You are on page 1of 13

Renewable and Sustainable Energy Reviews 48 (2015) 635647

Contents lists available at ScienceDirect

Renewable and Sustainable Energy Reviews


journal homepage: www.elsevier.com/locate/rser

Recent studies on soot modeling for diesel combustion


Hamid Omidvarborna a, Ashok Kumar a, Dong-Shik Kim b,n
a
b

Department of Civil Engineering, The University of Toledo, Toledo, OH, USA


Department of Chemical and Environmental Engineering, The University of Toledo, Toledo, OH, USA

art ic l e i nf o

a b s t r a c t

Article history:
Received 20 August 2014
Received in revised form
15 March 2015
Accepted 3 April 2015
Available online 28 April 2015

This paper analyzes published works on the emission models of diesel and BD1 fuels. To the best our
knowledge, this is the rst comprehensive survey that reviews various modeling aspects of soot emitted
from the combustion of diesel and BD fuels. The pros and cons of past and recent soot models, the
chronological advancement of diesel combustion chemistry, and soot modeling approaches are highlighted in this review. Soot models are divided into three main groups of empirical, semi-empirical, and
detailed soot model. Phenomenological model is also explored as a soot model which is one of the most
extensively investigated soot models in recent years. Soot formation mechanism is discussed with an
emphasis on their molecular structure. In a vast majority of the papers reviewed, acetylene was used as a
soot precursor, and also as a reactant for soot mass growth and aromatics formation in diesel soot
modeling studies. Thus, it is recommended that the formation and consumption of acetylene and
aromatic compounds should be included in the diesel soot modeling. For BD, aromatic compounds are
found at very low concentrations during the combustion, so the contribution of aromatic compounds to
soot formation may be reduced or excluded in BD soot modeling. Unlike diesel, oxygen in BD fuels is
found very important in soot oxidation, thus, formation and consumption of oxygen molecules, radicals
and OH2 should be incorporated in the soot modeling as well. Finally, regardless of their structures,
simple molecules such as MB3 and MD4 are found practical as BD surrogates in many modeling papers.
Published by Elsevier Ltd.

Keywords:
Soot formation
Particulate matter
Combustion modeling
Precursor formation
Oxygenated fuel

Contents
1.
2.
3.

4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Soot composition and structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Soot formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
Precursors for soot formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.
Nucleation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.
Mass growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.
Coagulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5.
Oxidation process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.
Empirical and semi-empirical soot models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.
Recent studies on soot modeling with emphasis on phenomenological studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3.
Soot formation mechanism from oxygenated fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Corresponding author. Tel.: 1 419 530 8084; fax: 1 419 530 8086.
E-mail address: dong.kim@utoledo.edu (D.-S. Kim).
1
Biodiesel.
2
Hydroxide bonds.
3
Methyl butanoate.
4
Methyl decanoate.

http://dx.doi.org/10.1016/j.rser.2015.04.019
1364-0321/Published by Elsevier Ltd.

636
636
637
637
638
638
638
638
638
639
640
643

636

H. Omidvarborna et al. / Renewable and Sustainable Energy Reviews 48 (2015) 635647

5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 644
Acknowledgement. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 645
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 645

1. Introduction

2. Soot composition and structure

The call for emission reduction has been mandated by many


governments. For past few years, it has been one of the top priorities
for combustion research centers to investigate combustion processes
and emission reduction methods through optimizing the engines and
fuels [13]. Soot modeling is regarded as an important part of
understanding the process of soot formation, which in turn contributes to the development of effective emissions reduction techniques.
Diesel is known as a source of emission species such as
particulate matter (PM), polycyclic aromatic hydrocarbons (PAHs),
heavy metals, and nitrogen oxide (NOx) [4,5]. These species are
produced from combustion and found in emissions mainly in the
form of aerosols, and are recognized as health hazards [6,7].
Among these diesel emission components, PM5 has been a serious
concern for human health due to its direct and broad impact on
the respiratory organs [4,5]. In earlier times, health professionals
associated PM10 (diametero 10 mm) with chronic lung disease,
lung cancer, inuenza, asthma, and the cause of increased mortality rate [8]. However, recent scientic studies suggest that these
correlations be more closely linked with ne particles (PM2.5) and
ultra-ne particles (PM0.1) [9], because the ne and ultra-ne
particles can easily penetrate deep into the lungs. To address these
problems, a great deal of air quality research has been performed
on toxicity and chemistry of PM over the last 40 years [10,11]. It is
generally reported that the majority of PM is originated from soot,
(highly carbonaceous material which weighs typically higher than
50% in PM mass), which is usually formed in fuel-rich or lowoxygen regions of a diesel engine [1215].
A better understanding of the soot formation made it possible
to formulate mathematical models that predict the concentration
or mass of soot in the emissions, and validate the proposed mechanisms, and in turn good models are helpful in better understanding soot characteristics and formation mechanisms. Signicant advances have been made on the mechanisms of soot
formation in the last two decades [1619]. However, it appears
that there is still a gap between the existing soot models and
actual soot formation processes. The gap becomes even greater
when it comes to soot formation from combustion of oxygenated
fuels (BD fuels) due to the varying compositions and diverse types
of feedstock [2022].
Soot modeling can be improved as long as the formation and
oxidation mechanisms are clearly understood and accordingly
more realistic assumptions are made. This review begins with a
brief description of soot formation fundamentals and summarizes
the progress of soot modeling for diesel combustion since the early
1970s, while giving more emphasis on the modeling work over the
last 20 years. It is followed by different modeling approaches and
comparison of those approaches along with the highlights of
theoretical and empirical results. It also investigates the models'
specications and parameters, and then the accuracy of their
predictions for the performance of the model with regard to the
soot concentration. Not all of the available models in the literature
are considered here, but the review focus is mainly on the practical
soot models for vehicle combustion of diesel developed over the
past two decades.

To better understand the mechanisms of soot formation found


in the literature, it is worthwhile to briey review the chemistry of
the soot. Soot is a solid substance consisting of roughly eight parts
of carbon and one part hydrogen (soot density is 1.84701 g/cm3
[23] and the reports by most other authors fall near this value). In
urban areas soot is mostly formed as a result of fuel combustion in
engines and its characteristics do not appear to be functions of fuel
and other operating conditions [24]. Soot becomes a part of black
carbon/smoke when present in sufciently large particle size and
quantity in exhaust gases. Soot nucleates from the vapor phase to
a solid phase in fuel-rich regions at elevated temperatures [13,14].
HCs or other available molecules may condense on, or be absorbed
by soot depending on the surrounding conditions [25].
A newly formed soot particle initially has the highest hydrogen
content, and the C/H ratio is as low as one. However, as the soot
matures, the carbon fraction increases. Trace amounts of zinc,
phosphorus, calcium, iron, silicon, and chromium are also often
detected in emitted soot from diesel engines [13,14,26,27].
Soot is found to be in the size of sub-microns and in the form of
necklace-like agglomerates [16]. Fig. 1a is a typical scanning
electron microscope (SEM) image of diesel soot showing these
agglomerates are composed of collections of smaller particle units
in spherical or close to spherical shape [17]. X-ray diffraction
(XRD), as illustrated in Fig. 1b [28], indicates that the carbon atoms
of a primary soot particle are packed into hexagonal face-centered
arrays, commonly referred to as platelets. Platelets are arranged in
layers to form crystallites, and there are typically two to ve
platelets per crystallite [28]. When analyzed under high-resolution
transmission electron microscopy (HRTEM), two distinct parts of a
primary diesel soot particle can be identied: an outer shell and an
inner core, as shown in Fig. 1c [29]. The platelet model mentioned
above applies to the outer shell. However, the inner core contains
ne particles with a spherical nucleus surrounded by carbon networks with a bending structure. It shows that the outer shell,
which is composed of graphitic crystallites, is of a rigid structure,
while the inner core is chemically and structurally less stable due
to the thermodynamic instability of its structure. Arrangements of
crystallites which contain inner/outer shells and ne particles in
collected soot (with different sizes observed under HRTEM analysis) are shown in Fig. 1c.
In summary, the formation of soot, i.e. the conversion of HC
fuel molecules into carbonaceous agglomerates, is an extremely complicated process. It is a kind of gaseous-solid phase
transition where the solid phase exhibits no unique physical
and chemical structures, and the transition occurs through
various chemical reaction and physical interaction steps. A
number of approaches to soot modeling exist, but there is a
trade-off between the capability of predicting the details of
soot formation and computational time. Another issue in soot
formation modeling is the complexity of simultaneous chemical and physical phenomena, such as precursor formation
from the gas phase chemistry, primary particle inception,
nucleation, particle growth, coagulation and particle oxidation which are hard to describe in a series of mathematical
formula. So, simplied soot models that can produce more
realistic results in reduced computational time are highly
desired for engine design and emission control.

Particulate matter.

H. Omidvarborna et al. / Renewable and Sustainable Energy Reviews 48 (2015) 635647

3. Soot formation
The formation of soot is a complex process, an evolution of
matter in which a number of molecules undergo many chemical
and physical reactions within a few milliseconds. It is still not

637

clearly understood how soot particles and their precursors are


formed despite the broad and extensive studies published in the
literature [16,17,24,28]. Many details of soot formation chemistry
remain unanswered and controversial, but there have been a few
agreements which are summarized here [18]:
Soot begins with some precursors or building blocks.
Nucleation of heavy molecules occurs to form particles.
Surface growth of a particle proceeds by adsorption of gas
phase molecules.
Coagulation happens via reactive particleparticle collisions.
Oxidation of the molecules and soot particles reduces soot
formation.

Fig. 1. (a) SEM image of soot aggregates in diesel exhaust collected from a Toledo
Area Regional Transit Authority (TARTA) bus (b) Substructure of a soot particle (c1)
Microstructure of diesel soot particles [26] (c2) HRTEM image of collected soot
from a combustion chamber.

Many references proposed various soot formation processes in


which many of them have in common. These widely agreed
mechanisms proceed in three steps and are depicted in Fig. 2.
Large aromatic rings are formed mainly through addition of light
HCs (acetylene) molecules in the molecular scale. Primary soot
particles are supposed to be formed either by surface growth or
coagulation of these larger aromatic compounds.
Tree and Svensson [25] presented ve steps of soot formation
as depicted schematically in Fig. 3. In which, acetylene and PAH
molecules are involved during precursor formation after fuel
pyrolysis. Nucleation, surface growth, coalescence and agglomeration are considered as afterwards steps.
According to the Tree and Svensson [25] soot formation
mechanism, the HC fuel is degraded into small HC separate
radicals. Later on HC radicals are added for growth of unsaturated HCs when they contain a sufciently large number of
carbon atoms in their structure. An increase in the acetylene
concentration acetylene mainly helps the formation of larger
aromatic rings. The growth is supposed to happen by coagulation of larger aromatic structures forming primary soot particles. These primary particles quickly coagulate, simultaneously
picking up molecules from the gas phase for surface growth.
Five common steps in soot formation are briey discussed in
next sub-sections.
3.1. Precursors for soot formation

Fig. 2. A conceptual description of progression of soot formation in three steps [17].

The species that are considered to be the onset for soot


formation and growth are referred to as precursors. Soot inception
is a mechanism through which the precursors are resulted from
fuel combustion to form soot particles. Inception is poorly understood because the nascent soot particles are extremely small
(about 1 nm in diameter) thus making experimental investigations
very difcult [18]. Among them acetylene has received great attention [30]. Acetylene has been identied by Glassman [24] and
later conrmed by Richter and Howard [18] as a very important precursor for soot formation in diesel combustion, most
likely because the rst aromatic rings are formed from C2 and C3
additions [31].
In 1990s, Frenklach and Wang [32] proposed that the addition of acetylenes lead to the formation of rst aromatic rings,
and those aromatic rings are the soot precursors. Due to limited
formation of some intermediate molecules via acetylene and
complexity of experimental studies, PAHs molecules instead of

Fig. 3. Schematic diagram of the soot formation step process from gas phase to solid agglomerated particles in ve steps [25].

638

H. Omidvarborna et al. / Renewable and Sustainable Energy Reviews 48 (2015) 635647

acetylene are much more considered as soot precursor in diesel


combustion [33,34]. PAH formation and its growth appear to
depend mainly on the type of fuel. Some of the reaction sequences which depict the formation of rst aromatic rings are
summarized elsewhere [26].
3.2. Nucleation
The next step, nucleation or inception of particles from heavy
PAH molecules, bridges the transition from gaseous media in a
combustion process to heavy molecules that eventually turn into
nascent soot. The molecular mass of nascent soot is approximately
2000 atomic mass unit (amu) [16] with an effective diameter of
about 1.5 nm (can be detected by HRTEM) [18], while it is
commonly believed that nucleation starts at lower amu around
300700 [35].
3.3. Mass growth
Soot surface growth is the overall mechanism through which
soot particle masses grow via the addition of gas species such as
acetylene and PAH molecules/radicals. There is no clear distinction
between the end of the nucleation and the beginning of surface
growth and in reality the two processes are concurrent. Frenklach
[35,36] introduced the surface growth reaction mechanism back in
1980s. Soot particles undergo surface reactions with gaseous
species via the hydrogen abstraction carbon addition (HACA)
process [37,38]. For HACA growth, the soot surface property is
an important factor in soot mass growth. CH bonds on the surface
of the soot interfere with H and OH radicals to form reactive sites,
where gaseous molecules (particularly acetylene) can be added to
the surface of the soot particle [37,39].
3.4. Coagulation
As depicted in Figs. 2 and 3, during nucleation, particle
growth happens through the coagulation step, i.e., a combination of two or more particles to form a larger particle, sometimes called coalescence [17,25]. The results of experiments
depict that particle coagulation process occurs almost immediately after the soot particle formation, or when soot particles are
relatively small or young [17]. Sticking collisions between
particles during the mass growth process signicantly increase
the particle size and decrease the number of particles without
changing the total mass of soot present. Sometimes individual
or primary particles stick together to form large groups of
primary particles which maintain their shape. In this case the
process is called agglomeration. So, the coagulation process
forms a large particle by combining small particles, where during agglomeration the primary particles stick to each other,
forming a group of chain-like aggregates. An example of agglomeration is easily found in the collection of exhaust soot from
a diesel engine. In soot exhaust, soot consists of primary
particles which are spherical in shape, and they are agglomerated to form long chain-like structures as shown in Fig. 1a.
3.5. Oxidation process
Soot oxidation is the result of the processes that reduce the
mass of soot by converting the solid soot particles or part of them
back into gases (e.g. CO and CO2). Oxidation is similar to the
surface growth in a sense that the surface area of the particles
affects the rate of oxidation. Oxidation takes place on the surfaces
of soot particles and decreases the mass of soot and reduces the
mass of carbon accumulated in the soot particles [17]. Unlike the
surface growth of soot, which occurs in a specic step, oxidation

happens all the time during and after soot formation. Oxidizing
elements are O, O2, and OH under fuel-rich conditions, but in fuellean media H2O, CO2, NO, N2O, and NO2 are also possible oxidants
[19]. More oxidation models in which oxidants other than O2 are
involved are presented elsewhere [19,40].
A comprehensive review for the fundamentals of soot formation mechanism is beyond the scope of this review. More in-depth
reviews were provided by earlier studies [16,17,24,28].

4. Modeling
Soot mechanism is difcult to be mathematically modeled
because of the large number of primary components of diesel
fuel, quite complex combustion mechanisms, and the heterogeneous interactions during soot formation [41].
Soot models are broadly categorized into three subgroups
[42]. Empirical (equations that are adjusted to match experimental soot proles), semi-empirical (combined mathematical
equations and some empirical models which used for particle
number density and soot volume and mass fraction), and
detailed theoretical mechanisms (covers detailed chemical
kinetics and physical models in all phases) are usually available
in literatures for soot models.
Empirical models use correlations of experimental data to
predict trends in soot production [4345]. Empirical models are
easy to implement and provide excellent correlations for a given
set of operating conditions. However, empirical models cannot be
used to investigate the underlying mechanisms of soot production.
So, these models are not exible enough to handle changes in
operating conditions. They are only useful for testing previously
established designed experiments under specic conditions.
Second, semi-empirical models solve rate equations that are
calibrated using experimental data [43,46,47]. Semi-empirical
models reduce computational costs primarily by simplifying the
chemistry in soot formation and oxidation. Semi-empirical models
reduce the size of chemical mechanisms and use simpler molecules, such as acetylene as precursors.
Detailed theoretical models use extensive chemical mechanisms containing hundreds of chemical reactions in order to predict
concentrations of soot. Detailed theoretical soot models contain all
the components present in the soot formation with a high level of
detailed chemical and physical processes.
Such comprehensive models (detailed models) usually take
high nancial burden for programing and operating, and much
computational time to produce a converged solution. On the
other hand, empirical and semi-empirical models ignore some
of the details in order to make complex model simple and to
reduce the computational cost and time. Thanks to recent
technological progress in computation, it becomes more feasible
to use detailed theoretical models and obtain more realistic
results. However, further advancement of comprehensive theoretical models must be preceded by the more detailed and
accurate formation mechanisms.
On the other hand, models that are based on a phenomenological description have found wide use recently. Phenomenological
soot models, which may be categorized as semi-empirical models,
correlate empirically observed phenomena in a way that is consistent with the fundamental theory, but is not directly derived
from the theory. Phenomenological models use sub-models developed to describe the different processes (or phenomena) observed
during the combustion process. These sub-models can be empirically developed from observation or by using basic physical and
chemical relations. Advantages of phenomenological models are
that they are quite reliable and yet not so complicated. So, they are
useful, especially when the accuracy of the model parameters is

H. Omidvarborna et al. / Renewable and Sustainable Energy Reviews 48 (2015) 635647

low. For example, as presented by Argachoy and Pimenta [48], the


phenomenological models can predict the soot formation even
when several operating conditions are changed in a system and
the accuracy cannot be guaranteed. Examples of sub-models of
phonological empirical models could be listed as spray model, liftoff model, heat release model, ignition delay model, etc. [47,48].
4.1. Empirical and semi-empirical soot models
This section presents diesel soot models published mainly on
empirical and semi-empirical approaches. Several models proposed in these categories consider two competing reactions, soot
formation and soot oxidation, as a two-step approach [40,47,49].
Both formation and oxidation rates are highly temperature dependent, and they are represented by Arrhenius type expressions. Soot
formation rates are proportional to fuel vapor pressure while
formation expressions contain no dependence on the type, composition or structure of fuel. Also, the two-step models contain no
information on particle size or agglomeration of soot, both of
which affect the surface area of soot available for a given mass of
soot produced as explained in Section 3. The oxidation expression
includes only O2, leaving out other important oxidation mechanisms such as OH reactions.
A study on soot formation by Tesner et al. in 1971 [46] was one
of the rst soot models in these categories that include a
branched-chain process and soot particle formation. Tesner et
al.'s model implemented an idea that soot is formed as a result of
adsorption of radical nuclei on the precursor surface. In the same
year, an empirical model was proposed by Khan et al. [44,50]. This
model predicts soot emissions from diesel engines based on the
assumption that the rate of soot production is entirely a function
of soot nucleation rate. It means that the rates of particle growth
and oxidation were neglected in this model. Also, the model
assumes that the diameter of a soot particle is not a function of
engine operating conditions at different speeds or loads, which is
regarded as unrealistic. The model includes some modeling parameters determined by comparing the output of the model with
experimental data [42,44].
A few years later, Hiroyasu et al. [49] proposed one of the most
widely used models that are essentially based on empirical
formulas for predicting the formation and oxidation of soot
particles. They found that their two-step soot model was primarily
affected by pressure, temperature, and equivalence ratio, which is
the actual fuel to oxidant ratio normalized by the stoichiometric
fuel to oxidant ratio. The model includes a soot formation and
oxidation rate, which incorporates the available fuel mass and
oxygen partial pressure. Eq. (1) calculates the rst order rate of net
soot formation (dms/dt) using a combination of soot formation rate
and soot oxidation rate for Hiroyasu's model. Then, the soot
formation rate equation (Eq. (2)) follows an Arrhenius-type relation with the mass of vaporized fuel (mfg). Oxidation rate of soot
(dmsc/dt) in Eq. (3) also follow an Arrhenius-type equation as a
function of mass of soot in the system (ms) and oxygen pressure
P o2 . The oxidation rate of soot is almost a second order (1.8) of
pressure whereas the formation rate of soot is a half order of
pressure

639

and soot oxidation, respectively. Af and Ac were determined by


matching the calculated soot and experimental soot in the
exhaust. The Hiroyasu's model has been very helpful in providing
knowledge on the bulk distribution and transport of the soot in
the high-temperature combustion environments of conventional
diesel engines [51]. Moreover, owing to ease of implementation
into computational uid dynamic (CFD) codes, this model and its
modications have acquired wide popularity in the community
engaged in multidimensional diesel combustion simulations [52].
The two-step approach of Hiroyasu's model [49] is regarded
oversimplied for the diesel soot formation processes because
they proposed a two-step empirical soot model for predicting the
formation and oxidation of soot particles and the model underpredicted the peak in-cylinder soot concentration [51]. Soot
formation formula of Hiroyasu's model contains no dependence
on the type, composition or structure of fuel. The oxidation
expression includes only oxygen molecules in the model [53].
Hiroyasu's model is regarded very practical and simple, but it
needs more parameters to be upgraded for further studies [54
56].
Another basic semi-empirical model developed for diesel
engines was proposed by Nishida et al. [57,58], Belardini et al.
[59], and Patterson et al. [51]. In 1993, Gorokhovski et al. [60]
assumed that soot is generated from a stable HC intermediate
species. It means that the soot surface growth rate was determined
by experimental data on a nal soot volume.
The Hiroyasu-Nagle and Strickland (HNS) soot model has been
another very popular two-step semi-empirical model for soot
formation in diesel engines [47,61]. The rate of soot formation
and oxidation were expressed again in an Arrhenius-type equation
as follows:
Rate of formation Af P 0:5 e  Ef =RT

Rate of oxidation X O2 Ao P 1:8 e  Eo =RT

dms dmsf dmsc


dt
dt
dt

where Af and Ao are the pre-exponential factors; Ef and Eo are the


activation energies; X O2 is oxygen molar fraction; R is the ideal gas
constant; and T is the gas temperature. Hiroyasu's approach is
adopted to describe soot formation while its oxidation is estimated
by the Nagle-Strickland and Constable model [54,55].
Moss [62] also presented a semi-empirical two-step model
which is slightly different from the previous models. In this model,
not only nucleation and oxidation rates, but also coagulation and
growth rates, were considered and implemented in the process of
soot formation. The superiority of Moss's model is not surprising
since it simulates the processes of nucleation, surface growth, and
coagulation, whereas other two-step models, such as Khan et al.
[50], rely on a simple kinetic expression for soot nucleation rates.
Lindstedt outlined reaction steps for the formation and growth
of soot particles [63]. Detailed gas phase chemistry and simplied
steps of nucleation, surface growth and particle agglomeration
were incorporated in the model. The soot nucleation and surface
growth reactions are linked to the gas phase chemistry. Lindstedt
paid much attention to the problems in modeling the soot mass
growth. Lindstedt considered four models for the soot growth
reactions [42]. More details on the recent soot models are
provided later in Section 4.2 that addresses detailed mechanisms
and phenomenological soot models as well.

dmsf
Af mf g P 0:5 e  Esf =RT
dt

4.2. Recent studies on soot modeling with emphasis on


phenomenological studies

dmsc
Po
Ac ms 2 P 1:8 e  Esc =RT
dt
P

where msf is the mass of the formed soot, and mfc is the mass of
oxidized soot. Esf and Esc are activation energies of soot formation

Soot formation phenomenon is far from being fully understood


today and models available for simulation of soot in combustion
devices remain of relatively limited success, despite signicant
progresses made over the last decade. Since Hiroyasu et al. [47,64]

640

H. Omidvarborna et al. / Renewable and Sustainable Energy Reviews 48 (2015) 635647

and Khan et al. [50,65] presented ones of the earliest models for
the prediction of soot production from a diesel engine, variety of
soot models with different levels of complexity have been proposed and applied to soot formulation.
As mentioned earlier soot formation and oxidation are very
complex, and it is not possible to exactly model a complex process
only with simplied models. Unlike the empirical and semiempirical models which extremely simplify the soot formation
process, the detailed model describes the formation, growth, and
oxidation of soot with a detailed chemical reaction mechanism.
The extremely high demand of computing time of detailed soot
models makes them unrealistic for simulation of diesel engine
combustion. Hence, most of the investigations conducted in a real
conguration such as multidimensional diesel engines utilize
coarse modeling schemes to take advantage of easy implementation and low computational cost. This section reviews the published papers that focused on the detailed models and those with
the emphasis on phenomenological methods.
With recent advances in computer technology and developments in mathematical sub-models, it is now possible to
obtain useful predictions and visualizations of complex systems. Therefore, numerical simulation of such complex systems
is considered in many areas. In soot modeling, numerical
simulations can also be divided into two main classes: phenomenological modeling and multidimensional CFD modeling.
In phenomenological modeling, the spatial variations are often
simplied by zero-dimensional or one-dimensional models,
while multidimensional modeling is designed to take into
account all the spatial variations of reactive uid ow in diesel
engines simultaneously [26]. In order to improve the accuracy
and predictability, phenomenological multi-step soot models
have been implemented in many empirical, semi-empirical and
detailed models. In other words, phenomenological models
describe the complex process of soot formation and oxidation
in terms of several global steps that are particularly advantageous for practical combustion simulations.
In 1994, Fusco et al. [66] proposed a phenomenological
soot model to overcome some limitations of the previous soot
models for combustion conditions of a diesel engine. The
model accounts for the number of carbon atoms of the major
constituent molecules in the fuel and incorporates the physical process of inception, surface growth, coagulation and
oxidation into the eight-step phenomenological soot model.
They also compared their model with the existing two-step
empirical models and criticized the non-applicability of the
two-step empirical models for a wide range of operation
conditions in diesel engines. The model consists of four
differential equations balancing between the rates of particle
number change, soot precursor radicals, acetylene and soot
volume fraction. Just like the previous formulations, Arrhenius-type rate expression has been used for most of the
Inert products
(3)
(1)
Pyrolysis

Oxidation

Soot precursor
radicals

Inert products
Inception

(5)

Fuel

(2)

(4)

Oxidation

Oxidation

Soot
particles

(6)
Growth species
(acetylene)

(7)

Surface
growth

(8)

Coagulation

Soot
particles

Inert products
Fig. 4. Schematic diagram of eight-step phenomenological soot model presented
by Fusco et al. [66].

Fig. 5. Nine steps in soot modeling presented by Tao et al. [72].

processes except for coagulation and oxidation steps. The


surface growth species, which was assumed to be acetylene,
enhances the mass of the soot particles. Schematic diagram of
the phenomenological model is presented in Figs. 4 and 5.
Fig. 4 shows that the portion of the fuel is converted to soot
precursor radicals (R1) and growth species (C 2 H 2 by R2). Radicals and growth species are considered to be separated species, although they could be the same species at least at the
beginning of the soot formation process. A portion of the
precursor radicals are oxidized (R3) and the rest are converted
to soot particles (R5). The growth species increase the mass of
soot particles by R6. Oxidation is assumed that it does not
affect the particle number density. Growth species (C 2 H 2 ) may
disappear, and the mass of soot particles may decrease via
oxidation (R4 and R7), respectively. The number of soot
particles can decrease due to coagulation with other soot
particles (R8). Their modeling result demonstrated that more
soot is produced when more acetylene is available. This result
means that the nal amount of soot depends on the balance
between soot formation and oxidation in the both solid and
gas phases.
In 1998, a modied version of the phenomenological model
of soot formation by Kazakov and Foster [43] has been implemented into the model developed by Fusco et al. [66]. The model
includes major generic processes involved in soot formation
during combustion; formation of soot precursors, soot particle
nucleation, coagulation, surface growth, and oxidation. After
Kazakov and Foster, Fusco's original model [66] extended by Liu
et al. [67] to produce a nine-step model as presented below:
1.
2.
3.
4.
5.
6.
7.
8.
9.

Acetylene formation from fuel pyrolysis.


Soot precursor formation from acetylene.
Particle inception from soot precursors.
Soot particle coagulation.
Surface growth from acetylene.
Oxidation by O2.
Oxidation by OH.
Acetylene oxidation by O2.
Precursor oxidation by O2.

The phenomenological model covers oxidation of precursor


(acetylene) and fuel by either O2 or OH. Also, the role of acetylene
in inception and surface growth was very important in Liu et al. to

H. Omidvarborna et al. / Renewable and Sustainable Energy Reviews 48 (2015) 635647

develop a nine-step model [67]. The nine-step model had a fundamental weakness which is still unable to express the role of fuel
composition and structure whereas each of the acetylene formation rates is reported to be dependent on fuel structure [25].
Conventional diesel fuel contains around 3035% PAHs, and
numerous studies in recent years have demonstrated that aromatics such as PAHs play a key role in the soot formation process
in diesel engines [68,69]. Idicheria and Pickett [69] focused on the
role of PAH in soot formation in diesel combustion. They came to
the conclusion that PAH chemistry might play an important role
both for accurate prediction of soot mass and distribution. Therefore, the chemical kinetic mechanisms of the mixture of n-heptane
and aromatics (to present as a diesel surrogate) included in recent
modeling studies are considered more accurate to simulate soot
formation than previous models. Since it is not practically possible
to have kinetic reaction mechanisms for all of the hundreds or
even thousands of species present in conventional diesel fuel,
n-heptane is used in many modeling studies of diesel combustion
as a diesel surrogate [30,70,71]. N-heptane has a cetane rating of
56 that is typical of ordinary diesel fuels, and that is why the
combustion process in diesel engines has often been simulated
using n-heptane as a surrogate diesel fuel [71].
In Tao et al.'s [70] model, diesel fuel is assumed to be singlecomponent, and its oxidation chemistry is represented only by the
n-heptane kinetics. The chemical mechanism simplied to a size of
65 species and 273 elementary reactions. Here formation reactions
of PAHs (up to few aromatic rings) from acetylene were considered
as initiator for soot formation in soot modeling.
In 2002, Frenklach assumed the formation and growth of
PAHs as the rst step in soot formation [37]. In this model, the
particles grow via surface reactions similar to the growth reactions for PAHs, i.e., mainly by the HACA mechanism. When
particles collide with each other, either they form new spherical particles or agglomerates. In that paper, the discussion
shifted from phenomenological possibilities to specics of reaction pathways.
In a parallel study with Liu et al. [67], nine-step phenomenological soot model was updated for predicting soot formation and
oxidation processes in diesel engines by Tao et al. [72]. The brand
new model presented by Tao and coworkers consist of two parts,
detailed chemical reaction mechanism and a phenomenological
semi-empirical soot model.
Tao's model retains the main features of his original model
[43,66], but contains three major modications: (1) fuel pyrolysis leads solely to acetylene formation; (2) the soot precursor is formed merely via acetylene (i.e., not directly from
fuel); (3) an OH-related soot oxidation step is added. In the
earlier study of Liu et al. [67], the OH concentrations were
calculated using the concept of chemical equilibrium, the
assumption of which was unfortunately unrealistic when
applied to transient diesel combustion processes. The updated
nine-step soot model [72] was successfully applied to analyze
the soot distribution structure in a conventional diesel for a
benchmark heavy-duty diesel engine (Cummins) based on
which a comparison to the two-step soot model was attempted.
In 2010, Vishwanathana and Reitz [73] presented a practical
model framework only based on four fundamental steps: soot
inception through a four-ring PAH species, surface growth of
acetylene, coagulation of acetylene to form soot, and soot oxidation via oxygen and OH. They concluded that the soot model is
fairly sensitive to the PAH chemical mechanism [74].
Simultaneously, Cheng et al. [75] presented an improved
detailed soot model for the numerical investigation of soot
formation, mass concentration, and size distribution in diesel
engines. The effects of soot precursors, including isomers of

641

acetylene and PAHs, and the physical processes of PAH deposition


on the particle surface, soot formation, and particle surface growth
were considered into the model. They found that large amounts of
small-size soot particles (in the range of 540 nm) were produced
at the initial stage of combustion by the pyrolysis reactions and
polymerization of the HC fuel. In the intermediate stage of
combustion, soot particles continued to grow by particle coagulation, surface growth, and the deposition of PAHs. In the nal stage
of combustion, the particle size distribution stabilized in the range
of 5 to 20 nm due to the inuence of further oxidation reactions.
Jia et al. [76] quantitatively validated and improved the
phenomenological soot model over wide operating conditions
of homogeneous charge compression ignition (HCCI) combustion. The phenomenological model developed in this research
as summarized in Table 1 is based on the work of Tao et al.
[77]. By comparing experimental results, necessary improvements have been made to the model for describing the soot
formation process under various conditions. The complex
processes of soot formation and oxidation are divided into
several steps including acetylene (C 2 H 2 ) formation from pyrolytic decomposition of fuel, precursor formation via C 2 H 2
conversion, particle inception from precursor, particle surface
growth by C 2 H 2 , particle coagulation, particle surface oxidation via oxygen (O 2 ) and OH. Later, a six-step phenomenological soot model with particle dynamics was developed by
Pang et al. [78]. The sub-model for soot formation was constructed based on Jia's soot model [104] by introducing
necessary improvements and optimizations. The schematic
representation in Fig. 6 shows the structure of the six-step
phenomenological soot model developed in their study. Soot
formation and oxidation process are divided into several steps
including soot precursor formation via C 2 H 2 , A3 (aromatic
structure with 3 rings) and A4 conversion, particle inception
from soot precursor, particle surface growth by C 2 H 2 and A1,
particle coagulation, particle surface oxidation via O 2 and OH,
and precursor oxidation. The new model retains the main
features of the original one [76] but two major modications
are as follows:
1. PAHs (A3, A4) are used as precursor species.
2. Particle surface growth by A1 is added in the new soot model.

In Fig. 6, carbon atoms for soot precursor and soot particle


are represented by C(PR) and C(S), respectively. The rates of
reactions, including soot precursor formation (RS(1), RS(2), RS
(3)), particle inception (RS(4)), particle surface growth (RS(5),
RS(6)), and soot precursor oxidation (RS(10)) were assumed to
be in the form of Arrhenius equation. In this approach, this
phenomenological soot model has gained signicant improvements in performance by incorporating the PAH chemistry into
the model. A modied skeletal PAH mechanism for the phenomenological soot formation was integrated into a primary
reference fuel (PRF) oxidation mechanism where A3 and A4
were the soot precursor species. The new skeletal PAH mechanism is capable of describing the formation process of PAHs
beyond A1 and up to A4.
As pointed out by Vishwanathan and Reitz [73] and Ra and
Reitz [79], the PRF mechanism developed earlier by Ra and Reitz
[80] under-predicts the concentration of C2H2, which is considered
as an important species in soot inception and growth processes. If
a PAHs mechanism is applied, C2H2 is also an important species
that affects the PAH growth process through the well-known
HACA growth mechanism [81]. Table 1 summarizes the recent
modeling approaches and their results that are presented in
this paper.

642

H. Omidvarborna et al. / Renewable and Sustainable Energy Reviews 48 (2015) 635647

Table 1
Concise description of works done on soot modeling mainly for diesel engines.
Author name
(year)
Belardini et al.
[41]
Frenklach et al.
[82,83]

Model specication

N-heptane is selected as a diesel surrogate.


Assuming acetylene as the crucial pyrolitic species.
The model combines recent developments in gas-phase reactions, aromatic
chemistry, soot particle coagulation, soot particle aggregation, and develops
a new sub-model for soot surface growth.
Daly and Nag
A new gas phase kinetic model using Westbrook's gas phase n-heptane
[84]
model and Frenklach's soot model [82,83].
614 Species and 2883 reactions are involved in the complex reaction
mechanism.
Tao et al. [70]
Diesel fuel is assumed to be single-component, and its oxidation chemistry
is represented by the n-heptane kinetics.
The chemical mechanism reduced to a size of 65 species and 273
elementary reactions.
Kong et al. [30] A reaction mechanism of n-heptane is coupled with a reduced NOx
mechanism to simulate diesel fuel oxidation and NOx formation.
The soot emission process is simulated by a phenomenological soot
model that uses a competing formation and oxidation rate formulation.
Acetylene is selected as a soot precursor.
Boulanger et al. A phenomenological three-equation soot model in diesel engine
[85]
combustion.

Results/notes

The model over-predicts both acetylene and soot volume fraction.


Better results can be obtained by model's constant optimization.
The surface growth and oxidation of soot particles are described
consistently with the kinetics of gaseous PAHs.
It represents the state of the art in detailed soot modeling for diesel
combustion.

Molecular precursors of soot produced during the rich burning of the


sprays contribute to soot formation.

Both experiments and models reveal that soot emissions peak when the
start of injection (SOI) occurs.

Some distinct features of this new soot model are:


 No soot is formed at low temperature.
 Minimal model parameter adjustment for application to different fuels.
 There is no need to prescribe the soot particle size.

A reduced n-heptane chemistry mechanism has been extended to include Soot formation and growth regions are not adequately represented by
PAH species up to four fused aromatic rings (pyrene).
using acetylene alone as the soot inception species.
Various soot inception species have been tested.
A simpler model that only considers up to phenanthrene (A3) as the soot
inception species has good possibilities for better soot location
predictions.
Tao et al. [72]
Nine-step phenomenological soot model.
Nine-step model is not only computationally efcient but also
Model includes a detailed chemical reaction mechanism and a
fundamentally sound.
phenomenological semi-empirical soot model.
After calibration, the model is suitable to be integrated with genetic
algorithms for system optimization over a controllable range of
operations.
Mosbach et al.
A detailed model for the formation of soot in internal combustion engines A detailed chemical kinetic mechanism describing the combustion of
PRFs is extended to include small PAHs such as pyrene, which function as
[86]
describing not only bulk quantities such as soot mass, number density,
soot precursor species for particle inception in the soot model.
volume fraction, and surface area but also the morphology and chemical
composition of soot aggregates.
Even with a detailed chemical mechanism, soot formation and oxidation
Jia et al. [76])
An improved phenomenological soot model coupled with a reduced
n-heptane chemical to describe soot formation and oxidation processes in still remain as challenges.
HCCI combustion.
The phenomenological soot model coupled with reduced fuel chemical
mechanism showed satisfactory agreement with the experiments.
Vishwanathana Reduced n-heptane and PAH chemistry mechanisms are formulated from The model is based on four fundamental steps: soot inception through a
the literature.
four-ring PAH species, surface growth through acetylene, soot
and Reitz
Acetylene was selected as a soot precursor.
coagulation, and oxygen- and OH-induced soot oxidation.
[73,74]
Sukumaran
A multistep soot model coupled with reaction mechanisms for fuel
Soot emissions from the engine are highly sensitive to local temperature
et al. [87]
oxidation and PAH formation.
and chemical compositions.
N-heptane mechanism is combined with a detailed PAH mechanism, by
choosing pyrene as precursor.
The overall reaction mechanism consists of 68 species and 145 reactions
and is used with a multistep soot model.
Pang et al. [78] 12 Species and 26 reactions for the formation of PAH are integrated into a The results prove a very good agreement with experimental data and the
skeletal mechanism for the oxidation of PRF (n-heptane and iso-octane).
necessity of including PAHs chemistry for soot modeling.
Six-step phenomenological soot model with PAHs (A3 and A4) as a
precursor.
Particle surface growth by A1 is added in the new soot model.
Naik et al. [88] N-hexadecane, heptamethylnonane, 1-methylnaphthalene, and decalin
A new pseudo-gas soot model coupled with the fuel chemistry to
are used to represent standard European diesel.
simulate an in-cylinder soot nucleation, growth, and oxidation processes.
A validated detailed surrogate mechanism containing 392 species and
2579 reactions was employed to model the chemistry of fuel combustion
and emissions.
Analyses are conducted under low-temperature combustion (LTC)
condition.
Cheng et al. [75] An improved soot model coupled with a detailed mechanism of reduced
The particle emissions increase with increasing engine load.
diesel surrogate fuel (n-heptane / toluene).
Chemical kinetic mechanism contains 70 species and 313 reactions.
Particle concentration and average particle size signicantly increase at
Isomers of acetylene and PAHs are selected as precursors.
the starting stage of the combustion process and quickly stabilize.
Vishwanathan
and Reitz
[34]

H. Omidvarborna et al. / Renewable and Sustainable Energy Reviews 48 (2015) 635647

643

Fig. 6. Schematic representation of the improved phenomenological soot model [78].

4.3. Soot formation mechanism from oxygenated fuels


Investigations of alternative fuels for internal combustion
engines have recently become important due to the growing
concerns about the future availability of oil reserves and
environmental sustainability. Among them, biofuel is receiving
increasing public and scientic attention especially from the
transportation sector driven by its carbon neutrality and comparability with existing engines [13,14,20,27,89,90].
Several investigators have concluded that the structure of
oxygenated fuel has an effect on the amount of soot reduction
possibly achieved with a given amount of the oxygen atoms
included in the fuel structure [13,91,92,93]. The oxygenated
fuels such as BD usually consist of fatty acid methyl esters
(FAMEs)/fatty acid ethyl esters (FAEEs), and they should be
considered for their reactions in soot modeling. FAMEs and
FAEEs are produced through the transesterication process of
vegetal oils or animal fat with methanol or ethanol as a catalyst.
FAME has lower energy content than diesel due to its high
oxygen/low carbon contents, and as a result, its combustion and
fuel consumption can be affected accordingly. The main fatty
acids in rapeseed and soybean oils are oleic (C18:1 monounsaturated) and linoleic (C18:2 polyunsaturated) acids [94].
Compared to regular diesel, the oxygenated structure of BD
enhances oxidation process in soot formation and dramatically
reduces the mass of soot. On the other hand, kinetic studies of BD
surrogates showed that early CO2 production from the methyl
ester (ME) group in methyl decanoate (MD) has important impacts
on ignition and soot production, because if the oxygen in the fuel
immediately produces CO2, it becomes less effective in reducing
soot production [95]. Therefore, it is necessary to develop well
validated models for the combustion and the oxidation of the
oxygenated components of BD to account for the effect of extra
oxygen on soot formation. Among the emission studies conducted
on oxygenated fuels, Miyamoto et al. [91] concluded that soot
reduction was related to the oxygen content of the fuel and not to
the type of fuel. Also, the study of Song et al. [93] on soot
emissions from oxygenated fuels showed that the operating
conditions of a diesel engine would change the morphology of
soot and it would produce smaller particles in idle mode while
using oxygenated fuel.
Not many papers have been published on soot formation
modeling for oxygenated fuels such as BD compared to regular
diesel modeling [9698]. Although the effect of oxygen content
on soot formation and oxidation process are considered highly

107
108
109
105
105
91
110
110

Fig. 7. Reduction of PM, smoke, or integrated jet-soot as a function of oxygen


weight percent in the fuel from numerous experiments reported in the literature
[25,91,105,107110].

important in soot generation from BD combustion, it is also


expected that the widely differing physical properties of BD
from regular diesel will inuence the combustion mechanism
and the related emissions formation. But, a basic understanding
from the kinetic modeling of oxygenated molecules presented in
these studies may provide clues for soot reduction processes in
BD combustion.
BD surrogates are mixtures of one or more simple fuel
components that are designated to emulate physical and chemical properties of BD. While surrogate mixtures can demonstrate more than one characteristic of the fuel to be simulated,
more often than not many other components are required to
emulate the wide variety of properties that are of interest to
researchers. As mentioned earlier n-heptane is normally used
for the diesel surrogate while usually M9D (methyl 9 decenoate),
n-butanol, and MB (even if the small size of these molecules
prevents them from the combustion chemistry of the large
molecules present in BD) are extensively used for BD [96,99].
As mentioned above, unsaturated esters are the most abundant
esters in BD, but very few studies have been dedicated to
combustion modeling for unsaturated esters [100].
As stated above, the oxygen content is an important factor to be
investigated in BD study because the oxygen content in fuel not
only provides more oxygen to burn carbon, but also displaces and
reduces the amount of carbon that needs to be burned. In addition,

644

H. Omidvarborna et al. / Renewable and Sustainable Energy Reviews 48 (2015) 635647

considering the fact that PAH is a main cause for soot formation,
BD has an advantage of low PAH formation as BD was observed to
emit lower PAH emissions than normal diesel [101]. It was
reported that potential for soot precursor formation disappears
almost completely at an oxygen-fuel ratio of 25 wt% [102], and
drops to an insignicant level with extra 3040 wt% oxygen in a
fuel [103]. It means that the BD's molecular structure and its
oxygen content should be the main factors for soot precursor
formation. A summary listing of references and fuels including BD
is presented by Tree and Svensson [25] (Fig. 7), which is sufcient
to draw several conclusions on the effects of oxygen contents on
soot formation. It should be recognized that the results are
gathered from the in-cylinder soot data while others are obtained
from the exhaust emission data. Some of the ndings are as
follows:
1) Soot emissions are reduced by increasing fuel oxygen contents,
based on Smith's study [104]. The reason would be because of
oxidization of acetylene to relatively inert products in the
presence of enough O2 and OH.
2) More than one study has demonstrated a complete or near
complete elimination of soot when fuel oxygen content reaches
30% or more.
3) Scattered data points indicate that mass fraction of oxygenated
fuel can have various amounts of percent PM reduction.
4) Miyamoto et al. [91,105] and Nabi et al. [106] have reported a
nearly linear decrease in soot concentration as fuel oxygen
contents increase. The linear relationship is very interesting
because the required oxygen content to burn the fuel should be
related to molar quantities for oxygen and carbon, not the mass
of them.
According to Curran et al. [103], n-heptane was used as a
representative surrogate of diesel fuel, and methanol, ethanol,
dimethyl ether, dimethoxymethane and MB were used as oxygenated fuel additives to simulate the oxygenated contents. It was
found that when the overall oxygen content in the fuel reached
approximately 3040% by mass, production of soot precursors fell
apparently to zero. Later on Mueller and coworkers [111] explored
characteristics of soot and soot-precursor formation from oxygenated fuels (di-butyl maleate and tri-propylene glycol methyl
ether). They dened four goals for their study:
1. To introduce the oxygen ratio for accurate quantication of
reactant-mixture stoichiometry for both oxygenated and nonoxygenated fuels.
2. To provide experimental results demonstrating that some oxygenates are more effective at reducing diesel soot than others.
3. To present results of numerical simulations showing that
detailed chemical-kinetic models without complex uid
mechanics can capture some of the observed trends in the
sooting tendencies of different oxygenated fuels.
4. To provide further insight into the underlying mechanisms by
which oxygenate structure and in-cylinder processes can affect
soot formation in diesel engines.
In 2013, a simplied chemical reaction mechanism was developed for modeling the combustion process and soot emissions
for both non-oxygenated and oxygenated HC fuels by Wang
et al. [96]. The nal mechanism consists of 76 species and 349
reactions [96]. They reported that soot emission can be greatly
reduced by addition of n-butanol. By blending n-butanol into a
non-oxygenated HC fuel, air entrainment is enhanced by reducing the overall fuel to air ratio by introducing extra available
oxygen atoms through the n-butanol molecule. The predicted

soot emissions under various conditions agree quite well with


the experimental results.
Inspection of the published results leads to the conclusions as
follows [25]:

 There is a reduction in soot emissions with increased fuel


oxygen in all cases.

 More than one investigator has demonstrated a complete or








near complete elimination of soot when the fuel oxygen


content reaches 2735%.
The scattering of experimental data veries that a given fuel
oxygen mass fraction can have various levels of soot reduction.
Some authors [91,105,106,107,110] had observed a nearly linear
decrease in soot concentration as oxygen contents increase
while others [92,108] found decreasing slopes with decreasing
benets for soot reduction when oxygen contents increased.
PAHs are known as soot precursors in diesel fuels, whereas they
are detected at very low-concentrations in BD combustion
[19,26,106,107].
BD has an advantage of low soot formation as BD was observed
to emit lower PAH emissions than diesel fuel. Oxygen content
plays an important role in lowering emissions of PAHs and it
makes BD reactions more complex than diesel fuels. Therefore,
including the detailed precursor formation and oxidation reaction in the BD combustion helps better model the soot formation mechanism.

The high oxygen content of BD makes extra oxygen available to


facilitate the combustion of fuel, especially in the areas of very rich
in fuel. It can have a favorable effect on less occurrence of pyrolysis
and more enhancement of soot oxidation in comparison with
regular diesel combustion. In addition, a wide diversity of feedstock selections for BD necessitates the inclusion of the effect of
basic components of BD on combustion in the soot formation
mechanisms and modeling. Because the major fuel components of
BD from soybean oil, for example, are much different from the BD
from tallow oil or waste cooking oil, and these feedstock-specic
effect should be included in the soot modeling of BD.

5. Conclusion
Fundamental concepts and models about soot mechanism in
diesel and BD emission from combustion are examined in this
review paper. Fuel combustion process is very complex, and their
detailed mechanisms on soot are not quite well understood. From
the literature review, it can be observed that the emission of soot
from the BD-fueled engine is less than ULSD. However, more
experimental and theoretical studies are needed to describe the
complicated process of soot mechanisms in BD combustion.
Soot formation and oxidation steps were incorporated in
early studies on soot mechanism. Later on, due to emerging
new computational systems along with modern experimental
and analytical tools, common steps in soot mechanism were
identied. Following these steps results in more accurate
results compared to experimental data as described earlier.
Among the proposed models for different combustion systems, empirical and semi-empirical soot models are found
relatively simple and practical for specic systems where
experimental results have been implemented into the models.
In recent years, phenomenological models have been found to
be more effective tool for simple and easy prediction of soot
mechanism, but case by case adjustment of the implemented
parameters may be needed. Detailed models were introduced
as the most accurate and comprehensive models. These
models require a great deal of computational time and cost.

H. Omidvarborna et al. / Renewable and Sustainable Energy Reviews 48 (2015) 635647

It is concluded that modeling of acetylene and its isomers as a


starting component for formation of soot precursors seems to
be a reasonable approach to the diesel soot modeling. The size
of the model is important in determining the amount of
computational time in detailed models, because the rate of
soot precursor formation for each fuel will be dependent on
fuel structure.
Regarding soot formation in BD combustion, molecules such
as MB, MD, n-heptane, and MS were identied as common fuel
surrogates in modeling. The degrees of oxygenation and
saturation (i.e., the number of double bonds) of BD fuels appear
to be the important factors to be included in BD soot modeling.
This is because of the number and position of double bonds
which may effect the reaction pathways and mechanisms. For
the BD soot modeling, due to lower emission of PAHs, the
oxidation of precursors such as PAHs may be excluded. On the
other hand the effect of early CO 2 production in BD combustion
on soot formation should be considered.
Modeling of soot formation has to address all these aspects of
regular and BD combustions. In order to develop more robust
and reliable models for soot mechanism, it is recommended that
more reasonable assumptions be made based on a better understanding of chemical and physical interactions in soot mechanism. A fundamental challenge in soot modeling for BD is the
inability to predict differences in soot formation for different
feedstock types and their blends with regular diesel. Further
research needs to be carried out to understand the relationship
between the type of BD feedstock and performance and
emission.

Acknowledgement
The authors express their gratitude to the United States
Department of Transportation (USDOT) under Grant number
DTRT12-G-UTC21 and Mineta National Transit Research Consortium (MNTRC) for funding the BD study. The views expressed in this paper are those of the authors alone and do not
represent the views of the funding organizations.
References
[1] Xue J, Grift TE, Hansen AC. Effect of biodiesel on engine performances and
emissions. Renew Sustain Energy Rev 2011;15:1098116.
[2] Enweremadu CC, Rutto HL. Combustion, emission and engine performance
characteristics of used cooking oil biodiesela review. Renew Sustain Energy
Rev 2010;14:286373.
[3] Shahabuddin M, Liaquat AM, Masjuki HH, Kalam MA, Mojur M. Ignition
delay, combustion and emission characteristics of diesel engine fueled with
biodiesel. Renew Sustain Energy Rev 2013;21:62332.
[4] Krzyzanowski M, Kuna-Dibbert B, Schneider J. Health effects of transportrelated air pollution. Copenhagen, Denmark: World Health Organization,
WHO Regional Ofce for Europe, Schergsvej 8; 2005.
[5] Kunzli N, Kaiser R, Medina S, Studnicka M, Chanel O, Filliger P, et al. Publichealth impact of outdoor and trafc-related air pollution: a European
assessment. Lancet 2000;356:795801.
[6] Kaden DA, Hites RA, Thilly WG. Mutagenicity of soot and associated
polycyclic aromatic hydrocarbons to Salmonella typhimurium. Cancer Res
1979;39:41529.
[7] Durant JL, Busby WF, Laeur AL, Penman BW, Crespi CL. Human cell
mutagenicity of oxygenated, nitrated and un-substituted polycyclic aromatic
hydrocarbons associated with urban aerosols. Mutat Res 1996;371:12357.
[8] Stber W, Abel UR. Lung cancer due to diesel soot particles in ambient air?
Int Arch Occup Environ Health 1996;68:S361.
[9] Schwartz J. Air pollution and daily mortality: a review and meta analysis.
Environ Res 1994;64:3652.
[10] Mauderly JL. Toxicological and epidemiological evidence for health risks
from inhaled engine emissions. Environ Health Perspect 1994;102:16571.
[11] Salvi S, Blomberg A, Rudell B, Kelly F, Sandstrom T, Holgate ST. Acute
inammatory responses in the airways and peripheral blood after shortterm exposure to diesel exhaust in healthy human volunteers. Am J Respir
Crit Care Med 1999;159:7029.

645

[12] Kumar A, Kim DS, Omidvarborn H, Kuppili SK. Combustion chemistry of


biodiesel for use in urban transport buses: experiment and modeling. report
12  17. MNTRC; 2014 http://transweb.sjsu.edu/PDFs/research/1146-biodie
sel-bus-fuel-combustion-chemistry.pdf.
[13] Omidvarborn H, Kumar A, Kim DS. Characterization of particulate matter
emitted from transit buses fueled with B20 in idle modes. J Environ Chem
Eng 2014;2:233542.
[14] Shandilya KK, Kumar A. Particulate emissions from tailpipe during idling of
public transit buses fueled with alternative fuels, Environ Prog Sustain
Energy 2013; 32: 11341142.
[15] Omidvarborn H, Kumar A, Kim DS, Shandilya KK. Analysis of particulate
matter from the exhaust of biodiesel transit buses under idling conditions.
Air and waste management association annual meeting and conference
exhibition, vol 106; 2014 pp. 271225.
[16] Haynes BS, Wagner HG. Soot formation. Prog Energy Combust Sci
1981;7:22973.
[17] Bockhorn H. Soot formation in combustion. In: Bockhorn H, editor. Springer
series in chemical Physics, vol. 59. Berlin: Springer; 1994.
[18] Richter H, Howard JB. Formation of polycyclic aromatic hydrocarbons and
their growth to soot, a review of chemical reaction pathways. Prog Energy
Combust Sci 2000;26:565608.
[19] Stanmore BR, Brilhac JF, Gilot P. The oxidation of soot: a review of
experiments, mechanisms and models. Carbon 2001;39:224768.
[20] Murugesan A, Umarani C, Subramanian R, Nedunchezhian N. Production and
analysis of bio-diesel from non-edible oilsa review. Renew Sustain Energy
Rev 2009;13:65362.
[21] Misra RD, Murthy MS. Jatropathe future fuel of India. 1364-0321. Renew
Sustain Energy Rev 2011;15:13509.
[22] Kumar N, Varun, Chauhan SR. Performance and emission characteristics of
biodiesel from different origins: a review. Renew Sustain Energy Rev
2013;21:63358.
[23] Choi MY, Hamins A, Mulholland GW, Kashiwagi T. Simultaneous optical
measurement of soot volume fraction and temperature in premixed ames.
Combust Flame 1994;99:17486.
[24] Glassman I. Soot formation in combustion processes. Symp (Int) Combust
1989;22:295311.
[25] Tree DR, Svensson KI. Soot processes in compression ignition engines. Prog
Energy Combust Sci 2007;33:272309.
[26] Xi J, Zhong BJ. Review: soot in diesel combustion systems. Chem Eng
Technol. 2006;29:66573.
[27] Omidvarborna H, Kumar A, Kim DS, Venkata PKP, Bollineni VSP. Characterization and exhaust emission analysis of biodiesel in different temperature
and pressure: laboratory study. J Hazard Toxic Radioact Waste, 19, 2015
(http://dx.doi.org/10.1061/(ASCE)HZ.2153-5515.0000237).
[28] Glassman I. Combustion. 3rd ed. San Diego, California: Academic Press; 1996.
[29] Ishiguro T, Takatori Y, Akihama K. Microstructure of diesel soot particles
probed by electron microscopy: rst observation of inner core and outer
shell. Combust Flame 1997;108:2314.
[30] Kong SC, Sun Y, Rietz RD. Modeling diesel spray ame lift-off, sooting
tendency and NOx emissions using detailed chemistry with phenomenological soot model. ASME J Gas Turbines Power 2007;129:24551.
[31] Miller JA, Melius CF. Kinetic and thermodynamic issues in the formation of
aromatic compounds in ames of aliphatic fuels. Combust Flame
1992;91:2139.
[32] Frenklach M, Wang H. Soot formation in combustion. In: Bockhorn H, editor.
Springer series in chemical physics, vol. 59. Berlin: Springer; 1994. p. 16592.
[33] Richter H, Benish TG, Mazyar OA, Green WH, Howard JB. Formation of
polycyclic aromatic hydrocarbons and their radicals in a nearly sooting
premixed benzene ame. Proc Combust Inst 2000;28:260918.
[34] Vishwanathan G, Reitz RD. Modeling soot formation using reduced polycyclic aromatic hydrocarbon chemistry in n-heptane lifted ames with
application to low temperature combustion. J Eng Gas Turbines Power
2009;131:032801.17.
[35] Frenklach M, Ebert LB. Comment on the proposed role of spheroidal carbon
clusters in soot formation. J Phys Chem 1988;92:5613.
[36] Frenklach M, Clary DW, Gardiner WC, Stein SE. Detailed kinetic modeling of
soot formation in shock-tube pyrolysis of acetylene. Proc Combust Inst
1985;20:887901.
[37] Frenklach M. Reaction mechanism of soot formation in ames. Phys Chem
Chem Phys 2002;4:202837.
[38] Frenklach M, Wang H, Bockhorn H, editors. Detailed mechanism and
modeling of soot particle formation Soot formation in combustion: mechanisms and models. Berlin: Springer-Verlag; 1991. p. 16290.
[39] Harris SJ, Weiner AM. Surface growth of soot particles in premixed ethylene/
air ames. Combust Sci Technol 1983;31:15567.
[40] Nagle J. Strickland-Constable RF. Oxidation of carbon between 1000
2000 1C. In: Proceedings of the fth carbon conference, vol. 1 Pergammon
Press; 1962, p. 154.
[41] Belardini P, Bertoli C, Beatrice C, Danna A, Giacomo ND. Application of a
reduced kinetic model for soot formation and burnout in three-dimensional
diesel combustion computations. Sympos (Int) Combust 1996;26:251724.
[42] Kennedy IM. Models of soot formation and oxidation. Prog Energy Combust
Sci 1997;23:95132.
[43] Kazakov A, Foster DE. Modeling of soot formation during DI diesel combustion using a multi-step phenomenological model. SAE paper 982463; 1998.

646

H. Omidvarborna et al. / Renewable and Sustainable Energy Reviews 48 (2015) 635647

[44] Khan IM, Greeves G, Probert DM. Air pollution control in transport engines.
Inst Mech Eng 1971;C142/71:20517.
[45] Micklow GJ, Gong W. A multistage combustion model and soot formation
model for direct-injection diesel engines. Proc Inst Mech Eng Part D:
J Automobile Eng 2002;216:495504.
[46] Tesner PA, Snegiriova TD, Knorre VG. Kinetics of dispersed carbon formation.
Combust Flame 1971;17:25360.
[47] Hiroyasu H, Kadota T. Models for combustion and formation of nitric oxide
and soot in direct injection diesel engines. SAE technical paper 760129; 1976.
[48] Argachoy C, Pimenta AP. Phenomenological model of particulate matter
emission from direct injection diesel engines. J Braz Soc Mech Sci Eng
2005;27:26673.
[49] Hiroyasu H, Kadota T, Arai M. Development and use of a spray combustion
modeling to predict diesel engine efciency and pollutant emissions: Part
1 combustion modeling. Bull JSME 1983;26:56975.
[50] Khan IM, Greeves G. A method for calculating the formation and combustion
of soot in diesel engines. chapter 25. In: Afgan NH, Beer JM, editors. Heat
Transfer in Flames. Washington DC: Scripta; 1974.
[51] Patterson M, Kong S, Hampson G, Reitz R. Modeling the effects of fuel
injection characteristics on diesel engine soot and NOx emissions. SAE
technical paper 940523; 1994.
[52] Komninos NP, Rakopoulos CD. Modeling HCCI combustion of biofuels: a
review. Renew Sustain Energy Rev 2012;16:1588610.
[53] Vander Wal RL. Soot oxidation: dependence upon initial nanostructure.
Combust Flame 2003;134(12):19.
[54] Srinivas S, Reitz RD, Foster DE, Tao F. Comparison of three soot models
applied to multi-dimensional diesel combustion simulations. JSME Int
J 2005;48:6718.
[55] Golovitchev VI, Tao F, Chomiak J. Numerical evaluation of soot formation
control at diesel-like conditions by reducing fuel injection timing. SAE paper
1999-01-35521; 1999.
[56] Rakopoulos CD, Rakopoulos DC, Giakoumis EG, Kyritsis DC. Validation and
sensitivity analysis of a two zone diesel engine model for combustion and
emissions prediction. Energy Convers Manage 2004;45:147195.
[57] Nishida K, Hiroyasu H. Simplied three-dimensional modeling of mixture
formation and combustion in a D.I. diesel engine. SAE technical paper
890269; 1989.
[58] Yoshizaki T, Nishida K, Hiroyasu H. Approach to low NOx and smoke
emission engines by using phenomenological simulation. SAE technical
paper 930612 1993.
[59] Belardini P, Bertoli C, Ciajolo A, D'Anna A, Del Giacomo N. Three dimensional
calculations of DI diesel engine combustion and comparison whit in cylinder
sampling valve data. SAE technical paper 922225; 1992.
[60] Gorokhovski M, Borghi R. Numerical simulation of soot formation and
oxidation in diesel engines. SAE technical paper 930075; 1993.
[61] Cheng X, Chen L, Hong G, Yan F, Dong S. Modeling study of soot formation
and oxidation in DI diesel engine using an improved soot model. Appl Therm
Eng 2014;62:30312.
[62] Moss JB. Modelling soot formation for turbulent ame prediction, soot
formation in combustion. Springer Ser Chem Phys 1994;59:55168.
[63] Lindstedt PR. Simplied soot nucleation and surface growth steps for nonpremixed ames. Soot formation in combustion. Springer Ser Chem Phys
1994;59:41741.
[64] Hiroyasu H. Diesel engine combustion and its modeling, Diagnostics and
Modeling of Combustion in Reciprocating Engines. Conference on Modeling
and Diagnostics for Advanced Engine Systems (COMODIA) 1985;85:5375.
[65] Khan I, Greeves G, Wang C. Factors affecting smoke and gaseous emissions
from direct injection engines and a method of calculation. SAE Technical
Paper 1973:730169.
[66] Fusco A, Knox-Kelecy AL, Foster DE. Application of a phenomenological soot
model to diesel engine combustion. Conference on Modeling and Diagnostics for Advanced Engine Systems (COMODIA) 1994;94:5716.
[67] Liu Y, Tao F, Foster DE, Reitz RD. Application of a multiple-step phenomenological soot model to HSDI diesel multiple injection modeling. SAE Paper
2005 2005-01-0924.
[68] He C, Ge Y, Tan J, You K, Han X, Wang J. Characteristics of polycyclic aromatic
hydrocarbons emissions of diesel engine fueled with biodiesel and diesel.
Fuel 2010;89:20406.
[69] Idicheria C, Pickett L. Formaldehyde visualization near lift-off location in a
diesel jet. SAE technical paper 2006-01-3434 2006.
[70] Tao F, Golovvitchev VI, Chomiak J. Application of complex chemistry to
investigate the combustion zone structure of DI diesel sprays under enginelike conditions (DE-3) diesel engine combustion 3-modeling). Proceedings of
the conference on modeling and diagnostics for advanced engine systems
(COMODIA) 2001:92100.
[71] Kolaitis DI, Founti MA. On the assumption of using n-heptane as a surrogate
fuel for the description of the cool ame oxidation of diesel oil. Proc Combust
Inst 2009;32:3197205.
[72] Tao F, Reitz RD, Foster DE, Liu Y. Nine-step phenomenological diesel soot
model validated over a wide range of engine conditions. Int J Thermal Sci
2009;48:122334.
[73] Vishwanathana G, Reitz RD. Development of practical soot modeling
approach and its application to low-temperature diesel combustion. Combust Sci Technol 2010;182:105082.
[74] Vishwanathan G, Reitz RD. Application of a semi-detailed soot modeling
approach for low temperature diesel combustion. International

[75]

[76]

[77]
[78]

[79]
[80]
[81]
[82]

[83]
[84]
[85]

[86]

[87]

[88]

[89]

[90]

[91]

[92]
[93]

[94]

[95]
[96]

[97]
[98]

[99]

[100]

[101]

[102]

[103]

[104]

multidimensional engine modeling user's group meeting at the SAE congress. Detroit (MI); April 12, 2010.
Cheng X, Chen L, Yan F, Dong S. Study on soot formation characteristics in
the diesel combustion process based on an improved detailed soot model.
Energy Convers Manage 2013;75:110.
Jia M, Peng ZJ, Xie MZ. Numerical investigation of soot reduction potentials
with diesel homogeneous charge compression ignition combustion by an
improved phenomenological soot model. Proc Inst Mech Eng Part D
J Automob Eng 2009;223:395412.
Tao F, Foster DE, Reitz RD. Soot structure in a conventional non-premixed
diesel ame. SAE paper 2006-01-0196; 2006.
Pang B, Xie MZ, Jia M, Liu YD. Development of a phenomenological soot
model coupled with a skeletal PAH mechanism for practical engine simulation. Energy Fuels 2013;27:1699711.
Ra Y, Reitz RD. A combustion model for IC engine combustion simulations
with multi-component fuels. Combust Flame 2011;158:6990.
Ra Y, Reitz RD. A reduced chemical kinetic model for IC engine combustion
simulations with primary reference fuels. Combust Flame 2008;155:71338.
Frenklach M, Wang H. Detailed modeling of soot particle nucleation and
growth. Int Symp Combust 1991;23:155966.
Appel J, Bockhorn H, Frenklach M. Kinetic modeling of soot formation with
detailed chemistry and physics: laminar premixed ames of C-2 hydrocarbons. Combust Flame 2000;121:12236.
Frenklach M, Wang H. In: Bockhorn H, editor. Mechanisms and models in
soot formation in combustion. Heidelberg: Springer; 1994.
Daly D, Nag P. Combustion modeling of soot reduction in diesel and alternate
fuels using CHEMKINs. SAE technical paper 2001-01-1239; 2001.
Boulanger J, Liu F, Neill WS, Smallwood GJ. An improved soot formation
model for 3D diesel engine simulations. J Eng Gas Turbines Power
2007;129(3):87784.
Mosbach S, Celnik MS, Raj A, Kraft M, Zhang HR, Kubo S, et al. Towards a
detailed soot model for internal combustion engines. Combust Flame
2009;156:115665.
Sukumaran S, Van Huynh C, Kong SC. Modeling soot emissions in diesel
spray using multistep soot model with detailed PAH chemistry. In: International multidimensional engine modeling user's group meeting at the SAE
congress; April 23rd, 2012.
Naik C, Puduppakkam K, Meeks E. Simulation and analysis of in-cylinder soot
formation in a low temperature combustion diesel engine using a detailed
reaction mechanism. SAE Int J Engines 2013;6(2):1190201.
Murugesan A, Umarani C, Subramanian R, Nedunchezhian N. Bio-diesel as an
alternative fuel for diesel enginesa review. Renew Sustain Energy Rev
2009;13:65362.
Bergthorson JM, Thomson MJ. A review of the combustion and emissions
properties of advanced transportation biofuels and their impact on existing
and future engines. Renew Sustain Energy Rev 2015;42:1393417.
Miyamoto N, Ogawa H, Nurun M, Obata K, Arima T. Smokeless, low NOx, high
thermal efciency, and low noise diesel combustion with oxygenated agents
as main fuel. SAE Paper 980506; 1998.
Beatrice C, Bertoli C, Giacomo ND. New ndings on combustion behavior of
oxygenated synthetic diesel fuels. Combust Sci Technol 1998;137:3150.
Song J, Cheenkachorn K, Wang J, Perez J, Boehman AL, Young PJ, et al. Effect
of oxygenated fuel on combustion and emissions in a light-duty turbo diesel
engine. Energy Fuels 2002;16:294301.
Demirbas A. Biodiesel production from vegetable oils via catalytic and
noncatalytic supercritical methanol transesterication methods. Prog Energy
Combust Sci 2005;31:46687.
Herbinet O, Pitz WJ, Westbrook CK. Detailed chemical kinetic oxidation
mechanism for a biodiesel surrogate. Combust Flame 2008;154(3):50728.
Wang H, Reitz RD, Yao M, Yang B, Jiao Q, Qiu L. Development of an
n-heptane-n-butanol-PAH mechanism and its application for combustion
and soot prediction. Combust Flame 2013;160:50419.
Choi CY, Reitz RD. A numerical analysis of the emissions characteristics of
bio-diesel blended fuels. J Eng Gas Turbines Power 1999;121:317.
Kuleshov A, Mahkamov K. Multi-zone diesel fuel spray combustion model
for the simulation of a diesel engine running on biofuel. Proc Instit Mech
Eng, Part A, J Power Energy 2008;222:30921.
Fisher EM, Pitz WJ, Curran HJ, Westbrook CK. Detailed chemical kinetic
mechanisms for combustion of oxygenated fuels. Proc Combust Inst
2000;28(2):157986.
Tran LS, Sirjean B, Glaude PA, Fournet, R, Battin-Leclerc F. Progress in detailed
kinetic modeling of the combustion of oxygenated components of biofuels.
Energy 2012;43(1):418.
Kitamura T, Ito T, Senda J, Fujimoto H. Detailed chemical kinetic modeling of
diesel spray combustion with oxygenated fuels. SAE paper 2001-01-1262;
2001.
Flynn PF, Durrett RP, zur Loye AO, Akinyemi OC, Dec JE, Westbrook CK. Diesel
combustion: an integrated view combining laser diagnostics, chemical
kinetics, and empirical validation. SAE paper 1999-01-0509; 1999.
Curran HJ, Fisher EM, Glaude PA, Marinov NM, Pitz WJ, Westbrook CK, et al.
Detailed chemical kinetic modeling of diesel combustion with oxygenated
fuels. SAE Paper 2001-01-0653; 2001.
Smith OI. Fundamentals of soot formation in ames with application to
diesel engine particulate emissions. Prog Energy Combust Sci 1981;7:27591.

H. Omidvarborna et al. / Renewable and Sustainable Energy Reviews 48 (2015) 635647

[105] Miyamoto N, Ogawa H, Arima T, Miyakawa K. Improvement of diesel


combustion and emissions with addition of various oxygenated agents to
diesel fuels. SAE Paper 962115; 1996.
[106] Nabi MN, Minami M, Ogawa H, Miyamoto N. Ultra low emission and high
performance diesel combustion with highly oxygenated fuel. SAE paper
2000-01-0231; 2000.
[107] Beatrice C, Bertoli C, Giacomo ND, Guido C, Migliaccio M. In-cylinder soot
evolution analysis in a transparent research DI engine fed by oxygentated
fuels. SAE Paper 2002-01-2851; 2002.
[108] Liotta FJ, Montalvo DM. The effect of oxygenated fuels on emissions from a
modern heavy-duty diesel engine. Paper 932734. SAE 1993.

647

[109] Cheng AS, Dibble RW, Buchholz BA. The effect of oxygenates on diesel engine
particulate matter. SAE paper 2002-01-1705; 2002.
[110] Musculus MP, Dec JE, Tree DR. Effects of fuel parameters and diffusion ame
lift-off on soot formation in a heavy-duty DI diesel engine. SAE paper 200201-0889; 2002.
[111] Mueller C, Pitz W, Pickett L, Martin G, Siebers DL, Westbrook CK. Effects of
oxygenates on soot processes in DI diesel engines: experiments and numerical simulations. SAE technical paper 2003-01-1791.

You might also like