, several studies that were examining the effects of mixing experienced a
reduction in jet penetration due to wall interactions [36, 39, 43]. The Coanda2 effect was
taking place in which a fluid would attract itself to the nearby cylinder walls. Jennings
2

The Coanda effect is the name of the phenomenon described above where the flow of a fluid is altered
due to the entrainment of another.

Chapter 2. The Process of Injection and Ignition in a DING Engine

(a) Gas injection with minimized plume interac-

18

(b) Gas injection with merging plumes

tions

Figure 2.5: Simulated in-cylinder plume interactions [43]

and Jeske [43] in an attempt to minimize the interaction from the cylinder head simulated
different injection angles in the nozzle head. Their findings corroborated that of Ouellette
and Hill, that worse mixing was achieved at steeper angles than at shallow angles near
the head. Another approach they took to minimizing wall interactions was to adjust the
nozzle tip height at which the gas would be injected. It seems intuitive that the optimal
location in this case was halfway between the two chamber surfaces where the volume
of air surrounding the gas sheet is maximized. It was also noted that increasing the tip
height had a greater affect on reducing jet deflection than did the nozzle angle since this
did not affect the radial momentum flux.
Maximizing the air entrainment and rate of mixing in essence translates to maximizing the surface area of the injected gas plumes that comes in contact with a sufficient
volume of air. Although having multiple small holes is more beneficial, it is also crucial
to have sufficient space between each plume to allow for air entrainment to take place
since they can become more susceptible to the Coanda effect. Increasing nozzle holes

Chapter 2. The Process of Injection and Ignition in a DING Engine

19

while maintaining injection area in too close a proximity can lead to plume merger and
hence reduce the effective surface area (see Figure 2.5). Thus, Jennings and Jeske concluded that to obtain the best mixing, the injection nozzle should be equidistant from the
cylinder walls with the smallest injection angle to minimize interactions as well as have
the maximum number of holes that form separate gas plumes to maximize entrainment.
The ERDL combustion bomb used in this paper follows the above principles for injector
placement and design. Brombacher [24], using the combustion bomb with identical geometry, performed laser shadowgraph measurements of the jet penetration. At distances
close to the nozzle exit his results are in good agreement with Equation 2.2 but suggests
behaviour indicative of interactions as the jet approaches the side walls.

2.2

Auto-ignition of Natural Gas

Auto-ignition in DI engines is characterized by an ignition delay time which is the time


between the start of injection and the start of ignition. The start of injection is typically
taken as the the time when the pintle lifts off the seat whereas the start of ignition is
much more challenging to determine. To find the ignition delay time, two methodologies
that are widely used are luminous delay and pressure delay. The luminous delay is
commonly defined as the time from injector opening until the first luminosity is detected.
The pressure delay is widely defined as the intersection of the measured pressure in the
chamber with the extrapolation of the pre-injection pressure curve and its corresponding
3-sigma line, where the pre-injection pressure is measured for 10 ms prior to injection.
Another form of pressure delay would be to specify an increment in pressure rise, such as
a 0.25 atm increase in the chamber, which is intended to account for 0.10 atm of noise.
Generally, an acceptable ignition delay time is said to be less than 2 ms with anything
later than that resulting in combustion late in the expansion stroke; reducing the power
output and fuel conversion efficiency yielding poor engine performance [44]. This would

Chapter 2. The Process of Injection and Ignition in a DING Engine

20

also lead to increased emissions that would occur due to overleaning or misfiring. Diesel
fuel can easily meet this criteria under typical compression ratios of 16 : 1 to 22 : 1
due to its favourable auto-ignition properties. As mentioned, natural gas has a high
octane number of around 130 RON and is very adverse to auto-ignition. Furthermore,
the temperature dependence and type of reaction of natural gas differs significantly from
that of diesel fuels and hence cannot be evaluated based on the typical Cetane Number
system [45]. For these reasons, natural gas requires further investigation regarding its
ignition properties and will need some form of ignition assist to be able to perform
adequately under diesel-like conditions.

2.2.1

The Effect of Temperature and Pressure

Natural Gas requires chamber temperatures of 1200K 1300K to achieve auto-ignition


under the 2ms threshold without assistance. With BDC temperatures of 325K, this requires compression ratios in excess of 26 : 1 compared to cetane at 10 : 1 (See Figure 2.6).
Fraser et al. [46] performed auto-ignition tests in a combustion bomb to evaluate the effects of chamber conditions. They found that in-cylinder temperatures had a large effect
on ignition delay times, decreasing delays with increasing temperatures. Various natural gas blends were tested and show a stronger dependence than methanol and cetane
(see Figure 2.7). The challenge in attaining high in-cylinder temperatures is further
aggravated by the fact that the fuel is being injected at much colder temperatures. It
was shown that the mixing of the cold fuel with hot gases cause a drop in pressure and
the formation of endothermic radicals prior to ignition that further drop the temperature [47]. This also relates the delay time with engine load. At higher loads the engine
runs hotter, reducing delay time, but at low loads and cold start conditions delay times
would increase.
It was also found that changing the chamber pressure had a small effect on ignition
delay. Although higher pressures could help with the delay time, it was found to have

Chapter 2. The Process of Injection and Ignition in a DING Engine

21

Figure 2.6: Temperature and pressure map for natural gas auto-ignition under 2ms [46]

less than a first order chemical kinetic effect, much less than the effect of changing
temperature [48]. Knowing the trends from the temperature and pressure effects can
help improve auto-ignition but cannot solely determine its success. It is still a function
of the mixing and interactions previously described as well as the gas composition and
the means by which those temperatures are reached.

2.2.2

The Composition Effects of Natural Gas

The varying nature of the composition of Natural Gas brings into question its effects
on the combustion process. As mentioned, methane usually comprises 90 95% by volume but can have extremes of 75 98% depending on the production process. The
remaining components are ethane, propane, carbon dioxide and nitrogen with trace
amounts of higher order hydrocarbons and other gases. It is widely accepted that natural
gas composition has a significant effect on combustion with several studies supporting
this [44, 4649]. Kubesh et al [49] tested ten different synthetic gas blends primarily

Chapter 2. The Process of Injection and Ignition in a DING Engine

22

Figure 2.7: Pressure Delay vs Temperature for various fuels [46]

composed of methane and found that the octane number can vary by at least 20.
Looking more carefully, it was observed that natural gas that had a larger concentration of the minor species seemed to react faster and reduce ignition delay. As Naber et
al. [48] explains, most alkane fuels under diesel conditions produce HO2 radicals through
the major alkyl radicals. This creates chain branching that accelerates the overall reaction. The methyl radical in methane does not produce HO2 nor does it decompose
thermally but instead must be oxidized directly resulting in significantly slower chain
branching. Furthermore, the oxidation of the methyl radical has a relative large activation energy making it a slow chain propagating step [50].
As a result, having larger concentrations of higher order hydrocarbons that more
readily decompose and promote chain branching would reduce the ignition delay time.
Aesoy and Valland [44] found that NG composition with differing amounts of Ethane and
Propane can shorten ignition delay by a factor of 2 3 times when compared with pure
Methane; namely with delays decreasing with the increasing order of the hydrocarbons

Chapter 2. The Process of Injection and Ignition in a DING Engine

(a) Ignition delay time for natural gas blends

23

(b) Relative ignition temperature of two compound


blends

Figure 2.8: Ignition delay time and ignition temperature of various additives with
methane. The blends in Figure 2.8a mainly being CH4 /C2 H6 /C3 H8 : methane(100/0/0),
NGB(91/5.3/1.8), NGD(84/5.2/8.0) [44]

(see Figure 2.8). Conversely, the presence of inert gases such as N2 and CO2 are said
to hinder the combustion process. They cause a thermal effect of heat absorption as
opposed to a direct chemical effect, the degree of which is dictated by each gas heat
capacity and concentration.
Intuitively, the use of additives could be used to improve the ignition properties.
Agarwal and Assanis [47] performed simulations with hydrogen peroxide (H2 O2 ) as an
additive that resulted in a 35% decrease in ignition delay time over pure methane for a
95%/5% blend of CH4 /H2 O2 . They attributed its effectiveness to hydrogen peroxides
rapid decomposition in the early stages of combustion. Practically, however, the use
of additives is far more difficult. Aside from an increase in production costs, there is
uncertainty in ensuring that the correct proportion of NG to additive is delivered with
each injection, leading to high cycle-to-cycle variation. This is largely due in part to
the additive and higher order hydrocarbons condensing at the high storage pressures of
natural gas and therefore not uniformly mixing.

Chapter 2. The Process of Injection and Ignition in a DING Engine

2.2.3

24

Ignition Assist Methods and Hot Surface Ignition

To overcome the large temperature requirement for NG to auto-ignite, ignition assist


methods are required for it to successfully burn. A spark ignition would be difficult to
achieve under direct injection conditions since this would require accurate timing and
location to strike an ignitable mixture. This is problematic with natural gas due to
the mixing issues described in Section 2.1 and would result in high cyclic variation and
misfires. Another option that has been more commonly explored is that of pilot igniting
with diesel fuel, which can readily be ignited at temperatures of 900K under normal
CI engine conditions. This method entails the use of a dual fuel system and a more
complex injector system. Westport Research Inc. in collaboration with the University
of British Colombia have designed a dual fuel injector that involves two concentric shells
that deliver an early spray of pilot diesel followed by a CNG injection [18]. This managed
to successfully reduce emissions but also experienced varying performance in regards to
the alignment of the gas and diesel plumes which required optimization to minimize
negative interactions [51].
The method of hot surface ignition was also performed with some success. This is
typically done with a continuously powered glow plug that is placed in the combustion
chamber. The gas is then injected towards the hot surface and spontaneously ignites once
a flammable mixture reaches it. Aesoy and Valland [52] observed that the reaction starts
at the glow plug and propagates into the jet as a diffusion flame; also noting an effect from
the gas composition and glow plug location. Abate [25] observed multiple interactions
taking place. Both found that the optimal positioning of the glow plug to reduce ignition
delay lies in the middle of the gas jet, in the mixing flow region. Injection angle also
played a crucial role in regards to the hot surface contacting a stoichiometric mixture.
Furthermore, significant cooling of the glow plug was observed due to the injection of
much cooler gas.
Fabbroni [26], continuing Abates work, reported that a glow plug shield was crucial

Chapter 2. The Process of Injection and Ignition in a DING Engine

25

for obtaining acceptable delays. Comparing different sized shields on a two-plume injector, he noted multiple effects. First, the shield design needed to protect from the cooling
of swirl flow and injection yet allow the gas to come in contact with the hot surface.
Second, a larger shield greatly deflected the nearest gas plume (see Figure 2.9). This
often resulted in incomplete combustion since the flame was not able to spread across
both plumes due to an overlean region forming between them. Also, in agreement with
Abate, a decreased jet momentum would reduce the cooling effect of the gas on the hot
surface but the reduced mass would have to be compensated with multiple small holes.
Hence, the shield was designed as a cylindrical sleeve that would closely surround the
glow plug yet have a passageway facing the the direction of the injected plumes.
The shield also proved to serve as a pre-ignition chamber, increasing the residence time
of the gas in contact with the glow plug. It was also observed that the flame propagates
outward from the shield and ignites the plumes in one of three ways: either it propagates
along one plume and it ignites the edge of the other through a flammable path present
between them, or propagates up to the wall through one plume and passes to the second
plume in the dilution region and travels back toward the base of the second jet, or the
flame emerging from the hole ignites both plumes. The first two methods were dominant
with the use of the smaller shield and produced higher combustion efficiencies. With a
larger shield, the overlean region between plumes only allowed for the third mechanism
to take place if both plumes were to ignite.
In this study, hot surface ignition is used to expand on the work of Wager [28] and
Fabbroni [26]. The smaller glow plug shield design and position determined by Fabbroni
is considered to be the optimal case for the ERDL combustion bomb. Although in
practise continuously powered glow plugs experience very short life spans, owing to the
complexities of diesel pilot ignition, they provide a simpler alternative by removing the
diesel interactions and provide a more robust system characteristic of its basic design.
This allows for the analysis of the effect of nozzle hole geometry and potential soot

Chapter 2. The Process of Injection and Ignition in a DING Engine

26

Figure 2.9: Injected gas plume deflection due to glow plug shield [26]
formation from directly injected natural gas with less variables to consider.

2.3

The Effect of Nozzle Hole Geometry on Natural


Gas Combustion

The success of ignition is largely dependent on the rate at which the gas and air form
a flammable mixture. Given that the injected gas jet is only slightly affected by the
in-cylinder bulk gas motion, other parameters must be considered. The effect of nozzle
hole geometry, specifically elliptical holes, was examined in the hopes of faster mixing to
achieve higher combustion efficiencies.

2.3.1

Elliptical Jet Structure

The interest in nozzle holes with an elliptical geometry stems from a process unique to
it. The phenomenon of axis switching takes place in flows exiting elliptical holes where

Chapter 2. The Process of Injection and Ignition in a DING Engine

27

the major and minor axis exchange positions. In a circular hole, the viscous forces acting
on the jet and its shape are self-similar and maintain its symmetry as it penetrates. An
elliptical hole, on the other hand, creates a vortex that is inherently unstable due to
the variation in the azimuthal plane caused by self-induction and is not self-similar [53].
Hussain et al [54] expanded on this and explained it extensively. The velocity of a local
segment of this vortex is proportional to its curvature. As a result, a segment with a
larger curvature will move faster than that of a smaller one. This leads to the elliptical
vortex changing shapes and moving across planes (i.e. axis switching). As shown in
Figure 2.10, the plane of the vortex ring upon exit is parallel to that of the nozzle. Due
to its instability, the major axis will move ahead of the minor axis and form a fold that
induces a velocity in the minor axis inverting them. Axis switching can occur several
times over the life of the vortex ring before it collapses under the shear forces of the
surrounding fluid.
The above model is valid for elliptical holes with aspect ratios less than 3.5 : 1,
with anything greater experiencing more complex vortex splitting. Nonetheless, studies
have shown that a significantly larger amount of air has been entrained for turbulent
jets through an elliptical nozzle, experiencing between 3 to 8 times when compared to
circular nozzle [53]. The majority of the entrainment occurs in the minor axis where the
ambient air is drawn in as the major axis stretches outward. Schadow et al. [55] expanded
upon this for underexpanded jets and found that the same phenomenon applies and is
amplified by the shock structure of the jet. They observed that axis switching occurred
multiple times very rapidly near the exit and the mixing rate had increased.

2.3.2

Elliptical Nozzle Fuel Injectors

Studies have been done regarding diesel direct injection with non-circular holes based on
the above principles [5658]. They examined the effect of elliptical nozzle holes in vertical
and horizontal orientations as well as varied aspect ratios and compared them to their

Chapter 2. The Process of Injection and Ignition in a DING Engine

28

Figure 2.10: Axis switching phenomenon from elliptical jets [54]

circular counterpart. Generally what resulted is a negligible difference in performance.


Looking more closely, a visible trend was a reduction in NOx for the elliptical nozzle with
an increase in soot emissions and an increase in combustion delay. This was concurrent
with a lower rate of heat release (ROHR) observed in the diffusion burning period of
combustion. It was followed by the observation that at higher aspect ratios, these effects
were more apparent but were still not significantly discernible from the circular holes.
The above studies did not show a benefit for the use of elliptical nozzle with liquid
diesel sprays but must now be applied with a gaseous fuel. The experiments of Section 2.3.1 that showed significant mass entrainment were done for gaseous fluids and
noted that the majority of the entrainment is experienced near the nozzle exit. For diesel
fuel sprays, the liquid near the exit would tend to have much higher momentum and be
less scattered to entrain the air.
Following this rationale, Wager [28] performed experiments on the ERDL combustion
bomb to evaluate the effectiveness of a single hole elliptical nozzle with that of the circular
one from Fabbronis study. The results showed that the elliptical nozzle hole produced
jets with smaller rich cores, suggesting better mixing, but the overall area the jet envelops
was similar to that of the circular one. Lower peaks were observed with the elliptical hole
when comparing the ROHR and the curve showed a smoother transition to the diffusion
burning phase (see Figure 2.11); lower peak temperatures would lead to less NOx being

Chapter 2. The Process of Injection and Ignition in a DING Engine

29

Figure 2.11: Normalized ROHR curves vs CAD for elliptical and round nozzle holes [28]
produced but this was not quantified in the study. The elliptical nozzle also produced
more luminous flames but no conclusion was drawn relating this to soot production since
it also could not be quantified. Finally, the elliptical nozzle hole showed slightly lower
combustion efficiency which was hypothesized to be due to increased overleaning of the
fuel. In summation, a more in-depth look at elliptical nozzle holes was needed to quantify
its performance characteristics with respect to circular ones.

2.4

Soot Formation from Direct Injection Natural


Gas

Particulate emissions are generally thought to be greatly reduced with the use of direct
injection natural gas and this study will attempt to quantify the soot that comes from
this process. Generally, particulate emissions are comprised of a non-volatile fraction
and a volatile fraction. The non-volatile fraction is made up of solid carbon particles

Chapter 2. The Process of Injection and Ignition in a DING Engine

(a) Benzene

(b) Ethanol

30

(c) Acetylene

Figure 2.12: Soot derived from pyrolysis of benzene, ethanol, and acetylene [60]

known as soot, or black carbon. The volatile fraction is made up of organic species
from unburned fuel and lubricating oil and can either adsorb on surfaces or re-vaporize.
Despite the widespread belief that natural gas is a clean fuel, burning it under diesel-like
conditions still yields some particulates due to the nature of the soot forming process. As
mentioned, particulates are a regulated emission and pose a health risk; their standards
are, like all emissions, increasingly more stringent.

2.4.1

Principles of Soot Formation

Soot particles are primarily made up of Carbon with around 1% Hydrogen by weight
whose empirical formula is approximated by C8 H. As described by Palmer and Cullis [59],
they form spherical particles that can range between 100 to 500
A but can be as large as
2000
A. Each particle is comprised of about 104 crystallites with each crystallite consisting
of 5 10 sheets of graphite each in turn being 20 30
A in size. These particles can then
coagulate and agglomerate into larger fractal-like structures, with some variations due to
fuel structure (see Figure 2.12).
Best described by Heywood [14], the combustion in diesel engines of a directly injected
fuel can be separated into two parts. Initially when the fuel is injected, it vaporizes

Chapter 2. The Process of Injection and Ignition in a DING Engine

31

into a plume that has a rich core and an outer overlean region; between the two, a
flammable mixture is present (see Figure 2.4). Ignition takes place in this flammable
zone as premixed combustion, like that of spark ignition engines. Once ignited, the flame
propagates through the chamber and the remaining rich cores mix with the air in a
diffusion burning process. Typically in diesel engine, the majority of soot is formed near
the very rich fuel cores, with pyrolysis playing a large role. The central precursors to soot
arise from the fuel breaking down either via oxidation or pyrolysis to form acetylene and
polycyclic aromatic hydrocarbons (PAH). These go on to create particulate nucleation
from which the soot grows, and the particles can then agglomerate into larger particles
and attach themselves into fractal-like chain structures. The soot can then either oxidize
or escape as emissions.
Premixed combustion in engines has relatively no soot in comparison with its DI
counterpart. The reason being that the reactions that promote the formation of soot
are slower than that which oxidize it. Consequently an increase in temperature further
decreases soot. The tendency to soot under premixed conditions was shown to be independent of fuel structure but can in fact be related to the number of C-C bonds in the
fuel, increasing with more bonds [50].
In diffusion flames, as described by Glassman et al. [50], different isotherms exist that
can be used to map the flame from the fuel rich core to the very lean surroundings. The
incipient particle formation for soot is seen to occur at the 1650K isotherm. Several
studies were performed to vary the flame temperature by the addition of inert gases to
serve as a diluent. They conclude that as the difference between the flame temperature,
Tf , and the nucleation temperature, Tn , decreases; so too does the soot production. Also,
the soot that is produced creates a luminous visible front as it oxidizes whose luminosity
can be indicative of the amount of soot present. Therefore, an increase in temperature
increases both soot production and oxidation reactions, the net result of which determines
the final particulate emissions.

Chapter 2. The Process of Injection and Ignition in a DING Engine

32

Unlike premixed flames, the fuel structure holds a key role in dictating the path and
propensity to sooting. In general, the tendency for a fuel to soot follows the relation
described below:
Aromatics > Alkynes > Alkenes > Alkanes

(2.3)

The commonality to both premixed and diffusion burning processes lies in the formation
of acetylene and PAHs that lead to sooting but the routes at which they achieve this differ.
Soot formation in diesel engines is a very complex process. The rate at which it is formed
and oxidized depends on many factors that affect several reaction pathways. Clearly a
strong dependence on temperature exists but it is also influenced by the pressure, velocity,
diffusivity, and composition of the flame and fuel jet. Practically, most of the oxidation of
the soot occurs in the early stages of combustion and this can further be promoted with
better mixing of the fuel, with burnout of the particles stopping at temperatures less than
1300K (see Figure 2.13). The limitation on temperature with regards to soot burn-up is
dependent on the trade-off with NOx emissions that are exponentially aggravated with
temperature.

2.4.2

Soot Formation in a DING Engine

From the previous discussion, it is evident that natural gas has a low propensity to
soot being primarily composed of methane. The fact that it is used in direct injection
combustion means it undergoes both a premixed and diffusion burning phase and will
produce soot, albeit at much lower levels. A model simulating combustion bomb ignition of direct injection natural gas examined the effects of injection parameters on soot
production [61]. It showed that increasing EGR3 would increase the soot production and
reduce NOx , further confirming this trade-off. It also showed that the turbulent flows
from the incoming jet and in-cylinder gases played a key role in driving the concentrated
3

Exhaust Gas Recirculation (EGR) is a common technique used in reducing in-cylinder temperatures
to reduce NOx emissions.

Chapter 2. The Process of Injection and Ignition in a DING Engine

33

(a) Early and late burning elements with intermedi-(b) Fast and slow mixing for early burning elements
ate mixing

Figure 2.13: Soot particle burn-up rate in a diesel combustion environment [14]

soot to leaner regions to further oxidize. Hence, changing injection pressure and timing
significantly changed the in-cylinder fluid mechanics that impacts soot production.
An examination of the injection parameters was performed with a diesel pilot ignited
heavy duty engine [62]. It also found that particulate emissions increased with higher
EGR and later injection timing. Shorter delays between the pilot diesel and gas injection
would offset some soot emissions at higher EGR at the cost NOx . More importantly, it
was noted that PM emissions were reduced by higher peak heat release rates and shorter
combustion durations. It ultimately concluded that the best performance was achieved
by optimizing injection parameters for each operating condition of an engine.
Expanding on this work, Jones et al. [63] was able to quantify the particulate emissions, more specifically the black carbon, that come from the pilot fuel to that which
comes from the natural gas and oil. Using accelerator mass spectroscopy and a biodiesel/diesel blend for carbon tracing, they varied the load, EGR, and fraction of pilot

Chapter 2. The Process of Injection and Ignition in a DING Engine

34

Figure 2.14: Soot from pilot diesel for a given set of engine parameters: overall equivalence
ratio (), %EGR, pilot flow rate (low = 0.15kg/hr, high = 0.25kg/hr) [63]

fuel injected while keeping all other factors constant. They found that over the entire
range of tests, soot from the pilot injections ranged from 4 40%, meaning the remaining
60 96% was coming from the natural gas and oil (see Figure 2.14). At high load, the
highest amount of soot was generated but only about 5% of which was contributed from
the diesel pilot. At that condition, the oil was said to contribute only 20% of the total
PM emissions with the conservative assumption that no oil was burned in the combustion
chamber. Clearly the majority of the particulates were coming from the combustion of
natural gas.
Relating these findings to the ERDL combustion bomb, particulates were suspected
to be present during previous testing but was not quantified due to the scope of the
study. Wager [28] observed a more luminous flame with the use of elliptical nozzle holes,
suggesting a greater oxidation of soot. This could mean that either less soot was oxidized
with the circular holes or more soot was produced with the elliptical ones. The elliptical

Chapter 2. The Process of Injection and Ignition in a DING Engine

35

holes also experienced lower peak ROHR values which the previous discussion has shown
can decrease soot oxidation. Knowing that particulates from natural gas are significant
in exhaust emissions, this study will attempt to quantify the engine-out soot emissions
at different injector configurations.

Chapter 3
Experimental Apparatus
The apparatus used in this study is based from the original design of Cheung [23]. Over
years of testing, the design underwent improvements and minor modifications to be optimized to its current state. This chapter will therefore summarize the apparatus used
in testing and focus more heavily on the changes made in this generation of testing. An
overview of the CFR engine and Bomb is described in the literature written by Fabbroni
et al. [64] and a more detailed discussion of the generational changes can be noted in the
works of Wager [28], Fabbroni [26], and Abate [25].

3.1

The CFR Engine and ERDL Combustion Bomb

The experiments in this work are performed using a Cooperative Fuels Research (CFR)
engine. The CFR engine, typically used as the ASTM standard for automotive fuel
testing, is a single cylinder engine that offers the flexibility to change the compression
ratio and to modify components easily. The compression ratio is changed by either
increasing or decreasing the height of the cylinder, changing the clearance volume. The
current engine was overhauled and fit with an aluminum piston and modern piston rings
36

Chapter 3. Experimental Apparatus

37

to improve sealing and minimize blowby1 . The stroke remained unchanged at 4.5000 but
the bore was increased to 3.2600 for the overhaul. A plate with a counterbore underneath
each valve was added on top of the piston which occupies approximately 3.274 in3 to
further increase its compression ratio. This was necessary to compensate for the volume
increased by the addition of the combustion bomb. The engine is motored through a 3phase 208 V DC motor that is capable of attaining speeds of up to 400 rpm. This provides
the means of attaining rapid compression to simulate automotive engine conditions.
The ERDL combustion bomb was designed to mount on the side of the CFR engine
connecting through its spark plug port (see Figure 3.1). The air enters through the usual
intake port and is then driven into the bomb during the compression stroke. Natural
gas is injected through the back of the bomb and combustion occurs within the bomb
chamber. The gas is then expanded back into the CFR cylinder through the spark plug
port and evacuated during the exhaust stroke. The combustion bomb was designed with a
0.87500 thick A-1 optical grade quartz on the front side to provide a 200 optically accessible
window to observe the combustion event. An exploded view of the bomb is shown in
Figure 3.2.
The bomb is designed to simulate that of a medium sized diesel engine. Its combustion
chamber has a diameter of 50.8 mm (2.000 ) and a depth of 12.7 mm (0.500 ). The compression ratio of the engine incorporates the volume of the bomb chamber and the tangential
intake entry port as well as that in the CFR cylinder. This results in a clearance volume
of 44.60 cm3 , a displacement volume of 615.52 cm3 , and an effective compression ratio of
14.80 : 1.
The bulk gas swirl motion is generated by the tangential intake port of the bomb.
Lacking the means for direct measurement, the angular velocity of the gas was estimated
by Abate [25] using high speed imaging to track the motion of the ignited gas with the
1

Blowby is an undesirable condition at which the gases in the cylinder blow past the pistons sealing
rings

Chapter 3. Experimental Apparatus

Figure 3.1: 3-D view of CFR cylinder and ERDL combustion bomb [25]

Figure 3.2: Exploded view of ERDL combustion bomb and its components [28]

38

Chapter 3. Experimental Apparatus

39

flame front as the reference point. He approximated the bombs swirl velocity to be 3622
rpm at an engine speed of 230 rpm.
The CFR engine and bomb is run such that there is 1 fired cycle in every 5 engine
cycles; meaning that for every fired cycle there is 4 motored cycles that follow. The
reason for this is twofold: for one, it reduces the thermal stresses on the quartz window;
and just as importantly it clears the combustion chamber from any residual exhaust gases
from the previous fired cycle. It is said that skip firing under these conditions nominally
removes 99.99% of the residual gas in the chamber [64].

3.1.1

Combustion Bomb Heating

Given that the engine is skip fired, it generally remains quite cool relative to normal
engine conditions. This would create unfavourable and unrepresentative scenarios to
simulate standard engine operation. As a result, the combustion bomb is wrapped with
a rope heater that is pressed against it with plates that are fastened to the bombs body.
The heater is controlled via an Omega CN9000A temperature controller that has a PID
controller capable of self-tuning to optimize its response. A K-type thermocouple is
placed inside the rear block of the combustion bomb to monitor its temperature. This
temperature is referred to as the bomb wall temperature.

3.1.2

Intake Air Heating

To maintain hot in-cylinder temperatures, the air in the intake manifold is preheated.
The temperature of the induced air has a significant effect on the bombs peak temperatures and pressures. The air is heated using a 1000 W Chromalox heater that is placed
upstream of the intake port in the manifold. It is monitored with a K-type thermocouple
and controlled by a CAL9900 temperature controller that has a similar self-tuning feature
to the bomb heaters controller. The intake air can be heated anywhere from 50 C to
350 C.

Chapter 3. Experimental Apparatus

3.1.3

40

Intake Air Supercharging

The intake manifold is connected to a compressed air line to simulate supercharging


which is typical in diesel engines. Intake pressure from the regulator can range from 0
to 30 psig with the option of also operating under naturally aspirated conditions. The
pressure is monitored through a GM Manifold Absolute Pressure (MAP) sensor and is
set to the desired intake pressure using the regulator and the LabView interface. The
pressure is also recorded during motored and fired tests through the data acquisition
system. Its calibration curve can be found in Appendix A.

3.1.4

Engine Oil System

An external oil pump is used to lubricate the engine coupled with an external heating pad
placed underneath the oil sump. the oil pressure is adjustable through a bypass valve on
the oil pump but is maintained at 40 psig. The temperature is maintained between 40 C
and 50 C and is monitored by a K-type thermocouple placed in the oil sump. Heating
the oil assists in keeping the engine warm as well as reduces the cranking resistance of
the engine since the viscosity of the oil decreases with heat. The push-rods, rocker arms,
and valve stems are lubricated manually due to the exposed design of the CFR engine
components.

3.1.5

Engine Coolant System

Coolant is circulated through the engine block with a Haake N3 heating bath. Undiluted
ethylene glycol is sent through the engines coolant passages and, following the same
rationale of the previous systems, is heated with the Haake bath between 110 120 C.
The temperature is monitored with a K-type thermocouple that is placed in the return
line of the coolant system after it circulates through the engine.

Chapter 3. Experimental Apparatus

41

Figure 3.3: Glow plug position relative to injector nozzle [25]

3.2

Hot Surface Ignition System

As discussed in Chapter 2, natural gas is adverse to auto-ignition under diesel conditions


and requires some form of ignition assistance. Experiments conducted with the CFR
engine have used hot surface ignition as a means of inducing the combustion of natural
gas in the bomb via a glow plug. Much work has been done in the past to characterize
the performance of the glow plug and will be explained briefly.

3.2.1

Glow Plug Position and Orientation

The combustion bomb was designed to allow the use of the glow plug at three locations
that are at different radial distances from the injector nozzle. Early experimental analysis
concluded that the optimal location for ignition was at the nearest glow plug port. The
distance between the exit of the gas jet and the central axis of the glow plug at this location is 6.17 mm. Furthermore, the glow plug is inclined at 20 due to space restrictions
on the bomb itself, hence the distance of the nozzle varies slightly with the depth of the
chamber (see Figure 3.3). The most recent studies by Wager [28] and Fabbroni [26] all
used the nearest position exclusively and will continue to be used in this study.

Chapter 3. Experimental Apparatus

3.2.2

42

Glow Plug Surface Temperature

From the previous discussion, to achieve the successful ignition of natural gas, the glow
plug surface temperature would have to be higher than 1300K. Commercial glow plugs
used in diesel engines typically achieve lower temperatures and are only meant to be
used at start up as opposed to continuous operation. Initially the surface temperature
was controlled by a feedback system designed by Cheung [23] that used a relationship
between the electrical resistance and surface temperature to control the power of the
glow plug. This system was plagued with several issues, one of which was a very slow
response time.
As a result, Fabbroni [26] operated the glow plug continuously at the maximum
power to achieve ignition. Although they are typically rated at 5V and 50W, to achieve
the desired temperatures the glow plugs are powered to 120W but consequently have
a reduced lifespan. The glow plugs used throughout this study are Denso DG-143, the
same that were used and characterized in the work of Wager [28]. Operating at this
power yields a surface temperature of over 1400K, which was measured using an S-type
thermocouple and a glow plug shield simulator (shown in Appendix D).

3.2.3

Glow Plug Shielding

The conditions of operation of the CFR engine with the combustion bomb are ones that
create a very high swirl environment in the chamber. As a result, the glow plug is highly
susceptible to convective cooling by the bulk gas motion in the chamber. This cools
the glow plug the most at the time of ignition where it should be the hottest. Past
work has shown that a shield around the glow plug is essential to achieving successful
ignition [25, 26]. The shield basically consists of a thin cylindrical metal sheath that fits
closely to the glow plug and is fixed to the rear of the bomb wall and spans the length
of the glow plug that is in the chamber. It also incorporates a hole in its side facing the

Chapter 3. Experimental Apparatus

43

injector to accept fuel and air to contact the hot surface. It was observed that the shield
not only prevents convective cooling of the glow plug but also increases the residence
time of the fuel and air to stay in contact with the high temperature surface to improve
ignition.

3.3

Natural Gas Injector

The natural gas injector used in this study is based on the design by Green and Wallace [65] which was also used by Wager [28]. It uses a solenoid that is controlled by an
external driver circuit to rapidly lift a pintle that presses against a Vespel polymide seat
at the tip of the nozzle. A triggering pulse is sent to the driver which then sends a high
voltage pulse to the solenoid in the injector to create the lift. The injector sits at the rear
of the bomb block and is seated such that only the tip of the nozzle protrudes into the
chamber. The nozzle is very easily removable and can be interchanged with nozzles that
have varying hole patterns and geometry. The gas temperature and pressure are measured upstream of the injector body with a K-type thermocouple and AST4700 pressure
transducer, respectively. The calibration curve of the AST transducer can be found in
Appendix A.

3.3.1

Nozzle Hole Geometry and Configuration

Due to the modularity of the injector nozzle, experiments that explore the effects of nozzle
hole patterns and geometry can be easily performed. The current study will expand on
the work by Wager [28] who compared the performance of a single round hole with that
of a single elongated slot with a 2:1 aspect ratio. Following his rationale, the major axis
of the elongated hole is aligned with the axial direction of the chamber. This is owing to
the observation that elliptical jets tend to spread more quickly in the minor axis plane
and will try to avoid wall interactions in this particular combustion chamber.

44

Chapter 3. Experimental Apparatus

(a) Circular hole

(b) Elliptical hole in the single (c) Elliptical hole in 3 and 9


hole nozzle

hole nozzle

Figure 3.4: Enlarged image of injector nozzle holes

When producing the holes, the goal was to maintain the same area as its circular
counterpart to achieve comparable flow rates. Also, machining a true ellipse at such
small dimensions is not feasible due to several practical limiting factors. In light of
this, an elongated slot with rounded ends is created through a wire electric discharge
machining (EDM) process that plunges the cutting tool multiple times. A total of 5
plunges are performed to remove any cusps that may form along the wall to create a slot
with a 2:1 aspect ratio. The depth of the cut is 0.061500 and other details of the nozzle
can be found in Appendix D. Hence the reader should note that the holes referred to in
this study as elliptical holes are in fact elongated slots but Wagers work still showed
notable differences in performance between the two. An enlarged image was captured of
each hole, a sample of which is given in Figure 3.4.
In this study, nozzles with a 1 hole, 3 hole, and full scale 9 hole pattern of circular
and elliptical geometry will be tested. The elliptical 3 and 9 hole nozzles were newly
designed and machined for this study, the drawings of which are found in Appendix D,
while the rest were made previously. The holes are all spaced 40 apart to maximize
the number of individual plumes around the nozzle. The ramp up from a single hole
will likely accentuate the differences in emissions between the round and elliptical hole

45

Chapter 3. Experimental Apparatus

(a) Cross section of injector (b) Projected view of injector nozzle (c) Cross section of injector
nozzle tip for 3-hole pattern

nozzle tip for 9-hole pattern

Figure 3.5: Injector nozzle and hole patterns

geometries. With the possibility that plume interactions may exist, the 3 hole nozzle,
which maintains 40 intervals between holes, will be used as an intermediary that may
identify trends moving forward with the full scale nozzle (see Figure 3.5).

3.4

Fast Flame Ionization Detector

To detect the presence of hydrocarbons within the exhaust, a Cambustion HFR400 fast
flame ionization detector (FFID) is used. When a hydrocarbon is burned, it produces
ions that are nearly proportional to the amount of carbon atoms it contains. An FID uses
this principle by taking a sample of gas and burning it in a carbonless hydrogen flame by
which a high voltage ion collector produces a current. The current across the collector is
proportional to the concentration of hydrocarbons via the rate of ionization. The fast FID
greatly increases response time through the use of better fuel and sample handling as well
as closer proximity of the ionization chamber to the sample. The HFR400 also isolates
the ionization chamber from pressure fluctuations that may occur from the sample source.
This is critical in engine applications, which typically always have pressure variations,
for the FFID to have an accurate reading of hydrocarbons. Further details of this system

Chapter 3. Experimental Apparatus

46

can be found in the manual [66].

3.5

Isokinetic Particulate Sampling System

In order to gather a representative sample of particulates exiting the engine, exhaust must
be collected isokinetically. Isokinetic sampling occurs when the velocity at the entrance
of the sampling probe is equal to that of the gas flowing around it. Under idealized
frictionless flow conditions, this can be achieved via Bernoullis equation by balancing
the static pressure of the sampling probe and the exhaust duct at the point at which
the sample is drawn. Dennis et al. [67] however, demonstrated that significant friction
losses can take place that result in a velocity difference between the probe and exhaust
stream that appear isokinetic through pressure differentials. Adhering to this and other
practical limitations of the current setup, an isokinetic sampling probe was designed and
will be further described in the next section. It should also be noted that a new custom
built stainless steel exhaust manifold was installed to avoid any contamination of the
exhaust gas.

3.5.1

Sampling Probe Design

A schematic of the isokinetic sampling probe that was designed is shown in Figure 3.6.
Initially this was designed similar to that used by Dipchand [68], utilizing the concept of
balancing the static pressures at the tip of the probe with that of the exhaust manifold
at the sampling point. This was achieved by having two concentric stainless steel tubes
welded together. The smaller inner tube is the true probe by which the sample is collected
and passed to the filter. The larger outer tube seals a portion of the inner tube and acts
as means to measure the pressure near the probe. Within the sealed portion of the
outer tube 0.5 mm holes are placed around the diameter of the inner tube approximately
2.54 cm from the tip of the probe. Another 0.5 mm hole is placed in the manifold wall at

Chapter 3. Experimental Apparatus

47

Figure 3.6: Isokinetic sampling system schematic


the location of sampling. The differential pressure is then measured through a manometer
at the manifold wall and at the other end of the outer tube that is outside the exhaust
manifold (as shown by the dotted lines in Figure 3.6). Having a hole diameter of 0.5 mm
to measure the static pressure was determined by Prandtl et al. [69] to be sufficiently
small such that the fluid in the hole itself remains at rest and unaffected by the passing
flow of the gas through the duct. There were several issues with this design that led to
unreliable readings in the manometer, which will be explained in Section 4.1.5. In light
of this, the sampling probe was set to draw a constant average flow that was determined
by the total flow passing through the system.

3.5.2

Particulate Filter System

The filter assembly is placed downstream of the probe and upstream of the vacuum pump,
with everything upstream of the filter being made of stainless steel. A 47mm Gelman 2220
stainless steel filter holder is used with either a Pall Emfab or Tissuquartz filter which has
a typical aerosol retention of 99.9% 2 . The filters are kept in a Class 100 clean room set
2

Following ASTM D 2986-95A 0.3m (DOP) at 32L/min/100cm2 filter media.

Chapter 3. Experimental Apparatus

48

to the ambient conditions as per section 1065.190 of the U.S. EPA engine test procedure,
which requires an ambient temperature of 22 1 C and a dew point temperature of
9.5 1 C for gravimetric analysis 3 . The filters are measured and dessicated before and
after testing in the clean room using a Santorius SE2-F microbalance with a readability
of 0.1 g. Typically, a batch of filters would take approximately 2 days to acclimate to
the clean room conditions.

3.6

Data and Image Acquisition Systems

Through the course of this study operational parameters and experimental data is collected simultaneously with a high speed camera used to capture combustion images. Both
systems are run through a single 3.00 GHz Pentium 4 computer that runs on Windows
XP. The data and image acquisition systems are independent from each other through
the computer but are synchronized through an external timing circuit. Each component
of the system will be further described in the following sections.

3.6.1

DAQ Measurements and Hardware

The data acquisition (DAQ) system is comprised of a National Instrument AT-MIO16E-2 acquisition board. The board has 8 differential analog inputs, 8 digital I/O lines,
2 analog outputs, and 2 counters. It uses a 12-bit analog to digital converter for a
quantization error of approximately 0.02% of the voltage range. Full specifications of the
data acquisition board can be found in the user manual [70]. The board is connected to
an SC-2070 termination board which is in turn connected to a series of BNC connectors to
allow for quick connection of inputs and outputs. The board also has an on-board cold
junction reference for thermocouple measurements that is selectable through jumpers.
Further details on the SC-2070 board is available in its manual [71].
3

This results in a relative humidity in the range of 39.56 51.10%.

Chapter 3. Experimental Apparatus

49

To isolate the potential for signal cross-talk between channels, the boards scan clock
is set to its maximum 500000 Hz, which corresponds to a settling time of 2s. Based on
the experience of Fabbroni [26], it is still recommended to sample the signals in order of
increasing voltage to further minimize this effect.

3.6.2

Image Acquisition System

The combustion images are captured using a Princeton Instruments EEV 1152x298 Intensified Charge Coupled Device (ICCD). This produces an image through an array where
each pixel is a photosensitive capacitor that holds a charge proportional to the amount
of incident light on it. The array is then sent to the computer through the Princeton
Instruments ST-138 controller. The controller is the means by which the camera and
computer interface. The intensifier is controlled by the Princeton Instruments PG-200
Pulse Generator. It serves two purposes: for one it controls the amplification of light
proportional to the voltage set on the PG-200 by the user; and second, it behaves like a
mechanical shutter, not allowing light to pass when no voltage is applied. The triggering
of acquisition is all controlled through an external timing circuit that sends pulses to the
ST-138 and PG-200.
To have the camera operate a sufficiently high speeds, the ICCD is run in Kinetics
Mode through the WinView software. Instead of capturing an image onto the CCD
array and transferring it to the ST-138 controller, which is relatively time consuming,
the array is broken up into multiple segments, each of which is a frame. The ST-138
controller blocks all but one segment when an image is captured and records it to the
array. It then shifts the array so that the next segment is exposed and another image is
captured. In this manner the ICCD is able to capture 11 frames in quick succession and
is then transferred to the controller. More information on the camera, controller, pulse
generator, and software can be found in their manuals [7275].
The initial stages of combustion in a direct injection engine undergo premixed com-

Chapter 3. Experimental Apparatus

50

bustion which typically emits light in shorter wavelengths, mainly the blue spectrum.
The later diffusion burning stage tends to emit more in the yellow-orange spectrum as
more soot is produced and burned. To better capture these initial stages, a Kodak 47B
Wratten filter is mounted onto the lens which allows light from the 450-500 nm wavelength to pass. It is important to keep in mind that although the diffusion burning
process appears orange, it in fact emits across the entire spectrum of light. Differentiating between which phase of burning is taking place may not be clear from the images
alone and may need accompanying data to support it.

3.6.3

Timing Circuit

The external timing circuit triggers the start and the increment of data acquisition and
more importantly synchronizes the collection of data with the captured images from the
high speed camera. An AVL 360C shaft encoder connected to the CFR engine generates
5 pulses per crank angle degree, equivalent to a pulse every 0.2 degrees or 1800 pulses
per revolution. These crank degree marker (CDM) pulses are sent both directly to the
DAQ and indirectly to the ICCD camera.
The AVL also sends a single pulse every time the engine reaches TDC, once per
revolution. A secondary encoder wheel is setup to send a pulse every non-firing TDC,
once every cycle (i.e. once every two revolutions). Both of these pulses are sent to an
AND gate that mark the start of a new cycle and trigger the beginning of the data
acquisition. Once the DAQ begins recording, it uses the CDM pulses as its clock source,
recording a set of data at every pulse for a total of 3600 points per channel per cycle.
The CDM pulses that are indirectly sent to the camera are sent to the Variable
Window Data Acquisition (VWDA) circuit which in turn sends a pulse to the Adjustable
Burst Fire (ABF) circuit. The camera triggering circuit behaves like an AND gate
receiving pulses from the ABF and from the injector circuit and sending a pulse to the
PG-200 pulse generator to start capturing images.

Chapter 3. Experimental Apparatus

51

The VWDA allows users to select a window of image acquisition anywhere between
0-3600 of the cycle corresponding to 0-720 . The ABF circuit specifies the timing between
the frames of an image by removing some of the pulses passed on to it by the VWDA
window. Thus, setting the ABF to 0104 will pass one CDM pulse for every 4 that it
receives, spacing the frames 4 CDM apart. Both the VWDA and ABF are reset with
every non-fired TDC pulse. This allows full flexibility on the timing of injection, data
acquisition, and image acquisition while maintaining the correct timing between each
system.

Chapter 4
Validation of Experimental
Apparatus
Before performing experiments, it is important to characterize the performance of the
various components that make up the overall system. It is essential to have an understanding of those components to be able to properly identify the engines conditions
leading up to and at the time of injection and ignition. The following chapter describes
the characterization of the engine under motored conditions, the performance of the natural gas injector, and the performance of the fast FID used to measure the unburned
fuel.

4.1

Motored Engine Conditions

To properly interpret the data gathered during fired tests it is important to estimate the
in-cylinder conditions at the time of the combustion event. In order to do this, the engine
is motored at the same speed and with the same intake conditions that would be used
during a fired test. From the set of motored data, correlations are made to estimate the
temperature and mass in the cylinder at the time of combustion and are used later in
the calculations when processing the data of the fired tests.
52

53

Chapter 4. Validation of Experimental Apparatus

Over the course of the testing the intake conditions and operational parameters are
kept constant. The in-cylinder conditions, however, increase gradually with time from
the heat generated by compression and the glow plug power. This increase is reflected
in the bomb wall temperature that is recorded. A set of motored data therefore entails
an average of 100 acquired engine cycles and sets are collected at several intervals of
the bomb wall temperature that ranges from 200 245 C. This range will be further
explained in Section 4.1.2. The operating conditions are summarized in Table 4.1.
Table 4.1: Operating conditions of motored tests
Parameter

Value

Units

Engine Speed

230

rpm

Intake Air Temperature

220

Intake Air Pressure (abs)


Coolant Temperature
Bomb Wall Temperature
Glow Plug Power

4.1.1

194.3

kPa

120

200-245

120

Combustion Bomb Pressure

The motored pressure trace of the combustion bomb operating under the previously
mentioned conditions are recorded through a Kistler piezoelectric pressure transducer
and is shown in Figure 4.1. This trace encompasses all of the sets of data gathered
under the bomb wall temperature range listed and shows the average value. Although
the pressure transducer is placed in the bomb chamber, Chengs [27] model showed the
systems pressure to equalize very rapidly, making it safe to assume a uniform pressure
between the cylinder and the bomb.

Chapter 4. Validation of Experimental Apparatus

54

Figure 4.1: Pressure trace of combustion bomb under motored conditions averaged over
the entire bomb wall temperature range

Looking at the trace, it can be seen that at the start of the cycle, when the intake valve
opens, the pressure quickly increases to the boosted pressure of the intake manifold. Once
the intake valve closes, the pressure rapidly increases as the air is getting compressed. It
then decreases during expansion until the exhaust valve opens and pressure reaches that
of the exhaust manifold, which is at roughly 2 inH2 O below ambient. Over the entire
range of bomb wall temperatures recorded, the average peak pressure is 4354 22 kP a
and occurs at 355.79 0.17 CAD1 . Examining the cycle-to-cycle variation for each set
at a specific bomb wall temperature yields an average variation in pressure and location
of 5.37 kP a and 0.20 CAD, respectively. This shows that the CFR produces a very
consistent pressure for the given intake condition.
1

Standard error calculated with 95% confidence limits

Chapter 4. Validation of Experimental Apparatus

55

Figure 4.2: Temperature trace of combustion bomb under motored conditions averaged
over the entire bomb wall temperature range

4.1.2

Combustion Bomb Temperature

Following the same rationale as for the pressure measurements, bomb temperature measurements were recorded for the bomb wall temperature range. Measuring temperature
accurately, however, is significantly more challenging. These temperatures were recorded
using a 0.000500 K-type thermocouple placed in the bomb from the top at the mid-point
between the centre of the chamber and the wall. The resulting trace across all wall
temperature conditions are shown in Figure 4.2. The peak temperature occurred at
792.96 3.86 K and 354.93 0.78 CAD. The cycle-to-cycle variation in temperature
and location of a given set are on average quite small at 1.07 K and 0.64 CAD.
Unlike the pressure measurements, temperature measurements cannot be performed
during fired tests due to the fragility of the thermocouple. Given that that there is a
gradual increase in chamber temperature during operation,there is a need to account
for this change through a parameter that can be readily measured. Owing to this,

Chapter 4. Validation of Experimental Apparatus

56

Figure 4.3: Peak, TDC, and BDC temperature correlations


correlations have been made relating the measured bomb wall temperature to the chamber
peak temperature, temperature at TDC, as well as temperature at BDC. Bomb wall
temperature was chosen since it is the only parameter that changes during the operation
and is indirectly related to the chamber pressure. The temperature at these key points
are later used in calculations when processing the data. A plot of these correlations is
presented in Figure 4.3 and the following equations were derived:

2
+ 3.476 Twall + 318.8
Tpeak = 6.101 103 Twall

(4.1)

2
+ 3.493 Twall + 303.7
TT DC = 6.053 103 Twall

(4.2)

TBDC = 6.729 101 Twall + 488.4

(4.3)

It can be seen that the chamber temperature increases very slowly in comparison to
the wall temperature. There is, however, some uncertainty relating to the temperature
measurements. In particular, the response time of the thermocouple with respect to the

Chapter 4. Validation of Experimental Apparatus

57

rapidly changing cylinder environment was explored. Fabbroni [26] attempted to quantify
the error associated with this and found that the measured temperature was slightly
lower and lagging behind the models prediction. This finding is somewhat expected as
the residence time of the thermocouple at a given temperature is along the same order
of magnitude as its response time.

Furthermore, the temperature distribution across the chamber itself was measured in
the past to be varying significantly. The temperature was measured moving along the
vertical axis of the combustion bomb starting from the top wall. First, the temperature
was seen to increase moving toward the centre of the chamber. Second, from its design,
the gas enters with high swirl and moves clockwise around it starting from the top. With
this in mind, the temperature of the gas at the bottom of the chamber is cooler than at
the top owing to heat losses as it travels.

Also, there is likely an effect due to hot and cold regions that would be caused by the
glow plug and the nozzle tip, respectively. The glow plug increases the local temperature
significantly and may contribute to heating the gas that passes over it. The injector
nozzle tip, on the other hand, protrudes through the centre of the chamber and may
have a heat sink effect. Both of these factors are difficult to quantify with the current
equipment and are only mentioned as possibilities.

For the motored data collected, the thermocouple was placed from the top at the
midpoint between the centre and the wall in an attempt to record an average temperature.
The glow plug and injector were placed in the chamber as it would be during a test to
try and compensate for their effects. Despite the uncertainties listed, the temperature
measurements are considered to be acceptable for its intended purpose. Holding all else
equal during testing, the relative performance of each nozzle can be evaluated under
comparable circumstances.

Chapter 4. Validation of Experimental Apparatus

58

Figure 4.4: Motored intake manifold absolute pressure trace

4.1.3

Intake Flow Rate

The air flow rate of the engine was measured using a TSI Model 2013 laminar flowmeter
at the intake upstream of the heater. This provides a reliable method of measuring the
volumetric and mass flow rate of the engine. The intake manifold is boosted to an absolute
pressure of around 15 psig and measured with a MAP sensor at the intake. Figure 4.4
shows the manifold pressure with respect to CAD. A sharp drop can be seen when the
intake valve opens and the charged air enters the cylinder which is near atmospheric
pressure. Once the intake valve closes, the line re-pressurizes to its set value.
The volumetric air flow rate is shown in Figure 4.5. Intuitively, it behaves inversely
to the MAP trace, where the peak flow rate occurs during the IVO window. Running the
engine boosted, as opposed to naturally aspirated, greatly reduces the pressure fluctuations in the manifold. The average flow rate per cycle was measured to be 3.09 SCF M .
Taking this curve and integrating with respect to time then yields a mass flow rate
of 0.90 gAir /cycle. Given that this is measured directly and with the uncertainties in

Chapter 4. Validation of Experimental Apparatus

59

Figure 4.5: Motored intake air flow rate

the temperature measurements, this value is considered to be the more accurate value of
mass inducted into the system as opposed to calculating it through the ideal gas law. The
calibration curves for the TSI flowmeter and MAP sensor can be found in Appendix A.

4.1.4

Gas Exchange and Blowby Losses

Normally, an ideal engine cycle would undergo polytropic compression characterized by


a ratio of specific heats of approximately 1.3. Due to heat losses and blowby, the process
diverges from the ideal cycle. A P-v diagram is plotted in Figure 4.6 from the motored
cycle data along with a polytropic line demonstrating this divergence.
With the motored data, it is possible to calculate the mass in the cylinder directly
through the ideal gas law. By measuring the pressure and temperature and knowing
the cylinder volume at all points throughout the cycle, the calculation becomes trivial.
Since the total mass induced into the system is calculated from the intake flowmeter, it is
more relevant to plot the mass fraction with respect to CAD using the calculated values.

Chapter 4. Validation of Experimental Apparatus

60

Figure 4.6: Motored P-v trace with polytropic compression line

This creates a map for the cycle by which to find the actual mass in the chamber. The
resulting curve is plotted in Figure 4.7.
It should be noted that the pressure and temperature used for this calculation were
measured in the combustion bomb. Hence, the mass fraction plotted reflects the bomb
chamber more closely than that of the cylinder. The calculated value assumes a uniform
temperature across the entire volume where in fact the conditions in the cylinder may
differ slightly. Further examination of Figure 4.7 shows a distinct step shortly after BDC
during compression and expansion. This is likely due to the flow of air being induced from
the cylinder to the bomb and vice versa as it passes through the tangential passageway.
This was not apparent in the past due to the significant amount of blowby that was being
experienced which led to a very convoluted curve. Having a newly honed cylinder and
improved sealing around the bomb greatly reduced this effect. The calculated blowby in
this case ranges from 3 8% versus 20 35% from the past.

Chapter 4. Validation of Experimental Apparatus

61

Figure 4.7: Motored mass fraction in combustion bomb

4.1.5

Particulate Sampling Flow

As mentioned in Section 3.5.1, the initial probe design based on that of Dipchand [68]
encountered some issues. Two things were noted during the initial testing of the system:
some fluctuations were apparent in the manometer at all times and the manometer reading seemed to be affected very little by changes in the vacuum pump. Another limitation
of the system, was the inability to change the vacuum in relation to the fluctuations due
its slow response. Owing to these issues, the sampling was to be dictated by setting the
vacuum pump at a constant flow rate representing the average flow of an engine cycle.
This would take into account any fluctuations in the system as well as try to target a
representative value in the absence of real-time flow regulation. Using the ratio of the
area of probe entrance to that of the exhaust pipe, the flow rate that should pass through
the probe under isokinetic conditions can be determined by the following relation:
Qprobe
(AV )probe
Aprobe
=
=
Qexh
(AV )exh
Aexh

(4.4)

Chapter 4. Validation of Experimental Apparatus

62

Equation 4.4 reduces to be a function of the areas only, having the density and
velocity of the gases being equal. With the dimensions known, the flow rate that must
pass through the probe is 1% of that which passes through the exhaust, or in this case
0.0309 SCF M . The vacuum pump was calibrated with a Precision Scientific 63111 wet
test meter and was maintained at this value during all tests.

4.2

Natural Gas Injector Performance

Extensive work was done in the past to characterize the performance of the natural
gas injector and some of its key points will be briefly mentioned. The circuitry and
the injector body remain unchanged and its performance can be assumed to be the
same. The gas injectors response time was found to be very consistent at 0.6 ms and
was independent of the fuel pressure. This test was carried out by injecting gas into a
very small chamber along with the Kistler pressure transducer. The injectors response
time reflects the time between the injector receiving a signal and the point where the
transducer rises above 3 of its steady value. This value is used to incorporate this effect
when calculating the ignition delay of the fired tests.
As described in Chapter 2, the gas is injected under choked conditions. Choking will
occur at the point which has the smallest cross-sectional area. For predictable injector
performance, this should occur at the exit plane of the nozzle. However, as the needle
lifts from the seat, there is a moment where choking will occur upstream of the exit hole
at the pintle. For longer durations, the time that the the gas is choked at the seat is small
compared to the overall duration. For shorter durations, the behaviour would differ due
to the initial upstream choking. This is important to note although it does not have a
large impact on this study since the characterization of the nozzles were performed in a
different manner. The next section will describe the characterization of the nozzles that
will be tested.

Chapter 4. Validation of Experimental Apparatus

4.2.1

63

Injector Mass Flow Testing

Typically in the past, experiments were carried out with one or two nozzles where the
intake pressure, bomb wall temperature, fuel pressure, and injection duration were varied.
This placed a larger emphasis on examining the effects of the varying parameters with
the nozzles performance. As a result, it was necessary for them to characterize the nozzle
mass flow rate as a function of injection duration and pressure. In the current study, a set
of optimal conditions were selected based on those experiments and were kept constant
for each of the six nozzles. The goal here is to focus more on the difference in performance
between the design of the nozzles themselves, namely hole geometry, hole pattern, and
injection duration. Keeping all other factors constant isolates any difference observed
to the listed parameters. As a result, the injector mass tests for the six nozzles were
performed at the three specific injection durations that are outlined in the test matrix
(see Section 6.2).
The mass test was performed using the same procedure that was used in previous
works:
1. The test cylinder was evacuated using a vacuum pump.
2. The cylinder was weighed on a digital scale with a precision of 0.02 g.
3. The cylinder was connected to the injector and its valve opened.
4. Using LabView, MassTest C.vi was run after specifying the injection duration and
number of injections.
5. The valve on the cylinder was closed and disconnected from the injector.
6. The cylinder was weighed and the final mass recorded.
The average mass per injection was determined by dividing the difference in masses
recorded by the number of injections. Typically the number of injections varied between

64

Chapter 4. Validation of Experimental Apparatus


35.00

Mass of Fuel per Injection [mg]

30.00

25.00

3C

20.00

3E
1C
1E

15.00

9C
9E
10.00

5.00

0.00

1.00

1.50

2.00

2.50

3.00

Injection Duration [CAD]

Figure 4.8: Mass of Fuel per Injection vs Injection Duration

75 200 per test and tests were repeated 3 times. The injector showed excellent consistency within each case. The results are plotted in Figure 4.8 and the values presented in
the test matrix (see Table 6.2). The nomenclature used to identify the nozzles are intuitive: the number indicates the number of holes (i.e. 1/3/9) and the letter distinguishes
between the elliptical or circular hole geometry (i.e. E/C). When referring to the nozzles
as a whole, they are sometimes listed in the order of 3/1/9 as a convention to match the
order in which they were tested. Also, recall from Section 3.3.1 that the holes are always
spaced 40 apart to observe any plume interactions, if any.
Several trends can already be seen between the nozzles. As expected, the values
generally increase linearly with injection duration. They also increase with an increasing
number of holes. This as well is expected since more holes create a larger area through
which the gas can flow in the alloted duration. An anomaly occurs with the shortest
duration of the 3C nozzle but this is attributed to the Vespel seat where the pintle sits

Chapter 4. Validation of Experimental Apparatus

65

and will be explained later. Another, less apparent trend is that the slope of the values
appear to increase with an increasing number of holes. It can be seen that the 1 hole
nozzles have a slow increase with injection duration while the 9 hole ones increase much
faster. This is also due to the fact that there is a much greater area for the flow to pass
in the time it is open.
A key design parameter for the circular and elliptical nozzles to be comparable was to
match the same exit orifice area. In theory, with all else being equal this should produce
equivalent mass flow rate for a given number of holes. Practically, however, it was very
difficult to consistently machine such small holes to the specifications listed (the smallest
dimension being around 215 m). The total area for each nozzle is presented in Table 4.2.
The 1 hole and 9 hole circular and elliptical areas are in very good agreement but the 3
hole circular and elliptical areas vary by almost 24%. Looking at the mass test results,
there is a contradiction in performance. Despite being very close in hole area, 9E injects
consistently more than 9C and 1C consistently more than 1E. Furthermore, although the
3C nozzle is 24% larger in area than the 3E, it injects less gas at the two lower durations
than the 3E and the least of all nozzles at the lowest duration. But interestingly enough
at the highest duration, the 3C and 3E inject essentially the same amount of gas.
Table 4.2: Nozzle hole area and % difference between the circular and elliptical geometry
Area [mm2 ]
Nozzle Hole #

% Difference
Circular Elliptical

0.3780

0.2873

23.99

0.1239

0.1239

-0.01

0.8532

0.8408

1.45

The variation in performance can be explained by two factors: the first is a machining
issue and the second relating to the Vespel seat design. When the nozzles were designed,

Chapter 4. Validation of Experimental Apparatus

66

they had a tolerance of 0.00100 . Between the sealing face at the base of the nozzle, the
lip at which the Vespel sits, and the surface of the Vespel where the pintle sits; there can
be as much as, but not limited to, 0.00300 . The reason as to why it might be larger than
0.00300 is due to the difficulties associated with machining a surface that is 0.04100 wide
in a blind hole over 200 deep. This tolerance stack-up proved to be a significant challenge
and resulted in custom seats being machined for each nozzle.
The second factor affecting the injector performance is the design of the Vespel seat.
The seat itself rests on a thin surface near the tip of the nozzle. At the very end, the
seat has an undercut with four large holes cut crosswise on the same plane as the outer
exit holes. The rationale in the past was to have the seat holes cut large enough so as to
not have any interference with the flow regardless of its alignment with the outer hole.
Having only mainly tested 1 hole injectors in the past this did not appear to be an issue.
Having tested multi-hole injectors with a much larger surface area, the alignment of the
inner holes with the outer holes proved to be a significant issue, as seen in the 3 hole
nozzle. The 9 hole nozzles were redesigned to have a modified seat that eliminated the
undercut and this revised design should be applied to all nozzles moving forward. A
drawing of this simplified seat and the design it superseded can be found in Appendix D.

4.2.2

Injected Gas Plume and Chamber Bulk Gas Interactions

To better understand potential interactions that can occur between the injected gas
plumes and the chamber bulk gases, interactions were investigated using the 1 hole
circular geometry nozzle. The diffusion time of methane with air was first examined.
To compare its time scales with that of the jet penetration, injection, and burn duration;
an approximation of the characteristic laminar diffusion time was calculated from the
relationship in Glassman [50]:
tD =

rj2
D

(4.5)

67

Chapter 4. Validation of Experimental Apparatus

where rj2 is the radius of the nozzle hole and D is the binary diffusion coefficient. For
methane and air, the resulting characteristic time is 11.88 ms. From Cheng [27], who
modelled the injection of natural gas in the combustion bomb, the time for the plume
to reach the bomb wall was found to be 0.9 ms. Furthermore, at the longest injection
duration of the 1C nozzle, the actual injection time and the resulting burn duration lasted
1.57 ms and 5.21 ms, respectively. Clearly the diffusion process is much slower than the
other time scales involved in the injection process which illustrates the importance of the
early stages of air entrainment to improve mixing.
In addition to this, the effect of the turbulence in the combustion chamber on the
injected gas plume was explored. To characterize its behaviour, the momentum flux ratio
is calculated. The bulk gas motion in the chamber can be approximated as a solid body
in rotation with a uniform density being made up entirely of air [76]. As a result the
angular momentum can be described as:
b =

mb Db2 Nb
Ib
=
b
240

(4.6)

where the mb is the mass of the air in the chamber, Db is the diameter of the chamber,
and Nb is the rotational speed of the gas in the chamber. The rotational speed of the gas
at TDC was found by Cheng [27] to be 4621 rpm. Knowing the mass of the air and the
dimension of the chamber yields an angular momentum of 1.40 104

kgm2
.
s

The linear momentum of the injected gas is described as:


pf = mf vf

(4.7)

The velocity was approximated by using the value found by Cheng to be 477 m/s.
Using the lowest mass injected as a conservative approach yields a linear momentum of
3.00

kgm
.
s

The momentum flux ratio is then:


b
= 4.66 105
pf

(4.8)

It is apparent that the injected jets own turbulence is much larger than that of the
bulk gas in the chamber. This would suggest that the mixing process in the initial stages

Chapter 4. Validation of Experimental Apparatus

68

is indeed heavily dominated by the injected jet itself, with the bulk gas motion having
little effect on it. This is in agreement with the discussion of Section 2.1.2.

4.3

FFID Calibration and Performance

The fast FID used in this study was used to measure the unburned fuel at the end of
each cycle. The calibration was performed through LabView, running the FFIDCal C.vi
to record the readings. A mixture of methane and air of known concentration is fed to
the FFID and the reading is averaged over a span of 10 seconds sampling at 1kHz. The
average removes any fluctuations in the signal that can occur from noise or from the
pressure fluctuations in the supply line. The mixture concentration is controlled by two
separate Matheson rotameters whose calibration curves can be found in Appendix A.
The procedure to calibrate the FFID was as follows:
1. Open the FFIDCal C.vi and light the FFID following the standard operating procedure as per its manual [66].
2. Once the temperature has stabilized, open the valves connecting the methane and
air to the rotameters. Ensure both supply pressures to the rotameters are the same.
3. Connect the mixture supply line to the end of the probe and open the overflow
valve to exhaust any excess mixture.
4. Open the rotameters to the desired mixture and input the scale values into LabView
5. Set the sampling rate and time interval to record the data.
6. Run the program. It will convert the scale values to ppm as well as display the
data recorded over the set time interval and the final averaged value.
7. Record the final averaged value along with the concentration in a spreadsheet.

69

Chapter 4. Validation of Experimental Apparatus

45000

y = 4623.3x
R = 0.9853

40000

y = 3906.6x
R = 0.9978

35000

y = 3209.4x
R = 0.9994

Concentration [ppm]

30000

y = 2502.1x
R = 0.9754
25000

y = 2923.9x
R = 0.9994

20000

15000

y = 1344.7x
R = 0.9949

10000

5000

0
0

6
FFID Output [V]

10

12

Figure 4.9: FFID Calibration Curves


8. Repeat steps 4 to 7 to generate a curve.
The calibration curve generated is later used to process the data from the FFID
during the fired experiments and is used in the calculations. Throughout the experiments,
multiple calibration curves were generated. This was due to a loss of calibration that
occurred due to clogging of the FIDs sampling line during fired tests. The FID was
very sensitive to fuel and air pressure as well as to the pressures in its sampling and
combustion chambers. For this reason it can be challenging to match the calibration
once it has been lost. The curves used are plotted in Figure 4.9.

Chapter 5
Combustion Data Processing and
Heat Release Calculations
The procedure used to analyze the combustion data is largely unchanged from that of
past works [26, 28]. This procedure will be explained in detail in this chapter before
proceeding to the final set of results.

5.1

Determining Ignition Delay

The procedure used to calculate the ignition delay is that used by Fabbroni [26]. For a
given set of fired data, an average motored pressure trace is generated from all non-fired
cycles of the test. Then for each fired cycle, a pressure difference trace is created by
subtracting the fired cycle pressure from the average motored pressure. The motored,
fired, and differential pressure trace is shown with the injector current signal in Figure 5.1.
The start of injection is found by averaging the injector current signal during the
initial portion of the compression stroke then locating the point where the signal crosses
above the line representing 3 of its mean. The injector current signal is quite smooth
up to this point and does not pose an issue. The start of ignition is found with a similar
rationale by locating the point at which the pressure differential trace crosses the 3 line
70

Chapter 5. Combustion Data Processing and Heat Release Calculations71

Figure 5.1: Motored, fired, and differential pressure traces for a typical fired cycle
of its mean permanently. In the case of the pressure trace, spurious spikes were observed
to occur that can lead to a severe underestimation of the the ignition delay. This could
have been due to either the injection event itself or to early flame kernel development.
As a result, the approach to detecting the point at which the line crosses is approached
differently. The differential pressure trace is examined starting at its maximum value
and working backwards. The 3 line in this case is determined by averaging the base
pressure differential signal for 100 CAD before the start of injection. This was shown
to provide a sufficient number of points to describe the variability of the signal prior to
injection without being severely effected by local spurts. A closer view of the pressure
differential signal and its 3 line is shown in 5.2.

5.2

Heat Release Calculations

To quantitatively evaluate the combustion process, the net rate of heat release (ROHR)
is calculated for the combustion bomb system. This calculation is attractive as it avoids

Chapter 5. Combustion Data Processing and Heat Release Calculations72

Figure 5.2: Injector current and pressure differential with mean and 3 lines during
ignition event
the need to account for heat transfer losses during the combustion process. Instead, the
calculation uses empirical pressure, temperature, and volume with respect to crank angle
and retains the simplicity of treating the combustion chamber as a single zone system.
What results is the difference between the heat release in the chemical energy of the fuel
and the heat transferred to the walls. This is derived starting with the basic differential
form of the first law:
dU
dV
dmb
dQnet
=
+p
hb
d
d
d
d
where

dU
d

(5.1)

is the rate of change of sensible energy and p dV


is the work done on the
d

b
piston by the gas. The third term, hb dm
, represents the energy leaving the system
d

through the blowby losses, represented by the subscript b. Here it is assumed that
kinetic and potential energy terms can be neglected and that the sensible energy of the
fuel injected is quite small compared to the mass in the cylinder and is also neglected.
Also, the fluid in this case can be treated as an ideal gas which results in

dU
d

= d(mcv T )

Chapter 5. Combustion Data Processing and Heat Release Calculations73


and that the specific heats are a function of temperature only (i.e. cv = f (T ), cp = f (T )).
Hence, Equation 5.1 can be rewritten as:
dQnet
dT
dcv dT
dm
dV
dmb
= mcv
+ mT
+ cv T
+p
hb
d
d
dT d
d
d
d
The term

dcv
d

was expanded through the chain rule since

dcv
dT

(5.2)

can be calculated directly.

Taking the ideal gas law P v = mRT and differentiating with respect to yields Eq. 5.3.
dT
T dp T dV
T dm
=
+

d
p d V d
m d
Also keeping in mind that

dm
d

(5.3)

b
b
= dm
and that hb dm
= cpb Tb dm
, this can be
d
d
d

combined with Eq. 5.2 and written as:


dQnet
=
d







dcv V dP
dcv P dV
dm
2 dcv
cv + T
+ cp + T
+ cpb Tb + T
dT R d
dT R d
dT d

(5.4)

The blowby gas passes through a small crevice between the cylinder and piston and
its enthalpy will differ from that of the bulk gas in the chamber. To compensate for this,
the temperature at the halfway point of the thermal boundary layer formed between the
wall and the chamber is used for this calculation. Hence, Tb =

T Twall
.
2

The calculation presented in Eq. 5.4 produces a net rate of heat release curve with
respect to crank angle. The integral of this curve, then, produces the net heat released per
cycle. Most of the terms, however, cannot be drawn directly from the raw experimental
data. The volume and its first derivative are readily available through the engines
geometry. The calculations for the remaining terms will be further explained in the next
section.

5.3

Modelling In-Cylinder Conditions

To carry out the ROHR calculations, knowledge of pressure, temperature, and mass are
required during the fired cycles. Pressure and temperature can be readily measured

Chapter 5. Combustion Data Processing and Heat Release Calculations74


during motored tests and used to calculate the mass at a given point in the cycle. Temperature, however, cannot be recorded during fired tests as the thermocouple would
quickly be destroyed. As mentioned briefly, a functional model of the temperature and
mass distribution is required. Data collected from the motored tests are used for the
correlations and will be described in more detail.

5.3.1

In-Cylinder Mass Model

Under motored conditions, the mass can be calculated at each point in the cycle to generate a plot with respect to crank angle. This is possible since pressure, temperature, and
volume are measured at all points in the cycle. The average total mass of air entering
each cycle is also measured through the flowmeter during those tests. As mentioned in
Section 4.1.2, there are some discrepancies associated with the thermocouple measurements, especially near TDC. For this reason the total mass entering the system is taken
from the readings of the flowmeter. The calculated mass values are converted to a mass
fraction and were plotted in Figure 4.7. This mass fraction provides a map with respect
to crank angle by which to track the mass in the combustion bomb and is referenced to
the total mass entering from the flowmeter.
In the past, the plot for mass fraction was very disjointed as a result of the significant
blowby losses that were being experienced. This led to the use of convoluted correlations
to predict the system mass. With the engine overhauled and the blowby losses greatly
minimized, the mass fraction profile is much smoother and more consistent. As a result,
the total mass entering the engine is mapped out against the mass fraction profile directly
from the motored data.
It is interesting to note that four distinct regions can still be identified in Figure 4.7,
similar to past work but structured differently. The first zone begins at the start of the
cycle until the intake valve closes at 214 CAD. The mass can be seen entering the system
during the intake stroke at a steady pace. Once the IVC event occurs, there is a distinct

Chapter 5. Combustion Data Processing and Heat Release Calculations75


step in the curve which is attributed to the mass being pushed through the connecting
orifice into the combustion bomb during the compression stroke. The second zone, then,
describes the region from IVC to the peak mass fraction. Through the same rationale, the
third zone shows the gas leaving the bomb to enter the cylinder and the profile descends
from the peak until the exhaust valve opens at 500 CAD. Another distinct step occurs
at this point. The fourth and final zone encompasses the region from EVO to the end of
the cycle, representing the exhaust stroke.

5.3.2

In-Cylinder Temperatures

With the mass fraction determined through the motored profile and the volume and
pressure known, the temperature can be calculated at all points in the cycle. The temperature during the intake stroke is assumed to be constant and equal to TBDC which is
calculated from Eq. 4.3. The temperature is then calculated with the ideal gas law for
the compression and expansion strokes and is considered to be frozen during the exhaust
stroke at T540 . This was shown to have good agreement with the experimental motored
temperature curve.

5.3.3

Specific Heat of the Working Fluid

The specific heat as a function of temperature and its derivative is calculated from Equation 5.5 from C
engel and Boles [77]. This correlation is valid from 273 to 1800K, where
T is evaluated in Kelvin. The working fluid in the system is assumed to be air. With
a relatively small amount of natural gas being injected relative to the surrounding bulk
gas, this simplification should have a minimal effect on the calculations.

cp = 28.11 + 0.1967 102 T + 0.4802 105 T 2 1.966 109 T 3

(5.5)

Chapter 5. Combustion Data Processing and Heat Release Calculations76

5.3.4

Signal Derivative

Calculating the derivative from an experimentally measured signal is considerably more


challenging than from a known function. The noise inherent in a measured signal can
cause inaccuracies when calculating the derivative whos effects vary depending on the
method chosen. Simply calculating the difference quotients from the raw data can lead
to large amplification of the noise level and produce significant error in the heat release
calculations. On the other hand, smoothing the curve through segment averaging can
lead to underestimation of the peak values, the magnitude of which depends on the length
of the segment.
The Savitzky-Golay filter is applied to the measured data to remedy those issues. It
works by calculating a polynomial over some local interval of the data. The polynomial
regression determines the smoothed value of a point in the underlying function and its
performance is determined by the order of the polynomial as well as the size of the local
interval. This filtering method is ideal to smooth the signal and preserve the resolution
of the data at the peaks.

5.4

Combustion Efficiency Calculations

To evaluate the performance of combustion, two measures of efficiency are used. The
first is the fuel utilization which is defined as the percent of unburnt hydrocarbons that
escapes the combustion event. The FFID described in Section 4.3 is used to measured
the concentration of hydrocarbons in the exhaust after each cycle. The signal from the
FFID is converted to a molar concentration through the calibration curves presented.
It does not follow the traditional definition of combustion efficiency since the FID is
unable to detect oxidized carbon compounds as effectively (i.e. CO). Nonetheless, the
fuel utilization is defined in this study by Equation 5.6.
F uel% = 1

mf uel,unburned
mf uel,injected

(5.6)

Chapter 5. Combustion Data Processing and Heat Release Calculations77

Another measure of combustion performance is the thermal efficiency. In this case, the
work on the piston is evaluated during the compression and expansion stroke by finding
the net heat released by integrating the net ROHR curve that is generated from the
fired data. This value is then divided by the total energy input from the fuel calculated
from the Lower Heating Value and mass injected. Thus, it is actually the gross indicated
thermal efficiency that is calculated. It is described in Equation 5.7.
GiT E = 1

5.5

QN et
mf uel LHV

(5.7)

Particulate Data Processing

The particulate filters were sampled at a constant average flowrate that was determined
in Section 4.1.5 to approximated isokinetic sampling. Each filter would be placed in the
line for approximately 10 min having 0.0309 SCF M of exhaust pass through it. For the
gravimetric analysis, the Pall Emfab filters were used to collect samples. Each filter is
stored in the Class 100 clean room to acclimate where it is weighed before and after each
test. This provides a simple baseline to quantify the engine out particulate emissions.
The Pall Tissuquartz filters are used to sample exhaust under the same conditions as
their Emfab counterparts, however these filters are collected for OC/EC testing and sent
out to the National Resource Canada laboratory. This test will determine the ratio of
organic carbon to elemental carbon in the exhaust.
With the measured weights of particulates, the values are converted to a basis relative
to the power output. This is the conventional form when listing automotive emissions,
called specific emissions. Lacking a measure of brake power, using GiT E then yields the
PM emissions in reference to the gross indicated power and is described in Equation 5.8.

Chapter 5. Combustion Data Processing and Heat Release Calculations78

sP M =

m
PM
m
PM
=
P
m
f uel LHV GiT E

(5.8)

All tests operated at 230 rpm with 200 fired cycles recorded. With Equation 5.7 and
knowing the mass of particulates collected, the fuel injected per cycle, and the LHV of
the fuel; the calculation is straightforward.

Chapter 6
Experimental Results and Discussion
This chapter presents the results and a discussion on the findings of this study. It
is introduced by a description of the experimental procedure, the test matrix, and a
preamble explaining the presentation of the results.

6.1

Experimental Conditions

The experimental procedure for operating the CFR engine, gas injector, ICCD, and FFID
in conjunction with the DAQ remains unchanged from previous studies [26, 28]. A key
difference in the current study is a large improvement on the number of fired cycles
collected. Each test consists of a total of 1000 acquired cycles, 200 of which are fired.
This provides a more realistic basis for evaluating performance. It offers a significantly
larger amount of points to observe any statistical discrepancies as well as a long enough
time frame to observe any changes in performance, if any, with respect to the length of
operation.
Stemming from this, a key difference in measurement from past studies is the bomb
wall temperature. In the past, this was recorded as a constant throughout the test since
only 50 cycles were recorded. Currently, with the duration of a test recording 1000 cycles,
the temperature increases slightly during operation. Consequently, the wall temperature
79

80

Chapter 6. Experimental Results and Discussion

is recorded throughout the experiment and the average wall temperature is used in the
calculations.
All tests are carried out with the same intake conditions that are held constant and
equal to that of the motored test. These are summarized in Table 6.1. They also all
begin when the bomb wall temperature reaches 200 C. The cycles are then recorded and
the the engine is allowed to cool back to 200 C before the start of the next test.
Table 6.1: Operating conditions of fired tests
Parameter

Value

Units

Engine Speed

230

rpm

Intake Air Temperature

220

Intake Air Pressure (abs)

6.2

194.3

kPa

Coolant Temperature

120

Glow Plug Power

120

Test Matrix

The test matrix outlining each case is described in Table 6.2. To reiterate, in the past,
one or two nozzles would be tested and the operating parameters would be changed to
study their effects. The objective of this study is to focus more on the performance
of the nozzles themselves. Hence, a set of optimal parameters were chosen from past
works and kept constant throughout all the tests. The results would then reflect more
the differences in nozzle hole geometry and hole pattern distribution and be comparable
relative to one another. In addition, the values for the total mass injected, mass injected
per hole, and mass injected per area is listed beside the injection duration. The reason
for this will be further explained in Chapter 6.

Chapter 6. Experimental Results and Discussion

81

Due to the small amount of fuel injected, the lowest injection duration and all of
the 1 hole nozzles were not tested with Tissuquartz filters. The filters used in each test
is marked in the matrix as Em and Qz for Emfab and Tissuquartz, respectively. The
nomenclature used in the Nozzle Number column is the same that was described in
Section 4.2.1. The results that emerged from the this test matrix will now be presented.

6.3

Results Preamble

The following sections present the results of the study from the tests performed under
the outlined test matrix. From the data collected, several performance indicators can
be determined to evaluate each nozzle. These indicators are analyzed from the methods
described in Chapter 5.
To evaluate the relative performance of each nozzle, the results must be plotted against
the appropriate parameters. Plots were initially generated with respect to injection
duration, which was the initial control parameter of interest. However, since each nozzle
injects slightly different amounts of natural gas at a set injection duration, the results
were then plotted with respect to three parameters: mass injected, mass injected per
hole, and mass injected per nozzle hole area. These proved to be a more relevant basis
for the performance evaluation of each nozzle and will be discussed accordingly. It should
be noted that an increase in injection duration consequently leads to an increase in mass
injected and will be used synonymously.
Comparing the data on a mass injected basis examines the correlation between a
measured result and the amount of fuel injected. Subsequently, evaluating the results on
a mass per hole basis could better reveal the effects of the nozzle hole patterns and their
dispersion of the fuel. Finally, comparing on a mass per area basis would show the effects
of the differences in flow area, revealing what may differ with nozzle hole geometry. The
last point would be most apparent in the 3 hole nozzle case where the difference in nozzle

Hole Pattern

Elliptical

Circular

9C

9E

Elliptical

Circular

1C

1E

Elliptical

Circular

3C

3E

Hole Geometry

Nozzle Number

23.52
17.13

1.4

31.37
2.0

3.0

14.09

1.4

26.84

3.0

19.18

4.03

1.4

2.0

5.40

7.31

2.0

3.0

6.14

1.4

9.03

3.0
6.84

7.03

1.4

2.0

14.00

2.0

17.91

2.36

1.4
3.0

10.52

18.00

2.0

3.0

1.90

2.61

3.49

1.57

2.13

2.98

4.03

5.40

7.31

6.14

6.84

9.03

2.34

4.67

5.97

0.79

3.51

6.00

20.38

27.98

37.31

16.51

22.48

31.45

32.55

43.62

59.02

49.57

55.24

72.92

24.47

48.75

62.36

6.24

27.84

47.63

Injection Duration [CAD] Mass per Injection [mg] Mass per Hole [mg] Mass per Area [mg/mm2 ]

Table 6.2: Experimental Test Matrix

Em

Em/Qz

Em/Qz

Em

Em/Qz

Em/Qz

Em

Em

Em

Em

Em

Em

Em

Em/Qz

Em/Qz

Em

Em/Qz

Em/Qz

Filter

Chapter 6. Experimental Results and Discussion


82

Chapter 6. Experimental Results and Discussion

83

hole area is significantly larger between the two nozzle hole geometries.
Furthermore, it is important to monitor any trends that are apparent in the two
design parameters; namely, the nozzle hole pattern and the nozzle hole geometry. Nozzle
hole pattern refers to the comparison of the 3, 1, and 9 hole nozzles while nozzle hole
geometry refers to the comparison between the elliptical and circular holes for a given
pattern.
Thus the discussion is broken down by comparing the results for each nozzle hole
pattern and geometry to the mass injected, mass per hole, and mass per area. Each set
of data is plotted as a single point with error bars representing the 95% confidence limits
of the set. A summary of the test cases and their values is listed in Table 6.2.
Before proceeding to discuss the results, the overall equivalence ratio should be briefly
mentioned. The fuel to air ratio is simply the mass flow rate of fuel to the mass flow rate
of air in the system. The equivalence ratio, , is the operating fuel to air ratio, (F/A),
over the stoichiometric fuel to air ratio, (F/A)s . This provides a measure of proximity to
stoichiometric conditions. Typically, diesel engines operate at an overall equivalence of
= 0.5 0.6 at full load. The current conditions of the CFR yields an equivalence ratio
that can range from = 0.02 0.13 depending on the nozzle and injection duration.
This would reflect very light loading conditions, the significance of which will arise in the
discussion.
As a final note before moving on to the results, the figures generated from the results
are presented in its entirety in Appendix B. The figures presented in this chapter are
those selected from that more complete set to reinforce key points that are discussed.

6.4

Ignition Delay

The ignition delay was the first parameter to be examined. As mentioned, achieving delay
times of less than 2 ms is critical for acceptable engine performance. Before proceeding to

84

Chapter 6. Experimental Results and Discussion


2.0

1.8

Ignition Delay [ms]

1.6

3C
3E

1.4

1C
1E

9C
9E

1.2

1.0

0.8
1

11

16
21
Mass of Fuel Injected [mg]

26

31

36

Figure 6.1: Ignition Delay vs Mass of Fuel Injected

the main results, which are based on the set averages, a quick note should be mentioned
on the cycle-to-cycle variation. The ignition delay was chosen as the parameter to be
examined since it would also be indicative of the other performance measures. The
CFR engine during almost all tests performed with very little cycle-to-cycle variation in
ignition delay, having a 95% confidence interval of 0.05 ms or better with only very
few being 0.08 or better. Owing to this, the cycle-to-cycle variation was deemed to be
acceptable and was not considered to have a significant effect on the data. Returning to
the data set, all of the cases tested achieved ignition delay times of less than 2 ms. The
results are plotted in Appendix B, Figures B.1 to B.3.

6.4.1

Ignition Delay vs Nozzle Hole Pattern

Generally speaking, the ignition delay was quite consistent between nozzle hole patterns,
judging on a mass basis (see Figure 6.1). In this case, the delay times remained fairly

85

Chapter 6. Experimental Results and Discussion

2.0

1.8

Ignition Delay [ms]

1.6
3C

3E

1.4

1C
1E
9C
9E

1.2

1.0

0.8
0

5
6
Mass per Hole [mg]

10

Figure 6.2: Ignition Delay vs Mass of Fuel Injected per Hole


2.0

1.8

Injection Delay [ms]

1.6
3C

3E

1.4

1C
1E
9C
9E

1.2

1.0

0.8
0

10

20

30
40
50
Mass per Area [mg/mm2]

60

70

80

Figure 6.3: Ignition Delay vs Mass of Fuel Injected per Area

Chapter 6. Experimental Results and Discussion

86

constant for the most part between 1.4 1.6 ms. An exception arises with the 3E
nozzle,which has an ignition delay time that is slightly lower than the others. The
ignition delay time also seems to decrease at the very high mass injections provided by
the 9 hole nozzles. This is expected as higher operating temperatures result from a higher
mass of fuel injected, reducing the ignition delay time.
Re-examining the data with respect to mass injected per hole, the plot now differs
significantly where an inversing trend is apparent (see Figure 6.2). The nozzles that
injected the most overall fuel actually inject the least fuel per hole. The main implication
of this is that the plume of fuel injected will be different for each nozzle pattern. Taking
the 1 hole nozzles for example, which inject the most fuel per hole, would suggest that
there will be a larger fuel rich core that is very localized to a region of the chamber (i.e.
the region in which it was injected) and a very lean environment that surrounds it. In
this regard, the 3 hole nozzles would have a smaller fuel rich core and would encompass a
third of the chamber. The 9 hole nozzles, then, would provide the most uniform mixing
as they distribute the largest amount of fuel the most evenly.
Returning to the ignition delay time, no strong trend is apparent with the 1 and 3 hole
nozzles but a more visible difference can be seen in the 9 hole nozzle. In this case, there
seems to be a slight decrease in delay time with increasing mass. This is particularly
interesting because upon examination of the rest of the plot as a whole, there seems to
be no significant change with an increase in mass.
Now re-plotting the results with respect to mass injected per area yields a similar plot
to that on a mass per hole basis (see Figure 6.3). The reason for the similarity lies in the
fact that the difference in areas between the nozzles of the same nozzle pattern are very
small except for the 3 hole case where the areas differed by 24% (see Section 4.2.1). As a
result, excluding the 3 hole case, the position of the data of each nozzle pattern will stay
virtually the same relative to one another. Thus, the changes will be apparent in the
position of each pattern relative to one another. In terms of the ignition delay, the same

Chapter 6. Experimental Results and Discussion

87

trend is apparent as in the previous plots. Namely, the longer durations of the 9 hole
nozzles have a slight decrease in delay time due to the higher operating temperatures.

6.4.2

Ignition Delay vs Nozzle Hole Geometry

Returning to the plots, an emphasis will now be placed on the hole geometry and its
effect on a given pattern. Returning to Figure 6.2, the geometry does not seem to have
an effect on the 1 hole pattern, for the most part. The data for the 1E and 1C nozzles
show a fairly constant value of ignition delay regardless of the mass injected. Wager [28]
also noted very similar performance in ignition delay when comparing 1E and 1C.
Examining the 3 hole pattern reveals that there is a tendency for the the 3E to have
a slightly lower delay time than the 3C on average. It does not appear to have a strong
trend with respect to mass and is fairly linear. It is interesting to point out that the value
of 3C at the 3 CAD duration and that of the 3E at 2 CAD duration have an equivalent
mass per area. These two points can be considered the most comparable case for this
pattern with that basis. From that, it can be observed that the 3E produced a shorter
ignition delay time.
The 9 hole nozzles produced the distinct trend of decreasing ignition delay with
increasing mass associated with their higher operating temperatures. This was prevalent
in the larger injection durations that produced more that 20 mg per injection. It is also
worth noting that the 9E nozzle seemed to consistently produce lower delay times than
the 9C. Even at the longest duration of the 9C nozzle, the second longest duration of the
9E nozzle manages to post a lower ignition delay time. These results would suggest that
the 9 hole nozzles are capable of delivering an ignitable mixture much faster through
better mixing than the other two patterns and that the 9E nozzle in particular offers
even better mixing than the 9C.

Chapter 6. Experimental Results and Discussion

6.5

88

Combustion Image Data

When analyzing the combustion images it is important to remember that more fuel
produces a higher luminosity flame and that the flames propagation, in this case, depends
heavily on the nozzle hole pattern used. For this reason it is less relevant to compare
the differences between nozzle hole patterns when examining the images. Hence, the
discussion of this section will be divided based on a given nozzle hole pattern and will
focus more on the differences between nozzle hole geometries.
The images collected of the combustion event provided an excellent view of the flame
as it developed and propagates throughout the chamber. The glow plug is located underneath the centre of the window, below the nozzle tip. The swirl produced in the
engine rotates counterclockwise starting from the top of the window and its effect will
be apparent as the flame will be swept in that direction.
Each image strip shows the progression of the flame across 11 frames spaced equally
apart at 0.8 CAD intervals. They are then compiled into a series of 10 image strips placed
side by side. Due to the low luminosity of the 1 hole nozzles, the image intensity in those
instances were scaled to the 5%-95% range of the image. The 3 and 9 hole nozzles were
much more luminous and did not require this scaling. Histograms of the image were
also created to quantify the illumination of the pixels as the frames progress through
the cycle. They present the number of pixels that are illuminated in a frame beyond a
specified percentage of the maximum illumination. It can be seen that the initial starting
point of the histogram represents the number of pixels occupied by the glow plug. The
slight decrease shortly after can be attributed to the cooling of the glow plug after the
injection event.
Due to the age of the equipment, the ICCD would tend to miss one or two images
on occasion. This problem would tend to be aggravated at the higher luminosity cases,
indicating that it is an internal issue to the ICCD system. For this reason the image
number does not necessarily correlate to the fired cycle number and cannot be directly

Chapter 6. Experimental Results and Discussion

89

linked to the measured data. Furthermore, at high luminosities, pixels can become oversaturated and leak their charge to adjacent pixels in a process known as blooming. This
is apparent in a few images giving the appearance of two frames merging. Nonetheless,
very interesting information was gathered from the images that will be later shown to
support the other results that were gathered.

6.5.1

Images from the 1 Hole Nozzle Pattern

The 1 hole nozzle, as expected, produced the least luminous flames. A typical set of
fired images are shown in Figure 6.4 for the circular and elliptical case, respectively.
Visually there is very similar performance between the circular and elliptical geometry
with regards to the flame propagation. From the image histograms the circular geometry
consistently produced a higher luminosity flame. This is contrary to the findings of
Wager [28]. This is likely due to the fact that the circular nozzles injects slightly more
fuel than the elliptical one. At the lower injection durations the luminosity would peak
a second time near the end of the frame (see Figure 6.5). This was more apparent
in the circular nozzle but also occurred in the elliptical one. Given that this scenario
takes place in a very lean environment, one explanation could be that the flame passes
through rich and lean pockets in the chamber causing a change in the luminosity. This
is not considered to be a sign of compression ignition taking place since the rate of heat
release curve shows a smoothly declining profile (this will be discussed in greater detail
in Section 6.6).

6.5.2

Images from the 3 Hole Nozzle Pattern

Images gathered from test with the 3 hole nozzles offered somewhat expected results
as well and are presented in Figure 6.6. Luminosity is higher than in the 1 hole case
and, between the elliptical and circular geometries, similar luminosity peak values and
profiles are produced. From Figure 6.7, the histograms show a similar profile of increasing

Chapter 6. Experimental Results and Discussion

90

luminosity with respect to crank angle. The luminosity in this case rose after ignition
and would typically stay high until it decreases near the end of the strip. It should be
noted that the 3 hole cases had a large tendency to soot leading to fouling of the window.
This is apparent in the bottom right corner of the images where a darkened region can
be seen. The build up occurs at this corner as the soot is carried in the direction of the
swirl. This undoubtedly would reduce the luminosity detected by the camera and the
histograms produced. The region is quite small, however, and the images still provide
insight into the flames behaviour.
The tendencies of the 3 hole nozzles, though initially surprising, can be explained.
Having a concentrated fuel rich region would tend to produce a sooting environment in
the absence of air. With this pattern, a third of the chamber is exposed to the fuel and
the remaining air in the chamber surrounds it. Recall that the 3 hole nozzle produces
the most fuel per hole and hence the richest fuel cores. For this same reason, a longer
diffusion burning process can be taking place which would span the combustion over a
longer duration.

6.5.3

Images from the 9 Hole Nozzle Pattern

The 9 hole nozzle pattern produced the most interesting results and displayed a unique
behaviour compared to the other two patterns. Images in these cases were much more
luminous than those produced by the other patterns and are displayed in Figure 6.8.
Here, the luminosity would peak early in the frames and dissipate more rapidly than in
the 3 hole case. Interestingly enough, the flame re-emerges near the end of the frames with
high luminosity on several occasions. This occurs with both the circular and elliptical
geometries and is further reinforced by their histograms in Figure 6.9. This behaviour
was also apparent in the tests conducted at the 2 CAD injection duration, although not
as frequently.
This can be explained through the fuel distribution in regards to a nozzles geometry.

Chapter 6. Experimental Results and Discussion

91

Looking once again at the nozzles fuel per hole, it is apparent that the 9 hole nozzles
distribute the fuel much more evenly than that of the 3 hole nozzles. This lends itself
to more rapid mixing and combustion in the chamber since there is relatively more air
interlaced between the plumes. The images of the flame re-emerging seems to be a
sign of compression ignition taking place in the cylinder. Taking the longest duration
as an example, one possible explanation for this occurrence can be the following. The
gas injected into the cylinder is a transient process starting at 355 CAD and ending at
358 CAD. After it starts, combustion would begin and consume the fuel and air available
at a rate faster than the mixing process. During this time, the injection continues, causing
a fuel rich zone that escapes combustion until it is mixed with the remaining air. At the
same time, the pressure in the cylinder peaks from the previous combustion event leading
to the compression ignition of the the remaining fuel. This will be further reinforced by
the discussion in Section 6.6.

Chapter 6. Experimental Results and Discussion

92

(a) Circular

(b) Elliptical

Figure 6.4: Set of combustion images for 10 consecutive fired cycles for a 1 hole nozzle,
2 CAD duration (scaled to 5%-95% intensity range)

Chapter 6. Experimental Results and Discussion

93

(a) Circular

(b) Elliptical

Figure 6.5: Typical image histogram for a 1 hole nozzle, 2 CAD duration showing the
number of pixels above 50% intensity

Chapter 6. Experimental Results and Discussion

94

(a) Circular

(b) Elliptical

Figure 6.6: Set of combustion images for 10 consecutive fired cycles for a 3 hole nozzle,
3 CAD duration, 10 consecutive fired cycles

Chapter 6. Experimental Results and Discussion

95

(a) Circular

(b) Elliptical

Figure 6.7: Typical image histogram for a 3 hole nozzle, 3 CAD duration showing the
number of pixels above 50% intensity

Chapter 6. Experimental Results and Discussion

96

(a) Circular

(b) Elliptical

Figure 6.8: Set of combustion images for 10 consecutive fired cycles for a 9 hole nozzle,
3 CAD duration

Chapter 6. Experimental Results and Discussion

97

(a) Circular

(b) Elliptical

Figure 6.9: Typical image histogram for a 9 hole nozzle, 3 CAD duration showing the
number of pixels above 50% intensity

Chapter 6. Experimental Results and Discussion

6.6

98

ROHR and Net Heat Release: Curve Profiles

The rate of heat release (ROHR) was calculated for each set of fired tests using Equation 5.4 and the methods described in Chapter 5. Curves are generated showing the
net rate of heat release in J/CAD and the net heat released in J with respect to crank
angle. The profile and peak values of the curves give some insight as to how combustion
is taking place in the engine and a means to characterize it in addition to the images
presented. This section, much like the last, will break up the discussion in terms of nozzle hole pattern. The complete set of curves are presented in Appendix B, Figures B.10
to B.13.

6.6.1

Curves from the 1 Hole Nozzle Pattern

The calculated ROHR for the 1 hole nozzle produced a very smooth curve. Soon after
the injection event, a sharp rise in the rate of heat release can be seen and is followed by
a sharp decline soon after. There is no significant difference in the profiles between the
elliptical and circular geometries. The peak value of ROHR and the net heat released is
mainly dependent on the amount of fuel injected. Another thing to note is that the peak
of the ROHR curve, which is closely followed by the peak pressure, occurs much before
TDC in the cycle. Such behaviour is not desirable during normal engine operation and
is a function of the injection timing. This is an important point to keep in mind and will
be re-examined at the end of this section.

6.6.2

Curves from the 3 Hole Nozzle Pattern

The profile for 3 hole nozzles behaved much the same way as the 1 hole nozzles. Higher
peak values and a sharper rise resulted from the increase in fuel injected. These profiles
also experienced a sharp rise and decline that occurs before TDC. In terms of the hole
geometry, the circular nozzle seems to produce a sharp decline, whereas the elliptical

Chapter 6. Experimental Results and Discussion

99

has a somewhat smoother taper at the end of the decline. The net heat values however,
remain very comparable. Figure 6.10 shows the curves for the 3 hole nozzles.

6.6.3

Curves from the 9 Hole Nozzle Pattern

The 9 hole nozzles, once again, presented behaviour that differed from the other two
patterns. At the lowest injection duration, the ROHR curves produced were much wider
than in the other nozzles and the energy released was much smoother during the combustion process (see Figure 6.11). The peak pressure, in this case, occurred a bit after
TDC. This was an improvement but still undesirable for engine performance. At the 2
CAD duration, meaning a greater amount of fuel injected, the behaviour was similar to
that of the 3 hole nozzle. Now the curve experiences a sharp rise before TDC and an
equally sharp fall soon after. The profile was not as wide as that of the 1.4 CAD duration.
Intuitively, the peak values were increasingly higher as the amount of fuel injected was
increased.
The highest injection duration for the 9 hole case presented a unique scenario. First, to
reiterate, peak values increased as a result of increasing fuel in the chamber, as expected.
Also, similar to the past cases, the peak ROHR occurred before TDC and also experienced
a sharp rise and fall within a short span. More interestingly though, the behaviour at
the end of the curves decline was exclusive to this case and showed two distinct features.
For one, the dip after the peak was much lower than in the other cases, resulting a large
negative value for the rate of heat release. Second, there is a distinct increase shortly
after the dip that occurs after TDC. This rise can be associated with the re-emergence
of the flame taking place near the end of the frames in the combustion images and is
likely a sign compression ignition taking place. The negative value for the ROHR was
also apparent in the other nozzles but is greatly accentuated in this case. This suggests
that heat absorption is taking place in the system which is clearly erroneous. To explain
this occurrence, a discussion on the behaviour of typical diesel engines will be presented.

Chapter 6. Experimental Results and Discussion

100

(a) Circular

(b) Elliptical

Figure 6.10: Typical ROHR and Net Heat curves for a 3 hole nozzle, 3 CAD duration

Chapter 6. Experimental Results and Discussion

101

Normally in a diesel direct injection system, the injection starts at around 15 BTDC.
Following a typical1 rate of heat release curve, premixed combustion begins at 12 BTDC
and shifts toward diffusion burning until 15 ATDC. This seems be a much longer process
than what is taking place with the natural gas combustion in this study. However,
converting these values on a time basis for a typical diesel engine at 1800 rpm and the
values of this study operating at 230 rpm, the injection event takes place 1.38 ms prior
to TDC in a typical diesel engine as opposed to 3.62 ms in the CFR.
What is important to note is the occurrence of the peak pressure in the chamber.
According to Heywood [14], peak pressure should typically occur around 10 ATDC. This
would generally be the timing for the maximum brake torque (MBT). In the case of the
CFR engine, the peak pressure was examined to be very close to TDC. Furthermore, a
study by Lapuerta et al. [78] examined the sensitivity of calculating the ROHR for diesel
engines and found that having an early peak pressure would result in a calculation that
produces higher compression work and lower expansion. This affects the ROHR curve
by giving the appearance of heat absorption during the expansion stroke and a negative
value in the curve thereafter.
These findings present several implications. It is clear that the location of the peak
pressure is having an adverse affect on the ROHR curve. In essence, the energy being
released is being used to work against the compression of the piston, causing the sharp
rise and fall of the curve as well as the negative dip. This is felt in the calculations with
Equation 5.4 through large values of

dP
d

that resulted. And, upon examining the typical

start of injection in a diesel engine, the injection timing of the CFR is indeed much earlier.
As a result, this would likely lower the value of the net heat released that is calculated
but having this injection timing constant still allows for the relative comparison of the
nozzles performance.
It should also be mentioned that under the current conditions, the richest operation
1

Values for a typical diesel engine taken from Heywood [14]

Chapter 6. Experimental Results and Discussion

102

yielded an overall equivalence ratio of = 0.13, which was noted to be very lean compared
to normal diesel operation. This would indeed provide much faster mixing which would
further convolute any differences between the premixed and diffusion burning phases.
Nonetheless, the peak values of each nozzle are compared in the next section to evaluate
their performance in regards to one another.

Chapter 6. Experimental Results and Discussion

103

(a) Circular

(b) Elliptical

Figure 6.11: Typical ROHR and Net Heat curves for a 9 hole nozzle, 1.4 CAD duration

Chapter 6. Experimental Results and Discussion

104

(a) Circular

(b) Elliptical

Figure 6.12: Typical ROHR and Net Heat curves for a 9 hole nozzle, 3 CAD duration

Chapter 6. Experimental Results and Discussion

6.7

105

ROHR and Net Heat Release: Nozzle Performance

The following section examines the peak values for the ROHR and net heat released for
each nozzle to observe any trends between hole pattern and geometry. They are plotted
with respect to the mass of fuel injected, the mass per hole, and the mass per area. The
discussion will now be broken down in terms of hole pattern and geometry, like that of
the ignition delay. The complete set of plots for the peak ROHR and net heat values are
presented in Appendix B, Figures B.14 to B.19.

6.7.1

Heat Release vs Nozzle Hole Pattern

By first evaluating the performance of the nozzles on a total mass injected basis, a very
straightforward trend can be identified (see Figures 6.13 and 6.16). By increasing the
fuel in the chamber, the peak ROHR and net heat values increase. What can be seen
is that the plots are very similar in distribution. This is somewhat expected given the
similarity of the curve profiles. Each hole pattern seems to exhibit its own behaviour
which is reflected in the fact that they carry with them a distinct slope at which they
increase. The max ROHR for the 1 hole nozzles, for example, seem to rise slightly faster
than that of the 3 hole, which remains fairly flat. The 9 hole nozzles seem to increase
very rapidly with increasing mass in comparison. Looking at the net heat released, the
1 hole increases steadily with increasing mass but the 3 hole remains quite flat. The 9
hole, on the other hand increases rapidly and even seems to taper slightly at the higher
duration.
When regrouping the data on a mass per hole basis, the hole patterns are further
segregated (see Figures 6.14 and 6.17). The 1 hole nozzles form a distinct linear pattern
that increases with fuel mass. This is also displayed in the net heat plots. The 3 hole
pattern also reinforces what was seen in the mass plots and shows the least increase with

106

Chapter 6. Experimental Results and Discussion

700

600

Max ROHR [J/CAD]

500

400

3C

3E
1C
1E

300

9C
9E

200

100

0
1

11

16
21
Mass of Fuel Injected [mg]

26

31

36

Figure 6.13: Max ROHR vs Mass of Fuel Injected


700

600

Max ROHR [J/CAD]

500

400

3C

3E
1C
1E

300

9C
9E

200

100

0
1

5
6
Mass per Hole [mg]

10

Figure 6.14: Max ROHR vs Mass of Fuel Injected per Hole

107

Chapter 6. Experimental Results and Discussion

700

600

Max ROHR [J/CAD]

500

400

3C
3E
1C
1E

300

9C
9E
200

100

0
1

11

21

31
41
51
Mass per Area [mg/mm2]

61

71

81

Figure 6.15: Max ROHR vs Mass of Fuel Injected per Area


800

700

600

Net Heat [J]

500
3C

3E

400

1C
1E
9C

300

9E
200

100

0
1

11

16
21
Mass of Fuel Injected [mg]

26

31

Figure 6.16: Net Heat vs Mass of Fuel Injected

36

108

Chapter 6. Experimental Results and Discussion

800

700

600

Net Heat [J]

500
3C

3E

400

1C
1E
9C

300

9E
200

100

0
1

5
6
Mass per Hole [mg]

10

Figure 6.17: Net Heat vs Mass of Fuel Injected per Hole


800

700

600

Net Heat [J]

500
3C

3E

400

1C
1E
9C

300

9E
200

100

0
1

11

21

31
41
51
Mass per Area [mg/mm2]

61

71

81

Figure 6.18: Net Heat vs Mass of Fuel Injected per Area

Chapter 6. Experimental Results and Discussion

109

mass injection in both the ROHR and the net heat plots. The 9 hole pattern, contrary
that, displays the largest increase with respect to mass injected.
It should be noted that at the lowest duration of the 9 hole, the peak ROHR is
similar to that of the 3 hole case. However, the net heat release between the 3 and 9 hole
nozzle is vastly different, with the 9 holes heat release being over twice as much. This is
the result of the ROHR curve profiles, which suggest that the 9 hole nozzle has a wider
curve yielding a higher net heat release at the same peak ROHR. Examining Figures 6.15
and 6.18 further reinforces the trends described. In this case the linearity of the 3 hole
nozzles is slightly more apparent.
The behaviour with respect to nozzle hole pattern is linked to its distribution of the
fuel in the chamber. The 3 hole nozzle has a very small increase with an increase in
fuel. This is likely related to its ability to utilize the air around it. Given the rich core,
it is probable that it is consuming all of the available air already, with little room for
improvement. The 1 hole and 9 hole nozzles are surrounded with air that is more readily
available and can be utilized more efficiently with an increase in mass.

6.7.2

Heat Release vs Nozzle Hole Geometry

The differences in nozzle heat release with respect to the elliptical and circular geometries
are more apparent when comparing them on a mass per area basis, shown in Figures 6.15
and 6.18. Generally the 1 hole nozzle did not experience any apparent difference in values
or trends and performed fairly linearly with increasing mass. The 3C and 3E nozzles
however do show a slight difference in performance. Looking at the ROHR values of each
set individually, a deviation can be seen in the slopes of the 3E and 3C nozzles. That
of the 3E appears to have less of an incline while the 3C is slightly steeper. Looking
at the net heat values, the slopes are generally the same but the values of the 3E case
are slightly higher. Eventually there will be a limit by which the net heat release will
taper off but not seeing this limit yet, the 3 hole data is open to some interpretation.

Chapter 6. Experimental Results and Discussion

110

The flatter profile of the 3E data might be a sign that the most air is being utilized with
little room for improvement. There is a clear difference in the flow of gas between the 3E
and 3C nozzles. This is especially highlighted by their mass injected and hole area from
the injector tests. Recall the conclusion was that the 3E injected a larger amount of fuel
despite having a significantly smaller area. With that in mind and coupled with the fact
that the 3E produced shorter ignition delay times, it would further suggest better mixing
taking place in the 3E over the 3C.
The 9 hole nozzles proved to have the most consistent performance. The 9C and
9E seem to increase in unison and taper slightly at the highest duration. Between the
two, the 9C consistently produced higher peak ROHR and higher net heat release values.
This would suggest that the circular hole pattern in this case produces better mixing and
better combustion than the elliptical geometry. However, looking back at the discussion
regarding the ignition delay times, the 9E nozzle would clearly produce a lower delay time,
especially at the higher injection durations. There may be a difference in the behaviour
of the gases at the time of ignition and as the flame propagates in the chamber. Adhering
to this rationale, this would suggest that the elliptical geometry may produce an ignitable
periphery faster at the time of ignition then experience a more steady burn after. The
circular geometry, on the other hand, may produce a more powerful premixed burn as a
result of the initial slower ignition process, resulting in a higher peak ROHR.

6.8

Engine Efficiency Calculations

As a measure of performance, two types of efficiencies are calculated from the data
gathered from each nozzle. The first, is the fuel utilization to measure the amount of
unburnt hydrocarbons in the exhaust, which is found through the FFID, and is plotted
in Figures 6.19 to 6.20. The second, is the indicated thermal efficiency found by the
ROHR calculations and is plotted in Figures 6.21 to 6.22. The methodology explained in

Chapter 6. Experimental Results and Discussion

111

Section 5.4 is used to generate the values described. This discussion, as well, is divided in
terms of nozzle hole pattern and hole geometry. The complete set of plots for the nozzle
efficiencies are presented in Appendix B, Figures B.20 to B.25.

6.8.1

Efficiency vs Nozzle Hole Pattern

The fuel utilization was generally very good for the most part being over 94% for the
majority of cases. Looking at Figure 6.19, it can be seen that the utilization increases
with increasing fuel injected in all cases. It is also interesting to note the effectiveness
seems to be relatively independent of nozzle hole pattern and more strongly related to
the mass injected. For example, the higher mass flow cases for the 1 hole pattern has
similar fuel conversion efficiencies to that of the higher mass cases of the 3 and 9 hole
patterns. This can possibly be a result of higher localized temperatures from the increase
in fuel resulting in the higher consumption of fuel. The poorer performance of the 3 hole
case at lower mass can possibly be a combination of lower temperature and a fuel rich
environment.
The relative slope at which the efficiency increases seems to be different in regards
to the hole pattern on a mass basis. However, looking at Figure 6.20 on a mass per
area basis, the the slope at which the fuel utilization increases differs much less. The 1
hole and 3 hole nozzles appear to increase comparably while the 9 hole deviates slightly,
increasing at a faster rate. The 9 hole, with a combination of the amount of fuel and
its dispersion of it, would reach higher in-cylinder temperatures at a lower duration,
resulting in a higher percent of fuel being consumed.
Now looking at the gross indicated thermal efficiency calculated on a mass basis in
Figure 6.21 shows a grouping between the 1 and 3 hole nozzles and a segregation of the 9
hole pattern. Observing Figure 6.22, there is clearly a lot more scatter between the points.
The 3 hole nozzles efficieny in particular decreases significantly with increasing mass.
This can be traced back to the relatively flat profile of the net heat data in Figure 6.16.

112

Chapter 6. Experimental Results and Discussion

1.00

0.98

0.96

Fuel Utilization

0.94
3C
3E

0.92

1C
1E
9C

0.90

9E
0.88

0.86

0.84
1

11

16
21
Mass of Fuel Injected [mg]

26

31

36

Figure 6.19: Fuel Utilization vs Mass of Fuel Injected


1.00

0.98

0.96

Fuel Utilization

0.94
3C
3E

0.92

1C
1E
9C

0.90

9E
0.88

0.86

0.84
1

11

21

31
41
51
Mass per Area [mg/mm2]

61

71

81

Figure 6.20: Fuel Utilization vs Mass of Fuel Injected per Area

113

Chapter 6. Experimental Results and Discussion

0.8

Gross Indicated Thermal Efficiency

0.7

0.6

3C
3E

0.5

1C
1E
9C
9E

0.4

0.3

0.2
1

11

16
21
Mass of Fuel Injected [mg]

26

31

36

Figure 6.21: Indicated Thermal Efficiency vs Mass of Fuel Injected


0.8

Gross Indicated Thermal Efficiency

0.7

0.6

3C
3E

0.5

1C
1E
9C
9E

0.4

0.3

0.2
1

11

21

31
41
51
Mass per Area [mg/mm2]

61

71

81

Figure 6.22: Indicated Thermal Efficiency vs Mass of Fuel Injected per Area

Chapter 6. Experimental Results and Discussion

114

Having the mass of fuel increase at a rate faster than the net heat release would yield a
decreasing gross indicated thermal efficiency from Equation 5.7. The same holds true for
the 9 hole nozzle where the net heat release tapers off slightly at the highest duration and
a decrease in thermal efficiency can be seen. This can suggest that a mixing-controlled
burning is taking place more so than at the shorter durations. This becomes difficult
to distinguish due to the issue experienced with MBT timing and the location of peak
pressure discussed in Section 6.6. Unfortunately, the magnitude of the MBT timing issue
is unknown with the data that is available.

6.8.2

Efficiency vs Nozzle Hole Geometry

The fuel utilization efficiency, when comparing nozzle hole geometries, are again quite
similar. Performance between the elliptical and circular geometries are quite comparable.
The most notable case would be at the 9 hole geometry, where a difference in the slope
of the 9E and 9C is apparent (see Figure 6.19). It is difficult to make a a conclusion in
this regard since it seems to be more strongly related to the mass injected, and hence
the temperature, than to nozzle hole pattern or geometry.
The indicated thermal efficiency for the 1 hole and 3 hole nozzle shows a similar
performance between the elliptical and circular geometries. On a mass per hole basis,
the same trend is observed for both geometries and similar indicated thermal efficiency
values are achieved. On a mass per area basis (see Figure 6.22), the 3E seems to yield
a higher efficiency than the 3C. In the 9 hole nozzles, the same profile is visible in the
elliptical and circular holes, with a peak at the 2 CAD duration and a dip at the 3 CAD.
The 9C nozzle, however, consistently produced a higher indicated thermal efficiency. This
is undoubtedly a reflection of the ROHR and net heat values that were observed from
the 9C and 9E nozzles.
One thing to note, the 3C nozzle experienced odd behaviour at the lowest duration
during the calibration. It yielded only 2.36 mg of fuel per injection. During the discussion

Chapter 6. Experimental Results and Discussion

115

of the results it can be seen that its values for the ROHR and net heat do not correspond
to such a low delivery of fuel. This is also seen in the fact that the indicated thermal
efficiency in that case was over 100% and was disregarded from the graph. A possible
explanation for this may be the fact that the lift height of the pintle on the Vespel seat
was likely creating an area that was approaching that of the exit holes. This could have
potentially cause choking at the seat during calibration which could have shifted to the
exit holes during operation due to thermal expansion. Regardless, this emphasizes the
need for stricter tolerance control of the nozzles and the injectors Vespel seats.

6.9

Particulate Emissions

Particulate matter (PM) emissions are presented as a power specific value as discussed
in Section 5.5. The results from the OC/EC testing will be presented first, then followed
by the gravimetric tests. For all of the filters collected, the tests were performed in the
same manner, collecting 1000 cycles of exhaust flow, 200 of which are fired. As mentioned
previously, lacking the measurement for the brake power, the particulate emissions are
presented in reference to the gross indicated power being produced by the engine. The
plots generated for this section is presented in its entirety in Appendix B, Figures B.26
to B.38.

6.9.1

Organic Carbon vs Elemental Carbon

Tests were carried out to determine the amount of organic carbon and elemental carbon
in the particulate emissions samples. The organic/elemental carbon (OC/EC) tests were
performed with a smaller sample size as exploratory work. For this reason, only the 3
and 9 hole nozzles at the two higher injection durations were tested. These cases were
chosen to generate samples with the most particulates for the analysis. The findings of
these tests proved to be very insightful. Due to the small sample size, all points were

Chapter 6. Experimental Results and Discussion

116

plotted on the graph instead of a statistically representative point. Furthermore, the data
is labelled as circular or elliptical with the nozzle number and duration implied by the
value on the x-axis. The point of this was to emphasize the differences in performance
of the hole geometry, if any.
Proceeding to the results, it was noticed that a very high organic carbon content was
found in the samples. The range would typically be around 80%-100% of the total mass
on the filters. This is shown in Figure 6.23. Typically the inverse would be true, with a
much lower organic carbon content. During preliminary testing, a significant amount of
oil was captured on the filters at elevated coolant temperatures. This likely altered the
viscosity of the oil making it prone to pass through the piston rings. The test conditions
selected for this study appeared to have stopped this from occurring, although at the
time could only be confirmed visually. The high organic content is likely impacted by
the trace amounts of oil vapours being captured and condensed on the filter.
Looking at the elemental carbon content, which is plotted in Figure 6.24, the clear
trend was that the 9 hole produced less elemental carbon than the 3 hole nozzles. This
followed the rationale of the previous section. Between the elliptical and the circular
geometries with the 3 hole nozzles, it appears that the 3E nozzle produced slightly more
consistent results but overall seems somewhat comparable. The 9 hole nozzles performed
with even more consistency. In this case, the lower durations of the 9C and 9E nozzles
produced no elemental carbon. At the longer duration, the 9E seemingly produced less
soot than its circular counterpart. This appears to be a promising result to show better
mixing taking place in favour of the elliptical geometry. It should be re-emphasized that
the sample size is too small to have complete confidence in these conclusions, however
it served its purpose in presenting evidence to move forward with further testing of this
nature.
The organic carbon content is plotted in Figure 6.25, and exhibits slightly different
behaviour than the elemental carbon. For one, there is less scatter being observed in the

117

Chapter 6. Experimental Results and Discussion


100%

90%

Distribution of Organic/Elemental Carbon [%]

80%

70%

60%

50%

Elemental
Organic

40%

30%

20%

10%

0%
3C 2CAD

3C 3CAD

3E 2CAD
3E 3CAD
9C 2CAD
9C 3CAD
Nozzle Number and Injection Duration

9E 2CAD

9E 3CAD

Figure 6.23: Distribution of Organic and Elemental Carbon


3 hole nozzle pattern compared to the values of the elemental carbon emissions. The 9
hole nozzles still produce less emissions than the 3 hole nozzles, however now there is a
slight decrease with increasing mass of fuel injected, contrary to the elemental emissions.
Also, what is interesting to note is that the 9E nozzle produces higher organic carbon
emissions than the 9C, while the opposite is true for the elemental carbon. As mentioned,
the organic carbon content is likely influenced by the oil in the chamber and is difficult
to sequester the portion that comes from the fuel.

118

Chapter 6. Experimental Results and Discussion

specific Elemental Carbon [mg/kWh]

Circular
Elliptical

0
5

10

15

20
Mass of Fuel Injected [mg]

25

30

35

Figure 6.24: Elemental Carbon Emissions vs Mass of Fuel Injected


20

18

specific Organic Carbon [mg/kWh]

16

14

12

10

C Organic
E Organic

0
5

10

15

20
Mass of Fuel Injected [mg]

25

30

35

Figure 6.25: Organic Carbon Emissions vs Mass of Fuel Injected

Chapter 6. Experimental Results and Discussion

6.9.2

119

PM Emissions vs Nozzle Hole Pattern

For the gravimetric filter tests, all cases tested were weighed in the Class 100 clean
room. In this case, the particulate emissions of each nozzle hole pattern is very clearly
grouped together. This is is highlighted in Figure 6.26. Typically with engine emissions,
one would see a decrease in emissions with decreasing fuel consumption. Figure 6.26
shows a contrary trend as emissions decrease with an increase in fuel injected. Although,
when looking at the overall mass of particulates in Figure 6.27, there is a slight increase
with increasing fuel mass. To explain this, it is important to investigate the results and
behaviour of each nozzle hole pattern that was examined so far.
Looking at the 1 hole nozzles, a relatively high value for specific PM is occurring. First
it should be noted that the 1 hole nozzles produce the least amount of power output.
Second, there is a very small increase in the net mass of particulates collected across all of
the nozzles. As a result, the rate of change of particulates measured is much smaller than
that of the power output with increasing fuel. Hence, when normalized to the output
power, the 1 hole nozzles have the worst performance. This trend is also apparent when
looking purely at the 1 hole case, which can be seen by the decreasing slope of its points.
When examining the 3 hole geometry, a slight increase in PM emissions can be seen
with increasing fuel. This can also be attributed to the relatively weak increase in power
output. From previous discussions, this was linked to the fuels inability to utilize the
air due to its larger rich core. Following this rationale, the positioning of the 3 hole
data relative to the other two is not surprising. It produces more power than the 1 hole
nozzles to produce less g/kWh and produces much less power out than the 9 hole nozzles.
What can be contributing to its overall position on the plot would be the fact that the
conditions of the 3 hole nozzle should be the most prone to sooting. Recalling that rich
operation of diesel engines generates more soot, the 3 hole would produce the locally
richest regions.
The 9 hole nozzle, consistently produced the least amount of soot. The reason for

Chapter 6. Experimental Results and Discussion

120

this can be twofold. First, following the same reasoning as the other two cases, the 9
hole generates a much larger power output for a small increase in particulates. The
increase in power is followed in suit by higher in-cylinder temperatures which would
further consume any particulates being generated at a greater degree than the other
cases. The decrease with increasing mass would be dominated by the increasing power
with the relative steadiness of the particulate emissions.
To reiterate, it is clear from the OC/EC tests that the majority of what is collected
on the filter is organic carbon. This is likely coming from two sources: the products of
incomplete combustion and oil in the cylinder, the latter of which is suspected to be a
larger contributor. For this reasons, it becomes more difficult to isolate the differences
between the elliptical and circular geometries with the gravimetric tests alone, however
useful trends in the total particulates are still apparent.

6.9.3

PM Emissions vs Nozzle Hole Geometry

Comparing the particulate emissions with respect to hole geometry does not seem to
reveal a pattern in the 1 and 3 hole cases. It is evident that the amount of fuel injected
and the dispersion of the fuel through the hole pattern are much more influential than
the hole geometry. This reinforces what was seen in the previous discussions and further
emphasizes the subtleties of varying geometry. The 9 hole nozzle, however, did produce
a trend. The 9C consistently produced fewer particulates than the 9E nozzle (see Figure 6.26). Here, the 9E is seen to produce slightly more particulates and slightly less
power out resulting in the 9C having a lower specific PM. The difference is quite small
but seems consistent in the data. This may be related to a more in depth aspect of the
combustion process that was previously mentioned: the behaviour of the elliptical plumes
of gas may favour the early stages of combustion whereas the circular plumes may favour
the later diffusion burning stage. Overall, the 9 hole nozzles were seen to perform the
best, having the lowest specific PM.

Chapter 6. Experimental Results and Discussion

121

The central driving force of this study was to quantify the amount of particulate emissions from the CFR engine operating with direct injection natural gas. The PM emissions
are comprised of two parts, elemental carbon and organic carbon, both of which were captured on the filters. Lower particulate emissions were observed with the 9 hole nozzles
compared to the 3 hole nozzles, which reinforces that better mixing is taking place as a
result of the 9 hole pattern. The combustion images confirmed this result by showing
soot buildup at the corner of the chamber in the cases involving the 3 hole nozzles. Also,
luminosity in direct injection combustion is typically indicative of soot production. This
is confirmed through the histograms, which show a larger luminous portion in the 3 hole
nozzles over the 9 hole nozzles. Performing OC/EC tests differentiated the amount of
organic carbon and elemental carbon in the filters. An unusually high organic carbon
content was measured and is attributed to an oil control issue in the CFR engine. Thus,
the total PM emissions are affected by the organic carbon content coming from the oil
and is difficult to isolate the portion coming from the fuel. The elemental carbon content
can be attributed to the fuel alone and the 9E nozzle produced lower emissions than the
9C nozzle in this regard. This shows a promising potential in the elliptical hole geometry
that has yet to be fully explored.

122

Chapter 6. Experimental Results and Discussion

40

35
1 Hole Nozzles

specific PM [mg/kWh]

30

25
3C
3E

3 Hole Nozzles

20

1C
1E
9C

15

9E
10
9 Hole Nozzles
5

0
1

11

16
21
Mass of Fuel Injected [mg]

26

31

36

Figure 6.26: Specific PM Emissions vs Mass of Fuel Injected


0.30

Mass Collected on Filter [mg]

0.25

0.20
3C
3E
1C
1E
0.15

9C
9E

0.10

0.05
1

11

16
21
Mass of Fuel Injected [mg]

26

31

36

Figure 6.27: Mass Collected on Filter vs Mass of Fuel Injected

Chapter 7
Conclusions and Recommendations
The goal of this study was to examine the performance of six natural gas injector nozzles
each having a different design. The two central design parameters that were examined
were the injectors nozzle hole pattern and nozzle hole geometry. The nozzle hole patterns
evaluated included 1, 3, and 9 hole design. The 3 and 9 hole nozzle patterns had the
same 40 intervals between nozzle holes so that interactions between the injected gas
plumes were similar. Two nozzle hole geometries were evaluated, standard circular holes
and elliptical holes. The rationale for evaluating the elliptical hole geometry is based on
the principles found in literature, namely that a larger amount of air is entrained into
fluid jets from an elliptical orifice.
The tests were performed over 200 fired cycles at an optimal set of conditions that
were selected from previous work on the CFR engine. The purpose of the tests was to
compare the performance of the six nozzle designs, and especially to evaluate the potential
merit of using an elliptical nozzle. Injection durations of 1.4, 2.0, and 3.0 CAD were used
to achieve three different load settings. High speed images of the combustion event
were taken and histograms of the images produced. Key engine parameters were also
measured or calculated to provide further insight into each nozzles performance. Those
parameters include: ignition delay time, combustion images and histograms, ROHR and
123

Chapter 7. Conclusions and Recommendations

124

net heat release curve profiles, ROHR and net heat release peak values, fuel utilization
and gross indicated thermal efficiency, and particulate emissions. The conclusions will
now be summarized by parameter in the order listed.

7.1

Conclusions of Experimental Results

Ignition Delay
Peformance of the six nozzle designs was first characterized by the ignition delay
time, the key measure of ignition performance. All nozzles designs achieved an
average ignition delay time of less than 2 ms, the maximum value required for
satisfactory engine operation. The cycle-to-cycle variation of the ignition delay
times of each data set proved to be very minimal at 0.05 ms or better for the
majority of cases.
Nozzle Hole Pattern The 9 hole nozzle pattern performed the best overall, while
the 1 hole and 3 hole pattern produced similar ignition delays. The better
performance of the 9 hole pattern was attributed to the higher in-cylinder
temperatures that result from the greater amount of fuel delivered by the 9
hole nozzle.
Nozzle Hole Geometry The 3E nozzle posted a slightly lower ignition delay time
than the 3C when compared on a mass per area basis. The 9E also consistently
posted shorter delay times than the 9C. These results suggest that the elliptical
geometry provides better mixing, at least in the early stages of combustion,
and this leads to shorter ignition delays.

Chapter 7. Conclusions and Recommendations

125

Combustion Images
Images were taken of the combustion event and histograms were produced from
each set of images.
Nozzle Hole Pattern The combustion images revealed that the 1 hole nozzles
were indeed operating very lean. This was shown with the returning luminosity
after the initial ignition and was attributed to the fame travelling through rich
and lean pockets within the chamber. The 3 hole nozzles produced the longer
high luminosity segments which was indicative of the high sooting this nozzle
pattern experienced. This undoubtedly was a result of very rich localized
combustion which is reflected in its high mass of fuel injected per nozzle hole.
The 9 hole nozzles experienced the fastest rise in luminosity that would rise
and fall earlier than the 3 hole nozzle but re-emerge near the end. This was
attributed to compression ignition taking place in the unburnt fuel-air mixture
and was supported by the rate of heat release curves.
Nozzle Hole Geometry No significant difference between the elliptical and circular geometries was apparent in the combustion images.
Rate of Heat Release and Net Heat Release
The rate of heat release curve profiles demonstrated behaviours more apparent in
the nozzle hole pattern than in the nozzle hole geometry.
Nozzle Hole Pattern The 1 hole nozzle experienced the smoothest profile as it
rises and falls. The 3 and 9 hole nozzles experienced a very sharp rise and
fall in ROHR in a short period. The 9 hole experienced unique behaviour
in the sense that a small rise could be seen after the decline from the peak,
suggesting compression ignition is taking place. It was also noticed that the
timing for the peak pressure in the cycle was much earlier than MBT timing.
This was speculated to be the reason for the very large drop in ROHR that

Chapter 7. Conclusions and Recommendations

126

becomes negative. Needless to say the peak values increase with an increase
in fuel injected.
Comparing the peak ROHR and net heat release values of each nozzle revealed
further trends. The 1 hole nozzles displayed a steady increase with increasing
fuel injected, the 3 hole nozzles displayed a relatively at profile, while the 9
hole nozzles displayed the largest increase with increasing fuel. This behaviour
is attributed to the utilization of air of each nozzle hole pattern. The 1 hole
nozzle is readily surrounded with air but has a fuel rich core, hence its net
heat release increases steadily with more fuel. The 3 hole profile creates a very
rich localized area, and hence cannot utilize much more air. The 9 hole nozzle,
injecting the least fuel per hole, best distributes the fuel in the chamber and
can readily utilize more air with increasing fuel.
Nozzle Hole Geometry Looking at the net heat released with respect to the
nozzle hole geometry, the profile of the 3E can suggest better mixing as it
fully utilizes the air available while the 3C nozzle has yet to plateau, making
it difficult to draw a conclusion. The 9C has a higher peak value but a longer
ignition delay time than the 9E. This suggests that the 9E nozzle produces an
ignitable periphery more quickly but that the 9C produces a more powerful
premixed burn as a result of the slower ignition.
Fuel Utilization Efficiency
The fuel utilization efficiency was seen to be strongly dependent on the amount of
fuel injected, which correlates to a higher in-cylinder temperature. No strong trend
was visible in terms of the either nozzle hole design or nozzle geometry.

Chapter 7. Conclusions and Recommendations

127

Gross Indicated Thermal Efficiency


The gross indicated thermal efficiency was shown to be decreasing with increasing
mass of fuel injected. This was attributed to the relatively flat heat release profile
that does not increase as much as the increase in fuel delivered.
Nozzle Hole Pattern The 9 hole nozzles experienced a peak indicated thermal
efficiency that tapers off at the highest injection duration.
Nozzle Hole Geometry The differences between the elliptical and circular geometries are a reflection of the differences in the net heat released which may
be affected by the MBT timing position.
Particulate Emissions
The main objective for this study was to quantify the particulate matter (PM)
emissions from the CFR engine operating with direct injection natural gas, which
was done through isokinetic sampling of the exhaust.
Nozzle Hole Pattern The organic carbon and elemental carbon content was
evaluated for the two longest injection durations of the 3 hole and 9 hole
nozzles. It was apparent that a large portion of organic carbon was detected
in the exhaust which is attributed to an oil control issue in the CFR system.
The amount of organic carbon that comes from the fuel and from the oil is
not easily discernible. The elemental carbon, on the other hand, is most likely
completely from the fuel. Based on that, the 9 hole nozzles produced less PM
emissions than the 3 hole nozzles.
Nozzle Hole Geometry The 9E nozzle produced less elemental carbon than the
9C nozzle. The gravimetric analysis showed the best performance occurred
with the 9 hole nozzles, however a comparison of the total value between the
nozzle hole geometries is somewhat obscured by the high organic content.

Chapter 7. Conclusions and Recommendations

128

Overall Conclusions
The overall conclusions of this study can be described by the following:
Nozzle Hole Pattern This study consistently showed that the best fuel distribution was achieved with the 9 hole nozzle pattern. The 9 hole nozzles outperformed the 1 hole and 3 hole nozzles in every aspect of the evaluation. Mixing
performance was clearly demonstrated between the nozzle hole patterns. The
1 hole nozzle generally produced the lowest luminosity and heat released concurrent with a very lean flame. The 3 hole nozzle exhibited the behaviour
of very rich combustion. The 9 hole nozzle, meanwhile, exhibited the most
optimal combustion behaviour during this study.
Nozzle Hole Geometry It should be noted that the nozzle hole pattern and the
mass of fuel injected are more dominant factors than the nozzle hole geometry.
Nonetheless, a key takeaway in regards to the geometry, is that the holes with
an elliptical geometry demonstrated better mixing than those with a circular
geometry. This was most apparent by the lower ignition delay times and the
reduced elemental carbon emissions of the elliptical geometry. This study
reinforces the need for further exploration into fully characterizing the effects
of elliptical nozzle holes in direct injection engines, an area which has yet to
show its full potential

7.2

Recommendations

The universal goal of the work performed on this modified CFR engine is to better
understand direct injection natural gas combustion in reciprocating engines. As such,
it is ultimately tied to advances that would better simulate real engine conditions. The
work done in this study took a step forward by testing a full scale 9 hole nozzle. The
results not only showed that it was possible, but also that it yielded the best performance

Chapter 7. Conclusions and Recommendations

129

of any other nozzle hole pattern. Moving forward, tests should be conducted using only
full scale 9 hole nozzles.
The first thing that needs to be explored is the effect of operating at MBT timing.
As discussed, it was suspected that having an early peak pressure led to rate of heat
release values that were negative, suggesting heat absorption during the expansion stroke.
Changing the injection timing to operate at MBT would conclude whether this is indeed
the issue. This requires no hardware change and can be implemented immediately.
An underlying theme was that the injector hole pattern and mass of fuel injected were
the dominant parameters, showing that the hole geometry is much more subtle. In regards
to this, very precise control of the natural gas injector is required to better emphasize
the differences between elliptical and circular hole geometries. This means that great
importance must be placed in better characterizing the injector nozzles performance.
This encompasses two main things: the implementation of the simplified seat design to all
nozzles moving forward (see Figure D.7), and a sensitivity analysis in regards to varying
tolerance stack up errors. Tighter tolerances must be required to reduce inconsistencies
between nozzles. Although said quite plainly, this is no trivial task.
Owing to this, steps should be taken to progress toward engine operation without
the combustion bomb (i.e. directly inside the cylinder). This would improve conditions
on multiple counts. For one, there would be reduced heat losses from an improved
surface area to volume ratio. For the same reason, the particulate emissions would likely
reduce. Also, the plate installed on the piston could be removed and the oil control
rings would minimize the oil in the cylinder. A higher compression ratio could also
be used to better reflect diesel engine conditions and increase in-cylinder temperatures.
This would also remove the need to skip fire the engine and would allow for continuous
operation. Eventually the next step from this would be to operate under different loading
conditions. There are several hardware limitations regarding in-cylinder operation, the
main one being the physical installation of the nozzle for direct injection into the CFR

Chapter 7. Conclusions and Recommendations

130

cylinder.
Although not discussed in the paper, during preliminary tests, the injector would
seize at high bomb wall temperatures and continue working once cooled down. It was
concluded that the seat was operating near its rated temperature and the operation
during testing was kept below that temperature. This would cause issues during incylinder operation that would reach higher temperatures than the bomb. At that stage,
a redesign of the injectors sealing system may be required. Regardless, the effect of high
temperatures on the Vespel seat should also be examined.

References
[1] U.S. Energy Information Administration, International energy audit 2006, 2008.
http://www.eia.doe.gov/iea/overview.html.

[2] M. J. Economides and D. A. Wood, The state of natural gas, Journal of Natural
Gas Science and Engineering, vol. 1, no. 1-2, pp. 113, 2009.

[3] D. F. Merrion, Heavy duty diesel emissions, fifty years; 1960 to 2010, ASME
2002 Internal Combustion Engine Division Fall Technical Conference, vol. 39,
no. ICEF2002-478, pp. 115, 2002.

[4] T. V. Johnson, Diesel emissions in review, SAE Technical Paper Series, vol. 201101-0304, 2011.

[5] T. V. Johnson, Diesel emission control in review, SAE Technical Paper Series,
vol. 2008-01-0069, 2008.

[6] S. Yeh, An empirical analysis on the adoption of alternative fuel vehicles: The case
of natural gas vehicles, Energy Policy, vol. 35, pp. 58655875, 2007.

[7] M. Hekkert, F. Hendriks, A. Faaij, and M. Neelis, Natural gas as an alternative


to crude oil in automotive fuel chains well-to-wheel analysis and transition strategy
development, Energy Policy, vol. 33, no. 5, pp. 579594, 2005.
131

References

132

[8] R. Nichols, The challenges of change in the auto industry: Why alternative fuels?,
Journal of Engineering for Gas Turbines and Power, vol. 116, no. 4, pp. 727732,
1994.

[9] K. Hamai, H. Mitsumoto, Y. Iwakiri, K. Ishihara, and M. Ishii, Effects of Clean


Fuels (Reformulated Gasoline, M85, and CNG) on Automotive Emissions, SAE
Technical Paper Series, vol. 922380, 1992.

[10] L. Conti, M. Ferera, R. Garlaseo, E. Volpi, and G. Cornetti, Rationale of dedicated


low emitting CNG cars, SAE Technical Paper Series, vol. 932763, 1993.

[11] B. Finley and T. Daly, A three year comparison of natural gas and diesel transit
buses, SAE Technical Paper Series, vol. 1999-01-3738, 1999.

[12] B. Bartunek and U. Hilger, Direct Induction Natural Gas (DING): A Diesel-Derived
Combustion System for Low Emissions and High Fuel Economy, SAE Technical
Paper Series, vol. 2000-01-2827, 2000.

[13] B. Dhaliwal, N. Yi, and D. Checkel, Emissions effects of alternative fuels in lightduty and heavy-duty vehicles, SAE Technical Paper Series, vol. 2000-01-0692, 2000.

[14] J. Heywood, Internal Combustion Engine Fundamentals. New York, NY: McGrawHill Inc., 1988.

[15] J. Kubesh, Development of a throttleless natural gas engine, SAE Technical Paper
Series, vol. 2001-01-2522, 2001.

[16] D. Meyers, G. Bourn, J. Hedrick, and J. Kubesh, Evaluation of six natural gas
combustion systems for lng locomotive applications, SAE Technical Paper Series,
vol. 972967, 1997.

References

133

[17] J. Harrington, S. Munshi, C. Nedelcu, P. Ouellette, J. Thompson, and S. Whitfield,


Direct injection of natural gas in a heavy-duty diesel engine, SAE Technical Paper
Series, vol. 2002-01-1630, 2002.
[18] K. Hodgins, P. Hill, P. Ouellette, and P. Hung, Directly injected natural gas fueling
of diesel engines, SAE Technical Paper Series, vol. 961671, 1996.
[19] M. Kamel, E. Lyford-Pyke, M. Frailey, M. Bolin, N. Clark, R. Nine, and S. Wayne,
An emission and performance comparison of the natural gas Cummins Westport
Inc. C-Gas Plus versus diesel in heavy-duty trucks, SAE Technical Paper Series,
vol. 2002-01-2737, 2002.
[20] M. Frailey and P. Norton, An evaluation of natural gas versus diesel in mediumduty buses, SAE Technical Paper Series, vol. 2000-01-2822, 2000.
[21] H. Jaaskelainen and J. Wallace, Performance and emissions of a natural gas-fueled
16 valve DOHC four-cylinder engine, SAE Technical Paper Series, vol. 930380,
1993.
[22] J. Kubesh and D. Podnar, Ultra low emissions and high efficiency from an onhighway natural gas engine, SAE Technical Paper Series, vol. 981394, 1998.
[23] A. Cheung, Design of an optical access engine for combustion research, Masters
thesis, Department of Mechanical and Industrial Engineering, University of Toronto,
1997.
[24] E. Brombacher, Flow visualiztion of natural gas fuel injection, Masters thesis,
Department of Mechanical and Industrial Engineering, University of Toronto, 1997.
[25] V. Abate, Natural gas ignition delay study under diesel engine conditions in a
combustion bomb with glow plug assist, Masters thesis, Department of Mechanical
and Industrial Engineering, University of Toronto, 2001.

References

134

[26] M. Fabbroni, Flame propagation in natural gas fuelled direct injection engines,
Masters thesis, Department of Mechanical and Industrial Engineering, University
of Toronto, 2004.
[27] S. X. Cheng, Modeling Injection and Ignition in Direct Injection Natural Gas Engines. PhD thesis, University of Toronto, 2008.
[28] D. Wager, The influence of elliptical nozzle holes on mixing and combustion in
direct injection natural gas engines, Masters thesis, Department of Mechanical
and Industrial Engineering, University of Toronto, 2008.
[29] W. Rizk, Experimental studies of the mixing processes and flow configurations in
two-cycle engine scavenging, Proceedings of the Institution of Mechanical Engineers,
Series E, vol. 172, pp. 417422, 1958.
[30] J. Turner, The starting plume in neutral surroundings, Journal of Fluid Mechanics, vol. 13, pp. 356368, 1959.
[31] S. Abramovitch and A. Solan, The initial development of a submerged laminar
round jet, Journal of Fluid Mechanics, vol. 59, no. 4, pp. 791801, 1973.
[32] P. Witze, The impulsively started incompressible turbulent round jet, Sandia
Laboratories Report, no. SAND80-8617, 1980.
[33] A. Birch, D. Brown, M. Dodson, and J. Thomas, The turbulent concentration field
of a methane jet, Journal of Fluid Mechanics, vol. 88, no. 3, pp. 431449, 1978.
[34] G. Hyun, M. Nogami, K. Hosoyama, J. Senda, and H. Fujimoto, Flow characteristics in transient gas jet, SAE Technical Paper Series, vol. 950847, 1995.
[35] B. Ewan and K. Moodie, Structure and velocity measurements in underexpanded
jets, Combustion Science and Technology, vol. 45, no. 6, pp. 275288, 1986.

References

135

[36] P. Ouellette and P. Hill, Visualization of natural gas injection for a compression
ignition engine, SAE Technical Paper Series, vol. 921555, 1992.
[37] A. Birch, D. Brown, M. Dodson, and F. Swaffield, The structure and concentration
decay of high pressure jets of natural gas, Combustion Science and Technology,
vol. 36, pp. 249261, 1984.
[38] P. Hill and P. Ouellette, Transient turbulent gaseous fuel jets for diesel engines,
Transactions of the ASME: Journal of Fluids Engineering, vol. 121, no. 1, pp. 93
101, 1999.
[39] J. Abraham, V. Magi, J. MacInnes, and F. Bracco, Gas versus spray injection:
Which mixes faster?, SAE Technical Paper Series, vol. 940895, 1994.
[40] P. Hill and P. Ouellette, Turbulent transient gas injections, Transactions of the
ASME: Journal of Fluids Engineering, vol. 122, no. 4, pp. 743753, 2000.
[41] M. Jennings and F. Jeske, Analysis of the injection process in direct injected natural
gas engines: Part 1 - Study of unconfined and in-cylinder plume behavior, Journal
of Engineering for Gas Turbines and Power, vol. 116, no. 4, pp. 799805, 1994.
[42] J. Abraham and F. Bracco, Effects of combustion on in-cylinder mixing of gaseous
and liquid jets, SAE Technical Paper Series, vol. 950467, 1995.
[43] M. Jennings and F. Jeske, Analysis of the injection process in direct injected natural
gas engines: Part II - Effects of injector and combustion chamber design, Journal
of Engineering for Gas Turbines and Power, vol. 116, no. 4, pp. 806813, 1994.
[44] V. sy and H. Valland, The influence of natural gas composition on ignition in
a direct injection gas engine using hot surface assisted compression ignition, SAE
Technical Paper Series, vol. 961934, 1996.

References

136

[45] D. Siebers, Ignition delay characteristics of alternative diesel fuels: Implications on


cetane number, SAE Technical Paper Series, vol. 852102, 1985.
[46] R. Fraser, D. Siebers, and C. Edwards, Autoignition of methane and natural gas
in a simulated diesel environment, SAE Technical Paper Series, vol. 910227, 1991.
[47] A. Agarwal and D. Assanis, Multi-dimensional modeling of natural gas ignition under compression ignition conditions using detailed chemistry, SAE Technical Paper
Series, vol. 980136, 1998.
[48] J. Naber, D. Siebers, C. Westbrook, J. Caton, and S. Di Julio, Natural gas autoignition under disel conditions: Experiments and chemical kinetic modeling, SAE
Technical Paper Series, vol. 942034, 1994.
[49] J. T. Kubesh, S. R. King, and W. E. Liss, Effect of gas composition on octane
number of natural gas fuels, SAE Technical Paper Series, vol. 922359, 1992.
[50] I. Glassman and R. A. Yetter, Combustion. Oxford, UK: Academic Press Inc., 4 ed.,
2008.
[51] S. Dumitrescu, P. Hill, G. Li, and P. Ouellette, Effects of injection changes on
efficiency and emissions of a diesel engine fueled by direct injection of natural gas,
SAE Technical Paper Series, vol. 2000-01-1805, 2000.
[52] V. sy and H. Valland, Hot surface assisted compression ignition of natural gas
in a direct injection diesel engine, SAE Technical Paper Series, vol. 960767, 1996.
[53] C. Ho and E. Gutmark, Vortex induction and mass entrainment in a small-aspectratio elliptic jet, Journal of Fluid Mechanics, vol. 179, pp. 383405, 1987.
[54] F. Hussain and H. Husain, Elliptic Jets. Part 1. Characteristics of unexcited and
excited jets, Journal of Fluid Mechanics, vol. 208, pp. 257320, 1989.

References

137

[55] K. Schadow, E. Gutmark, S. Koshigoe, and K. Wilson, Combustion-related shearflow dynamics in elliptic supersonic jets, AIAA Journal, vol. 27, no. 10, pp. 1347
1353, 1989.

[56] L. Jacobsson and J. Chomiak, Injection orifice shape: Effects on combustion and
emission formation in diesel engines, SAE Technical Paper Series, vol. 972964, 1997.

[57] A. Matsson, L. Jacobsson, and S. Andersson, The effect of elliptical nozzle holes on
combustion and emission formation in a heavy duty diesel engine, SAE Technical
Paper Series, vol. 2000-01-1251, 2000.

[58] A. Matsson and S. Andersson, The effect of non-circular nozzle holes on combustion
and emission formation in a heavy-duty diesel engine, SAE Technical Paper Series,
vol. 2002-01-2671, 2002.

[59] H. B. Palmer and C. F. Cullis, The formation of carbon from gases, Chemistry
and Physics of Carbon, vol. 1, pp. 265325, 1965.

[60] R. L. Vander Wal, Soot nanostructure: Definition, quantification and implication,


SAE Technical Paper Series, vol. 2005-01-0964, 2005.

[61] S. Hong, D. Assanis, and M. Wooldridge, Multi-dimensional modeling of NO and


soot emissions with detailed chemistry and mixing in a direct injection natural gas
engine, SAE Technical Paper Series, vol. 2002-01-1112, 2002.

[62] G. McTaggart-Cowan, W. Bushe, S. Rogak, P. Hill, and S. Munchi, PM and N Ox


reduction by injection parameter alterations in a direct injected, pilot ignited, heavy
duty natural gas engine with EGR at various operating conditions, SAE Technical
Paper Series, vol. 2005-01-1733, 2005.

References

138

[63] H. Jones, G. McTaggart-Cowan, S. Rogak, W. Bushe, S. Munchi, and B. Buchholz,


Source apportionment of particulate matter from a diesel pilot-ignited natural gas
fuelled heavy duty DI engine, SAE Technical Paper Series, vol. 2005-01-2149, 2005.
[64] M. Fabbroni, X. Cheng, V. Abate, and J. S. Wallace, An optically-accessible combustion apparatus for direct-injection natural gas ignition studies, Proceedings of
the 2007 Fall Technical Conference of the ASME Internal Combustion Engine Division, vol. ICEF2007-1763, 2007.
[65] C. Green and J. Wallace, Electrically actuated injectors for gaseous fuels, SAE
Technical Paper Series, vol. 892143, 1989.
[66] Cambustion Ltd., 1, Portugal Place, Cambridge, England, HFR400 Fast FID Quick
User Guide, 1.1 ed.
[67] R. Dennis, W. R. Samples, D. M. Anderson, and L. Silverman, Isokinetic sampling
probes, Industrial and Engineering Chemistry, vol. 49, 1957.
[68] J. S. Dipchand, An investigation into the effects of CNG-gasoline operation on EGO
sensor behavior and deterioration, Masters thesis, Department of Mechanical and
Industrial Engineering, University of Toronto, 2001.
[69] L. Prandtl and O. Tietjens, Applied hydro- and Aeromechanics. Dover Publications,
Inc., 1957.
[70] National Instruments, AT-MIO E Series User Manual. March 1995 Edition.
[71] National Instruments, SC-207X Series User Manual. November 1995 Edition.
[72] Princeton Instruments, Inc., ITE/CCD Detector Operation Manual, version 1, revision c ed., January 1998.
[73] Princeton Instruments, Inc., ST-138 Controller Operation Manual, version 2, revision a ed., September 1994.

References

139

[74] Princeton Instruments, Inc., PG-200 Programmable Pulse Generator Operation


Manual, version 1, revision e ed., June 1994.
[75] Princeton Instruments, Inc., WinView Users Manual, version 1.4, revision c ed.,
February 1996.
[76] J. A. Wyman, An investigation of the mixing of fuel and air in the prechamber
of an indirect injection compression ignition engine using a water analog model,
Masters thesis, Department of Mechanical and Industrial Engineering, University
of Toronto, 1994.
[77] Y. A. C
engel and M. A. Boles, Thermodynamics: An Engineering Approach.
McGraw-Hill Inc., 5 ed., 2005.
[78] M. Lapuerta, O. Armas, and V. Berm
udez, Sensitivity of diesel engine thermodynamic cycle calculation to measurement errors and estimated parameters, Applied
Thermal Engineering, vol. 20, no. 9, pp. 843861, 2000.

Appendix A
Equipment Calibration Curves

A.1

AST Pressure Transducer

The AST4700 pressure transducer was a new unit bought to replace the Schaevitz transducer used in the past. The unit has a range of 02500 psig and an accuracy of 0.1% of
full scale with an output of 0 5 V . The calibration was performed by the manufacturer
and the curve supplied is shown in Figure A.1.

A.2

Kistler Pressure Transducer

The Kistler piezoelectric pressure transducer operates by measuring a change in pressure.


It was calibrated through the AST pressure transducer by connecting both to a chamber
that is pressurized. The chamber is then charged to a pressure and the reading on the
Kistler is allowed to stabilize. The pressure is then suddenly dropped to ambient and
the resulting voltage from the Kistler is recorded. This process is repeated for multiple
points to generate the calibration curve shown in Figure A.2.
140

141

Appendix A. Equipment Calibration Curves

3000

2500
y = 500.28x

Pressure [kPa]

2000

1500

1000

500

0
0.00

1.00

2.00

3.00
AST Transducer Output [V]

4.00

5.00

6.00

Figure A.1: AST4700 pressure transducer calibration curve

9000

8000

Change in Pressure [kPa]

7000

y = 1096.7x
R = 1

6000

5000

4000

3000

2000

1000

0
0

3
4
Change in Voltage [V]

Figure A.2: Kistler piezoelectric pressure transducer calibration curve

142

Appendix A. Equipment Calibration Curves


350

300
y = 61.864x
R = 0.9994

Absolute Pressure [kPa]

250

200

150

100

50

0
0.00

1.00

2.00

3.00
GM MAP Output [V]

4.00

5.00

6.00

Figure A.3: GM MAP sensor calibration curve

A.3

GM Manifold Absolute Pressure Sensor

The GM MAP snesor is used to measure the pressure at the intake manifold. It was
calibrated using the AST pressure transducer by connecting both to a chamber that is
pressurized, similar to the calibration of the Kistler transducer. Multiple points from 1
to 3 bar were measured and plotted in Figure A.3

A.4

TSI 2013 Laminar Flowmeter

The TSI 2013 laminar flowmeter was calibrated using a Singer gas meter prover to generate the curve and confirmed the curve supplied by the manufacturer. The flow rates
were well characterized by the logarithmic curve shown in Figure A.4 and agreed well
with that of past works. The resulting equation is used to characterize the flow rates of
the CFR engine.

143

Appendix A. Equipment Calibration Curves

4.5

4.0

3.5

Output Voltage [V]

3.0

y = 2.3449x0.1541
R = 0.9969

2.5

2.0

1.5

1.0

0.5

0.0
0

10
15
Volumetric Flow Rate [SCFM]

20

25

Figure A.4: TSI 2013 flowmeter calibration curve

A.5

Matheson Rotameters

The two Matheson rotameters used to prepare the calibration mixture for the FFID are
characterized through a data set supplied by the manufacturer. The data set for both
tubes are supplied by the manufacturer. Tube 6 is used for Air and the calibration was
performed with the stainless steel float, presented in Figure A.5. Tube 602 is used for
Methane with a calibration using the glass float, presented in Figure A.6. The gases
supplied to the tubes were both set at 10 psig. Having equal tube pressure eliminates
the need for a pressure correction term using the calibration curves.

Appendix A. Equipment Calibration Curves

144

Figure A.5: Matheson rotameter calibration curve, Tube 6 with stainless steel float [28]

Figure A.6: Matheson rotameter calibration curve, Tube 602 with glass float [28]

Appendix B
Complete Set of Data Figures
The following appendix is a summary of all of the figures produced from the results.
Although some are repeated from Chapter 6 all of the figures are presented here in its
entirety for completeness. This allows for a more consistent organization of the figures
and easier indexing for the reader. The figures are presented in the order in which the
topic was presented in the discussion in Chapter 6: ignition delay, combustion images and
histograms, ROHR curve profiles, ROHR and net heat nozzle performance values, fuel
utilization and gross indicated thermal efficiencies, OC/EC test results, and gravimetric
particulate results.

145

146

Appendix B. Complete Set of Data Figures

2.0

1.8

Ignition Delay [ms]

1.6

3C
3E

1.4

1C
1E

9C
9E

1.2

1.0

0.8
1

11

16
21
Mass of Fuel Injected [mg]

26

31

36

Figure B.1: Ignition Delay vs Mass of Fuel Injected


2.0

1.8

Ignition Delay [ms]

1.6
3C

3E

1.4

1C
1E
9C
9E

1.2

1.0

0.8
0

5
6
Mass per Hole [mg]

10

Figure B.2: Ignition Delay vs Mass of Fuel Injected per Hole

147

Appendix B. Complete Set of Data Figures

2.0

1.8

Injection Delay [ms]

1.6
3C

3E

1.4

1C
1E
9C
9E

1.2

1.0

0.8
0

10

20

30
40
50
Mass per Area [mg/mm2]

60

70

80

Figure B.3: Ignition Delay vs Mass of Fuel Injected per Area

Appendix B. Complete Set of Data Figures

148

(a) Circular

(b) Elliptical

Figure B.4: Set of combustion images for 10 consecutive fired cycles for a 1 hole nozzle,
2 CAD duration (scaled to 5%-95% intensity range)

Appendix B. Complete Set of Data Figures

149

(a) Circular

(b) Elliptical

Figure B.5: Typical image histogram for a 1 hole nozzle, 2 CAD duration showing the
number of pixels above 50% intensity

Appendix B. Complete Set of Data Figures

150

(a) Circular

(b) Elliptical

Figure B.6: Set of combustion images for 10 consecutive fired cycles for a 3 hole nozzle,
3 CAD duration

Appendix B. Complete Set of Data Figures

151

(a) Circular

(b) Elliptical

Figure B.7: Typical image histogram for a 3 hole nozzle, 3 CAD duration showing the
number of pixels above 50% intensity

Appendix B. Complete Set of Data Figures

152

(a) Circular

(b) Elliptical

Figure B.8: Set of combustion images for 10 consecutive fired cycles for a 9 hole nozzle,
3 CAD duration

Appendix B. Complete Set of Data Figures

153

(a) Circular

(b) Elliptical

Figure B.9: Typical image histogram for a 9 hole nozzle, 3 CAD duration showing the
number of pixels above 50% intensity

Appendix B. Complete Set of Data Figures

154

(a) Circular

(b) Elliptical

Figure B.10: Typical ROHR and Net Heat curves for a 1 hole nozzle, 3 CAD duration

Appendix B. Complete Set of Data Figures

155

(a) Circular

(b) Elliptical

Figure B.11: Typical ROHR and Net Heat curves for a 3 hole nozzle, 3 CAD duration

Appendix B. Complete Set of Data Figures

156

(a) Circular

(b) Elliptical

Figure B.12: Typical ROHR and Net Heat curves for a 9 hole nozzle, 1.4 CAD duration

Appendix B. Complete Set of Data Figures

157

(a) Circular

(b) Elliptical

Figure B.13: Typical ROHR and Net Heat curves for a 9 hole nozzle, 3 CAD duration

158

Appendix B. Complete Set of Data Figures

700

600

Max ROHR [J/CAD]

500

400

3C

3E
1C
1E

300

9C
9E

200

100

0
1

11

16
21
Mass of Fuel Injected [mg]

26

31

36

Figure B.14: Max ROHR vs Mass of Fuel Injected


700

600

Max ROHR [J/CAD]

500

400

3C

3E
1C
1E

300

9C
9E

200

100

0
1

5
6
Mass per Hole [mg]

10

Figure B.15: Max ROHR vs Mass of Fuel Injected per Hole

159

Appendix B. Complete Set of Data Figures

700

600

Max ROHR [J/CAD]

500

400

3C
3E
1C
1E

300

9C
9E
200

100

0
1

11

21

31
41
51
Mass per Area [mg/mm2]

61

71

81

Figure B.16: Max ROHR vs Mass of Fuel Injected per Area


800

700

600

Net Heat [J]

500
3C

3E

400

1C
1E
9C

300

9E
200

100

0
1

11

16
21
Mass of Fuel Injected [mg]

26

31

Figure B.17: Net Heat vs Mass of Fuel Injected

36

160

Appendix B. Complete Set of Data Figures

800

700

600

Net Heat [J]

500
3C

3E

400

1C
1E
9C

300

9E
200

100

0
1

5
6
Mass per Hole [mg]

10

Figure B.18: Net Heat vs Mass of Fuel Injected per Hole


800

700

600

Net Heat [J]

500
3C

3E

400

1C
1E
9C

300

9E
200

100

0
1

11

21

31
41
51
Mass per Area [mg/mm2]

61

71

81

Figure B.19: Net Heat vs Mass of Fuel Injected per Area

161

Appendix B. Complete Set of Data Figures

1.00

0.98

0.96

Fuel Utilization

0.94
3C
3E

0.92

1C
1E
9C

0.90

9E
0.88

0.86

0.84
1

11

16
21
Mass of Fuel Injected [mg]

26

31

36

Figure B.20: Fuel Utilization vs Mass of Fuel Injected


1.00

0.98

Fuel Utilization

0.96
3C
3E

0.94

1C
1E
9C

9E

0.92

0.90

0.88
1

5
6
Mass per Hole [mg]

10

Figure B.21: Fuel Utilization vs Mass of Fuel Injected per Hole

162

Appendix B. Complete Set of Data Figures

1.00

0.98

0.96

Fuel Utilization

0.94
3C
3E

0.92

1C
1E
9C

0.90

9E
0.88

0.86

0.84
1

11

21

31
41
51
Mass per Area [mg/mm2]

61

71

81

Figure B.22: Fuel Utilization vs Mass of Fuel Injected per Area


0.8

Gross Indicated Thermal Efficiency

0.7

0.6

3C
3E

0.5

1C
1E
9C
9E

0.4

0.3

0.2
1

11

16
21
Mass of Fuel Injected [mg]

26

31

36

Figure B.23: Indicated Thermal Efficiency vs Mass of Fuel Injected

163

Appendix B. Complete Set of Data Figures

0.8

Gross Indicated Thermal Efficiency

0.7

0.6

3C
3E

0.5

1C
1E
9C
9E

0.4

0.3

0.2
1

5
6
Mass per Hole [mg]

10

Figure B.24: Indicated Thermal Efficiency vs Mass of Fuel Injected per Hole
0.8

Gross Indicated Thermal Efficiency

0.7

0.6

3C
3E

0.5

1C
1E
9C
9E

0.4

0.3

0.2
1

11

21

31
41
51
Mass per Area [mg/mm2]

61

71

81

Figure B.25: Indicated Thermal Efficiency vs Mass of Fuel Injected per Area

164

Appendix B. Complete Set of Data Figures

100%

90%

Distribution of Organic/Elemental Carbon [%]

80%

70%

60%

50%

Elemental
Organic

40%

30%

20%

10%

0%
3C 2CAD

3C 3CAD

3E 2CAD
3E 3CAD
9C 2CAD
9C 3CAD
Nozzle Number and Injection Duration

9E 2CAD

9E 3CAD

Figure B.26: Distribution of Organic and Elemental Carbon


6

specific Elemental Carbon [mg/kWh]

Circular
Elliptical

0
5

10

15

20
Mass of Fuel Injected [mg]

25

30

35

Figure B.27: Elemental Carbon Emissions vs Mass of Fuel Injected

165

Appendix B. Complete Set of Data Figures

specific Elemental Carbon [mg/kWh]

Circular
Elliptical

4
Mass per Hole [mg]

Figure B.28: Elemental Carbon Emissions vs Mass of Fuel Injected per Hole
6

specific Elemental Carbon [mg/kWh]

Circular
Elliptical

20

25

30

35

40
45
Mass per Area [mg/mm2]

50

55

60

65

Figure B.29: Elemental Carbon Emissions vs Mass of Fuel Injected per Area

166

Appendix B. Complete Set of Data Figures

20

18

specific Organic Carbon [mg/kWh]

16

14

12

10

C Organic
E Organic

0
5

10

15

20
Mass of Fuel Injected [mg]

25

30

35

Figure B.30: Organic Carbon Emissions vs Mass of Fuel Injected


20

18

specific Organic Carbon [mg/kWh]

16

14

12

10

C Organic
E Organic

0
1

4
Mass per Hole [mg]

Figure B.31: Organic Carbon Emissions vs Mass of Fuel Injected per Hole

167

Appendix B. Complete Set of Data Figures

20

18

specific Organic Carbon [mg/kWh]

16

14

12

10

C Organic
E Organic

0
20

25

30

35

40
45
Mass per Area [mg/mm2]

50

55

60

65

Figure B.32: Organic Carbon Emissions vs Mass of Fuel Injected per Area
40

35
1 Hole Nozzles

specific PM [mg/kWh]

30

25
3C
3E

3 Hole Nozzles

20

1C
1E
9C

15

9E
10
9 Hole Nozzles
5

0
1

11

16
21
Mass of Fuel Injected [mg]

26

31

36

Figure B.33: Specific PM Emissions vs Mass of Fuel Injected

168

Appendix B. Complete Set of Data Figures

40

35

specific PM [mg/kWh]

30

25
3C
3E

20

1C
1E
9C

15

9E
10

5
6
Mass per Hole [mg]

10

Figure B.34: Specific PM Emissions vs Mass of Fuel Injected per Hole


40

35

specific PM [mg/kWh]

30

25
3C
3E

20

1C
1E
9C

15

9E
10

11

21

31
41
51
Mass per Area [mg/mm2]

61

71

81

Figure B.35: Specific PM Emissions vs Mass of Fuel Injected per Area

169

Appendix B. Complete Set of Data Figures

0.30

Mass Collected on Filter [mg]

0.25

0.20
3C
3E
1C
1E
0.15

9C
9E

0.10

0.05
1

11

16
21
Mass of Fuel Injected [mg]

26

31

36

Figure B.36: Mass Collected on Filter vs Mass of Fuel Injected


0.30

Mass Collected on Filter [mg]

0.25

0.20
3C
3E
1C
1E
0.15

9C
9E

0.10

0.05
0

5
6
Mass per Hole [mg]

10

Figure B.37: Mass Collected on Filter vs Mass of Fuel Injected per Hole

170

Appendix B. Complete Set of Data Figures

0.30

Mass Collected on Filter [mg]

0.25

0.20
3C
3E
1C
1E
0.15

9C
9E

0.10

0.05
0

10

20

30
40
50
Mass per Area [mg/mm2]

60

70

80

Figure B.38: Mass Collected on Filter vs Mass of Fuel Injected per Area

Appendix C
MATLAB Script Description
This section will describe each of the MATLAB scripts used in this study to process the
data. Several steps are taken when processing the raw data to generate the final set
of numbers that characterize the performance of the CFR system. First, the motored
performance of the CFR system must be characterized. The resulting information is then
output for use in the processing of the final fired tests. The raw data is recorded into a
text file through the National Instruments LabView software. This text file is processed
through MATLAB into a set of cycle-to-cycle data which is then further processed into
a final set average used to evaluate a nozzles overall performance. The images acquired
through the Princeton Instrument WinView software are processed directly through a
separate MATLAB script. A detailed explanation of each script will be presented here.

MotorAvgV2.m This function reads a set of motored data files and returns an average
set of data for a complete cycle. It requires a file generated from LabView using the
MotoredDAQ3 C.vi and outputs a text file of the generated mass fraction profile vs
CAD use to process the fired data. The script has options to generate cycle-to-cycle
and total averaged statistics and plots of in-cylinder temperature, pressure, mass
fraction, air flow rate, and intake pressure. It calls the LPFil.m script to filter the
data.
171

Appendix C. MATLAB Script Description

172

LPFil.m This script takes an input signal X and filtering window N (where N must
be odd), and produces a smoothed output signal by applying a running average.
Larger N values will produce greater smoothing at the expense of resolution of the
output.
DAQAnalysisV8.m This function generates the main results for a single set of firing
test data. Revisions include new calibration constants, increased robustness, and
the option for batch file processing. Each file generated by LabView during a fired
set of tests is placed into an array. In addition to the raw data, the fired data file
contains the nozzle number, number of acquired and fired cycles, and operational
parameters (i.e. coolant temperature, oil temperature, etc). The file requires the
mass fraction profile generated by the MotorAvgV2.m script. It also calls on the
Ignition.m and derivative.m scripts. The text file produced contains the following
fired cycle-to-cycle data: operational parameters, ignition delay, instantaneous and
cumulative heat release, and combustion efficiencies.
Ignition.m This script is called upon during the DAQAnalysisV8.m data processing.
It requires the cycle crank angle, fired bomb pressure, motored bomb pressure,
injector current trace, engine speed, and injector delay for a given fired cycle. It
returns the engines position at the start of the injector pulse, at the start of the
pressure rise indicating ignition, the ignition delay time, and the actual injection
duration. It also has the option of plotting injector and pressure traces at the time
of injection and ignition.
derivative.m This script generates the first or second derivative of a set of experimentally collected data points. The Savitzky-Golay filter is used by which it fits a
polynomial to a portion of the data then calculates the derivative of the midpoint
of that portion. The script requires the signal to be process, the length of the window of data (must be odd), the order of the polynomial, the order of the derivative

Appendix C. MATLAB Script Description

173

desired, and the vector by which to reference the derivative. The output is the
derivative of the signal with respect to the reference specified. This script requires
the Signal Processing Toolbox.
DAQAnalysisPostProcessingV3.m This function reads the file generated from the
DAQAnalysisV8.m script and generates cycle-to-cycle statistics on the final data
as well as values for the total set averages. It generates an excel file with the set
averaged values of each file in a column for the final analysis. User has the option
to plot cycle-to-cycle or set averaged ROHR and net heat release curves, ignition
delay, and burn duration with respect to crank angle. A file array can be used for
batch file processing.
SPEImageProcessingV2.m This script reads an SPE image file generated from the
Princeton Instrument WinView software during a fired test and generates images
in PNG format after cropping out the black spaces and compiling the strips into
sheets. It reads the SPE image through the readSPE.m script from a third party
developer into an image array. From the image array, the histograms are generated
for pixels above a user specified value and composite images comprising of 10 fired
cycles are generated and saved to a sequentially numbered set of PNG images with
crank angle position displayed. User has the option of scaling the image to 5 95%
intensity range and to specify a file array for batch processing. This script requires
the Image Processing Toolbox.

Appendix D
Drawings
This section presents the drawings of newly designed components for this study. Many
of the parts used were the same as that of past works and the latest drawings for those
parts can be found in Fabbroni [26]. The drawings presented here are that of new parts
that have been implemented during the course of this study. These include two new
injector nozzles, two revised designs for the Vespel injector seats, and a glow plug shield
simulator for temperature tests. The 9E nozzle seat prototype in Figure D.9 was first
used as the 9E seat. However, in light of the potential interference of the channels with
the flow of gas, this was superseded by the the simplified seat design in Figure D.7. The
benefits of reduced interference can be better seen in Figure D.8, which shows the seat
nested in the tip of the injector nozzle.

174

0.040

0.330

Figure D.1: Injector nozzle with 3 hole elliptical geometry, Page 1


5

0.400

304 Stainless Steel


Charles Habbaky
Jan 09, 2012

MATERIAL

DRAWN BY
DATE

DIMENSIONS ARE IN INCHES


TOLERANCES:
DECIMALS
ANGULAR
.X
0.01
0.5
.XX
0.05
.XXX
0.010
.XXXX 0.0005

UNLESS OTHERWISE SPECIFIED:

0.125

2.288
2.188
1.180
1.110

1.275
1.375

0.125

SECTION A-A

40

0.248

0.748

0.166

1.000

1.000

0.185

0.115

DRILL 0.115
C.B. 0.185 X 0.125 DEEP
8 PLACES ON 1.138 PCD
FOR A 4-40 CAP SCREW

TITLE:

SCALE: 1:1
1

REV

SHEET 1 OF 3

SIZE DWG. NO.


Injector Nozzle Ellipse3

3-Hole Elliptical
Nozzle Body

University of Toronto - Engines


Research and Development Laboratory

DETAIL B
SCALE 5 : 1

Appendix D. Drawings
175

Figure D.2: Injector nozzle with 3 hole elliptical geometry, Page 2


5

0.400

0.166
Drill and Ream

0.125
Drill and Ream for Press Fit

45

0.569
0.402

0.569
0.402

0.569

0.569

304 Stainless Steel


Charles Habbaky
Jan 09, 2012

MATERIAL

DRAWN BY
DATE

DIMENSIONS ARE IN INCHES


TOLERANCES:
DECIMALS
ANGULAR
.X
0.01
0.5
.XX
0.05
.XXX
0.010
.XXXX 0.0005

UNLESS OTHERWISE SPECIFIED:

1.138 PCD for


4-40 cap screw

1.375

1.275

0.185 X 0.125

0.115 THRU

0.402

0.402

3-Hole Elliptical Base


Hole Detail

SCALE: 1:1
1

REV

SHEET 2 OF 3

SIZE DWG. NO.


Injector Nozzle Ellipse3

TITLE:

University of Toronto - Engines


Research and Development Laboratory

Appendix D. Drawings
176

0.100

SECTION D-D
SCALE 5 : 1 0.042

40

DETAIL B
SCALE 10 : 1

0.166
Drill and Ream

0.125 Drill and Ream


for press fit Vespel seat

Figure D.3: Injector nozzle with 3 hole elliptical geometry, Page 3


304 Stainless Steel
Charles Habbaky
Jan 09, 2012

MATERIAL

DRAWN BY
DATE

DIMENSIONS ARE IN INCHES


TOLERANCES:
DECIMALS
ANGULAR
.X
0.01
0.5
.XX
0.05
.XXX
0.010
.XXXX 0.0005

UNLESS OTHERWISE SPECIFIED:

0.0090

0.0955

0.0090

0.0090

3-Hole Elliptical
Nozzle Hole Detail

SCALE: 1:1

REV

SHEET 3 OF 3

SIZE DWG. NO.


Injector Nozzle Ellipse3

TITLE:

University of Toronto - Engines


Research and Development Laboratory

DETAIL C
SCALE 20 : 1

BLIND HOLE
USING EDM

Appendix D. Drawings
177

0.040

0.330

Figure D.4: Injector nozzle with 9 hole elliptical geometry, Page 1


5

0.400

0.166

1.138

1.000

1.000

0.185

0.115

DRILL 0.115
C.B. 0.185 X 0.125 DEEP
8 PLACES ON 1.138 PCD
FOR A 4-40 CAP SCREW

SECTION A-A

1.375

1.275

0.125

0.125

304 Stainless Steel


Charles Habbaky
Jan 09, 2012

MATERIAL

DATE

DIMENSIONS ARE IN INCHES


TOLERANCES:
DECIMALS
ANGULAR
.X
0.01
0.5
.XX
0.05
.XXX
0.010
.XXXX 0.0005

UNLESS OTHERWISE SPECIFIED:

DRAWN BY

0.125

2.288
2.188
1.180
1.110

40

0.248

0.748

TITLE:

SCALE: 1:1
1

REV

SHEET 1 OF 3

SIZE DWG. NO.


Injector Nozzle Ellipse9

9-Hole Elliptical
Nozzle Body

University of Toronto - Engines


Research and Development Laboratory

DETAIL B
SCALE 5 : 1

Appendix D. Drawings
178

Figure D.5: Injector nozzle with 9 hole elliptical geometry, Page 2


5

0.400

0.166
Drill and Ream

0.125
Drill and Ream for Press Fit

45

0.569
0.402

0.115 THRU

1.375

1.275

304 Stainless Steel


Charles Habbaky
Jan 09, 2012

MATERIAL

DRAWN BY
DATE

DIMENSIONS ARE IN INCHES


TOLERANCES:
DECIMALS
ANGULAR
.X
0.01
0.5
.XX
0.05
.XXX
0.010
.XXXX 0.0005

UNLESS OTHERWISE SPECIFIED:

9-Hole Elliptical Base


Hole Detail

SCALE: 1:1

REV

SHEET 2 OF 3

SIZE DWG. NO.


Injector Nozzle Ellipse9

TITLE:

0.569

0.569

University of Toronto - Engines


Research and Development Laboratory

0.402

0.402

1.138 PCD for


4-40 cap screw

0.185 X 0.125

0.569
0.402

Appendix D. Drawings
179

SECTION D-D
SCALE 5 : 1

40

0.100

0.042

DETAIL B
SCALE 10 : 1

0.166
Drill and Ream

0.125 Drill and Ream


for press fit Vespel seat

Figure D.6: Injector nozzle with 9 hole elliptical geometry, Page 3


304 Stainless Steel
Charles Habbaky
Jan 09, 2012

MATERIAL

DRAWN BY
DATE

DIMENSIONS ARE IN INCHES


TOLERANCES:
DECIMALS
ANGULAR
.X
0.01
0.5
.XX
0.05
.XXX
0.010
.XXXX 0.0005

UNLESS OTHERWISE SPECIFIED:

0.0090

0.0955

0.0090

0.0090

9-Hole Elliptical
Nozzle Hole Detail

SCALE: 1:1

REV

SHEET 3 OF 3

SIZE DWG. NO.


Injector Nozzle Ellipse9

TITLE:

University of Toronto - Engines


Research and Development Laboratory

DETAIL C
SCALE 20 : 1

BLIND HOLE
USING EDM

Appendix D. Drawings
180

0.025

0.100

Figure D.7: Simplified nozzle seat, Page 1


5

0.166

0.125 for Press


Fit in Nozzle

0.041

0.125

118

45

45
0.005

0.005

Charles Habbaky
Apr 27, 2012
DATE

Polymide
(Dupont Vespel)
DRAWN BY

MATERIAL

DIMENSIONS ARE IN INCHES


TOLERANCES:
DECIMALS
ANGULAR
.X
0.01
0.5
.XX
0.05
.XXX
0.010
.XXXX 0.0005

UNLESS OTHERWISE SPECIFIED:

Simplified Nozzle
Pintle Seat

SCALE: 10:1
1

REV

SHEET 1 OF 2

SIZE DWG. NO.


Simplified Nozzle Seat

TITLE:

University of Toronto - Engines


Research and Development Laboratory

Appendix D. Drawings
181

Figure D.8: Simplified nozzle seat, Page 2


4

DETAIL D
With Seat Insert

Charles Habbaky
Apr 27, 2012

DRAWN BY
DATE

MATERIAL

DIMENSIONS ARE IN INCHES


TOLERANCES:
DECIMALS
ANGULAR
.X
0.01
0.5
.XX
0.05
.XXX
0.010
.XXXX 0.0005

UNLESS OTHERWISE SPECIFIED:

Simplified Pintle Seat


in Nozzle Body

SCALE: 10:1

REV

SHEET 2 OF 2

SIZE DWG. NO.


Simplified Nozzle Seat

TITLE:

University of Toronto - Engines


Research and Development Laboratory

SECTION C-C
SCALE 1 : 1
of Nozzle Body

Appendix D. Drawings
182

Figure D.9: 9E nozzle seat prototype


5

0.025 (Tip
Depth 0.0376)

0.005

0.010
0.005

0.010

118

0.166

0.125

0.041

0.095

0.125

0.037
0.077
0.100

0.200

Charles Habbaky
Apr 26, 2012
DATE

Polymide
(Dupont: Vespel)
DRAWN BY

MATERIAL

DIMENSIONS ARE IN INCHES


TOLERANCES:
DECIMALS
ANGULAR
.X
0.01
0.5
.XX
0.05
.XXX
0.010
.XXXX 0.0005

UNLESS OTHERWISE SPECIFIED:

0.057

45

0.020

SCALE: 10:1

SIZE DWG. NO.


9ENozzleSeat

SHEET 1 OF 1

REV

9E Nozzle Seat
(Superceded by
"Simplified Nozzle Seat")

University of Toronto - Engines


Research and Development Laboratory
TITLE:

SECTION A-A

0.041

Appendix D. Drawings
183

Figure D.10: Glow plug shield simulator


5

0.23622 (6mm)

2.000 (3/8"-16 bolt must be


threaded to at least 1 5/8")

0.27559 (7mm)

316 Stainless Steel


Charles Habbaky
Mar 29, 2012

MATERIAL

DRAWN BY
DATE

SHIELD DIMENSIONS TAKEN


FROM "COMBUSTION BOMB
SMALL GLOW PLUG SHIELD
DESIGN" BY FABBRONI

DIMENSIONS ARE IN INCHES


TOLERANCES ARE 0.01

UNLESS OTHERWISE SPECIFIED:

GP Shield Simulator

SCALE: 1:1
1

REV

SHEET 1 OF 1

SIZE DWG. NO.


GP Shield Simulator

TITLE:

University of Toronto - Engines


Research and Development Laboratory

1.050 0.050 (assumes Denso DG-143


Glow Plug with a 0.95" element length)

0.23819 (6.05mm)

Appendix D. Drawings
184

Appendix E
Thermocouple Assembly
Instructions
The following will outline the procedure used to assemble the 0.000500 thermocouple
probe. The assembly is very delicate and the instructions should be read in its entirety
before starting the assembly. Needless to say, the thermocouple wire is extremely sensitive
and breaks very easily.

E.1

Preliminary Setup

The following are the preliminary instructions that should be prepared prior to the assembly. They include any tools and parts that are needed.

1. Clean a workspace table and secure a white surface to work on to be able to see
the thermocouple wire easily. This can be accomplished by taping sheets of paper
to the table
2. Open the Themocouple Assembly Kit and keep it near you.
3. Place the following on the table:
185

Appendix E. Thermocouple Assembly Instructions

186

(a) 1x Thermocouple assembly block (Aluminum block stand with set screws to
support 1/800 stainless steel tubing)
(b) 1x 600 stainless steel tube (thermocouple shell)- 1/800 OD, 0.06900 ID
(c) 1x Omega ceramic 2-hole sleeve (TRM-164116-6 or TRA-164116-6)
(d) 1x Omega male K-type high temperature connector with contact washer
(HMPW-K-M)1
(e) 1x Omega K-type 0.000500 Thermocouple wire (CHAL-0005)
(f) 1x OmegaBond 300 High Temperature Cement (OB-300)
(g) 1x petri dish to mix cement
(h) 1x Syringe with water (for cement)
(i) 1x Instant Krazy Glue brand glue2
(j) 1x Loctite 406 Surface Insensitive glue
(k) 1x Hastings Lens 10x Loupe (Bausch&Lomb)
(l) 1x Loupe head strap (Bausch&Lomb)
(m) 1x Goo Gone sticker removal solution
(n) 1x cotton swab
(o) 1x 5/6400 Allen key
(p) 1x sharp knife
(q) 1x tweezers
(r) 1x wire cutter
(s) 1x small Philips screw driver
(t) 2x needles (any gauge)
1
2

Model with contact washer is critical to avoid breakng the wire during tightening.
Must get a glue that adheres very well to metal.

Appendix E. Thermocouple Assembly Instructions

187

(u) 1x red steel scribing ink


(v) 2x 800 0.00900 spring temper wire (also known as music wire)
(w) 1x electrical tape (or tape with a weak adhesive)
(x) 1x Bench mounted overhead lamp
(y) 1x K-type thermocouple reader
(z) 1x K-type extension wire (one end to connect to the reader and the other with
stripped wire leads exposed)

E.2

Assembly Instructions

With everything prepared from the previous section. The assembly procedure will now
be described. Several different methods have been tested but the following proved to
yield the best results.
1. Disassemble the K-type connector (d) until only base unit remains: remove cover,
remove set screw and contact washer, and remove red rubber tube ring.
2. Place red rubber tube ring from connector on SS tube (b) such that it would later
insert nicely in the connector.
3. Insert the ceramic sleeve (c) in the stainless steel tube. It should fit nicely such
that both ends are somewhat flush with the SS tube.
4. Push the ceramic to expose 1/200 of the ceramic on the side with the red rubber
ring.
5. Place a small amount of Loctite 406 (j) on the exposed ceramic, and push back
into the SS tube gently with a needle or tweezers. This ensures that the ceramic
does not drift during assembly. Allow a few seconds to dry.

Appendix E. Thermocouple Assembly Instructions

188

6. Insert the SS tube and ceramic assembly (a.k.a. T/C probe) into the Thermocouple
assembly block and tighten set screw to keep in place. Ensure the two holes are
aligned horizontally to match the T/C wire placement.
7. Cut two pieces of 0.00900 spring temper music wire (v) roughly 800 in length. This
will be used to pull the T/C wire through.
8. Thread both music wires through the ceramic such that both ends are exposed.
This should be done with ease since the spring temper wire is far more rigid.
9. On the ends of the music wire that stick out of the tip of the T/C probe (the end
opposite of the red rubber ring), use the red scribing ink (u) to mark the wire 1/200
from the end. This will identify when the wire is almost completely through when
threading.
10. Place the set of 0.000500 T/C wire (e) still attached to the white plastic support
such that the terminal ends (taped by yellow/red stickers) face the tip of the probe
that is in the assembly block. Turn on the overhead lamp (x) to better see the
wire.
11. Open the Instant Krazy Glue (i) and keep close to the work.
12. Using a sharp knife (p), cut one terminal of the T/C wire as close to the sticker
as possible while placing a finger on the wire to ensure it does not spring back.
Release the wire but keep track of where it is.
13. Grab the end of the corresponding music wire (while keeping it threaded in the
ceramic) and dip in the Krazy glue bottle. Once the wire is removed, beads of glue
should be visible.
14. Gently grab the 0.000500 wire of the cut terminal leaving the end exposed and wrap
around the glue covered music wire until it sticks.

Appendix E. Thermocouple Assembly Instructions

189

15. Gently run fingers over the glued wires to wipe off excess glue. No beads should be
visible when dry to be able to fit through ceramic holes.
16. Hold for a few seconds until dry
NOTE1: Do not release the music wire when it is in tension. It will whip back and
break wire. GENTLY return the wire to an un-tense position before releasing it.
NOTE2: Once the glue is applied to the wire, Steps 13-15 should be performed
quickly while it is still wet.
17. Repeat steps 12-16 for the other terminal.
18. VERY Gently and Slowly pull the exposed end of the music wire on the connector
side of the T/C probe with tweezers. If the music wire is not exposed on the
connector side because of the earlier glue application, gently push the wires until
exposed.
NOTE: Feeding the wires is a VERY VERY DELICATE process.
With tweezers alternate between pulling both wires such that neither are highly
tensioned. As you pull the wires, push the white plastic T/C wire that still supports
the junction that remains stuck with the white sticker. A slot underneath the
Thermocouple assembly block allows enough clearance for the plastic support to
slide underneath. Pull the music wire through until the red ink appears.
19. Use wire cutters (r) to trim the length of the exposed music wire to a more manageable size without disturbing the assembly.
20. VERY CAREFULLY pull the remainder of the wire without allowing the music
wire to spring out once it is through. This will break the wires.
21. Once through, use electrical tape (w) to hold the music wire (still attached to the
T/C wire) to the table.

Appendix E. Thermocouple Assembly Instructions

190

22. Slide the assembly block forward toward the junction that is still held by the white
sticker as close as possible without tensioning the T/C wire.
23. Dip a cotton swab (n) in the Goo Gone (m) sticker removal solution and dab gently
of the white sticker that holds the junction. Wait a minute for liquid to set in.
24. With a sharp knife, VERY CAREFULLY peel the white sticker off. It should come
off easily and should not stick to the T/C wire. If it still sticks to the T/C wire
then not enough time has past for the liquid to penetrate. Try to secure the wire
while peeling the wire if this occurs.
25. Use one needle (t) and pin it between the wires near the junction. Pull the wire
through such that the junction is as close to the probe end with the needle spacing
between them. This can be done with a combination of pushing the assembly block
and pulling one of the music wires.
NOTE: DO NOT lose the junction! Use the Hastings loupe (k) to locate if necessary.
26. In a petri dish (g), add the Omegabond 300 cement (f) and use a syringe filled with
water (h). Mix as per cement instructions. Note that not much cement is needed.
27. Remove the needle that is between the junction and the probe tip. Using the second
needle, grab some cement from the petri dish and drop cement very gently over the
T/C wire to fill the gap between the junction and the probe, sealing the ceramic.
Allow a few minute for cement to harden enough to hold the junction in place (full
cure requires 24 hours in air).
NOTE: DO NOT cover the junction since this will tamper the temperature readings.
Use the Hastings loupe (k) and head strap (l) to better see the junction and apply
the cement.
NOTE2: Remove any excess cement from the SS tube such that its OD does not
exceed 1/800 so that it can be removed from the assembly block and connected via

Appendix E. Thermocouple Assembly Instructions

191

Swagelok fitting.
28. Following this, loosen the set screw and simultaneously turn the T/C probe with
one of the terminals to switch their orientation, holding it by the music wire. The
reason for this is to align the correct terminal with that on the connector, which is
packaged in the reverse manner.
29. Slide the disassembled connector underneath the probe such that the red rubber
ring rests properly in its place.
30. Starting with one terminal, loosely place the wire above the terminal plate. Use
the Hastings lens to see the positioning of the wire and make sure the wire is not
over the threads where the screw will be placed.
31. With tweezers carefully place the contact washer over the T/C wire.
32. Still using tweezers, place the screw gently in its place disturbing as little as possible.
33. Tighten using the screw driver, but be extremely cautious in not disturbing the
assembly. Hold the connector in place as you tighten the screw firmly without
moving the assembly.
34. Using a portable thermocouple reader (y) and extension wire (z), touch the stripped
wire leads to their respective terminals on the connector and the temperature should
match the ambient. If so, the connection is complete. If not the connection is broken
and the following should be done:
(a) Use the Hastings lens to examine the terminals.
(b) If a wire is broken at the screw terminal, you can attempt to loosen the the
contact washer and slide it back under with the hasting lens and needle. This
is very difficult but possible. Check the connection.

Appendix E. Thermocouple Assembly Instructions

192

(c) If the wires seem intact, try further tightening the screw to ensure that the
washer is pushing the wire to the terminal and recheck the connection (0.000500
wires can slip through very tight clearances).
(d) If the connection is still broken, then the wire likely broke inside the tube at
which point it should be easily able to slide out (using the Hastings lens and
needle to pull it). This would require you to restart completely. The only
thing that would be salvaged is the K-type connector and its parts. The T/C
wire, SS tube, and ceramic would all be wasted.
35. Following a successful connection, CAREFULLY screw the cover of the K-type
connector tightly and add cement to the crevices between the probe and the connector. This will secure the probe and connector in place with respect to each other
and further restrict movement between the two.
36. Allow the assembled thermocouple probe to dry for 24 hours as per the instructions
of the cement.
NOTE: Even with cement reinforcing the probe, the assembly is still very fragile
and should be handled with care.

Вам также может понравиться