elsevier editorial system(tm) for energy conversion and

61
Elsevier Editorial System(tm) for Energy Conversion and Management Manuscript Draft Manuscript Number: Title: Particulate Emissions from Biodiesel Fuelled CI Engines Article Type: Review Article - by invitation only Keywords: Particulate; biodiesel; size-number distribution; toxic potential; soot morphology. Corresponding Author: Prof. Avinash Kumar Agarwal, PhD Corresponding Author's Institution: IIT Kanpur First Author: Avinash Kumar Agarwal, PhD Order of Authors: Avinash Kumar Agarwal, PhD; Tarun Gupta, PhD; Pravesh c Shukla, MTech; Atul Dhar, PhD Abstract: Compression ignition engines are the most popular prime-movers for transportation sector as well as for stationary applications. Petroleum reserves are rapidly and continuously depleting at an alarming pace and there is an urgent need to find alternative energy resources to control both, the global warming and the air pollution, which is primarily attributed to combustion of fossil fuels. In last couple of decades, biodiesel has emerged as the most important alternative fuel candidate to mineral diesel. Numerous experimental investigations have confirmed that biodiesel results in improved engine performance, lower emissions, particularly lower particulate mass emissions vis-à-vis mineral diesel and is therefore relatively more environment friendly fuel, being renewable in nature. Environmental and health effects of particulates are not simply dependent on the particulate mass emissions but these change with varying physical and chemical characteristics of particulates. Particulate characteristics are dependent on largely unpredictable interactions between engine technology, after-treatment technology, engine operating conditions as well as fuel and lubricating oil properties. This review paper presents an exhaustive summary of literature on the effect of biodiesel and its blends on exhaust particulate's physical characteristics (such as particulate mass, particle number-size distribution, particle surface area-size distribution, surface morphology) and chemical characteristics (such as elemental and organic carbon content, speciation of polyaromatic hydrocarbons, crustal and anthropogenic trace metals, sulfates, nitrates etc.) in order to comprehensively assess the effects of biodiesel usage on the environment as well as on the human health. Control of particulate emissions using various engine control parameters such as intake air boosting using turbocharging, high pressure fuel injections and multiple injections, exhaust gas recirculation (EGR), after-treatment devices etc. in combination with the use of biodiesel has also been critically reviewed and included in this review article. Suggested Reviewers: Suresh K Aggarwal PhD Professor, Department of Mechanical Engineering, University of Illinois@chicago, USA [email protected] Chang Sik Lee PhD Professor, Department of Mechanical Engineering, Hanyang University, South Korea [email protected]

Upload: buituyen

Post on 03-Jan-2017

224 views

Category:

Documents


3 download

TRANSCRIPT

Page 1: Elsevier Editorial System(tm) for Energy Conversion and

Elsevier Editorial System(tm) for Energy Conversion and Management Manuscript Draft Manuscript Number: Title: Particulate Emissions from Biodiesel Fuelled CI Engines Article Type: Review Article - by invitation only Keywords: Particulate; biodiesel; size-number distribution; toxic potential; soot morphology. Corresponding Author: Prof. Avinash Kumar Agarwal, PhD Corresponding Author's Institution: IIT Kanpur First Author: Avinash Kumar Agarwal, PhD Order of Authors: Avinash Kumar Agarwal, PhD; Tarun Gupta, PhD; Pravesh c Shukla, MTech; Atul Dhar, PhD Abstract: Compression ignition engines are the most popular prime-movers for transportation sector as well as for stationary applications. Petroleum reserves are rapidly and continuously depleting at an alarming pace and there is an urgent need to find alternative energy resources to control both, the global warming and the air pollution, which is primarily attributed to combustion of fossil fuels. In last couple of decades, biodiesel has emerged as the most important alternative fuel candidate to mineral diesel. Numerous experimental investigations have confirmed that biodiesel results in improved engine performance, lower emissions, particularly lower particulate mass emissions vis-à-vis mineral diesel and is therefore relatively more environment friendly fuel, being renewable in nature. Environmental and health effects of particulates are not simply dependent on the particulate mass emissions but these change with varying physical and chemical characteristics of particulates. Particulate characteristics are dependent on largely unpredictable interactions between engine technology, after-treatment technology, engine operating conditions as well as fuel and lubricating oil properties. This review paper presents an exhaustive summary of literature on the effect of biodiesel and its blends on exhaust particulate's physical characteristics (such as particulate mass, particle number-size distribution, particle surface area-size distribution, surface morphology) and chemical characteristics (such as elemental and organic carbon content, speciation of polyaromatic hydrocarbons, crustal and anthropogenic trace metals, sulfates, nitrates etc.) in order to comprehensively assess the effects of biodiesel usage on the environment as well as on the human health. Control of particulate emissions using various engine control parameters such as intake air boosting using turbocharging, high pressure fuel injections and multiple injections, exhaust gas recirculation (EGR), after-treatment devices etc. in combination with the use of biodiesel has also been critically reviewed and included in this review article. Suggested Reviewers: Suresh K Aggarwal PhD Professor, Department of Mechanical Engineering, University of Illinois@chicago, USA [email protected] Chang Sik Lee PhD Professor, Department of Mechanical Engineering, Hanyang University, South Korea [email protected]

Page 2: Elsevier Editorial System(tm) for Energy Conversion and

Anirudh Gautam PhD Executive Director, Engine Development Directorate, Research Designs and Standards Organisation, Lucknow [email protected] Opposed Reviewers:

Page 3: Elsevier Editorial System(tm) for Energy Conversion and

IIINNNDDDIIIAAANNN IIINNNSSSTTTIIITTTUUUTTTEEE OOOFFF TTTEEECCCHHHNNNOOOLLLOOOGGGYYY KKKAAANNNPPPUUURRR DEPARTMENT OF MECHANICAL ENGINEERING

KANPUR-208016, INDIA

Dr. Avinash Kumar Agarwal, Tel: + 91 512 2597982 (O), + 91 512 2598682 (R)

Poonam and Prabhu Goyal Endowed Chair Professor Fax: + 91 512 259 7408

Email:[email protected] http://home.iitk.ac.in/~akag

September 4th, 2014

Editor,

Energy Conversion and Management

Dear Sir,

I am submitting a manuscript entitled “Particulate Emissions from Biodiesel Fuelled CI Engines", by

Avinash Kumar Agarwal*, Tarun Gupta, Pravesh C Shukla, Atul Dhar for inclusion in “Energy

Conversion and Management”. This paper is being submitted against the invitation by Ms. Shannon

Qu on 14th March, 2014.

Submission of this article implies that the work described has not been published previously (except in

the form of an abstract or as part of a published lecture or academic thesis), that it is not under

consideration for publication elsewhere, that its publication is approved by all authors and that, if

accepted, it will not be published elsewhere in the same form, in English or in any other language,

without the written consent of the Publisher.

Looking forward to your kind consideration.

Best regards

Dr A K Agarwal

Cover letter

Page 4: Elsevier Editorial System(tm) for Energy Conversion and

Energy Conversion and Management

Submission Checklist

Please save a copy of this form to your computer, complete and

upload as ‘Checklist for New Submissions’.

Your manuscript will be considered incomplete unless all the below

requirements have been met. Please initial each step to confirm.

(1) I __Avinash K Agarwal --- confirm that the work described has not been

published previously (except in the form of an abstract or as part of a published

lecture or academic thesis), that it is not under consideration for publication

elsewhere, that its publication is approved by all authors and that, if accepted, it will

not be published elsewhere in the same form, in English or in any other language,

without the written consent of the Publisher.

Authors found to be deliberately contravening the submission guidelines on

originality and exclusivity shall not be considered for future publication in this

journal.

(2) The source document is editable, i.e. Word, _AKA_____ enter initials

(3) The source document is double-spaced

__AKA____ enter initials

(4) The source document has been prepared prepared in 12 or 10 point font size,

preferably 12 points

__AKA____ enter initials

(5) The source document is in one column per page

__AKA____ enter initials

(6) The figures and tables have been supplied: either integrated with the text file or

as separate files.

___AKA___ enter initials

*Checklist for New Submissions

Page 5: Elsevier Editorial System(tm) for Energy Conversion and

Particulate Emissions from 1

Biodiesel Fuelled CI Engines 2

3

Avinash Kumar Agarwal1*, Tarun Gupta2, Pravesh C Shukla2, Atul Dhar1 4

Engine Research Laboratory, 5

Departments of 1Mechanical and 2Civil Engineering 6

Indian Institute of Technology Kanpur, Kanpur-208016, India 7

*Corresponding author’s email: [email protected] 8

9

Abstract 10

Compression ignition engines are the most popular prime-movers for transportation 11

sector as well as for stationary applications. Petroleum reserves are rapidly and 12

continuously depleting at an alarming pace and there is an urgent need to find 13

alternative energy resources to control both, the global warming and the air pollution, 14

which is primarily attributed to combustion of fossil fuels. In last couple of decades, 15

biodiesel has emerged as the most important alternative fuel candidate to mineral 16

diesel. Numerous experimental investigations have confirmed that biodiesel results in 17

improved engine performance, lower emissions, particularly lower particulate mass 18

emissions vis-à-vis mineral diesel and is therefore relatively more environment friendly 19

fuel, being renewable in nature. Environmental and health effects of particulates are 20

not simply dependent on the particulate mass emissions but these change with varying 21

physical and chemical characteristics of particulates. Particulate characteristics are 22

dependent on largely unpredictable interactions between engine technology, after-23

treatment technology, engine operating conditions as well as fuel and lubricating oil 24

*ManuscriptClick here to view linked References

Page 6: Elsevier Editorial System(tm) for Energy Conversion and

2

properties. This review paper presents an exhaustive summary of literature on the effect 25

of biodiesel and its blends on exhaust particulate’s physical characteristics (such as 26

particulate mass, particle number-size distribution, particle surface area-size 27

distribution, surface morphology) and chemical characteristics (such as elemental and 28

organic carbon content, speciation of polyaromatic hydrocarbons, crustal and 29

anthropogenic trace metals, sulfates, nitrates etc.) in order to comprehensively assess 30

the effects of biodiesel usage on the environment as well as on the human health. 31

Control of particulate emissions using various engine control parameters such as intake 32

air boosting using turbocharging, high pressure fuel injections and multiple injections, 33

exhaust gas recirculation (EGR), after-treatment devices etc. in combination with the 34

use of biodiesel has also been critically reviewed and included in this review article. 35

Keywords: Particulate; biodiesel; size-number distribution; toxic potential; soot 36

morphology. 37

1. Introduction 38

Fossil fuels have dominated transportation sector since the invention of internal 39

combustion engines in early nineteenth century. Conventional petroleum resources are 40

finite and they contribute enormously to the ever-rising green house gas emissions to 41

the atmosphere thus renewable alternative fuels are being globally developed and 42

explored frenetically by researchers. Depletion of fossil fuels is eminent in near future. 43

In addition, environmental pollution concerns due to combustion of fossil fuels provides 44

a unique and significant motivation for developing renewable alternative fuels, which 45

have the potential to sustain ever-growing fuel demand for transportation sector. In 46

order to effectively control environmental pollution and mitigate its harmful effects, 47

exhaust characterization at the engine outlet as well its impact of its transformational 48

products in the atmosphere is extremely essential [1]. Health effects of exhaust 49

particulates depend on chemical composition and physical characteristics, which 50

Page 7: Elsevier Editorial System(tm) for Energy Conversion and

3

determines their true residence time and availability as sorption sites inside the human 51

respiratory system [2-6]. 52

Biodiesel has emerged as a strong diesel alternative, and comprises of fatty acid alkyl 53

esters derived from transesterification of triglycerides present in vegetable oils/ animal 54

fats. Biodiesel has been well accepted as renewable alternative to mineral diesel 55

globally. Large numbers of scientific studies have reported successful operation of 56

compression ignition (CI) engines with biodiesels derived from different feedstock. 57

Biodiesel can either be used as a full replacement of mineral diesel or it can also be 58

blended with mineral diesel in any proportion[7]. 59

Formation of particulates and gaseous emissions depend not only on physical and 60

chemical properties of the fuels but this process is also greatly influenced by complex in-61

cylinder processes such as air-fuel mixing, combustion chamber geometry and 62

temperature and pressure condition of cylinder charge during combustion [8, 9]. 63

Particulate formation in a diesel engine is very sensitive to relative air-fuel mixture 64

strength (λ) in rich, premixed reaction zones of the combustion chamber, where soot 65

precursors are initially generated [8]. Physical properties of biodiesel are also important, 66

which directly affect spray atomization, droplet size distribution and fuel-air mixing in 67

the combustion chamber. Injection delay and spray tip penetration is relatively longer 68

for biodiesel compared to mineral diesel, whereas spray cone angle, spray area and 69

spray volume are relatively smaller. Relatively higher viscosity and surface tension of 70

biodiesel is responsible for larger sauter mean diameter (SMD) of the spray droplets 71

[10]. In last two decades, advancements in diesel technology such as application of very 72

high fuel injection pressures, split injection, turbo-charging, after-treatment devices 73

have resulted in considerable reduction in engine-out emissions but these technologies 74

have also increased engine control complexities and sensitivity of engine towards 75

changes in fuel properties and lubricating oil properties. It is universally accepted that 76

Page 8: Elsevier Editorial System(tm) for Energy Conversion and

4

biodiesel blend usage in CI engines reduces PM mass emission but mixed trends are 77

reported for the physical and chemical characteristics of biodiesel particulates 78

depending upon the engine technology, biodiesel feedstock, biodiesel blend 79

concentration, type of after-treatment device used, engine management system 80

optimization etc. Several toxicological studies [2, 3] reported that biodiesel use in engine 81

results in lower toxicity of particulates, especially lower mutagenic potential as 82

compared to mineral diesel. A review paper [11] on effect of diesel exhaust on human 83

health concluded that evidence from scientific studies so far is insufficient to adequately 84

validate the diesel particulate-lung cancer hypothesis. This further emphasizes the need 85

to study biodiesel vs. diesel exhaust critically under various engine operating conditions 86

typically encountered in real world situation. 87

This review is an attempt to present an exhaustive literature review of the effects of 88

biodiesel and biodiesel blends on the exhaust particulate’s chemical characteristics (such 89

as elemental and organic carbon, speciation of polyaromatic hydrocarbons, crustal and 90

anthropogenic trace metals, sulfates and nitrates) and physical characteristics (such as 91

particulate mass emission, particle number-size distribution, particle surface area-size 92

distribution, and particulate morphology). This review critically assesses the effect of 93

biodiesel on the environment as well as human health under varying engine operating 94

conditions, varying fuel injection parameters and strategies, using various after-95

treatment control technologies to assess the possible threat emanating from a large 96

scale biodiesel usage. 97

2 Chemical Characterization of Particulates 98

2.1 Chemical Composition of Particulates 99

Diesel engine undergoes heterogeneous combustion phenomenon. Fuel is injected into 100

the combustion chamber towards the end of the compression stroke in conventional CI 101

engines. Modern diesel engines are equipped with common rail direct injection system, 102

Page 9: Elsevier Editorial System(tm) for Energy Conversion and

5

which have spit injection, i.e. fuel is injected in pilot injection, main injection and post 103

injection in the same engine cycle. When fuel is injected into the combustion chamber at 104

very high injection pressure, it breaks into large number of small droplets under the 105

influence of high combustion chamber pressure, prevailing at the end of compression 106

stroke. Compressed air in the combustion chamber offers resistance to the high pressure 107

fuel droplets, which results in further fragmentation of small droplets into further finer 108

droplets. When a relatively larger droplet breaks into several smaller droplets, total 109

surface area of the droplet increases significantly. This higher surface area provides 110

superior interaction between fuel droplets and surrounding hot high pressure 111

combustion chamber air, which eventually results in higher degree of completion of 112

combustion of the fuel injected. 113

Heywood [12] explained soot formation in the engine combustion chamber. By 114

considering a single fuel droplet inside the combustion chamber, soot formation process 115

can was explained. A single fuel droplet comes into contact with hot, high pressure air in 116

the combustion chamber (Figure 1). 117

118

(a) Diesel droplet (b) Biodiesel droplet 119

Figure 1: Evaporation of single diesel and biodiesel droplet inside combustion chamber 120

A small quantity of fuel evaporates from the fuel droplet surface and forms rich fuel-air 121

mixture closer to the surface of the droplet. Mixture composition becomes progressively 122

leaner as the distance from the surface of the droplet increases as seen in figure 1. 123

Wherever the air availability is good enough for combustion of fuel droplets, combustion 124

tends to be complete, resulting in low particulate formation. Near the droplet surface, 125

Page 10: Elsevier Editorial System(tm) for Energy Conversion and

6

where rich mixture formed (λ <1.0), combustion tends to be relatively incomplete, which 126

results in higher particulate formation. Similarly, there is no air/ oxygen available 127

inside the core of the droplet. Under the influence of hot burning droplets, fuel closer to 128

the droplet surface and the fuel present inside the droplet is either not able to burn or 129

burn partially, thus forming pyrolyzed fuel remnants, which act as precursors for soot 130

formation. Biodiesel has slightly higher temperature required of vaporization therefore 131

it can be assumed that the vaporization of biodiesel droplet is slightly lower as compared 132

to mineral diesel droplet. On the other hand, there is absolutely no oxygen present 133

inside diesel droplet (Figure 1a). Absence of oxygen molecules inside the fuel droplet 134

results in unburned hydrocarbons and pyrolyzed carbonaceous compounds. In case of 135

biodiesel, combustion is aided partially inside the droplet (Figure 1b) by the oxygen 136

atoms present in the biodiesel molecules (~11% w/w) [13]. 137

Diesel particulate mainly comprise of four components, namely (i) elemental carbon 138

(EC), (ii) organic carbon (OC), (iii) sulfate and (iv) ash, which mainly includes trace 139

metals [14]. Figure 2 shows typical composition of particulate emitted by CI engines. 140

141

142

Figure 2: Typical diesel particulate composition [15] 143

Elemental carbon is known as 'soot' and it mainly comprises of 'carbon'. Elemental 144

carbon is crystalline in structure and mostly forms central part of particulate [16]. 145

Organic carbon mainly consists of hydrocarbons, which either remained unburned 146

during combustion, primarily originating from fuel or lubricating oil or form due to 147

Carbon 41%

Unburned Oil

25%

Unburned fuel 7%

Ash and Others

14%

Sulphate and Water

13%

Page 11: Elsevier Editorial System(tm) for Energy Conversion and

7

condensation of organic vapors as a left-over of incomplete combustion. Organic fraction 148

of the particulate is of great concern due to its harmful effects on humans. Diesel and 149

biodiesel exhaust particulates both consists of significant amount of organic fraction. 150

However, scientific studies have shown that biodiesel exhaust particulates have 151

significantly lower organic fraction. 152

In a study, Rounce et al. [17] reported that concentration of acetaldehyde, formaldehyde, 153

benzene, and 1,3-butadiene was lower for rapeseed methyl ester (RME) in comparison to 154

ultra-low sulfur diesel (ULSD). RME produced relatively lower solid particulate and 155

higher liquid particulate as compared to ULSD. Particulate number concentration 156

reduced for the entire size range. However, RME produced a higher proportion of nano-157

particulates of smaller size range and such nano-particulates had higher soluble organic 158

fraction (SOF), which is a marker of toxicity. They reported that DPF captured ~99% 159

solid particulates in terms of mass and particle number for both ULSD and RME. 160

Similarly, DPF reduced~88% and ~80% liquid portion of particulates for ULSD and 161

RME. In another study, Zhu et al. [18] reported significantly reduced smoke opacity 162

with increase in biodiesel proportion in the test fuel, while the total particle number 163

concentration actually increased. It was observed that sulfate and SOF increased in 164

particulates with increasing biodiesel blend concentration, whereas solid particulates 165

actually reduced in number. Nucleation mode particle number concentration was 166

observed to be higher for biodiesel and total particle number concentration reduced with 167

ULSD. The contribution of lubricating oil was suggested to be as high as 80- 90% in 168

the SOF portion of the particulate [19]. 169

Chuepeng et al. [20] did experiments with B30 and reported that B30 produced lower 170

particulates on a mass basis at all engine operating conditions compared to ULSD. They 171

also observed lower EC content in particulate of B30 as compared to ULSD. They 172

suggested that presence of oxygen in biodiesel limits the in-cylinder particle formation 173

Page 12: Elsevier Editorial System(tm) for Energy Conversion and

8

by influencing both the carbon chain formation and its oxidation. At a high engine load, 174

higher amount of fuel-borne oxygen leads lower EC formation for biodiesel. Young et al. 175

[21] reported that non-volatile particulate concentration emitted by heavy-duty diesel 176

engine increased with increasing load from 25% to 75% and it decreased with increasing 177

biodiesel blend from 2- 20%. Williams et al. [19] measured particulate mass for different 178

engine operating conditions and determined volatile mass fraction of the particulates 179

using thermo-gravimetric analysis (TGA) from a CRDI V6 engine fuelled with RME 180

blended with ULSD (B30). They reported higher volatile organic fractions for idle and 181

low load conditions. EC fraction was lower for B30 compared to ULSD. Li et al. [22] 182

compared exhaust emissions and particulate size distribution for diesel, fresh cooking oil 183

(FCO) and waste cooking oil (WCO) at two engine operating conditions (23 kW and 47 184

kW). PM emission was almost equal to diesel at lower load however it reduced 185

significantly when engine was operated at higher loads. FCO showed higher PM 186

emissions at both conditions and both fuels showed lower nuclei mode particles after 187

DOC compared to mineral diesel. This suggested that DOC was not very effective in 188

reducing nano-particles emanating from diesel as compared to those from biodiesel. 189

Schönborn et al. [23] studied the effect of different molecular structures of fatty esters 190

on NOx and soot formation. They concluded that different biodiesels have different 191

physical and chemical properties depending on the fatty ester composition, which affect 192

the combustion in diesel engines. Lapuerta et al. [24] evaluated the effect of 193

unsaturation level of biodiesel on NOx and PM emissions. They reported significant 194

reduction in PM mass and smoke opacity for biodiesel (B100) however PM mass 195

decreased by ~20% and NOx increased by ~10%, as biodiesel became more unsaturated. 196

They also reported smaller mean diameter particulates from unsaturated biodiesel. 197

Zhang et al. [25] investigated the particle size distribution in a biodiesel blend fuelled 198

CRDI engine exhaust and found that B100 resulted in no emission of nucleation mode 199

Page 13: Elsevier Editorial System(tm) for Energy Conversion and

9

particles. Soewono et al. [26] investigated particulate morphology and microstructure 200

using diesel and B20 by employing transmission electron microscopy (TEM) and Raman 201

spectroscopy. B20 particulates showed higher structural disorder compared to diesel 202

particulates for the same engine operating conditions and structural order of soot 203

improved for higher engine loads. Tan et al. [27] tested five test fuels namely mineral 204

diesel, B10, B20, B50 and B100 for particle size-number distribution and reported that 205

the nucleation mode particles increased with increasing biodiesel blends concentration 206

vis-a-vis mineral diesel. Number of accumulation mode particles decreased with 207

increasing biodiesel content in the test fuel. They explained that higher number of 208

nucleation mode particles emitted with biodiesel may be a result of higher degree of 209

saturation of condensed matter in presence of lesser number of soot nuclei. Evaporation 210

and mixing characteristics of biodiesel is worse than mineral diesel, which leads to an 211

increase in SOF, which form nucleation mode particles. Biodiesel's fuel oxygen content 212

helps in producing higher number of ultra-fine and nano-particles. Tinsdale et al. [28] 213

investigated the impact of biodiesel on particle numbers, sizes and mass emissions from 214

a diesel engine. Accumulation mode particles and carbonaceous mass decreased and 215

organic mass in particulate increased by using B30. FAME (B30) led to increased 216

nucleation mode particles as compared to mineral diesel. Song et al. [29] studied 217

particulate emissions from oxidized and heated biodiesel and compared the results with 218

non-oxidized biodiesel and ULSD. They reported that particle mass and particle number 219

reduced with the use of biodiesel for heated and oxidized biodiesel. 220

2.2 Elemental and Organic Carbon (EC/OC) 221

Elemental carbon and organic carbon are the two main components exhaust 222

particulates. A large number of studies have investigated the EC and OC content of 223

diesel and biodiesel particulates. As a result of heterogeneous combustion in CI engines, 224

unburned and partially burnt hydrocarbons are emitted. Under high temperature and 225

Page 14: Elsevier Editorial System(tm) for Energy Conversion and

10

pressures encountered in the combustion chamber, a fraction of hydrocarbons, which are 226

present in locally rich regions, undergo pyrolysis. Most of the hydrogen atoms get 227

striped off the hydrocarbon chain and only carbon atoms remain. This result in the 228

formation of carbon core, which is also called as 'soot' (also known as EC). Carbon 229

undergoes cyclization, sheet like structure formation and eventually nano-tube like 230

structures called spherules are formed. Volatile organic materials usually condense over 231

the solid and dry soot and the particles grow. This condensed organic material is 232

extremely harmful for humans. This condensed organic matter contain hundreds of 233

organic compounds formed as a result of complex organic species formation pathways 234

during fuel pyrolysis in the engine combustion chamber. Some of the organic compounds 235

are known carcinogens such as polyaromatic hydrocarbons (PAHs), Benzene-Toluene-236

Xylene (BTX). A detailed discussion on PAHs emissions has been included in later part 237

of this review. Poitras et al. [30] observed a significant reduction in PM as well as 238

OC/EC ratio upon using biodiesel blends during an experimental study conducted to 239

study the impact of B0 (Diesel), B2, B5, B10, B20, B50 and B100 on particulate 240

emissions. Gangwar et al. [31] performed comparative study of diesel and biodiesel PM 241

mass and chemical composition. They reported that OC/ EC ratio decreased with 242

increasing engine load. BSOF was higher for B20 compared to mineral diesel for same 243

engine operating condition and it decreased with increasing engine load for both test 244

fuels. Although PM emission was lower for B20, PAHs emissions were same for both 245

fuels. Schauer et al. [32] analysed chemical composition of PM from four gasoline 246

vehicles under three driving cycles namely; cold-cold unified driving cycle (UDC), hot 247

UDC and steady-state driving cycle. They analyzed particulate composition for EC, OC, 248

sulfates, nitrates, organic compounds etc. using gas chromatography-mass spectroscopy 249

(GC-MS). The average mass emission rates varied from <0.1 to 1.3 mg/km for a hot UDC 250

and steady-state driving cycles, while it ranged between 1.0 to 7.1 mg/km for cold-cold 251

Page 15: Elsevier Editorial System(tm) for Energy Conversion and

11

UDC. EC was the main component in particulates for cold-cold UDC cycle and OC 252

consisted of different types of compounds for different cycles. Pabkin et al. [33] also 253

investigated physical, chemical and toxicological characteristics of particulates emitted 254

from a heavy-duty diesel engines equipped with advanced after-treatment devices. They 255

reported significant reduction in particle bound organics in the vehicles equipped with 256

advanced emission control devices. They observed insignificant reduction in hopanes 257

and steranes. Liu et al. [34] used two engine models (2004 model with EGR and 2007 258

model with EGR) crankcase condenser and a DPF and analyzed the particulate samples 259

for C1, C2 and C10-13 particle (EC, OC) phase and semi-volatile organic species. In 2004 260

model, formaldehyde, acetaldehyde and naphthalene were the major fractions out of 150 261

analyzed organic species. The concentration of the above compounds reduced 262

significantly in the 2007 model engine. In another study, Agarwal et al. [6] evaluated 263

comparative toxicity of nano-particles emitted from diesel and B20 fuelled engine. OC, 264

EC content of the particulates was determined for primary and secondary emissions 265

from diesel and B20 fuelled engine. 266

267

Figure 3: EC/OC ratio for diesel and biodiesel (primary and secondary emissions) [6] 268

A photochemical chamber was used for generating secondary emissions. It was found 269

that EC was higher for diesel for both primary and secondary emissions. EC/OC ratio 270

Page 16: Elsevier Editorial System(tm) for Energy Conversion and

12

was also higher for diesel at higher engine loads. Figure 3 shows the EC/OC ratio 271

obtained for diesel and B20 exhaust [6]. 272

2.3 Trace Metals in Particulates 273

Trace metal emissions from compression ignition engines is of a major environmental 274

and health concern. Lubricating oil, fuel and engine friction/ wear generated debris are 275

major source of trace metal emitted as a part of particulates from the engines. 276

Concentration of trace metals in the fuels such as mineral diesel varies depending upon 277

various factors namely type of crude oil used for production of diesel, synthesis processes 278

and catalysts used in refining process etc. Trace metals are categorized as 279

'anthropogenic metal emissions' and 'crustal metal emissions'. Fe, Ca, Mg, Na are the 280

major crustal metal emissions. Several studies have been performed for trace metal 281

emission evaluation from diesel engines. Transition metal containing particulates can 282

even penetrate deep into the human body. These trace metals raise the level of reactive 283

oxygen species (ROS) activity in cell structures, which in-turn elevates the oxidative 284

stress [35-39]. Pillay et al. [40] compared the trace metals in Neem biodiesel vis-à-vis 285

two commercial grade biodiesels and concluded that Neem biodiesel has relatively lower 286

trace metal content compared to the other two. Some metals like Mn, Cu, Pb etc. were 287

observed to be in higher concentration in Neem biodiesel though. They suggested that 288

further refinement of biodiesel for de-metallization need to be undertaken for 289

sustainable biodiesel usage. Betha et al. [41] characterized trace metal emissions from 290

waste cooking oil-derived biodiesel (B100), ultra-low sulfur diesel and B50 blend. Mg, K, 291

and Al were present in significantly higher concentration in diesel as well as biodiesel. 292

In biodiesel, Zn, Cr, Cu, Fe, Ni, Mg, Ba and K were observed to be in higher 293

concentration compared to biodiesel. However, Co, Pb, Mn, Cd, Sr, and As were observed 294

to be higher in mineral diesel. They carried out risk assessment study and found that 295

Page 17: Elsevier Editorial System(tm) for Energy Conversion and

13

B100 exhaust has possibly higher health risk compared to ULSD. Figure 4 shows the 296

emission of anthropogenic trace metal emissions from ULSD, B50 and B100 [41]. 297

298

Figure 4: Concentration of carcinogenic trace metals in diesel and biodiesel exhaust [41] 299

Gangwar et al. [31] evaluated trace metal content in diesel and Karanja biodiesel (B20) 300

particulates. They observed that Si, Cu, and Mg were present in higher concentration in 301

diesel particulate as compared to biodiesel particulate. However, Na, Fe, Ca, Pb, Mn and 302

Cr were found to be higher in biodiesel particulate as compared to diesel particulate. 303

Table 1 shows the metal concentration in diesel and Karanja biodiesel particulates. 304

Table 1: Trace metals concentration in diesel and biodiesel exhaust particulates [31] 305

Metals in

particulate

(mg/g)

No Load Full Load

1800 rpm 2400 rpm 1800 rpm 2400 rpm

DE B20 DE B20 DE B20 DE B20

Na 29.4-29.8 34.6-35.0 51.55-51.59 20.25-20.29 5.50-5.54 11.0-11.4 1.91-1.95 1.37-1.41

Ca 6.22-6.26 11.60-11.64 24.44-24.48 31.70-21.74 1.80-1.84 2.27-2.31 0.39-0.43 0.22-0.26

Fe 12.40-12.44 8.81-8.85 9.62-9.66 8.27-8.31 0.82-0.86 3.26-3.30 0.74-0.78 0.32-0.36

Pb 6.51-6.55 2.31-2.35 2.57-2.61 5.11-5.15 0.61-0.65 1.00-1.04 0.33-0.37 0.61-0.65

Si 6.75-6.79 2.04-2.08 2.40-2.44 0.63-0.67 1.01-1.05 0.84-0.88 1.14-1.18 0.04-0.08

Cu 2.60-2.64 1.81-1.85 2.13-2.17 2.07-2.11 1.89-1.93 1.93-1.97 0.31-0.35 0.11-0.13

Mg 1.89-1.93 0.81-0.85 1.74-1.78 2.23-2.27 1.52-1.56 0.57-0.61 0.27-0.31 0.04-0.08

B 4.30-4.34 3.31-3.35 0.32-0.36 0.07-0.11 0.37-0.41 1.50-1.54 0.14-0.18 0.01-0.03

Mn 0.23-0.27 0.56-0.6 0.057-0.061 0.27-0.31 0.01-0.03 0.17-0.21 0.01-0.05 0.01-0.03

Page 18: Elsevier Editorial System(tm) for Energy Conversion and

14

Cr 0.40-0.44 0.05-0.09 0.15-0.19 0.13-0.17 0 0.01-0.03 0.01-0.03 0.01-0.03

306

Agarwal et al. [7] have also determined trace metal concentrations in diesel and Karanja 307

oil and reported relatively higher concentration of Ca, Co, Cr, Cu, Fe, Mg, Ni, Pb and Zn 308

in mineral diesel compared to Karanja oil. 309

310

2.4 Significance of Desulphurized Fuel 311

Diesel is a petroleum product derived from crude oil. Crude oil contains significant 312

amount of sulfur ranging from 0.5 to 5.0% (w/w) [42]. For automotive applications, 313

sulfur content of diesel should be very low therefore diesel is refined in such a manner 314

that sulfur content is reduced significantly. Several studies show that sulfur content of 315

diesel increases formation of particulate matter. Sulfur leads to formation of sulfur 316

dioxide, which results in formation of sulfates. Sulfate acts as nuclei for the 317

condensation of volatile organic compounds present in the diesel exhaust. The process, 318

in which condensed matter get adsorbed on to the existing nuclei is called heterogeneous 319

nucleation. On the other hand, main advantage of biodiesel is that it has no sulfur [13] 320

hence sulfate origin particulates are formed. Modern diesel vehicles equipped with after-321

treatment devices like diesel oxidation catalyst (DOC), diesel particulate filter (DPF) 322

etc. are vulnerable to sulfur content of the fuel. Catalysts like platinum (Pt) promote 323

formation of sulfate particulates in diesel exhaust, which is undesirable. 324

2.5 Unregulated Emissions 325

The current emission regulations are based on controlling emission of regulated gases 326

(CO, THC and NOx) and particulate matter (PM). Vehicles/ engines also emit large 327

number of other emissions, most of which are in very small quantities except CO2 and 328

moisture, also are currently categorized as unregulated emissions. In some regulations, 329

CO2 is now included as regulated emission. Unregulated emissions are important from 330

health stand point. Ravindra et al. [43] indicated in their research that there should be 331

Page 19: Elsevier Editorial System(tm) for Energy Conversion and

15

emission regulations for carcinogenic compounds like PAHs. PAHs, carbonyl compounds 332

and BTEX are harmful species emitted in diesel exhaust in traces. It is important to 333

perform chemical speciation of these organic species being emitted by the engine. 334

2.5.1 Carbonyl Compounds 335

Diesel engines emit large number of different harmful compounds and many compounds 336

are still unknown. The term carbonyl refers to the carbonyl functional group, which is a 337

divalent group consisting of a carbon atom double-bonded to oxygen. Carbonyls are such 338

compounds, which have significant presence in engine exhaust. Most studies have 339

measured carbonyl emissions by derivatives of 2,4- di-nitro-phenyl-hydrazine (DNPH) 340

[44-48]. Carbonyl emissions lead to formation of secondary organic aerosols (SOA) by 341

forming oligomers [49]. Contribution of carbonyls in diesel particles also enhances its 342

responses physiologically [50]. Figure 5 shows the basic structures of carbonyl group and 343

carbonyl compounds such as Aldehydes and Ketones. 344

345

Figure 5: Carbonyl Group, Aldehyde and Ketone 346

Pang et al. [51] investigated characteristics of carbonyl emissions from a diesel engine 347

fuelled with biodiesel-ethanol-diesel blend. They reported that acetaldehyde was the 348

carbonyl compound in highest concentration emitted, followed by formaldehyde, acetone, 349

propaldehyde and benzaldehyde respectively. They reported 1-12% higher total carbonyl 350

emissions with biodiesel-ethanol-diesel blend depending on engine operating condition. 351

They also observed that carbonyl emissions increased with increasing engine speed 352

while minimum carbonyl emissions were found at 50% engine load, when the engine 353

was operated at a constant speed [52]. Ho et al. [52] measured and quantified 15 354

Page 20: Elsevier Editorial System(tm) for Energy Conversion and

16

different carbonyl emissions and formaldehyde was found to be the most dominant 355

compound, followed by acetaldehyde and acetone. They reported that formaldehyde was 356

54.8- 60.8% of the total carbonyl compounds present in the exhaust. They took samples 357

at various locations in city of Hong Kong and reported that formaldehyde concentration 358

was quite high compared to theoretical value expected in summer, which suggests 359

significant effect of photochemical reactions in formaldehyde production in ambient. 360

2.5.2 Benzene, Toluene, Ethyl-Benzene and Xylene (BTEX) 361

Petroleum derivatives such as gasoline contain these compounds (BTEX), which have 362

harmful effects on humans. Cheung et al. [53] investigated BTEX emission from a diesel 363

engine fuelled with mineral diesel, biodiesel and biodiesel blends with methanol (5%, 364

10%, and 15%) at constant engine speed 1800 rpm for five different loads. They reported 365

that biodiesel had lower BTEX emissions compared to mineral diesel and higher blends 366

of methanol in biodiesel further reduced BTEX emissions. Higher oxygen content in the 367

fuel leads to oxidation of BTEX. They observed that higher engine load results in lower 368

BTEX emission in the engine exhaust. Di et al. and Takada et al. [54, 55] also reported 369

lower BTEX emissions at higher engine loads. Ballesteros et al. [56] used biodiesel and 370

reported relatively lower aromatic emissions. Correa and Arbilla [57] found a strong 371

correlation between carbonyl emissions and biodiesel content (r2 > 0.96). They reported 372

that esters in biodiesel may be a main source of these carbonyl emissions. On the other 373

hand, Liu et al. and Cheung et al. [34, 53] indicated that carbonyl emissions increase 374

with increasing biodiesel content at lower engine load, however it decreases at higher 375

engine loads. Xue et al. [58] summarized that biodiesel reduces the emission of aromatic 376

and poly-aromatic compounds. They also suggested that carbonyl emissions increase in 377

general with biodiesel content because biodiesel provides extra oxygen in the fuel 378

molecules. 379

2.5.3 Polycyclic Aromatic Hydrocarbons (PAHs) 380

Page 21: Elsevier Editorial System(tm) for Energy Conversion and

17

PAHs are well known carcinogens and are produced as a result of incomplete 381

combustion of fuel in diesel engines. Ravindra et al. [43] prepared a database to identify 382

and characterize the PAHs emissions in their study. They also discussed factors 383

affecting PAHs emissions. Most of the probable human carcinogenic PAHs are found to 384

be adsorbed on the particulate matter surface. There are no strict regulations for PAH 385

emissions but these pollutants should get high priority due to their huge negative 386

impact on human health. Figure 6 shows the priority listed PAHs. Some of compounds 387

shown in Figure 6 are considered as 'probable human carcinogen' (B2), while some are 388

not listed as 'human carcinogens' (D) [43]. The toxicity of these PAH compound is highly 389

dependent on their molecular structure. Two isomers of PAHs with different structure 390

show quite different toxicity. Therefore EPA has divided these PAH compounds into 391

different categories. 392

393

Figure 6: Priority list PAHs [43] 394

* (Not included in priority list); D (not listed as to human carcinogens); B2 (probable human carcinogen) 395

Page 22: Elsevier Editorial System(tm) for Energy Conversion and

18

Lea-Langton et al. [59] collected particulate samples for diesel, biodiesel and cooking oil 396

for comparing and analyzing particulate bound PAH emissions from a heavy duty DI 397

diesel engine. Most of the particulate bound PAH were found to be lower in both biofuels 398

compared to mineral diesel, especially at low load conditions and most of the larger 399

PAHs such as benzo(a)anthracene, chrysene, benzo(b)fluoranthene and 400

benzo(k)fluoranthene were oxidized by DOC. They also reported that flouranthene was 401

absent in mineral diesel but was present in particulates, which was an evidence of 402

pyrolytic formation of flouranthene in engine combustion chamber. Zielinska et al. [1] 403

reviewed physical and chemical transformations of primary diesel emissions. They 404

concluded that transformation of primary diesel emissions in atmosphere is very 405

important in the context of human health. Primary diesel exhaust reacts mainly with 406

OH radicals, ozone, NOx radicals and sunlight. Monocyclic aromatics of primary diesel 407

exhaust reacts with OH radicals and form various aromatic compounds such as phenols, 408

glyoxal, quinones, nitro-PAHs, and aromatic aldehydes etc. Agarwal et al. [6] compared 409

the toxic potential of diesel and biodiesel (B20) fuels for primary and secondary 410

emissions. By measuring particulate size-number distribution, size-surface area 411

distribution, elemental and organic carbon content, particle bound PAHs and toxic 412

equivalent factor, toxicity and potential health hazards of these emissions were 413

assessed. They reported that toxicity of biodiesel exhaust was comparatively lower than 414

mineral diesel exhaust. Lu et al. [60] reported that waste cooking biodiesel resulted in 415

reduction of PAH emissions in comparison to LSD and ULSD. They reported that ULSD 416

resulted in nearly 8.6% lower PAH emissions compared to LSD. Biodiesel significantly 417

reduces the PAH emissions in the particulate as compared to ULSD and LSD and it was 418

reported to be lower by 32.5% and 38.1%, respectively compared to LSD and ULSD. 419

In recent decades, it is reported that PAHs presents in the diesel particulates are one of 420

the main factor, which adversely affect human health. PAHs include various kinds of 421

Page 23: Elsevier Editorial System(tm) for Energy Conversion and

19

poly-aromatic compounds, which manifest different toxic properties. EPA has listed 16 422

PAHs as carcinogenic, probable carcinogenic and possible carcinogenic and the 423

molecular structure of these is shown in Figure 6 [43]. Researchers [61, 62] have 424

performed speciation of PAHs adsorbed on to the diesel particulates. Each PAH has a 425

different toxic potential for carcinogenic effects therefore speciation of PAHs is 426

important [63]. There are some studies, which report the toxic potentials of individual 427

PAH species. 428

429

Figure 7: Total toxic equivalent potential of PAHs emitted by diesel and biodiesel (B20) 430

fuelled engine [6] 431

Agarwal et al. [6] calculated individual PAH content from the total PAH load by using 432

procedure given by Pan et al. [62] (Figure 7). They also evaluated the toxic equivalent 433

factors (TEFs) of 8 PAHs and 2 nitro-PAHs. They performed this experimental study on 434

diesel and biodiesel (B20) for primary and secondary emissions from a CRDI engine. For 435

secondary emissions, they used a UV light illuminated photo-chemical chamber with a 2 436

hours residence time. They observed that the trend of total toxic equivalent potential 437

was similar to the particle bound PAH emissions. Slightly higher toxic potential for 438

diesel was observed compared to biodiesel. Similarly, primary particulates showed lower 439

toxicity compared to secondary emissions in terms of PAH toxicity. 440

Page 24: Elsevier Editorial System(tm) for Energy Conversion and

20

441

Figure 8: Toxicity equivalent factors (TEFs) for diesel and biodiesel blends [65] 442

Karavalakis et al. [64] studied the impact of five biodiesels on PAHs, nitro-PAHs and 443

oxy-PAHs emissions. 10% (v/v) Biodiesels were blended with mineral diesel. A Euro-3 444

CRDI engine was tested on New European Driving Cycle (NEDC) and results were 445

compared with Artersias Driving Cycle (ADC). They observed that biodiesel addition to 446

mineral diesel leads to increase in emission of lower molecular weight PAHs. This 447

indicated relatively lower toxicological potential of biodiesel blends. However, higher 448

molecular weight PAHs showed both increasing and decreasing trends. It was observed 449

that nitro-PAHs were higher for biodiesel blends and oxy-PAHs increased with 450

increasing biodiesel blend concentration. Increasing engine speed and load reduced 451

emission of most of PAHs. Karavalakis et al. [61] tested Euro-2 diesel engine with diesel 452

and different biodiesel blends (B5, B10 and B20) for two vehicle driving cycle (ADC and 453

NEDC). They reported 11 PAHs and 5 nitro-PAHs emissions in the exhaust from diesel 454

and biodiesel blends. Lower molecular weight PAHs like phenanthrene, anthracene, 455

pyrene were the dominating PAHs, when biodiesel was blended with diesel. In general, 456

biodiesel blending resulted in lower emission of PAHs and nitro-PAHs. Bakeas et al. [65] 457

investigated PAH emissions from a Euro-4 CRDI engine fitted with DOC under NEDC 458

and ADC driving cycles (Figure 8). They used soyabean biodiesel, a palm-based biodiesel 459

and an oxidised biodiesel obtained from used frying oils, which were blended with ULSD 460

Page 25: Elsevier Editorial System(tm) for Energy Conversion and

21

(B30, B50 and B80). They observed that all types of biodiesel reduced overall PAHs 461

toxicity except waste frying oil, which offsets the advantage of using waste frying oil as 462

feedstock for biodiesel production (Figure 8). Ravindra et al. [43] reported that larger 463

molecular weight PAHs formed by pyro-synthesis of lower molecular weight PAHs in 464

addition to contributions from lubricating oil. Rhead and Hardy [66] explained that 465

PAHs are complex organic molecules, which include hydrogen and carbon atoms and at 466

least two benzene rings. PAH formation takes place because of the incomplete fuel 467

combustion, and unburned lubricating oil [43, 67]. Riddle et al. [68] explained that PAHs 468

are mutagenic and carcinogenic. PAHs are widely spread compounds in the atmosphere. 469

According to USEPA, 16 PAHs are classified as priority pollutants. Miet et al. [69] 470

explained that nitro-PAHs are formed in the engine as precursor PAHs due to 471

incomplete combustion. Nitro-PAHs can also be formed as a result of radical reactions of 472

OH and NO3 with PAHs. Heeb et al. [70] explained that Nitro-PAHs contribute to 473

mutagenicity and genotoxicity of diesel particulates. Nisbet et al. [63] reported the toxic 474

equivalent factor of different PAHs species. Pan et al. [62] compared the PAHs and 475

nitro-PAHs emissions from diesel, biodiesel (B100,soy methyl ester), and B20. They 476

performed experiment on a Cummins B5.9 engine. The values for PAHs and nitro-PAHs 477

emissions for diesel and biodiesel blends were obtained by using gas chromatography 478

(GC) coupled with mass spectrometry (MS). Bagley et al. [71] reported that particle-479

bound PAHs and 1-nitropyrene reduced by use of biodiesel. 480

2.5.5 Effect of Biodiesel on Unregulated Emissions 481

Karavalakis et al. [61] carried out investigations on regulated and unregulated 482

emissions from a Euro-2 IDI diesel passenger vehicle (Toyota Corolla 2.0 TD CR: 23:1, 483

61 kW @ 4000 rpm 174 Nm @ 2000 rpm, 1998 model) using LSD and soy methyl ester 484

blends and compared the results of experiments performed under ADC and NEDC test 485

cycles. Unregulated emissions of PAHs, Nitro-PAHs and carbonyl compounds were 486

Page 26: Elsevier Editorial System(tm) for Energy Conversion and

22

measured. For PAH analysis, glass fiber filters were used for particulate phase 487

sampling. Gas chromatograph mass spectrophotometer (GC-MS) was used for PAHs and 488

nitro-PAHs determination. They identified and quantified 13 carbonyl compounds in the 489

exhaust and reported relatively higher concentration of carbonyl compound in ADC 490

compared to NEDC. Formaldehyde was the major compound in both cases, followed by 491

acetaldehyde, butylaldehyde, benzaldehyde, valeraldehyde and p-tolualdehyde. 492

Formaldehyde mainly originated from incomplete combustion of saturated aliphatic 493

hydrocarbons. Lower saturated aromatic hydrocarbons in the biodiesel blends were 494

responsible for lower Formaldehyde emissions for higher blends. Thus, carbonyl 495

emissions were affected by biodiesel blending ratio. 496

Tan et al. [72] performed experiment on a light duty diesel engine using five different 497

fuels having different sulfur content and investigated effect of sulfur on regulated and 498

unregulated emissions. The investigations were conducted for three unregulated 499

emissions namely formaldehyde, acetaldehyde and SO2. They found that formaldehyde 500

emission was not detected by their instruments. Acetaldehyde emission decreased with 501

increasing load and decreased with increasing fuel sulfur content. SO2 emission 502

increased continuously with increasing engine load and decreased with lowering sulfur 503

content of the fuel. Concentration of formaldehyde was so low that it could be measured 504

only at low engine loads. This suggested that low combustion chamber temperature, 505

prevailing at low engine load conditions has higher formaldehyde emissions. 506

Formaldehyde is an intermediate combustion product. Formaldehyde emission 507

decreases with increasing engine load and combustion chamber temperature [72]. 508

Cheung et al. [73] conducted an experiment on a four cylinder DI engine with ULSD and 509

four different ethanol blends (blend-1, blend-2, blend-3 and blend-4 containing 6.1%, 510

12.2%, 18.2% and 24.2% ethanol v/v i.e. oxygen content 2%, 4%, 6% and 8% w/w). 511

Ethanol is an oxygenated compound, and can be blended with diesel. The objective of 512

Page 27: Elsevier Editorial System(tm) for Energy Conversion and

23

this study was to investigate regulated and unregulated emissions from ULSD-ethanol 513

blends. They found that unburnt ethanol and acetaldehyde emissions in exhaust 514

increased but formaldehyde, ethene, ethyne, 1,3-butadiene and BTX decreased with 515

increasing load. They observed that formaldehyde emission decreased with increasing 516

alcohol content in ULSD, possibly because of increased H/C ratio [74]. However, 517

emissions depend on several factors such as fuel blends composition, fuel's oxygen 518

content, engine technology, test cycles, etc. They also measured BTX emissions. BTX 519

emissions reduced with increasing engine load. This was because at higher combustion 520

chamber temperature, benzene and its derivatives oxidize. At low engine load, higher 521

benzene emissions were observed. Toluene and xylene also showed the same trend as 522

benzene. Combustion chamber temperature and oxygen content of fuel are therefore 523

very important factors for BTX emissions. 524

525

2.6 Effect of After-treatment Devices (DOC and DPF) 526

Biodiesel is an alternate fuel for mineral diesel however it has some properties, which 527

are different from diesel. Some studies [6, 75] showed that particulate characteristics of 528

biodiesel are different than that of mineral diesel. Several researchers performed 529

studies on the effect of biodiesel exhaust on the after-treatment devices. Biodiesel 530

contains some trace metals, which result in catalytic activity in the exhaust. 531

Composition of biodiesel particulates is quite different compared to ones from diesel. 532

Agarwal et al. [6] suggested that biodiesel contain complex compounds, which are 533

relatively more difficult to oxidize during combustion process. Shi et al. [76] investigated 534

effect of diesel oxidation catalyst (DOC) on exhaust from engine fuelled with mineral 535

diesel and B20 and reported that DOC reduces CO and HC emissions by 90-95% and 36-536

70% respectively (Table 2). Total carbon emission decreases by ~22-32% with use of 537

DOC. OC reduction was ~35-97% and EC reduction was ~3-65%. OC/ EC ratio of PM2.5 538

Page 28: Elsevier Editorial System(tm) for Energy Conversion and

24

from mineral diesel slightly increased at lower loads after DOC, however this parameter 539

showed an opposite trend for B20. 540

Table 2: Effect of DOC on fuel based emission factors (EF) of TC, EC and OC in PM2.5, 541

diesel fuel [76] 542

Speed

(rpm)/load

%

EC/(mg/g) OC/(mg/g) TC/(mg/g)

Diesel

Diesel +

DOC

Diesel

Diesel +

DOC

Diesel

Diesel +

DOC

2125/25 0.016 0.007 0.15 0.108 0.165 0.115

2125/50 0.05 0.037 0.148 0.118 0.198 0.155

2125/75 0.084 0.082 0.144 0.073 0.229 0.155

2690/25 0.032 0.029 0.306 0.228 0.337 0.258

2690/50 0.135 0.12 0.23 0.162 0.365 0.282

2690/75 0.206 0.134 0.221 0.182 0.427 0.317

543

Zhu et al. [77] evaluated particulate and unregulated emissions with and without DOC 544

in a Euro-5 diesel engine fuelled with biodiesel and biodiesel-ethanol blends. DOC was 545

found to be quite effective in reducing particle mass emissions, particle numbers as well 546

as unregulated emissions. However, DOC was not equally effective in emission 547

reduction for hydrocarbon compounds. They concluded that combination of biodiesel-548

ethanol with DOC is effective in reducing particulate emission and unregulated 549

emissions. Bagley et al. [71] tested an IDI diesel engine for emission reduction with 550

oxidation catalytic converter fuelled with diesel and soy biodiesel. They reported that 551

vapor phase PAH emissions reduced up to 90% by the use of oxidation catalytic 552

converter for both fuels. Particle and vapor phase mutagenic compounds reduced up to 553

50% by use of oxidation catalytic converter. Williams et al. [78] investigated adverse 554

effects of trace metals present in biodiesel on the after-treatment devices, which 555

included several diesel particulate filter (DPF) substrates, DOC and selective catalytic 556

reduction (SCR) catalysts. They observed no thermal-mechanical degradation of 557

Page 29: Elsevier Editorial System(tm) for Energy Conversion and

25

cordierite, aluminum titanate or silicon carbide DPFs with 150,000 miles equivalent 558

exposure to biodiesel ash and thermal aging. Performance of DOC was adversely 559

affected at 150,000 miles equivalent aging, and resulted in increased level of HC and CO 560

emissions after DOC. Vertin et al. [79] conducted tests to observe the impact of soy 561

biodiesel blend (B20) and two kinds of ULSD on a cordierite DPF. They observed that 562

B20 particulates were more reactive to DPF compared to diesel particulates. DPF 563

showed 80% higher efficiency with B20 and pressure drop was also not very high. Austin 564

et al. [80] performed active regeneration experiment on a B20 fuelled diesel engine 565

equipped with DOC and DPF. B20 particulates showed five times greater reaction ratio 566

in active regeneration compared to ULSD. The researchers concluded that due to the 567

higher reaction rate of B20 particulates, lower amount of fuel is required for 568

regeneration. Asti et al. [81] studied the effect of biodiesel on particulates during active 569

and passive DPF regeneration and found that particulate emissions decreased with 570

increasing biodiesel content in the fuel. They observed temperature gradient in DPF 571

during active regeneration with biodiesel however, no appreciable temperature gradient 572

was observed during passive regeneration with biodiesel. Ash content of DPF was also 573

higher with biodiesel. Parihar et al. [82] focused on physical characterization of diesel 574

and biodiesel particulates from a CRDI engine using MOUDI (10 stages) and observed 575

that submicron particle mass concentration was higher for higher loads. PM2.5 576

contributes approximately 75-90% of the total particulate mass. Di Iorio et al. [83] 577

analyzed the impact of biodiesel on particulate emissions and DPF regeneration and 578

observed that biodiesel leads to lower particulate emissions, which require less frequent 579

regeneration. On the other hand, regeneration of DPF with biodiesel require higher 580

quantity of fuel to be injected due to biodiesel's lower calorific value. They suggested 581

that using a 'flexible management system' is required for optimum regeneration, which 582

can take care of differences in fuel properties. Pidgeon et al. [84] investigated the effect 583

Page 30: Elsevier Editorial System(tm) for Energy Conversion and

26

of biodiesel blend on catalyzed particulate filter (CPF) performance and observed that 584

the soot reactivity of the CPF increases with increasing blend ratio of biodiesel. They 585

also reported that the PM oxidation increases with increasing biodiesel blend ratio at 586

constant CPF temperature. 587

3 Physical Characterization of Particulates 588

Biodiesel is a good alternative fuel for mineral diesel. As a partial or complete 589

replacement of mineral diesel, biodiesel needs to be critically evaluated for its physical, 590

chemical and thermal properties as well as its combustion products/ emissions. Physical 591

characterization of emissions includes measurement of the emitted particulate’s mass 592

and size-number distribution. Section 3.1 and 3.2 provide insights into particulate mass 593

and number emissions from diesel engine fueled with biodiesel and its blends with 594

mineral diesel. 595

3.1 Particulate Mass Emissions 596

Diesel engines are one of the biggest source of carbonaceous particulate emissions in the 597

environment. These primary particulate emissions cause several adverse effects in the 598

environment as well as on human health. These particles are formed due to incomplete 599

combustion of fuel in fuel-rich regions of the diesel engine. Formation of particulate 600

occurs mainly due to the insufficient oxygen availability in fuel-rich regions during 601

heterogeneous combustion. It is therefore essential to optimize the fuel injection 602

parameters in order to reduce particulate formation. Biodiesel is an oxygenated fuel and 603

some of its properties are different than mineral diesel, causing lower particulate 604

formation in an engine. Particulate mass emission is very important from regulated 605

emission control point of view and has been explained in two different aspects. First, the 606

effect of fuel injection strategy, and second, the effect of biodiesel blend composition of 607

particulate emission is summarized here. Most studies on biodiesel particulate show 608

lower particulate mass emissions compared to mineral diesel. Biodiesel fuel's molecular 609

Page 31: Elsevier Editorial System(tm) for Energy Conversion and

27

oxygen helps in improving combustion in the engine combustion chamber thus lower 610

particulate emissions are reported [24, 85] however some studies have also reported 611

identical or increased PM mass emissions with biodiesel usage [86, 87]. An EPA report 612

encompassing 39 scientific studies on heavy duty engines concluded that use of B100 613

and B20 leads to ~50% and 10% PM mass reduction respectively vis-a-vis mineral diesel 614

[85]. A scientific study suggested that reduction in smoke from Neem oil methyl ester 615

(NOME)-diesel and castor oil methyl ester (COME)-diesel blends was mainly due to 616

additional oxygen present in biodiesel, which reduced PM formation [88]. Biodiesel 617

blended with LSD emitted lower PM mass emission as compared to LSD, however no 618

concrete trend was reported for biodiesel blended with ULSD as compared to ULSD [89]. 619

Another scientific study reported relatively lower PM mass emissions from 100% 620

biodiesel and biodiesel blends in comparison to mineral diesel in a DI engine equipped 621

with electronically controlled fuel pump [90]. They also suggested that fuel oxygen in 622

biodiesel/ blends was the main reason for reduced PM mass emission [90]. Another 623

study reported that Soy biodiesel usage resulted in 77% lower PM mass emission in 624

comparison to mineral diesel [91]. Correlations derived from several test fuels suggested 625

that PM mass emission decreased with increasing fuel oxygen content as well as 626

adiabatic flame temperature [91]. Another study on RME reported that particulate 627

emissions significantly depend upon engine operating conditions. At low engine loads, 628

relatively higher PM mass emissions were observed with biodiesel compared to mineral 629

diesel. Interestingly, at higher engine loads, a reduction in PM mass emission was 630

observed with biodiesel [92]. Table 3 summarizes several scientific studies showing the 631

effect of different biodiesels on PM mass emission vis-a-vis baseline mineral diesel. 632

Table 3: Summary of effect of biodiesel on PM mass emissions 633

Test fuel Change in

PM (% w/w)

Reference

Page 32: Elsevier Editorial System(tm) for Energy Conversion and

28

B35 (Soybean oil biodiesel) -25 Wang et al., 2000 [93]

B100 (Soybean oil biodiesel) -55* Canaki, 2007[94]

B100 (Soybean oil biodiesel) -60* Canaki and Van Gerpen 2003 [95]

B100 (Soybean oil biodiesel) -77 Sharp et al., 2005 [91]

B100 (Soybean oil biodiesel) -52* Qi et al., 2009 [96]

B10 (Rapeseed oil biodiesel) -24 Kousoulidou et al., 2010 [97]

B20 (Rapeseed oil biodiesel) ~0 Turrio-Baldassarri et al., 2004 [86]

B100 (Rapeseed oil methyl ester) -50 Krahl et al., 2007 [98]

B100 (mixture of Rapeseed and Recycled

Cooking oil methyl ester)

-80 to -65* Grimaldi et al., 2002 [99]

B100 (Karanja oil biodiesel) -50* Raheman et al., 2004 [100]

B10 (Palm oil biodiesel) -17 Kousoulidou et al., 2010 [97]

Ethanol: Methyl soyate: Diesel (5:20:75) blend Shi et al., 2006 [101]

B100 (Yellow grease biodiesel) -64* Canaki et al., 2003 [95]

B50 -12* Xiaoming et al., 2005 [102]

B20 -13* Xiaoming et al., 2005 [102]

*: Change in smoke opacity 634

Another study reported that both, B100 and B20 reduced PM emissions, and PM 635

reduction is generally independent of the feedstock used for biodiesel production. Even 636

lower blends of biodiesel were effective in reducing PM mass emissions [103]. Lower 637

particulate emissions were also seen for biodiesel fuelled engine fitted with an oxidation 638

catalyst [104]. Reduction in smoke was also reported with increasing blend 639

concentration of Linseed oil biodiesel in another study [105] and higher reduction in 640

smoke level was reported at higher engine loads [106]. Lower PM mass emissions were 641

also reported for 100% Palm oil biodiesel vis-a-vis mineral diesel [107]. 642

Page 33: Elsevier Editorial System(tm) for Energy Conversion and

29

643

Figure 9: Effect of biodiesel blending on PM emissions [108] 644

A review article on criteria pollutants from biodiesel engines concluded that increased 645

biodiesel concentration in test fuel in a heavy-duty diesel engine application reduces 646

HC, CO, and PM emissions substantially, along with slightly increased NOx emissions 647

[108]. Figure 9 shows these findings for heavy-duty and medium-duty engines. Results 648

from light-duty engines were more fluctuating, and showed some increase in CO, PM 649

and NOx emissions with increasing biodiesel concentration in the test fuel. Another 650

scientific study reported that PM mass emissions mostly decreased with increasing 651

biodiesel content in the test fuel at all test modes, reaching peak reduction of ~49-62% 652

with 100% biodiesel vis-a-vis baseline diesel [30]. Saanum et al. [109] reported that PM 653

emissions were slightly higher for biodiesel than diesel (marine gas oil: MGO) at low 654

engine loads, but the trend reverses at higher engine loads and biodiesel emits lower PM 655

mass emissions compared to mineral diesel. The filter smoke number (FSN) was 656

significantly lower for biodiesel at all loads tested. 657

3.1.1 Effect of Injection Strategies 658

It is desirable to have superior combustion in diesel engines to have lower particulate 659

emissions. This can be achieved by optimizing injection parameters such as fuel 660

injection pressure, injection timing etc. A diesel engine with mechanical pump system 661

does not provide flexibility in injection therefore modern diesel engines are equipped 662

Page 34: Elsevier Editorial System(tm) for Energy Conversion and

30

with CRDI system, which provides excellent flexibility for fuel injection such as high 663

pressure (up to 2000 bar) injection, split and multiple injections, injection duration 664

control and injection timing control. PM mass emissions can be drastically reduced by 665

optimizing fuel injection parameters [110]. Many studies have reported significant 666

reduction in PM mass emissions upon using modern high pressure flexible fuel injection 667

systems. This section provides insights into the effect of different injection strategies on 668

particulate emission from CI engines fuelled with mineral diesel and biodiesel. 669

Suryawanshi et al. [111] reported that for various blends of Karanja oil methyl ester, 670

smoke opacity was lower compared to mineral diesel operation. Smoke was further 671

reduced by retarded injection timings. Corgard et al. [112] observed that utilization of 672

biodiesel blends resulted in lower smoke emission from a HSDI diesel engine. They 673

retarded the injection timings for reducing increased NOx levels due to use of biodiesel. 674

Nearly two-third reduction in smoke produced by mineral diesel was achieved by 30% 675

biodiesel blend with retarded injection timings [112]. 676

Combustion chamber visualization and computational studies on soot formation 677

confirmed that it gets accumulates in the spray tip region [113-116]. In single injection 678

pulses, high momentum of injected spray droplets for longer duration continuously 679

results in supply of fuel droplets to relatively low temperature region of the combustion 680

chamber in comparison to split injection, which results in higher soot formation [113]. In 681

a split-injection mode, the fuel injected in the second pulse, wherein fuel droplets enter 682

into a relatively fuel-lean and high-temperature region formed due to the combustion of 683

the fuel injected during the first injection pulse. In split injection, soot formation is 684

considerably reduced because the fuel injected is rapidly consumed in combustion 685

process before it starts accumulating in a fuel-rich soot-producing zone [113]. For 686

achieving effective reduction in soot formation, split injection, and optimization of time 687

interval between the two injection pulses is very critical. Separation between two 688

Page 35: Elsevier Editorial System(tm) for Energy Conversion and

31

injection pulses should be long enough so that the soot formation zone of the first 689

injection pulse is not replenished with the incoming fuel from the second injection pulse. 690

Separation between the two injection pulses should be short enough to ensure that high 691

temperature and pressure conditions are available to the incoming fuel droplet for its 692

rapid combustion, resulting in lower soot formation [113]. 693

Yamane et al. [117] observed that PM emissions from biodiesel showed a high level of 694

SOF compared to mineral diesel (gas oil) at lower engine loads. They investigated spray 695

jet penetration of biodiesel and mineral diesel corresponding to injection conditions 696

prevailing at lower engine loads. They observed shorter spray penetration for biodiesel 697

due to its higher kinematic viscosity and density, which resulted in inferior air-fuel 698

mixing [117]. Inferior air-fuel mixing was reported to be the main reason for higher PM 699

mass emissions. Ye et al. [118] concluded that impact of injection strategy and biodiesel 700

fueling on PM mass emissions strongly depends on the engine load in a CRDI engine. 701

They also suggested that use of biodiesel and increased fuel injection pressure 702

effectively reduced PM emissions at low load conditions, however biodiesel didn't show 703

significant effect at moderate and higher engine loads on PM mass emissions. Yehliu et 704

al. [119] reported increased brake specific PM emissions with B100 vis-a-vis diesel at 705

some operating conditions but reduction at other conditions for both single and split 706

injection mode in a CRDI engine. At some operating conditions, increased PM mass was 707

attributed to increase in SOF fraction of particulates due to relatively lower volatility of 708

biodiesel in comparison to mineral diesel. Particulates from B100 mainly consisted of 709

condensed organics, because the particle number concentration dropped dramatically in 710

comparison to particle concentration for mineral diesel, when a thermo-denuder was set 711

to 400°C. instead of 30°C, in order to remove organics [119]. Kegl et al. [120] reported 712

50% reduction in smoke along with other regulated pollutants for a B100 fuelled bus 713

Page 36: Elsevier Editorial System(tm) for Energy Conversion and

32

engine at a retarded fuel injection timing (19° bTCD) in comparison to mineral diesel 714

engine at a normal fuel injection timing (23° bTDC) in an ESC test-cycle. 715

3.1.2 Effect of Fuel Composition 716

After 1970's energy crisis, many researchers started exploring a suitable alternative fuel 717

for mineral diesel. Biodiesel was considered as one of the potential alternative fuels for 718

diesel. One major advantages of biodiesel is that it has inherent oxygen content in the 719

fuel molecule itself, which enhances the probability of complete combustion. On the 720

other hand, biodiesel's evaporative property are not as good as mineral diesel, therefore 721

it has low evaporation rate at relatively lower engine loads, because of lower in-cylinder 722

temperatures. Shi et al. [101] reported significant reduction in PM emissions from blend 723

of ethanol:methylsoyate:diesel (5:20:75) compared to that from mineral diesel [101]. Zhu 724

et al. [18] observed that PM emissions from biodiesel fuelled engine operation were 725

lower than mineral diesel fuelled engine. PM emissions further reduced with an 726

increasing ethanol/ methanol concentration in biodiesel-alcohol blend at medium and 727

high engine loads. PM reduction by addition of alcohols to biodiesel was due to higher 728

oxygen content of alcohol-biodiesel blend in comparison to biodiesel, which improved the 729

combustion process and reduced PM emissions. Also, alcohol in the blended fuel reduced 730

the cetane number hence increased ignition delay period therefore higher fuel quantity 731

burned in premixed combustion phase, resulting in lower PM emissions [18]. However, 732

15% alcohol blends lead to higher PM emissions than biodiesel (B100) and mineral 733

diesel at low loads [18]. Yoon et al. [121] observed significantly lower filter smoke 734

number (FSN) for biodiesel-ethanol blend (90:10) in comparison to mineral diesel in 735

double injection strategy. Lower soot emissions were primarily due to higher oxygen 736

content of the blended fuel and absence of soot precursors (sulfur and aromatics 737

contents) [121]. Particle number concentration of larger particles, which contribute 738

dominatingly to PM mass for biodiesel-ethanol blend, were significantly lower than 739

Page 37: Elsevier Editorial System(tm) for Energy Conversion and

33

mineral diesel [121]. Lu et al. [122] reported that in his experiments of port injection of 740

ethanol in a biodiesel fuelled engine, CO and HC emissions increased compared to 741

biodiesel (B100) operated engine, and 35–85% reduction in NOx and smoke was 742

observed. Kohoutek et al. [123] used ULSD and got 30% lower PM mass and number 743

emissions in comparison to 410 ppm sulfur containing diesel in a modern DI diesel 744

engine. At high engine load, specific emission of PM1.8 from biodiesel decreased by 68.4% 745

and 50.3%, compared with LSD and ULSD respectively [60]. PM mass emission reduced 746

by 20.6% upon using biodiesel compared to LSD and was slightly on the higher side as 747

compared to ULSD at lower engine loads. 748

Dwivedi et al. [124] reported that total PM mass reduced by ~20% with B20 in 749

comparison to mineral diesel in a CIDI engine. Due to better lubricity of biodiesel, trace 750

metal content in the particulate also reduced for B20. Benzene soluble organic fraction 751

(BSOF) was found to be higher for B20. Kim et al. [125] also found 20% lower PM mass 752

for B20 fuelled engine in comparison to mineral diesel. 15% biodiesel and 5% ethanol 753

blended with mineral diesel reduced PM mass further. Important observation was that 754

total number of particles reduced for biodiesel but number of nuclei mode particles was 755

higher for biodiesel compared to mineral diesel. Rakopoulos et al. [126] reported that 756

peak value of smoke opacity reduced by 40% and 73% respectively for the biodiesel and 757

n-butanol blends during transient tests in a CI engine. Relative fuel-bound oxygen plays 758

dominant role here. In a review, Graboski et al. [127] concluded that PM mass emission 759

reduction was proportional to the fuel oxygen content as long as cetane number was 760

higher than 45 or density was lower than 0.89 kg/ l. 761

Page 38: Elsevier Editorial System(tm) for Energy Conversion and

34

762

Figure 10: Effect of carbon chain length on total PM mass emissions (450 bar injection 763

pressure, 4 bar IMEP, 7.1°bTDC SOI) [23] 764

Schönborn et al. [23] studied the effect of carbon chain length and degree of 765

unsaturation on PM mass emissions. Total particulate mass emitted by mineral diesel 766

was higher than biodiesel (Figure 10). Lower particulate mass emission from biodiesel 767

are attributed to higher fuel oxygen content, which helps in oxidation of soot particles 768

and soot precursors. The behenic acid methyl ester (22 carbon atoms fatty acid chain) 769

showed a distinctively higher emission of PM mass than other lower fatty acid chain 770

length molecules. It is possibly due to spray formation and fuel-air mixing being affected 771

by fuel's high viscosity and low volatility. 772

3.2 Particulate Size-Number Distribution 773

Most emission regulations globally prescribe PM mass measurement and control. Size 774

affects the behavior of particulates in the engine as well as in the environment [15]. 775

Adverse health effects due to particulates are more severe for smaller nuclei mode 776

particulates. This fact has been recognized and new emission legislations, beyond Euro-777

5 also prescribe limit on total particle number concentration along with particulate 778

mass. Table 4 shows typical contribution of different size particles emitted by diesel 779

engines to PM mass and total numbers. Almost 90% particulate emitted from diesel 780

engines originates as nuclei mode particles. Parihar et al. [82] suggested that 781

contribution made by PM2.5 particles to the total PM mass varies from 75-95%, 782

Page 39: Elsevier Editorial System(tm) for Energy Conversion and

35

depending on engine load, for mineral diesel as well as B20. Particles larger than 10 µm 783

contributed up to 10-15% to cumulative mass whereas particles in size range 2.5-10 µm 784

contribute up to 20-30% of the total PM mass. Li et al. [22] reported that densities of 785

particles smaller than 50 nm varied from 1.1 g/cm3 to 1.6 g/cm3. For particles having 786

mobility diameter closer to 1 µm, particle densities varied from 0.2 to 0.6 g/cm3. Hence, 787

methods employed to control PM mass do not necessarily result in particulate number 788

reduction. In this scenario, it is important to characterize the effect of biodiesel on 789

particulate number emissions. 790

Table 4: Size based classification of engine exhaust particulates [1, 15] 791

Size (nm) % Number % Mass

Nuclei mode 5-50 ~90 1-20

Accumulation mode 50-700 1-10 60-94

Coarse mode 700-10000 0-2 5-20

792

Kawano et al. [128] reported that particle size distribution of both diesel and RME was 793

mono-modal, and this distribution mostly covered accumulation mode particles for all 794

engine loads. The size distribution of accumulation mode particles of RME shifted 795

towards smaller sizes compared to diesel, and peak position of particle size distribution 796

was almost constant for varying engine loads. An increase in engine load increased the 797

peak number concentration in case of diesel and reduced the peak concentration for 798

RME [128] and these findings were in line with the other studies which suggested that 799

biodiesel particulates are mostly comprised of SOF, which gets destroyed in engine in 800

high temperature conditions prevailing at higher engines loads. Raahede [129] reported 801

a decreasing trend for particle number concentration, when moving from reference 802

diesel to biodiesel (B20) and this trend was attributed to higher fuel bound oxygen. 803

Puzun et al. [130] studied the particle size-number distribution of rapeseed biodiesel 804

blends and mineral diesel in a high-pressure CRDI engine. They reported that 805

Page 40: Elsevier Editorial System(tm) for Energy Conversion and

36

particulate sizes emitted from this engine were mostly smaller than 300 nm [130]. For 806

diesel, particle number concentrations showed single peak distributions dominating 807

accumulation mode particles [130]. For biodiesel blends at lower and intermediate loads, 808

double peak in particle size-number distribution was seen with increasing concentration 809

of biodiesel in the fuel and nuclei mode particle numbers increased significantly; 810

particles with sizes more than 50 nm (accumulation mode) decreased and peak of 811

number concentration shifted towards smaller particles [130]. Tan et al. [27] reported 812

that peak value of nucleation mode particles below 30 nm increased with increasing 813

biodiesel concentration in fuel blend. Nucleation mode particle size peak becomes larger 814

with increasing biodiesel blend ratio, and accumulation mode particle size peak value 815

becomes smaller. Three mechanisms were thought to lead to greater nucleation mode 816

particle formation: (i) high super-saturation encourages formation of new particles by 817

nucleation at less solid soot surface, (ii) increased viscosity and lower volatility of 818

biodiesel causes higher SOF, and (iii) oxygen content of biodiesel causes carbonaceous 819

particle to transform from fine particles to ultra-fine particles or nano-particles. 820

Agarwal et al. [5] reported that particulate number emissions for B100 were higher than 821

mineral diesel however, these were comparable for B20 and diesel at lower loads. Jung 822

et al. [13] investigated the effect of biodiesel on oxidation of particulates using soy 823

methyl ester (SME) and diesel (#2) at 1400 rpm engine speed and 75% load. 824

Accumulation mode particle number concentration and particle volume distribution 825

were lowered by 38% and 82%, respectively for SME. They also reported lower particle 826

numbers than mineral diesel in large particle size range above 50 nm, but very similar 827

numbers was observed in nuclei region below 50 nm. This indicated towards lower PM 828

mass for B100 because of lower soot pyrolysis due to presence of fuel oxygen. High 829

numbers of engine out nuclei mode particles with B100 are considered to be due to 830

presence of large number of condensed droplets of high boiling point hydrocarbons, 831

Page 41: Elsevier Editorial System(tm) for Energy Conversion and

37

which are also responsible for high VOF. They reported that rate of oxidation of SME 832

particulates was 6 times faster than diesel particulates [13], which verified the above 833

hypothesis. 834

3.2.1 Effect of Injection Strategy 835

Most modern engines are now equipped with advanced fuel injection systems controlled 836

by electronic control unit (ECU). These injection systems are designed in order to reduce 837

total PM mass emissions. However there is strong need to investigate the particle 838

number emissions also for varying injection parameters because it is proved beyond a 839

reasonable doubt that finer diesel particulates have higher toxicity. Several studies have 840

been carried out to study the particle number emissions from different types of engines, 841

fuels and engine operating conditions and some of them are discussed in this section. 842

Desantes et al. [131] investigated the effect of fuel injection pressure, SOI timings and 843

EGR on engine exhaust particle size-number distribution from a heavy-duty diesel 844

engine [131]. Increasing fuel injection pressure reduced number of accumulation mode 845

particle and favored nuclei mode particle formation. Increasing fuel injection pressure 846

improved air-fuel mixing, which reduced the spread of fuel-rich zones responsible for 847

formation of carbonaceous soot particles. Reduction in the concentration of carbonaceous 848

particles resulted in reduction in number of accumulation mode particles. Application of 849

EGR suppressed the nucleation mode particles and increased number of accumulation 850

mode particles. Advanced injection timings slightly reduced number of accumulation 851

mode particles without shifting the position peak concentration [131]. 852

Page 42: Elsevier Editorial System(tm) for Energy Conversion and

38

853

Figure 11: Effect of fuel injection pressure on number of nuclei mode particles (2-67nm) 854

for biodiesel blends [132] 855

Sinha et al. [132] reported increase in total number concentration of nuclei mode 856

particles with increasing fuel injection pressure for diesel as well as biodiesel blends 857

(Figure 11). As biodiesel percentage increases in the blend, it increases fuel's oxygen 858

content, leading for formation of fewer number of carbonaceous accumulation mode 859

particles. Since lower number of carbonaceous particles are available for adsorption of 860

SOF, the partial pressure of these organic fractions increases, leading to higher 861

numbers of nuclei mode particles. Also, biodiesel produces higher soluble organic 862

fractions, which add to the partial pressure of gaseous hydrocarbons forming SOFs, 863

further enhancing nucleation process. Sinha et al. [132] also reported reduced 864

accumulation mode particle numbers (50-1000 nm) at all fuel injection pressures except 865

600 bar from Soybean biodiesel (B100) vis-à-vis ULSD. 866

3.2.2 Effect of Fuel Composition 867

Several studies showed that biodiesel and mineral diesel have different particle size-868

number distributions [18,23,133]. It is important from regulatory norms stand point to 869

evaluate the particle number emission for biodiesel. Pham et al. [133] reported that 870

saturated short-chain length FAMEs reduce NOx and particulate number concentration, 871

Page 43: Elsevier Editorial System(tm) for Energy Conversion and

39

but led to higher BSFC as well as higher reactive oxygen species (ROS) emissions. 872

Unsaturated FAMEs emit lower particulate and ROS, but higher NOx. 873

Higher particulate mass (Figure 10) for diesel originated from larger number of 874

accumulation mode particles (Figure 12). 875

876

Figure 12: Effect of carbon chain length on particulate size distribution (450 bar 877

injection pressure, 4 bar IMEP, 7.1°bTDC SOI) [23] 878

Progressively higher number of nucleation mode particles was seen for fatty acid esters 879

with longer fatty acid chains (Figure 12). It is possible that the nuclei mode particles 880

consist of high boiling point constituents of fuel which remains unburned and condenses 881

in the exhaust gas [23]. Schönborn et al. also reported that increase in alcohol chain 882

length from 1 to 2 carbon atoms reduced NOx emissions, but increased total PM mass 883

emission, even under constant ignition delay and similar heat release conditions [23]. 884

Zhu et al. [18] compared the particulate size-number distribution for biodiesel and 885

biodiesel-alcohol blends with mineral diesel [18]. At all engine loads, total particulate 886

number concentration from biodiesel was higher than that from mineral diesel. Addition 887

of ethanol/ methanol in biodiesel reduced the total number concentration of particles 888

dramatically below the level of mineral diesel fuelled engine operation [18]. They also 889

suggested like many other researchers that blending alcohols with biodiesel decreases 890

carbon content and increases oxygen content of the fuel, leading to reduction in number 891

Page 44: Elsevier Editorial System(tm) for Energy Conversion and

40

of nuclei mode particles and total particulate number concentration [18,134]. Lapuerta 892

et al. [135] reported that higher unsaturation level of biodiesel led to retarded start of 893

combustion, higher NOx emissions, higher heat release rate, lower smoke opacity, lower 894

particulate mass emissions and lower particulate size distribution. Song et al. [29] 895

reported that number of nucleation mode particles decreased or remained constant, and 896

number of accumulation mode particles above 30 nm increased for oxidized biodiesel 897

blends compared to non-oxidized biodiesel blend. The particulate mass-size distribution 898

for oxidized biodiesel blend reduced by 5 - 23.4% compared to non-oxidized biodiesel 899

blend at all engine loads. Total particle number concentration for oxidized biodiesel 900

blend compared with non-oxidized biodiesel blend was found to be strongly dependent on 901

test conditions. Nuszkowski et al. [136] reported that use of cetane improving additives 902

resulted in lower number of particulates in entire particle size range measured in the 903

study. The reduced particle number concentrations were in the diameter range of 6-56 904

nm and 100-205 nm, respectively. Addition of biodiesel reduced the particle number 905

concentration in the diameter range of 6-56 nm and 100-487 nm and was not affected in 906

other size ranges during transient engine operation [136]. Sinha et al. [132] investigated 907

the effect of using biodiesel on particulate emissions from a HSDI engine. They observed 908

that particle number density increased and particle size-mass distribution decreased 909

with increasing blending ratio of biodiesel. Zhang et al. [25] studied particle size-number 910

distribution of diesel (1135 ppm sulfur) and biodiesel (64 ppm sulfur) blends. Number 911

concentration of nuclei mode particles were three orders of magnitude higher for B60 912

and other lower biodiesel blends including mineral diesel in comparison to B100. They 913

gave the hypothesis that in case of high sulfur containing fuel, hydrated H2SO4 nuclei 914

acts as precursor for nuclei mode particulate formation (when the blend concentration 915

was lower than B60). For B100, nuclei mode particulate concentration was low due to its 916

extremely low sulfur content (64ppm). 917

Page 45: Elsevier Editorial System(tm) for Energy Conversion and

41

4. Summary 918

It is essential to evaluate the effect of biodiesel on the particulate emissions from diesel 919

engines before it can be implemented on a large scale worldwide. Diesel particulate 920

consists of elemental carbon, organic carbon, trace metals, and organic compounds. Most 921

of the diesel particles are nano-particles. Their composition varies and depends strongly 922

upon engine operating conditions. A large number of researchers characterized biodiesel 923

particles for their physical and chemical characteristics and these studies are 924

summarized below: 925

1. Biodiesel has an advantage and emits lower particulate mass emissions compared to 926

mineral diesel for most engine technologies and all engine load-speed conditions. Most 927

studies suggested that there is a large reduction in total particulate mass emissions by 928

using biodiesel or biodiesel blends with mineral diesel. 20% blend of biodiesel with diesel 929

showed good result in terms of lower particulate mass emission. B100 further reduced 930

particulate emissions but not in the same proportion as that of B20. 931

2. Total particle number emissions also reduced with use of biodiesel but particle 932

number emissions near the nano-size range were higher for biodiesel. 933

3. Biodiesel is an oxygenated fuel, and fuel oxygen helps in improving combustion inside 934

the combustion chamber, resulting in lower PM mass emissions. Presence of fuel oxygen 935

reduces pyrolysis reactions in the combustion chamber. Pyrolysis of fuel and lubricating 936

oil in oxygen deficient regions of the combustion chamber is the main reason for 937

particulate formation in the engine. 938

4. Biodiesel shows lower organic carbon (OC) content of the particulate emitted by an 939

order of magnitude compared to mineral diesel. No significant reductions were observed 940

for elemental carbon (EC) content of the particulate. For biodiesel, most studies reported 941

higher EC/ OC ratio along with lower total particulate mass emissions, which indicated 942

to its lower environmental toxicity compared to mineral diesel. 943

Page 46: Elsevier Editorial System(tm) for Energy Conversion and

42

5. For biodiesel, generally PAH emissions were found to be relatively lower but some of 944

the specific PAHs adsorbed by particulate were slightly higher compared to mineral 945

diesel. This indicates towards presence of structurally strong PAHs in biodiesel. Possibly 946

these PAHs are difficult to oxidize in the engine combustion chamber, even under higher 947

temperature and pressure conditions. 948

6. Biodiesel helps particulate oxidation in DOC/ DPF. Biodiesel inherently contains 949

some trace metals, which possibly act as catalysts in the after-treatment devices and 950

lower particulate emissions. Lower particulate emissions from biodiesel also result in a 951

longer useful life of after-treatment devices. 952

Overall, Biodiesel emits relatively lower particulate mass emissions which have lesser 953

environmental and health related toxicity, and impacts the exhaust gas after-treatment 954

devices life positively, in addition to protecting the environmental, being a green fuel. 955

References 956

[1] B. Zielinska. Atmospheric transformation of diesel emissions. Experimental and 957

Toxicologic Pathology. 57 (2005) 31-42. 958

[2] O. Schroeder, J. Krahl, J. Bünger. Environmental and health effects caused by the 959

use of biodiesel. SAE Technical Paper 1999-01-3561 (1999). 960

[3] J. Krahl, A. Munack, O. Schröder, H. Stein, J. BüNGER. Influence of biodiesel and 961

different designed diesel fuels on the exhaust gas emissions and health effects. SAE 962

Technical Papers 2003-01-3199 (2003). 963

[4] J. Krahl, A. Munack, Y. Ruschel, O. Schröder, J. Bünger. Exhaust gas emissions and 964

mutagenic effects of diesel fuel, biodiesel and biodiesel blends. SAE Technical Papers 965

2008-01-2508 (2008). 966

[5] A.K. Agarwal, T. Gupta, A. Kothari. Particulate emissions from biodiesel vs diesel 967

fuelled compression ignition engine. Renewable and Sustainable Energy Reviews. 15 968

(2011) 3278-300. 969

Page 47: Elsevier Editorial System(tm) for Energy Conversion and

43

[6] A.K. Agarwal, T. Gupta, N. Dixit, P.C. Shukla. Assessment of toxic potential of 970

primary and secondary particulates/aerosols from biodiesel vis-a-vis mineral diesel 971

fuelled engine. Inhalation toxicology. 25 (2013) 325-32. 972

[7] A.K. Agarwal, T. Gupta, A. Kothari. Toxic potential evaluation of particulate matter 973

emitted from a constant speed compression ignition engine: A comparison between 974

straight vegetable oil and mineral diesel. Aerosol Science and Technology. 44 (2010) 724-975

33. 976

[8] P.F. Flynn, R.P. Durrett, G.L. Hunter, A.O. zur Loye, O. Akinyemi, J.E. Dec, et al. 977

Diesel combustion: an integrated view combining laser diagnostics, chemical kinetics, 978

and empirical validation. SAE Technical Paper 1999-01-0509 (1999). 979

[9] A. Dhar, A.K. Agarwal. Effect of Multiple Injections on Particulate Size-Number 980

Distributions in a Common Rail Direct Injection Engine Fueled with Karanja Biodiesel 981

Blends. SAE Technical Paper 2013-01-1554 (2013). 982

[10] X. Wang, Z. Huang, O.A. Kuti, W. Zhang, K. Nishida. Experimental and analytical 983

study on biodiesel and diesel spray characteristics under ultra-high injection pressure. 984

International journal of heat and fluid flow. 31 (2010) 659-66. 985

[11] J.F. Gamble, M.J. Nicolich, P. Boffetta. Lung cancer and diesel exhaust: an updated 986

critical review of the occupational epidemiology literature. Critical reviews in toxicology. 987

42 (2012) 549-98. 988

[12] J.B. Heywood. Internal combustion engine fundamentals. Mcgraw-hill New York 989

1988. 990

[13] H. Jung, D.B. Kittelson, M.R. Zachariah. Characteristics of SME biodiesel-fueled 991

diesel particle emissions and the kinetics of oxidation. Environmental science & 992

technology. 40 (2006) 4949-55. 993

[14] P. Eastwood. Particulate emissions from vehicles. John Wiley & Sons 2008. 994

Page 48: Elsevier Editorial System(tm) for Energy Conversion and

44

[15] D.B. Kittelson. Engines and nanoparticles: a review. Journal of Aerosol Science. 29 995

(1998) 575-88. 996

[16] R. Stevenson. The morphology and crystallography of diesel particulate emissions. 997

Carbon. 20 (1982) 359-65. 998

[17] P. Rounce, A. Tsolakis, A. York. Speciation of particulate matter and hydrocarbon 999

emissions from biodiesel combustion and its reduction by aftertreatment. Fuel. 96 (2012) 1000

90-9. 1001

[18] L. Zhu, C. Cheung, W. Zhang, Z. Huang. Emissions characteristics of a diesel engine 1002

operating on biodiesel and biodiesel blended with ethanol and methanol. Science of the 1003

Total Environment. 408 (2010) 914-21. 1004

[19] P. Williams, M. Abbass, G. Andrews, K. Bartle. Diesel particulate emissions. 1005

Combust Flame. Combustion and Flame. 75 (1989) 1-24. 1006

[20] S. Chuepeng, H. Xu, A. Tsolakis, M. Wyszynski, P. Price, R. Stone, et al. Particulate 1007

Emissions from a Common Rail Fuel Injection Diesel Engine with RME-based Biodiesel 1008

Blended Fuelling Using Thermo-gravimetric Analysis. SAE Technical Paper 2008-01-1009

0074 (2008). 1010

[21] L.H. Young, Y.J. Liou, M.T. Cheng, J.H. Lu, H.H. Yang, Y.I. Tsai, et al. Effects of 1011

biodiesel, engine load and diesel particulate filter on nonvolatile particle number size 1012

distributions in heavy-duty diesel engine exhaust. Journal of hazardous materials. 199 1013

(2012) 282-9. 1014

[22] H. Li, A. Lea-Langton, G. Andrews, M. Thompson, C. Musungo. Comparison of 1015

exhaust emissions and particulate size distribution for diesel, biodiesel and cooking oil 1016

from a heavy duty DI diesel engine. SAE Technical Paper 2008-01-0076 (2008). 1017

[23] A. Schönborn, N. Ladommatos, R. Allan, J. Williams, J. Rogerson. Effect of the 1018

molecular structure of individual fatty acid alcohol esters (biodiesel) on the formation of 1019

Page 49: Elsevier Editorial System(tm) for Energy Conversion and

45

NOx and particulate matter in the diesel combustion process. SAE International 1020

Journal of Fuels and Lubricants. 1 (2009) 849-72. 1021

[24] M. Lapuerta, O. Armas, J. Rodriguez-Fernandez. Effect of biodiesel fuels on diesel 1022

engine emissions. Progress in energy and combustion science. 34 (2008) 198-223. 1023

[25] X. Zhang, H. Wang, L. Li, Z. Wu. Characteristics of Particulates and Exhaust Gases 1024

Emissions of DI Diesel Engine Employing Common Rail Fuel System Fueled with Bio-1025

diesel Blends. Fuel. 2013 (2008) 03-19. 1026

[26] A. Soewono, S. Rogak. Morphology and microstructure of engine-emitted 1027

particulates. Diesel Engine. SAE Techinical Paper 2009-01-1906 (2009). 1028

[27] P.Q. Tan, D.M. Lou, Z.Y. Hu. Nucleation mode particle emissions from a diesel 1029

engine with biodiesel and petroleum diesel fuels. SAE Technical Paper 2010-01-0787 1030

(2010). 1031

[28] M. Tinsdale, P. Price, R. Chen. The impact of biodiesel on particle number, size and 1032

mass emissions from a Euro4 diesel vehicle. SAE Technical Paper 2010-01-0796 (2010). 1033

[29] H. Song, M.H. Lee, H. Kang, K. Kim. The effects of oxidation deterioration biodiesel 1034

on particulate emission from heavy duty diesel engine. SAE Technical Papers 2012-01-1035

1600 (2012). 1036

[30] M.-J. Poitras, D. Rosenblatt, T.W. Chan, G. Rideout. Impact of Varying Biodiesel 1037

Blends on Direct-Injection Light-Duty Diesel Engine Emissions. SAE Technical Paper 1038

2012-01-1313 (2012). 1039

[31] J.N. Gangwar, T. Gupta, A.K. Agarwal. Composition and comparative toxicity of 1040

particulate matter emitted from a diesel and biodiesel fuelled CRDI engine. Atmospheric 1041

Environment. 46 (2012) 472-81. 1042

[32] J. Schauer, C. Christensen, D. Kittelson, J. Johnson, W. Watts. Impact of ambient 1043

temperatures and driving conditions on the chemical composition of particulate matter 1044

Page 50: Elsevier Editorial System(tm) for Energy Conversion and

46

emissions from non-smoking gasoline-powered motor vehicles. Aerosol Science and 1045

Technology. 42 (2008) 210-23. 1046

[33] P. Pakbin, Z. Ning, J.J. Schauer, C. Sioutas. Characterization of particle bound 1047

organic carbon from diesel vehicles equipped with advanced emission control 1048

technologies. Environmental science & technology. 43 (2009) 4679-86. 1049

[34] Y.Y. Liu, T.C. Lin, Y.J. Wang, W.L. Ho. Carbonyl compounds and toxicity 1050

assessments of emissions from a diesel engine running on biodiesels. Journal of the Air 1051

& Waste Management Association. 59 (2009) 163-71. 1052

[35] L.K. Limbach, P. Wick, P. Manser, R.N. Grass, A. Bruinink, W.J. Stark. Exposure of 1053

engineered nanoparticles to human lung epithelial cells: influence of chemical 1054

composition and catalytic activity on oxidative stress. Environmental science & 1055

technology. 41 (2007) 4158-63. 1056

[36] K. Donaldson, L. Tran, L.A. Jimenez, R. Duffin, D.E. Newby, N. Mills, et al. 1057

Combustion-derived nanoparticles: a review of their toxicology following inhalation 1058

exposure. Particle and Fibre Toxicology. 2 (2005) 10. 1059

[37] N. Künzli, I.S. Mudway, T. Götschi, T. Shi, F.J. Kelly, S. Cook, et al. Comparison of 1060

oxidative properties, light absorbance, and total and elemental mass concentration of 1061

ambient PM2. 5 collected at 20 European sites. Environmental health perspectives. 114 1062

(2006) 684. 1063

[38] A. Baulig, M. Garlatti, V. Bonvallot, A. Marchand, R. Barouki, F. Marano, et al. 1064

Involvement of reactive oxygen species in the metabolic pathways triggered by diesel 1065

exhaust particles in human airway epithelial cells. American Journal of Physiology-1066

Lung Cellular and Molecular Physiology. 285 (2003) L671-L9. 1067

[39] B. González-Flecha. Oxidant mechanisms in response to ambient air particles. 1068

Molecular aspects of medicine. 25 (2004) 169-82. 1069

Page 51: Elsevier Editorial System(tm) for Energy Conversion and

47

[40] A. Pillay, M. Elkadi, S. Fok, S. Stephen, J. Manuel, M. Khan, et al. A comparison of 1070

trace metal profiles of neem biodiesel and commercial biofuels using high performance 1071

ICP-MS. Fuel. 97 (2012) 385-9. 1072

[41] R. Betha, R. Balasubramanian. Emissions of particulate-bound elements from 1073

stationary diesel engine: Characterization and risk assessment. Atmospheric 1074

Environment. 45 (2011) 5273-81. 1075

[42] M. Carrales, R.W. Martin. Sulfur content of crude oils. US Department of the 1076

Interior, Bureau of Mines 1975. 1077

[43] K. Ravindra, R. Sokhi, R. Van Grieken. Atmospheric polycyclic aromatic 1078

hydrocarbons: source attribution, emission factors and regulation. Atmospheric 1079

Environment. 42 (2008) 2895-921. 1080

[44] J.D. McDonald, E.B. Barr, R.K. White, J.C. Chow, J.J. Schauer, B. Zielinska, et al. 1081

Generation and characterization of four dilutions of diesel engine exhaust for a 1082

subchronic inhalation study. Environmental science & technology. 38 (2004) 2513-22. 1083

[45] J.J. Schauer, M.J. Kleeman, G.R. Cass, B.R. Simoneit. Measurement of emissions 1084

from air pollution sources. 2. C1 through C30 organic compounds from medium duty 1085

diesel trucks. Environmental science & technology. 33 (1999) 1578-87. 1086

[46] J.J. Schauer, M.J. Kleeman, G.R. Cass, B.R. Simoneit. Measurement of emissions 1087

from air pollution sources. 5. C1-C32 organic compounds from gasoline-powered motor 1088

vehicles. Environmental science & technology. 36 (2002) 1169-80. 1089

[47] D. Grosjean, E. Grosjean, A.W. Gertler. On-road emissions of carbonyls from light-1090

duty and heavy-duty vehicles. Environmental science & technology. 35 (2001) 45-53. 1091

[48] A. Kristensson, C. Johansson, R. Westerholm, E. Swietlicki, L. Gidhagen, U. 1092

Wideqvist, et al. Real-world traffic emission factors of gases and particles measured in a 1093

road tunnel in Stockholm, Sweden. Atmospheric Environment. 38 (2004) 657-73. 1094

Page 52: Elsevier Editorial System(tm) for Energy Conversion and

48

[49] K.W. Loeffler, C.A. Koehler, N.M. Paul, D.O. De Haan. Oligomer formation in 1095

evaporating aqueous glyoxal and methyl glyoxal solutions. Environmental science & 1096

technology. 40 (2006) 6318-23. 1097

[50] M.C. Madden, L.A. Dailey, J.G. Stonehuerner, D.B. Harris. Responses of cultured 1098

human airways epithelial cells treated with diesel exhaust extracts will vary with the 1099

engine load. Journal of Toxicology and Environmental Health Part A. 66 (2003) 2281-97. 1100

[51] X. Pang, X. Shi, Y. Mu, H. He, S. Shuai, H. Chen, et al. Characteristics of carbonyl 1101

compounds emission from a diesel-engine using biodiesel–ethanol–diesel as fuel. 1102

Atmospheric Environment. 40 (2006) 7057-65. 1103

[52] S.S.H. Ho, K.F. Ho, S.C. Lee, Y. Cheng, J.Z. Yu, K.M. Lam, et al. Carbonyl 1104

emissions from vehicular exhausts sources in Hong Kong. Journal of the Air & Waste 1105

Management Association. 62 (2012) 221-34. 1106

[53] C. Cheung, L. Zhu, Z. Huang. Regulated and unregulated emissions from a diesel 1107

engine fueled with biodiesel and biodiesel blended with methanol. Atmospheric 1108

Environment. 43 (2009) 4865-72. 1109

[54] Y. Di, C. Cheung, Z. Huang. Experimental investigation on regulated and 1110

unregulated emissions of a diesel engine fueled with ultra-low sulfur diesel fuel blended 1111

with biodiesel from waste cooking oil. Science of the Total Environment. 407 (2009) 835-1112

46. 1113

[55] K. Takada, F. Yoshimura, Y. Ohga, J. Kusaka, Y. Daisho. Experimental study on 1114

unregulated emission characteristics of turbocharged DI diesel engine with common rail 1115

fuel injection system. SAE Technical Paper 2003-01-3158 (2003). 1116

[56] R. Ballesteros, J. Hernandez, L. Lyons, B. Cabanas, A. Tapia. Speciation of the 1117

semivolatile hydrocarbon engine emissions from sunflower biodiesel. Fuel. 87 (2008) 1118

1835-43. 1119

Page 53: Elsevier Editorial System(tm) for Energy Conversion and

49

[57] S. Machado Corrêa, G. Arbilla. Carbonyl emissions in diesel and biodiesel exhaust. 1120

Atmospheric Environment. 42 (2008) 769-75. 1121

[58] J. Xue, T.E. Grift, A.C. Hansen. Effect of biodiesel on engine performances and 1122

emissions. Renewable and Sustainable Energy Reviews. 15 (2011) 1098-116. 1123

[59] A. Lea-Langton, H. Li, G.E. Andrews. Comparison of particulate PAH emissions for 1124

diesel, biodiesel and cooking oil using a heavy duty DI diesel engine. SAE Technical 1125

Paper 2008-01-1811 (2008). 1126

[60] T. Lu, Z. Huang, C. Cheung, J. Ma. Size distribution of EC, OC and particle-phase 1127

PAHs emissions from a diesel engine fueled with three fuels. Science of the Total 1128

Environment. 438 (2012) 33-41. 1129

[61] G. Karavalakis, S. Stournas, E. Bakeas. Effects of diesel/biodiesel blends on 1130

regulated and unregulated pollutants from a passenger vehicle operated over the 1131

European and the Athens driving cycles. Atmospheric Environment. 43 (2009) 1745-52. 1132

[62] J. Pan, S. Quarderer, T. Smeal, C. Sharp. Comparison of PAH and nitro-PAH 1133

emissions among standard diesel fuel, biodiesel fuel, and their blend on diesel engines. 1134

Proceedings of the 48 th ASMS Conference on Mass Spectrometry and Allied Topics, 1135

Long Beach, CA 2000. 1136

[63] I.C. Nisbet, P.K. LaGoy. Toxic equivalency factors (TEFs) for polycyclic aromatic 1137

hydrocarbons (PAHs). Regulatory toxicology and pharmacology. 16 (1992) 290-300. 1138

[64] G. Karavalakis, G. Fontaras, D. Ampatzoglou, M. Kousoulidou, S. Stournas, Z. 1139

Samaras, et al. Effects of low concentration biodiesel blends application on modern 1140

passenger cars. Part 3: Impact on PAH, nitro-PAH, and oxy-PAH emissions. 1141

Environmental Pollution. 158 (2010) 1584-94. 1142

[65] E. Bakeas, G. Karavalakis, G. Fontaras, S. Stournas. An experimental study on the 1143

impact of biodiesel origin on the regulated and PAH emissions from a Euro 4 light-duty 1144

vehicle. Fuel. 90 (2011) 3200-8. 1145

Page 54: Elsevier Editorial System(tm) for Energy Conversion and

50

[66] M. Rhead, S. Hardy. The sources of polycyclic aromatic compounds in diesel engine 1146

emissions. Fuel. 82 (2003) 385-93. 1147

[67] H. Richter, J. Howard. Formation of polycyclic aromatic hydrocarbons and their 1148

growth to soot—a review of chemical reaction pathways. Progress in energy and 1149

combustion science. 26 (2000) 565-608. 1150

[68] S.G. Riddle, C.A. Jakober, M.A. Robert, T.M. Cahill, M.J. Charles, M.J. Kleeman. 1151

Large PAHs detected in fine particulate matter emitted from light-duty gasoline 1152

vehicles. Atmospheric Environment. 41 (2007) 8658-68. 1153

[69] K. Miet, K. Le Menach, P.-M. Flaud, H. Budzinski, E. Villenave. Heterogeneous 1154

reactivity of pyrene and 1-nitropyrene with NO2: Kinetics, product yields and 1155

mechanism. Atmospheric Environment. 43 (2009) 837-43. 1156

[70] N.V. Heeb, P. Schmid, M. Kohler, E. Gujer, M. Zennegg, D. Wenger, et al. 1157

Secondary effects of catalytic diesel particulate filters: conversion of PAHs versus 1158

formation of nitro-PAHs. Environmental science & technology. 42 (2008) 3773-9. 1159

[71] S.T. Bagley, L.D. Gratz, J.H. Johnson, J.F. McDonald. Effects of an oxidation 1160

catalytic converter and a biodiesel fuel on the chemical, mutagenic, and particle size 1161

characteristics of emissions from a diesel engine. Environmental science & technology. 1162

32 (1998) 1183-91. 1163

[72] P.Q. Tan, Z.Y. Hu, D.M. Lou. Regulated and unregulated emissions from a light-1164

duty diesel engine with different sulfur content fuels. Fuel. 88 (2009) 1086-91. 1165

[73] C. Cheung, Y. Di, Z. Huang. Experimental investigation of regulated and 1166

unregulated emissions from a diesel engine fueled with ultralow-sulfur diesel fuel 1167

blended with ethanol and dodecanol. Atmospheric Environment. 42 (2008) 8843-51. 1168

[74] E. Zervas. Regulated and non-regulated pollutants emitted from two aliphatic and a 1169

commercial diesel fuel. Fuel. 87 (2008) 1141-7. 1170

Page 55: Elsevier Editorial System(tm) for Energy Conversion and

51

[75] P.C. Shukla, T. Gupta, A.K. Agarwal. A Comparative Morphological Study of 1171

Primary and Aged Particles Emitted from a Biodiesel (B20) vis-à-vis Diesel Fuelled 1172

CRDI Engine. Aerosol and Air Quality Research. 14 (2014) 934-42. 1173

[76] X. Shi, K. He, W. Song, X. Wang, J. Tan. Effects of a diesel oxidation catalyst on 1174

gaseous pollutants and fine particles from an engine operating on diesel and biodiesel. 1175

Frontiers of Environmental Science & Engineering. 6 (2012) 463-9. 1176

[77] L. Zhu, C. Cheung, W. Zhang, J. Fang, Z. Huang. Effects of ethanol–biodiesel blends 1177

and diesel oxidation catalyst (DOC) on particulate and unregulated emissions. Fuel. 113 1178

(2013) 690-6. 1179

[78] A. Williams, R. McCormick, J. Luecke, R. Brezny, A. Geisselmann, K. Voss, et al. 1180

Impact of Biodiesel Impurities on the Performance and Durability of DOC, DPF and 1181

SCR Technologies. SAE International Journal of Fuels and Lubricants. 4 (2011) 110-24. 1182

[79] K. Vertin, S. He, A. Heibel. Impacts of B20 Biodiesel on Cordierite Diesel 1183

Particulate Filter Performance. SAE Technical Paper 2009-01-2736 (2009). 1184

[80] G. Austin, J. Naber, J. Johnson, C. Hutton. Effects of Biodiesel Blends on 1185

Particulate Matter Oxidation in a Catalyzed Particulate Filter during Active 1186

Regeneration. SAE International Journal of Fuels and Lubricants. 3 (2010) 194-218. 1187

[81] M. Asti, E.M. Borla, F. Parussa, C.R.F. Scpa. Biodiesel Influence on Particulate 1188

Matter Behavior during Active and Passive DPF Regeneration. SAE Technical Paper 1189

2011-24-0204 (2011). 1190

[82] A. Parihar, V. Sethi, P. Parikh, D. Radhakrishna. Physical Characterization of 1191

Particulate Emissions from Compression Ignition Engine Operating on Diesel and 1192

Biodiesel Fuels. SAE Technical Paper 2011-26-0026 (2011). 1193

[83] S. Di Iorio, C. Beatrice, C. Guido, P. Napolitano, A. Vassallo, C. Ciaravino. Impact 1194

of Biodiesel on Particle Emissions and DPF Regeneration Management in a Euro5 1195

Automotive Diesel Engine. SAE Technical Paper 2012-01-0839 (2012). 1196

Page 56: Elsevier Editorial System(tm) for Energy Conversion and

52

[84] J. Pidgeon, J. Johnson, J. Naber. An Experimental Investigation into Particulate 1197

Matter Oxidation in a Catalyzed Particulate Filter with Biodiesel Blends on an Engine 1198

during Active Regeneration. SAE Technical Paper 2013-01-0521 (2013). 1199

[85] EPA. A comprehensive analysis of biodiesel impacts on exhaust emissions. 1200

Assessment and Standards Division (Office of Transportation and Air Quality of the US 1201

Environmental Protection Agency). Report EPA420-P-02-001 (2002). 1202

[86] L. Turrio-Baldassarri, C.L. Battistelli, L. Conti, R. Crebelli, B. De Berardis, A.L. 1203

Iamiceli, et al. Emission comparison of urban bus engine fueled with diesel oil and 1204

‘biodiesel’blend. Science of the Total Environment. 327 (2004) 147-62. 1205

[87] Y.-f. Lin, Y.-p.G. Wu, C.-T. Chang. Combustion characteristics of waste-oil produced 1206

biodiesel/diesel fuel blends. Fuel. 86 (2007) 1772-80. 1207

[88] M.N. Nabi, M.M.Z. Shahadat, M.S. Rahman, M.R.A. Beg. Behavior of diesel 1208

combustion and exhaust emission with neat diesel fuel and diesel-biodiesel blends. SAE 1209

Technical Paper 2004-01-3034 (2004). 1210

[89] M. Alam, J. Song, R. Acharya, A. Boehman, K. Miller. Combustion and emissions 1211

performance of low sulfur, ultra low sulfur and biodiesel blends in a DI diesel engine. 1212

SAE Technical Paper 2004-01-3024 (2004). 1213

[90] R. McGill, J. Storey, R. Wagner, D. Irick, P. Aakko, M. Westerholm, et al. Emission 1214

performance of selected biodiesel fuels. SAE Technical Paper 2003-01-1866 (2003). 1215

[91] C.A. Sharp, T.W. Ryan, G. Knothe. Heavy-duty diesel engine emissions tests using 1216

special biodiesel fuels. SAE Technical Paper 2005-01-3671 (2005). 1217

[92] B.M. Spessert, I. Arendt, A. Schleicher. Influence of RME and Vegetable Oils on 1218

Exhaust Gas and Noise Emissions of Small Industrial Diesel Engines. SAE Technical 1219

Paper 2004-32-0070 (2004). 1220

Page 57: Elsevier Editorial System(tm) for Energy Conversion and

53

[93] W. Wang, D. Lyons, N. Clark, M. Gautam, P. Norton. Emissions from nine heavy 1221

trucks fueled by diesel and biodiesel blend without engine modification. Environmental 1222

science & technology. 34 (2000) 933-9. 1223

[94] M. Canakci. Combustion characteristics of a turbocharged DI compression ignition 1224

engine fueled with petroleum diesel fuels and biodiesel. Bioresource technology. 98 1225

(2007) 1167-75. 1226

[95] M. Canakci, J.H. Van Gerpen. Comparison of engine performance and emissions for 1227

petroleum diesel fuel, yellow grease biodiesel, and soybean oil biodiesel. Transactions of 1228

the ASAE. 46 (2003) 937-44. 1229

[96] D. Qi, L. Geng, H. Chen, Y. Bian, J. Liu, X. Ren. Combustion and performance 1230

evaluation of a diesel engine fueled with biodiesel produced from soybean crude oil. 1231

Renewable Energy. 34 (2009) 2706-13. 1232

[97] M. Kousoulidou, G. Fontaras, L. Ntziachristos, Z. Samaras. Biodiesel blend effects 1233

on common-rail diesel combustion and emissions. Fuel. 89 (2010) 3442-9. 1234

[98] J. Krahl, A. Munack, O. Schröder, H. Stein, L. Herbst, A. Kaufmann, et al. The 1235

influence of fuel design on the exhaust gas emissions and health effects. SAE Technical 1236

Paper 2005-01-3772 (2005). 1237

[99] C.N. Grimaldi, L. Postrioti, M. Battistoni, F. Millo. Common rail HSDI diesel 1238

engine combustion and emissions with fossil/bio-derived fuel blends. Evaluation. 2011 1239

(2002) 12-20. 1240

[100] H. Raheman, A. Phadatare. Diesel engine emissions and performance from blends 1241

of karanja methyl ester and diesel. Biomass and Bioenergy. 27 (2004) 393-7. 1242

[101] X. Shi, X. Pang, Y. Mu, H. He, S. Shuai, J. Wang, et al. Emission reduction 1243

potential of using ethanol–biodiesel–diesel fuel blend on a heavy-duty diesel engine. 1244

Atmospheric Environment. 40 (2006) 2567-74. 1245

Page 58: Elsevier Editorial System(tm) for Energy Conversion and

54

[102] L. Xiaoming, G. Yunshan, W. Sijin, H. Xiukun. An experimental investigation on 1246

combustion and emissions characteristics of turbocharged DI engines fueled with blends 1247

of biodiesel. SAE Technical Paper 2005-01-2199 (2005). 1248

[103] R.L. McCormick, C. Tennant, R.R. Hayes, S. Black, J. Ireland, T. McDaniel, et al. 1249

Regulated emissions from biodiesel tested in heavy-duty engines meeting 2004 emission 1250

standards. SAE Technical Paper 2005-01-2200 (2005). 1251

[104] E. Verhaeven, L. Pelkmans, L. Govaerts, R. Lamers, F. Theunissen. Results of 1252

demonstration and evaluation projects of biodiesel from rapeseed and used frying oil on 1253

light and heavy duty vehicles. SAE Technical Paper 2005-01-2201 (2005). 1254

[105] A.K. Agarwal, L. Das. Biodiesel development and characterization for use as a fuel 1255

in compression ignition engines. Journal of engineering for gas turbines and power. 123 1256

(2001) 440-7. 1257

[106] D. Qi, H. Chen, L. Geng, Y. Bian. Experimental studies on the combustion 1258

characteristics and performance of a direct injection engine fueled with biodiesel/diesel 1259

blends. Energy Conversion and Management. 51 (2010) 2985-92. 1260

[107] Y.-C. Lin, W.-J. Lee, T.-S. Wu, C.-T. Wang. Comparison of PAH and regulated 1261

harmful matter emissions from biodiesel blends and paraffinic fuel blends on engine 1262

accumulated mileage test. Fuel. 85 (2006) 2516-23. 1263

[108] C. Robbins, S.K. Hoekman, E. Ceniceros, M. Natarajan. Effects of biodiesel fuels 1264

upon criteria emissions. SAE Technical Paper 2011-01-1943 (2011). 1265

[109] I. Saanum, M. Bysveen, J.E. Hustad. Study of particulate matter-, NOx-and 1266

hydrocarbon emissions from a diesel engine fueled with diesel oil and biodiesel with 1267

fumigation of hydrogen, methane and propane. SAE Technical Paper 2008-01-1809 1268

(2008). 1269

Page 59: Elsevier Editorial System(tm) for Energy Conversion and

55

[110] A.K. Agarwal, V. Chaudhury, P.C. Shukla. Macroscopic Spray Parameters of 1270

Karanja Oil and Blends: A Comparative Study. SAE Technical Paper 2012-28-0028 1271

(2012). 1272

[111] J.G. Suryawanshi, N. Deshpande. Effect of injection timing retard on emissions 1273

and performance of a Pongamia oil methyl ester fuelled CI engine. SAE Technical Paper 1274

2005-01-3677 (2005). 1275

[112] D.D. Corgard, R.D. Reitz. Effects of Alternative Fuels and Intake Port Geometry 1276

on HSDI Diesel Engine Performance and Emissions. Development. 1 (2001) 0652. 1277

[113] R. Reitz. Controlling DI diesel engine emissions using multiple injections and 1278

EGR. Combustion science and technology. 138 (1998) 257-78. 1279

[114] A. Uludogan, J. Xin, R. Reitz. Exploring the use of multiple injectors and split 1280

injection to reduce DI diesel engine emissions. SAE Technical Paper 962058 (1996). 1281

[115] Z. Han, A. Uludogan, G.J. Hampson, R.D. Reitz. Mechanism of soot and NOx 1282

emission reduction using multiple-injection in a diesel engine. SAE Technical Paper 1283

960633 (1996). 1284

[116] J.E. Dec, C. Espey. Ignition and early soot formation in a DI diesel engine using 1285

multiple 2-D imaging diagnostics. Presented at the Society of Automotive Engineers 1286

International Congress and Exposition, Detroit, MI, 27 Feb-2 Mar 19951995. 1287

[117] K. Yamane, A. Ueta, Y. Shimamoto. Influence of physical and chemical properties 1288

of biodiesel fuels on injection, combustin and exhaust emission characteristics in a direct 1289

injection compression ignition engine. International Journal of Engine Research. 2 1290

(2001) 249-61. 1291

[118] P. Ye, A.L. Boehman. An investigation of the impact of injection strategy and 1292

biodiesel on engine NOx and particulate matter emissions with a common-rail 1293

turbocharged DI diesel engine. Fuel. 97 (2012) 476-88. 1294

Page 60: Elsevier Editorial System(tm) for Energy Conversion and

56

[119] K. Yehliu, A.L. Boehman, O. Armas. Emissions from different alternative diesel 1295

fuels operating with single and split fuel injection. Fuel. 89 (2010) 423-37. 1296

[120] B. Kegl. Effects of biodiesel on emissions of a bus diesel engine. Bioresource 1297

technology. 99 (2008) 863-73. 1298

[121] S.H. Yoon, J.W. Hwang, C.S. Lee. Effect of injection strategy on the combustion 1299

and exhaust emission characteristics of a biodiesel-ethanol blend in a di diesel engine. 1300

Journal of engineering for gas turbines and power. 132 (2010). 1301

[122] X. Lu, J. Ma, L. Ji, Z. Huang. Simultaneous reduction of NOx emission and smoke 1302

opacity of biodiesel-fueled engines by port injection of ethanol. Fuel. 87 (2008) 1289-96. 1303

[123] P. Kohoutek, N. Metz, K. Richter, A. Volkswagen, M. AG. Particulate matter 1304

emissions from modern diesel engines, impact of fuel quality and PM-development from 1305

road transport in Germany. ILSI-Symposium vom1999. 1306

[124] D. Dwivedi, A.K. Agarwal, M. Sharma. Particulate emission characterization of a 1307

biodiesel vs diesel-fuelled compression ignition transport engine: A comparative study. 1308

Atmospheric Environment. 40 (2006) 5586-95. 1309

[125] H. Kim, B. Choi. The effect of biodiesel and bioethanol blended diesel fuel on 1310

nanoparticles and exhaust emissions from CRDI diesel engine. Renewable Energy. 35 1311

(2010) 157-63. 1312

[126] C.D. Rakopoulos, A.M. Dimaratos, E.G. Giakoumis, D.C. Rakopoulos. 1313

Investigating the emissions during acceleration of a turbocharged diesel engine 1314

operating with bio-diesel or n-butanol diesel fuel blends. Energy. 35 (2010) 5173-84. 1315

[127] M. Graboski, R. McCormick, T. Alleman, A. Herring. The effect of biodiesel 1316

composition on engine emissions from a DDC series 60 diesel engine. National 1317

Renewable Energy Laboratory (Report No: NREL/SR-510-31461). (2003). 1318

[128] D. Kawano, H. Ishii, Y. Goto, A. Noda. Application of biodiesel fuel to modern 1319

diesel engine. Training. 2012 (2006) 09-5. 1320

Page 61: Elsevier Editorial System(tm) for Energy Conversion and

57

[129] C. Raahede. Particle characteristics - PAH and Gaseous emission from light duty 1321

CI engines fueled with biodiesel and biodiesel blends. SAE Technical Paper 2010-01-1322

1277 (2010). 1323

[130] A. Puzun, S. Wanchen, L. Guoliang, T. Manzhi, L. Chunjie, C. Shibao. 1324

Characteristics of particle size distributions about emissions in a common-rail diesel 1325

engine with biodiesel blends. Procedia Environmental Sciences. 11 (2011) 1371-8. 1326

[131] J. Desantes, V. Bermúdez, J. García, E. Fuentes. Effects of current engine 1327

strategies on the exhaust aerosol particle size distribution from a heavy-duty diesel 1328

engine. Journal of Aerosol Science. 36 (2005) 1251-76. 1329

[132] A. Sinha, J. Chandrasekharan, J. Nargunde, W. Bryzik, K. Acharya, N. Henein. 1330

Effect of biodiesel and its blends on particulate emissions from HSDI diesel engine. SAE 1331

Technical Paper 2010-01-0798 (2010). 1332

[133] P. Pham, T. Bodisco, S. Stevanovic, M. Rahman, H. Wang, Z. Ristovski, et al. 1333

Engine performance characteristics for biodiesels of different degrees of saturation and 1334

carbon chain lengths. SAE International Journal of Fuels and Lubricants. 6 (2013) 188-1335

98. 1336

[134] Y. Di, C. Cheung, Z. Huang. Comparison of the effect of biodiesel-diesel and 1337

ethanol-diesel on the particulate emissions of a direct injection diesel engine. Aerosol 1338

Science and Technology. 43 (2009) 455-65. 1339

[135] M. Lapuerta, O. Armas, J. Rodríguez-Fernández. Effect of the degree of 1340

unsaturation of biodiesel fuels on NOx and particulate emissions. SAE International 1341

Journal of Fuels and Lubricants. 1 (2009) 1150-8. 1342

[136] J. Nuszkowski, K. Flaim, G. Thompson. The effect of cetane improvers and 1343

biodiesel on diesel particulate matter size. SAE International Journal of Fuels and 1344

Lubricants. 4 (2011) 23-33. 1345