Sei sulla pagina 1di 262

The Pennsylvania State University

The Graduate School

John and Willie Leone Family Department of Energy and Mineral Engineering

INVESTIGATION OF IMPACT OF FUEL INJECTION STRATEGY AND

BIODIESEL FUELING ON ENGINE EMISSIONS AND PERFORMANCE

A Dissertation in

Energy and Mineral Engineering

by

Peng Ye

© 2011 Peng Ye

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

December 2011
The dissertation of Peng Ye was reviewed and approved* by the following

André L. Boehman
Professor of Fuel Science and Materials Science and Engineering
Dissertation Advisor
Chair of Committee

Randy L. Vander Wal


Associate Professor of Energy and Mineral Engineering and Materials Science
and Engineering

Daniel C. Haworth
Professor in Mechanical and Nuclear Engineering

Yaw D. Yeboah
Professor and Department Head of Energy and Mineral Engineering

R. Larry Grayson
Professor of Energy and Mineral Engineering
George H., Jr., and Anne B. Deike Chair in Mining Engineering
Graduate Program Officer of Energy and Mineral Engineering;

*Signatures are on file in the Graduate School

ii
Abstract

Both biodiesel fueling and changes of fuel injection pressure have significant

impacts on diesel engine emissions. The investigations of their impacts on engine exhaust

NOx and particulate matter emissions were conducted with an 8-cylinder common-rail

turbocharged direct injection diesel engine using ultra low sulfur diesel fuel and soybean

methyl ester (SME) – based biodiesel blends. The engine was running at moderate speed

and different loads. Three fuel injection parameters: start of injection, fuel injection

pressure and fuel injection duration were investigated to investigate their impact on

engine emissions. With the control of fuel injection strategy, it is shown in this work that

the biodiesel engine NOx emission penalty can be eliminated.

A fuel spray, mixture stoichiometry field and lift-off length model was employed

to explain the variations of NOx emission from biodiesel fueling and change of fuel

injection strategy. Linear correlations between the average oxygen equivalence ratio of

the fuel-air mixture at the autoignition zone near the lift-off length and brake specific

NOx emissions were observed for all load conditions, regardless of fuel type. This

confirms that the dominant factor that determines NOx emissions is the ignition event

controlled by the oxygen equivalence ratio at the autoignition zone.

The impact of late in-cylinder (post) injection combustion with biodiesel on

lubricating oil dilution was investigated in this work. It is shown that this injection

strategy could effectively decrease engine NOx emissions, while increase the CO and

unburned hydrocarbon emissions. The lubricating oil dilution depends on the post

injection timing: an increase in the lubricating oil dilution can be only observed if the

post injection timing is later than 45º after top dead center.

iii
The impacts of fuel injection pressure on diesel and biodiesel soot morphology

and oxidative reactivity were investigated. It is shown that compared with engine

condition and fuel injection pressure, biodiesel has much less significant impact on soot

morphology. For soot oxidative reactivity, it is found that both diesel and biodiesel soot

from higher fuel injection pressure have higher reactivity, and biodiesel soot has higher

reactivity than diesel soot when both of them are obtained from the same injection

pressure.

The optimized apparent heat release pattern for improved engine thermal

efficiency was investigated with a zero-dimensional engine thermodynamic simulation.

The results suggest that the optimized apparent heat release is a “wide and low” peak.

The reason for this kind of heat release is that it can decrease the in-cylinder temperature

and consequently decrease heat loss.

iv
Table of Contents

List of Figures ................................................................................................................... vii

List of Tables. .................................................................................................................. xiv

Nomenclature .................................................................................................................... xv

Acknowledgements ........................................................................................................ xviii

Chapter I. Introduction ................................................................................................... 1

Chapter II. Literature Review.......................................................................................... 7

2.1 Biodiesel background ............................................................................................ 7


2.2 Impact of biodiesel fueling on engine emissions ................................................ 13
2.2.1 Nitric oxides (NOx) ....................................................................................... 15
2.2.2 Particulate Matter (PM) ................................................................................. 19
2.3 Origin of biodiesel emission effect ..................................................................... 22
2.3.1 Biodiesel NOx effect...................................................................................... 22
2.3.2 Biodiesel particulate matter effect ................................................................. 27
2.4 Fuel Injection strategy with biodiesel fueling ..................................................... 29
2.5 Post injection strategy ......................................................................................... 33
2.6 Biodiesel and fuel injection pressure on PM morphology and soot oxidative
reactivity .............................................................................................................. 36
2.6.1 PM morphology ............................................................................................. 36
2.6.2 Soot oxidative reactivity ................................................................................ 40
2.7 Impact of biodiesel fueling, injection strategy and heat release pattern on Engine
efficiency ............................................................................................................. 42
Chapter III. Technical Approach .................................................................................... 49

3.1 Experiment apparatus.......................................................................................... 49


3.2 Engine emission bench: ...................................................................................... 50
3.3 Test fuel .............................................................................................................. 51
3.4 Engine test conditions ......................................................................................... 53
3.5 Biodiesel lubricating oil dilution ........................................................................ 57
3.6 Soot characterization ........................................................................................... 60
Chapter IV. Investigation of injection strategy and biodiesel fueling on diesel engine
emissions and performance ........................................................................ 64

v
4.1 Impact of injection timing and fuel injection pressure on NOx emissions ......... 64
4.2 Impact of injection timing and fuel injection pressure on particulate matter (PM)
emissions ............................................................................................................. 68
4.3 NOx-PM trade-off ............................................................................................... 72
4.4 Brake thermal efficiency ..................................................................................... 78
4.5 Apparent heat release analysis ............................................................................ 83
Chapter V. Origin of biodiesel NOx effect ................................................................... 92

Chapter VI. Investigation of late in-cylinder injection strategy on engine emissions and
lubricant fuel dilution ............................................................................... 108

6.1 Exhaust Emissions ............................................................................................ 108


6.2 Exhaust Temperature and BSFC ....................................................................... 110
6.3 Heat Release Analysis....................................................................................... 112
6.4 Oil Dilution Level, Total Acid Number (TAN) and mfMEP ........................... 120
6.5 Analysis of Hydrocarbons from Lubricant and Exhaust................................... 123
Chapter VII. Impact of fuel injection pressure on soot morphology and oxidative
reactivity ................................................................................................... 129

7.1 Morphology analysis......................................................................................... 129


7.2 Analysis of soot oxidative reactivity................................................................. 139
Chapter VIII. Optimized heat release pattern to achieve high engine efficiency ............ 149

8.1 Model formulation ............................................................................................ 149


8.2 Computational procedure .................................................................................. 156
8.3 Optimized heat release pattern probing ............................................................ 163
Chapter IX. Conclusions and recommendations for future work ................................. 193

References…. .................................................................................................................. 199

Appendix A: Matlab code ............................................................................................... 229

Appendix B: Repeatability of thermogravimetric analysis ............................................. 236

Appendix C: Repeatability of Raman analysis .............................................................. 239

Appendix D: Repeatability of Morphology analysis ...................................................... 240

Appendix E: Repeatability of particle size distribution (SMPS) result .......................... 242

vi
List of Figures
Figure II-1: A diagram representing the transesterification process, replicated from [6].
The general scheme of the process is to modify a triglyceride (a fat or oil
which is a glyceride with three fatty acids, represented as R1, R2 andR3)
into a fatty acid methyl ester. ....................................................................... 8

Figure II-2: NOx, PM, CO and HC % changes as biodiesel blending percentage increase
as determined by the EPA through statistical regression of publically
available data on highway heavy-duty truck engines [18, 51]. Taken from
[44] ............................................................................................................. 14

Figure II-3: Diesel soot agglomerate composed of spherical primary particles, taken from
Ref. [168].................................................................................................... 37

Figure II-4: A sketch of limit-pressure cycle .................................................................... 46

Figure II-5: Indicated efficiency map with α and β .......................................................... 47

Figure III-1: FAME composition of B100 soybean oil methyl ester ................................ 52

Figure III-2: Molecular structures[217], names, abbreviations, chemical formulas, and


Chemical Abstracts Service (CAS) numbers of compounds comprising the
biodiesel used in this work. ........................................................................ 53

Figure III-3: Thermophoretic sampling apparatus ............................................................ 57

Figure III-4: FTIR spectrums for oils with different biodiesel dilution............................ 58

Figure III-5: Correlation between absorbance of Peak 1746 cm-1 from FTIR and dilution
level ............................................................................................................ 59

Figure III-6: Demonstration of curve fitting for Raman spectrum ................................... 63

Figure IV-1: NOx emissions at 25% load. Four injection timings (9˚, 5˚, 1˚ before top
dead center and 3˚ after top dead center) were tested for both ultra low
sulfur diesel (ULSD) and B40 blend. Three fuel injection pressures (52, 60
and 80 MPa) were tested at each injection timing...................................... 65

Figure IV-2: NOx emissions at 50% load. Four injection timings (9˚, 7˚, 5˚, 3˚ 1˚ before
top dead center and 1˚, 3˚ after top dead center) were tested for both ultra
low sulfur diesel (ULSD) and B40 blend. Three fuel injection pressures (72,
90 and 108 MPa) were tested at each injection timing for ULSD, and six
fuel injection pressures (72, 80, 90, 98, 108 and 118 MPa) were tested at
each injection timing for B40. .................................................................... 66

vii
Figure IV-3: NOx emissions at 75% load. Four injection timings (9˚, 5˚, 1˚ before top
dead center and 3˚ after top dead center) were tested for both ultra low
sulfur diesel (ULSD) and B40 blend. Three fuel injection pressures (87,
109 and 131 MPa) were tested at each injection timing............................. 67

Figure IV-4: PM emissions at 25% load. Four injection timings (9˚, 5˚, 1˚ before top dead
center and 3˚ after top dead center) were tested for both ultra low sulfur
diesel (ULSD) and B40 blend. Three fuel injection pressures (52, 60 and 80
MPa) were tested at each injection timing. ................................................ 69

Figure IV-5: PM emissions at 50% load. Four injection timings (9˚, 5˚, 1˚ before top dead
center and 3˚ after top dead center) were tested for both ultra low sulfur
diesel (ULSD) and B40 blend. Three fuel injection pressures (72, 90 and
108 MPa) were tested at each injection timing. ......................................... 71

Figure IV-6: PM emissions at 75% load. Four injection timings (9˚, 5˚, 1˚ before top dead
center and 3˚ after top dead center) were tested for both ultra low sulfur
diesel (ULSD) and B40 blend. Three fuel injection pressures (87, 109 and
131 MPa) were tested at each injection timing. ......................................... 72

Figure IV-7: NOx-PM emission trade-off at 25% load. The legend shows the fuel and
injection timing for each test. Three fuel injection pressures (52, 66 and 80
MPa) were tested at each injection timing for each fuel. Fuel injection
pressure is not distinguished in the plot. .................................................... 73

Figure IV-8: NOx-PM emission trade-off at 50% load. The legend shows the fuel and
injection timing for each test. Three fuel injection pressures (72, 90 and
108 MPa) were tested at each injection timing for each fuel. Fuel injection
pressure is not distinguished in the plot ..................................................... 74

Figure IV-9: NOx-PM emission trade-off at 75% load. The legend shows the fuel and
injection timing for each test. Three fuel injection pressures (87, 109 and
131 MPa) were tested at each injection timing for each fuel. Fuel injection
pressure is not distinguished in the plot ..................................................... 75

Figure IV-10: PM-NOx emission trade-off at the start of injection timing at 9˚ before top
dead center for ultra low sulfur diesel and B40. (a) 25% load; (b): 50% load;
(c): 75% load; (d): 75% load after Soxhlet of PM filters ........................... 77

Figure IV-11: Brake thermal efficiency at 25% load for ultra low sulfur diesel (left) and
B40 (right). The multiple points at each start of injection timing are due to
the multiple fuel injection pressures. .......................................................... 79

Figure IV-12: Brake thermal efficiency at 50% load for ultra low sulfur diesel (left) and
B40 (right). The multiple points at each start of injection timing are due to
the multiple fuel injection pressures. .......................................................... 81

viii
Figure IV-13: Brake thermal efficiency at 75% load for ultra low sulfur diesel (left) and
B40 (right). The multiple points at each start of injection timing are due to
the multiple fuel injection pressures. .......................................................... 82

Figure IV-14: Apparent heat release profile of different fuels and different fuel injection
pressures at 25% load. The SOI: a) 9˚BTDC, b) 5˚BTDC, c) 1˚BTDC, d)
3˚ATDC. ..................................................................................................... 85

Figure IV-15: Apparent heat release profile of different fuels and different fuel injection
pressures at 50% load. The SOI: a) 9˚BTDC, b) 5˚BTDC, c) 1˚BTDC, d)
3˚ATDC. ..................................................................................................... 88

Figure V-2: Oxygen equivalence ratio (φΩ) field for ULSD and B40 at different injection
pressures at the SOI of 7ºBTDC, mid-load ................................................ 99

Figure V-3: Oxygen equivalence ratio (φΩ) field for ULSD and B40 at different injection
pressures at the SOI of 9ºBTDC, mid-load .............................................. 100

Figure V-4: The correlation between brake specific NOx emissions and average oxygen
equivalence ratio of the fuel-air mixture at lift-off length at 25% load,
regardless of fuel type. The legend: xB is x degree before TDC; xA is x
degree after TDC. ..................................................................................... 102

Figure V-5: The correlation between brake specific NOx emissions and average oxygen
equivalence ratio of the fuel-air mixture at lift-off length at 50% load,
regardless of fuel type. The legend: xB is x degree before TDC; xA is x
degree after TDC. ..................................................................................... 103

Figure V-6: The correlation between brake specific NOx emissions and average oxygen
equivalence ratio of the fuel-air mixture at lift-off length at 75% load,
regardless of fuel type. The legend: xB is x degree before TDC; xA is x
degree after TDC. ..................................................................................... 104

Figure V-7: semi-ln plot of absolute value of slope vs. SOI .......................................... 105

Figure V-8: semi-ln plot of Y-axis intercept (at X=0) vs. SOI ....................................... 106

Figure VI-1: Emissions of B20 engine test with various post injection strategies. Samples
are differentiated with post injection conditions while rest parameters were
kept the same. The fuel for the non-ULSD test was B20. ........................ 109

Figure VI-2: Exhaust temperature and BSFC profile of the 100-hour test. Each point is
the average of one test period. The dash lines are used to divide zones of
each post injection strategy. ..................................................................... 111

Figure VI-3: Pressure traces of test with different post injection strategies at both TDC
(top) and late crank angle (bottom) .......................................................... 113

ix
Figure VI-5: Bulk cylinder temperature of tests with different post injection strategies.
The arrow points to the lowest acceptable combustion temperature. ....... 117

Figure VI-7: GC-MS results of diesel, B100, lighter cut from 432˚C distillation of fresh
and used lubricant (top); Region of interest in the GC-MS results (bottom).
The numbers at the diesel zone means the number of carbon in relevant
paraffin. .................................................................................................... 124

Figure VI-8: GC-MS results of hydrocarbons collected from exhaust gases for various
post injection strategies ............................................................................ 126

Figure VI-9: FAME composition of B100 and of the biodiesel in used lubricant. ........ 127

Figure VII-1: A transmission electronic microscopy (TEM) image of diesel soot at 30%
load and 50 MPa fuel injection pressure .................................................. 130

Figure VII-2: A transmission electronic microscopy (TEM) image of B20 soot at 30%
load and 50 MPa fuel injection pressure .................................................. 131

Figure VII-3: Primary particle sizes for both diesel and B20 soot ................................. 133

Figure VII-4: Fractal dimensions for both diesel and B20 soot...................................... 134

Figure VII-5: Number of primary particles for both diesel and B20 soot ...................... 135

Figure VII-6: Particle size distribution of diesel and B20 combustion exhaust, 30% load,
before thermodenuder............................................................................... 136

Figure VII-7: Particle size distribution of diesel and B20 combustion exhaust, 30% load,
after thermodenuder ................................................................................. 137

Figure VII-8: Particle size distribution of diesel and B20 combustion exhaust, 60% load,
before thermodenuder............................................................................... 138

Figure VII-9: Particle size distribution of diesel and B20 combustion exhaust, 60% load,
after thermodenuder ................................................................................. 139

Figure VII-10: Thermogravimetric analysis result of 30% load diesel and B20 soot, m/m0
indicates the residual mass weight percentage. ........................................ 140

Figure VII-11: Thermogravimetric analysis result of 60% load diesel and B20 soot, m/m0
indicates the residual mass weight percentage. ........................................ 141

Figure VII-12: Apparent oxidative reactivity of 30% load diesel and B20 soot ............ 143

Figure VII-13: Apparent oxidative reactivity of 60% load diesel and B20 soot ............ 144

Figure VII-14: AD1/Ag ratio with fuel injection pressure for both diesel and B20 soot . 146

x
Figure VII-15: D1 FWHM values with fuel injection pressure for both diesel and B20
soot ........................................................................................................... 147

Figure VIII-1: comparison of simulated motor pressure trace with engine data, running at
1500 rpm, 30% load, single injection timing at 1º before top dead center,
100 MPa fuel injection pressure ............................................................... 158

Figure VIII-2: Representation of apparent heat release (AHR) at 1500 rpm, 30% load, 1º
before top dead center injection, 100 MPa fuel injection pressure with a
convolution of two peaks: one Gaussian peak and one ExpConvExp peak.
Top: difference between the engine data and fit curve; middle: comparison
of engine data and fit curve; bottom: deconvolution of artificial peaks. .. 159

Figure VIII-3: Representation of apparent heat release (AHR) at 1500 rpm, 60% load, 1º
before top dead center injection, 100 MPa fuel injection pressure with a
convolution of two peaks: one Gaussian peak and one ExpConvExp peak.
Top: difference between the engine data and fit curve; middle: comparison
of engine data and fit curve; bottom: deconvolution of artificial peaks. .. 160

Figure VIII-4: comparison of simulated pressure trace with engine data, running at 1500
rpm, 30% load, single injection timing at 1º before top dead center, 100
MPa fuel injection pressure ...................................................................... 163

Figure VIII-5: Examples of apparent heat release with: h1/h1b = 0.2, h2/h2b=1.29;
h1/h1b = 0.6, h2/h2b=1.17; h1/h1b = 1, h2/h2b=1 .................................. 166

Figure VIII-6: Effect of start of combustion on indicated thermal efficiency, h1/h1b = 0.2,
h2/h2b=1.29; h1/h1b = 0.6, h2/h2b=1.17; h1/h1b = 1, h2/h2b=1 ............ 167

Figure VIII-7: Effect of start of combustion on brake thermal efficiency, h1/h1b = 0.2,
h2/h2b=1.29; h1/h1b = 0.6, h2/h2b=1.17; h1/h1b = 1, h2/h2b=1 ............ 168

Figure VIII-8: Examples of apparent heat release with: h1/h1b = 1, h2/h2b=1; h1/h1b = 2,
h2/h2b=0.66; h1/h1b = 3, h2/h2b=0.34; h1/h1b = 4, h2/h2b=0.02 .......... 169

Figure VIII-9: Effect of start of combustion on indicated thermal efficiency, h1/h1b = 1,


h2/h2b=1; h1/h1b = 2, h2/h2b=0.66; h1/h1b = 3, h2/h2b=0.34; h1/h1b = 4,
h2/h2b=0.02 .............................................................................................. 170

Figure VIII-10: Effect of start of combustion on brake thermal efficiency, h1/h1b = 1,


h2/h2b=1; h1/h1b = 2, h2/h2b=0.66; h1/h1b = 3, h2/h2b=0.34; h1/h1b = 4,
h2/h2b=0.02 .............................................................................................. 171

Figure VIII-11: Examples of apparent heat release with: w1/w1b = 0.6, h2/h2b=1.17;
w1/w1b = 0.8, h2/h2b=1.11; w1/w1b = 1, h2/h2b=1 ............................... 172

Figure VIII-12: Effect of start of combustion on indicated thermal efficiency, w1/w1b =


0.6, h2/h2b=1.17; w1/w1b = 0.8, h2/h2b=1.11; w1/w1b = 1, h2/h2b=1 . 173

xi
Figure VIII-13: Effect of start of combustion on brake thermal efficiency, w1/w1b = 0.6,
h2/h2b=1.17; w1/w1b = 0.8, h2/h2b=1.11; w1/w1b = 1, h2/h2b=1 ........ 174

Figure VIII-14: Examples of apparent heat release with: w1/w1b = 1, h2/h2b=1; w1/w1b
= 2, h2/h2b=0.66; w1/w1b = 3, h2/h2b=0.34; w1/w1b = 4, h2/h2b=0.02 175

Figure VIII-15: Effect of start of combustion on indicated thermal efficiency, w1/w1b = 1,


h2/h2b=1; w1/w1b = 2, h2/h2b=0.66; w1/w1b = 3, h2/h2b=0.34; w1/w1b
= 4, h2/h2b=0.02 ...................................................................................... 176

Figure VIII-16: Effect of start of combustion on brake thermal efficiency, w1/w1b = 1,


h2/h2b=1; w1/w1b = 2, h2/h2b=0.66; w1/w1b = 3, h2/h2b=0.34; w1/w1b
= 4, h2/h2b=0.02 ...................................................................................... 177

Figure VIII-17: Examples of apparent heat release with different peak distances ......... 178

Figure VIII-18: Effect of start of combustion on indicated thermal efficiency with


different peak distance, ∆L = l2-l1 ............................................................ 179

Figure VIII-19: Effect of start of combustion on brake thermal efficiency with different
peak distance ............................................................................................ 180

Figure VIII-20: Examples of apparent heat release profiles for sharp peak (h1/h1b = 8)
versus low and wide peak (w1/w1b = 8) .................................................. 181

Figure VIII-21: Indicated thermal efficiency for sharp peak (h1/h1b = 8) versus low and
wide peak (w1/w1b = 8) ........................................................................... 182

Figure VIII-22: Brake thermal efficiency for sharp peak (h1/h1b = 8) versus load and
wide peak (w1/w1b = 8) ........................................................................... 183

Figure VIII-23: Heat transfer loss for sharp peak (h1/h1b = 8) versus load and wide peak
(w1/w1b = 8) ............................................................................................ 184

Figure VIII-24: Comparison of cylinder pressure for sharp peak (h1/h1b = 8) versus load
and wide peak (w1/w1b = 8) apparent heat release.................................. 185

Figure VIII-25: Indicated work for sharp peak (h1/h1b = 8) versus low and wide peak
(w1/w1b = 8) ............................................................................................ 186

Figure VIII-26: P-V diagram for sharp peak (h1/h1b = 8) versus low and wide peak
(w1/w1b = 8) ............................................................................................ 187

Figure VIII-27: Comparison of two apparent heat release (AHR) profiles with the same
start of combustion timing at 362.2º (2.2º after top dead center), indicated
efficiency of AHR1 = 49.74%, indicated efficiency of AHR2 = 47.23% 188

xii
Figure VIII-28: Comparison of cylinder pressure traces of two apparent heat release
(AHR) profiles with the same start of combustion timing at 362.2º (2.2º
after top dead center), indicated efficiency of AHR1 = 49.74%, indicated
efficiency of AHR2 = 47.23%.................................................................. 189

Figure VIII-29: Comparison of two apparent heat release (AHR) profiles with the same
start of combustion timing at 350º (10º before top dead center), indicated
efficiency of AHR1 = 50.02%, indicated efficiency of AHR2 = 50.84% 190

Figure VIII-30: Comparison of cylinder pressure traces of two apparent heat release
(AHR) profiles with the same start of combustion timing at 350º (10º
before top dead center), indicated efficiency of AHR1 = 50.02%, indicated
efficiency of AHR2 = 50.84%.................................................................. 191

Figure B-1: Repeatability of thermogravimetric analysis for diesel soot, 30% load and 75
MPa .......................................................................................................... 238

Figure E-1: Repeatability of particle size distribution of exhaust gases running with B20
at 30% load and 50 MPa fuel injection pressure ...................................... 242

xiii
List of Tables

Table I-1: Emission regulation for light and heavy duty diesel engines [3] ....................... 2

Table II-1: Chemical structure of common fatty acids, taken from [38] ............................ 9

Table II-2: Chemical composition of biodiesel derived from different feedstock, taken
from [39]..................................................................................................... 10

Table II-3: Physical and chemical properties of common fatty acid methyl esters
(FAMEs), taken from [39].......................................................................... 10

Table II-4: The effects of property differences between soy methyl ester biodiesel or its
blend and petroleum diesel on engine parameters and fuel combustion. A
positive/negative sign in the “difference” column means increase/decrease;
a positive/negative sign in the rest column means increase (or
advance)/decrease (or retarding) of the property. Taken from [44] ........... 12

Table III-1: Engine specifications ..................................................................................... 49

Table III-2: Engine operating parameters in Task 1 and 2 ............................................... 54

Table III-3: Engine operating parameters in Task 3 ......................................................... 55

Table III-4: Engine operating parameters in Task 4 ......................................................... 56

Table V-1: Input parameters during calculation ............................................................... 94

Table VI-1: Equivalence ratio ϕ, mean pressure P mean temperature T and ignition delay
τ ID calculated for each post injection case. .............................................. 118

Table VI-2: Calculated UHC emission rate and consumed HC by combustion at late
crank angle. B20 is a fuel mixture blended with 20% v/v SME based B100
and 80% v/v ultra low sulfur diesel. ......................................................... 120

Table VIII-1: Engine basic parameters used by friction model. ..................................... 155

Table C-1: Repeatability of Raman analysis for B20 soot, 30% load and 50 MPa fuel
injection pressure ...................................................................................... 239

Table D-1: Repeatability of number of primary particles and fractal dimension for B20
soot, 60% load at 75 MPa fuel injection pressure .................................... 240

Table D-2: repeatability of measurement of diameter of primary particles of diesel soot,


60% load at 125 MPa fuel injection pressure ........................................... 241

xiv
Nomenclature

AZ = autoignition zone

AHR = apparent heat release

CO = carbon monoxide

PM = Particulate Matter

NOx = Nitrogen Oxides

IVC = inlet valve close

EVO = outlet valve open

LCA = late crank angle

LOL = Lift-Off Length

l1,l2 = locations of artificial heat release peaks

h1, h2 = height of artificial heat release peaks

w1 = width of artificial Gaussian heat release peaks

l1b, l2b, h1b, h2b, w1b = location, height and width of apparent heat release at baseline

TDC = top dead center

TAN = total acid number

FAME = Fatty Acid Methyl Esters

SME = Soybean Methyl Esters

SMPS = scanning mobility particle sizer

SOI = Start of Injection

SOC = Star of Combustion

ULSD = Ultra Low Sulfur Diesel

UHC = Unburned Hydrocarbon

xv
mfMEP = mechanic fraction mean effective pressure

IMEPg = gross indicated mean effective pressure

τID = ignition delay

ϕ = overall equivalence raito

x = axial coordinate

r = radical coordinate

ϕ(x) = average equivalence ratio of a fuel spray at x

(A/F)st = stoichiometric air fuel ratio

ϕΩ = oxygen equivalence ratio

x+ = penetrate length scale of fuel jet

ρf = fuel density

ρa = air density

d = diameter of orifice

Ca = area-contraction coefficient

nC = nuber of carbon atoms in fuel

nH = number of hydrogen atoms in fuel

nO = number of oxygen atoms in reactants

H = lift-off length

Uf = velocity of injected fuel

Zst = stoichiometric mixture fraction of fuel

p = mean pressure

P = in-cylinder pressure

V = chamber volume

xvi
T = in-cylinder temperature

T = mean temperature

mɺ post = rate of the mass of post injected fuel

η = engine thermal efficiency

θ = crank angle

γ = Cp/Cv, ratio of constant pressure heat capacity and constant volume heat capacity

qt = total heat added

qw = engine wall heat transfer

xvii
Acknowledgements

I want to express my sincere gratitude to my academic advisor, Dr. André L.

Boehman, for his support, guidance, enthusiasm, patience and encouragement during this

study. Thanks to his excellent direction, I could explore new field of knowledge and

complete my Ph.D. study.

I would also like to express my gratitude to Dr. Daniel C. Haworth, Dr. Randy L.

Vander Wal and Dr. Yaw D. Yeboah for their valuable comments and suggestions to

improve the quality of my research and dissertation as my committee.

I also want to thank Dr. Magin Lapuerta Amigo from University of Castilla-La

Mancha, Spain and Dr. John R. Agudelo from Universidad de Antioquia, Colombia for

their generous help and instructions in the work of soot oxidative reactivity and engine

efficiency analysis.

I also want to thank Dr. Joseph M. Perez and his tribology group in the

Department of Chemical Engineering for their generous help in the lubricant analysis.

I also want to thank my group members at the Diesel Combustion and Emissions

lab for their help. In particular, I am grateful to Vince Zello, Greg Lilik and Bhaskar

Prabhakar for their help in the engine test stand, to Chenxi Sun for her help in soot TEM

xviii
imaging, to Eduardo Barrientos for his help in TGA operating, to Dongil Kang and

Vickey Kalaskar for their general help and discussions during my research.

I also want to thank Dr. Dania A. Fonseca for her generous help in GC analysis

and Joe Stitt for the Raman spectroscopy analysis in the Material Characterization Lab.

I want to give my special thanks to my family for their support during my work.

The financial support and valuable discussions from Infineum, USA, GE

Transportation, Volvo Powertrain and Oak Ridge National Lab are heartily

acknowledged.

xix
Chapter I. Introduction

Since it was first developed in the late 19th century, the diesel engine has been

widely used in heavy duty transportation for its higher output torque and brake thermal

efficiency than gasoline engine [1]. Through the past century, although the fundamental

design of the diesel engine hasn’t been significantly changed, the regulation for diesel

engine exhaust emissions, including carbon monoxide (CO), unburned hydrocarbon

(UHC), nitrogen oxides (NOx) and particulate matter (PM), are becoming increasingly

strict. Among these exhaust gases, the amount of CO and UHC from conventional engine

operating mode are much lower than NOx and PM, therefore the latter are more of

interest to regulators and researchers. Table I-1 presents the current and future planned

emission regulations for both light duty and heavy duty vehicles from authorities

including environmental protection agency (EPA) and California government. Significant

difficulties can be expected for current emission control technologies. For example, the

EPA Tier 2 Bin 5 NOx regulation is likely beyond the reach of today’s de-NOx

aftertreatment considering cold start, and in 2010 only 30% of current new cars in

California are able to meet the SULEV program [2]. For heavy duty transportation, very

few engines before 2010 are able to meet the US 2010 regulation [3], i.e., most on-road

engines need to be improved in emission control.

1
Table I-1: Emission regulation for light and heavy duty diesel engines [3]

NOx PM CO HCHO
(g/mile) (g/mile) (g/mile) (g/mile)
Bin 5 (2009 0.07 0.01 4.2 0.018
Light duty
effective)
(EPA Tier
Bin 4 0.04 0.01 2.1 0.011
2)
Bin 3 0.03 0.01 2.1 0.011
Bin 2 0.02 0.01 2.1 0.004
Bin 1 0 0 0 0
Light duty SULEV 0.02 0.01 1 0.004
(California) program
NOx PM CO NMHC
Heavy duty (mg/kWh) (mg/kWh) (mg/kWh) (mg/kWh)
US 2010 260 13 - 182

In addition to the emission regulation, current engine technologies are also facing

the challenge of sustainable development. Given that it is widely believed that the fossil

energies – coal and petroleum, which have supported the past 200 year’s industrial

development – may not be sufficient to satisfy the increased need for energy in the future,

studies on energy sustainability are more and more popular. Typically, sustainable

energy development involves three aspects: energy saving on demand side, replacement

of fossil fuel with renewable fuels and efficiency improvement in energy applications [4].

For engineering practice, the evaluations of applications of renewable fuels and

efficiency improvements are especially important, which is therefore the focus of the

present work. With respected to diesel engines, evaluation of biodiesel (primarily on the

emission impact as discussed previously) as a renewable fuel and engine thermal

efficiency improvements are investigated.

Biodiesel is a fuel which is comprised of monoalkyl esters of long chain fatty

acids derived from vegetable oils or animal fats [5], and it has been studied as an

alternative fuel to petroleum diesel for compression-ignition engines, both in neat form

2
(i.e., 100% esters) or in blended form with diesel fuel [6]. Compared with petroleum

diesel fuel, biodiesel has many advantages: it is renewable and biodegradable; it has a

high flash point which improves transportation safety [7]; and it can reduce exhaust

levels of some regulated emissions including CO, UHC and PM [8-11]. However, some

disadvantages of biodiesel are limiting the growth of its usage, such as decreased

oxidative stability [12], degraded cold-flow performance compared with petroleum diesel

[7], and increased emissions of NOx [8-11, 13]. Among these, the increase of NOx

emissions has especially drawn researcher’s attention and attempts to explain the origin

of the biodiesel NOx effect have been made [14-18], which may contribute to potential

approaches to biodiesel NOx reduction.

Traditionally, there are two ways for control of tailpipe NOx emissions: control of

in-cylinder combustion, and after treatment devices like NOx absorber catalysts and

selective catalytic reduction [19, 20]. The former can be affected by a variety of factors

including the physical and chemical properties of fuel, the engine control strategy (e.g.

fuel injection control), exhaust gas recirculation (EGR) ratio, and boost pressure. The

latter can be affected by the type of catalyst, exhaust gas composition and temperature of

the treatment device. The interest of this work focuses mainly on the engine in-cylinder

combustion control via fuel injection strategy, i.e., the impact of start of injection (SOI),

fuel injection pressure, fuel injection duration and multiply injections on the combustion

and, consequently, engine emissions.

Experiments conducted by previous researchers have shown that fuel injection

strategy can significantly affect engine emissions [21, 22]. For instance, it has been

shown that advancing the SOI can increase NOx emissions and decrease PM emissions

3
[21, 23]. Note this observation was also claimed to contribute to the biodiesel NOx effect

in traditional mechanical (pump line nozzle type) fuel injection systems [23] given that

biodiesel has a higher bulk modulus of compressibility. Nevertheless, the modern

common-rail injection system will not be affected by the fuel compressibility [13]. Fuel

injection pressure can also significantly affect the engine emissions. Increased fuel

injection pressure can increase NOx emissions and decrease PM emissions [19, 20]. This

observation is of special interest with biodiesel fuel. Since biodiesel has a lower heating

value than petroleum diesel [24], a higher brake specific fuel consumption (increased

injection quantity per unit time) for biodiesel is observed [6], and the default engine

calibration will achieve the increased injection quantity by increasing both the injection

pressure and the injection duration. Therefore, this work also intends to investigate which

is the more significant factor for the biodiesel NOx effect: biodiesel’s chemical properties,

increased injection pressure or increased injection duration. Furthermore, it has been

reported by many researchers that multi-injection strategy, especially multi-injection with

post injection [a fuel injection considerable after top dead center (ATDC)], can improve

engine emissions [13, 25-30]. Therefore, a thorough investigation of the post injection

strategy is also conducted in this work.

The other aspect of sustainable energy development – efficiency improvement in

energy applications (i.e., brake thermal efficiency of a diesel engine) is also an interesting

topic. Although biodiesel fueling introduces higher brake specific fuel consumption, the

brake thermal efficiency of a diesel engine operating with biodiesel has been observed to

be the same level as with diesel fuel [31]. Meanwhile, variation of fuel injection strategy,

which consequently affects combustion phasing, can be a significant factor to affect

4
engine brake thermal efficiency [22]. Theoretical investigation by Hoppie et al. [32]

suggests that neither the Otto nor the Diesel cycle is optimum for realistic engines, and an

optimum rate of heat release may be attainable via hypergolic combustion. However, the

artificial heat release profiles employed by them are not suitable for practical engineering.

Hence in this work, the author applies an analysis of optimized heat release pattern,

which can be controlled by fuel properties and engine injection strategy, intending to

provide suggestions for high efficiency engine design. While it is acknowledged that

other factors such as turbo, EGR system and aftertreatment devices can also affect the

total brake thermal efficiency of a diesel engine, they are not included in the discussion of

this work.

In summary, this work investigates the impact of biodiesel fueling and fuel

injection strategy, including start of injection, fuel injection pressure/duration and split

injection on diesel engine emissions and performance. The origins of the biodiesel NOx

effect and the change of NOx emissions by fuel injection strategy are investigated in this

work. In addition, a theoretical attempt to achieve high engine efficiency by optimizing

the apparent heat release pattern is also conducted in this work.

Objective: in this work the intention is to understand the relation between engine

emissions, biodiesel fueling and fuel injection strategy. The significance of factors

contributing to the biodiesel NOx effect is evaluated. In this work it is also intended to

understand the impact of post injection on lubricating oil dilution with biodiesel fueling,

and the impact of fuel injection pressure on diesel engine soot characteristics. The

optimized heat release pattern toward high engine thermal efficiency is also investigated.

5
Hypothesis:

Main hypotheses of this work include:

• Biodiesel NOx penalty can be eliminated by engine fuel injection control

• The impact on NOx emission from engine fuel injection strategy control and

biodiesel fueling can be explained by the equivalence ratio at autoignition

zone near the flame lift-off length

• Biodiesel fueling can introduce lubricating oil dilution with post injection

strategy

• Soot from high injection pressure has higher oxidative reactivity since the

decreased soot precursor concentration can prolong the soot inception time.

• Optimized heat release design is a sharp peak at top dead center

To achieve the objectives of this research plan, the study is divided into 5 tasks:

Task 1: Investigation of the impact on engine emissions of both biodiesel fueling

and injection strategy at different engine speeds and loads.

Task 2: Investigation of the origin of engine NOx emission variations from

biodiesel or injection strategy by evaluating the adapted spray model.

Task 3: Investigation of the impact of post injection on engine emissions and

lubricating oil dilution.

Task 4: Investigation of the impact of fuel injection pressure on diesel engine soot

oxidative reactivity with both diesel and biodiesel blends.

Task 5: Investigation of optimized heat release pattern to achieve high engine

thermal efficiency.

6
Chapter II. Literature Review

2.1 Biodiesel background

Throughout this thesis biodiesel implies a fuel composed of glyceride-free mono-

alkyl esters of long-chain fatty acids converted from triglycerides such as biologically-

based fats and oils [33-35]. In fact, attempts to use biologically derived hydrocarbons

(not only biodiesel) in internal combustion engines have been conducted for more than a

century. For instance, in 1900 when Dr. Diesel exhibited his great invention – the Diesel

engine – at the Paris World Exposition for the first time, the engine was running on 100%

peanut oil [33, 34]. From the 1920s to the 1940s numerous applications of vegetable oils

as diesel engine fuels were reported simply based on the theme of providing tropical

colonies of European countries with an independent source of fuel [36]. The more

appropriate form of biologically-derived fuel (ethyl esters of palm oil), which meets the

present definition of biodiesel, was initially included in a Belgian patent issued to

Chavanne as early as 1937 [37]. However, the petroleum derived diesel fuel immediately

became the primary fuel for diesel engines, and the concept of biologically-derived fuel

hadn’t drawn people’s attention until the 1970s due to the energy crises. Since then

vegetable oil-based fuels have been reconsidered among the alternative energy sources.

In addition, the feedstocks for biodiesel have also been expanded to a variety of vegetable

oils including soybean, rapeseed, sunflower, etc., which also creates a variation in both

physical and chemical properties of biodiesel [34].

The production of biodiesel mainly follows the transesterification process, which

can crack the large molecules (triglyceride) of vegetable oils/fats and transform them to

7
fatty acid methyl esters (FAME) with similar molecular weight as diesel fuel. Figure II-1

presents a traditional transesterification process for biodiesel production. This reaction

requires methanol as a reactant and alkali or acid as catalyst, and the by-product of the

reaction is glycerol. Demirbas [35] proposes that the same reaction can also be promoted

with supercritical methanol, with which no catalysts are needed, and consequently

purification of the final product can also be simplified.

Figure II-1: A diagram representing the transesterification process, replicated from


[6]. The general scheme of the process is to modify a triglyceride (a fat or oil which
is a glyceride with three fatty acids, represented as R1, R2 andR3) into a fatty acid
methyl ester.

8
With different structures (length, saturation) in the fatty acid, the physical and

chemical properties of FAME may vary. Table II-1 presents the chemical structure of

common fatty acids.

Table II-1: Chemical structure of common fatty acids, taken from [38]

Fatty acid Systematic name Structure* Formula


Lauric Dodecanoic 12:0 C12H24O2
Myristic Tetradecanoic 14:0 C14H28O2
Palmitic Hexadecanoic 16:0 C16H32O2
Stearic Octadecanoic 18:0 C18H36O2
Arachidic Eicosanoic 20:0 C20H40O2
Behenic Docosanoic 22:0 C22H44O2
cis-9-Hexadecenoic 16:1 C16H30O2
Palmitoleic Acid
Oleic cis-9-Octadecenoic 18:1 C18H34O2
cis-9,cis-12- 18:2 C18H32O2
Linoleic Octadecadienoic
cis-9,cis-12,cis-15- 18:3 C18H30O2
Linolenic Octadecatrienoic
- cis-11-Eicosenoic 20:1 C20H38O2
* xx:y indicates xx carbons in the fatty acid chain with y double bounds

Although these FAMEs are found in biodiesel, the chemical composition of

biodiesel is different if the biodiesel is derived from different feedstock, as shown in

Table II-2. It can be seen that generally C16:0, C18:0, C18:1 and C18:2 are the major

FAMEs species in biodiesel. In the United States, the primary feedstock for biodiesel is

soybean oil [12], which predominantly contains unsaturated FAME – C18:1, C18:2 and

C18:3, as well as, a small amount of saturated FAME – C16:0 and C18:0. The difference

in physical and chemical properties of FAME can also be significant, as shown in Table

II-3. Saturated FAME with a long chain always leads to higher melting point and

kinematic viscosity, which is expected. However, the degree of saturation can

significantly affect certain properties including melting point, oil stability index and

cetane number.

9
Table II-2: Chemical composition of biodiesel derived from different feedstock, taken from [39]

Fatty acid composition (%)


feedstock
14:0 16:0 18:0 20:0 22:0 16:1 18:1 18:2 18:3 20:1 Other
Canola oil 0 4 2 0 0 0 61 22 10 1 0
Palm oil 1 45 4 0 0 0 39 11 0 0 0
Soybean oil 0 11 4 0 0 0 23 54 8 0 0
Sunflower oil 0 6 5 0 1 0 29 58 1 0 0
Corn oil 0 11 2 0 0 0 28 58 1 0 0
Cottonseed oil 1 23 2 0 0 1 17 56 0 0 0
Chicken fat 1 25 6 0 0 8 41 18 1 0 0
Beef tallow 4 26 20 0 0 4 28 3 0 0 14

Table II-3: Physical and chemical properties of common fatty acid methyl esters (FAMEs), taken from [39]

Kinematic
Melting pointa Heat value Oil stability indexc Lubricitye (µm) at
FAME viscosityb at 40 ᵒC Cetane No.d
(ᵒC) (MJ/kg) (h), at 110 ᵒC 60 ᵒC
(mm2/s)
C14:0 19 39.45 3.3 >40 - 353
C16:0 31 39.45 4.38 >40 86 357
C16:1 −34 39.3 3.67 2.1 51 246
C18:0 39 40.07 5.85 >40 101 322
C18:1 −20 40.09 4.51 2.5 59 290
C18:2 −35 39.7 3.65 1 38 236
C18:3 −52 39.34 3.14 0.2 23 183
a: Ref. [7, 40]; b: Ref. [41]; c: Ref. [7]; d: Ref. [7]; e: Ref. [42]

10
Although cetane numbers of the C18:2 and C18:3 FAMEs are relatively low, as a

mixture, biodiesel is often observed to have a higher cetane number than petroleum diesel

[39], which improves its performance in diesel engines. High unsaturation can improve

the cold weather performance, but also significantly decrease the oxidation stability.

Chemical properties of a fuel (e.g., hydrocarbon structure, aromatic content,

cetane number) dominate the fuel’s effect on combustion and pollutant formation [43].

Physical properties (e.g., viscosity, density) generally affect injection timing, fuel

atomization and fuel evaporation, which eventually have indirect effects on combustion

and emission formation. In other words, diesel engine may have a different response

when biodiesel is used. Table II-4 presents a summary by Sun et al. [44] on the impact of

soy methyl ester (SME) based biodiesel and its blend with petroleum diesel on physical

and chemical properties and engine response. This summary also reveals the difficulty in

biodiesel research that multi-parameters can change simultaneously with biodiesel

fueling, so that the understanding of each parameter’s impact is not straightforward. For

instance, the lower heating value of biodiesel can lead to an increase injection pressure

and duration by the default engine calibration, which may also have impact on engine

emissions [21, 22]. This issue will also be addressed in this work.

11
Table II-4: The effects of property differences between soy methyl ester biodiesel or its blend and petroleum diesel on engine
parameters and fuel combustion. A positive/negative sign in the “difference” column means increase/decrease; a
positive/negative sign in the rest column means increase (or advance)/decrease (or retarding) of the property. Taken from [44]
Difference Injection Injection Fuel spray Fuel spray Fuel spray Ignition Heat Combustion
timing pressure penetration angle atomization delay release temperature
Physical Properties
Liquid densitya + + + - - +
Bulk modulus of + + + +
compressibilityb
Speed of soundb + + + +
a
Liquid viscosity + + - - +
a
Surface tension + + + -
Vapor pressurea - - - +
c
Volatility - -
Liquid specific - + -
a
heat
Vapor specific - + -
a
heat
Heat of + -
a
vaporization
Chemical Properties
Chain lengthd + -
Oxygen contentc + + - + +
e
Aromatics content - + + +
f
Sulfur content - +
Saturation (iodine - + +
d
value)
Cetane numberg + -
c
Heating value - -
a: Ref [45]; b:Ref. [23, 46]; c: Ref. [47]; d: Ref.[48]; e: Ref. [49]; f: Ref.[15]; g: Ref. [50]

12
at the same time, the “sustainability” concept of biodiesel has also been criticized

since the methanol needed for biodiesel production majorly comes from petroleum [35].

Nevertheless, approximately 95% of the carbon in biodiesel are derived from renewable

sources such as oil and fat [33]. Hence, it is still of interest to investigate the impact of

biodiesel on engine emissions and performance to approach the future of sustainable

energy.

2.2 Impact of biodiesel fueling on engine emissions

The review by Graboski and McCormick in 1998 [6] and a more recent review by

Lapuerta et al. in 2008 [15] show consistently that the majority of diesel engine tests with

biodiesel fueling show reductions of exhaust levels of some regulated emissions

including unburned hydrocarbons (UHC), carbon monoxide (CO), and particulate matter

(PM), but increases of emissions of nitric oxides (NOx). For example, Figure II-2 shows

the statistical analysis of trends of the impact of biodiesel blending on emissions from

heavy-duty engines, conducted by the Environmental Protection Agency (EPA) [51].

However, despite the “statistical analysis” by the EPA, the impact of biodiesel on engine

emissions as observed in Figure II-2 may be limited to certain engine conditions: in this

case it is a heavy-duty truck engine on highway. The understanding of biodiesel’s impact

on diesel engine emissions requires a thorough investigation of previous engine studies.

13
Figure II-2: NOx, PM, CO and HC % changes as biodiesel blending percentage
increase as determined by the EPA through statistical regression of publically
available data on highway heavy-duty truck engines [18, 51]. Taken from [44]

In addition, as discussed in the introduction, since NOx and PM emissions from

current technology diesel engines are close to the limits permitted by regulation and both

limits will be even more stringent in the near future, these two emissions will be critical

factors in the development of new diesel engines and aftertreatment systems. For the

other regulated emissions, CO and UHC, no further development in engine technology

seems to be necessary to meet future limits [15].

14
2.2.1 Nitric oxides (NOx)

In their comprehensive review, Lapuerta et al. [15] states that a majority of the

literature report an increase of NOx emissions with biodiesel use and a minority of other

observations. In this sense, they divide the impact of biodiesel fueling on NOx emissions

into 4 groups:

• Group I: Increase.

• Group II: Increase only in certain operating conditions

• Group III: No significant difference

• Group IV: Decrease with biodiesel fueling

Similar categorization is also adapted in another review by Sun et al. [44], except

that they claim that the majority of the literature they reviewed reported inconsistent

trends in the effect on NOx emissions from the use of biodiesel, in other words,

considering the distribution of engine configurations and operating conditions, it is

difficult to arrive at a deterministic conclusion on the impact of biodiesel fueling on NOx

emissions. Nevertheless, observations associated with engine configurations and

operating conditions are still meaningful for biodiesel research.

Group I: In Graboski and McCormick’s review [6] of ten older two-stroke engines

running with biodiesel derived from soy and rapeseed, all engines running with rapeseed-

oil biodiesel show a consistent increase in NOx emissions, and two third of the engines

running with soybean-oil biodiesel show a consistent increase in NOx emissions. The

remaining one third engines running with soybean-oil biodiesel showed both increases

and decreases in NOx emissions depending on the operating conditions, which should be

categorized into Group II. FEV Engine Technology [52] carried out an experimental

15
study in a 7.31L Navistar engine running the 13-mode US Heavy-Duty test cycle using

different soybean blends and observed a maximum of 8% increase in NOx emissions

with B100. Marshall et al. [53] tested a Cummins L10E engine under steady conditions

with different biodiesel blends, and a maximum of 16% increase in NOx emissions was

observed with B100. However, in their transient test condition B20 blend gave a higher

increase in NOx emissions than B30. Szybist et al. [14] tested a Yanmar L70 EE air-

cooled, four-stroke, single cylinder direction injection diesel engine operating at steady

3600 rpm and 75% load and reported around 10% increase in NOx emissions with pure

soybean-oil biodiesel. Other experiments have reported increased NOx emissions with

biodiesel fueling [18, 54, 55].

Group II: Some authors claimed that the effect of biodiesel on NOx emissions

depends on the type of engine and its operating conditions. Serdari et al. [56] tested on-

road emissions from three different vehicles using high sulfur diesel fuel (1800ppm) and

10% sunflower-oil biodiesel blends. They found both increases and decreases in NOx

emissions, and attributed such differences to the different engine technology and

maintenance conditions. Hamasaki et al. [57] tested a single-cylinder engine at 2000 rpm

and different loads with three waste-oil biodiesel fuels. They measured slight decreases in

NOx emissions at low loads but increases at high loads. Similar observations have also

been reported by Zhang and Boehman [13]. They conducted an experiment in a

DDC/VM Motori 2.5L, 4-cylinder, turbocharged, common rail, direct injection light-duty

diesel engine and found that NOx increases with biodiesel fuel were only observable at

high engine load. At low engine load, no significant difference in NOx emissions can be

observed between the baseline diesel, B20 and B40. Yehliu et al. [25] conducted

16
experiments in the same engine with a thorough investigation of different load and speed

conditions, and found that at high speed and low load, biodiesel fueling decreased NOx

emissions. Staat and Gateau [58] tested a 6-cylinder engine following the ECE R49 test

cycle and an urban transient cycle named AQA F21 established by the French Agency of

Air Quality. They observed a 9.5% increase in NOx emissions in the ECE R49 cycle, yet

a 6.5% reduction in the transient urban cycle. The above observations appear to be

consistent with the review by Tat [59], that NOx emissions with biodiesel fuels are

usually higher than those from diesel fuel when they are measured in an engine test bench

but not when they are measured from vehicles. Tat [59] attributed this observation to the

reason that engine loads are usually lower in vehicles than those imposed in experimental

test rigs. Similar observations have also been reported by McCormick [60, 61] who

measured NOx emission reductions of around 5% when using 20% soybean-oil biodiesel

blends.

Attempts have been made to explain these “conditional” biodiesel NOx effects. Li

and Gulder [62] claimed that the higher cetane number of biodiesel has a more sensible

effect on NOx emission at low load than at high load. However, this understanding may

no longer work when high-cetane ultra low sulfur diesel (ULSD) fuel is used instead of

the conventional low sulfur diesel fuel, i.e., blending of biodiesel with ULSD may not

significantly change the cetane number of fuel. Tat [59] proposed another explanation:

the injection pump tended to advance the injection timing at low load, but he observed

that this advance was higher with diesel than with biodiesel fuel in a certain load range,

leading to increased NOx emissions with diesel fuel at these load condition. This

explanation, however, is not applicable for diesel engines with a common-rail injection

17
system, whose start of injection is well controlled and consistent for both petroleum

diesel and biodiesel, as in [13, 25, 63]. Although the same authors attributed the NOx

reduction with biodiesel to the specific engine configuration, it might arise from the

decreased apparent heat release from premixed combustion with biodiesel [63], due to its

lower energy content and a shortened ignition delay.

Group III: Durbin et al. [64] tested four different engines with diesel, pure

biodiesel and a 20% biodiesel blend. The engines were chosen to represent a wide variety

of heavy-duty engines: turbocharged and naturally aspirated, direction and indirect

injection, and no significant differences in NOx emissions were found. Mandpe et al. [65]

tested a Euro III emission certified common-rail diesel engine on road with Jatropha

Curcas seed oil biodiesel and found no significant difference in NOx emissions between

biodiesel and petroleum diesel. Nabi et al. [66] conducted an experiment with a single-

cylinder 9.8 kW engine at steady-state with neem-oil biodiesel at different EGR ratios.

Although they observed decreased NOx emissions with increased EGR ratio, they found

no significant differences in NOx emissions between diesel and biodiesel. As suggested

by Lapuerta et al.[15], the observation by Nabi et al. [66] may be actually due to the low

unsaturation level of neem oil, since Graboski et al. [67] have shown that NOx emissions

increased with unsaturation.

Group IV: A minor number of papers have reported decreases in NOx emissions

when using biodiesel fuels. Armas et al. [63] conducted an experiment in a DDC/VM

Motori 2.5L, 4-cylinder, turbocharged, common rail, direct injection light-duty diesel

engine at high speed and low load, with different start of injection. For each start of

injection they also fine-tuned the ECU to match the combustion phasing between

18
biodiesel blends and baseline diesel. They found slight NOx reduction with biodiesel

consistently for all start of injection timings. Lapuerta et al. [68] tested pure and blended

biodiesel from sunflower and cardoon oils in an indirection injection 1.9L engine

operating at five selected steady modes, and they observed a slight decrease in NOx

emissions with biodiesel.

Overall, although no unanimity has been found, the majority of observations in

the literature show a slight increase in NOx emissions with biodiesel fueling. Under

certain conditions, no significant difference or decreased NOx emissions were observed

with biodiesel, which was likely due to specific engine configuration or fuel composition,

which suggests that such results are not generally representative.

2.2.2 Particulate Matter (PM)

It has been consistently reported in the literature that a noticeable decrease in

diesel PM emissions can be observed with biodiesel fueling [6, 8, 9, 11, 58, 69-72]. The

Environmental Protection Agency collected a number of engine emissions data with

biodiesel [51] and proposed a statistical correlation between PM emissions level and

biodiesel blending:

PM
= e −0.006384⋅% B (1)
PM D

PM
where indicates the PM emission ratio between biodiesel and diesel, and %B is the
PM D

biodiesel blending level in percentage. The shape of Equation (1) can be also seen in

Figure II-2.

19
This trend proposed by EPA suggests that the maximum PM emission reduction is

around 50% with pure biodiesel. This observation is consistent with several other

investigations [54, 69, 73-75]. Other authors found even higher PM emission reductions

with biodiesel. Van Gerpen and Canakci [76] observed as high as 65% reductions of PM

emissions with soybean oil and waste oil based biodiesel fuels. Schumacher et al. [77]

tested PM emissions from a DDC series 60 diesel engine and reported more than 60%

PM emission reduction with biodiesel.

The reductions in PM emissions have been shown in general, as being more

effective with lower biodiesel concentrations in the blends [15]. In other words, the

reduction of PM emissions for each percentage of biodiesel blending, i.e., relative

reduction, has been observed to be the highest for partial blends [78-81].

The effects of biodiesel on PM emissions together with other parameters, such as

engine configurations and load conditions have also been investigated. In the review by

Krahl [82] on diesel engine tests with rapseed-oil biodiesel, the authors found that the PM

emission reductions were lower ( or there were no significant reductions) in heavy-duty

engines than in light-duty engines. Lapuerta et al. [79] investigated the PM emissions of a

typical European passenger car diesel engine operating at low-middle load conditions

and found that the PM emission reduction with biodiesel is larger at low load. Contrary

observations have also been reported [83]. Leung et al. [84] tested a single-cylinder

engine with a diesel fuel and a pure rapeseed oil based biodiesel at different load

conditions. They found larger decreases for biodiesel at high load. In addition, the

property of diesel fuel can also affect the NOx reduction by biodiesel. Boehman et al. [75]

reported a PM emission of 20% blending soybean oil based biodiesel with both low

20
sulfur diesel (325 ppm sulfur) and ultra low sulfur diesel (15 ppm sulfur) and found that

the reduction of PM emissions by biodiesel is more significant for low sulfur diesel.

Similar observations have also been reported [51].

A small number of studies did not find significant reductions (or even found

slightly increases) in PM emissions. Armas et al. [63] found increased PM emissions in

their experiment with a DDC/VM Motori 2.5L, 4-cylinder, turbocharged, common rail,

direct injection light-duty diesel engine at high speed and low load. Boehman et al. [75]

found a slight increase of PM emission for soybean oil based biodiesel blending with

ultra low sulfur diesel. Similar observations that PM emissions were increased with

biodiesel have also been reported in other places [74, 85]. This phenomenon has been

generally attributed to the higher soluble organic fraction (SOF) of the particulate [63,

75], which is widely accepted to be increased when using biodiesel [6, 67, 69, 74, 77] due

to its lower volatility of the unburned hydrocarbons (Table II-4). Meanwhile, the

insoluble fraction (ISF) was decreased with biodiesel fueling [75].

In summary, it is widely accepted that biodiesel fueling can decrease the PM

emissions of diesel engine. Therefore, it has been suggested that this impact of biodiesel

on PM emissions may be combined with engine injection strategy to optimize the NOx-

PM emission trade-off [15].

21
2.3 Origin of biodiesel emission effect

2.3.1 Biodiesel NOx effect

NOx formation chemistry:

To identify the factors that can affect NOx emissions, it is necessary to understand

the mechanism of NOx formation. Although a large number of reactants and reaction

pathways have been found to participate during NOx formation [86, 87], the key

reactions at the pressure, temperature and time scales in a diesel engine consist of three

major reactions [86], referred as to the extended Zeldovich mechanism:

N 2 + O 
k1
→ NO + N
O2 + N 
k2
→ NO + O (2)
OH + N 
k3
→ NO + H

Typically the time scales in a diesel engine do not allow these reactions to reach

equilibrium. Therefore, longer residence time of in-cylinder gases at high temperature

(above 1800K since the thermal NO formation below 1800K is generally negligible [87])

can lead to higher NOx emissions. A simple second order kinetic expression can be used

to describe the NO formation when its concentration is below equilibrium:

d [ NO ]
= k1[O ][ N 2 ] + k 2 [O2 ][ N ] + k3[OH ][ N ] (3)
dt

where [NO], [O], [N2] and [OH] are the molar concentrations of NO, O, N2 and OH,

respectively. The rate constant k can be generally expressed as [88]:

Eai
ki = BiT α i exp( − ), i = 1, 2,3 (4)
RuT

22
where Bi, αi, Ru and Eai are constants for each ki. Equations 2 and 3 suggest that the key

parameters that control the NOx formation include: 1) concentration of O-atom, N2, N-

atom and OH, 2) in-cylinder temperature, or 3) the residence time of the in-cylinder

mixture at high temperatures. Hence, any variation in NOx emissions must be related to

one or more of the above factors.

Biodiesel NOx effect:

To date, a significant number of studies have been conducted in the attempt to

explain the NOx variation and to identify the key parameters that affect it when

alternative fuels (e.g., biodiesel) are used, both from experiments [14, 15] and numerical

simulation [16, 17, 89]. Literature reviews by Mueller et al. [18], Lapuerta et al. [15] and

Sun et al. [44] have discussed the various proposed hypotheses that may account for

engine NOx variation with biodiesel fueling, which can be summarized into seven

categories:

(1) Injection timing. Alternative fuels have different compressibility than

petroleum diesel. For example, biodiesel has higher bulk modulus of

compressibility while FT diesel has lower bulk modulus of compressibility [23].

A higher bulk modulus can lead to a faster traveling of the pressure wave in the

fuel line of pump-line-nozzle fuel injection systems, causing an advance in the

start of injection (SOI) that can consequently lead to a longer residence time

and/or higher in-cylinder temperature. The advanced SOI of biodiesel and

retarded SOI of FT diesel have been observed to affect the NOx emissions [23, 90,

91]. Although this factor has been considered as the most solid argument for the

23
biodiesel NOx effect [15], the injection timing of a modern common-rail injection

system is not affected by the bulk modulus of compressibility of the fuel [13].

(2) Injection pressure and fuel spray. The lower heating value of biodiesel

results in an increase in the brake specific fuel consumption, which will be treated

as an increase of the fuel injection pressure/duration according to default engine

calibration [55]. This change in fuel injection parameters can yield an increase in

NOx emissions [21, 22]. Furthermore, the change of fuel properties, e.g. viscosity,

density and surface tension when biodiesel is used can increase the average

droplet diameter [92], leading to a lengthened ignition delay as well as a possibly

longer diffusion combustion period [6].

(3) Combustion phasing. That biodiesel can increase NOx emissions also can

be due to its impact on combustion phasing. The higher cetane number of

biodiesel can advance the start of combustion (SOC), which may lead to longer

residence time at high temperature for the combustion products [18]. However, as

discussed previously for FT diesel, some authors also claim that the NOx

reduction is due to its higher CN [25, 63, 91, 93] since very high CN indicates

short ignition delay, leading to less premixed burning [94, 95]. This inconsistency,

that both biodiesel and FT diesel have increased CN, leads to different direction in

change of NOx emissions, suggests that the impact of high cetane number on

NOx emissions is a combination of two effects.

(4) Premixed-burn fraction. The fact that biodiesel contains oxygen increases

the premixing quality (since fewer oxidizer molecules are required to reach

stoichiometric proportion) during the ignition delay, resulting in better

24
combustion and higher temperature. The oxygen content in biodiesel also

increases the local oxygen concentration, which can promote NOx formation as

discussed previously. Song et al. [96, 97] tested oxygenated diesel fuel (via

oxygenated fuel additives) and found increases in NOx emissions. However, this

explanation is questioned by other authors [68, 69, 89]. In addition, this

explanation is not suitable for FT diesel since it does not contain oxygen in the

fuel.

(5) Kinetics. Different chemical kinetic pathways for NOx formation exist for

different fuels. For example, it is argued that an increased level of CH exists at the

autoignition zone (AZ) during biodiesel combustion [98], which leads to the

production of N-atoms in the jet core as well as the promotion of NOx formation.

(6) Adiabatic flame temperature. Some authors claimed that biodiesel has a

slightly higher adiabatic flame temperature than conventional diesel, so thermal

NOx production is higher [16, 69, 99]. Although the decreased NOx emissions for

FT diesel appears to be consistent to this theory: FT fuel has a lower adiabatic

flame temperature [100], due to its lower C/H ratio [101], than the petroleum

diesel fuel. No unanimity can be found for the biodiesel since others state that the

adiabatic flame temperature is higher for diesel fuel [44, 66, 102].

(7) Radiation heat transfer. Soot radiation is the primary means of heat loss

from an in-cylinder flame. It has been suggested that since biodiesel can decrease

the PM emissions, the actual flame temperature is higher, leading to an increase

of NOx formation [103]. However, it has been suggested that soot emissions may

25
not correlate with flame luminosity and the soot radiation appears to be smaller

than other effects [18].

Mueller et al. [18] also pointed out that the biodiesel NOx increase is related to a

number of coupled mechanisms, which also have influences on each other, rather than a

single phenomenon. Note that many of those mechanisms are affected by both fuel

properties and engine operating conditions. For example, the premixed-burn quality can

be affected by the droplet size, which is controlled not only by the injection strategy and

fluid surface tension [1], but also by the difference in the rate of mass transfer during the

injection of the multi-component fuel [104]. Among those mechanisms, Mueller et al. [18]

suggest that the most important ones responsible for the biodiesel NOx increase are

longer residence time and higher in-cylinder temperature from either advances in

combustion phasing or lower radiative heat loss. They proposed, based on the models

established by Siebers and Naber [105, 106], that the origin of the NOx increase with

biodiesel is based on reacting mixtures that are closer to stoichiometric (less rich) for

biodiesel-containing fuels: a) during ignition (i.e., during the premixed volumetric

autoignition event from the start of combustion to the end of premixed burning), and b) in

the standing premixed autoignition zone (AZ) near the flame lift-off length (LOL) at

higher loads. According to Mueller et al. [18], this is so far the most generalized

explanation which is consistent with the observations from biodiesel, high cetane number

diesel and engine load conditions. In additional to their heavy-duty single-cylinder engine

test, in this work this theory will be tested with a light-duty diesel engine for the impact

of both biodiesel and engine fuel injection strategy.

26
2.3.2 Biodiesel particulate matter effect

As discussed previously, it is widely accepted that biodiesel can decrease the PM

emissions. It is also believed that the factors contributing to the PM emission reductions

include:

• Oxygen content in biodiesel: the oxygen content of the biodiesel molecule,

which can promote complete combustion even in regions of fuel-rich diffusion

flames, has been believed to contribute to the PM emission reduction with

biodiesel [107]. Song et al. [96, 97] tested oxygenated diesel fuel (via oxygenated

fuel additives) in a VW turbocharged direct injection diesel engine, and found

decreased PM emissions with increased oxygen content of fuel for both engines.

Similar observation has also been reported by Frijters and Baert [108]. Sison et al.

[109] tested different oxygenates, including biodiesel, with diesel fuel in a single-

cylinder optically accessible diesel engine and observed decreased soot formation.

The kinetic model developed by Flynn et al. [110] based on the conceptual model

of Dec [111] suggested that a sharp decrease in the formation of soot precursors

would be observed when the oxygen content in the fuel increased: as the oxygen

content of the fuel increased, larger fractions of the fuel carbon were converted to

CO in the rich premixed region, rather than to soot precursors.

• Absence of aromatics in biodiesel: The aromatic content in petroleum

diesel fuel has been considered to serve as soot precursors [69, 71, 112, 113]. This

can be investigated in two ways: 1) decrease the aromatic content in petroleum

diesel, (Schmidt and Van Gerpen [113] artificially blended diesel fuel with

octadecane (C18H38) and observed a significant reduction in PM emissions.); 2)

27
increase the aromatic content in biodiesel. Mueller et al. [18] doped the pure

biodiesel with phenanthrene, a 3-ring aromatic species, and compared the

emissions at both decreased oxygen environment and normal air environment

with a heavy-duty single-cylinder optically accessible diesel engine. They found

increased PM emissions for all load conditions for the doped biodiesel in a

decreased oxygen environment. With a normal air environment, the doped

biodiesel has an increased PM emission only at low-moderate load condition.

Other authors have treated both the oxygen content and absence of aromatics in

the biodiesel to the PM emission reductions [114, 115].

• Higher oxidative reactivity of biodiesel soot: Boehman et al. [75] collected

the particulate matter from the exhaust of a six-cylinder Cummins ISB 5.9L direct

injection turbocharged diesel engine, with both diesel fuel and a 20% blend

soybean oil based biodiesel, and then they employed thermogravimetry to analyze

the oxidative reactivity of the soot particles (de-volatized particulate matter).

They found that the soot particles for B20 are more reactive than for diesel fuel,

which suggests that the particulate matter will be more rapidly consumed by the

oxidation in the exhaust, leading to decreased PM emissions. Song et al. [116]

expanded this observation to pure biodiesel and used high resolution transmission

electronic microscopy (TEM) technique to confirm the faster oxidation of the soot

for pure biodiesel. Yehliu et al. [117] applied a counting algorithm to identify the

initial nanostructure of soot particles from high resolution TEM images, and they

found more short and curvy graphene layers in soot from biodiesel combustion,

28
i.e., it has a higher oxidative reactivity due to its higher degree of disorder [118-

120].

There have been other explanations proposed for the biodiesel PM effect, as well.

For instance, biodiesel can introduce an advance of combustion either due to advanced

injection timing [90] or higher cetane number [102] can prolong the residence time of

soot particles in a high-temperature atmosphere, which in the presence of oxygen

promotes further oxidation [72, 112]. Others also claimed that the usually lower final

boiling point of biodiesel provided lower probability of soot or tar being formed from

heavy hydrocarbon fractions unable to vaporize [69]. High sulfur content of diesel fuel

has also been suggested as a reason for its high PM emissions [121]. Nevertheless, these

explanations have been believed to be insignificant [108, 115, 122], as well as not

applicable since new fuel standards came into effect.

2.4 Fuel Injection strategy with biodiesel fueling

Fuel injection strategy is an important aspect in the optimization of diesel engine

missions. For instance, in earlier discussion the significant impact of injection timing and

injection pressure on engine NOx emissions has been shown. Some studies have

attempted to optimize diesel engine combustion with biodiesel fuel by adjusting the

injection strategy to improve exhaust emissions [123, 124]. Generally speaking, for

conventional diesel combustion, the control of fuel injection strategy includes three

aspects 1) injection timing (or start of injection); 2) injection pressure or injection

duration; 3) number of fuel injection. Among those, although the advance of injection

29
timing with biodiesel fueling was a concern for earlier engine research, it is now believed

as an insignificant factor for modern engines with common-rail fuel injection systems

[13]. The application of common-rail fuel injection systems also provides availability of

multiple fuel injection design, which has become a fuel injection strategy. A large

number of researchers have found that multiple fuel injections can reduce engine NOx

emissions [125-129]. However, some authors observed increased hydrocarbon emissions

and slight decreased engine thermal efficiency with multiple fuel injections [25]. In

addition, the complex combustion behavior of multiple fuel injections is not suitable for

fundamental engine combustion study, and therefore it is not a main focus in the fuel

injection strategy investigated here. Nevertheless, multiple fuel injections with post

injection, which is a fuel injection event considerably after top dead center (TDC) will be

investigated in this work and discussed in a later section.

Meanwhile, fuel injection pressure is affected by fluid properties such as density,

compressibility, and speed of sound [76, 130]. Biodiesel fueling in a diesel engine can

also affect the fuel injection pressure or/and injection duration, regardless of the type of

fuel injection system. As shown by Heywood [1], the fuel mass flow rate thought the

injection nozzle can be related to:

∆θ
m f = C D An 2 ρ f ∆P (5)
360 N

where m f is the mass of fuel delivered per cycle, C D is the injector tip nozzle discharge

coefficient, An is the nozzle flow area, ρ f is the fuel density, ∆P is the pressure

difference across the nozzle, ∆θ is the injection duration, and N is the engine speed.

Hence, for an engine running at steady state (i.e., N is constant), ρ f , ∆P and ∆θ control

30
the mass fuel flow. Steady state operating also requires a constant engine consumption,

and since biodiesel has around 13% heat value lower than petroleum diesel fuel, with

albeit a slight increase (around 3%) in fuel density [34], an increased fuel injection

volume (i.e., the value of ∆P∆θ ) is needed with biodiesel fueling [107].

The fuel injection pressure and ambient gas pressure, together with fuel properties

including density, surface tension, bulk modulus and boiling point can affect fuel spray

atomization and penetration, which consequently affect air mixing and fuel evaporation

rate [131-134]. Allen and Watts [48] stated that the Sauter mean diameter [135] of methyl

ester biodiesel, which can be used to describe the spray droplet size, varies from 5% to 40%

higher for biodiesel than petroleum diesel fuel, which was suggested to reduce the

premixed phase of combustion, causing an increase in the diffusion phase of combustion

[6]. Similar observations have also been reported by Suh et al. [136] They concluded that

the higher viscosity and surface tension of biodiesel were the reason for the its higher

droplet size, lower axial mean velocity and lower droplet number during single injection.

Meanwhile, no impact from viscosity or surface tension on spray penetration length was

found. Naber and Siebers [105, 106] experimentally measured the spray penetration

length of diesel fuel in various environments and found that it is dominantly controlled by

fuel injection pressure, fuel density and ambient gas density. These changes of spray

atomization and formation can affect fuel combustion and consequently gaseous

emissions. Although a significant number of studies attempted to investigate the

combustion character with computational fluid dynamics simulations of the fuel spray

process [137-141], the investigations have been computationally intensive and time

consuming. From this consideration, Siebers et al. [105, 106, 142-145] have

31
comprehensively developed a simple and fast mathematical model that describes the

equivalence ratio distribution in a fuel spray and the flame lift-off length based on

mathematical modeling and experimental correlations. Even without full implementation

of the complex physical and chemical process of fuel spray injection and ignition, this

model has been shown to be able to effectively describe fuel combustion and gas

emissions [18]. Hence, it will be adapted in this work in the attempt to explain the

variation of the engine NOx emissions, and a more detailed description of the model will

be presented in a later section.

In addition, given the aforementioned fuel injection behavior, the default

calibration of diesel engine fuel injection system tends to increase both fuel injection

pressure and duration when biodiesel is used. As discussed earlier, this “system response”

would also contribute to biodiesel’s emission effect. It has been shown that the increase

of fuel injection pressure leads to increased NOx emissions and decrease PM emissions

[21, 22] since the formation and duration of the fuel spray can significantly affect

combustion phasing [146]. A question still remains that for biodiesel usage in a common-

rail diesel engine, which aspect, including fuel chemistry, the increased injection pressure

and duration, is the most significant factor that account to biodiesel’s emission effect.

This question will also be addressed in this work.

32
2.5 Post injection strategy

As discussed earlier, multiple fuel injection strategy has been investigated by

many researchers as an approach to control engine exhaust gas emissions [125-129]. The

more conventional way to conduct multiple fuel injections is usually through a split fuel

injection strategy: a pilot fuel injection with a small quantity of the fuel starting at from

35º to 15º before top dead center (TDC) and a main injection with the majority quantity

of the fuel starting at around TDC [25, 136]. This split fuel injection will appear as two

stage combustion events in the apparent heat release profile. Since the fuel mass injected

at each event is lower than that from a single injection event, the local combustion

temperature and residence time at high temperature can be reduced, leading to decreased

engine NOx emissions.

At the same time, it is acknowledged that this kind of split injection: pilot + main

injection does not significantly alter the fuel combustion behavior of diesel engine,

resulting in an insignificant improvement in NOx-PM emission trade-off [25]. To

improve the emissions trade-off, researchers have attempted to take advantage of

multiple-injection capacity in new combustion methods, which generally represent

advanced compression-ignition combustion modes. Popular methods proposed for

advanced combustion modes include homogeneous-charge compression-ignition (HCCI)

and low-temperature combustion (LTC) [147], both of which rely extensively on engine

fuel injection control. For instance, HCCI combustion can be achieved by early and

multiple-stage fuel injection to create a homogenous (or close to homogeneous) fuel/air

mixture before ignition [148-152]. However, HCCI combustion has been observed to

have difficulty in extending operating ranges, short ignition delay, non-homogeneous

33
mixture formation and other problems [147]. From this consideration, some introduced a

new method used to depress NOx and soot formation by controlling the combustion

temperature [153]. It employed a very high EGR ratio to dilute the oxygen content in

engine intake air to reduce the combustion temperature. Later this approach was named

as low-temperature combustion (LTC) [154, 155]. Although LTC is formed on the

similar basis to HCCI, that the fuel/air mixture is more uniform than the conventional

diesel combustion, it still possesses moderate inhomogeneity of the whole mixture

compared to HCCI. This leads to a combustion process that is easier to control with a

flexible fuel injection strategy. In fact, LTC can be achieved by either very early injection

before TDC or late post injection after TDC [22, 147, 156]. However, the late post

injection often showed much better engine thermal efficiency than the early injection

method [157], and the higher exhaust temperature from post injection can also be utilized

to re-generate aftertreatment devices [19]. Therefore, a large number of authors have

investigated the application of post injection in diesel engines [26, 28-30, 158, 159]. Yun

and Reitz [28] tested the post injection strategy combined with a main injection in a

single-cylinder version of a 2.4L five-cylinder light-duty common-rail DI diesel engine

and found reductions of both NOx and PM emissions. Fang and Lee [159] tested a single

pulse post injection in an optically accessible single-cylinder high speed direction-

injection (HSDI) diesel engine and also found improvement in both NOx and PM

emissions.

At the same time, it has been observed that post injection can introduce increased

CO and unburned hydrocarbon (UHC) emissions [13, 25-30]. Therefore, a combination

of biodiesel fueling and post-injection strategy appears to be attractive. The increased

34
NOx emissions with biodiesel (as discussed earlier) can be reduced by using a post

injection strategy, while biodiesel’s lower CO and UHC emissions can compensate the

disadvantages of post injection strategy. Fang and Lee [160] tested biodiesel in an

optical-accessible single-cylinder high-speed direction-injection (HSDI) diesel engine

with an early pilot and post main injection strategy, and they found with biodiesel fueling

both NOx emissions and soot luminosity decreased. However, a few potential issues may

also arise. For example, unexpected lubricant dilution may be introduced [161] since the

emissions of UHC, which may be dissolved into the oil film on the cylinder wall, are

significantly increased. Fuel impingement on the cylinder wall has also been observed for

post injection, and it becomes stronger with biodiesel fueling [162]. In addition, the

complexity of the lubricant dilution challenge may increase as a multitude of

hydrocarbon species from exhausts can be observed in an engine using post injection

with diesel fuel [19]. This issue will become more severe when biodiesel fueling is used,

since biodiesel tends to concentrate in the lubricant [20, 163, 164] due to its higher

boiling point [20] than diesel fuel. High levels of oil dilution by biodiesel has a negative

impact on engine performance since it can increase mechanical friction [161], engine

wear [165], and metal corrosion [166], which may pose additional challenges for

lubricant development. However, a systematic understanding of these issues is not

available due to the lack of published studies combining both biodiesel and post injection

strategy.

Since the impact of post injection on engine performance and fuel oil dilution,

especially with biodiesel fueling, is still not well understood, the work presented here will

investigate the impact of post injection with biodiesel fueling on exhaust emissions, fuel

35
combustion and lubricating oil dilution. In addition, the hydrocarbon species in both

exhaust and used lubricant were also characterized.

2.6 Biodiesel and fuel injection pressure on PM morphology and soot


oxidative reactivity

2.6.1 PM morphology

Diesel soot agglomerates are composed of primary particles with geometries very

similar to spheres. Although both the mean diameter of the primary particles and their

standard deviation have small variations within the engine operating conditions [167],

diesel agglomerates are usually treated as uniform in size (monodisperse) [168].

Meanwhile, the soot agglomerates often show irregular shape which may be not easy to

describe with traditional terminology (Figure II – 3).

36
Figure II-3: Diesel soot agglomerate composed of spherical primary particles, taken
from Ref. [168]

The description of this kind of irregular formation presents difficulties for

making quantitative observations. Recently researchers have applied fractal geometry

[169] in the investigation of diesel engine soot agglomerates and to obtain quantitative

observations [168, 170, 171]. Generally speaking, the morphology analysis of diesel soot

agglomerates includes three aspects: 1) primary particle size 2) fractal dimension of the

agglomerates 3) particle size distribution of exhaust gases. The first two analyses can be

conducted based on transmission electronic microscopy (TEM) images of diesel

particulate matter obtained from thermophoretic sampling [168], and the particle size

distribution information can be obtained with scanning mobility particle sizer (SMPS)

[63]. The detailed description of test procedure will be provided in a later section.

37
These morphology characterizations essentially reflect diesel particle

agglomerates’ ability to adsorb unburned hydrocarbons. For instance, for agglomerates

with smaller primary particle size or lower fractal dimension, it will possess higher

surface area per unit mass, which may adsorb more hydrocarbons and possess higher

active surface area for soot oxidation [172]. It has been observed that both fuel properties

and engine operating condition have significant impacts on diesel particulate matter

morphology [73, 170, 173-176]. Lapuerta et al. [177] investigated the impact of engine

operating conditions on the size of primary particles of diesel soot agglomerates, and they

found the primary particle size decreases with increased air/fuel ratio, and increased

engine speed. No significant impact from EGR ratio was observed though. Smekens et al.

[178] investigated the impact of biodiesel fueling on primary particle size of diesel

engine soot agglomerates and found decreased primary particle size with biodiesel.

Mathis et al. [176] investigated the effect of fuel injection pressure on primary particle

size and found a reduction in primary soot particle diameter when fuel injection pressure

increases or start of injection advances. Lapuerta et al. [170] found that the decrease of

air/fuel ratio also decreases the fractal dimension of soot agglomerates, however, engine

speed did not have a significant impact on the fractal dimension. Zhu et al. [179]

investigated the fractal dimension of soot agglomerates from both a light-duty and a

heavy-duty diesel engines. They found that soot agglomerates from light-duty diesel

engine have lower fractal dimension. Neer and Koylu [174] tested particulate

morphology with a medium-duty John Deere 5059T diesel engine and found neither

engine speed nor load can significantly affect agglomerate fractal dimension. Similar

observations have also been reported by Soewono and Rogak [175]. They tested a VW

38
1.9L 4-cylinder turbocharged direct injection diesel engine and found no significant

impact on fractal dimension from engine speed or load. They also conducted experiments

with B20 blend biodiesel and found no impact on fractal dimension either. However,

Lapuerta et al. [170] found an increase in fractal dimension with increased engine load in

a 2.2L 4-cylinder direction-injection diesel engine (Nissan YD22). Other researchers

investigated the impact on particle size distribution from fuel properties and engine

operating conditions. Tsolakis [173] measured the particle size distribution with a heavy-

duty single-cylinder direction injection diesel engine and found decreased mean particle

diameter with high engine speed. Mathis et al. [176] investigated the impact of fuel

injection pressure on particle size distribution. They found the mean particle diameter

significantly decreases with high fuel injection pressure. Maricq [180] investigated the

impact of biodiesel fueling on particle size distribution and found a slight increase in the

number of smaller particles. Similar observations have also been reported in other studies.

[73, 173]

In summary, the number of existing studies on the impact of biodiesel fueling and

engine fuel injection strategy on the morphology of soot agglomerates is still limited, and

no unanimous conclusion can be summarized from previous work. Hence, the impact of

biodiesel fueling and engine fuel injection strategy on particulate morphology will be

investigated in this work.

39
2.6.2 Soot oxidative reactivity

As discussed earlier, it is widely known that biodiesel fueling can increase the

oxidative reactivity of diesel engine soot [75, 116, 117, 181]. This is due to increase of

degree of disorder of the soot [118-120], given a carbon atom at an edge site is much

more reactive than one in the basal plane of a graphene layer. As explained by Vander

Wal and Mueller [182], the increase of fuel oxygen content can increase the degree of

amorphous nanostructure with narrower fringe length distributions and broader and larger

tortuosity. Engine operating conditions can also significantly affect soot oxidative

reactivity. Braun et al. [183] tested the soot nanostructure with a two-stroke diesel engine.

They found that soot from an idling condition is less ordered in the crystalline structure.

Similar observations have also been reported by [175]. Al-Qurashi and Boehman [184]

investigated the impact of EGR ratio and observed that EGR soot is less ordered than

non-EGR soot. Yehliu [185] compared the oxidative reactivity of soot from single

injection and split injection, and found higher oxidative reactivity for soot from a single

injection strategy. Yehliu also found increased oxidative reactivity with retarded injection

timing. Although they suggested that fuel injection pressure had an impact on soot

oxidative reactivity, detailed investigation hasn’t been conducted. Zheng et al. [186]

investigated the impact of fuel injection duration on soot nanostructure, and they

observed no significant difference in the degree of disorder between different injection

durations.

It is believed that soot nanostructure depends upon its formation conditions, such

as temperature and residence time [187]. Vander Wal and Tomasek [188] observed that

soot formed at 1250ºC in a furnace appeared to be amorphous in its nanostructure, while

40
soot formed at 1650ºC shows a graphite-like nanostructure. In fact, Al-Qurashi and

Boehman [184], in the observation of higher disorder of EGR soot compared with non-

EGR soot, and Barun et al. [183], in the observation of higher disorder of idling soot,

both claimed that cylinder temperature is the key factor that determines soot crystal

structure. However, Seong and Boehman [189] thoroughly investigated the impact of

engine cylinder temperature and adiabatic flame temperature on soot oxidative reactivity

by intake oxygen enrichment. Although they found increased oxidative reactivity for soot

from lower adiabatic flame temperature combustion in some cases, this statement is not

generally valid. For some cases soot from similar adiabatic flame temperature

combustion possesses significantly difference in oxidative reactivity. Meanwhile, in

another investigation of the degree of disorder of soot collected from flame burner by the

same authors [190], they concluded that it was soot inception time that determined the

soot degree of disorder: the longer the soot inception time, the more amorphous the soot

particles. This statement is consistent with various experimental observations since soot

inception time is shorted with increasing flame temperature due to the enhanced soot

formation, and oxygenated fuel combustion has prolonged soot inception time due to

decreased soot precursor concentration [191-193]. In the present work we also believed

this is the key parameter that determines the soot oxidative reactivity, although Seong

[190] did not provide a theoretical justification. On the other hand, Hurt et al. [194]

investigated the nanostructure of primary soot particles with a thermodynamic approach

and found that during soot formation, the nanostructure has a “self-organization” feature,

i.e., it will transform itself into a more ordered state to decrease the total energy, given

that the soot precursors always possess more disordered structure than mature soot

41
particles [195, 196]. Hence, the reason that longer soot inception time corresponds to

higher degree of disorder of soot particles may be due to a decreased time for soot self-

organization, which can leave more amorphous (original state of soot precursors) soot

particles.

In summary, studies on the impact of fuel injection pressure on soot oxidative

reactivity is limited. Since higher fuel injection pressure can significantly decrease soot

emissions, it is expected that the soot inception time will increase. According to the

aforementioned theory, soot from higher fuel injection pressure should possess higher

oxidative reactivity, which is one of the hypotheses of this work.

2.7 Impact of biodiesel fueling, injection strategy and heat release


pattern on Engine efficiency

Engine thermal efficiency is the ratio between the power output and the rate of

energy consumption of fuel, which is a product of fuel consumption speed and lower

heating value. Since it is customary to use the brake power for the determination of

thermal efficiency in experimental engine studies, the efficiency obtained is really a

brake specific thermal efficiency. This parameter is more appropriate than brake specific

fuel consumption in comparing engine performance with different fuels.

With biodiesel fueling, the majority of existing literature shows no significant

difference in engine brake specific thermal efficiency [10, 76, 173, 197-201]. Graboski et

al. [67] tested a large number of methyl esters from different feedstocks and even pure

esters in a 11.1L, 254 kW engine following the transient cycle for heavy-duty engines (40

CFR Part 86 subpart N). From the results obtained, they showed that neither the oil origin,

42
nor the length of the carbon chain, nor the number of double bonds provides significant

differences in thermal efficiency. Nabi [202] investigated the impact of fuel oxygen

content on engine thermal efficiency, and found that efficiency decreases with increased

oxygen content only when oxygen content of the fuel is higher than 20%, otherwise the

impact was insignificant. A minor number of authors have found some decrease [63] or

some improvement [203] in thermal efficiency with biodiesel fueling. Armas et al. [63]

tested pure biodiesel derived from soybean oil in a 2.5L 4-cylinder turbocharged

common-rail direction injection DDC/VM Motori light-duty diesel engine at a steady

high speed and low load mode, with three start of injection timings. They found

decreased brake thermal efficiency for all injection timings. Some authors also observed

a conditional impact of biodiesel fueling on engine thermal efficiency. Bittle et al. [204]

tested palm olein biodiesel in a turbocharged common-rail direct injection medium-duty

diesel engine at different loads and speed, and only found decreased thermal efficiency at

high load, regardless of engine speed. From low to middle load, no significant differences

in thermal efficiency were found. Other than those, some authors also claimed synergies

when blending biodiesel. Labeckas and Slavingkas [205] tested a 4.75L engine under

different steady modes using 5%, 10%, 20%, 35% blends and pure rapeseed oil biodiesel.

The thermal efficiency appeared to reach a maximum for 5-10% blends. Murillo et al. [11]

conducted similar study with biodiesel from used cooking oil in a marine outboard 3-

cylinder naturally aspirated engine and found decreased thermal efficiency with blends

while highest thermal efficiency with pure biodiesel. Nevertheless, generally it is

accepted that biodiesel fueling won’t significantly affect diesel engine brake specific

thermal efficiency.

43
Engine thermal efficiency also relies on engine operating condition and control

strategy. Kowalewicz [206] showed that the engine brake specific thermal efficiency

increased with load with experiments in a single-cylinder naturally aspirated direct

injection diesel engine for both diesel and rapeseed oil methyl ester. Similar observations

have also been reported by Bittle et al. [204], and they also found that at low load the

brake thermal efficiency decreased with increased engine speed, but maintained at stable

level regardless engine speed at mid-high loads. Aoyagi et al. [207] investigated the

impact of boost pressure on engine thermal efficiency in a single-cylinder direct injection

diesel engine, and they found that increased boost pressure led to higher brake thermal

efficiency. Reddy and Ramesh [208] also investigated the impact of gas swirl on engine

thermal efficiency in a single-cylinder direct injection diesel engine at steady state, and

they observed no significant impact of gas swirl on thermal efficiency. Meanwhile, fuel

injection strategy has been observed to exert a significant impact on engine thermal

efficiency [63, 207]. Armas et al. [63] investigated engine thermal efficiency at different

single injection timings of -2º, 0 and 2º after TDC and observed decreased thermal

efficiency with retarded injection timing. Karra et al. [22] tested a John Deere 4045

HF475 4-cylinder turbocharged common-rail direction injection diesel engine using a

single injection strategy with start of injection ranging from -20º to 5º after TDC at steady

moderate speed and load. They found that the highest engine thermal efficiency was

observed when the start of injection is in the range of 10º to 5º before TDC, which led to

an apparent heat release peak at around TDC. Similar observations have also been

reported by Shudo et al. [209]. They conducted hydrogen assisted combustion in a 4-

cylinder spark ignition gas engine modified from an automobile gasoline engine (Nissan

44
CA 20S) and observed highest thermal efficiency at an apparent heat release peak close

to TDC. Fuel injection pressure can also affect engine thermal efficiency. Theoretical

development by Nagase and Fukushima [210] suggested that an increase in fuel injection

pressure can lead to higher thermal efficiency. However, Reddy and Ramesh [208] tested

various fuel injection pressure at different engine loads and observed that the maximum

engine thermal efficiency happened at moderate injection pressure, i.e., too high or too

low fuel injection pressure would have decreased thermal efficiency. In addition, thermal

efficiencies with post injection strategy have also been investigated and were often

reported to decrease from conventional engine combustion [160, 211]. Nevertheless, it

can be also improved by adjusting fuel injection strategy. In the low temperature

combustion work done by Fang et al. [212], an improvement of thermal efficiency was

observed when the starts of injection was retarded to 3º after TDC.

Although it is possible to investigate the strategy to improve engine thermal

efficiency through extensive engine tests by varying engine controls such as fuel injection

strategy, it requires tremendous effort to conduct such a test program, as well as, engine

modification for un-conventional combustion modes. Hence, theoretical development

might be a more appropriate approach in the attempt to improve engine thermal

efficiency. For example, Heywood [1] suggests that modern diesel engine pressure-

volume diagram can be represented with a “limit-pressure cycle”, as shown in Figure II-

4:

45
Figure II-4: A sketch of limit-pressure cycle

This limit-pressure cycle can be seen as a combination of Otto cycle and Diesel

cycle. It illustrates the maximum ideal thermal efficiency. From [1] one can calculate the

ideal indicated efficiency as:

1 αβ γ − 1
η f ,i = 1 − γ −1
[ ] (6)
rc αγ ( β − 1) + α − 1

where: rc = V4 / V2 is the compression ratio

γ = c p / cv is the ratio of heat capacity

and: α = p3 / p2 , β = V3b / V3a

Although the ratio of heat capacity γ is a weak function of temperature, the scale

of its variation is insignificant for this specific investigation. With an assumption of

46
constant γ value of 1.35 one can prepare an indicated thermal efficiency map as shown in

Figure II-5:

0.65
0.6
0.55
0.5
ηf,i

0.45
0.4
0.35

1
3
1.9
5 1.7
7 1.5
1.3
9 1.1
α
β

Figure II-5: Indicated efficiency map with α and β

It appears from Figure II-5 that β has a more significant impact on engine

efficiency than α. It can also be observed that: 1) increase of α can improve thermal

efficiency, although its effect is less obvious, 2) decrease of β close to 1 can increase

thermal efficiency significantly. Based on the above observations, an intuitive conclusion

can be drawn: to improve engine thermal efficiency, one should tend to reduce the

difference between V3a and V3b , i.e., adjust the limit-pressure cycle to Otto cycle.

This is actually suggesting that a maximum engine thermal efficiency may be

observed if the combustion heat release rate is a shape with its peak located at TDC, and

47
this understanding of location appears to be consistent with the highest thermal efficiency

locations of several experimental studies [22, 209]. However, engine energy conversion

is a much more complex process including combustion efficiency, heat loss, friction and

leakage. In fact, Hoppie et al. [32] showed that neither the Otto nor the Diesel cycle is

optimum for realistic engines. Meanwhile, the majority of existing theoretical researches

[210, 213-216] focused on parametric studies in energy efficiency. Therefore, in this

work an investigation of optimized heat release pattern design to improve engine thermal

efficiency is included, the conclusion of which could be used to direct the design of both

conventional and advanced combustion modes.

48
Chapter III. Technical Approach

3.1 Experiment apparatus

Engine tests were performed with an 8-cylinder 6.4L Ford “Powerstroke”

turbocharged direct injection diesel engine. The engine was equipped with two variable

geometry turbochargers and a common-rail fuel injection system (180 MPa maximum

injection pressure, 5 maximum injections within one combustion cycle). The injection

strategy was controlled through an electronic interface. The engine has a maximum brake

power of 261 kW at 3000 rpm, and a peak torque of 881 Nm at 2000 rpm. The

compression ratio is 17.2. The original engine calibration complied with EPA Tier 2

emission standards. Table III-1 summarizes the engine specifications:

Table III-1: Engine specifications

item Specifications
Engine type Direction injection 8-cylinder diesel engine
Bore 9.82 mm
Stroke 10.50 mm
Displacement 795.25 × 8 cm3
Compression ratio 17.20
Vc 49.09 cm3
Connect rod length 17.60 cm
Crank radius 5.25 cm
Injection system Common-rail, 180MPa maximum injection
pressure, 5 maximum injection pulse per cycle
Air charging Variable geometry turbocharger
Peak torque 880 N.m at 3000 rpm

49
3.2 Engine emission bench:

Gaseous emissions: The exhaust gases went through a heated filter and a heated

sample line, with both held at 190 ˚C. Exhaust gases contents were measured with an

AVL CEB-II emissions bench: NOx emissions were measured using an EcoPhysics

chemiluminescence analyzer, unburn hydrocarbons were measured using ABB flame

ionization detector and CO was measured by Rosemount infrared analyzer. For each

condition, the engine was operated for fifteen minutes to stabilize at the operating

condition. Data were recorded every fifteen seconds for a period of 20 minutes. The

measurements were performed three times for each condition to assess repeatability and

experimental uncertainty.

Particulate matter (PM) emissions: PM emissions were sampled through a

Sierra Instruments BG-3 micro-dilution tunnel using a dilution ratio of 10:1 at a total

flow rate of 75 slpm and a sample flow rate of 7.5 slpm for 5 min. The PM samples were

collect on Pallflex 47mm filters. The filters were weighed on a Sartorius M5P electronic

microbalance before and after sampling. The scale was located in a constant relative

humidity (45%) chamber. The filters were placed in the chamber 48 hour prior to mass

analysis before and after sampling. Three sample filters were taken per mode. The

average and standard deviation of the three samples was used to represent the result error.

Scanning mobility particle sizer (SMPS): A TSI 3936 SMPS was used to

analyze the size distribution of the PM. The SMPS instrument included a TSI series 3080

electrostatic classifier with a differential mobility analyzer (DMA), a series 3776

condensation particle counter (CPC), and a TSI series 3065 thermal denuder. A PC

running Aerosol Instrument Manager Software collected and managed the sampled data.

50
SMPS sampled diluted exhaust gas after BG-3 dilution tunnel, and it had an internal

dilution ratio of 10:1. The final dilution of the SMPS is 100:1 compared with raw exhaust

gas. Three samplings were conducted for each mode. The reported result is the average

value of three samples. In addition, a thermodenuder was used to remove the

hydrocarbon in the exhaust gases.

3.3 Test fuel

The baseline fuel for this test is ultra-low sulfur diesel fuel (ULSD). Tasks 1 and 2

used B40 blends (v/v %) using soybean methyl ester (SME)-based biodiesel obtained

from PeterCremer, L.P. Task 3 and 4 used B20 blends (20% v/v). Figure III-1 shows the

fatty acid methyl ester (FAME) composition of the SME (as neat biodiesel, B100)

measured by gas chromatography – mass spectrometry (GC-MS). The B100 used in the

experiment primarily consisted of methyl linoleate (C18:2), methyl oleate (C18:1),

methyl stearate (C18:0) and methyl palmitate (C16:0), the molecular structures of which

are shown in Figure III-2.

51
Figure III-1: FAME composition of B100 soybean oil methyl ester

52
Figure III-2: Molecular structures[217], names, abbreviations, chemical formulas,
and Chemical Abstracts Service (CAS) numbers of compounds comprising the
biodiesel used in this work.

3.4 Engine test conditions

Tasks 1 and 2: To simplify the problem of exploring the relationship between

injection strategy, biodiesel fueling and NOx emissions, a single fuel injection strategy

was used in Tasks 1 and 2. The engine was operated at the speed of 1500 rpm at 25%, 50%

and 75% of the maximum torque. The same injection timings were used for both test

fuels, and a total of seven injection timings were studied. At each start of injection, three

fuel injection pressures and injection durations were tested with each fuel. Table III-2

53
summarizes the engine operating parameters for Tasks 1 and 2. The fuel injection

pressures were chosen based on: 20% lower than default, default and 20% higher than

default when the engine was running on ultra low sulfur diesel (ULSD). The injection

duration was automatically adjusted by the electronic control unit (ECU) given that the

load and fuel injection pressure were both fixed, i.e., a higher fuel injection pressure led

to shortened injection duration. When B40 is used, both the injection pressure and

duration are increased by default because of the higher rate of fuel consumption to

achieve the same power.

Table III-2: Engine operating parameters in Task 1 and 2

Parameter Value
Speed (RPM) 1500
EGR ratio (%) 10.6 ± 2.5
Number of Injection 1
Boost pressure (Bar) 1.947 ± 0.1
Load (N.m) 220 475 660
Start of Injection (˚ATDC) -9,-5,-1,3 -9,-7,-5,-3,-1,1,3 -9,-5,-1,3
52 72 87
Injection pressure, with ULSD (MPa) 66 90 108
80 108 131
52 72, 80±1 87
Injection pressure, with B40 (MPa) 66 90, 98±1 108
80 108, 121±1 131

Task 3: one 100-hour test was performed with the same engine. During the test

period, the engine was operated at the condition of 1500 rpm with 270 N.m torque (30 %

load). A stable EGR level around 10% was used during this test. The engine started with

fresh lubricant, which is obtained from Infineum USA L.P. The engine operation was

divided into multiple time blocks ranging from 3 to 8 hours to accumulate the 100 hour

test duration. At the end of each test block 50 ml oil sample was taken. The engine

control setting for the post injection was adjusted during the test, as shown in Table III-3.

54
The default engine control only included a pilot and a main injection. The introduction of

post injection brought the total number of injections to three, and the injection quantities

and timings of the pilot and main injections were controlled by the engine’s default

calibration. The timing of the post injection ranged from 45º to 70º after top dead center

(TDC) to cover the zone at which both late combustion and lubricant dilution were

observed. Engine states, e.g., exhaust temperature and brake specific fuel consumption

(BSFC), were recorded every two hours, and the data were averaged within one test block.

An extra post injection experiment at 45º after top dead center (ATDC) with ULSD was

conducted separately to provide a comparison to the biodiesel fuel.

Table III-3: Engine operating parameters in Task 3

Engine Speed 1500 rpm


Engine Torque 270 N.m (30% load)
0h – 32h 45º ATDC at 5mg/stk
32h – 48h 60º ATDC at 5mg/stk
Post Injection
48h – 56h 60º ATDC at 8mg/stk
Configurations
56h – 75h 70º ATDC at 5 mg /stk
75h – 100h NONE

Task 4: The same engine tested in Tasks 1 – 3 will also be used in Task 4.

Similar to Task 3, B20 soybean oil based biodiesel and ULSD blends will also be used in

this task. Table III-4 summarizes the operating conditions. Two load conditions were

tested. Single injection at 1º before top dead center was used (the injection timing was

chosen based on default engine calibration). Three fuel injection pressures for each load

were tested.

55
Table III-4: Engine operating parameters in Task 4

Engine Speed 1500 rpm


Engine Torque 270 N.m (30% load) and 530 N.m (60%
load)
Start of Injection 1º before top dead center, single injection
Fuel injection pressure 50, 75 and 100 MPa for 30% load, 75, 100
and 125 MPa for 60% load

A vacuum pump was used to draw exhaust gas through a teflo filter for soot

collection. Before sampling, the engine was maintained at steady state for 15 minutes.

The time for each collection was 45 min. Soot was gently scratched off the filter after the

collection.

Diesel particulate were collected onto transmission electronic microscopy (TEM)

grids with a thermophoretic sampling apparatus, as shown in Figure III-3. The TEM grid

was mounted at the tip of the apparatus, which will be forced into raw exhaust gases

during engine running. The sampling time is around 0.6 second, and the grid was

controlled to be parallel to the flow direction of prevent direct particle compact.

56
Figure III-3: Thermophoretic sampling apparatus

Task 5: This task was conducted with a zero-dimensional thermodynamic

simulation with Matlab. A detailed description will be presented in a later chapter.

3.5 Biodiesel lubricating oil dilution

Biodiesel oil dilution was measured with a Fourier transform infrared

spectrometer (FTIR). It was conducted with a Nicolet Avatar 360 FTIR system. First of

all, a background (or air) spectrum was obtained. 1 ml sample of used oil was then placed

on the device crystal and the obtained spectrum was subtracted with the background

spectrum. The scanning region is from 500cm-1 to 4000 cm-1. The raw spectrum was then

mathematically converted to the absorption spectrum. FTIR can measure the level of

biodiesel dilution in lubricant and identify the functional groups in lubricating oil.

57
Figure III-4 shows the FTIR spectrums of engine oil with different artificially

biodiesel content. It can be observed that with higher biodiesel dilution, the intensities of

peaks at around 1746 cm-1 and 1180 cm-1 tend to increase. This observation can be used

to identify and quantitatively measure the biodiesel dilution of any used oil.

Figure III-4: FTIR spectrums for oils with different biodiesel dilution.

The authors artificially mixed 3%, 5%, 8% and 10% v/v diluted lubricant

standards with B100 and measured their IR absorbance. A linear correlation between the

absorbance of the peak at 1746 cm-1 and dilution level was observed, as shown in Figure

III-5, Equation (7) was regressed from the correlation and used to calculated the oil

dilution level.

Dilution = 33.161 × I1746 − 0.192 (7)

58
10
Y = 33.161*x-0.192
R-Square = 0.9977
8
Dilution Level (%)

0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35


Absorbance of Peak 1746 cm-1

Figure III-5: Correlation between absorbance of Peak 1746 cm-1 from FTIR and
dilution level

Total acid number (TAN) of the oil sample was measured according to ASTM

standard method D974 based on a color indication titration method. In addition, the used

lubricant after 100-hour test was distilled at 432 ˚C (which was the highest boiling point

of the biodiesel that was used, determined by simulated distillation method). The lighter

cut was collected and analyzed with GC-MS. Hydrocarbons from the engine exhaust

were collected in chilled dichloromethane and also analyzed with GC-MS. A detailed

description of the collection method can be found in Ref.[218]

Mechanical friction mean effective pressure (mfMEP) during engine operation

was calculated with Equation (8), which was based on the suggestion by Heywood [1].

mfMEP = IMEPg − BMEP − PMEP (8)

59
where IMEPg is the gross indicated mean effective pressure, BMEP is brake mean

effective pressure and PEMP is the pumping loss.

3.6 Soot characterization

3.6.1 Removal of organic fraction

Since raw soot particles collected from engine exhaust often adsorb certain

amount of hydrocarbon on the surface, which may affect instrument, a thermal treatment

procedure was employed before the analysis of soot particle. The raw particulates were

placed in a Perkin Elmer TGA 7 thermogravimetric analysis under nitrogen environment.

The temperature was ramped at a speed of 5ºC/min to 450º, and maintained at isothermal

condition for 30 min. The residual soot after this thermal treatment can be used to

conduct other experiments.

3.6.2 Thermogravimetric analysis (TGA)

An SDT Q600 from TA Instruments thermogravimetric analyzer was employed to

evaluate soot oxidative reactivity. Each diesel raw PM sample was placed in a 90 µL

alumina sample crucible and was heated at a rate of 10ºC/min to 500ºC and kept for 60

min to drive off volatile compounds under N2 gas in a flow rate of 100 cm3/min. After

30º min isothermal treatment, the sample was heated at a rate of 5ºC/min to 550ºC, and

then N2 gas was switched to ultra zero air (99.8% purity) at a flow rate of 100 cm3/min

for isothermal experiments. The mass loss of each sample during isothermal oxidation

was normalized with respect to the weight after thermal treatment, and the normalized

mass loss was compared for each sample to evaluate soot oxidative reactivity. Two

60
repeated tests were conducted for each sample, and the reported result was an average of

the two tests.

3.6.3 Transmission Electron Microscopy (TEM)

The grids collected from thermophoretic sampling were loaded onto a single-tilt

holder and analyzed in the TEM mode using a JEOL EM-2010F with field emission

source at 200 kV accelerating voltage. All images were taken in a Gatan

DigitalMicrograph with a slow scan CCD camera using 150 µm of condenser aperture

size.

3.6.4 Morphology analysis

The procedure for morphology analysis follows the work by Lapuerta et al. [168],

where a more detailed description can be found. Given this analysis requires a large

amount of sample to be statistically significant, TEM samples were examined to produce

more than 25 images at random locations on the samples to characterize many soot

agglomerates. Furthermore, 10 of the total images for each sample, which tend to have

clear structure of primary particles, were chosen. For each chosen images, 10 of random

primary particles were chosen to measure the size. A totally 100 primary particles were

measured, and the average and derivation were calculated.

The fractal dimension analysis tested all of the images obtained. These images

require a de-background procedure first. After removal of the background, with input on

primary particle size combined with an iteration procedure, the fractal dimension of each

sample can be obtained.

61
3.6.5 Raman Spectroscopy

A confocal Raman spectrometer (WITec CRM200) was used to determine the

degree of disorder of the soot samples. The excitation laser was an Argon ion laser

(λ=514.5nm), focused onto the sample using a microscope objective lens (100x or 40x).

The scattered light was collected by the same objective lens collected in a backscattering

configuration (180 degree). The microscope is equipped with a XY stage, driven by

piezo-electric actuators, for selecting the sample area of interest. The spectrum collection

method is configured by setting the integration time (seconds), software accumulations,

and hardware accumulations in the software WITec Project. All spectra were collected at

the same source power level, and post-processed via a software package (Igor Pro 6.22,

Wavemetrics Inc.). The software was used for the determination of spectral parameters.

The Raman spectra were first corrected by a linear baseline. Then, the corrected spectra

were curve fitted by three Lorentzian-shaped bands (1580, 1350, 1200 cm-1) and one

Gaussian-shaped band (1500 cm-1) based on the approach by Seong and Boehman [189,

190], as shown in Figure III-6. The ratio of AD1/AG and D1 FHWM can represent the

density of edge sites [219] and the distribution of crystalline sizes [220], respectively.

According to an early research [190], those two values correlate with the soot oxidative

reactivity, hence they will be employed for the investigation in this work.

A total of five locations for each sample were investigated. The reported results

were averaged data from the four locations.

62
Figure III-6: Demonstration of curve fitting for Raman spectrum

63
Chapter IV. Investigation of injection strategy and biodiesel
fueling on diesel engine emissions and performance

4.1 Impact of injection timing and fuel injection pressure on NOx


emissions

Figure IV-1 presents the NOx emissions at 25% load for ULSD and B40 both

with different start of injection timings and fuel injection pressures. Consistent with

previous research [21, 22, 55], at the same SOI the NOx emissions for both B40 and

ULSD increase with fuel injection pressure. Meanwhile, at the same fuel injection

pressure the advance of SOI can significantly increase NOx emissions, which is also

consistent with previous studies [23]. At the same SOI and fuel injection pressure, B40

has significantly higher NOx emissions than ULSD, which is also expected [10, 61]. One

can observe that the difference in NOx emissions due to the change of fuel injection

pressure tends to decrease with retarded SOI, and at some SOIs (e.g., 9˚ and 5˚ BTDC)

the NOx emissions for B40 at low fuel injection pressure can be lower than that of ULSD

at higher fuel injection pressure.

64
Figure IV-1: NOx emissions at 25% load. Four injection timings (9˚, 5˚, 1˚ before
top dead center and 3˚ after top dead center) were tested for both ultra low sulfur
diesel (ULSD) and B40 blend. Three fuel injection pressures (52, 60 and 80 MPa)
were tested at each injection timing

Figure IV-2 presents the NOx emissions at 50% load. To increase statistical

significance, more SOIs were tested in the range of interest (9˚ BTDC to 3˚ ATDC).

Similar to 25% load, at the same SOI, higher fuel injection pressure leads to increased

NOx emissions, and at the same fuel injection pressure, advancing SOI leads to increased

NOx emissions. B40 significantly increases NOx emissions compared with ULSD with

the same injection strategy. The additional fuel injection pressures considered for B40

were for the purpose of comparing the effects of fuel injection pressure and injection

duration. For instance, to compare the NOx emissions for ULSD at 90 MPa fuel injection

pressure, both 90 and 98 MPa fuel injection pressure for B40 are tested: matched fuel

injection pressure can lead to prolonged injection duration, and matched injection

duration requires increased fuel injection pressure, which is due to the increased fuel
65
consumption with biodiesel. It can be seen that matched injection duration increases NOx

emissions more than matched fuel injection pressure, which suggests that fuel injection

pressure plays a more significant role in the impact on NOx emissions from injection

strategy. The difference of NOx emissions from the change of fuel injection pressure also

tends to decrease with retarded SOI, which is consistent with 25% load. Furthermore,

lower NOx emissions for B40 than ULSD can also be observed for some cases at the

same SOI due to the difference in fuel injection pressure, e.g., at the SOI of 9˚ and 7˚

BTDC, NOx emission of B40 at 72 MPa fuel injection pressure is lower than that of

ULSD at 108 MPa fuel injection pressure.

Figure IV-2: NOx emissions at 50% load. Four injection timings (9˚, 7˚, 5˚, 3˚ 1˚
before top dead center and 1˚, 3˚ after top dead center) were tested for both ultra
low sulfur diesel (ULSD) and B40 blend. Three fuel injection pressures (72, 90 and
108 MPa) were tested at each injection timing for ULSD, and six fuel injection
pressures (72, 80, 90, 98, 108 and 118 MPa) were tested at each injection timing for
B40.

66
Figure IV-3 presents the NOx emissions at 75% load. The impacts of SOI, fuel

injection pressure and biodiesel on NOx emissions are consistent with those at 25% and

50% loads. The difference in NOx emission due to changes of fuel injection pressure

decreases with retarded SOI, and at certain SOI the NOx emissions of B40 is also lower

than ULSD as the fuel injection pressure decreases. For instance, at the SOI of 9˚ and 5˚

BTDC, NOx emission of B40 at 87 MPa fuel injection pressure is lower than that of

ULSD at 131 MPa fuel injection pressure.

Figure IV-3: NOx emissions at 75% load. Four injection timings (9˚, 5˚, 1˚ before
top dead center and 3˚ after top dead center) were tested for both ultra low sulfur
diesel (ULSD) and B40 blend. Three fuel injection pressures (87, 109 and 131 MPa)
were tested at each injection timing.

Overall, the impact of SOI, fuel injection pressure and biodiesel on NOx

emissions are consistent regardless of engine load. The general observations include: 1)

for both ULSD and B40, NOx emissions increase with advanced SOI; 2) NOx emissions
67
increase with increased fuel injection pressure; 3) B40 has higher NOx emissions than

ULSD at the same SOI and fuel injection pressure; 4) the difference in NOx emissions

due to fuel injection pressure becomes smaller with retarded SOI; 5) at advanced SOI (e.g.

before 3˚ BTDC) it is possible to reduce the NOx emission of B40 below ULSD by

decreasing the fuel injection pressure.

4.2 Impact of injection timing and fuel injection pressure on


particulate matter (PM) emissions

The level of particulate matter emissions is also an important concern for diesel

engines. Figure IV-4 presents the total PM emissions at 25% load for ULSD and B40,

both at different start of injection timings and fuel injection pressures. The general trend

is that the PM emissions increase with retarded SOI or decreased fuel injection pressure,

which is consistent with previous research [21] and consistent with the trend for NOx

emissions (i.e. PM-NOx trade-off [15]). At the same SOI and fuel injection pressure, the

PM emissions for B40 are lower than ULSD, which is also expected [6, 72, 201]. An

interesting observation is that the reduction of PM emissions with biodiesel at lower fuel

injection pressure is more significant than that at higher fuel injection pressure, and the

reduction of PM emissions at retarded SOI is also more significant than that at advanced

SOI. This is surprising given that such differences have not been observed for NOx

emissions, but it may suggest that the reduction of PM emissions by biodiesel is more

significant during slower combustion (a low and wide peak in the apparent heat release).

One may note the unusual decrease of PM emissions when retarding SOI from 1˚ BTDC

68
to 3˚ ATDC at a fuel injection pressure of 52 MPa. This could be due to a difference in

the soluble organic fraction (SOF) [15].

Figure IV-4: PM emissions at 25% load. Four injection timings (9˚, 5˚, 1˚ before top
dead center and 3˚ after top dead center) were tested for both ultra low sulfur diesel
(ULSD) and B40 blend. Three fuel injection pressures (52, 60 and 80 MPa) were
tested at each injection timing.

An interesting observation is that the retarding of SOI timing from 1 ºBTDC to 3º

ATDC slightly decreased the PM emissions for both ULSD and B40 at the lowest fuel

injection pressure. Since the PM emissions come from a combined adverse effects of PM

formation and oxidation, and a retarding of SOI timing could increase the exhaust

temperature and consequently promote PM oxidation while increasing PM formation.

These combined competing effects may be the reason of slight decrease of PM effect at

retarded SOI timing (3 ºATDC to 1 ºBTDC).

Figures IV-5 and IV-6 show the PM emissions at 50% and 75% load,

respectively. A consistent observation with 25% load is that the PM emissions increase
69
with retarded SOI or decreased fuel injection pressure. However, contrary to 25% load,

no significant difference in PM emissions can be observed between B40 and ULSD. At

75% load with fuel injection pressures of 109 and 131 MPa, the PM emissions for B40

are even slightly higher than ULSD. Similar observations have also been reported.

Lapuerta et al.[79] showed that at low load biodiesel could introduce a larger PM

decrease. In their review on rapeseed oil biodiesel, Krahl et al.[82] noted that the PM

emissions reductions were lower (or no significant reductions) in heavy-duty engines

than in light-duty engines. This may also suggest that the reduction of PM emissions by

biodiesel is more significant during slower combustion. The slightly higher PM emissions

of B40 at 75% load may be due to the increase of SOF with biodiesel,[6, 69, 74, 77]

which is probably caused by the lower volatility of the unburned hydrocarbons. It also

appears that the impact of SOI or the fuel injection process on PM emission becomes less

significant with increased load. For instance, the impact of SOI and fuel injection

pressure on PM emissions for B40 is less significant and consistent at 75% load (Figure

IV-6).

70
Figure IV-5: PM emissions at 50% load. Four injection timings (9˚, 5˚, 1˚ before top
dead center and 3˚ after top dead center) were tested for both ultra low sulfur diesel
(ULSD) and B40 blend. Three fuel injection pressures (72, 90 and 108 MPa) were
tested at each injection timing.

71
Figure IV-6: PM emissions at 75% load. Four injection timings (9˚, 5˚, 1˚ before top
dead center and 3˚ after top dead center) were tested for both ultra low sulfur diesel
(ULSD) and B40 blend. Three fuel injection pressures (87, 109 and 131 MPa) were
tested at each injection timing.

4.3 NOx-PM trade-off

An important evaluation of diesel engine emissions is obtained by comparing the

NOx-PM emissions trade-off. If the trade-off is not improved, a reduction of one

emission always leads to an increase of the other emission. Figure IV-7 presents the

NOx-PM emissions trade-off at 25% load for ULSD and B40 both at different start of

injection timings and fuel injection pressures. It can be seen that at this load, the NOx and

PM emissions generally fall into the same trade-off curve, regardless of fuel, start of

injection or fuel injection pressure. The trade-off curve shows a hyperbolic behavior, i.e.,

at high PM (NOx) emissions points, a significant change in the PM (NOx) emissions by

72
controlling the fuel injection pressure only has an insignificant impact on the NOx (PM)

emissions. The lowest trade-off curve location (closest to the origin) is that at the SOI of

3˚ ATDC: 1) for B40 the best fuel injection pressure is moderate, or 66 MPa in this case;

2) for ULSD the best fuel injection pressure is moderate to high, or 66 – 80 MPa in this

case.

Figure IV-7: NOx-PM emission trade-off at 25% load. The legend shows the fuel
and injection timing for each test. Three fuel injection pressures (52, 66 and 80 MPa)
were tested at each injection timing for each fuel. Fuel injection pressure is not
distinguished in the plot.

Figure IV-8 presents the NOx-PM emissions trade-off at 50% load. One can

observe that the hyperbolic feature of the trade-off curve is weakened, i.e., the test points

become more scattered. Figure IV-9 presents the NOx-PM emissions trade-off at 75%

load. The scattering becomes more significant, and the trade-off seems able to be

improved with engine injection control. For instance, at 50% load ULSD at 1˚ BTDC SOI
73
with fuel injection pressure of 108 MPa has both lower NOx and PM emissions than at 9˚

BTDC SOI with fuel injection pressure of 90 MPa. At 75% load ULSD at 1˚ BTDC SOI

with fuel injection pressure of 90 MPa has both lower NOx and PM emissions than at 9˚

BTDC SOI.

Figure IV-8: NOx-PM emission trade-off at 50% load. The legend shows the fuel
and injection timing for each test. Three fuel injection pressures (72, 90 and 108
MPa) were tested at each injection timing for each fuel. Fuel injection pressure is
not distinguished in the plot

74
Figure IV-9: NOx-PM emission trade-off at 75% load. The legend shows the fuel
and injection timing for each test. Three fuel injection pressures (87, 109 and 131
MPa) were tested at each injection timing for each fuel. Fuel injection pressure is
not distinguished in the plot

From the previous NOx emission analysis (Figures IV-1 to IV-3), it can be seen

that decreasing the fuel injection pressure is an available method to counter biodiesel

NOx effect with the start of injection timing at 9˚ BTDC. Although this manipulation in

injection strategy may introduce an increase of PM emissions, it may then be countered

by biodiesel’s PM effect. To more clearly understand its impact, the NOx –PM trade-off

for only the SOI timing at 9˚ BTDC at different loads are summarized, as shown in

Figure IV-10. At 25% load (Figure IV-10(a)), the decrease of fuel injection pressure

does not significantly increase the PM emissions for B40, and B40 with a fuel injection

pressure of 52 MPa has the same NOx and total PM emissions as ULSD with a fuel

injection pressure of 66 MPa. At 50% load (Figure IV-10(b)), the total PM emissions for
75
B40 are generally lower than those of ULSD. Therefore, B40 with decreased fuel

injection pressure appears to be the lowest NOx-total PM emissions combination for this

condition. The generally higher total PM emissions for B40 at 75% load (Figure IV-

10(c)) may be due to the higher hydrocarbons adsorbed on soot surfaces. Hence, the

authors performed a Soxhlet process to clean the solute organic fraction (SOF) of the PM.

Figure IV-10(d) presents the NOx-soot emission (both PM filters for ULSD and B40

have been cleaned with Soxhlet) trade-off for both fuel. It can be seen that the soot

emissions for B40 is generally lower than those for ULSD, verifying a higher SOF, i.e.,

hydrocarbons adsorbed on the soot surface for B40. Similar to cases at lower loads, the

decrease of fuel injection pressure for B40 is an effective way to counter the biodiesel

NOx penalty without resulting of higher soot emissions than ULSD.

76
Figure IV-10: PM-NOx emission trade-off at the start of injection timing at 9˚ before top dead center for ultra low sulfur
diesel and B40. (a) 25% load; (b): 50% load; (c): 75% load; (d): 75% load after Soxhlet of PM filters

77
Overall, for the case at early SOI timing, decreasing the fuel injection pressure for

biodiesel is an effective way to solve the biodiesel NOx penalty while maintaining the

level of PM emissions. It provides a fast and easily applicable way to implement

biodiesel in a modern engine without a significant alteration of engine emission profile.

Although the NOx-PM trade-off at early SOI may not be the best trade-off point (Figures

IV-7 to IV-9 show the best trade-off points are around the SOI timing at 1˚ BTDC), the

significance of earlier SOI will be more apparent if the engine brake thermal efficiency is

considered, which will be discussed in the following section.

4.4 Brake thermal efficiency

The brake thermal efficiency investigated here is the percentage of fuel energy

that is converted into engine brake energy. Figure IV-11 presents the thermal efficiency

at 25% load for ULSD and B40 both at different start of injection timings and fuel

injection pressures. At the same SOI, the variation of fuel injection pressure investigated

in this work (52 to 80 MPa) does not have a significant impact on thermal efficiency for

either ULSD or B40. Although it appears that B40 has a slower brake thermal efficiency

than ULSD, the differences are in the standard deviation zone. Meanwhile, the thermal

efficiency tends to decrease with retarded SOI after 5˚ BTDC, suggesting decreased

efficiency at retarded SOI.

78
Figure IV-11: Brake thermal efficiency at 25% load for ultra low sulfur diesel (left) and B40 (right). The multiple points at
each start of injection timing are due to the multiple fuel injection pressures.

79
The same trend has also been observed for higher load. Figures IV-12 and IV-13

present the thermal efficiency of ULSD and B40 at 50% and 75% load, respectively. It

can be shown that the highest engine brake efficiency seems to be at the SOI between 9˚

to 5˚ BTDC, and the efficiency decreases with retarded SOI later than 5˚ BTDC.

Similarly, no significant difference in thermal efficiency between ULSD and B40 can be

observed, nevertheless, engine brake thermal efficiency seems to be higher for B40 at

high load. In addition, the impact on thermal efficiency from variation of SOI is more

significant at higher load (Figures IV-12 and IV-13).

80
Figure IV-12: Brake thermal efficiency at 50% load for ultra low sulfur diesel (left) and B40 (right). The multiple points at
each start of injection timing are due to the multiple fuel injection pressures.

81
Figure IV-13: Brake thermal efficiency at 75% load for ultra low sulfur diesel (left) and B40 (right). The multiple points at
each start of injection timing are due to the multiple fuel injection pressures.

82
Given no significant difference in brake thermal efficiency from the variation of

fuel injection pressure and biodiesel fueling, and higher brake thermal efficiency can only

be observed with early SOI, for a modern engine which intends to run at a high thermal

efficiency the biodiesel NOx penalty can be fixed with decreased fuel injection pressure.

4.5 Apparent heat release analysis

The apparent heat release profile can provide valuable information about

combustion phasing. Figure IV-14 presents the apparent heat release profiles for ULSD

and B40 at 25% load with different SOI and fuel injection pressures. The single peak in

apparent heat release profile suggests that the fuels are primarily consumed by premixed

combustion. The shift of heat release profile is consistent with the change of SOI. Similar

to previous observations [55], at the same SOI, variation of fuel injection pressure can

significantly alter the heat release profile, i.e., higher fuel injection pressure can lead to

earlier start of combustion (SOC) and higher, as well as, narrower heat release peak. The

earlier SOC and higher heat release peak yields longer residence times and/or higher in-

cylinder temperature, leading to an increase in NOx emissions [18]. This can be

explained by the increased fuel droplet velocity [105, 106] and decreased droplet size

[221] due to increased fuel injection pressure, which can lead to better overall mixing

between fuel and air and shortened ignition delay. Higher heat release rate can introduce

higher in-cylinder temperature, yielding increased NOx emissions. It further confirms the

significance of residence time and higher in-cylinder temperature at lower load [18, 55].

e.g., the timing of maximum cylinder temperature [14].

83
When the fuel is injected before TDC, the apparent heat release profiles are

similar with the same fuel injection pressure regardless of SOI, i.e., a similar highest heat

release rate with a noisy long tail (diffusion flame). However, significant differences can

be observed for the fuel injected after TDC. At the same fuel injection pressure, the

maximum heat release rate of fuel with an SOI after TDC is apparently higher than those

with an SOI before TDC. Furthermore, the “tail” after the main combustion event is also

smaller than those with SOI before TDC. These observations suggest that the fraction of

premixed combustion of fuel with a SOI after TDC is higher than those with the SOI

before TDC. This can be explained in that the decreasing in-cylinder temperature and

pressure due to piston motion during expansion stroke can prolong the ignition delay,

resulting in longer time for premixing.

84
Figure IV-14: Apparent heat release profile of different fuels and different fuel injection pressures at 25% load. The SOI: a)
9˚BTDC, b) 5˚BTDC, c) 1˚BTDC, d) 3˚ATDC.

85
Figure IV-14 also shows the impact of biodiesel on the heat release profile. At

the same fuel injection pressure, B40 tends to have an earlier SOC, and this advancing of

SOC is more significant with later SOI. For instance, at the SOI of 9˚ BTDC the SOC for

B40 and ULSD are similar to each other, but at the SOI of 3˚ ATDC the SOC for B40 is

earlier than for ULSD. This may be attributed to the slightly higher cetane number of

biodiesel, which will lead to a shorter ignition delay, and its impact will be more

observable at prolonged ignition delay condition, i.e., with the SOI after TDC. In addition,

the maximum heat release rate of B40 seems to be lower than that of ULSD at the same

SOI and fuel injection pressure. This decreased heat release is widely observed for

biodiesel in premixed combustion [34, 91, 222, 223], which is mostly attributed to its

lower volatility in addition to the shorter ignition delay [44]. It is also consistent with the

theory proposed by Mueller et al. [18] given the oxygen content in biodiesel can make the

mixture at autoignition zone closer to stoichiometric during ignition, which can lead to

more rapidly increased in-cylinder temperature and consequently to a more advanced

combustion event (shorter ignition delay), as well as higher NOx emissions.

Figure IV-15 presents the apparent heat release profiles of ULSD and B40 at 50%

load. Unlike at 25% load, the diffusion burn constitutes the dominate fraction of

combustion at 50% load. The heat release profiles at different SOIs are generally similar,

and the shifting of the profile is consistent with the variations in SOI as well. At the same

SOI, the impact of fuel injection pressure is also similar to 25% load, i.e., higher fuel

injection pressure can lead to higher maximum heat release rate and narrower heat release

peak (for both premixed and diffusion burns). Higher fuel injection pressure can also

advance the SOC, and the work presented here confirms it by expanding the applicable

86
range of SOI from 9˚ ~ 5˚ BTDC [55] to 9˚ BTDC ~ 3˚ ATDC. Similar to 25% load, the

heat release peak of premixed burn is higher at the SOI of 3˚ ATDC, suggesting a

prolonged ignition delay. Meanwhile, the heat release profiles for B40 and ULSD with

the same SOI and injection pressure are mostly overlapping with each other. This

suggests that biodiesel does not have a significant impact on the combustion phasing. The

same observation has been reported by Zhang and Boehman [13], where the same authors

have shown that the SOI will not be affected by the biodiesel using a common-rail

injection system, and the slight difference in cetane number between ULSD and B40 will

not be strong enough to alter the combustion phasing. In addition, from the presented

work, it is suggested that the difference in combustion phasing between B40 and ULSD

is smaller at higher load.

87
Figure IV-15: Apparent heat release profile of different fuels and different fuel injection pressures at 50% load. The SOI: a)
9˚BTDC, b) 5˚BTDC, c) 1˚BTDC, d) 3˚ATDC.

88
Figure IV-16 presents the apparent heat release profiles for ULSD and B40 and

75% load. The fraction of diffusion burn is further increased. Similar to Figures 16 and

17, the heat release profiles at different SOIs are generally similar, and the shifting of the

profile is consistent with the variations in SOI. The fraction of premixed burn is stable

with the SOI from 9˚ to 1˚ BTDC, but higher with the SOI of 3˚ ATDC. At the same SOI,

higher fuel injection pressure can lead to advanced SOC, higher maximum heat release

rate and narrower heat release peak (for both premixed and diffusion burns). In addition,

the heat release profiles for B40 and ULSD are also similar to each other with the same

SOI and fuel injection pressure, suggesting the insignificant impact of biodiesel on

combustion phasing is valid for higher load as well.

89
Figure IV-16: Apparent heat release profile of different fuels and different fuel injection pressures at 75% load. The SOI: a)
˚BTDC, b) 5˚BTDC, c) 1˚BTDC, d) 3˚ATDC

90
In summary, it is observed that both fuel injection strategy and biodiesel

properties can significantly affect diesel engine NOx emissions. Advance of start of

injection or increase in the fuel injection pressure can significantly increase NOx

emissions. Both matched fuel injection pressure and matched fuel injection duration tests

with biodiesel fueling significantly increase NOx emission, and the matched fuel

injection duration (higher fuel injection pressure) has more increased NOx emissions than

the matched fuel injection pressure case. This suggests that fuel injection pressure has a

more significant impact on the NOx emissions than the fuel injection duration with

biodiesel fueling. In addition, the impact of biodiesel’s properties is even most significant

factor since a big increase in NOx emissions can be observed between diesel and B40

with the same fuel injection pressure at the same start of injection. It is also observed that

for engine with advanced start of injection (to achieve high engine thermal efficiency),

the decrease of fuel injection pressure can be an effective method to decrease the NOx

emissions when biodiesel is used. This approach can reduce the NOx emissions of engine

fueling with biodiesel blends back to the same level of NOx emissions of engine fueling

with diesel fuel. Because biodiesel generates less particulate matter (PM) emissions,

although the decrease of fuel injection pressure may introduce increased PM emissions, it

will still be at the same level as / or even lower than the PM emissions of engine fueling

with diesel. However, this approach is more effective for engine running at conventional

high efficiency mode (fuel injection timing earlier than 5º before TDC). Therefore, the

hypothesis that biodiesel NOx emission penalty can be eliminated by fuel injection

strategy is conditionally validated.

91
Chapter V. Origin of biodiesel NOx effect

As discussed in the Introduction, the behavior of the fuel spray in the autoignition

zone may explain the origin of the NOx increase in this experiment. Siebers et al. [105,

106, 142-144] have comprehensively developed a numerical model that describes the

equivalence ratio distribution in a fuel spray at the end of the premixed combustion.

Based on their analysis, the cross-sectional average equivalence ratio in a non-reacting,

isothermal jet at any position, x, along its axis is given by:

2( A / F ) st
φ ( x) = (9)
x
1 + 16( + ) 2 − 1
x

where (A/F)st is the stoichiometric ambient-gas/fuel ratio by mass and x+ is the

penetration length scale for the jet, which is defined as:

ρ f d Ca
x+ = (10)
ρ a a tan(θ )
2

where ρf is the density of injected fuel, ρa is the density of air, d is the orifice diameter, Ca

is the area-contraction coefficient of the orifice which varies with injection pressure, a is

a constant with a value of 0.75[145], and θ/2 is the spreading half-angle of the jet. The

spreading half-angle of the jet has been shown to be a function of ρf and ρa [105]:

θ ρ ρ
tan( ) = c[( a ) 0.19 − 0.0043 f ] (11)
2 ρf ρa

Here c is a constant for a defined orifice diameter which is correlated with

injection pressure [105]. The equivalence ratio from Equation (9), however, does not

92
account for oxygen heteroatoms, if any, present in the fuel. Therefore the oxygen

equivalence ratio φΩ is used to characterize the mixture stoichiometry. The oxygen

equivalence ratio is defined as the amount of oxygen required to convert all carbon atoms

to CO2 and all hydrogen atoms to H2O, divided by the amount of total oxygen available

in both reactants:

1
2nC + nH
φΩ = 2 (12)
nO

where nC, nH and nO are the numbers of carbon, hydrogen and oxygen atoms in all the

reactants, respectively. Therefore, the oxygen equivalence ratio and equivalence ratio

have this relation:

φ
φΩ = (13)
1−φ
1−
φΩ , f

where φΩ,f is the oxygen equivalence ratio of the fuel alone. It can be shown that for

diesel fuel, φΩ,f is infinitely large, or φΩ = φ. Table V-1 summarizes the input parameters

for the calculation.

93
Table V-1: Input parameters during calculation

Parameter ULSD B40


a
Chemical Structure C14.48H26.23 C16.15H30.38O0.8
3 b
ρf (kg/m ) 843 856
Zstc 0.0099 0.0089
dd (mm) 0.111
Cae 0.91+7.5785*10-5*(Pf-720)
Cdf 0.8
g
c 0.255
ρah (kg/m3) at 40˚C 1.14
Lower heating value
42.7 40.6
(MJ/kg)i
Cetane numberj ~45 ~47
Kinematic viscosity(cSt) 2.25 3.01
a: B100 obtained from GC-MS results, the chemical formula of diesel fuel was calculated
by the suggestion of C/H = 0.552[18] and the average molecular weight of around 200
g/mol[224]; b: density; c: stoichiometric mixture fraction; d: orifice diameter; e: obtained
and correlated with injection pressure (Pf) from[105]; f: discharge coefficient; g:
spreading angle constant; h: air density; i: obtained from GREET fuel cycle analysis
model[225]; j: based on the data from[18]

The local gas density at fuel injection is calculated as:

PI V
ρa = ⋅ ρ aφ ⋅ B (14)
Pa VI

where PI is intake pressure, Pa is atmospheric pressure, ρaφ is air density at 40˚C at 1 bar.

VB is the chamber volume at bottom dead center; VI is the chamber volume at the

injection crank angle. It is assumed that at the intake the air is ideal gas and the boost

temperature is constant at 40˚C.

The radial variation of equivalence ratio can be obtained given a known centerline

axial variation based on the suggestions by various researchers [105, 142, 226] as:

r ln(0.08)
φΩ ( x, r ) = 1.3φΩ ( x) exp[( )2 ] (15)
x tan 2 (θ / 2)

94
where φΩ ( x) was calculated from Equations (9) and (13), r is the radial coordinate, and

tan(θ/2) was calculated with Equation (11). The lift-off length is calculated based on

Siebers and Higgins’s model[145, 227]

H = 7.04 *108 Tg −3.74 ρ a −0.85d 0.34U f Z st −1 (16)

where Tg is cylinder temperature at fuel injection, which was obtained by heat release

calculation, ρa is the air density from Eq.(14), d is the orifice diameter, Zst is the

stoichiometric mixture fraction for fuel, and Uf is the velocity of injected fuel based on

Bernoulli’s equation:

Pf − Pa
U f = Cv 2( ) (17)
ρf

where Cv is the ratio of discharge coefficient Cd and area-contraction coefficient Ca [105].

The results of the computation are shown as follows. Figure V-1 illustrates the

oxygen equivalence ratio variation of the fuel spray and lift-off length for diesel and B40

at different injection pressures at the SOI of 5 ºBTDC. Figure V-1 shows that the change

of injection pressure does not have a significant impact on the oxygen equivalence ratio

field. The lift-off length, however, significantly increases when the injection pressure

increases due to the increase of injected fuel velocity, which results in an oxygen

equivalence ratio that is closer to stoichiometric at the autoignition zone near the lift-off

length. This explains the NOx increase when the injection pressure was increased, which

is also consistent with the observation of higher apparent heat release rate. When B40

was used at the same injection pressure, the equivalence ratio field shrinks and the lift-off

length slightly increases, leading to an equivalence ratio that is closer to stoichiometric at

autoignition, which is consistent with the observations of Mueller et al.[18] regarding the

95
biodiesel NOx increase. Note that the oxygen equivalence ratio at AZ of diesel at 108

MPa injection pressure is close to that of B40 at 72 MPa injection pressure, which may

explain the similarity of NOx emissions level between these two conditions.

96
φΩ) field for ULSD and B40 at different injection pressures at the SOI of 5ºBTDC, mid-
Figure V-1: Oxygen equivalence ratio (φ
load. The isolines divide relevant equivalence zones. The lift-off length (LOL) for each fuel and injection pressure is
represented with a horizontal line.

97
Figure V-2 illustrates the oxygen equivalence ratio variation of the fuel spray and

lift-off length for diesel and B40 at the SOI of 7ºBTDC. Compared with the case at the

SOI of 5ºBTDC, the lift-off lengths were slightly increased and the equivalence ratio

field was slightly extended due to lower cylinder temperature and gas density at the

injection region, resulting in similar equivalence ratios at the AZ. This suggests that the

difference in residence time at high cylinder temperature, which is not included in the

present model, may be the primary reason for the NOx difference when the SOI is

changed. Nevertheless, given the same residence time, the change of oxygen equivalence

ratio at the AZ is consistent with the change of NOx emissions: both for the increase of

injection pressure and the change of fuel to B40 led the oxygen equivalence ratio at the

AZ to be closer to stoichiometric. One can also observe that diesel at 108 MPa injection

pressure has a closer to stoichiometric equivalence ratio at the autoignition zone than B40

at 72 MPa injection pressure, which is also consistent with the higher NOx emissions for

diesel at 108 MPa than for B40 at 72 MPa injection pressure. Figure V-3 illustrates the

oxygen equivalence ratio variation of the fuel spray and lift-Off length for diesel and B40

at the SOI of 9ºBTDC. The equivalence ratio field was further extended and the lift-off

lengths were further increased. The observations of the relation between equivalence ratio

at AZ and NOx emissions for 9 ºBTDC are consistent with those at the SOI of 5 and 7

ºBTDC.

98
φΩ) field for ULSD and B40 at different injection pressures at the SOI of 7ºBTDC, mid-
Figure V-2: Oxygen equivalence ratio (φ
load

99
φΩ) field for ULSD and B40 at different injection pressures at the SOI of 9ºBTDC, mid-
Figure V-3: Oxygen equivalence ratio (φ
load

100
Consequently, one can estimate the average oxygen equivalence ratio at the AZ

with Equations (9) and (16) and construct a φΩ ( H ) – BSNOx emission correlation

diagram. Figure V-4 shows the correlation between brake specific NOx emissions and

average oxygen equivalence ratio of the fuel-air mixture at the autoignition zone (location

near lift-off length) at 25% load, regardless of fuel type, i.e., both B40 and ULSD are

presented with the same symbol. The variation for the same fuel and SOI is due to the

change of fuel injection pressure, the effect of which has been thoroughly investigated

[55]. At the same SOI, linear correlations between NOx emissions and average oxygen

equivalence ratio at the autoignition zone can be observed for all of the four SOI cases.

When the mixture is less rich, i.e., the oxygen equivalence ratio of the mixture is closer to

stoichiometric, and the NOx emissions are higher. It can also be seen that with retarded

SOI, the correlated straight line shifts downward, and the slope decreases, which suggests

that the impact of oxygen equivalence ratio at the AZ is more significant on NOx

emissions with earlier SOI. The relation between the slope, Y-axis intercept (at X=0) and

SOI will be discussed later.

101
Figure V-4: The correlation between brake specific NOx emissions and average
oxygen equivalence ratio of the fuel-air mixture at lift-off length at 25% load,
regardless of fuel type. The legend: xB is x degree before TDC; xA is x degree after
TDC.

Figures V-5 and V-6 present the correlation between brake specific NOx

emissions and average oxygen equivalence ratio of the fuel-air mixture at AZ 50% and 75%

load, respectively. At the same SOI, the linear correlations are still valid regardless of

fuel type for moderate and high load. Similar to low load, the slope and Y-axis intercept

(at X=0) is larger with earlier SOI, suggesting the same trend of impact of oxygen

equivalence ratio at AZ on NOx emissions at all load conditions.

102
Figure V-5: The correlation between brake specific NOx emissions and average
oxygen equivalence ratio of the fuel-air mixture at lift-off length at 50% load,
regardless of fuel type. The legend: xB is x degree before TDC; xA is x degree after
TDC.

103
Figure V-6: The correlation between brake specific NOx emissions and average
oxygen equivalence ratio of the fuel-air mixture at lift-off length at 75% load,
regardless of fuel type. The legend: xB is x degree before TDC; xA is x degree after
TDC.

Overall, this work validates the significance of oxygen equivalence ratio of the

fuel-air mixture at AZ in a light duty diesel engine with single injection extended to a

broader range of SOI and loads. Combined with the work by Mueller et al.[18] in a heavy

duty single-cylinder diesel engine, the present work further confirms that, at the same

SOI, the most important factor determining engine exhaust NOx emissions is the ignition

event that is controlled by the oxygen equivalence ratio of the fuel-air mixture at the

autoignition zone.

To probe the impact of SOI, one can construct the plots of slope – SOI and Y-axis

intercept (at X=0) – SOI from the linear correlations in Figures V-4 to V6. Especially, if

one constructs a semi-ln plot of the slope and SOI, i.e., a ln scale of the absolute value of

104
slope in Y-axis and a linear scale of the SOI in X-axis, as in Figure V-7, extremely

strong linear correlation can be observed for all three load conditions. The extra sample

points in the moderate load case are intended to increase the accuracy of the regression.

Figure V-8 shows the semi-ln plot of the Y-axis intercept (at X=0) and SOI. Extremely

strong linear correlations can be observed for all load conditions as well.

Figure V-7: semi-ln plot of absolute value of slope vs. SOI

105
Figure V-8: semi-ln plot of Y-axis intercept (at X=0) vs. SOI

Therefore, it is possible to establish a general form of an equation to indicate the

NOx emissions with SOI and the oxygen equivalence ratio at AZ for each load. After one

describes the slope and Y-intercept with the exponential correlation of SOI, the brake

specific NOx emissions can be predicted as:

At 25% load

BSNOx = −0.9574e −0.1127 SOI φΩ + 6.7849e −0.0897 SOI (18)

At 50% load

BSNOx = −1.2252e −0.1136 SOI φΩ + 7.4588e −0.0962 SOI (19)

At 75% load

BSNOx = −1.1963e −0.1257 SOI φΩ + 7.5489e −0.0858 SOI (20)

106
where BSNOx is in the unit of g/kWh and φΩ is the average oxygen equivalence ratio of

the fuel-air mixture at AZ. Equations (18) to (20) can be used to predict engine NOx

emissions given known fuel properties and engine injection strategy. Note the

contribution of start of injection (SOI) is interesting since it essentially follows a form of

first order kinetics, which is suggesting that the role of SOI may be a time-relevant factor.

However, this is beyond the scope of this work, therefore it may be included in the future

work. Although the specific parameters in Equations (18) to (20) may be not generalized

to be applied for any diesel engine, the correlation between NOx emissions and average

oxygen equivalence ratio of the fuel-air mixture at autoignition zone should hold

regardless of engine type. In addition, this model should not be applied to engine

operating at advanced combustion mode.

In summary, at the same injection time, strong correlations can be observed

between engine NOx emissions and the average oxygen equivalence ratio at the

autoignition zone near flame lift-off length regardless fuel type. Hence, the hypothesis

that the impact on NOx emission from engine fuel injection strategy control and biodiesel

fueling can be explained by the equivalence ratio at autoignition zone near the flame lift-

off length is validated.

107
Chapter VI. Investigation of late in-cylinder injection
strategy on engine emissions and lubricant fuel dilution

6.1 Exhaust Emissions

Figure VI-1 shows the impact of post injection strategy on NOx, CO and UHC

emissions. NOx emissions were decreased by around 50% when the post injection was

introduced. Compared with the introduction of post injection, changes of NOx emissions

from varying post injection timing, however, were insignificant in this experiment.

Reasons that may contribute to the NOx reduction when post injection was used includes

a retarding of the main start of injection (SOI) and a decrease in the peak heat release

[18]. A more detailed discussion will be provided in a later section.

108
25
UHC
20
15
10
5
Brake emissions (g/kWh)

0
CO
10

0
4 NOx

3
2
1
0
NONE AD(5mg/stk) AD(5mg/stk) AD(8mg/stk) AD(5mg/stk) g/stk) ULSD
45C 60C 60C 70C (5m
45CAD

Figure VI-1: Emissions of B20 engine test with various post injection strategies.
Samples are differentiated with post injection conditions while rest parameters were
kept the same. The fuel for the non-ULSD test was B20.

Both UHC and CO emissions had similar tendency with the variation of post

injection. An increase was observed for both UHC and CO emissions when the post

injection was introduced. When retarding the timing of post injection from 45º to 60º

ATDC, both UHC and CO emissions significantly increased. The highest emissions of

both UHC and CO were observed with the post injection at 60º ATDC and 8mg fuel

injection per stroke. Further retarding the post injection timing from 60º to 70º ATDC

while setting the injection quantity back to 5mg/stk increased UHC emissions but

decreased CO emissions. This suggests that fewer hydrocarbons from the post injection

were oxidized. Compared with diesel fuel at the post injection of 45º ATDC (5mg/stk),

109
B20 decreased UHC and CO emissions but increased NOx emissions, which confirms the

impact of biodiesel fuel on emissions reported in many previous studies [6, 8-11, 13].

6.2 Exhaust Temperature and BSFC

Exhaust temperature and BSFC are sensitive indicators of engine performance.

Figure VI-2 illustrates the exhaust temperatures and BSFC during the 100-hour test.

When there was no post injection during the 75 – 100 hour period, both the exhaust

temperature and BSFC were the lowest. The highest exhaust temperature was confirmed

at the beginning of the test when the post injection was set at 45º ATDC (5mg/stk). The

BSFC also increased when the post injection was set at 45º ATDC (5mg/stk), which was

consistent with the increased exhaust temperature and UHC emissions (Figure VI-1).

110
360
400 60CAD(5mg/stk)
70CAD(5mg/stk) 350

380 340

330
360
Exhaust Temperature (C)

320

BSFC (g/kWh)
340 310

45CAD(5mg/stk) 300
320
290
300
None 280

280 270

260
260 Exhaust Temperature -left Y
60CAD(8mg/stk) 250
BSFC - Right Y
240 240
0 10 20 30 40 50 60 70 80 90 100
Time (h)

Figure VI-2: Exhaust temperature and BSFC profile of the 100-hour test. Each
point is the average of one test period. The dash lines are used to divide zones of
each post injection strategy.

While retarding the post injection from 45º to 60º ATDC without changing the

injection quantity, the exhaust temperature decreased, but the BSFC still increased, which

suggested that more fuel escaped the cylinder. This observation was also consistent with

the trend of UHC emissions (Figure VI-1). Increasing the post injection quantity to

8mg/stk slightly increased the exhaust temperature but significantly increased BSFC,

suggesting that more fuel escaped the cylinder.

Setting the post injection quantity back to 5mg/stk but retarding the injection

timing to 70º ATDC decreased both exhaust temperature and BSFC. Compared with the

60º ATDC (5mg/stk), exhaust temperature was lower but BSFC was significantly higher.

111
These trends were also consistent with UHC emissions. This suggested that retarding post

injection timing decreased the combustion efficiency of the post injected fuel, which had

a significant effect on UHC emissions.

6.3 Heat Release Analysis

Heat release analysis is critical in the understanding of engine behavior. Figure

VI-3 illustrates the pressure traces of tests with different post injection strategies at

around TDC and late crank angle (LCA), i.e., later than 40º ATDC. It is obvious that the

peak pressure for no post injection was the highest, and the peak pressure for 45º ATDC

(5mg/stk) post injection including both B20 and diesel fuel was the lowest. This decrease

in peak pressure is consistent with the fact that a portion of work was done at LCA by the

combustion of post injected fuels, and tests with 45º ATDC (5mg/stk) post injection

generated the most LCA work. This is also confirmed by the pressure traces at LCA

(Figure VI-3), because tests with post injection had higher pressures in the later stage of

the combustion cycle. Thus, to maintain a constant power the engine decreased the work

by fuel from pilot and main injections. The pressure traces of 60º ATDC (8mg/stk) had a

lower peak pressure than 60º ATDC (5mg/stk) but higher LCA pressure, indicating more

energy output from the burning of increased post injected fuel. Late crank angle pressure

of 70º ATDC (5mg/stk) was similar to the test of no post injection, which suggested little

burning of post injected fuel.

112
None
9 45CAD(5mg/stk)
60CAD(5mg/stk)
60CAD(8mg/stk)
70CAD(5mg/stk)
8 45CAD(5mg/stk) ULSD
Crank Angle (deg.)

440 448 456 464 472 480

Crank Angle (deg.)

100

90
Cylinder Pressure (bar)

80

70

60

None
45CAD(5mg/stk)
50 60CAD(5mg/stk)
60CAD(8mg/stk)
70CAD(5mg/stk)
45CAD(5mg/stk) ULSD
40
350 355 360 365 370 375 380 385 390
Crank Angle (deg.)

Figure VI-3: Pressure traces of test with different post injection strategies at both
TDC (top) and late crank angle (bottom)

113
Apparent heat release and cumulative heat release are shown in Figure VI-4. The

combustion of fuel from pilot and main injections for each test are obvious, but for the

post injected fuel, the apparent heat release can only be observed in tests with post

injection at 45º ATDC (5mg/stk). This combustion, which consumed most of the post

injected fuel, explained the increase of exhaust temperature and pressures at LCA. The

slight increase of BSFC and UHC emissions suggested that this combustion at LCA was

not as complete as which occurred around TDC. The lack of significant combustion at

LCA in the other post injection tests also explains their significant increase of UHC

emissions and BSFC.

From cumulative heat release data, apart from the evidence of combustion at LCA

of fuel injected at 45º ATDC with 5mg per stroke, the energy adsorption from liquid

vaporization at around 60º – 80º ATDC in both tests with 60º ATDC post injection can be

clearly observed. The 60º ATDC (8mg/stk) post injection absorbed more energy due to

its higher injected quantity of fuel, which may explain why the increase of UHC

emissions was higher than the increase of quantity of fuel injected (more than twofold

increase in UHC compared with 60% increase in the amount of post injected fuel). Note

that the cumulative heat release increased at later crank angle, which suggested that fuel

still reacted with oxygen and generated thermal energy, though not in the same way as

the combustion at LCA in the test of 45º ATDC (5mg/stk). In contrast, the test of 70º

ATDC(5mg/stk) didn’t show a significant increase in its cumulative heat release before

140º ATDC, which suggests less reaction of post injected fuel at LCA.

114
100

NONE
80 45CAD(5mg/stk)
60CAD(5mg/stk)
60CAD(8mg/stk)
70CAD(5mg/stk)
Apparent Heat Release (J/deg.)

45CAD(5mg/stk) ULSD
60

40

20

-20
340 360 380 400 420 440

Crank Angle (deg.)

1200

1000

800
Cumulative Heat Release (J)

600

400

NONE
45CAD(5mg/stk)
200 60CAD(5mg/stk)
60CAD(8mg/stk)
70CAD(5mg/stk)
0 45CAD(5mg/stk) ULSD

-200
364 392 420 448 476

Crank Angle (deg.)

Figure VI-4: Apparent (top) and cumulative heat releases (bottom) of tests with
different post injection strategies

115
Thus, retarding post injection timing significantly decreased the combustion

efficiency at late crank angle. Furthermore, these results suggest that between 45º – 60º

ATDC there is a post injection timing threshold that determines whether combustion at

LCA consuming the posted injected fuel can occur. Figure VI-5 illustrates the computed

bulk cylinder temperature. Apparently the 45º ATDC (5mg/stk) cases had significant

temperature increase at LCA due to the combustion event. Though the tests with 60º

ATDC post injection timing didn’t have similar temperature increase as the test with 45º

ATDC injection, they still had gradual increases in temperature at LCA compared with

no post injection. The temperature increase at 70º ATDC (5mg/stk) was not as significant

as in the other cases. Hence, the bulk cylinder temperatures suggest that the post injected

fuel from different injection timings went through different combustion paths: that the

45º ATDC (5mg/stk) had high temperature combustion, that 60º ATDC (5mg/stk), 60º

ATDC (8mg/stk) and 70º ATDC (5mg/stk) only followed low temperature combustion

mechanisms [218, 228, 229].

116
2200

NONE
45CAD(5mg/stk)
2000 45CAD(5mg/stk)
60CAD(8mg/stk)
70CAD(5mg/stk)
1800 45CAD(5mg/stk) ULSD
Cylinder Temperature (K)

1600

1400

1200

1000

800
340 360 380 400 420 440 460 480 500
Crank Angle (deg.)

Figure VI-5: Bulk cylinder temperature of tests with different post injection
strategies. The arrow points to the lowest acceptable combustion temperature.

Cylinder temperature is the most critical parameter that determines the path of

combustion at LCA [14]. The fuel injected into the cylinder experiences an atomization –

combustion process, and the time between SOC and SOI is defined as ignition delay.

Here a predictive model (Equation 21) suggested by Assanis et al. [230] was employed to

predict the ignition delay of the post injected fuel in the combustion chamber of diesel

engine.

 Ea 
τ ID = 2.4φ −0.2 p −1.02 exp  ? (21)
 RuT 

117
where τID is ignition delay in millisecond (ms), ϕ is the overall equivalence ratio,

and p and T are the mean pressure and temperature during the ignition delay interval,

respectively. The interval was also corrected with outputs from the iterative ignition

delay calculation. Ea/Ru = 2100 was taken to be, as in [230].

Since the main combustion already finished before the SOI of post injection for

each test (Figure VI-4), the overall equivalence ratio of post injected fuel can be

calculated from the exhaust oxygen concentration, as shown in Equation (22).

O O ɺ
    m post
 F  sto  F  sto
φ= = (22)
O MAF + FC
* CO2 * ρO2
F ρexh

where ϕ is the overall equivalence ratio, (O/F)sto is the stoichiometric oxidizer/fuel

ratio, mɺ post is the rate of the mass of post injected fuel, MAF is mass air flow rate, FC is

fuel consumption, ρexh is exhaust gas density, ρO2 is exhaust oxygen density and CO2 is

exhaust oxygen concentration. The calculation assumed a complete combustion of fuel

from the main injection and neglected the impact of post injected fuel, given its small

quantity, on the total mass of exhaust and the oxygen concentration. Table VI-1 presents

the equivalence ratios, mean pressures, mean temperatures and ignition delays.

Table VI-1: Equivalence ratio ϕ, mean pressure P mean temperature T and ignition
delay τ ID calculated for each post injection case.

Post injection ϕ ഥ
P (bar) ഥ (K)
T τ୍ୈ (ms)
None - - - -
45º ATDC (5mg/stk) 0.134 20.43 1555.10 0.64
60º ATDC (5mg/stk) 0.125 10.50 1328.54 1.61
60º ATDC (8mg/stk) 0.210 10.76 1337.42 1.40
70º ATDC (5mg/stk) 0.123 7.50 1205.18 2.67

118
For post injection at 45º ATDC (5mg/stk), the injection timing was 405º, and the

ignition delay was 0.64 ms, which is 5.7º for an engine operating at 1500 rpm. This

timing was consistent with its SOC at LCA (Figure VI-4). Furthermore, it also

represented the timing with a critical temperature – 1500K. According to Sjӧberg and

Dec [231], 1500K is the lowest acceptable bulk cylinder temperature that high

temperature combustion can happen before the combustion is quenched by piston

expansion. As explained by the Sjӧberg and Dec, OH, which is required for the

dominating CO oxidation reaction (Reaction I, which produces 78% of total CO2 [231]),

becomes too low to initiate the reaction below 1500K.

CO + OH ⇒ CO2 + H (23)

For the post injection at 60º ATDC, the injection timing was 420º. The ignition

delay was 1.61 ms (14.5º) for the injection quantity of 5mg/stk, and 1.40 ms (12.6º) for

the injection quantity of 8mg/stk. These timings were consistent with the start of

temperature rise in Figure VI-5. In addition, the bulk cylinder temperatures at the start of

temperature rise were only around 1300K, which explains the significant change in the

combustion behavior when the post injection was retarded to 60º ATDC. For the post

injection at 70º ATDC (5mg/stk), the injection timing was 430º. The ignition delay was

2.67 ms, or 24.0º. The cylinder temperature when the fuel was ignited was about 1150K.

This further decreased the rate of oxidation of the post injected fuel.

Table VI-2 presents the calculated results of consumed hydrocarbons from the

post injected fuel, assuming the UHC from the pilot-main injection were negligible,

based on the injection quantities and emission measurements. The combustion of 45º

ATDC (5mg/stk) post injected fuel consumed the most hydrocarbons while the

119
combustion of 70º ATDC (5mg/stk) post injected fuel consumed the least, which is

consistent with the measured UHC emissions. The rates of fuel consumption in both tests

with post injection at 60º ATDC were similar due to their similar temperature

environment, which suggested that at the post injection timing of 60º ATDC, the injection

quantity didn’t significantly affect the fuel consumption.

Table VI-2: Calculated UHC emission rate and consumed HC by combustion at late
crank angle. B20 is a fuel mixture blended with 20% v/v SME based B100 and 80%
v/v ultra low sulfur diesel.

Studied cases UHC emission HC Post Consumed HC by


Fuel
(ATDC) (g/m) injected (g/m) Combustion (g/m)
NONE B20 0.11 0 -
45º (5mg/stk) B20 1.03 15 13.97
60º (5mg/stk) B20 6.13 15 8.87
60º (8mg/stk) B20 15.47 24 8.52
70º (5mg/stk) B20 11.88 15 3.12
45º (5mg/stk) ULSD 1.64 15 13.36

6.4 Oil Dilution Level, Total Acid Number (TAN) and mfMEP

Figure VI-6 illustrates the profiles of total acid number, dilution level of lubricant

and mfMEP during the 100-hour test. It shows that no oil dilution can be observed when

the post injection was at 45º ATDC (5mg/stk). Retarding the post injection timing to 60º

ATDC immediately increased oil dilution. However, the subsequent changes of post

injection timings and quantities did not have a significant influence on the rate of dilution.

When the post injection was disabled, the dilution stopped increasing. The slight decrease

observed at the beginning of the test without post injection may be due to the fuel

vaporization from the lubricant. Using the Clausius-Clapeyron equation [232] based on

the suggestion of normal boiling point from [233] and enthalpy of biodiesel from [234],

120
one can estimate the boiling point of biodiesel to be around 550˚C at 20 bar, around

500˚C at 10 bar, and around 470˚C at 7.5 bar. Therefore, the small quantity of post

injected fuel quickly vaporized since their boiling points were much lower than the local

temperature into which they were injected, which may suggest that the primary

mechanism of biodiesel dilution is the dissolution of gas phase hydrocarbons into the

lubricant film on cylinder wall. Furthermore, given the similar cylinder pressure and

cylinder wall temperature in the various tests, the hydrocarbon concentration in the liquid

would be similar regardless of the hydrocarbon concentration in the gases, if it was

simplified as a binary system (light hydrocarbon and heavy oil) and the oil film was

saturated [235]. This explains the observation that the change of hydrocarbon

concentration didn’t affect the quantity of fuel diluted into the oil, as shown by the oil

dilution profile that the change of post injection strategy didn’t have a significant

influence on the rate of dilution.

121
8
TAN - Left Y Dilution - Right Y mfMEP - Right Y

60CAD(5mg/stk) 140
7
4
Total Acid Number (mgKOH/g)

120
6

mfMEP (KPa)
Dilution (%)
100
5
2
45CAD(5mg/stk)

4 80

None

3 0 60
70CAD(5mg/stk)
60CAD(8mg/stk)
2 40
0 10 20 30 40 50 60 70 80 90 100
Time (h)

Figure VI-6: Total acid number (TAN), dilution level and calculated mechanic
friction mean effective pressure (mfMEP) profiles of the 100-hour test.

As shown in Figure VI-6, the total acid number (TAN) showed an initial decrease

and remained stable until around 30 hours. It then increased slightly and stayed at

constant level until around 60 hours. After that, the TAN kept increasing until the end of

the test, which was consistent with the effect of the acid control additives in the lubricant.

At 0 – 30 hours, the mfMEP showed an initial increase then a decrease, which was

attributed to a break-in process [236]. After 30 hours, there was no distinct trend. This

may be due to the increase of oil sludge, deposit and particulate matter content during

engine operation. Besides that, a weak correlation between the mfMEP and oil dilution

122
level can be observed, which may represent an increase of friction due to the decrease of

oil viscosity [236] from dilution.

6.5 Analysis of Hydrocarbons from Lubricant and Exhaust

Figure VI-7 illustrates the GC-MS results of the diesel base fuel, B100, and the

lighter cut from a 432˚C distillation of fresh and used lubricant. The region of interest in

the mass spectrum includes hydrocarbons that do not originate from the lubricant. The

results of used lubricant show evidently a portion of the paraffins from diesel and the

fatty acid methyl esters (FAME) components from biodiesel. This confirms that both

diesel and biodiesel were diluted into the lubricant. Other hydrocarbons in the used

lubricant have not been detected from the GC-MS analysis, which suggests that fuel

rather than combustion by-products was the primary source of lubricant dilution.

123
Diesel

B100

100h used lubricant

New lub

0 10 20 30 40 50 60 70 80
Ret.Time (min)

14
15
16
Diesel
17
18
19
20
21 22 23
C18:1
C18:2
B100 PN02738366

C16:0
C18:0
C20:1 C20:0

100h used lubricant

30 35 40 45 50 55
Ret.Time (min)

Figure VI-7: GC-MS results of diesel, B100, lighter cut from 432˚C distillation of
fresh and used lubricant (top); Region of interest in the GC-MS results (bottom).
The numbers at the diesel zone means the number of carbon in relevant paraffin.

124
Figure VI-8 shows the GC-MS analysis of hydrocarbons collected from the

exhaust for the tests with different post injection strategies. Consistent with the UHC

emission results, the exhaust hydrocarbons in the case without post injection and the case

of 45º ATDC (5mg/stk) were hardly detectable, but significant in the other three cases.

The majority of hydrocarbon components were paraffins that evaporated from the diesel

fuel, FAMEs from biodiesel and smaller molecular by-products from combustion. These

by-products such as benzene and toluene, however, have too low boiling points to be

diluted into the lubricant. Further smaller molecules were not detected by the approach

used. This also confirms the absence of high temperature combustion for the cases of 60º

ATDC (5mg/stk), 60º ATDC (8mg/stk) and 70º ATDC (5mg/stk).

125
(19)
(1) (11) (12) (17)
(20)
(2) (13)
(4) (8) (9) (10)
(7) (14) (15)
(3) 70CAD(5mg/stk)
(5) (6) (16) (18) (21)

60CAD(8mg/stk)

60CAD(5mg/stk)

45CAD(5mg/stk))

none latecycle

0 20 40 60
Ret.Time (min)
(1): C6H14 (5): C7H8 (9): C11H24 (13): C15H32 (17): C17H34O2 (21): C19H38O2
(2): C5H8O (6): C8H18 (10): C12H26 (14): C16H34 (18): C19H40
(3): C6H6 (7): C9H20 (11): C13H28 (15): C17H36 (19): C19H34O2
(4): C7H16 (8): C10H22 (12): C14H30 (16): C18H38 (20): C19H36O2

Figure VI-8: GC-MS results of hydrocarbons collected from exhaust gases for
various post injection strategies

Additionally, only paraffins with carbon numbers larger than 14 were able to be

diluted into the lubricant. This suggests an average operating temperature of the oil film

at around 400˚C according to the boiling point of C14H30 at 10 bar [233]. Smaller

molecules would vaporize under this temperature rather dilute into the lubricant.

Compared with B100 fuel, as shown in Figure VI-9, the FAME composition of

the biodiesel diluted in the used lubricant had a significant decrease in the most

unsaturated hydrocarbon methyl linoleate but various increases in the other FAME

126
components, the highest of which was methyl stearate. This can be explained by the

higher oxidation instability of unsaturated FAME compounds and the lower boiling point

of methyl palmitate.

Original Biodiesel
60
in the 100h used lubricant 53.6

50
Comparative Mass Fraction (%)

44.2

40

30 26.7 25.8
21.0
20 17.8

10 8.2
2.8
0
C16:0 C18:0 C18:1 C18:2
Biodiesel constituent

Figure VI-9: FAME composition of B100 and of the biodiesel in used lubricant.

Overall, post injection strategy can be an effective approach to reduce NOx

emission. However, it will increase the fuel consumption as well as CO and unburned

hydrocarbon emissions. The combustion behavior and oil dilution significantly depend on

the post injection strategy, including the start of injection and quantity of fuel post

injected. With biodiesel fueling, no significant oil dilution can be observed if the post

injection timing is earlier than 45º after top dead center. The hypothesis that biodiesel

fueling can introduce lubricating oil dilution with post injection strategy is a conditional

127
valid statement: if the post injection timing is earlier than 45º after top dead center, no

lubricating oil dilution can be observed; post injection timing later than 45º after top dead

center will introduce lubricating oil dilution.

128
Chapter VII. Impact of fuel injection pressure on soot
morphology and oxidative reactivity

7.1 Morphology analysis

Figures VII-1 and VII-2 show two examples of TEM images for soot

agglomerates from diesel and B20 combustion, respectively. Apparently these

agglomerates are composed of spherical primary particles, and the appearances of

agglomerates from diesel combustion and biodiesel combustion are very similar to each

other based on subjective observation.

129
Figure VII-1: A transmission electronic microscopy (TEM) image of diesel soot at 30%
load and 50 MPa fuel injection pressure

130
Figure VII-2: A transmission electronic microscopy (TEM) image of B20 soot at 30%
load and 50 MPa fuel injection pressure

At the same time, with the fractal analysis approach, a quantitative comparison

can be conducted. Figure VII-3 presents the comparison of primary particle size between

diesel and B20 soot. The result suggests that neither engine load nor fuel injection

pressure can significantly affect the primary particle size. This insignificant impact of

engine load on primary particle size is inconsistent with previous observations by

Lapuerta et al. [177], who claimed that the higher fuel/air ratio increases the size of

primary particle due to weakened mixing. However, they also suggested that higher in-

cylinder temperature can decrease the primary particle size. Hence, a combined effect of

131
higher fuel/air raito and higher in-cylinder temperature may be the reason for the

insignificance in the change of primary particle size with engine load. The insignificant

impact of fuel injection pressure is also inconsistent with previous observation by Mathis

[176], who claimed a reduction in primary particle size with increased fuel injection

pressure. Nevertheless, it is acknowledged that the number of studies on fractal analysis

of diesel engine soot is limited, and it is also possible that the impact of engine conditions

on soot primary particle size is engine-dependent.

Figure VII – 3 also compares the primary particle sizes of agglomerates between

diesel and B20 soot. For B20 soot particles, no significant impact of fuel injection

pressure or engine load on primary particle size can be observed either. However,

inconsistent with previous observation by Smekens et al. [178], who claimed decreased

primary particle size with biodiesel fueling, the primary particle sizes of B20 soot appear

to be higher than diesel soot. Again, it is possible that the impact of biodiesel fueling on

soot primary particle size is engine-dependent, and further investigation is needed to

thoroughly understand the impact of engine conditions and biodiesel fueling on soot

primary particle size.

132
60

55
30% load, diesel
60% load, diesel
Diamter of primary particles (nm) 50 30% load, B20
45 60% load, B20

40

35

30

25

20

15

10

0
50 75 100 125
fuel injection pressure (MPa)

Figure VII-3: Primary particle sizes for both diesel and B20 soot

Figure VII-4 compares the fractal dimension between diesel and B20 soot at

different fuel injection pressures and engine load. Consistently for both diesel and B20

soot, the increase of fuel injection pressure slightly decreases the fractal dimension. In

addition, increase of engine load significantly increases the fractal dimension, which is

consistent with previous researches [170]. It is also consistent with previous research by

Soewono and Rogak [175] that biodiesel fueling does not have a significant impact on

soot fractal dimension.

133
2.8
30% load, diesel
2.6 60% load, diesel
30% load, B20
2.4
60% load, B20
Fractal dimension

2.2

2.0

1.8

1.6

1.4

1.2

1.0
50 75 100 125
fuel injection pressure (MPa)
Figure VII-4: Fractal dimensions for both diesel and B20 soot

Since the fractal dimension characterizes the structure of soot agglomerates,

which are formed through the collision of soot primary particles and smaller

agglomerates in the exhaust flow [237], a possible reason for the higher fractal dimension

at higher load and lower injection pressure could be the higher temperature in the exhaust

flow where the restructuring and collision of soot agglomerates happens. Mikhailov et al.

[238] investigated the impact of temperature on the restructuring of oil soot, and found

that at higher temperature the soot will restructure to less anisotropic structure, i.e.,

higher fractal dimension. Since biodiesel won’t significantly affect engine in-cylinder

temperature [44], no significant impact on the fractal dimension with biodiesel fueling

can be observed.

134
Figure VII-5 presents the comparison of number of primary particles from fractal

analysis. It can be observed that the number of primary particles increases with higher

engine load, but decreases with higher fuel injection pressure. This is consistent with

previous investigations [170]. In addition, no significant impact of biodiesel fueling can

be observed. Note although this analysis intends to investigate the change of number of

primary particles of exhaust gas, it was conducted based on localized TEM image. A

more accurate investigation of particle number should be conducted with SMPS, which

will be provided.

600 30% load, diesel


60% load, diesel
Number of primary particles (#)

500 30% load, B20


60% load, B20

400

300

200

100

50 75 100 125
fuel injection pressure (MPa)

Figure VII-5: Number of primary particles for both diesel and B20 soot

Figure VII – 6 presents the particle size distributions of diesel and B20 exhaust at

30% load, before the thermodenuder. These particle size distributions show as symmetric

135
peaks in a semi-log plot (linear y-axis scale and log x-axis scale). The peak at smaller

particle diameters is dominantly hydrocarbon droplets, i.e., the nucleation mode peak,

and the peak at larger particle diameters is predominantly exhaust particulate matter, i.e.,

the accumulation mode peak [63]. It can be observed that an increase of fuel injection

pressure can decrease the accumulation model peak while increasing the nucleation

model peak. This can be due to the better mixing from increased fuel injection pressure.

In addition, only slight decrease in the particle concentrations can be observed with

biodiesel fueling. These results are consistent with the trends of number of primary

particle sizes presented earlier.

8
1.0x10
Diesel 100 MPa
Diesel 75 MPa
Diesel 50 MPa
Particle concentration (#/cm3)

7
8.0x10 B20 100 MPa
B20 75 MPa
B20 50 MPa
7
6.0x10

7
4.0x10

7
2.0x10

0.0
10 100
Particle diamter (nm)

Figure VII-6: Particle size distribution of diesel and B20 combustion exhaust, 30%
load, before thermodenuder

136
Figure VII-7 presents the particle size distributions of diesel and B20 exhaust at

30% load, after the thermodenuder. The function of the thermodenuder is to remove the

hydrocarbon content in the exhaust gases. As shown in Figure VII -7, although not

completely, the nucleation mode peaks (dominantly hydrocarbon droplets) have been

significantly reduced and the accumulation mode peaks are clearer.

7
1.2x10
Diesel 100 MPa
Diesel 75 MPa
1.0x10
7 Diesel 50 MPa
Particle concentration (#/cm3)

B20 100 MPa


B20 75 MPa
6
B20 50 MPa
8.0x10

6
6.0x10

6
4.0x10

6
2.0x10

0.0
10 100
Particle diamter (nm)

Figure VII-7: Particle size distribution of diesel and B20 combustion exhaust, 30%
load, after thermodenuder

Figure VII-8 presents the particle size distributions of diesel and B20 exhaust at

60% load, before the thermodenuder. It can be observed that the nucleation mode peaks

dominated the particle size distribution. Consistent with lower load, an increase in the

nucleation mode peak can be observed with higher fuel injection pressure. Figure VII-9

shows the particle size distributions of diesel and B20 exhaust at 60% load, after

137
thermodenuder. It clearly shows that an increase of fuel injection pressure can decrease

the accumulation model peak. One can also observe that the particle number

concentration at higher engine load is lower than that at lower engine load. Intuitively

this seems to be contrary with the trend of number of primary particle presented earlier,

but it should be acknowledged that even the number concentration of particles in the

exhaust is lower for higher engine load. An increased flow rate at higher engine load can

promote an increase in the total particle number.

7
8.0x10
Diesel 125 MPa
Diesel 100 MPa
Diesel 75 MPa
Particle concentration (#/cm3)

B20 125 MPa


7
6.0x10 B20 100 MPa
B20 75 MPa

7
4.0x10

7
2.0x10

0.0
10 100
Particle diamter (nm)

Figure VII-8: Particle size distribution of diesel and B20 combustion exhaust, 60%
load, before thermodenuder

138
6
3.0x10
Diesel 125 MPa
Diesel 100 MPa
6 Diesel 75 MPa
2.5x10
Particle concentration (#/cm3)

B20 125 MPa


B20 100 MPa
6
B20 75 MPa
2.0x10

6
1.5x10

6
1.0x10

5
5.0x10

0.0
10 100
Particle diamter (nm)
Figure VII-9: Particle size distribution of diesel and B20 combustion exhaust, 60%
load, after thermodenuder

7.2 Analysis of soot oxidative reactivity

Figure VII – 10 presents the thermogravimetric analysis result of 30% load diesel

and B20 soot. It appears that the fuel injection pressure has a significant impact on the

soot oxidative reactivity for both diesel and B20 soot. The increase of fuel injection

pressure from 50 MPa to 75 MPa significantly increases the oxidative reactivity, and

further increasing the fuel injection pressure to 100 MPa increases the oxidative reactivity,

though not as significantly as the change from low to moderate fuel injection pressure.

This trend is consistent with the PM emissions investigated earlier, where the most

significant change happens when the fuel injection pressure is increased from low to

139
moderate. Figure VII-10 also confirms one of the hypotheses in this work, i.e., soot from

higher fuel injection pressure has higher oxidative reactivity.

In addition, consistent with previous investigations [75, 116, 117, 181], the

oxidative reactivity of B20 soot is higher than diesel soot, given the same fuel injection

pressure.

diesel 50 MPa
1.0
diesel 75 MPa
diesel 100 MPa
B20 50 MPa
0.8 B20 75 MPa
B20 100 MPa

0.6
m/m0

0.4

0.2

0.0
0 10 20 30 40 50 60
time (m)

Figure VII-10: Thermogravimetric analysis result of 30% load diesel and B20 soot,
m/m0 indicates the residual mass weight percentage.

Figure VII – 11 presents the thermogravimetric analysis result of 60% load diesel

and B20 soot. The general observations for the behavior of oxidative reactivity for 60%

load soot are quite similar to the 30% load: increase the fuel injection pressure can

increase soot oxidative reactivity, and at the same fuel injection pressure B20 soot has a

140
higher oxidative reactivity than diesel soot. Figure VII-11 further confirms the

hypotheses proposed in this work.

diesel 75 MPa
1.0
diesel 100 MPa
diesel 125 MPa
B20 75 MPa
0.8 B20 100 MPa
B20 125 MPa

0.6
m/m0

0.4

0.2

0.0
0 10 20 30 40 50 60
time (m)
Figure VII-11: Thermogravimetric analysis result of 60% load diesel and B20 soot,
m/m0 indicates the residual mass weight percentage.

Since the soot oxidation is a reaction between carbon and oxygen, a simple first

order kinetic model can be employed to describe the mass consumption rate [239-242]:

dm
= − kcCasCO2 ⋅ dm (24)
dt

where m is the mass of soot, kc is rate constant of carbon-oxygen reaction, CO2 is the

concentration of oxygen and Cas is the concentration of active surface sites. The active

surface site parameter is included in the model since it (characterized as active surface

area [243]) has been observed to affect the oxidation rate [184, 240, 244] due to its

141
significant role in oxygen diffusion rate [245, 246]. Given in TGA a stable flow of air

was fed into the environment, therefore the concentration of oxygen can be considered as

a constant. This reorganization reduces Equation (24) into a first order chemical kinetic

model, if one defines an apparent reactivity as:

k a = kcCas CO2 (25)

Then Equation (24) becomes:

dm
= − k a ⋅ dm (26)
dt

Re-arrange Equation (26) as:

d ln( m) d ln( m / m0 )
= = −ka (27)
dt dt

where m0 is the initial mass weight of the soot sample. Equation (27) indicates that in a

semi-log plot (linear scale in x-axis and log scale in y-axis) of time vs. m/m0, the slope of

the curve represent the apparent reactivity, in other words, the absolute value of the first

order derivative of a time vs. ln(m/m0) plot will be the reactivity. Based on this

consideration, the apparent oxidative reactivity was calculated, as shown in Figure VII -

12 (for 30% load soot) and Figure VII-13 (for 60% load soot).

142
0.08
diesel 50 MPa
diesel 75 MPa
0.07
diesel 100 MPa
B20 50 MPa
0.06 B20 75 MPa
reactivity (1/min)

B20 100 MPa

0.05

0.04

0.03

0.02

0.01

0.00
0 10 20 30 40 50 60
time (m)
Figure VII-12: Apparent oxidative reactivity of 30% load diesel and B20 soot

143
0.12
diesel 75 MPa
0.11 diesel 100 MPa
diesel 125 MPa
0.10
B20 75 MPa
0.09 B20 100 MPa
B20 125 MPa
0.08
reactivity (1/min)

0.07

0.06

0.05

0.04

0.03

0.02

0.01

0.00
0 10 20 30 40 50 60
time (m)
Figure VII-13: Apparent oxidative reactivity of 60% load diesel and B20 soot

Immediately one can observe that the apparent reactivity changes with burn-off,

which is consistent with various previous studies [242-244, 247]. Meanwhile, the change

of apparent reactivity varies for soot produced from different fuel injection pressures. The

diesel soot with the lowest fuel injection pressure appears to give an apparent reactivity

which is insignificantly affected by the burn-off. With increased fuel injection pressure,

the apparent reactivity tends to increase with burn-off. This increase in reactivity may be

due to the increase in the active surface area with burn-off, which have been observed

earlier [244, 247]. The consumption of carbon atoms during oxidation, which could

expand the pore size and consequently increase the area that can be bond with oxygen,

can contribute to the increase of active surface area [243]. It can also explain why the

apparent reactivity of soot from higher fuel injection pressure (or biodiesel fueling) has a

144
faster increase than the soot from lower fuel injection pressure (or diesel fueling) since

they possess higher oxidative reactivity, i.e., carbon consumption rate.

A very interesting observation is that for the highest fuel injection pressure, the

reactivity of soot shows an initial increase with burn-off while it decreases when the

burn-off is close to completion. This kind of reduction in reactivity at very high burn-off

has also been reported by Song et al. [116]. They concluded that the oxidation of high

reactivity soot particle follows an internal burning process, forming central hollow

structures. The residual carbon shell at high burn-off then tends to undergo a coalescence

process due to physical factors to minimize strain energy. This physical coalescence

process may behave as a decrease in active surface area, leading to a decrease in apparent

reactivity. In fact, this decrease in active surface area at high burn-off for certain soots

has also been reported by Cheng and Harriott [244].

However, in both work by Song et al. [116] and Cheng and Harriott [244], the

reduction of reactivity is less significant than the increase of reactivity at earlier burn-off.

In the work presented here, the reactivity at high burn-off was decreased to be even lower

than the initial reactivity or the reactivity of soot from lowest fuel injection pressure. To

investigate this observation, Raman results were presented. Figure VII-14 shows the

AD1
ratio with fuel injection pressure for both diesel and B20 soot, and Figure VII-15
Ag

shows the D1 FWHM values with fuel injection pressure for both diesel and B20 soot.

AD1
Although Seong [190] claimed that the ratio of and D1 FWHM can be correlated
Ag

with soot oxidative reactivity, this correlation has not been observed in the present work.

145
On the contrary, no unanimous trends from Raman analysis of the initial soot

nanostructure can be observed corresponding to the oxidative reactivity.

2.6
diesel,30%
2.5 B20, 30%
diesel, 60%
B20, 60%
2.4

2.3
AD1/Ag

2.2

2.1

2.0

1.9
50 75 100 125
fuel injection pressure (MPa)

Figure VII-14: AD1/Ag ratio with fuel injection pressure for both diesel and B20 soot

146
190
diesel,30%
B20, 30%
180
diesel, 60%
B20, 60%
D1 FWHM (cm-1)

170

160

150

140

130
50 75 100 125
fuel injection pressure (MPa)
Figure VII-15: D1 FWHM values with fuel injection pressure for both diesel and
B20 soot

The similar results from Raman analysis for soot at all fuel injection pressures

may be because the Raman spectrum essentially shows an averaged result of the whole

soot carbon structure. If the soot from high fuel injection pressure is composed of both a

highly disordered and a highly ordered portion, while soot from low fuel injection

pressure is composed more uniform and moderately ordered particles. This will make the

initial reactivity of soot from high fuel injection pressure higher, but the reactivity after

burn-off lower than soot from low fuel injection pressure. These kinds of structures may

not show significant differences in Raman analysis. The reason for this kind of structure

may be: at high fuel injection pressure there is longer soot inception time, leading to more

disordered initial structure. Meanwhile, the high in-cylinder temperature from high fuel

injection pressure promotes the self-organization process [194]. Although the time for

147
self-organization is shorter, the organized portion of carbon structure may approach to

more ordered state than that at lower temperature environment. For low fuel injection

pressure there is short soot inception time, leading to longer self-organization time, i.e.,

self-organization process dominates the final carbon structure. This can lead to a more

uniformed degree of disorder.

However, given the scope of this work, the investigation of detailed soot carbon

structure is not included. To fully verify the aforementioned hypothesis, active surface

area and soot carbon crystalline structure analysis with burn-off should be conducted.

They will be included in the recommendation for the future work.

The hypothesis that soot from high fuel injection pressure has higher oxidative

reactivity since the decreased soot precursor concentration can prolong the soot inception

time is validated.

148
Chapter VIII. Optimized heat release pattern to achieve
high engine efficiency

8.1 Model formulation

A zero-dimensional engine thermodynamic simulation was conducted in this

work. The assumptions used in the simulation included:

1. Fuel combustion efficiency is 100%, i.e., all energy of the input fuel is

converted to heat completely, and the combustion process is not included in

this model.

2. There is no heat loss due to fuel vaporization, and physical processes

including gas swirl and fuel spray formation do not contribute to any heat

exchange.

3. No radiation heat transfer in the system

4. No gas leakage in the system.

5. The gas in the combustion chamber is an ideal gas.

6. No heat exchange during the gas exchange cycle

The above assumptions show that the system created here is close to an idealized

system. Nevertheless, the interest in this work focuses on the design of an optimized heat

release pattern, for which these assumptions are sufficient. Sub-models for each

assumption can be implemented easily if more thorough investigations is needed.

The crank angle of one complete cycle is from 0º to 720º, i.e., the top dead center

is 360º. First, in one complete cycle the total indicated work is:

149
EVO
Wi = ∫ P (θ ) dV (θ ) (28)
IVC

where: P (θ ) : cylinder pressure

π DB2
V (θ ) : chamber (gas) volume, V = Vc + [ Lc + r − ( r cos θ + L2c − r 2 sin 2 θ )] ,
4

Vc is the minimum volume and Lc is the length of connection rod

θ : crank angle in radian. Conversion between radian and degree and be easily

180°
done as: degree = × radian
π

IVC and EVO: intake value close and exhaust valve open

For an engine system, the rate of total heat addition is:

dqt dW dE dqw dM e
= + + +h (29)
dθ dθ dθ dθ dθ

where: qt: total heat added

W: work done by the gas; dW = PdV

E: internal energy of gas. dE = Cv ( MdT + TdM )

M: mass of gas in the chamber. M = n ⋅ MWg , n is the gas mole in the chamber,

MWg is the gas molar mass. Since it is air, we will use 28.97 g/mol

qw: heat transfer to the wall.

h: enthalpy per unit of mass. h = C pT

Me: mass of gas in leakage

Cv and Cp are heat capacity for constant volume and constant pressure,

respectively.

One can define:

150
Cp
γ= (30)
Cv

Also for ideal gas

PV = MRuT (31)

C p = Ru + Cv (32)

where T is the temperature of gas, Ru is universal gas constant. It is also easy to see:

dM dM e
=− (33)
dθ dθ

Put Equations (30) – (32) in Equation (29), rewrite:

dqt dV dT dM dqw dM
=P + Cv M + CvT + − C pT
dθ dφ dθ dθ dθ dθ
dV dT dM dqw
=P + Cv M − RuT +
dθ dθ dθ dθ
dV d ( PV / Ru ) PV dM dqw
=P + Cv − + (34)
dθ dθ M dθ dθ
dV C dV Cv dP PV dM dqw
=P + vP + V − +
dθ R dθ R dθ M dθ dθ
γ dV 1 P PV dM dqw
= P + V − +
γ − 1 dθ γ − 1 dθ M dθ dθ

From assumption, dM = 0 . However, one can easily implement a leakage model

inside the simulation. For instance, Hoppie et al. [32] suggested a rate of leakage as a

function of cylinder pressure, temperature and size of leaking hole.

Then the rate of total heat addition becomes:

dqt γ dV 1 dP dqw
= P + V + (35)
dθ γ − 1 dθ γ − 1 dθ dθ

From Heywood [1], the apparent heat release (AHR) during firing is:

γ dV 1 dP
AHR = [ P + V ] firing (36)
γ − 1 dθ γ − 1 dθ

151
The value of γ can be estimated with a polynomial function of temperature [1]:

γ = 1 / (1 − 1 / ( a1 + a2 * T + a3 * T 2 + a4 * T 3 + a5 * T 4 )) (37)

a1 = 3.6359
where

a2 = −1.33736 * 0.001

a3 = 3.29421* 0.000001

a4 = −1.91142 * 0.000000001

a5 = 0.275462 * 0.000000000001

The total energy input is:


dqt dqt
qt = ∫ ( dθ )dθ = ∫
0
firing
(

) dθ (38)

The overall heat transfer of the engine includes both convection and radiation.

However, since we are assuming complete combustion, the radiation term can be

neglected [1]. Hence, the heat transfer within the engine can be calculated as [248]:

dqw
= qɺw = hw [ Apis (T − Tpis ) + Acyw (T − Tcyw ) + Acyh (T − Tcyh )] (39)
dt

where: hw : convective heat transfer coefficient

Apis , Tpis : area and temperature of piston top

Acyw , Tcyw : area and temperature of cylinder wall

Acyh , Tcyh : area and temperature of cylinder head

The convective heat transfer coefficient hw can be predicted with the correlation

proposed by Woschni [249]:

hw = 3.26 ⋅ DB1− m ⋅ P m ⋅ T 0.75 −1.62 m wm (40)

152
where: DB : engine bore size

w : average cylinder gas velocity

m: an exponent used for correlation. Heywood [1] suggested a value of 0.8 in the

correlation. A later investigation by Han et al. [250] suggested that m possesses a

value between 0.7 to 0.8, which varies with engine.

For a four stroke direction injection compression ignition engine the average

cylinder gas velocity can be calculated as:

Vd Tr
w = Cw1S p + Cw 2 ( P − Pm ) (41)
PV
r r

where Vd : displacement volume

Tr , Pr and Vr : temperature, pressure and chamber volume at reference point, e.g.,

intake value closing

S p : mean piston speed, S p = 2 LN , L is piston stroke, N is engine speed

For gas exchange period: Cw1 = 6.18, C w2 = 0

For the compression period: Cw1 = 6.18, Cw2 = 0

For the combustion and expansion period: Cw1 = 6.18, Cw2 = 3.24*10-3

Finally, the temperatures of engine component can be estimated as [251]:

Tx = Tref + ax ( ρair S p )0.8 + bx (1000 ⋅ Y ) (42)

where Tref : reference temperature, here coolant temperature is used

ρ air : air density in the intake manifold

mf
Y : fraction of fuel in the gas mixture, Y =
ma + m f

153
When x = piston: ax = 6.9 , bx = 2.66

= Cylinder wall: ax = 1.8 , bx = 0.7

= Cylinder head: a x = 2.1 , bx = 0.91

Up to this point, it will be able to calculate the indicated thermal efficiency as:

Wi
ηi = (43)
qt

The brake specific thermal efficiency requires calculation of friction work. The

general friction work can be calculated as:


W friction = ∫T f (θ , P ) dθ (44)
0

The friction torque of a diesel engine includes piston assembly friction and

crankcase assembly friction. Rakopoulos et al. [252] developed a comprehensive model

for friction torque in a direct injection diesel engine, which includes six components, and

the total friction torque is the summation of fraction torque of all components:

Tf = Tf 1 + Tf 2 + Tf 3 + Tf 4 + Tf 5 + Tf 6 (45)

c
T f 1 = c1 µ ⋅ S p ⋅ wo ⋅ ( P + Pe ) ⋅ ( no + 0.4nc ) ⋅ DB ⋅ r ⋅ , friction torque for ring

hydrodynamic lubrication

c
T f 2 = c2 ⋅ π ⋅ DB ⋅ rc ⋅ wc ⋅ ( P + Pe ) ⋅ (1 − sin θ ) ⋅ r ⋅ , friction torque for ring mixed

lubrication

c c
T f 3 = c3 ⋅ µ ⋅ ⋅ DB ⋅ l p ⋅ r ⋅ , friction torque for piston skirt
h rω

154
nv ⋅ N s ⋅ r c
T f 4 = c4 ⋅ ⋅ , friction for valve train
ω rω

T f 5 = c5 ⋅ µ ⋅ ω , friction for auxiliaries and unloaded bearing

π DB2 rjb ⋅ P ⋅ cosθ


T f 6 = c6 ⋅ , friction for loaded bearing
4 ω

where µ is lubricant dynamic viscosity, wo is width of oil ring, wc is width of compression

ring, r is crank case radius, Pe is elastic ring pressure force, no is number of oil rings, nc is

number of compression rings, ω is radial speed, rjb is radius of journal bearing, lp is the

length of piston skirt, h is the oil film thickness, Ns is the preloading force of the valve

spring, and c is the instantaneous piston velocity:

r
) cosθ(
Lc
c = rω{sin θ ⋅ [1 + ]}
r 2 2 0.5
(1 − ( ) sin θ )
Lc

From [252] the multiplier at each friction torque component, assuming a middle

speed of 2500 rpm for the engine worked here, are: c1 = 21.69 , c2 = 0.20 , c3 = 0.45 ,

c4 = 0.27 , c5 = 0.65 , c6 = 0.19 . Table VIII-1 summarizes the basic parameters used by

the friction model

Table VIII-1: Engine basic parameters used by friction model.

Lc 17.6 cm
wo* 3 mm
wc* 2 mm
Pe* 15050 N/m2
no 1
nc 2
lp 52.5 mm
h* 1.5 mm
µ 0.0855 kg/m.s
Ns* 200 N

155
rjb* 15 mm
*: data based on Ref. [252], but scaled to fit a lighter duty diesel engine.

Note the values from Table VIII-1 may not be accurate for the real engine used in

the earlier tasks of this work. Nevertheless, it can still provide qualitative understanding

of the engine behavior. If we assumed a negligible pumping loss (e.g., a high intake

pressure from turbocharger), then the brake specific thermal efficiency will be:

Wi − W friction
ηi = (46)
qt

8.2 Computational procedure

8.2.1 Motor cycle calculation

The motor cycle can be calculated based on Equation (35), with defining dqt = 0 ,

or:

γ dV 1 dp dqw
P+ V + =0 (47)
γ − 1 dθ γ − 1 dθ dθ

Calculation of Equation (47) follows a numerical procedure. With a known

pressure P0 at the crank angle θ0 and chamber volume V0 , the pressure P after a change

of crank angle of ∆θ has the relation:

γ 1
P ∆V + V ( P − P0 ) + ∆qw = 0 (48)
γ −1 γ −1

The term ∆qw represent the heat transfer during the period when crank angle

changes ∆θ , so

∆θ
∆qw = qɺw ∆t = qɺw (49)
ω

156
Note as mentioned earlier, ω is the angular velocity. Combining Equation (48)

with Equation (49) and re-arranging, the pressure after ∆θ is:

∆θ
PV − qɺw (γ − 1)
P=
0
ω (50)
γ (V − V0 ) + V

qɺ w can be calculated from Equation (39). In addition, one must note that γ is a

function of temperature, which will be affected by the pressure value. Therefore, one

needs to perform an iteration procedure to re-calculate γ value with the new P value,

until P became stable ( δ P < 10−7 between each iteration).

The first point of this numerical procedure ( P0,i , V0,i ) should choose the crank

angle location at which the intake value closes. The subscript i means the initial value for

the procedure. The real engine used in Tasks 1 – 4 has IVC at 200º and EVO at 520º,

which will be used in this model. The initial pressure P0,i will use the intake air pressure

measured from real engine data.

Finally, the obtained motor pressure trace needs to be synchronized with volume

to eliminate the thermodynamic lag [253], i.e., the maximum motor pressure peak timing

should be synchronized with piston top dead center. The pressure trace data obtained

from real engine in this work were synchronized by timing and clocking signal which

were fed into the data acquisition system.

To compare the model with real engine data, the pressure trace of an engine

running with diesel at 1500 rpm and 30% load, single injection timing at 1º before top

dead center were used. The pressure trace data was obtained from Task 4. At this

injection timing the start of combustion is after top dead center so that the pressure trace

157
before top dead center can be used as the motor pressure. The value of P0,i used the intake

air pressure of this condition.

The comparison of simulation and real engine pressure trace also requires

adjustment of the value m in Equation (40). The initial value of m used 0.8, and with

adjustment it was found that a best match can be obtained when m = 0.761.

Figure VIII-1 presents a comparison of simulated motor pressure trace and real

engine data with combustion existed. A very good match can be observed between the

simulation and engine pressure trace. This serves as a validation of the employed model.

100
motor pressure (simulation)
90 Engine pressure trace

80

70

60
Pressure (bar)

50

40

30

20

10

0
0 100 200 300 400 500 600 700
CA (degree)

Figure VIII-1: comparison of simulated motor pressure trace with engine data,
running at 1500 rpm, 30% load, single injection timing at 1º before top dead center,
100 MPa fuel injection pressure

158
8.2.2 Apparent heat release representation

A typical apparent heat release profile of diesel combustion includes a premixed

flame peak and a diffusion flame peak. Figure VIII-2 shows that the apparent heat

release profile can be represented with a convolution of one Gaussian peak and one

ExpConvExp peak. The engine AHR data was chosen from a condition of 1500 rpm, 30%

load and 100 MPa fuel injection pressure, which is one case in Task 4. Note there’s no

negative value before the start of combustion given it was assumed no heat exchange due

to fuel vaporization. The convolution was conducted with Igor pro.


Difference (J/degree)

8
4
0
-4

150 Engine AHR


AHR (J/degree)

Fit curve

100

50
Artificial peaks (J/degree)

0
0
120
80
40
1
0
350 360 370 380 390 400
CA (degree)

Figure VIII-2: Representation of apparent heat release (AHR) at 1500 rpm, 30%
load, 1º before top dead center injection, 100 MPa fuel injection pressure with a
convolution of two peaks: one Gaussian peak and one ExpConvExp peak. Top:

159
difference between the engine data and fit curve; middle: comparison of engine data
and fit curve; bottom: deconvolution of artificial peaks.

Figure VIII-3 presents another example of the peak convolution. The engine

apparent heat release profile was obtained from the operating conditions: 1500 rpm, 60%

load, single injection timing at 1º before top dead center, 100 MPa fuel injection pressure.

Although the match is not as good as lower load, the convoluted peak can still represent

the major characteristic of the real heat release. Nevertheless, in the work presented here

we will only investigate the 30% load case.


Difference (J/degree)

10
0
-10

140
120 Engine AHR
AHR (J/degree)

100 Fit curve


80
60
40
20
Artificial peaks (J/degree)

0
0
120
80
40 1
0
350 360 370 380 390 400
CA (degree)

Figure VIII-3: Representation of apparent heat release (AHR) at 1500 rpm, 60%
load, 1º before top dead center injection, 100 MPa fuel injection pressure with a
convolution of two peaks: one Gaussian peak and one ExpConvExp peak. Top:
difference between the engine data and fit curve; middle: comparison of engine data
and fit curve; bottom: deconvolution of artificial peaks.
160
Figures VIII-2 and VIII-3 show that the apparent heat release can be artificially

generated by a convolution of two peaks. Through adjusting the two peaks’ locations,

heights and widths, one is able to test the effect of different apparent heat release patterns.

In addition, the close representation of the artificial peak to the real engine AHR profile

suggests that it should be able to be realized in engine combustion.

8.2.3 Firing cycle calculation

Once the apparent heat release is generated, the cylinder pressure during firing

cycle can be easily calculated. From Equation (36), using a similar numerical procedure

as 8.2.1:

γ ∆V 1 P − P0
AHR = P + V (51)
γ − 1 ∆θ γ − 1 ∆θ

Then the cylinder pressure after ∆θ from a previous pressure P0 is

AHR (γ − 1) ∆θ + P0V
P= (52)
γ (V − V0 ) + V

Similarly, an iteration procedure is needed since γ needs to be consistent with the

temperature. The initial point P0 should choose the motor pressure at start of combustion,

i.e., when AHR > 0 for the first time with increased crank angle. Note Equation (52) is

also only valid during firing, i.e., as long as AHR > 0. Meanwhile the rate of energy input

from fuel will be:

dqt ∆θ
= AHR + qɺw (53)
dθ ω

161
Since the total energy input should compensate both the pressure trace and heat

transfer during the firing. This should be also calculated only during firing. With a known

pressure trace during firing from Equation (52), qɺ w can be readily calculated. The total

energy input from fuel is:

dqt
qt = ∑ dθ ∆θ
AHR
(54)

After the firing crank angle (AHR = 0 again), Equation (50) from Section 8.2.1

should be used to calculate the pressure trace.

Figure VIII-4 compares simulated pressure trace and real engine data, running at

1500 rpm, 30% load, single injection timing at 1º before top dead center, 100 MPa fuel

injection pressure. A good match can be observed (R2 = 0.9648).

162
100
Simulated pressure trace
90 Engine pressure trace

80

70

60
Pressure (bar)

50

40

30

20

10

0
0 100 200 300 400 500 600 700
CA (degree)

Figure VIII-4: comparison of simulated pressure trace with engine data, running at
1500 rpm, 30% load, single injection timing at 1º before top dead center, 100 MPa
fuel injection pressure

After the pressure trace is obtained, it is straightforward to calculate friction work

and energy thermal efficiency. With an apparent heat release identical to the 30% load

engine data, the model gives an indicated efficiency of 49.7% and a brake specific

efficiency of 32.4%. These values are reasonable for this engine [from around 32% (low

load) to around 37% (high load), see Figures IV-11 to IV-13].

8.3 Optimized heat release pattern probing

163
As shown earlier (Figures VIII-2 and VIII-8), the artificially generated apparent

heat release can be deconvoluted to two peaks:

AHR = pG ( h1 , w1 , l1 ) + pE ( h2 , l2 , k1 , k 2 ) (55)

θ − l1
pG ( h1 , w1 , l1 ) = h1 exp[ −( )2 ]
w1


0,if θ<l
 2

pE ( h2 , l2 , k1 , k2 ) =  h2 (θ − l2 ) k 22 exp[ − k2 (θ − l2 )],if k1 =k 2
h kk
 2 1 2 {exp[ − k1 (θ − l2 )] − exp[( − k 2 (θ − l2 )]},otherwise
 k 2 − k1

Where, h1 and h2 are the peak height, w1 is the peak width, l1 and l2 are the

locations of the peaks (determining the start of combustion), and k1, k2 are two

parameters that control the shape of ExpConvExp peak, pE. It is noticed that pE has four

parameters, which introduces high complexity for adjustment. In this work, the

independent adjustment will be conducted primarily on the Gaussian peak pG through h1,

w1 and l1. The h2 of pE will be adjusted based on the principle that the total energy of

apparent heat release is constant, i.e.,

∫ AHR ⋅ dθ = constant (56)

The baseline of the AHR was the default engine output (Figure VIII-2), which

will be indicated by parameters h1b , w1b , l1b , h2b , l2b . To investigate the effect of start of

combustion (SOC), l1 is varied, and l2 is varied with l1 at the principle that the distance

between l1 and l2 is constant, i.e., l2-l1 = constant. Later the impact of distance between l1

and l2 on engine thermal efficiency will also be investigated.

164
8.3.1 The effect of peak height

Figure VIII-5 shows three apparent heat release profiles. These patterns have

different heights in the Gaussian peak. The height of ExpConvExp peak was adjusted

along with the height of the Gaussian peak, so the total area under AHR was constant. A

significant peak height of maximum heat release rate can be observed.

Figure VIII-6 investigates the indicated thermal efficiency of the heat release

profiles in Figure VIII-5. Each curve in Figure VIII-6 is generated corresponding to one

heat release rate pattern. The apparent heat release pattern was literally shifted around the

top dead center. At each location, both the start of combustion (which is noted as the first

location the apparent heat release is higher than 0.1 J/degree) and engine indicated

thermal efficiency were recorded. Figure VIII-6 then summarizes the change of

indicated efficiency with start of combustion. The first observation is that the maximum

indicated efficiencies are all around the start of combustion of 354º. With decreased peak

height, the SOC timing for the maximum indicated efficiency advances slightly. In

addition, the maximum indicated efficiency increases slightly with decreased peak height.

165
200
h1/h1b = 0.2
180 h1/h1b = 0.6
h1/h1b = 1
160

140
AHR (J/degree)

120

100

80

60

40

20

0
350 355 360 365 370 375 380 385 390 395 400
CA (degree)

Figure VIII-5: Examples of apparent heat release with: h1/h1b = 0.2, h2/h2b=1.29;
h1/h1b = 0.6, h2/h2b=1.17; h1/h1b = 1, h2/h2b=1

166
0.51
h1/h1b = 0.2
h1/h1b = 0.6
0.505 h1/h1b = 1

0.5
indicated efficiency

0.495

0.49

0.485

0.48

0.475
350 352 354 356 358 360 362 364 366
Start of Combustion (degree)

Figure VIII-6: Effect of start of combustion on indicated thermal efficiency, h1/h1b


= 0.2, h2/h2b=1.29; h1/h1b = 0.6, h2/h2b=1.17; h1/h1b = 1, h2/h2b=1

Figure VIII-7 presents the effect of SOC on brake thermal efficiency. Similar to

indicated efficiency observation, the maximum brake efficiency also requires an early

SOC. However, decrease of peak height seems to be able to increase the maximum brake

thermal efficiency for the particular friction model employed here. This could be due to

decreased friction from lower cylinder pressure, when a low height is used for the

apparent heat release peak.

167
0.34
h1/h1b = 0.2
h1/h1b = 0.6
0.335
h1/h1b = 1

0.33

0.325
Brake efficiency

0.32

0.315

0.31

0.305

0.3
350 352 354 356 358 360 362 364 366 368
State of Combustion (degree)

Figure VIII-7: Effect of start of combustion on brake thermal efficiency, h1/h1b =


0.2, h2/h2b=1.29; h1/h1b = 0.6, h2/h2b=1.17; h1/h1b = 1, h2/h2b=1

Figures VIII-8 to VIII-10 tested the impact of increasing the Gaussian peak

height. Consequently, with an increased Gaussian peak area, the ExpConvExp peak will

decrease. For the high Gaussian peak case, the ExpConvExp almost disappears, resulting

of a symmetry heat release rate. A consistent observation with Figure VIII-6 is that the

timing of SOC for the maximum indicated thermal efficiency retards with increased peak

height. However, a further increase of the peak height will slightly increase the maximum

indicated efficiency.

168
600
h1/h1b = 1
h1/h1b = 2
500 h1/h1b = 3
h1/h1b = 4

400
AHR (J/degree)

300

200

100

0
350 355 360 365 370 375 380 385 390 395 400
CA (degree)

Figure VIII-8: Examples of apparent heat release with: h1/h1b = 1, h2/h2b=1;


h1/h1b = 2, h2/h2b=0.66; h1/h1b = 3, h2/h2b=0.34; h1/h1b = 4, h2/h2b=0.02

169
0.51

0.505

0.5
Indicated efficiency

0.495

0.49

0.485
h1/h1b = 1
h1/h1b = 2
0.48 h1/h1b = 3
h1/h1b = 4

0.475
350 352 354 356 358 360 362 364 366
Start of Combustion (degree)

Figure VIII-9: Effect of start of combustion on indicated thermal efficiency, h1/h1b


= 1, h2/h2b=1; h1/h1b = 2, h2/h2b=0.66; h1/h1b = 3, h2/h2b=0.34; h1/h1b = 4,
h2/h2b=0.02

Figure VIII-10 shows the brake thermal efficiency with an increase of Gaussian

peak height. The timing of SOC for the maximum heat release retards with the increased

peak height. Meanwhile, the maximum brake efficiency does not show a monotonic trend.

With increase of peak height, the maximum brake efficiency decreases first then

increases.

170
0.34

0.335

0.33
Brake efficiency

0.325

0.32

0.315
h1/h1b = 1
h1/h1b = 2
0.31 h1/h1b = 3
h1/h1b = 4

0.305
350 352 354 356 358 360 362 364 366
Start of Combustion (degree)

Figure VIII-10: Effect of start of combustion on brake thermal efficiency, h1/h1b =


1, h2/h2b=1; h1/h1b = 2, h2/h2b=0.66; h1/h1b = 3, h2/h2b=0.34; h1/h1b = 4,
h2/h2b=0.02

8.3.2 The effect of peak width

Figures VIII-11 to VIII-13 investigated the effect of Gaussian peak width (this

may represent the duration of premixed combustion). By decreasing the peak width, the

ExpConvExp peak size can be slightly increased. The behaviors of indicated thermal

efficiencies are quite similar for decreased Gaussian peak. For brake thermal efficiency, a

decrease in Gaussian peak seems to be able to increase the maximum brake thermal

efficiency. Both the indicated and brake thermal efficiency requires early SOC to reach

the maximum value.

171
250
w1/w1b = 0.6
w1/w1b = 0.8
w1/w1b = 1
200
AHR (J/degree)

150

100

50

0
350 355 360 365 370 375 380 385 390 395 400
CA (degree)

Figure VIII-11: Examples of apparent heat release with: w1/w1b = 0.6, h2/h2b=1.17;
w1/w1b = 0.8, h2/h2b=1.11; w1/w1b = 1, h2/h2b=1

172
0.51
w1/w1b = 0.6
w1/w1b = 0.8
0.505
w1/w1b = 1

0.5
Indicated efficiency

0.495

0.49

0.485

0.48

0.475
352 354 356 358 360 362 364 366 368
Start of Combustion (degree)

Figure VIII-12: Effect of start of combustion on indicated thermal efficiency,


w1/w1b = 0.6, h2/h2b=1.17; w1/w1b = 0.8, h2/h2b=1.11; w1/w1b = 1, h2/h2b=1

173
0.34
w1/w1b = 0.6
w1/w1b = 0.8
0.335 w1/w1b = 1

0.33
Brake efficiency

0.325

0.32

0.315

0.31

0.305
352 354 356 358 360 362 364 366 368
Start of Combustion

Figure VIII-13: Effect of start of combustion on brake thermal efficiency, w1/w1b =


0.6, h2/h2b=1.17; w1/w1b = 0.8, h2/h2b=1.11; w1/w1b = 1, h2/h2b=1

Figure VIII-14 to VIII-16 also investigated the effect of peak width. For this

investigation, the Gaussian peak width is increased, which at the same time forced a

decrease for the peak height. With increased peak width, the timing of SOC for the

maximum indicated thermal efficiency advances, and the maximum value can also be

able to increase by about 0.7%. Similar observations can be found for brake thermal

efficiency. With increased peak width, the brake efficiency can be increased by 1%.

174
200
w1/w1b = 1
180 w1/w1b = 2
w1/w1b = 3
160 w1/w1b = 4

140
AHR (J/degree)

120

100

80

60

40

20

0
350 355 360 365 370 375 380 385 390 395 400
CA (degree)

Figure VIII-14: Examples of apparent heat release with: w1/w1b = 1, h2/h2b=1;


w1/w1b = 2, h2/h2b=0.66; w1/w1b = 3, h2/h2b=0.34; w1/w1b = 4, h2/h2b=0.02

175
0.515
w1/w1b = 1
w1/w1b = 2
0.51 w1/w1b = 3
w1/w1b = 4

0.505
Indicated efficiency

0.5

0.495

0.49

0.485

0.48
340 345 350 355 360 365
Start of Combustion (degree)

Figure VIII-15: Effect of start of combustion on indicated thermal efficiency,


w1/w1b = 1, h2/h2b=1; w1/w1b = 2, h2/h2b=0.66; w1/w1b = 3, h2/h2b=0.34; w1/w1b
= 4, h2/h2b=0.02

176
0.345

0.34

0.335

0.33
Brake efficiency

0.325

0.32

0.315
w1/w1b = 1
w1/w1b = 2
0.31 w1/w1b = 3
w1/w1b = 4

0.305
340 345 350 355 360 365
Start of Combustion (degree)

Figure VIII-16: Effect of start of combustion on brake thermal efficiency, w1/w1b =


1, h2/h2b=1; w1/w1b = 2, h2/h2b=0.66; w1/w1b = 3, h2/h2b=0.34; w1/w1b = 4,
h2/h2b=0.02

8.3.3 The effect of peak distance

Figures VIII-17 to VIII-19 investigates the effect of peak distance. The default

peak convolution was used for this purpose. It can be observed that this split-peak does

not have a significant impact on the maximum indicated and brake thermal efficiency.

177
200
∆L = 0 deg
180 ∆L = 2 deg
∆L = 4 deg
160
∆L = 6 deg

140
AHR (J/degree)

120

100

80

60

40

20

0
350 355 360 365 370 375 380 385 390 395 400
CA (degree)

Figure VIII-17: Examples of apparent heat release with different peak distances

178
0.51

0.505

0.5

0.495
Indicated efficiency

0.49

0.485

0.48

0.475
∆L = 1 deg
0.47 ∆L = 2 deg
∆L = 3 deg
0.465
∆L = 4 deg
0.46
350 352 354 356 358 360 362 364 366
Start of Combustion (degree)

Figure VIII-18: Effect of start of combustion on indicated thermal efficiency with


different peak distance, ∆L = l2-l1

179
0.34

0.33

0.32
Brake efficiency

0.31

0.3

∆L = 1 deg
∆L = 2 deg
0.29
∆L = 3 deg
∆L = 4 deg

350 352 354 356 358 360 362 364 366


Start of Combustion (deg)

Figure VIII-19: Effect of start of combustion on brake thermal efficiency with


different peak distance

Finally, two extreme cases were investigated: a very sharp peak and a very wide

but low peak (Figure VIII-20). Significant differences can be observed. The timing of

SOC for the maximum indicated and brake thermal efficiency is significantly advanced

with a low and wide peak, and the indicated and brake thermal efficiency of low and

wide peak appears to be generally higher than that of a sharp peak (Figures VIII-21 and

VIII-22).

180
1000 w1/w1b=8
h1/h1b = 8

800
AHR (J/degree)

600

400

200

0
350 360 370 380 390 400
CA (degree)
Figure VIII-20: Examples of apparent heat release profiles for sharp peak (h1/h1b =
8) versus low and wide peak (w1/w1b = 8)

181
0.52 w1/w1b=8
h1/h1b = 8

0.51
Indicated efficiency

0.50

0.49

0.48

0.47
325 330 335 340 345 350 355 360 365 370
Start of Combustion (degree)
Figure VIII-21: Indicated thermal efficiency for sharp peak (h1/h1b = 8) versus low
and wide peak (w1/w1b = 8)

182
0.36
w1/w1b=8
h1/h1b = 8
0.35
brake efficiency

0.34

0.33

0.32

0.31

325 330 335 340 345 350 355 360 365 370
Start of Combustion (degree)
Figure VIII-22: Brake thermal efficiency for sharp peak (h1/h1b = 8) versus load
and wide peak (w1/w1b = 8)

It is of interest to investigate the origin of engine thermal efficiency differences

from the change of apparent heat release pattern. Figure VIII-23 shows the heat transfer

loss of the two extreme heat release patterns (Figure VIII-20). It can be observed that the

difference of heat transfer loss between the two cases is not significant. However, the

cylinder pressure traces (Figure VIII-24) are significantly different between the two heat

release patterns. The “sharp” heat release gives a much higher maximum cylinder

pressure and narrower pressure peak, while the “wide and low” heat release pattern gives

a lower maximum cylinder pressure.

183
240
h1/h1b=8
w1/w1b=8
220

200
Heat loss (Joule)

180

160

140

120

100

80

325 330 335 340 345 350 355 360 365 370
Start of Combustion (degree)

Figure VIII-23: Heat transfer loss for sharp peak (h1/h1b = 8) versus load and wide
peak (w1/w1b = 8)

184
160
h1/h1b=8
140 w1/w1b=8

120

100
Pressure (bar)

80

60

40

20

300 320 340 360 380 400 420 440


CA (degree)

Figure VIII-24: Comparison of cylinder pressure for sharp peak (h1/h1b = 8) versus
load and wide peak (w1/w1b = 8) apparent heat release

The significant difference in the pressure trace between these two heat release

patterns leads to a significant difference in the indicated work. As shown in Figure VIII-

25, the “wide and low” apparent heat release pattern gives generally higher indicated

work, i.e., given similar energy input, the “wide and low” heat release pattern can output

higher work. This explains the higher engine thermal efficiency from the “wide and low”

apparent heat release pattern.

185
650
h1/h1b=8
w1/w1b=8
640
Indicated work (Joule)

630

620

610

600

590

580

570

325 330 335 340 345 350 355 360 365 370
Start of Combustion (degree)

Figure VIII-25: Indicated work for sharp peak (h1/h1b = 8) versus low and wide
peak (w1/w1b = 8)

Figure VIII-26 illustrates an example of comparison of P-V diagram between the

two extreme cases. Although the “sharp” heat release profile appears to benefit from the

higher pressure at low volume, the total area of P-V diagram from “low and wide” heat

release profile actually is higher given that the pressure from the “low and wide” heat

release profile is generally higher than that from the “sharp” heat release profile for a

significant range of volume. This illustrates the differences between indicated work from

the two extreme cases.

186
140 h1/h1b=8
w1/w1b=8
120

100
Pressure (Bar)

80

60

40

20

0
0 200 400 600 800
Volume (cm3)

Figure VIII-26: P-V diagram for sharp peak (h1/h1b = 8) versus low and wide peak
(w1/w1b = 8)

It must be acknowledged that the higher engine thermal efficiency from a heat

release profile that is significantly different from a “sharp” shape, i.e., the P-V diagram is

significantly different from the air standard Otto cycle, is counter intuitive. However,

Hoppie et al. [32] in their thermodynamic engine simulation concluded that the Otto

cycle is only an optimized cycle if all loss functions (friction, heat transfer, mass leakage,

etc.) are independent of pressure if maximum cylinder pressure is not constrained. The

model employed in this work indeed calculated the heat transfer loss, for which cases

Hoppie et al. [32] suggested neither Otto nor Diesel cycle is optimum. In fact, Lyn [254]

in his thermodynamic engine simulation concluded that for a given heat release period,

the cycle efficiency is a single valued function of peak pressure, and there appears to be

187
an optimized peak pressure beyond which no significant efficiency benefit can be

observed. For instance, Figure VIII-27 illustrates two kinds of heat release profiles:

AHR1 is more close to the “sharp” shape while AHR2 is more close to the “low and wide”

shape.

200 AHR 1
AHR 2

150
AHR (J/degree)

100

50

300 330 360 390 420 450


CA (degree)

Figure VIII-27: Comparison of two apparent heat release (AHR) profiles with the
same start of combustion timing at 362.2º (2.2º after top dead center), indicated
efficiency of AHR1 = 49.74%, indicated efficiency of AHR2 = 47.23%

With the model employed in this work, the engine cylinder pressure traces of the

two heat release profiles can be easily calculated, as shown in Figure VIII-28. It can be

observed that the peak pressure from the more “low and wide” heat release profile is

significantly lower than that from the more “sharp” heat release profile, and the indicated

thermal efficiency from the more “low and wide” heat release profile is also lower (47.23%

188
to 49.74%). The observation from Figure VIII-28 is entirely consistent with the

observation by Lyn [254], which indicates that if the majority of heat release occurs

during the expansion cycle, a “sharp” heat release profile is preferred since it can

generate higher peak pressure.

160

AHR 1
140
AHR 2

120
Pressure (bar)

100

80

60

40

20

0
300 320 340 360 380 400
CA (degree)

Figure VIII-28: Comparison of cylinder pressure traces of two apparent heat


release (AHR) profiles with the same start of combustion timing at 362.2º (2.2º after
top dead center), indicated efficiency of AHR1 = 49.74%, indicated efficiency of
AHR2 = 47.23%

At the same time, one can tend to advance the start of combustion. As shown in

Figure VIII-29, the exact same shapes of heat release profiles in Figure VIII-27 can be

shifted to an earlier start of combustion timing at 350º (10º before top dead center). The

cylinder pressure trances generated from these two heat release profiles can be found in

Figure VIII-30. It can be observed that the peak pressures of both heat release profiles

189
increase compared to the case with later start of combustion. This explains the increase of

thermal efficiency of both cases. However, the increase of peak pressure for the more

“low and wide” heat release profile has a more significant improvement on the cycle

efficiency. In addition, the extreme high cylinder pressure during the compression cycle

for the more “sharp” heat release profile is actually generating a negative impact on the

cycle efficiency, while its increased peak pressure has a less significant impact on the

cycle efficiency given the peak pressure from the later start of combustion is considerably

high already. These combined effects lead to a higher cycle efficiency for the more “low

and wide” heat release profile (50.84% to 50.02%).

200
AHR 1
AHR 2

150
AHR (J/degree)

100

50

300 330 360 390 420 450


CA (degree)

Figure VIII-29: Comparison of two apparent heat release (AHR) profiles with the
same start of combustion timing at 350º (10º before top dead center), indicated
efficiency of AHR1 = 50.02%, indicated efficiency of AHR2 = 50.84%

190
160

AHR 1
140
AHR 2

120
Pressure (bar)

100

80

60

40

20

0
300 320 340 360 380 400
CA (degree)

Figure VIII-30: Comparison of cylinder pressure traces of two apparent heat


release (AHR) profiles with the same start of combustion timing at 350º (10º before
top dead center), indicated efficiency of AHR1 = 50.02%, indicated efficiency of
AHR2 = 50.84%

The observations from Figures VIII 27 to VIII-30 suggest an explanation of the

unexpected high efficiency from the “low and wide” heat release profile: the “low and

wide” heat release profile with an early start of combustion can effectively make use of

the piston compression movement to reach a high peak cylinder pressure while maintain

a low negative work generated before the expansion cycle. This advantage is usually

difficult to be observed for the “sharp” heat release profile since it tends to either

generate high negative work if it has an early start of combustion or a lower peak

pressure if it has a late start of combustion.

191
In summary, this zero-dimensional engine efficiency simulation presents the

surprising result that an optimized apparent heat release pattern should possess the feature

of a low peak value and a wide profile. This suggests that the one hypothesis that a sharp

peak gives higher engine efficiency is not correct. The reason for this surprising result is

that a lower heat release peak can make use of piston compression to reach a high peak

cylinder pressure without introducing significant negative work.

It can be observed that it is extremely difficult to improve diesel engine efficiency

(even with a dramatic change in apparent heat release pattern as Figure VIII-20, the

maximum improvement in engine efficiency is around 2%). The chief observation from

this simulation work is that the most important consideration in the attempt to improve

engine thermal efficiency is to decrease the heat loss. A “wide and low” apparent heat

release pattern will certainly benefit due to increase of indicated work output. This kind

of “wide and low” apparent heat release can be actually approached from advanced

combustion, e.g., lean-fuel combustion [255]. It will be also interesting to investigate

other factors with this model, such as the maximum cylinder pressure, friction with

advanced lubricating oil and turbo-compounding. However, such considerations were not

included in the scope of this work.

192
Chapter IX. Conclusions and recommendations for future
work

From this work, the fully validated hypotheses include:

• The impact on NOx emission from engine fuel injection strategy control and

biodiesel fueling can be explained by the equivalence ratio at autoignition

zone near the flame lift-off length.

• Soot from high injection pressure has higher oxidative reactivity since the

decreased soot precursor concentration can prolong the soot inception time.

The conditionally validated hypotheses (only valid at certain conditions) are:

• For conventional diesel combustion at high efficiency (fuel injection timing

earlier that 5º before top dead center), biodiesel NOx penalty can be

eliminated by engine fuel injection control. With a decrease fuel injection

pressure, the biodiesel NOx emissions can be reduced to the same level as

petroleum diesel, while maintaining the same level of particulate emissions.

This approach is less effective for combustion with late fuel injection timing.

• Biodiesel fueling can introduce lubricating oil dilution with post injection

strategy. This statement depends on the timing of post injection, and only a

post injection timing later than 45º after top dead center can introduce

lubricating oil dilution.

193
The invalidated hypothesis is:

• Optimized heat release design is a sharp peak at top dead center. On the

contrary, a wide and low peak is the optimized heat release pattern to improve

engine thermal efficiency since it can decrease the heat loss during

combustion.

Other conclusions from this work include:

In a compression ignition direct injection engine, the increase of injection quantity

when biodiesel is used can be obtained by two means: (1) increase of injection pressure

and (2) extension of injection duration. Experiments of investigating both these two

factors have been conducted and the NOx increase is observed for both cases when

biodiesel is used, which suggests that both types of change in injection strategy (injection

pressure or duration) have similar impacts on NOx emissions. Nevertheless, a change of

injection pressure seems to have a higher impact on NOx emissions than a change of

injection duration.

Meanwhile, the impacts of injection strategy and biodiesel fueling on PM

emissions strongly depend on the engine load condition. At low load condition, both the

increase of fuel injection pressure and biodiesel fueling can significantly reduce PM

emissions. However, at moderate to high load conditions, the biodiesel fueling has a less

significant impact on PM emissions.

No significant difference in brake thermal efficiency is observed with biodiesel

fueling or change of fuel injection pressure. However, it is observed that retarding the

start of injection can reduce the brake thermal efficiency. The highest brake thermal

194
efficiency in this work has been observed at SOI between 9˚ to 5˚ before top dead center.

In addition, the NOx-PM emission trade-off of SOI timing at 9˚ BTDC shows that

decreasing the fuel injection pressure is an effective way to counter the biodiesel NOx

effect while maintaining a similar or even lower level of PM emissions than ULSD.

Therefore, for an engine which intends to run at a high brake thermal efficiency, a

decrease of fuel injection pressure is an applicable approach to solve the biodiesel NOx

penalty problem.

The heat release analysis shows that, for all load conditions, retarding the start of

injection after top dead center can increase the fraction of premixed combustion. At the

same SOI, higher fuel injection pressure leads to higher apparent heat release rate and a

slightly earlier start of combustion (SOC) due to the better mixing of fuel and air. At low

load conditions, the biodiesel fueling can advance the SOC and decrease the premixed

heat release. At moderate to high loads, however, the biodiesel fueling does not have a

significant impact on the heat release profile. Therefore in this study, combustion phasing

is not a primary factor that contributes to the biodiesel NOx effect for higher load.

It is confirmed that the key factor that determines the NOx emissions is the

closeness to stoichiometric conditions of the oxygen equivalence ratio of the fuel-air

mixture at the autoignition zone near the lift-off length, for all load conditions and a

range of start of injection from 9˚ BTDC to 3˚ ATDC. On the basis of the fuel spray

model by Pickett, Siebers and co-workers, the average oxygen equivalence ratios at lift-

off length are observed to be linearly correlated with brake specific NOx emissions at the

same SOI for all load conditions. In addition, the slopes and Y-axis intercepts (at NOx

emissions equal to 0) of the correlations are found to be exponentially correlated with

195
SOI, based on which general equations to predict NOx emissions are proposed. These

correlations are applicable for both petroleum diesel and biodiesel. In addition, given

their significantly lower requirement for CPU power than complicated computational

fluid dynamic models, these correlations may be implemented into the electronic control

unit of a modern engine to perform post (or even real time) evaluation of the engine NOx

emissions

Multi-injection strategy including post injection is an effective way to decrease

NOx emissions and increase exhaust temperature for diesel engine. Depending on the

timing and quantity of post injection, the engine showed a significant change in exhaust

temperature, brake specific fuel consumption, UHC and CO emissions. Between the post

injection timings from 45º to 70º ATDC, the consumption of post injected fuel was found

to primarily depend on the injection timing.

Post injection at 45º ATDC with the quantity of 5mg/stk had significant

combustion at late crank angle (LCA), i.e., later than 40º after top dead center.

Consequently it generated power – reducing the energy requirement for the pilot and

main injections – and produced a significantly higher exhaust temperature without an

apparent increase in BSFC. The UHC and CO emissions only slightly increased

compared with the test without post injection.

Post injection at 60º ATDC (5mg/stk), 60º ATDC (8mg/stk) and 70º ATDC

(5mg/stk) did not have apparent heat release at LCA. Compared with 45º ATDC

(5mg/stk), they barely generated any power, decreased the exhaust temperature and

significantly increased BSFC, UHC and CO emissions.

196
A model was employed to calculate the ignition delay of post injected fuel. The

temperature at which the ignition was supposed to happen was confirmed to coordinate

with the quality of combustion of post injected fuel at late crank angle. A bulk cylinder

temperature of 1500K was confirmed to be the lowest acceptable temperature for which

high temperature combustion of the post injected fuel could occur in the engine.

When the fuel was injected into an environment with temperature lower than

1500K, only slow fuel oxidation occurs. Hence, the fuel escaped the cylinder at a much

higher concentration than during high temperature combustion. GC-MS analysis showed

that the primary hydrocarbons in the exhaust gases were vaporized diesel and biodiesel

components regardless of the post injection strategy, and those hydrocarbons became the

primary source for lubricant dilution when they were dissolved into the oil film on the

cylinder wall. Additionally, in this experiment only alkanes with carbon number larger

than 14 from the diesel fuel and biodiesel were observed to dilute into the lubricant

because of the high temperature of the cylinder wall. The comparative mass fraction of

fatty acid methyl esters shifted towards more saturated and higher boiling point

components when biodiesel was diluted into the lubricant.

Fuel injection pressure can significantly affect the morphology of soot aggregates.

Increase of fuel injection pressure can decrease the fractal dimension of the aggregates

and decrease the number of primary particle. Increase of fuel injection pressure can also

increase the nucleation mode peak and decrease the accumulation mode peak in the

particle size distribution.

At the same fuel injection pressure, B20 soot has higher oxidative reactivity.

Meanwhile, the reactivity of soot particle is varies with burn-off. It often increases with

197
higher burn-off. However, for soot generated at very high fuel injection pressure, the

oxidative reactivity tends to firstly increase with burn-off and decreased when the burn-

off is close to completion.

The recommended future work includes:

• Investigate the kinetic role of start of injection timing in engine NOx

emissions. As shown in Chapter V, the start of injection timing serves as a

parameter similar to the time in a kinetic model.

• Investigate lubricating oil dilution in an advanced combustion mode, such as

low temperature combustion. The investigation of post injection in this work

(Chapter VI) is still conventional combustion. A significant increase of EGR

can promote advanced combustion, which may introduce more serious oil

dilution.

• Investigate the evolution of active surface area and carbon crystalline structure

of diesel and biodiesel soot from different fuel injection pressures. The

information of active surface area and carbon nanostructure during burn-off

can be critical in the explanation of evolution of instantaneous reactivity of

soot particles.

• Investigate the impact of maximum cylinder pressure, friction with advanced

lubricating oil and turbo compounding on engine efficiency. This can be

implemented to investigate the pathways of high engine thermal efficiency.

198
References

[1] J. Heywood, Internal Combustion Engine Fundamentals, McGraw-Hill, Inc, (1988).

[2] T.V. Johnson, Review of Diesel Emissions and Control, International Journal of

Engine Research, 10 (2009) 275-285.

[3] DieselNet, United States New Engine and Vehicle Emissions,

http://www.dieselnet.com/standards/, (2011).

[4] H. Lund, Renewable Energy Strategies for Sustainable Development, Energy, 32

(2007) 912-919.

[5] ASTM, Standard Specification for Biodiesel Fuel Blend Stock (B100) for Middle

Distillate Fuels, in, ASTM International Specification ASTM D 6751, 2008.

[6] M. Graboski, R. McCormick, Combustion of Fat and Vegetable Oil Derived Fuels in

Diesel Engines, Progree in Energy and Combustion Science, 24 (1998) 125-164.

[7] G. Knothe, "Designer" Biodiesel: Optimizing Fatty Ester (Composition to Improve

Fuel Properties, Energy & Fuels, 22 (2008) 1358-1364.

[8] R.L. McCormick, The Impact of Biodiesel on Pollutant Emissions and Public Health,

Inhalation Toxicology, 19 (2007) 1033-1039.

[9] J. Yanowitz, R.L. Mccormick, Effect of Biodiesel Blends on North American Heavy-

Duty Diesel Engine Emissions, European Journal of Lipid Science and Technology, 111

(2009) 763-772.

[10] M. Graboski, J.D.Ross, R.L.McCormick, Transient Emissions from No. 2 Diesel and

Biodiesel Blends in a DDC Series 60 Engine., SAE paper 961166 (1996).

199
[11] S. Murillo, J. Miguez, J. Porteiro, E. Granada, J. Moran, Performance and Exhaust

Emissions in the Use of Biodiesel in Outboard Diesel Engines, Fuel, 86 (2007) 1765-

1771.

[12] T.L.Allenman, R.L.Mccormick, Results of the 2007 B100 Quality Survey, NREL

Technical Report NREL/TP 540-42787, (2008).

[13] Y. Zhang, A.L. Boehman, Impact of Biodiesel on NOx Emissions in a Common Rail

Direct Injection Diesel Engine, Energy & Fuels, 21 (2007) 2003-2012.

[14] J.P. Szybist, A.L. Boehman, J.D. Taylor, R.L. Mccormick, Evaluation of

Formulation Strategies to Eliminate the Biodiesel NOx Effect, Fuel Processing

Technology, 86 (2005) 1109-1126.

[15] M. Lapuerta, O. Armas, J. Rodriguez-Fernandez, Effect of Biodiesel Fuels on Diesel

Engine Emissions, Progress in Energy and Combustion Science, 34 (2008) 198-223.

[16] G.A. Ban-Weiss, J.Y. Chen, B.A. Buchholz, R.W. Dibble, A Numerical

Investigation into the Anomalous Slight NOx Increase When Burning Biodiesel: a New

(Old) Theory, Fuel Processing Technology, 88 (2007) 659-667.

[17] C.Y. Choi, R.D. Reitz, A Numerical Analysis of the Emissions Characteristics of

Biodiesel Blended Fuels, Journal of Engineering for Gas Turbines and Power,

Transactions of the ASME, 121 (1999) 31-37.

[18] C.J. Mueller, A.L. Boehman, G. Martin, An Experimental Investigation of the Origin

of Increased Nox Emissions When Fueling a Heavy-Duty Compression-Ignition Engine

with Soy Biodiesel, SAE paper 2009-01-1792, (2009).

200
[19] K. Yamamoto, K. Takada, J. Kusaka, Y. Kanno, M. Nagata, Influence of Diesel Post

Injection Timing on Hc Emissions and Catalytic Oxidation Performance, SAE paper

2006-01-3442, (2006).

[20] M.M. Andreae, H.L. Fang, K. Bhandary, Biodiesel on Fuel Dilution of Engine Oil,

SAE paper 2007-01-4036, (2007).

[21] P.J. Tennison, R. Reitz, An Experimental Investigation of the Effects of Common-

Rail Injection System Parameters on Emissions and Performance in a High-Speed Direct-

Injection Diesel Engine, Journal of Engineering for Gas Turbines and Power,

Transactions of the ASME, 123 (2001) 167-174.

[22] P.K. Karra, M.K. Veltman, S.C. Kong, Characteristics of Engine Emissions Using

Biodiesel Blends in Low-Temperature Combustion Regimes, Energy & Fuels, 22 (2008)

3763-3770.

[23] A.L. Boehman, D. Morris, J. Szybist, E. Esen, The Impact of the Bulk Modulus of

Diesel Fuels on Fuel Injection Timing, Energy & Fuels, 18 (2004) 1877-1882.

[24] G. Knothe, Dependence of Biodiesel Fuel Properties on the Structure of Fatty Acid

Alkyl Esters, Fuel Processing Technology, 86 (2005) 1059-1070.

[25] K. Yehliu, A.L. Boehman, O. Armas, Emissions from Different Alternative Diesel

Fuels Operating with Single and Split Fuel Injection, Fuel, 89 (2010) 423-437.

[26] D.T. Hountalas, V. Lamaris, E. Pariotis, H. Ofner, Parametric Study Based on a

Phenomenological Model to Investigate the Effect of Post Fuel Injection on HDDI Diesel

Engine Performance and Emissions - Model Validation Using Experimental Data, SAE

paper 2008-01-0641, (2008).

201
[27] W.D. Ojeda, P. Zoldak, R. Espinosa, R. Kumar, Development of a Fuel Injection

Strategy for Partially Premixed Compression Ignition Combustion, SAE paper 2009-01-

1527, (2009).

[28] H.H. Yun, R.D. Reitz, An Experimental Investigation on the Effect of Post-Injection

Strategies on Combustion and Emissions in the Low-Temperature Diesel Combustion

Regime, Journal of Engineering for Gas Turbines and Power, Transactions of the ASME,

129 (2007) 279-286.

[29] M. Yao, H. Wang, Z. Zheng, Y. Yue, Experimental Study of Multiple Injections and

Coupling Effects of Multi-Injection and EGR in a HD Diesel Engine, SAE paper 2009-

01-2807, (2009).

[30] A. Vanegas, H. Won, C. Felsch, M. Gauding, N. Peters, Experimental Investigation

of the Effect of Multiple Injections on Pollutant Formation in a Common-Rail Di Diesel

Engine, SAE paper 2008-01-1191, (2008).

[31] K. Anand, R.P. Sharma, P.S. Mehta, Experimental Investigations on Combustion,

Performance, and Emissions Characteristics of a Neat Biodiesel-Fueled, Turbocharged,

Direct Injection Diesel Engine, Journal of Automobile Engineering, 224 (2009) 661-679.

[32] L.O. Hoppie, Y.K. Min, N. Srinivasan, Optimum Heat Release for a Reciprocating

Internal Combustion Engine, SAE paper 870572 presented in International Congree and

Exposition in Detroit, MI, (1987).

[33] G. Knothe, Biodiesel and Renewable Diesel: A Comparison, Progress in Energy and

Combustion Science, 36 (2010) 364-373.

[34] A.K. Agarwal, Biofuels (Alcohols and Biodiesel) Applications as Fuels for Internal

Combustion Engines, Progress in Energy and Combustion Science, 33 (2007) 233-271.

202
[35] A. Demirbas, Progress and Recent Trends in Biofuels, Progress in Energy and

Combustion Science, 33 (2007) 1-18.

[36] G. Knothe, Historical Perspectives on Vegetable Oil-Based Diesel Fuels, INFORM,

12 (2001) 1103-1107.

[37] A. Chavanne, Fuels from Vegetable Oils. (Procede De Transformation D'huiles

Vegetables En Vue De Leur Utilisation Comme Carburants), Belgian Patent BE422,877,

(1937).

[38] A. Srivastava, R. Prasad, Triglycerides-Based Diesel Fuels, Renewable &

Sustainable Energy Reviews, 4 (2000) 111-133.

[39] B.R. Moser, Biodiesel Production, Properties, and Feedstocks, In Vitro Cell DEV-

PL, 45 (2009) 229-266.

[40] F.D. Funstone, J.L. Harwood, A.J. Dijkstra, The Lipid Handbook, 3rd Ed., Roca

Raton, 2007.

[41] G. Knothe, K.R. Steidley, Kinematic Viscosity of Biodiesel Fuel Components and

Related Compounds. Influence of Compound Structure and Comparison to Petrodiesel

Fuel Components, Fuel, 84 (2005) 1059-1065.

[42] G. Knothe, K.R. Steidley, Lubricity of Components of Biodiesel and Petrodiesel.

The Origin of Biodiesel Lubricity, Energy & Fuels, 19 (2005) 1192-1200.

[43] A. Schonborn, N. Ladommatos, J. Williams, R. Allan, J. Rogerson, The Influence of

Molecular Structure of Fatty Acid Monoalkyl Esters on Diesel Combustion, Combustion

and Flame, 156 (2009) 1396-1412.

[44] J. Sun, J.A. Caton, T.J. Jacobs, Oxides of Nitrogen Emissions from Biodiesel-Fueled

Diesel Engines, Progress in Energy and Combustion Science, 36 (2010) 677-695.

203
[45] Y. Ra, R. Reitz, J. Mcfarlane, C. Daw, Effects of Fuel Physical Properties on Diesel

Engne Combustion Using Diesel and Biodiesel Fuels, SAE paper 2008-01-1379 SAE

Trans - Journal of Fuels and Lubricants, (2008).

[46] M.E. Tat, J.H. Van Gerpen, Speed of Sound and Isentropic Bulk Modulus of Alkyl

Monoesters at Elevated Temperatures and Pressures, Journal of American Oil Chemistry

Society, 80 (2003) 1249-1256.

[47] H.S. Hess, The Impact of Oxygenated Fuels on Diesel Combustion and Emissions,

in: Department of Energy and Geo-Environmental Engineering, The Pennsylvania State

University, University Park, PA, 2002.

[48] C. Allen, K. Watts, Comparative Analysis of the Atomization Characteristics of

Fifteen Biodiesel Fuel Types, Transactions of the ASAE, 43 (2000) 207-211.

[49] S. Takahashi, K. Wakimoto, N. Iida, D. Nikolic, Effects of Aroamatics Content and

90% Distillation Temperature of Diesel Fuels on Flame Temperature and Soot Formation,

SAE paper 2001-01-1950 SAE Trans. - Journal of Fuels and Lubricants, (2001).

[50] W.A. Eckerle, E.J. Lyford-Pike, D. Stanton, J. Wall, L. Lapointe, S. Whitacre,

Effects of Methyl-Ester Biodiesel Blends on NOx Production, SAE paper 2008-01-0078,

(2008).

[51] EPA, Assessment and Standards Division (Office of Transportation and Air Quality

of the Us Environmental Protection Agency). A Comprehensive Analysis of Biodiesel

Impacts on Exhaust Emissions, EPA 420-P-02-001, (2002).

[52] FEV Engine Technology, Emissiosn and Performance Characteristics of the Navistar

T444E DI Engine Fueled with Blends of Biodiesel and Low Sulphur Diesel, Final report

to National Biodiesel Board, (1994).

204
[53] W. Marshall, L. Schumacher, S. Howell, Engine Exhaust Emissions Evaluation of a

Cummins L10E When Fueled with a Biodiesel Blend, SAE paper 952363, (1995).

[54] J.P. Szybist, J.H. Song, M. Alam, A.L. Boehman, Biodiesel Combustion, Emissions

and Emission Control, Fuel Processing Technology, 88 (2007) 679-691.

[55] P. Ye, A.L. Boehman, Investigation of the Impact of Engine Injection Strategy on

the Biodiesel NOx Effect with a Common-Rail Turbocharged Direct Injection Diesel

Engine, Energy & Fuels, 24 (2010) 4215-4225.

[56] A. Serdari, K. Fragioudakis, C. Teas, F. Zannikos, S. Stournas, E. Lois, Effect of

Biodiesel Addition to Diesel Fuel on Engine Performance and Emissions, Journal of

Propulsion and Power, 15 (1999) 224-231.

[57] K Hamasaki, E Kinoshita, H Tajima, K Takasaki, D Morita, Combustion

Characteristics of Diesel Engines with Waste Vegetable Oil Methyl Ester, The 5th

international symposium on diagnostics and modeling of combustion in internal

combustion engines (COMODIA), (2001).

[58] F. Staat, P. Gateau, The Effects of Rapeseed Oil Methyl Ester on Diesel Engine

Performance, Exhaust Emissiosn and Long Term Behaviour - a Summary of Three Years

of Experimentation, SAE paper 950053, (1995).

[59] M.E. Tat, Investigation of Oxides of Nitrogen Emissions from Biodiesel-Fueled

Engines, in: Mechanical Engineering, Iowa State University, 2003.

[60] R.L. McCormick, T.L. Alleman Impact of Biodiesel on Pollutant Emissions from

Diesel Engines, National Renewable Energy Laboratory, (2005).

[61] R.L. McCormick, Effects of Biodiesel on Nox Emissions, National Renewable

Energy Laboratory, (2005).

205
[62] X. Li, O.L. Gulder, Influence of Diesel Fuel Cetane Number and Aromatic Content

on Engine Exhaust Emissions, Journal of Canadian Petroleum Technology, 37 (1998) 56-

60.

[63] O. Armas, K. Yehliu, A.L. Boehman, Effect of Alternative Fuels on Exhaust

Emissions During Diesel Engine Operation with Matched Combustion Phasing, Fuel, 89

(2010) 438-456.

[64] T. Durbin, J. Collins, J. Norbeck, M. Smith, Effects of Biodiesel, Biodiesel Blends,

and a Synthetic Diesel on Emissions from Light Heavy-Duty Diesel Vehicles,

Environmental Science & Technology, 34 (2000) 349-355.

[65] S. Mandpe, S. Kadlaskar, W. Degen, S. Keppeler, On Road Testing of Advanced

Common Rail Diesel Vehicles with Biodiesel from the Jatropha Curcas Plant, SAE paper

2005-26-356 (2005 ).

[66] M.N. Nabi, M.S. Akhter, M.M.Z. Shahadat, Improvement of Engine Emissions with

Conventional Diesel Fuel and Diesel-Biodiesel Blends, Bioresource Technology, 97

(2006) 372-378.

[67] M.S. Graboski, R.L. McCormick, T.L. Alleman, A.M. Herring, The Effect of

Biodiesel Composition on Engine Emissions from a DDC Series 60 Diesel Engine,

National Renewable Energy Laboratory; NREL/SR-510-31461, (2003).

[68] M. Lapuerta, O. Armas, R. Ballesteros, J. Fernandez, Diesel Emissions from

Biofuels Derived from Spanish Potential Vegetable Oils, Fuel, 84 (2005) 773-780.

[69] M. Lapuerta, O. Armas, R. Ballesteros, Diesel Particulate Emissions from Biofuels

Derived from Spanish Vegetable Oils, SAE paper 2002-01-1657 (2002).

206
[70] M.P. Dorado, E. Ballesteros, J.M. Arnal, J. Gomez, F.J. Lopez, Exhaust Emissions

from a Diesel Engine Fueled with Transesterified Waste Olive Oil, Fuel, 82 (2003) 1311-

1315.

[71] W.G. Wang, D.W. Lyons, N.N. Clark, M. Gautam, P.M. Norton, Emissions from

Nine Heavy Trucks Fueled by Diesel and Biodiesel Blend without Engine Modification,

Environmental Science & Technology, 34 (2000) 933-939.

[72] M. Cardone, M. Prati, V. Rocco, M. Seggiani, A. Senatore, S. Vitolo, Brassica

Carinata as an Alternative Oil Crop for the Production of Biodiesel in Italy: Engine

Performance and Regulated and Unregulated Exhaust Emissions, Environmental Science

& Technology, 36 (2002) 4656-4662.

[73] J.F. Mcdonald, S.T. Bagley, L.D. Gratz, J.H. Johnson, Effects of an Oxidation

Catalytic Converter and a Biodiesel Fuel on the Chemical, Mutagenic, and Particle Size

Characteristics of Emissions from a Diesel Engine, Environmental Science &

Technology, 32 (1998) 1183-1191.

[74] K.F. Hansen, M.G. Jensen, Chemical and Biological Characteristics of Exhaust

Emissions from a Di Diesel Engine Fuelled with Rapeseed Oil Methyl Ester (Rme), SAE

paper 971689, (1997).

[75] A. Boehman, J. Song, M. Alam, Impact of Biodiesel Blending on Diesel Soot and

the Regeneration of Particulate Filters, Energy & Fuels, 19 (2005) 1857-1864.

[76] J.H. Van Gerpen, M. Canakci, Comparison of Engine Performance and Emissions

for Petroleum Diesel Fuel, Yellow Grease Biodiesel, and Soybean Oil Biodiesel,

Transactions of the ASAE, 46 (2003) 937-944.

207
[77] L. Schumacher, W. Marshall, J. Krahl, W. Wetherell, M.S. Graboski, Biodiesel

Emissions Data from Series 60 DDC Engines, Transactions of the ASAE 44 (2006) 1465-

1468.

[78] M. Haas, K. Scott, T. Alleman, R. Mccormick, Engine Performance of Biodiesel

Fuel Prepared from Soybean Soapstock: A High Quality Renewable Fuel Produced from

a Waste Feedstock, Energy & Fuels, 15 (2001) 1207-1212.

[79] M. Lapuerta, O. Armas, R. Ballesteros, M. Carmona, Fuel Formulation Effects on

Passenger Car Diesel Engine Particulate Emissions and Composition, SAE paper 2000-

01-1850 (2000).

[80] O. Armas, J.J. Hernandez, M.D. Cardenas, Reduction of Diesel Smoke Opacity from

Vegetable Oil Methyl Esters During Transient Operation, Fuel, 85 (2006) 2427-2438.

[81] R. Last, M. Kruger, M. Durnholz, Emissions and Performance Characteristics of a 4-

Stroke, Direct Injected Diesel Engine Fueled with Blends of Biodiesel and Low Sulfur

Diesel Fuel, SAE paper 950054, (1995).

[82] J. Krahl, A. Munack, M. Bahadir, L. Schumacher, N. Elser, Review: Utilization of

Rapeseed Oil, Rapeseed Oil Methyl Ester or Diesel Fuel: Exhaust Gas Emissions and

Estimation of Environmental Effects, SAE paper 962096, (1996).

[83] T. Durbin, J. Norbeck, Effects of Biodiesel Blends and Arco Ec-Diesel on Emissions

from Light Heavy-Duty Diesel Vehicles, Environmental Science & Technology, 36

(2002) 1686-1691.

[84] D.Y.C. Leung, Y. Luo, T.L. Chan, Optimization of Exhaust Emissions of a Diesel

Engine Fuelled with Biodiesel, Energy & Fuels, 20 (2006) 1015-1023.

208
[85] C. Peterson, D. Reece, Emissions Testing with Blends of Esters of Rapeseed Oil

Fuel with and without a Catalytical Converter, SAE paper 961114, (1996).

[86] J.A. Miller, C.T. Bowman, Mechanism and Modeling of Nitrogen Chemistry in

Combustion, Progress in Energy and Combustion Science, 15 (1989) 287-338.

[87] S.R. Turns, Understanding Nox Formation in Nonpremixed Flames: Experiments

and Modeling, Progress in Energy and Combustion Science, 21 (1995) 361-385.

[88] K.K. Kuo, Principles of Combustion, John Wiley & Sons, Inc, U.S.A., (1986).

[89] W. Yuan, A.C. Hansen, M.E. Tat, J.H. Van Gerpen, Z. Tan, Spray, Ignition, and

Combustion Modeling of Biodiesel Fuels for Investigating Nox Emissions, Transactions

of the ASAE, 48 (2005) 933-939.

[90] J. Szybist, D. Morris, A. Boehman, E. Esen, Diesel Fuel Formation Effects on

Injection Timing and Emissions, paper presented in American Chemistry. Society, 48

(2003) 428-429.

[91] J.P. Szybist, S.R. Kirby, A.L. Boehman, NOx Emissions of Alternative Diesel Fuels:

A Comparative Analysis of Biodiesel and Ft Diesel, Energy & Fuels, 19 (2005) 1484-

1492.

[92] C. Allen, K. Watts, Comparative Analysis of the Atomization Characteristics of

Fiften Biodiesel Fuel Types, Transactions of the ASAE 2000, 43 (2000) 207-211.

[93] T.L. Allenman, R.L. Mccormick, Fischer-Tropsch Diesel Fuels - Properties and

Exhaust Emissions: A Literature Review, SAE paper 2003-01-0763, (2003).

[94] Y. Icingur, D. Altiparmak, Effect of Fuel Cetane Number and Injection Pressure on a

DI Diesel Engine Performance and Emissions, Energy Conversion and Management, 44

(2003) 389-397.

209
[95] C.F.Taylor, Internal Combustion Engines, International Textbook Company,

Seranton, Pennsylvania, 1989.

[96] J.H. Song, K. Cheenkachorn, J.G. Wang, J. Perez, A.L. Boehman, P.J. Young, F.J.

Waller, Effect of Oxygenated Fuel on Combustion and Emissions in a Light-Duty Turbo

Diesel Engine, Energy & Fuels, 16 (2002) 294-301.

[97] J. Song, V. Zello, A.L. Boehman, F.J. Waller, Comparison of the Impact of Intake

Oxygen Enrichment and Fuel Oxygenation on Diesel Combustion and Emissions, Energy

& Fuels, 18 (2004) 1282-1290.

[98] S. Garner, R. Sivaramakrishnan, K. Brezinsky, The High-Pressure Pyrolysis of

Saturated and Unsaturated C-7 Hydrocarbons, Proceedings of the Combustion Institute,

32 (2009) 461-467.

[99] N. Nabi, Z. Shahadat, S. Rhaman, B. Alam, Behavior of Diesel Combustion and

Exhaust Emission with Neat Diesel Fuel and Diesel-Biodiesel Blends, SAE paper 2004-

01-3034, (2004).

[100] Y. Huang, S. Wang, L. Zhou, Effects of Fischer-Tropsch Diesel Fuel on

Combustion and Emissions of Direct Injection Diesel Engine, Frontiers of Energy and

Power Engineering in China, 2 (2008) 261-267.

[101] N. Miyamoto, H. Ogawa, M. Shibuya, K. Arai, O. Esmilaire, Influence of the

Molecular Structure of Hydrocarbon Fuels on Diesel Exhaust Emissions, SAE paper

940676, (1994).

[102] A. Monyem, J.H. Van Gerpen, M. Canakci, The Effect of Timing and Oxidation on

Emissions from Biodiesel-Fueled Engines, Transactions of the ASAE, 44 (2001) 35-42.

210
[103] A. Cheng, A. Upatnieks, C.J. Mueller, Investigation of the Impact of Biodiesel

Fueling on Nox Emissions Using an Optical Direct Injection Diesel Engine, International

Journal of Engine Research, (2006).

[104] W.A. Sirignano, Fuel Droplet Vaporization and Spray Combustion Theory,

Progress in Energy and Combustion Science, 9 (1983) 291-322.

[105] D.L. Siebers, Scaling Liquid-Phase Fuel Penetration in Diesel Sprays Based on

Mixing-Limited Vaporization, SAE paper 1999-01-0528, 108 (1999).

[106] J.D. Naber, D.L. Siebers, Effects of Gas Density and Vaporization on Penetration

and Dispersion of Diesel Sprays, SAE paper 960034, (1996).

[107] C.D. Rakopoulos, D.T. Hountalas, E.G. Giakoumis, E.C. Andritsakis. Performance

and emissions of bus engine using blends of diesel fuel with bio-diesel of sunflower or

cottonseed oils derived from Greek feedstocks. Fuel, (2007)

[108] P.J.M. Frijters, R.S.G. Baert, Oxygenated Fuels for Clean Heavy-Duty Diesel

Engines, International Journal of Vehicle Design, 41 (2006) 242-255.

[109] N. Ladommatos, K. Sison, H.W. Song, H. Zhao, Soot Generation of Diesel Fuels

with Substantial Amounts of Oxygen-Bearing Compounds Added, Fuel, 86 (2007) 345-

352.

[110] P. Flynn, R. Durrett, G. Hunter, A. Loye, O. Akinyemi, J.E. Dec, Diesel

Combustions: An Integrated View Combining Laser Diagnostics, Chemical Kinetics, and

Emprical Validation, SAE paper 1999-01-0509, (1999).

[111] J.E. Dec, A Conceptual Model of DI Diesel Combustion Based on Laser-Sheet

Imaging, SAE paper 970873, (1997).

211
[112] D. Chang, J.H. Van Gerpen, Fuel Properties and Engine Performance for Biodiesel

Prepared from Modified Feedstockes, SAE paper 971684, (1997).

[113] K. Schmidt, J.H. Van Gerpen, The Effect of Biodiesel Fuel Composition on Diesel

Combustion and Emissions, SAE paper 961086 (1996).

[114] G. Knothe, C. Sharp, T. Ryan, Exhaust Emissions of Biodiesel, Petrodiesel, Neat

Methyl Esters, and Alkanes in a New Technology Engine, Energy & Fuels, 20 (2006)

403-408.

[115] C.Y. Choi, G.R. Bower, R.D. Reitz, Effects of Biodiesel Blended Fuels and

Multiple Injections on D.I .Diesel Engine, SAE paper 970218, (1997).

[116] J. Song, M. Alam, A. Boehman, U. Kim, Examination of the Oxidation Behavior of

Biodiesel Soot, Combustion and Flame, 146 (2006) 589-604.

[117] K. Yehliu, R.L. Vander Wal, A.L. Boehman, Development of an HRTEM Image

Analysis Method to Quantify Carbon Nanostructure, Combustion and Flame, In Press,

Corrected Proof (2011).

[118] R.E. Franklin, Crystallite Growth in Graphitizing and Non-Graphitizing Carbons,

Proceedings of the Royal Society of London Series a-Mathematical and Physical

Sciences, 209 (1951).

[119] H. Marsh, Introduction to Carbon Science, Butterworths, London, (1989) 107-152.

[120] D.M. Smith, A.R. Chughtai, The Surface-Structure and Reactivity of Black Carbon,

Colloids and Surfaces A, Physicochemical and Engineering Aspects, 105 (1995) 47-77.

[121] A. Duran, J.M. Monteagudo, O. Armas, J.J. Hernandez, Scrubbing Effect on Diesel

Particulate Matter from Transesterified Waste Oils Blends, Fuel, 85 (2006) 923-928.

212
[122] U. Tl, S. Kb, M. Rl, Effects of Cetane Number, Cetane Improver, Aromatics and

Oxygenates on 1994 Heavy-Duty Diesel Engine Emissions, SAE paper 941020 (1994).

[123] M. Bunce, D. Snyder, G. Adi, C. Hall, J. Koehler, B. Davila, S. Kumar, P.

Garimella, D. Stanton, G. Shaver, Stock and Optimized Performance and Emissions with

5 and 20% Soy Biodiesel Blends in a Modern Common Rail Turbo-Diesel Engine,

Energy & Fuels, 24 (2010) 928-939.

[124] M. Bunce, D. Snyder, G. Adi, C. Hall, J. Koehler, B. Davila, S. Kumar, P.

Garimella, D. Stanton, G. Shaver, Optimization of Soy-Biodiesel Combustion in a

Modern Diesel Engine, Fuel, 90 (2011) 2560-2570.

[125] E. Mancaruso, S.S. Merola, B.M. Vaglieco, Study of the Multi-Injection

Combustion Process in a Transparent Direct Injection Common Rail Diesel Engine by

Means of Optical Techniques, International Journal of Engine Research, 9 (2008) 483-

498.

[126] W.L. Hardy, R.D. Reitz, An Experimental Investigation of Partially Premixed

Combustion Strategies Using Multiple Injections in a Heavy-Duty Diesel Engine, SAE

paper 2006-01-0917, (2006).

[127] M. Zheng, R. Kumar, Implementation of Multiple-Pulse Injection Strategies to

Enhance the Homogeneity for Simultaneous Low-Nox and -Soot Diesel Combustion,

International Journal of Thermal Sciences, 48 (2009) 1829-1841.

[128] Z. Han, A. Uludogan, G.J. Hampson, R.D. Reitz, Mechanism of Soot and Nox

Emission Reduction Using Multiple-Injection in a Diesel Engine, SAE paper 960633,

(1996).

213
[129] Y. Wang, C. Zhang, J. Jiang, A Numerical Study of the Effect of Multi-Injection

Strategy on Nox Reduction in DI Diesel Engines, HT2005: Proceedings of the ASME

Summer Heat Transfer Conference 2005, Vol 1, (2005) 759-765.

[130] X. Lu, Y. Ge, S. Wu, X. Han, An Experimental Investigation on Combustion and

Emissions Characteristics of Turbocharged DI Engines Fueled with Blends of Biodiesel,

SAE paper 2005-01-2199, presented at the SAE Fuels and Lubricants Meeting, Rio de

Janeiro, Brazil, (2005).

[131] V. Pradeep, R.P. Sharma, Evaluation of Performance, Emissions and Combustion

Parameters of a CI Engine Fuelled with Bio-Diesel from Rubber Seed Oil and Its Blends,

SAE paper 2005-26-353, (2005).

[132] D. Taylor, B. Walsham, Combustion Processes in a Medium-Speed Diesel Engine,

Proceedings of the Institution of Mechanical Engineers 1847-1996, 184 (1969) 67-76.

[133] R. Burt, K. Troth, Penetration and Vaporization of Diesel Fuel Sprays, Proceedings

of the Institution of Mechanical Engineers 1847-1996 184 (1969) 147-170.

[134] I.V. Roisman, L. Araneo, C. Tropea, Effect of Ambient Pressure on Penetration of

a Diesel Spray, International Journal of Multiphase Flow, 33 (2007) 904-920.

[135] Sauter, Die Grössenbestimmung Der in Gemischnebeln Von

Verbrennungskraftmaschinen Vorhandenen Brennstoffteilchen, VDI-Forschungsheft 279

(1926).

[136] H.K. Suh, H.G. Roh, C.S. Lee, Spray and Combustion Characteristics of

Biodiesel/Diesel Blended Fuel in a Direct Injection Common-Rail Diesel Engine, Journal

of Engineering for Gas Turbines and Power, Transactions of the ASME, 130 (2008) -.

214
[137] B.Y. Guo, D.F. Fletcher, T.a.G. Langrish, Simulation of the Agglomeration in a

Spray Using Lagrangian Particle Tracking, Applied Mathematics Modelling, 28 (2004)

273-290.

[138] T. Takagi, Y.F. Ching, T. Kamimoto, T. Okamoto, Numerical-Simulation of

Evaporation, Ignition and Combustion of Transient Sprays, Combustion Science &

Technology, 75 (1991) 1-12.

[139] G. Gouesbet, A. Berlemont, Eulerian and Lagrangian Approaches for Predicting

the Behaviour of Discrete Particles in Turbulent Flows, Progress in Energy and

Combustion Science, 25 (1999) 133-159.

[140] J. Reveillon, L. Vervisch, Spray Vaporization in Nonpremixed Turbulent

Combustion Modeling: A Single Droplet Model, Combustion and Flame, 121 (2000) 75-

90.

[141] X. Jiang, G.A. Siamas, K. Jagus, T.G. Karayiannis, Physical Modelling and

Advanced Simulations of Gas-Liquid Two-Phase Jet Flows in Atomization and Sprays,

Progress in Energy and Combustion Science, 36 (2010) 131-167.

[142] L.M. Pickett, D.L. Siebers, Soot Formation in Diesel Fuel Jets near the Lift-Off

Length, International Journal of Engine Research, 7 (2006) 103-130.

[143] L.M. Pickett, D.L. Siebers, An Investigation of Diesel Soot Formation Processes

Using Micro-Orifices, Proceedings of the Combustion Institute, 29 (2002) 655-662.

[144] L.M. Pickett, D.L. Siebers, Soot in Diesel Fuel Jets: Effects of Ambient

Temperature, Ambient Density, and Injection Pressure, Combustion and Flame, 138

(2004) 114-135.

215
[145] D.L. Siebers, B.S. Higgins, L.M. Pickett, Flame Lift-Off on Direct-Injection Diesel

Fuel Jets: Oxygen Concentration Effects, SAE paper 2002-01-0890, 111 (2002).

[146] W.A. Sirignano, Fluid-Dynamics of Sprays - 1992 Freeman Scholar Lecture,

Journal of Fluids Engineering, Transactions of the ASME, 115 (1993) 345-378.

[147] X. Lu, D. Han, Z. Huang, Fuel Design and Management for the Control of

Advanced Compression-Ignition Combustion Modes, Progress in Energy and

Combustion Science, In Press, Corrected Proof.

[148] Y. Takeda, K. Nakagome, K. Niimura, Emission Characteristics of Premixed Len

Diesel Combustion with Extremely Early Staged Fuel Injection, SAE paper 961163,

(1996).

[149] K. Nakagome, N. Shimazak, K. Niimura, Combustion and Emission Characteristics

of Premixed Lean Diesel Combustion Engine, SAE paper 970898, (1997).

[150] T. Miyamoto, H. Akagawa, K. Tsujimura, Combustion and Emission

Characteristics of Multiple-Stage Diesel Combustion, SAE paper 980505, (1998).

[151] Y. Nishijima, Y. Asaumi, Y. Aoyagi, Premixed Lean Diesel Combustion (Predic)

Using Impingenment Spray System, SAE paper 2001-01-1892, (2001).

[152] G.K. Lilik, J.M. Herreros, A.L. Boehman, Advanced Combustion Operation in a

Compression Ignition Engine, Energy & Fuels, 23 (2009) 143-150.

[153] K. Akihama, Y. Takatori, K. Inagaki, S. Sasaki, A. Dean, Mechanism of the

Smokeless Rich Diesel Combustion by Reducing Temperature, SAE paper 2001-01-0655,

(2001).

216
[154] P. Mark, B. Musculus, J.E. Dec, L.M. Pickett, C.A. Idicheria, In-Cylinder Imaging

of Conventional and Advanced, Low-Temperature Diesel Combustion, 11th diesel engine

emissions reduction conference, Chicago IL (2005).

[155] T. Fang, R. Coverdill, C. Lee, R. White, Low Temperature Combustion within a

Small Diesel Engine Bore High Speed Direct Injection (Hsdi), SAE paper 2005-01-0919,

(2005).

[156] G.C. Martin, C.J. Mueller, D.M. Milam, M.S. Radovanovic, C.R. Gehrke, Early

Direct-Injection, Low-Temperature Combustion of Diesel Fuel in an Optical Engine

Utilizing a 15-Hole Dual-Row, Anrrow-Included-Angle Nozzle, SAE paper 2008-01-

2400, (2008).

[157] T.G. Fang, R.E. Coverdill, C.F.F. Lee, R.A. While, Low-Sooting Combustion in a

Small-Bore High-Speed Direct-Injection Diesel Engine Using Narrow-Angle Injectors,

Proceedings of the Institution of Mechanical Engineers Part D-Journal of Automobile

Engineering, 222 (2008) 1927-1937.

[158] D. Choi, P.C. Miles, H. Yun, R.D. Reitz, A Parametric Study of Low-Temperature,

Late-Injection Combustion in a Hsdi Diesel Engine, JSME International Journal Series B-

Fluids and Thermal Engineering, 48 (2005) 656-664.

[159] T. Fang, C.-F.F. Lee, Low Sooting Combustion of Narrow-Angle Wall-Guided

Sprays in an Hsdi Diesel Engine with Retarded Injection Timings, Fuel, 90 (2011) 1449-

1456.

[160] T.G. Fang, C.F.F. Lee, Bio-Diesel Effects on Combustion Processes in an Hsdi

Diesel Engine Using Advanced Injection Strategies, Proceedings of the Combustion

Institute, 32 (2009) 2785-2792.

217
[161] R. Oinuma, S. Takuma, T. Koyano, M. Takiguchi, Effects of Post-Injection on

Piston Lubrication in a Common-Rail, Small-Bore Diesel Engine, SAE Trans. 2005-01-

2166, (2005).

[162] T. Fang, Y.C. Lin, T.M. Foong, C.F. Lee, Biodiesel Combustion in an Optical

HSDI Diesel Engine under Low Load Premixed Combustion Conditions, Fuel, 88 (2009)

2154-2162.

[163] M.J. Thornton, T.L. Alleman, J. Luecke, R.L. Mccormick, Impacts of Biodiesel

Fuel Blends Oil Dilution on Light-Duty Diesel Engine Operation, SAE paper 2009-01-

1790, (2009).

[164] M. Morcos, G. Parsons, W. Hartgers, M. Boons, S. Roby, F. Lauterwasser,

Detection Methods for Accurate Measurements of the Fame Biodiesel Content in Used

Crankcase Engine Oil, SAE paper 2009-01-2661, (2009).

[165] H.L. Fang, S.D. Whitacre, E.S. Yamaguchi, M. Boons, Biodiesel Impact on Wear

Protection of Engine Oils, SAE paper 2007-01-4141, (2007).

[166] K.M. Richard, S. Mctavish, Impact of Biodiesel on Lubricant Corrosion

Performance, SAE paper 2009-01-2660, (2009).

[167] K.O. Lee, J. Zhu, S. Ciatti, A. Yozgatligil, M.Y. Choi, Sizes, Graphitic Structures

and Fractal Geometry of Light-Duty Diesel Engine Particulates, SAE paper 2003-01-

3169, (2003).

[168] M. Lapuerta, R. Ballesteros, F.J. Martos, A Method to Determine the Fractal

Dimension of Diesel Soot Agglomerates, Journal of Colloid and Interface Science, 303

(2006) 149-158.

[169] B. Mandelbrot, The Fractal Geometry of Nature, W.H.Freeman and Co., 1982.

218
[170] M. Lapuerta, R. Ballesteros, F.J. Martos, The Effect of Diesel Engine Conditions

on the Size and Morphology of Soot Particles, International Journal of Vehicle Design,

50 (2009) 91-106.

[171] U.O. Koylu, G.M. Faeth, T.L. Farias, M.G. Carvalho, Fractal and Projected

Structure Properties of Soot Aggregates, Combustion and Flame, 100 (1995) 621-633.

[172] J.O. Muller, D.S. Su, R.E. Jentoft, J. Krohnert, F.C. Jentoft, R. Schlogl,

Morphology-Controlled Reactivity of Carbonaceous Materials Towards Oxidation,

Catalysis Today, 102 (2005) 259-265.

[173] A. Tsolakis, Effects on Particle Size Distribution from the Diesel Engine Operating

on Rme-Biodiesel with EGR, Energy and Fuels, 20 (2006) 1418-1424.

[174] A. Neer, U.O. Koylu, Effect of Operating Conditions on the Size, Morphology, and

Concentration of Submicrometer Particulates Emitted from a Diesel Engine, Combustion

and Flame, 146 (2006) 142-154.

[175] A. Soewono, S. Rogak, Morphology and Raman Spectra of Engine-Emitted

Particulates, Aerosol Science and Technology, 45 (2011) 1206-1216.

[176] U. Mathis, M. Mohr, R. Kaegi, A. Bertola, K. Boulouchos, Influence of Diesel

Engine Combustion Parametes on Primary Soot Particle Diameter, Environmental

Science & Technology, 39 (2005) 1887-1892.

[177] M. Lapuerta, F.J. Martos, J.M. Herreros, Effect of Engine Operating Conditions on

the Size of Primary Particles Composing Diesel Soot Agglomerates, Journal of Aerosol

Science, 38 (2007) 455-466.

219
[178] A. Smekens, R.H.M. Godoi, P. Berghmans, R.E. Van Grieken, Characterisation of

Soot Emitted by Domestic Heating, Aircraft and Cars Using Diesel or Biodiesel, Journal

of Atmospheric Chemistry, 52 (2005) 45-62.

[179] J. Zhu, K.O. Lee, A. Yozgatligil, M.Y. Choi, Effects of Engine Operating

Conditions on Morphology, Microstructure, and Fractal Geometry of Light-Duty Diesel

Engine Particulates, Proceedings of the Combustion Institute, 30 (2005) 2781-2789.

[180] M.M. Maricq, Physical and Chemical Comparison of Soot in Hydrocarbon and

Biodiesel Fuel Diffusion Flames: A Study of Model and Commercial Fuels, Combustion

and Flame, 158 (2011) 105-116.

[181] J.H. Song, M. Alam, A.L. Boehman, Impact of Alternative Fuels on Soot

Properties and DPF Regeneration, Combustion Science & Technology, 179 (2007) 1991-

2037.

[182] R.L. Vander Wal, C.J. Mueller, Initial Investigation of Effects of Fuel Oxygenation

on Nanostructure of Soot from a Direct-Injection Diesel Engine, Energy & Fuels, 20

(2006) 2364-2369.

[183] A. Braun, F.E. Huggins, N. Shah, Y. Chen, S. Wirick, S.B. Mun, C. Jacobsen, G.P.

Huffman, Advantages of Soft X-Ray Absorption over TEM-EELS for Solid Carbon

Studies - a Comparative Study on Diesel Soot with EELS and NEXAFS, Carbon, 43

(2005) 117-124.

[184] K. Al-Qurashi, A.L. Boehman, Impact of Exhaust Gas Recirculation (EGR) on the

Oxidative Reactivity of Diesel Engine Soot, Combustion and Flame, 155 (2008) 675-695.

[185] K. Yehliu, Impact of Fuel Formulation and Engine Operating Parameters on the

Nanostructure and Reactivity of Diesel Soot, in, The Pennsylvania State University, 2010.

220
[186] Z. Li, C.L. Song, J.O. Song, G. Lv, S.R. Dong, Z. Zhao, Evolution of the

Nanostructure, Fractal Dimension and Size of in-Cylinder Soot During Diesel

Combustion Process, Combustion and Flame, 158 (2011) 1624-1630.

[187] R.L. Vander Wal, A.J. Tomasek, Soot Nanostructure: Dependence Upon Synthesis

Conditions, Combustion and Flame, 136 (2004) 129-140.

[188] R.L. Vander Wal, A.J. Tomasek, Soot Oxidation: Dependence Upon Initial

Nanostructure, Combustion and Flame, 134 (2003) 1-9.

[189] H.J. Seong, A.L. Boehman, Impact of Intake Oxygen Enrichment on Oxidative

Reactivity and Properties of Diesel Soot, Energy & Fuels, 25 (2011) 602-616.

[190] H.J. Seong, Impact of Oxygen Enrichment on Soot Properties and Soot Oxidative

Reactivity, Ph.D Dissertation, The Pennsylvania State University, 2010.

[191] Ö.L. Gülder, Soot Formation in Laminar Diffusion Flames at Elevated

Temperatures, Combustion and Flame, 88 (1992) 75-82.

[192] B.M. Kumfer, S.A. Skeen, R. Chen, R.L. Axelbaum, Measurement and Analysis of

Soot Inception Limits of Oxygen-Enriched Coflow Flames, Combustion and Flame, 147

(2006) 233-242.

[193] D.X. Du, R.L. Axelbaum, C.K. Law, Soot Formation in Strained Diffusion Flames

with Gaseous Additives, Combustion and Flame, 102 (1995) 11-20.

[194] R.H. Hurt, G.P. Crawford, H.S. Shim, Equilibrium Nanostructure of Primary Soot

Particles, Proceedings of the Combustion Institute, 28 (2000) 2539-2546.

[195] H.X. Chen, R.A. Dobbins, Crystallogenesis of Particles Formed in Hydrocarbon

Combustion, Combustion Science & Technology, 159 (2000) 109-128.

221
[196] R.L. Vander Wal, A TEM Methodology for the Study of Soot Particle Structure,

Combustion Science & Technology, 126 (1997) 333-357.

[197] A. Senatore, M. Cardone, V. Rocco, M.V. Prati, A Comparative Analysis of

Combustion Process in DI Diesel Engine Fueled with Biodiesel and Diesel Fuel, SAE

paper 2000-01-0691, (2000).

[198] A. Shaheed, E. Swain, Combustion Analysis of Coconut Oil and Its Methyl Esters

in a Diesel Engine, IMECHE conference transactions, (1998).

[199] M. Canakci, Performance and Emissions Characteristics of Biodiesel from Soybean

Oil, IMECHE conference transactions, (2005).

[200] M. Lapuerta, J.R. Fernandez, J.R. Agudelo, Diesel Particulate Emissions from

Used Cooking Oil Biodiesel, Bioresource Technology, 99 (2008) 731-740.

[201] A. Monyem, G. Van, Jh, The Effect of Biodiesel Oxidation on Engine Performance

and Emissions, Biomass Bioenerg, 20 (2001) 317-325.

[202] M.N. Nabi, Theoretical Investigation of Engine Thermal Efficiency, Adiabatic

Flame Temperature, Nox Emission and Combustion-Related Parameters for Different

Oxygenated Fuels, Applied Thermal Engineering, 30 (2010) 839-844.

[203] C. Kaplan, R. Arslan, A.Surmen, Performance Characteristics of Sunflower Methyl

Esters as Biodiesel, Energy Sources, Part A 28 (2006) 751-755.

[204] T.J. Jacobs, J.A. Bittle, J.K. Younger, Biodiesel Effects on Influencing Parameters

of Brake Fuel Conversion Efficiency in a Medium Duty Diesel Engine, Journal of

Engineering for Gas Turbines and Power, Transactions of the ASME, 132 (2010).

222
[205] L. Gvidonas, S. Stasys, The Effect of Rapeseed Oil Methyl Easter on Direct

Injection Diesel Engine Performance and Exhaust Emissions., Energy Convers Manage,

47 (2006) 1954-1967.

[206] A. Kowalewicz, Eco-Diesel Engine Fuelled with Rapeseed Oil Methyl Ester and

Ethanol. Part 2: Comparison of Emissions and Efficiency for Two Base Fuels: Diesel

Fuel and Ester, Proceedings of the Institution of Mechanical Engineers Part D, Journal of

Automobile Engineering, 220 (2006) 1275-1282.

[207] Y. Aoyagi, E. Kunishima, Y. Asaumi, Y. Aihara, M. Odaka, Y. Goto, Diesel

Combustion and Emission Using High Boost and High Injection Pressure in a Single

Cylinder Engine - (Effects of Boost Pressure and Timing Retardation on Thermal

Efficiency and Exhaust Emissions), JSME International Journal Series B, Fluids and

Thermal Engineering, 48 (2005) 648-655.

[208] A. Ramesh, J.N. Reddy, Parametric Studies for Improving the Performance of a

Jatropha Oil-Fuelled Compression Ignition Engine, Renewable Energy, 31 (2006) 1994-

2016.

[209] T. Shudo, Y. Nakajima, T. Futakuchi, Thermal Efficiency Analysis in a Hydrogen

Premixed Combustion Engine, JSAE Review, 21 (2000) 177-182.

[210] K. Nagase, Y. Fukushima, A Fuel-Injection System for Diesel-Engines by Injection

Pressure Control .1. Theories, Influence Upon Theoretical Thermal Efficiency and the

Result of Basic Experiments, Bulletin of the Japan Society of Mechanical Engineers, 29

(1986) 1788-1794.

223
[211] M. Zheng, U. Asad, G.T. Reader, Y.Y. Tan, M.P. Wang, Energy Efficiency

Improvement Strategies for a Diesel Engine in Low-Temperature Combustion,

International Journal of Energy Resourse, 33 (2009) 8-28.

[212] T. Fang, Y.C. Lin, T.M. Foong, C.F. Lee, Reducing Nox Emissions from a

Biodiesel-Fueled Engine by Use of Low-Temperature Combustion, Environmental

Science & Technology, 42 (2008) 8865-8870.

[213] H. Hiroyasu, T. Kadota, M. Arai, Development and Use of a Spray Combustion

Modeling to Predict Diesel-Engine Efficiency and Pollutant Emissions .1. Combustion

Modeling, Bulletin of the Japan Society of Mechanical Engineers, 26 (1983) 569-575.

[214] H. Hiroyasu, T. Kadota, M. Arai, Development and Use of a Spray Combustion

Modeling to Predict Diesel-Engine Efficiency and Pollutant Emissions .2.

Computational-Procedure and Parametric Study, Bulletin of the Japan Society of

Mechanical Engineers, 26 (1983) 576-583.

[215] H. Hiroyasu, O. Furukawa, M. Arai, S. Iida, H. Motonaga, Development and Use

of a Spray Combustion Modeling to Predict Diesel-Engine Efficiency and Pollutant

Emissions .3. An Analysis by the Method of System-Model Transformation, Bulletin of

the Japan Society of Mechanical Engineers, 26 (1983) 584-591.

[216] T.J. Williams, N.D. Whitehouse, Investigation into Some Aspects of the

Computation of Diesel Engine Combustion, Proceedings of the institution of Mechanical

Engineers 190 (1976) 467 - 476.

[217] NIST, Chemistry Web Boook, http://webbook.nist.gov/chemistry, (2011).

[218] Y. Zhang, Y. Yang, A.L. Boehman, Premixed Ignition Behavior of C-9 Fatty Acid

Esters: A Motored Engine Study, Combustion and Flame, 156 (2009) 1202-1213.

224
[219] A.C. Ferrari, J. Robertson, Interpretation of Raman Spectra of Disordered and

Amorphous Carbon, Physical Review B, 61 (2000) 14095-14107.

[220] T. Jawhari, A. Roid, J. Casado, Raman-Spectroscopic Characterization of Some

Commercially Available Carbon-Black Materials, Carbon, 33 (1995) 1561-1565.

[221] C.S. Lee, S.W. Park, An Experimental and Numerical Study on Fuel Atomization

Characteristics of High-Pressure Diesel Injection Sprays, Fuel, 81 (2002) 2417-2423.

[222] C.W. Yu, S. Bari, A. Ameen, A Comparison of Combustion Characteristics of

Waste Cooking Oil with Diesel as Fuel in a Direct Injection Diesel Engine, Proceedings

of the Institution of Mechanical Engineers Part D, Journal of Automobile Engineering,

216 (2002) 237-243.

[223] M. Munoz, F. Moreno, J. Morea, Emissions of an Automobile Diesel Engine

Fueled with Sunflower Methylester, Transactions of the ASAE, 47 (2004) 5-11.

[224] U.S. DOE EERE, Properties of Fuels, online resources,

http://www.afdc.energy.gov/afdc/pdfs/fueltable.pdf, (2009).

[225] Argonne National Laboratory, Greet Transportation Fuel Cycle Analysis Model,

Greet 1.8, http://www.transportation.anl.gov/modeling_simulation/GREET/index.html,

(2008).

[226] C.A. Idicheria, L.M. Pickett, Quantitative Mixing Measurements in a Vaporizing

Diesel Spray by Rayleigh Imaging, SAE paper 2007-01-0647, (2007).

[227] D.L. Siebers, B.S. Higgins, Flame Lift-Off on Direct Injection Diesel under

Quiescent Conditions, SAE paper 2001-01-0530, (2001).

[228] H.J. Curran, P. Gaffuri, W.J. Pitz, C.K. Westbrook, A Comprehensive Modeling

Study of N-Heptane Oxidation, Combustion and Flame, 114 (1998) 149-177.

225
[229] J.P. Szybist, A.L. Boehman, D.C. Haworth, H. Koga, Premixed Ignition Behavior

of Alternative Diesel Fuel-Relevant Compounds in a Motored Engine Experiment,

Combustion and Flame, 149 (2007) 112-128.

[230] D. Assanis, Z. Filipi, S. Fiveland, M. Syrimis, A Predictive Ignition Delay

Correlation under Steady-State and Transient Operation of a Direct Injection Diesel

Engine, Journal of Engineering for Gas Turbines and Power, transactions of the ASME,

125 (2003) 450-457.

[231] M. Sjoberg, J.E. Dec, An Investigation into Lowest Acceptable Combustion

Temperatures for Hydrocarbon Fuels in HCCI Engines, in: 30th International

Symposium on Combustion, Chicago, IL, 2004, pp. 2719-2726.

[232] S.I. Sandler, Chemical, Biochemical, and Engineering Thermodynamics, 4e, John

Wiley & Sons, Inc, U.S.A., (2006) 319.

[233] W. Yuan, A.C. Hansen, Q. Zhang, Vapor Pressure and Normal Boiling Point

Predictions for Pure Methyl Esters and Biodiesel Fuels, Fuel, 84 (2005) 943-950.

[234] J.W. Goodrum, Volatility and Boiling Points of Biodiesel from Vegetable Oils and

Tallow, Biomass & Bioenergy, 22 (2002) 205-211.

[235] X.Y. Zou, J.M. Shaw, Phase Behavior of Hydrocarbon Mixtures, in: Encyclopedia

of Chemical Processing, Taylor & Francis, 2006, pp. 2067-2075.

[236] K.S. Sanvordenker, Engine Friction-Effect of Oil Viscosity and Break-in on

Cranking Resistance, SAE paper 841095, (1984).

[237] P. Meakin, B. Donn, G.W. Mulholland, Collisions between Point Masses and

Fractal Aggregates, Langmuir, 5 (1989) 510-518.

226
[238] E.F. Mikhailov, S.S. Vlasenko, A.A. Kiselev, T.I. Ryshkevitch, Restructuring of

Soot Particles: Experimental Study, Journal of Aerosol Science, 27 (1996) 771-712.

[239] J.H. Song, Effect of Fuel Formulation on Soot Properties and Regeneration of

Diesel Particulate Filters, in: Fuel Science, The Pennsylvania State University, 2005.

[240] L.R. Radovic, P.L. Walker, R.G. Jenkins, Importance of Carbon Active-Sites in the

Gasification of Coal Chars, Fuel, 62 (1983) 849-856.

[241] R.G. Jenkins, S.P. Nandi, P.L. Walker, Reactivity of Heat-Treated Coals in Air at

500 Degrees C, Fuel, 52 (1973) 288-293.

[242] K. Al-Qurashi, A.D. Lueking, A.L. Boehman, The Deconvolution of the Thermal,

Dilution, and Chemical Effects of Exhaust Gas Recirculation (EGR) on the Reactivity of

Engine and Flame Soot, Combustion and Flame, 158 (2011) 1696-1704.

[243] A. Arenillas, F. Rubiera, C. Pevida, C.O. Ania, J.J. Pis, Relationship between

Structure and Reactivity of Carbonaceous Materials, Journal of Thermal Analysis and

Calorimetry, 76 (2004) 593-602.

[244] A. Cheng, P. Harriott, Kinetics of Oxidation and Chemisorption of Oxygen for

Porous Carbons with High Surface-Area, Carbon, 24 (1986) 143-150.

[245] P. Gilot, F. Bonnefoy, F. Marcuccilli, G. Prado, Determination of Kinetic Data for

Soot Oxidation - Modeling of Competition between Oxygen Diffusion and Reaction

During Thermogravimetric Analysis, Combustion and Flame, 95 (1993) 87-100.

[246] A.L. Boehman, J. Song, C.H. Jeon, Impacts of Oxygen Diffusion on the

Combustion Rate of in-Bed Soot Particles, Energy & Fuels, 24 (2010) 2418-2428.

[247] N.R. Laine, F.J. Vastola, P.L. Walker, The Importance of Active Surface Area in

Carbon Oxygen Reaction, Journal of Physical Chemistry, 67 (1963) 1461-1475.

227
[248] A. Agudelo, J.R. Agudelo, P. Benjumea, Diagnóstico De La Combustión De

Biocombustibles En Motores (Combustion Diagnosis of Biofuels in Engines), Imprenta

Universidad de Antioquia, 2007.

[249] G. Woschni, Universally Applicable Equation for the Instantaneous Heat Transfer

Coefficient in the Internal Combustion Engine, SAE paper 670930, (1967).

[250] B.S. Han, Y.J. Chung, Y.J. Kwon, S. Lee, Empirical Forrmula for Instantaneous

Heat Transfer Coefficient in Spark Ignition Engines, SAE paper 972995, (1997).

[251] O. Armas, Diagnóstico Experimental Del Proceso De Combustión En Motores

Diesel De Inyección Directa., in, Universidad Politécnica de Valencia, España, 1988.

[252] C.D. Rakopoulos, D.T. Hountalas, A.P. Koutroubousis, T.C. Zannis, Application

and Evaluation of Detailed Friction Model on a Di Diesel Engine with Extreme High

Peak Combustion Pressure, SAE paper 2002-01-0068, (2002).

[253] M. Lapuerta, O. Armas, S. Molina, Study of the Compression Cycle of a

Reciprocating Engine through the Polytropic Coefficient, Applied Thermal Engineering,

23 (2003) 313-323.

[254] [254] W.T. Lyn, Calculations of the Effect of Rate of Heat Release on the Shape of

Cylinder-pressure Diagram and Cycle Efficiency, Proceedings of the Institution of

Mechanical Engineers, (1960).

[255] S.H. Yoon, C.S. Lee, Lean Combustion and Emission Characteristics of Bioethanol

and Its Blends in a Spark Ignition (SI) Engine, Energy & Fuels, 25 (2011) 3484-3492.

228
Appendix A: Matlab code
Oxygen equivalence ratio field calculation, LLP.m

clear all;

nfC = 16.15; % fuel carbon atom number, averaged, diesel 14.48, B40 16.15
**************************************************Biodiesel
nfH = 30.38; % fuel hydrogen atom number, averaged, diesel 26.23; B40
30.38***************************************************Biodiesel
nfO = 0.8;% fuel oxygen atom number, averaged,diesel 0, B40
0.8;***************************************************Biodiesel

MWf = nfC*12+nfH*1+nfO*16; % averaged molecular weight (g/mol) of fuel


MWa = 28.97; % averaged molecular weight (g/mol) of air

Tg = 1172; % cylinder gas temperature at injection (K) 355 CAD


***********************************************
Pg = 112; % cylinder pressure at injection (Bar) 355 CAD
**************************************
VI = 49.8; % cylinder volume (cm3) 51.05 at injection 355 CAD, 52.93 at 353, 55.43 at 351,
49.80 at 3B, 49.17 at 1B ***************************,

Ru = 8.314472; % gas constant (J/K/mol)


rhoa_1 = 1.14; % density of air (kg/m3) at 1 bar at ~40 C
Pg_180 = 2.3; % boost pressure (bar), or cylinder pressure at 180 CAD

Pf = input('input Pf = '); % fuel injection pressure (bar)


***************************************************

Tbst = 60; % boost temperature (C)


rhoa_180 = Pg_180*rhoa_1*(40+273)/(Tbst+273); % density (kg/m3)of air at 180 CAD
Vb = 844.34; % volume (cm3) at bottom dead center
rhoa_I = rhoa_180*Vb/VI; % density of air (kg/m3) at injection
rhog = rhoa_I; % assume the density by fuel injection is negelectable, density of environmental
gas at injection

rhof = 855; %$ density of fuel (kg/m3), ULSD is 842, B100 is 877, B40 is 855
******************************************Biodiesel

dn = 0.111; % diameter of orifice (mm)


Yf_1 = 1; % fuel concentration at injector
XO2_2 = 0.21; % oxygen volume concentration at injector, 0.21 is the concentration of oxygen in
air
YO2_2 = 0.232; % oxygen mass concentration at injector
nf_st = 1; % fuel at stoichiometrix for nf_st*fuel + mO2_st*O2 = 0
mO2_st = nf_st*(nfC+nfH/4-nfO/2); % oxygen number at stoichiometric
mu = mO2_st*32/nf_st/MWf; % mu*Yf + YO2 = 0, reaction parameter
mf_st = nf_st*MWf; % mass of fuel at stoichiometric (g)
ma_st = mO2_st*32/YO2_2; % mass of air at stoichiemetric (g)
AFst = ma_st/mf_st; % stoichiometric air/fuel ratio
Zst = (1+mu*Yf_1/YO2_2)^-1; % stoichiometric mixture fraction
Ca = 0.91+(0.86-0.91)/(1380-720)*(Pf-720); % area contraction coeffecient Ca =0.82 for dn =
0.18, Ca =0.86 for dn = 0.1, 0.071, 0.05 mm @ 1400bar injection

229
Cd = 0.80; % discharged coefficient
Cv = Cd/Ca; % Bernolli's equation coefficient
Uf = Cv*sqrt(2*(Pf-Pg)*10^5/rhof); % velocity of injected fuel at orifice (m/s), Bernoulli's
equation, pressure is transferred to Pa, SI unit
c = 0.255; % speading angle coefficient interpolated from Siebers 1999 SAE
a = 0.75; % Constant from Siebers 1996 SAE
Tan_half_theta = c*((rhog/rhof)^0.19-0.0043*sqrt(rhof/rhog));% spray half angle
x_plus = sqrt(rhof/rhog)*dn*sqrt(Ca)/a/Tan_half_theta; % penetration length (mm), same unit as
dn
H = 7.04*10^8*Tg^-3.74*rhog^-0.85*(dn*10^3)^0.34*Uf*Zst; % calculate lift-off length, also SI
unit, except from that unit of dn is "um"
x = [x_plus:2:50 1000*H]; % buliding matrix for x
phi_x = 2.*AFst./(sqrt(1+16.*(x./x_plus).^2)-1);% equivalence ratio as f(x)
phi_omega_f_B40 = (2*nfC+0.5*nfH)/nfO; % oxygen equavalence ratio of the fuel alone, only
for biodiesel, here is B40
phi_x_omega_B40 = phi_x./(1-(1-phi_x)/phi_omega_f_B40); % oxygen equivalence ratio matrix
for B40
phi_x_omega_ULSD = phi_x; % oxygen equivalence ratio matrix for ULSD
r = -3:0.5:3; % building spray diameter matrix (mm)
num_initial = 1; %define the initial index
num = num_initial; % initial the index
numr = size(r); % count the number of row & column of r
numx = size(x); % count the number of row & column of x
numrr = numr(1,2); % total element in r
numxx = numx(1,2); % total element in x
phi_omega_x_r_ULSD = zeros(numrr, numxx); % initialized the phi_omega_x_r_ULSD
phi_omega_x_r_B40 = zeros(numrr,numxx); % initialized the phi_omega_x_r_B40
while num<=numrr
phi_omega_x_sr = 1.3.*phi_x_omega_ULSD.*exp((r(1,num)./x).^2*log(0.08)/Tan_half_theta^2); %
building oxygen equivalence ratio matrix based on f(x,r)
phi_omega_x_r_ULSD(num,:) = phi_omega_x_sr; % replace row of num of phi_omega_x_r with
data
num = num +1; %index num
end
num = num_initial; % reinitialize num index
while num<=numrr
phi_omega_x_sr = 1.3.*phi_x_omega_B40.*exp((r(1,num)./x).^2*log(0.08)/Tan_half_theta^2); %
building oxygen equivalence ratio matrix based on f(x,r)
phi_omega_x_r_B40(num,:) = phi_omega_x_sr; % replace row of num of phi_omega_x_r with
data
num = num +1; %index num
end
disp 'lift-off length';
disp(H);

Efficiency simulation, efficiency.m

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
% efficiency calculation for limited pressure cycle
% clear all;
% SOCtest = 0;
% C1 = 1; %control the ratio of premixed and diffusion burn
% C2 = 1;

230
gamma = 1.367;
alpha = [1:0.1:2]';
beta = 1:10;
rc = 17.2;
eta_ideal = alpha*beta; % initialize eta_otto matrix, efficiency in ideal cycle

for j = 1:size(alpha,1)
for i = 1:size(beta,2)
eta_ideal(j,i) = 1 - 1/rc^(gamma-1)*(alpha(j)*beta(i)^gamma-1)/(alpha(j)*gamma*(beta(i)-

1)+alpha(j)-1);
end
end
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
% this model assumes no leakage, but it can be easily added

theta = 0:0.1:719.9; % crank angle


bore = 9.82; % cm
Stroke = 10.5; % cm
CR = 17.2; % compression ratio
Vc = 49.089340; %cm3
lcr = 17.6; % cm, connect rod length
rc = 5.25; % cm, crank radius
rpm = 1500; % speed
omega = rpm/60*360; % dtheta/dt, degree per second
ncy = 8; % number of cylinder
dtheta = 0.1; % delta theta
V = Vc + pi*bore^2/4*(lcr + rc - (rc.*cos(theta*pi/180)+(lcr^2-rc^2.*sin(theta*pi/180).^2).^0.5));

% chamber volume with theta


A = pi*bore*(lcr + rc - (rc.*cos(theta*pi/180)+(lcr^2-rc^2.*sin(theta*pi/180).^2).^0.5)); %

instantaneous cylinder area, cm2


Ru = 8.314472; % gas constant, J/mol/K
IVC = 200; % timing of inlet valve close
EVO = 520; % timing of exhaust valve open, CA
Y = 0.7; % 0.027 is the mf/(mf+ma) value, 0.7 seems to give better match, but no sense in

temperature

pintake = 2.52835; % intake pressure,


bar %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
pexhaust = 1.5; % exhaust pressure, here we first assume no pumping loss
Tintake = 40 + 273; % intake temperature, K

gamma = fgamma(Tintake).*theta./theta; % initialize a gamma


qw = 0.*theta;

231
Sp = 2*Stroke*0.01*rpm/60; % average piston velocity, m/s
wmotor = 0.*theta; % the Woschni velocity, initialize, m/s, only for motor

for i = 1:size(theta,2)
if theta(i) < IVC
wmotor(i) = 6.18*Sp;
elseif theta(i) == IVC
n_IVC = i; % record the index of inlet valve close
wmotor(i) = 6.18*Sp;
% elseif theta(i) <360
% wmotor(i) = 2.28*Sp;
elseif theta(i) < EVO
wmotor(i) = 2.28*Sp;
elseif theta(i) == EVO
n_EVO = i; % index of exhaust value close
wmotor(i) = 2.28*Sp;
else
wmotor(i) = 6.18*Sp;
end
end

wcomb = wmotor; % initialize a combution w factor

%corr = 1.18/2.3789; % a correction for the MAF (don't have justification


%yet), for 1801
corr = 0.0407/0.0805; % for IVC (0.0671)
nintake = pintake*10^5*V(n_IVC)*10^-6/Ru/Tintake*corr;
Mwair = 28.97; % g/mol
Tw = 90+273; % cylinder wall temperature, K, coolant temp
MAF = Mwair*nintake; % g/cycle

for i = 1:size(theta,2)
if i <= n_IVC
pmotor(i) = pintake;
Tmotor(i) = Tintake;
elseif i<= n_EVO
% pmotor(i) = pmotor(i-1)*V(i-1)^gamma(i)/V(i)^gamma(i);
% Tmotor(i) = pmotor(i)*10^5*V(i)*10^-6/Ru/nintake;
% gamma(i) = fgamma(Tmotor(i));
% while abs(pmotor(i-1)*V(i-1)^gamma(i)/V(i)^gamma(i) - pmotor(i)) > 10^-3
% pmotor(i) = pmotor(i-1)*V(i-1)^gamma(i)/V(i)^gamma(i);
% Tmotor(i) = pmotor(i)*10^5*V(i)*10^-6/Ru/nintake;
% gamma(i) = fgamma(Tmotor(i));
% end % (adiabatic) process
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
pinter = pmotor(i-1);
Tmotor(i) = pmotor(i-1)*10^5*V(i)*10^-6/Ru/nintake;
% Tmotor(i) = ftemp(pmotor(i-1),V(i),MAF);
gamma(i) = fgamma(Tmotor(i));
pmotor(i) = (V(i)*pmotor(i-1) - 10*dtheta/omega*(gamma(i)-1)*dtheta*fconvection2

(pinter,Tmotor(i),wmotor(i),A(i),0))/(gamma(i)*(V(i)-V(i-1))+V(i));
while abs(pmotor(i) - (V(i)*pmotor(i-1) - 10*dtheta/omega*(gamma(i)-1)

232
*dtheta*fconvection2(pinter,Tmotor(i),wmotor(i),A(i),0))/(gamma(i)*(V(i)-V(i-1))+V(i))) >10^-3
Tmotor(i) = pmotor(i)*10^5*V(i)*10^-6/Ru/nintake;
% Tmotor(i) = ftemp(pmotor(i),V(i),MAF);
gamma(i) = fgamma(Tmotor(i));
pmotor(i) = (V(i)*pmotor(i-1) - 10*dtheta/omega*(gamma(i)-1)*dtheta*fconvection2

(pinter,Tmotor(i),wmotor(i),A(i),0))/(gamma(i)*(V(i)-V(i-1))+V(i));
pinter = pmotor(i);
end % with heat transfer

else
pmotor(i) = pmotor(n_EVO);
Tmotor(i) = Tmotor(n_EVO);

end
end

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%
% synchronize TDC with the maximum cylinder pressure

pmotor_old = pmotor; % record pmotor before a correction of TDC with maximum cylinder
pressure
Tmotor_old = Tmotor;
dpmotor = theta./theta; % initialized dqmotor
for i = 3400:3800
dpmotor(i) = (pmotor(i+1)-pmotor(i-1))/(theta(i+1)-theta(i-1));
end

for i = 3400:3800
if abs(dpmotor(i)) == min(abs(dpmotor))
n_TDCcorr = i; % the index the TDC correction
end
end

delta_n = 3601 - n_TDCcorr; % the correction

for i = delta_n:(size(theta,2) - delta_n)


pmotor(i) = pmotor_old(i-delta_n+1);
Tmotor(i) = Tmotor_old(i-delta_n+1);
end

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%

% arbitrary heat release

% for Gaussian peak

lgauss = 365.18 + SOCtest; % location, SOC


wgauss = 1.1683*1;
hgauss = 125.87*1;

% for ExpConvExp test


lece = 363.32 + SOCtest; % location, crank angle, SOC
hece = 619.69; % height, J/deg

233
k1 = 0.12822;
k2 = 0.77745;

ng = 1; % number of gaussian heat release


nece = 1; % recording the heat release period, number of ExpConvExp heat release

HRCAgauss = 0.*theta;
HRCAece = HRCAgauss;
HRCAgauss_index = HRCAgauss;
HRCAece_index = HRCAece;

for i = 1:size(theta,2)
fece(i) = hece/k2*fexpconvexp(theta(i) - lece ,k1,k2); % result of ExpConvExp
fgau(i) = hgauss*fgauss(theta(i),lgauss,wgauss); %result of gauss
if fgau(i) >= 0.1
HRCAgauss(i) = theta(i);
HRCAgauss_index(i) = 1;
ng = ng+1;
end
if fece(i) > 0.1
HRCAece(i) = theta(i);
HRCAece_index(i)=1;
nece = nece + 1;
end
end

AHR = C1*fgau + C2*fece; % total apparent heat release, J/Deg.


AHR_index = 1*HRCAgauss_index+C2*HRCAece_index;

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%

% calculate teh combustion

pcomb = pmotor; % initialize a pressure with combustion


Tcomb = Tmotor; % initialize a temperature with combustion

for i = 1:size(theta,2)
if AHR_index(i) > 0

Tcomb(i) = pcomb(i-1)*10^5*V(i)*10^-6/Ru/nintake;
gamma(i) = fgamma(Tcomb(i));
pcomb(i) = (V(i)*pcomb(i-1) + 10*AHR(i)*(gamma(i)-1)*dtheta)/(gamma(i)*(V(i)-V(i-
1))+V

(i)); % multiply 10 for unit conversion


while abs(pcomb(i) - ((V(i)*pcomb(i-1) + 10*AHR(i)*(gamma(i)-
1)*dtheta)/(gamma(i)*(V(i)-V

(i-1))+V(i)))) >10^-5

234
pcomb(i) = (V(i)*pcomb(i-1) + 10*AHR(i)*(gamma(i)-1)*dtheta)/(gamma(i)*(V(i)-V(i-

1))+V(i));
Tcomb(i) = pcomb(i)*10^5*V(i)*10^-6/Ru/nintake;
gamma(i) = fgamma(Tcomb(i));
end % to obtain the right gamma value
wcomb(i) = wmotor(i) + 3.24*10^-
3*Stroke*pi*bore^2/4/V(n_IVC)*Tintake/pintake*(pcomb(i)-

pmotor(i));
elseif theta(i) >= lgauss
if i<=n_EVO
wcomb(i) = wmotor(i) + 3.24*10^-
3*Stroke*pi*bore^2/4/V(n_IVC)*Tintake/pintake*(pcomb

(i-1)-pmotor(i));
pinter = pcomb(i-1);
Tcomb(i) = pcomb(i-1)*10^5*V(i)*10^-6/Ru/nintake;
gamma(i) = fgamma(Tcomb(i));
pcomb(i) = (V(i)*pcomb(i-1) - 10*dtheta/omega*(gamma(i)-1)*dtheta*fconvection2

(pinter,Tcomb(i),wcomb(i),A(i),Y))/(gamma(i)*(V(i)-V(i-1))+V(i));
while abs(pcomb(i) - (V(i)*pcomb(i-1) - 10*dtheta/omega*(gamma(i)-1)

*dtheta*fconvection2(pinter,Tcomb(i),wcomb(i),A(i),Y))/(gamma(i)*(V(i)-V(i-1))+V(i))) >10^-3
Tcomb(i) = pmotor(i)*10^5*V(i)*10^-6/Ru/nintake;
gamma(i) = fgamma(Tmotor(i));
wcomb(i) = wmotor(i) + 3.24*10^-3*Stroke*pi*bore^2/4/V(n_IVC)*Tintake/pintake*

(pcomb(i)-pmotor(i));
pcomb(i) = (V(i)*pcomb(i-1) - 10*dtheta/omega*(gamma(i)-1)*dtheta*fconvection2

(pinter,Tcomb(i),wcomb(i),A(i),Y))/(gamma(i)*(V(i)-V(i-1))+V(i));
pinter = pcomb(i);
end % with heat transfer
end
end
if i > n_EVO
pcomb(i) = pcomb(n_EVO);
Tcomb(i) = Tcomb(n_EVO-1);
end
end

qt = 0.*theta; % initialize a total energy addtion (essentially dQ/dtheta)


for i = 1: size(theta,2)
if AHR_index(i)>0
qt(i) = AHR(i) + dtheta/omega*fconvection2(pcomb(i),Tcomb(i),wcomb(i),A(i),0.027)/10; %

total energy addition, devide by 10 to convert the unit to J/deg.


end
end

CHR = 0; % cumulative heat release, J


tQ = 0; % total heat addition

for i = 1:size(theta,2)

235
if AHR_index(i)>0
CHR = CHR + (AHR(i)+AHR(i+1))/2*dtheta;
tQ = tQ + (qt(i)+qt(i+1))/2*dtheta;
end
end

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%
% indicated power

dV = 0.*V; % initialize a V, dV/dtheta


dV(1) = (V(2)-V(1))/dtheta;
dV(size(theta,2)) = (V(size(theta,2))-V(size(theta,2)-1))/dtheta;
for i = 2:size(theta,2)-1
dV(i) = (V(i+1)-V(i-1))/2/dtheta;
end

% pengine = xlsread('p_100MPa_25%'); % result engine pressure trace data, 25%, 1B, 100MPa,
CA and

pressure(bar)

indiQ = 38.6392; % a correction for the motor difference


indiQ_engine = 0;
for i = n_IVC:n_EVO
indiQ = indiQ + (pcomb(i)+pcomb(i-1))/2*(V(i)-V(i-1))*0.1;
% indiQ_engine = indiQ_engine + (pengine(i,2)+pengine(i-1,2))/2*(V(i)-V(i-1))*0.1;
end

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%
% calculate friction torque
% also friction power
frictiontorque2;

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%
% efficiency
eta_indi = indiQ/tQ; % indicated efficiency
eta_brake = (indiQ - qf)/tQ; % brake efficiency

Appendix B: Repeatability of thermogravimetric analysis

236
This section concerns the repeatability of the isothermal oxidation tests. Since

each sample was conducted repeatability two times, here we only present one example of

the repeatability of isothermal test.

The presented example is an isothermal oxidation test of diesel soot from 30%

load and 75MPa fuel injection pressure. Both test followed exactly the same procedure:

• Gas environment is N2

• Initialize temperature to 30ºC

• Ramp at a rate of 10ºC/min to 500 ºC

• Isotherm 30 min

• Ramp at a rate of 5ºC/min to 550 ºC

• Switch gas to air

• Isotherm 60ºC

Figure B-1 actually only present the isothermal test in air (the last step), which

can be used for oxidative reactivity analysis. One can observe that the repeatability of

isothermal test is satisfying.

237
1.0
test 1
test 2
0.8

0.6
m/m0

0.4

0.2

0.0
0 10 20 30 40 50 60
time (m)
Figure B-1: Repeatability of thermogravimetric analysis for diesel soot, 30% load
and 75 MPa

238
Appendix C: Repeatability of Raman analysis

This section concerns the repeatability of Raman analysis. Raman spectrum was

obtained for random five locations of the sample. The 3L + 1G curve fitting techniques

were conducted individually on each spectrum. An Example of the result of AD1/AG and

D1 FWHM was reported here.

The example is B20 soot, 30% load, 50 MPa fuel injection pressure. It can be

observed that the repeatability of Raman analysis is satisfying.

Table C-1: Repeatability of Raman analysis for B20 soot, 30% load and 50 MPa fuel
injection pressure

Test number AD1/AG D1 FWHM (cm-1)


1 2.24 174.07
2 2.28 175.56
3 2.24 178.09
4 2.34 174.94
5 2.35 172.85
Average 2.29 175.10
Standard derivation 0.05 1.46

239
Appendix D: Repeatability of Morphology analysis
The morphology analysis requires more than 25 images of the transmission

electronic microscopy. The primary particle size and fractal dimension of each images

are measured and averaged value is reported. Again, one example (Table D-1) of

repeatability test is presented here.

Table D-1: Repeatability of number of primary particles and fractal dimension for
B20 soot, 60% load at 75 MPa fuel injection pressure

Image number Number of primary particles Fractal dimension


1 100.17 1.90
2 117.17 2.18
3 113.18 1.83
4 141.87 2.16
5 105.10 1.85
6 134.48 1.91
7 112.58 1.96
8 106.56 1.98
9 105.31 2.03
10 101.37 1.97
11 98.68 1.83
12 102.86 2.05
13 106.55 1.98
14 100.23 1.87
15 133.12 2.22
16 209.89 2.00
17 114.52 1.91
18 162.23 1.97
19 140.36 2.13
20 109.02 1.83
21 104.10 1.87
22 107.93 2.08
23 116.33 2.15
24 131.97 1.86
25 108.92 1.88
26 108.03 2.02
Average 118.94 1.98
Standard derivation 24.39 0.12

240
The repeatability of measurement of diameter of primary particles is demonstrated

with diesel soot, 60% load at 125 MPa fuel injection pressure, as show in Table D-2:

Table D-2: repeatability of measurement of diameter of primary particles of diesel


soot, 60% load at 125 MPa fuel injection pressure

Primary particle number Diameter of primary particle (nm)


1 15.10
2 26.21
3 25.71
4 21.90
5 28.98
6 19.86
7 26.15
8 34.84
9 22.05
10 24.20
Average 24.50
Standard derivation 5.07

241
Appendix E: Repeatability of particle size distribution (SMPS)
result
Three samplings were conducted with the scanning mobility particle sizer for

each condition. The reported result was an averaged of the three results. An example

(engine exhaust running with B20 at 30% load and 50 MPa fuel injection pressure) is

presented in the Figure E-1. The repeatability of the particle size distribution is satisfying.

7
1.0x10
Test1
Test2
Particle concentration (#/cm3)

8.0x10
6 Test3

6
6.0x10

6
4.0x10

6
2.0x10

0.0
10 100
Particle diamter (nm)

Figure E-1: Repeatability of particle size distribution of exhaust gases running with
B20 at 30% load and 50 MPa fuel injection pressure

242
VITA

Peng Ye

Education

Aug. 2007 – Dec. 2011

Ph.D in Energy and Mineral Engineering, Fuel Science Option

The Pennsylvania State University

Dissertation: Investigation of impact of fuel injection strategy and biodiesel

fueling on engine emissions and performance

Aug.2009 – Dec. 2011

M.S. in Energy and Mineral Engineering, Petroleum Engineering Option

The Pennsylvania State University

Thesis: A pseudofunctionless density diffusivity equation approach to natural gas

reservoir engineering analysis

Sep. 2003 – Jul. 2007

B.S in Geology, Peking University, China

B.S. in Economics, Peking University, China

Potrebbero piacerti anche