Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

FL48CH07-Raman ARI 11 August 2015 15:22

Review in Advance first posted online


V I E W
E on August 19, 2015. (Changes may
R

still occur before final publication

S
online and in print.)

C E
I N

A
D V A

Modeling of Fine-Particle
Formation in Turbulent Flames
Venkat Raman1 and Rodney O. Fox2
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

1
Department of Aerospace Engineering, University of Michigan, Ann Arbor, Michigan 48109;
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

email: ramanvr@umich.edu
2
Department of Chemical and Biological Engineering, Iowa State University, Ames,
Iowa 50011-2230

Annu. Rev. Fluid Mech. 2016. 48:159–90 Keywords


The Annual Review of Fluid Mechanics is online at fine particles, nanoparticles, flame synthesis, aerosols, sooting flames,
fluid.annualreviews.org
nucleation, aggregation, turbulence
This article’s doi:
10.1146/annurev-fluid-122414-034306 Abstract
Copyright  c 2016 by Annual Reviews. The generation of nanostructured particles in high-temperature flames is
All rights reserved
important both for the control of emissions from combustion devices and
for the synthesis of high-value chemicals for a variety of applications. The
physiochemical processes that lead to the production of fine particles in
turbulent flames are highly sensitive to the flow physics and, in particular,
the history of thermochemical compositions and turbulent features they
encounter. Consequently, it is possible to change the characteristic size,
structure, composition, and yield of the fine particles by altering the flow
configuration. This review describes the complex multiscale interactions
among turbulent fluid flow, gas-phase chemical reactions, and solid-phase
particle evolution. The focus is on modeling the generation of soot particles,
an unwanted pollutant from automobile and aircraft engines, as well as metal
oxides, a class of high-value chemicals sought for specialized applications,
including emissions control. Issues arising due to the numerical methods
used to approximate the particle number density function, the modeling of
turbulence-chemistry interactions, and model validation are also discussed.

159

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

1. INTRODUCTION
In gas-phase flames, the presence of certain chemical and thermal conditions introduces pathways
in which the formation of solid particles is thermodynamically favored. Two such examples are the
formation of soot and metal-oxide particles. In the first example, the presence of fuel in excess of
the stoichiometric composition leads to the generation of aromatic rings that ultimately results in
solid particles with complex nanoscale structure. In the latter example, metallic chlorides or other
precursors intentionally added to the flame oxidize in the presence of air in high-temperature
regions, leading to incipient solid particles. Figure 1 shows an example of metal-oxide nanopar-
ticles generated in a turbulent flame. In addition to oxides, other nonoxide ceramic materials
have been produced in high-temperature, self-sustaining chemical reactors (e.g., Axelbaum et al.
1996, Calcote et al. 1991, Glassman et al. 1992) (see the sidebar Flame-Generated Fine Particles).
Flame-generated carbon black is used in automobile tires, titania in pigments, and catalysts and
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

polymers in various applications (see Pratsinis 1998 for more examples). Such particle generation
is intricately linked to the multiscale processes that dictate fluid flow, and variations in the
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

flow configuration can be used to generate a myriad of nanoparticle structures with tailored
properties.

FLAME-GENERATED FINE PARTICLES

Fine particles generated in flames have a characteristic size of less than a few micrometers. Incomplete combustion
of hydrocarbons in internal combustion engines can result in soot particles, which are undesirable when emitted
in automobile- and aircraft-engine exhaust. In contrast, metal-oxide and ceramic fine particles with favorable
chemical/physical properties for use as catalysts or high-performance materials can be produced in flame reactors.
In most cases, the toxicity/utility of flame-generated fine particles depends strongly on the particle size distribution
and morphology. The term nanoparticle is usually reserved for fine particles with a characteristic length of less than
a few hundred nanometers.

100 nm

Figure 1
Yttria-stabilized zirconia nanoparticles produced through flame-spray pyrolysis. The highly aggregated
structure of the nanoparticles results from primary particles formed by nucleation undergoing aggregation
with negligible sintering in the flame. Metal-oxide nanoparticles with high surface-to-volume ratios can
make effective chemical catalysts (Pratsinis 1998). Figure reproduced from Jossen et al. (2005).

160 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

The use of nanoparticles in technological applications is tied to their physiochemical charac-


teristics, including size, morphology, surface characteristics, and composition (see, e.g., Buesser &
Pratsinis 2012, Pratsinis 1998, Rosner 2005). For instance, in soot emissions, although the mass of
PM2.5 : particles with
soot is important, it has become increasingly clear that finer soot particles are carcinogenic because a size less than 2.5 μm
of their ability to pass more deeply into human lungs. The reduction of PM2.5 emissions is seen
as an important step toward air pollution control by the US Environmental Protection Agency
(http://www.epa.gov/pmdesignations/). Similarly, fine particles with a narrow size range, also
known as functional nanoparticles, are increasingly sought because their size and structure can be
used to control physical properties such as the melting point or the band gap for semiconductor
materials (see, e.g., Roth 2007). Titania particles, which are an important product of the flame
process, can possess rutile or anatase structure, with the former used as a pigment in paints and
the latter as photocatalysts (see, e.g., Akhtar et al. 1992, Buesser & Pratsinis 2012, Pratsinis 1998).
The final product properties are directly related to the flame process and are especially sensitive
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

to the exposure time to high-temperature flame conditions. Consequently, product tailoring is


Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

naturally achieved by designing flow configurations in which the particles traverse well-defined
trajectories through the reactor.
At a more fundamental level, fine-particle formation is directly coupled to the flame structure,
which controls the spatial distribution of chemical species in the gas phase. The formation of
precursors (soot) or the oxidation of injected precursors (oxides) is affected by the flame motion,
and the fluid mechanics governing the flame evolution. In laminar flows, in which particles follow
fixed trajectories (assuming steady flow), it is inferred that different initial inception regions lead
to different particle structures and characteristics (see, e.g., Kholghy et al. 2013). When the flow
is turbulent, fine particles are subject to its large spatial and temporal variations (see, e.g., Sung
et al. 2011, 2013). Understanding this multiscale, multiphysics link between fine particles and
the gas-phase turbulent flow is vital for both the design of aerosol synthesis and mitigation of
unwanted emissions.
Given the importance of these classes of nanoparticles, there is a significant body of litera-
ture developed around all aspects of particle formation. There are many prior reviews that have
summarized the chemical and physical transformations associated with particle formation (e.g.,
Buesser & Gröhn 2012, Buesser & Pratsinis 2012, Frenklach 2002b, Glassman 1989, Haynes &
Wagner 1981, Karataş & Gülder 2012, Kennedy 1997, Pratsinis 1998, Pratsinis & Vemury 1996,
Roth 2007, Stanmore et al. 2001, Wang 2011, Wang et al. 2005). However, these works focused
less on the essential interaction between turbulence and these processes. With the rapid advances
in computational science, there has been significant advancement in the modeling and predictive
simulation of complex turbulent reacting flows, and this has recently been extended to particulate
flows as well.
With this in mind, this review focuses on the flow-coupled physics and modeling of soot and
metal-oxide formation in turbulent flames. These two classes of fine particles provide a study in
contrast, each leveraging advances in computational modeling and diagnostic sciences in different
ways. Because soot-formation physics is relatively universal, fundamentally similar in, for example,
aircraft and automobile engines, there have been fundamental efforts to decipher every aspect of it,
often catalyzing advances in diagnostic and computational tools. Conversely, the utility of oxides
is in their structural and chemical diversity, and there has been an explosion of materials and
material properties that can be generated through flame processes. Given the evolving diversity
of the field, computational modeling of oxide formation has followed an engineering approach,
in which relatively low-fidelity models guide engineering design and optimization. Despite the
similarities in physics, these end goals differentiate the modeling approaches, which interestingly

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 161

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

mirrors the broader discussion in computational science on the balance between physical detail
and tractable uncertainty in the assumptions (see, e.g., Babuška et al. 2008 for more details).

2. PHYSICS OF PARTICLE FORMATION


This section provides an overview of the physics and chemistry of fine-particle formation. De-
tailed descriptions can be found in previous reviews (e.g., Buesser & Gröhn 2012, Buesser &
Pratsinis 2012, Frenklach 2002b, Glassman 1989, Haynes & Wagner 1981, Karataş & Gülder
2012, Kennedy 1997, Pratsinis 1998, Pratsinis & Vemury 1996, Roth 2007, Stanmore et al. 2001,
Wang 2011, Wang et al. 2005).

2.1. Key Physical Processes


Access provided by Stockholm University - Library on 08/24/15. For personal use only.

The conversion of gas-phase molecules into nanostructured particles occurs through a series
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

of complex interactions (see, e.g., Friedlander 2000, Fuchs 1964, Kraft 2005), each of which
alters the characteristics of the fine particles. Figure 2 shows the key physical processes that
are common to both soot and metallic oxides (see the sidebar Particle Formation Processes). In
general, gas-phase precursor molecules undergo chemical processes to form the nuclei, which

a Sintering/coagulation b 0.3 cm 0.5 cm 1.5 cm

Aggregation

High-temperature
turbulent flame
Growth and
surface oxidation

7.5 cm 12 cm 20 cm
Nucleation

Precursor
molecules

200 nm

Oxidizer Fuel Oxidizer


and
precursor

Figure 2
(a) Schematic of particle formation in a flame. At the base of the flame, an oxidizer is mixed with fuel and a precursor in the gas phase.
The resulting chemical reactions lead to the nucleation of primary particles. These primary particles can grow by surface addition or (in
the case of soot) be reduced in size by oxidation. Aggregation can occur once the particle number density is large enough. Metal-oxide
nanoparticles can undergo sintering and coagulation when they pass through high-temperature regions of the flame. For applications in
which nanoparticles are the desired product, it is usually necessary to minimize sintering by tailoring the turbulent mixing, which
reduces the particle surface area for a given particle volume. Panel a adapted from Sung et al. (2011). (b) Nanoparticle samples obtained
from flame synthesis at different heights from the burner exit. Panel b taken with permission from Arabi-Katbi et al. (2001).

162 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

PARTICLE FORMATION PROCESSES

 Nucleation: the formation of a solid-phase particle from the gas-phase precursors


 Surface growth: the molecular addition/abstraction of mass to/from the particle surface
 Aggregation: the combination of two existing particles into a single particle (or aggregate) without changing
the total surface area
 Coagulation: the combination of two existing particles into a single particle while minimizing the total surface
area
 Sintering: the reduction of the particle surface area at a constant volume

are the smallest solid-phase particles in the system. The particle then growths through surface
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

reactions and collisions with other particles, leading to aggregates. Depending on the conditions,
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

the constituent solid particles inside an aggregate can fuse to form larger primary particles. In the
case of soot, surface oxidation can destroy the particles. The relative importance of the different
physical processes depends on the type of fine particles and the process conditions. In particular,
the exposure time to particular conditions is important. Because the flow conditions change
significantly with the distance from the flame front or the primary reaction zone, the trajectory
of the fine particles relative to the flame is critical in determining the final particle properties.
The first step in the formation of fine particles is condensation of gas-phase precursors aggre-
gating to form the nucleus. In the case of metallic oxides, precursors are added to the flow, often at
concentrations comparable to the fuel or oxidizer components. In sooting flames, the precursors
that lead to soot formation are generated within the flow in fuel-rich regions close to the flame
front. In the gas phase, there needs to exist a set of precursors that promote the formation of the
so-called nucleus, which is the smallest-sized aggregate of molecules termed a solid particle. In
the case of soot particles, the pathway consists of the formation of ringed aromatics (e.g., Karataş
& Gülder 2012), although at least three different pathways that cause particle inception have
been proposed (D’Anna 2009, Wang 2011). For metal-oxide particles, the process consists of the
oxidation of the precursor, which is usually in the form of a chloride (Pratsinis et al. 1996, West
et al. 2009) or other stable compound (Menz et al. 2014, Yeh et al. 2004) at room temperature.
Soot nucleation occurs on the rich side of the flame, at which a higher hydrocarbon concentration
promotes the formation of the soot precursors. In metal-oxide-particle synthesis, the precursor
mass loading is higher, often constituting a significant fraction of the fuel flow rate. It has been
observed that a higher precursor concentration leads to an increase in the size of the primary
particles that constitute the aggregates (Pratsinis et al. 1996).
In the case of the flame spray process, in which the precursor is added in the form of a liquid
spray, droplet evaporation adds another layer of complexity, leading to differences in the structure
of the flame and the relative position of the precursor molecules with respect to the flame front
(Heine et al. 2006, Hu et al. 2011, Liu et al. 2011, Mädler et al. 2002). The addition of surfactants
to the droplets also alters droplet behavior (Hu et al. 2011, Liu et al. 2011), which has been used
to generate hollow alumina primary particles.
Fine-particle nucleation is followed by solid-phase processes that encompass agglomeration,
coalescence, surface growth, and oxidation (see the sidebar Aggregation, Agglomeration, Coa-
lescence, Coagulation, and Sintering). Agglomeration is the process by which colliding particles
adhere through the van der Waals force to form larger structures but with unchanged constituent
particles. In most high-temperature flames, pure agglomeration is not observed. Instead, some
level of chemical transformation also occurs in which the primary particles in contact with each

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 163

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

AGGREGATION, AGGLOMERATION, COALESCENCE, COAGULATION,


AND SINTERING

Various terminologies are used in the literature to refer to the process by which two colliding particles combine to
form a single particle. In the physics literature, the term aggregation is reserved for cases in which the individual
particle properties (except their velocities) do not change after collision. Thus, the volume and surface area of
the aggregate are v = v1 + v2 and a = a 1 + a 2 , where (v1 , a 1 ) and (v2 , a 2 ) are the volumes and surface areas of
the two incoming particles. However, in the literature on flame-generated particles, this process is referred to as
agglomeration. The terms coalescence and coagulation are reserved for particles with liquid-like properties that
minimize their surface area (i.e., v = v1 + v2 but a < a 1 + a 2 ). Metal-oxide nanoparticles, which are liquid-like
at high temperature, can undergo a two-step process of agglomeration followed by sintering in high-temperature
regions of the flame, resulting in an overall decrease in the particle surface area.
Access provided by Stockholm University - Library on 08/24/15. For personal use only.
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

other fuse to form nonspherical particles. The level of such sintering or coalescence depends on
the process conditions, with the rates highly sensitive to the surface area of the particles and the
local gas-phase temperature. Smaller primary particles will lead to higher sintering rates owing
to larger surface areas, as will prolonged exposure to high temperatures. Conversely, temperature
reduction or higher precursor concentrations prior to particle formation will reduce sintering
rates.
The agglomeration physics itself depends on the size of the nanoparticles that collide. If the
size is smaller or comparable to the mean free path of the gas molecules, the collision rate is
proportional to the volume swept by the colliding partners, similar to that expected from the
kinetic theory of gases. If the size of the aggregates is larger than the mean free path, the collisions
are governed by the diffusive motion of the aggregates. It is often observed that in the nucleation
region, the particles are small enough that the free-molecular regime is approached, whereas with
additional exposure time aggregates grow in size to reach the continuum regime in flame reactors.
Particles that grow to characteristic lengths greater than several micrometers can exhibit inertial
effects (see, e.g., Balachandar & Eaton 2010, Reade & Collins 2000, Sundaram & Collins 1997).
Perhaps the most important processes that determine the surface structure of the fine particles
are the surface growth and oxidation steps. During surface growth, gas-phase molecules are directly
deposited on the surface through either a chemical bond or physical adhesion. The resultant
Free-molecular
regime: regime in
nanoaggregates have modified surface reactivity (Appel et al. 2000, Blanquart & Pitsch 2009).
which the particle size Furthermore, the direct condensation of polycyclic aromatic hydrocarbon (PAH) molecules can
is smaller or equal to lead to mass addition and surface growth (Bisetti et al. 2012). In sooting flames, the interaction of
the gas-phase mean OH or O radicals leads to oxidation. In fact, most of the soot initially produced is later oxidized in a
free path turbulent flame. The emission of soot in a turbulent flame (discussed below) arises via inefficiencies
Continuum regime: in turbulent mixing, by which soot-rich pockets are able to escape oxidation. Depending on the
regime in which the flow conditions, metallic oxides also experience significant surface growth, by which the precursor
particle size is
approximately equal to
molecules chemically react with the surface of the aggregates (see, e.g., Kobata et al. 1991, Pratsinis
the gas-phase mean 1998).
free path Whereas the particle size and total mass are critical for soot emissions, metal-oxide processes are
PAH: polycyclic also tailored to produce specific particle crystalline structures and compositions (see, e.g., Buesser
aromatic hydrocarbon & Pratsinis 2012, Ifeacho et al. 2007). For example, titania particles exposed to higher gas-phase
temperatures or specific dopants (Vemury & Pratsinis 1995) transform into a rutile structure,
which is mostly used in pigments, whereas lower operating temperatures yield the anatase variety

164 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

that has desirable photocatalytic activity. The addition of specific dopants such as silicon dioxide
in small quantities also increases the yield of anatase titania even at higher temperatures. Similarly,
multiple flames and injectors can also be used to generate composite nanoparticles with more than
one type of metal oxide.

2.2. Particle Generation in Turbulent Flames


Turbulence is often a desired feature of a flow to increase fuel-conversion rates and to ensure spatial
compactness of the process. Because of the chaotic and unsteady nature of the flow, turbulence
introduces variations of gas-phase conditions over a range of length scales and timescales (see, e.g.,
Fox 2003, Pope 2000). For the particles, this changes the spatial trajectory and the history of gas
conditions encountered. The change in trajectory can affect collisions with other particles, and the
change in gas conditions affects the evolution of the particle through mass addition/removal and
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

surface physical/chemical properties. Because there are only a few macroscopic control variables
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

in most applications, the multiscale interactions between turbulence and particle evolution are
important factors in shaping the final product.
Here, we point out the important differences between the nucleation of metal-oxide particles
and that of carbonaceous soot particles. The design of aerosol systems for metallic oxides generally
utilizes precursors that have a strong tendency to oxidize to the final product (e.g., Wang 2011).
In other words, the critical limitation in these systems is the mixing that brings the precursors to
the high-temperature region and that exposes the fine particles to the environment for the time
period needed to control the resulting product properties. In this sense, fine-particle formation is
reaction-mixing limited (see, e.g., Gröhn et al. 2011, Pratsinis et al. 1996). For example, placing the
nucleation region close to the flame front leads to increased sintering that results in larger primary
particles, whereas dilution of the precursor stream with air reduces the particle number density
and collision rate. In other words, oxide formation is influenced by large-scale mixing and flame
behavior, but the fast precursor decomposition ensures that the final product is mostly insensitive to
the details of the turbulent flow. The production of soot is different because the precursor, namely
the fuel, is for the most part not alterable. In addition, the conversion of fuel-bound carbon to
soot particles is through a myriad of chemical transformations that are highly dependent on spatial
effects, including strain (see, e.g., Böhm et al. 2001). Consequently, the inception of soot particles
depends not only on the large-scale flow, but also on the small-scale structure of the turbulent
flame. This sensitivity to gas-phase dynamics introduces strong intermittency in the generation
and growth of soot particles (see the sidebar Intermittency).
Figure 3 shows experimental images of instantaneous soot volume fractions as well as
ensemble-averaged statistical quantities (Qamar et al. 2009). It is seen that at any given instant, soot
particles are concentrated in thin layers, and much of the domain actually contains no measurable
soot. From comparison with time-averaged statistics, the instantaneous peak volume fraction is

INTERMITTENCY

In high-Reynolds-number turbulent flows, intense bursts of small-scale vorticity occurring at highly localized
regions in space and time are referred to as intermittency (see, e.g., Pope 2000). For fine particles generated in
turbulent flames (e.g., soot), highly localized regions of high number concentration can result from turbulence
intermittency. Because fine particles have negligible diffusivity, these highly localized regions remain compact for
long periods, leading to pronounced spatial and temporal intermittency in the number concentration.

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 165

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

1.00
fv (ppb)
a b
140

35 0.99
130

Intermittency
30 0.98
120

110
0.97
25
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

100
0.96
x/d
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

20

90
2.0 c
Soot volume fraction, fv (ppb)
15
80
1.5

70
10
1.0

60
5
0.5
50

0 0
–10 0 10 50 80 110 140
r/d x/d

Figure 3
(a) Laser-induced incandescence measurement of instantaneous soot volume fractions in a jet flame. (b) Soot intermittency as a function
of the axial distance in the jet. (c) Average soot volume fraction measurements (symbols) with smoothed data (line). As seen from panel a,
the soot observed in a turbulent flame is highly intermittent in both space and time. Figure taken with permission from Qamar et al.
(2009).

roughly an order of magnitude higher than the average value. This is further confirmed in the soot
intermittency plot in Figure 3b, which measures the probability of not finding soot at a given axial
location. This factor remains close to unity, with a minimum of approximately 0.97, indicating
that roughly only 30 out of 1,000 samples contained measurable soot.
The spatial and temporal intermittency of soot in turbulent flames has been noted by many
authors (e.g., Coppalle & Joyeux 1994; Geigle et al. 2013, 2015; Haynes & Wagner 1981; Hu
et al. 2003). The soot intermittency itself is caused by the relatively slow timescales needed to
form appreciable soot mass compared with the high sensitivity to local strain in the formation of
the precursors that enhance soot inception and growth. Geigle et al. (2013, 2015) studied soot
formation at different pressures and concluded that with an increase in pressure, the intermit-
tency effects are not significantly diminished and that both the size of the soot structures and

166 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

the peak instantaneous soot volume fraction increase. This spatiotemporal structure of the soot
concentration field vastly increases the modeling complexity, as discussed in Section 3.2.
PBE:
population-balance
3. COMPUTATIONAL MODELING OF FLAME SYNTHESIS
equation
OF FINE PARTICLES
NDF: number density
The models for describing these complex turbulent reacting flows can be divided into two cat- function
egories: (a) the mathematical system that describes particle evolution in an inhomogeneous gas-
phase system and (b) the additional modeling approximation necessary to evaluate this mathemat-
ical system in the presence of turbulence. The mathematical description of fine-particle formation
incorporates not only all elements of reactive flow modeling (Fox 2003), but also the description of
the particle phase. This involves solving mass, momentum, and energy equations to describe the
fluid mechanics and gas-phase thermochemical equations to describe the transport of chemical
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

species and the generation of precursors that lead to the formation of the particle phase. Because
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

the number of fine particles is so large, a statistical approach is used to describe the fine-particle
population. This leads to the population-balance approach, in which a transport equation, the
population-balance equation (PBE), for the number density function (NDF) of the particles is
solved (see the sidebar Particle-Phase Population Balance Equation).

3.1. Modeling of Particle Evolution


Describing particle evolution in flames introduces two distinct challenges: The first is representing
so large a number of particles, and the second is describing the physical mechanisms that alter
the particles. As alluded to in the introduction to this section, the particle phase is modeled
using a PBE. This conservation equation governs the NDF, which describes the concentration of
particles as a function of a set of internal coordinates. Although many different choices of internal
coordinates have been considered (see, e.g., Blanquart & Pitsch 2009, Marchisio & Fox 2013), the
volume and surface area are commonly used. The NDF equation is high dimensional, typically
spanning N i + 4 coordinates, where Ni is the number of internal coordinates employed. We note
that the internal coordinates are used to distinguish particles for the purpose of modeling as well
as to generate information for use in process design and optimization. Consequently, the number
of internal coordinates employed depends on the physical description and the structural details of
interest in an application. Given the high dimensionality of the NDF, direct discretization using
grid-based approaches is intractable even for a modest number of internal coordinates (more
than one). Typically, the NDF is obtained indirectly by solving for its moments (see the sidebar
Moment Methods), and several moment methods are available for this (see, e.g., Donde et al.
2012; Frenklach 2002a; Frenklach & Harris 1987; Marchisio & Fox 2005, 2013; Mueller et al.
2009; Yu & Lin 2010). The use of moments to describe the NDF constitutes an implicit model as
well. In essence, the NDF could be recovered exactly given a large number (approaching infinity)
of moments. However, only a small number of moments are solved in practice, leading to errors
introduced by truncation, which limits the accuracy of the particle description.
With this description of the particle phase, models are required both for the progress toward
nucleation from the gas-phase species and for the evolution of the particle characteristics in the
particle phase (see the sidebar Detailed Mechanisms). Apart from fuel oxidation, fine-particle
formation requires additional pathways that describe the formation of precursors leading to the
inception of solid-phase particles. Although there is still considerable uncertainty, a generally
accepted theory for soot-particle inception is through the formation of PAHs. Here, chemical
reactions lead to the formation of aromatic hydrocarbons that then collide to form stable dimers,

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 167

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

PARTICLE-PHASE POPULATION BALANCE EQUATION

A PBE is a conservation equation for the NDF (see, e.g., Marchisio & Fox 2013, Ramkrishna 2000), denoted
by n(t, x, ξ), where each particle type is characterized by an internal coordinate vector ξ , and n is the number
concentration of particles of type ξ at time t and spatial location x. For flame-generated fine particles (e.g., Zucca
et al. 2006), typical internal coordinates are the volume v, surface area a, and surface chemical composition φs .
In addition to spatial transport (i), the PBE contains hyperbolic terms (ii) for continuous changes in the internal
coordinates (e.g., surface growth and oxidation), a source term for nucleation (iii), and integral terms for breakage
(iv) and aggregation (v):
 
∂n ∂ ∂n ∂ ∂ ∂ 
+ · u(ξ, φ)n − D(ξ, φ) + [Gv (ξ, φ)n] + [Ga (ξ, φ)n] + · Gφs (ξ, φ)n
∂t ∂x ∂x ∂v ∂a ∂φs
     
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

(i) (ii)

   
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

= J(ξ, φ) + N (ξ|ξ )α(ξ , φ)n(ξ )d ξ − α(ξ, φ)n(ξ)


   ξ
(iii)   
(iv)
∞ ∞
+ β(ξ − ξ , ξ , φ)n(ξ − ξ )n(ξ )d ξ −
  
β(ξ, ξ , φ)n(ξ)n(ξ )d ξ .
 0
 0

(v)

The gas-phase chemical composition φ is tightly coupled to all of the rate expressions that appear in the PBE.
For example, the particle velocity u(ξ, φ) is modeled as the sum of the gas-phase velocity and a thermophoretic
velocity (Talbot 1981) depending on the spatial gradient of the gas-phase temperature. The particle-phase diffusivity
D(ξ, φ) is inversely proportional to the particle diffusion diameter and proportional to the gas-phase temperature
(Einstein 1905) but is usually negligible for fine particles in turbulent flames. In addition, the physics of fine-particle
growth in turbulent flames leads to stiff rate expressions. This fact, combined with the integro-differential form
of the PBE, makes solving for the NDF a challenge. It is also worth noting that the PBE admits solutions in the
form of delta functions in ξ. For example, the solution with only aggregation and J(ξ, φ) = J0 (φ)δ(ξ − ξ0 ) (i.e.,
all nuclei have identical properties ξ0 ) will have the form of weighted delta functions centered at ξ = kξ0 , where k
is any positive integer.

which is often regarded as the first step in the formation of the particle phase (Frenklach 2002b).
The determination of a comprehensive pathway for soot inception, especially under practically
relevant conditions, has been the focus of a large number of studies (e.g., Bhatt & Lindstedt 2009,
Blanquart et al. 2009, Richter & Howard 2000, Saggese et al. 2014, Wang & Frenklach 1997).
Similar comprehensive models for metal-oxide-particle inception are relatively less common.
For instance, even though titania generation has been widely studied, the precursor oxidation is
often modeled using a simple global one-step reaction, assuming a precursor-limited formation of
fine particles (Pratsinis et al. 1990). Recently, West et al. (2009) formulated a detailed multistep
mechanism for titania in which chemical processes leading to the formation of molecules with
two titanium atoms are modeled and the collision of such molecules is taken to be the first step of
particle nucleation. Such detailed mechanisms have yet to be developed for other metal oxides.
The PBE for fine particles produced in turbulent flames involves a number of source terms
that describe the NDF associated with specific locations in internal coordinate space. In the
particle phase, models for coalescence, aggregation, and particle-surface processes are needed.
The aggregation or agglomeration process occurs through collisions, the rate of which depends

168 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

MOMENT METHODS

In a PBE, the NDF has high dimensionality because of the number of internal coordinates ξ, making the direct
numerical solution in a turbulent flow intractable. In practice, the lower-order moments of the NDF with respect
to the internal coordinates are sufficient for validation with experiments and for characterizing the fine particles
(see, e.g., Cheng et al. 2010; Marchisio et al. 2001, 2003; Piton et al. 2000). For example, the zero-order moment
m0 = n(ξ) dξ is the number concentration of particles, and the first-order moment m1 = ξ1 n(ξ)dξ, with ξ1
defined to be the particle mass, is the mass concentration of particles. Instead of solving for n(t, x, ξ), moment
methods solve for a finite set of moments mk (t, x) by introducing models to close the moment transport equations
(see Marchisio & Fox 2013 for more details).
The form of the moment transport equations is similar to other scalar transport equations used in turbulent-
flame simulations (Fox 2003). For example, starting from the PBE for n (t, x, v) (i.e., using only the particle volume
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

as the internal coordinate) and neglecting breakage, one finds that the moment transport equation is
 ∞ 
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

∞ ∞
∂mk ∂ ∂n
+ · v u(v, φ)n(v) dv −
k
v D(v, φ) dv =
k
v k J(v, φ)dv
∂t ∂x ∂x
 0
 0
  0
 
spatial transport nucleation
∞ ∞ ∞
1
(v + v) − v − v β(v, v  , φ)n(v  )n(v) dv  dv .
 k
k
+k v k−1
Gv (v, φ)n(v) dv + k
2
 0
   0 0
 
volume growth aggregation

In this equation, the spatial transport, volume growth, and aggregation terms are not closed. If thermophoresis
and diffusion are neglected, the spatial transport term is closed:
 ∞ ∞ 
∂ ∂n ∂
· v k u(v, φ)n(v) dv − v k D(v, φ) dv ⇒ · (umk ),
∂x 0 0 ∂x ∂x
where u is the gas-phase velocity. The growth and aggregation terms require knowledge of the NDF to evaluate
the integrals. In moment methods, the NDF is approximated using knowledge of a finite set of moments (see the
Supplemental Appendix for more information; follow the Supplemental Material link from the Annual Reviews
home page at http://www.annualreviews.org).

on the particle Knudsen number (see discussion in Section 2). In practice, this requires knowing the
size of the aggregates and the local gas-phase temperature. Surface processes that involve gas-phase
chemical species require further consideration. In the case of soot particles, both PAH molecules
and acetylene-based growth are accepted as the primary contributors (e.g., Frenklach 2002b,
Wang 2011). Similarly, titania particles generated from titanium tetrachloride precursors involve
the direct adsorption of precursor molecules on the particle surface followed by the desorption of
chlorine and oxidation of leftover titanium atoms. This heterogeneous growth on the surface has
been described using a first-order reaction rate (Ghoshtagore 1970) or a more detailed second-
order rate (Akroyd et al. 2011, Shirley et al. 2011). Hence, this surface-growth mechanism depends
on the local concentration of the precursor molecules, which provides a tight coupling between
gas-phase processes and the particle-phase evolution.

3.2. Modeling of Turbulence Interaction with Particle Evolution


The mathematical system described above requires solutions to transport equations for the velocity
(u), the gas-phase thermochemical composition vector that includes components that affect the

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 169

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

DETAILED MECHANISMS

In the context of flame-generated fine particles, detailed mechanisms describe the chemical transformations occur-
ring in (a) gas-phase combustion, (b) gas-phase oxidation of precursors, and (c) chemical reactions on the particle
surface often between surface-bound chemical entities (i.e., φs ) and gas-phase free radicals (contained in the gas-
phase composition vector φ). In particular, a reasonably complete detailed mechanism for fine-particle formation
should contain the chemical pathways and rate expressions leading from gas-phase species to the nucleation of fine
particles [i.e., J(ξ, φ)] and to surface growth [i.e., Gv (ξ, φ)].

particle evolution (φ), and the NDF moments m. These transport equations, in the form of partial
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

differential equations, can in principle be directly solved using numerical discretization techniques
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

and are termed direct numerical simulations (DNS). However, a turbulent reacting flow exhibits a
broad range of length scales and timescales that increases with increasing Reynolds and Damköhler
numbers (Fox 2003). Because the range of scales determines the number of computational mesh
points and the time step used, DNS is prohibitively expensive for practical flows. Typically,
some form of averaging is used to reduce the range of scales that are directly solved. In the
Reynolds-averaged Navier-Stokes (RANS) approach, ensemble-averaged statistics of the flow
field are solved, whereas in the large-eddy simulation (LES) approach, spatial filtering is used to
evolve only the large-scale features of the flow. This reduction in complexity necessitates modeling
the lost information, which is obtained using statistical properties of the underlying physics. In the
discussion below, we focus specifically on LES given its recent dominance in the field of turbulent
combustion (Haworth 2010, Pitsch 2006). The modeling issues identified are nevertheless equally
applicable to RANS approaches as well.
LES is based on the concept of spatial filtering, by which a low-pass filter is used to sepa-
rate the flow field into large scales that are evolved using filtered flow equations and small scales
whose effects are modeled (see, e.g., Pope 2000). Although this approach is clearly useful for
shear-dominated flows that are controlled by large scales, the importance of small-scale reaction
processes that support combustion and control fine-particle formation might indicate that LES
is less suited for such reacting flows. However, LES has been successful in predicting turbulent
reacting flows, as evidenced by validation exercises that demonstrate superior comparisons with
experimental data. The main reason for this success is that, in the absence of significant local extinc-
tion, the flame structure is reasonably described by conserved scalar approaches. Consequently,
the flame behavior is controlled predominantly by large-scale mixing, which is represented ex-
plicitly by LES (Pitsch 2006). However, flows that do exhibit local extinction, or flows with other
complexities, including particle formation, multiphase combustion, and shock-dominated physics,
require comprehensive modeling of the small-scale physics and their influence on the large-scale
DNS: direct flow. In this sense, LES modeling is similar to that used in the RANS approach.
numerical simulation If filtered variables are denoted by ˜·, then the governing equations for ũ, φ̃, and m̃ will contain
RANS: terms that represent correlations between these components. These terms cannot be represented
Reynolds-averaged directly by filtered quantities and need to be modeled. The impact of the combustion and par-
Navier-Stokes ticle formation on the velocity field is essentially through heat release and mass change from
LES: large-eddy the gas to the solid phase. The heat release effect is central to all combustion problems and in-
simulation troduces a density change. The mass-change problem is often encountered in multiphase flows
(see, e.g., Reveillon & Vervisch 2005), and a similar approach is used with particulates in which
a mass sink term appears in the continuity equation that accounts for the phase change (see, e.g.,

170 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

Mueller & Pitsch 2012). In the filtered or averaged equations for thermochemical composition
and NDF moments, two main closure terms arise: (a) the unresolved turbulent flux and (b) the
source-term closure. The first term describes the turbulent transport of scalars and the transfer
PDF: probability
of energy/chemical concentrations from the resolved scales to the unresolved small scales. The density function
second term arises from the fact that the chemical source term is nonzero only on thin structures
FGM:
at the small scales, and thus the average source term over a filter width cannot be accurately flamelet-generated
represented by the source term computed at the average thermochemical state because of the manifold
high degree of nonlinearity in the source term. The first term is almost always modeled using the
gradient-diffusion hypothesis (Pope 2000), except in premixed flames, in which countergradient
diffusion needs to be taken into account (e.g., Poinsot & Veynante 2001). The chemical source
term requires a more detailed description.
LES, similar to RANS, uses a one-point, one-time representation of the small-scale or un-
resolved physics (Langford & Moser 1999). In other words, we consider the set of variables for
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

which closure is needed at the small scales as ψ (where we have excluded velocity from this set for
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

the present discussion). Because these variables represent random variables in a turbulent flow, it
is possible to associate a joint probability density function (PDF) that provides their statistics. In
this one-point closure, if the joint PDF of ψ = {φ, m} is known as a function of space and time,
the chemical source term could be exactly closed. Note that the first moment of this PDF will
provide the filtered variables ψ̃. This description is central to turbulent combustion modeling,
and the different models (e.g., flamelet, conditional moment closures, transported PDF) represent
different approximations of this PDF. What is different about particle formation is the inclusion
of the particle NDF moments in addition to the thermochemical composition vector φ. This
joint PDF should then express the spatial correlation that creates the intermittency effects (see
Section 4). It is interesting to note that the NDF is necessary even for laminar flames in which the
flow-induced randomness is absent. The inclusion of the moments in the state space for turbulent
flows is a second layer of statistical representation.
Generally speaking, there are two different approaches to modeling turbulent combustion.
In the flame-structure-based approach, the thermochemical vector is represented in terms of a
reduced set of variables, and the mapping between the complete set and the reduced set is obtained
by the use of a flame structure. This approach includes flamelet models (e.g., Peters 2000, Pierce
& Moin 2004, Pitsch & Steiner 2000, Raman & Pitsch 2005), flamelet-generated manifold (FGM)
approaches (e.g., van Oijen & de Goey 2000, van Oijen et al. 2001), tabulated chemistry approaches
(e.g., Fiorina et al. 2010, Gicquel et al. 2000), and flame-surface-based descriptions (e.g., Duclos
et al. 1993, Hawkes & Cant 2000). The other technique is statistical, in which the PDF is directly
obtained by solving its transport equation (e.g., Colucci et al. 1998; Fox 2003; Pope 1981, 1985;
Raman & Pitsch 2006; Raman et al. 2004, 2006) or by splitting the PDF into a conditional PDF
and a marginal PDF of a mapping variable. Conditional moment closure (e.g., Klimenko & Bilger
1999, Steiner & Bushe 2001) and multi-environment closures (e.g., Fox & Raman 2004) are
examples of this latter approach. When dealing with the augmented state space that accounts for
soot or fine-particle evolution, the natural approach has been to extend these techniques directly.
However, there are some subtle differences in the nature of these extended space variables that
introduce additional complexities, with varying effects on each of the models mentioned above.
The flame-structure-based approach has become the most commonly employed technique for
LES, as the decoupled chemistry does not influence the computational cost of the simulations.
In these approaches, a laminar flame, such as a counterflow diffusion flame or a one-dimensional
premixed flame, is used to map the relation between a conserved or reacting scalar and the ther-
mochemical composition vector. It is then assumed that the turbulent flame is a collection of such
laminar flamelets at the small scale and is described by the distribution of the conserved scalar

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 171

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

at the subfilter scale. Hence, the PDF of the conserved scalar along with the mapping function
(usually in the form of a precomputed table) is sufficient to describe the flame. In practice, addi-
tional scalars (e.g., enthalpy) or parameters (e.g., the dissipation rate) are introduced to describe
the possible states that will be encountered in a turbulent flame.
The assumption that the flame structure is not affected by the details of the turbulent flow is valid
only as long as the reaction processes are small scale. Some of the physical processes that control
soot and fine-particle evolution are considerably slower, with length scales and timescales much
larger than those associated with the reaction layer in which much of the fuel oxidation takes place.
In this case, turbulence can introduce variabilities in thermochemical composition that will not
be observed in any laminar flame. Consequently, the state space accessed by the thermochemical
composition vector is problem dependent and generally insufficiently described by reduced-order
flame structures. In practice, this implies that the soot or fine-particle distribution and precursor
concentrations cannot be mapped based on simpler flames, as is done in the flamelet or FGM
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

descriptions.
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

In oxide processes, the irreversible generation of solid particles ensures that the particular time
history of evolution is important (Sung et al. 2011). Consequently, additional transport equations
to spatially and temporally evolve the fine-particle and precursor distributions are needed. From
the PDF viewpoint, this would imply that the joint PDF of the scalars used to describe the flame
structure and gas-phase precursor evolution as well as the moments is needed. The common
practice is to assume that the variables are independent, and the joint PDF is described as the
product of the marginal PDFs. In this sense, the statistical approach is more comprehensive
because it does not make any assumptions about the flame structure and provides the joint PDF
without simplifications, although its weakness is the description of the small-scale mixing process
(see, e.g., Fox 2003, Pitsch 2006, Raman et al. 2005).
Another important issue in soot formation is the phenomenon of differential diffusion (see the
sidebar Differential Diffusion). Because of their size, most fine particles do not diffuse significantly,
and thus the moment equations that describe the particle population do not contain a laminar
diffusion term. This leads to the differential diffusion of fine particles with respect to the gas
phase, which needs to be taken into account. In the flame-structure-based methods, the resolved-
scale differential diffusion is naturally treated by explicitly solving transport equations for the
filtered moments of the NDF that do not contain this diffusion term, whereas the mapping-
species equations contain the appropriate diffusion term. But the effect of small-scale differential
diffusion does not explicitly appear in the equations because it is assumed that the particles and gas
phase are essentially decorrelated at the small scales. This is implied by treating the joint PDF of
the gas-phase thermochemical composition and solid-phase moments of the NDF as the product
of two independent PDFs.
The treatment of differential diffusion is a long-standing issue in combustion modeling (e.g.,
Fox 1999, Pitsch & Peters 1998, Pope 2000) and is more complicated in the statistical approach.

DIFFERENTIAL DIFFUSION

When different chemical species have different diffusion coefficients, transport and reactions at the smallest scales
of a turbulent flame may be modified. Because of their large size relative to gas-phase chemical species (Einstein
1905), fine particles have negligibly small diffusion coefficients (i.e., infinite Schmidt number) and hence do not
diffuse with respect to the gas phase (Fox 2003). Thus, the differential diffusion of soot particles can leave them
stranded in fuel-rich regions of a turbulent flame in which the probability of being oxidized is low.

172 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

The conventional Fokker-Planck-type expression of the PDF transport equation does not allow
for differential diffusion at the large scales, but the small-scale part can be represented by a
conditional diffusion model that needs to be closed (Fox 2003, Pope 1985). A recent formulation
by McDermott & Pope (2007) recasts the PDF transport equation in such a way that large-scale
differential diffusion can in principle be described. However, no extension to fine-particle-laden
flows has emerged. More commonly, both the large- and small-scale differential diffusion are
ignored (Donde et al. 2013, Lindstedt & Louloudi 2005), which can alter the structure of soot in
composition space. Conditional moment closure approaches face a similar hurdle, in which the
conditional scalar dissipation rate needs to include differential diffusion effects (Kronenburg et al.
2000).
Finally, another critical issue in dealing with fine particles, especially soot particles, is the
intermittency associated with their spatial and temporal distributions. Because soot formation
is confined to narrow regions of thermochemical composition space, particles are oftentimes
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

confined to very thin regions of physical space as well, as confirmed by experiments. In addition,
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

soot itself is present only at sporadic intervals, so the formation and oxidation processes are strongly
sensitive to the local strain rate history they experience. The small-scale structure of soot is thus
expected to have a sparse spatial distribution, with a few subregions of very high soot content and
other regions with very small or absent soot content. In the flamelet-based approach, the marginal
PDF of the moments can be expressed as delta functions representing this intermittency, which
are then specified using the average and variance of the soot moments (Mueller & Pitsch 2011).
In the statistical approach, the PDF should technically yield this distribution directly, but this
in turn depends on the small-scale conditional diffusion model that describes the movement of
soot in composition space due to differential diffusion. Currently available models for small-scale
conditional diffusion are not capable of representing this type of distribution.
Because soot is strongly sensitive to the gas-phase composition for both its formation and
subsequent evolution (including oxidation), such effects become important in accurately capturing
the physics. Metal-oxide fine-particle formation is generally less sensitive, at least from existing
evidence, because the precursor oxidizes relatively easily. Although many soot-formation models
have been extended to metal oxides (Akroyd et al. 2011, Sung et al. 2011), the focus has been
more on predicting the precursor oxidation rate correctly rather than the spatial structure of the
oxidation process. As discussed in Section 5, some of these engineering approximations in oxide-
synthesis simulations are based on the simpler nature of precursor oxidation, but some are also
born out of the fact that the fidelity of validation data does not support a finer level of detail in
the models.

3.3. Numerical Issues in the Modeling of Particles


The numerical solution of the governing equations involves both spatial and temporal discretiza-
tion. Although any discretization technique for this is inherently an approximation, the error
introduced can be controlled by analyzing the convergence of the solution with a decrease in the
mesh spacing or time step used. However, LES and NDF formulations generally render conver-
gence studies inconsequential or intractable. Because these numerical issues affect the final utility
and the process of model validation, it is important to assess and minimize these errors.
The interaction of numerical errors and modeling is well known in the field of LES (see,
e.g., Gullbrand & Chow 2003, Kaul & Raman 2011, Kaul et al. 2009, Langford & Moser 1999,
Vasilyev et al. 1998). Errors in numerical discretization increase with a decrease in length scales
such that the largest errors are incurred at scales comparable to the mesh spacing. In practically
all LES, the filter width used to separate the large and small scales is equal to the local mesh

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 173

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

spacing. Consequently, numerical errors contaminate the filter-scale flow physics. All subfilter
models invariably use the filter-scale quantities and their gradients, which introduce numerical
errors into the evaluation of the unresolved terms. Of particular relevance are the models for
subfilter variance, which describe the level of scalar mixing at the small scales. Because of the high
sensitivity of soot formation to the gas-phase composition, small errors in variance models can
lead to disproportionate errors in soot models. More importantly, different computational grids
are now associated with different filter widths, which implies that varying ranges of length scales
are being resolved in these calculations. Hence, solution convergence by reducing mesh size is not
meaningful. Although explicit filtering techniques that decouple the mesh from the filter have been
proposed (e.g., Bose et al. 2010, Vasilyev et al. 1998), their utility in inhomogeneous flows with
boundary conditions has yet to be established. Other frameworks for LES based on a rigorous
statistical representation are also being considered (Fox 2012, Langford & Moser 1999, Pope
2010). In the interim, practical measures such as model reformulation have provided significant
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

improvement (Donde et al. 2012, Kaul et al. 2009).


Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

An additional, and often neglected, issue is the representation of the PBE in terms of transport
equations for the moments of the NDF. The primary use of the moments-based approach is its
computational efficiency, but the reconstruction of the NDF from the moments may not always
be unique or accurate. Although there is evidence that the use of the lower-order set of moments
(especially the first five for a univariate description) to close the higher-order moments is reason-
ably accurate (e.g., Frenklach 2002a, Frenklach & Harris 1987, Yuan et al. 2012), this may not be
a universal feature. Other studies show that the choice of method influences the prediction of soot
characteristics (e.g., Mueller et al. 2009). Because the convergence may not be smooth with regard
to the moment order, it is not clear if a convergence study that uses an increasing set of moments
to describe the NDF, even if it were possible, would be sufficient. In interpreting the validation
of particle models, and in their practical use, this source of numerical error needs to be included.
Finally, because of the special meaning associated with the moments, the numerical discretiza-
tion, especially of the convective term in the transport equations, should be considered carefully.
The convective term shares numerical challenges, particularly stability, common to hyperbolic
systems, and many numerical methods have been developed to address these without excessive
dissipation that might alter the turbulent flow (e.g., Herrmann et al. 2006, Jiang & Peng 2000,
Leonard 1979). Most of these techniques are designed for transporting a single scalar. Hence,
even in combustion-related problems that involve multiple scalars, the convective term is dis-
cretized independently for each scalar without taking into consideration the values and evolution
of other scalars in the system. On the contrary, moments that describe an NDF introduce special
constraints.
A moment vector m0 , m1 , . . . , m2N only makes sense physically if it is generated by a nonneg-
ative NDF. It can be shown that a realizable moment vector is confined to a convex set (Dette &
Studden 1997, Shohat & Tamarkin 1943) and that NDF reconstruction algorithms are defined
only for realizable moment vectors (Marchisio & Fox 2013). Therefore, it is imperative that the
numerical methods used to solve the moment transport equations yield realizable moments; how-
ever, the design of realizability-preserving methods is nontrivial (see, e.g., Vikas et al. 2011). In
practice, the spatial convective term in the moment transport equations often violates realizability
constraints (McGraw 2007, Wright 2007). Although Lagrangian schemes can be used to solve the
convection term (Attili & Bisetti 2013) (which partially overcomes the realizability issue), these
schemes have limited ability to produce spatially high-order approximations for the moments
owing to the small number of samples employed. Another approach for developing realizability-
preserving methods is to work with a reconstructed NDF (see the Supplemental Appendix for
more information).

174 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

4. COMPUTATIONAL ANALYSIS OF PARTICLE FORMATION


This section presents results from computational studies that use the methods and models discussed
in the previous sections. The studies are classified based on the approach to turbulence modeling
used in each. Results are selected for presentation to illuminate the complexities of soot and
metal-oxide fine-particle evolution in turbulent flames.

4.1. Direct Numerical Simulations


In the present context, DNS indicates the complete resolution of the turbulence and chemical
timescales, whereas the soot particles are represented using the NDF. DNS has been widely used
to study soot formation, in which the gas-phase chemistry plays a critical role in the formation
and subsequent growth of the particles. Yoo & Im (2007) performed two-dimensional DNS of a
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

nonpremixed flame, demonstrating that an increased soot volume fraction in turbulent flames, as
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

compared to laminar steady flames, might result from the rapid transport of soot particles out of
high-temperature regions rather than enhanced chemistry. In addition, they concluded that the
larger flame volume leads to increased production of soot. Further analysis by Lignell et al. (2007,
2008) showed that soot is not confined to small regions of mixture-fraction space but is present
over almost the entire range, indicating that a simple flamelet-type relation is not valid. They
also showed that there can be local flame events that cause the soot to break through the flame
region without being fully oxidized. In large-scale pool fires, the presence of radiation-enhanced
quenching can enhance soot breakthrough significantly (e.g., Henriksen et al. 2009).
These DNS calculations used simplified soot-precursor chemistry and particle-evolution
models to allow for a detailed description of the fuel-oxidation process. Recently, DNS using
detailed precursor chemistry has been conducted (Attili et al. 2014, 2015; Bisetti et al. 2012,
2014). A particular focus of these studies is the role of gas-phase precursors on the soot-formation
process. Here, soot inception is modeled using naphthalene dimers rather than the acetylene-
based model (Leung et al. 1991) that is often used. The simulations showed that naphthalene is
highly sensitive to the scalar dissipation rate, whereas acetylene showed little sensitivity in the
scalar dissipation rate. Because soot formation is a nonlinear function of the gas-phase precursors,
especially naphthalene and other PAHs, this sensitivity to the dissipation rate was found to
manifest as intermittency in the soot concentration fields. Figure 4 shows the highly intermittent
instantaneous soot and naphthalene concentrations. Very thin soot and precursor structures
are visible, which is consistent with the experimental data (see Section 2.2), implying that the
evolution of the soot concentration fields is sensitive to the structure of the dissipation and
strain rates in the flow. Because these are essentially small-scale quantities, models are needed to
describe not just their spatial structure but also their temporal evolution. (This aspect is further
discussed in Section 4.2.) Attili et al. (2014, 2015) and Bisetti et al. (2014) also determined that
turbulent motions cause soot to be predominantly transported to richer mixture fractions, whereas
only a small fraction is transported toward the lean side at which oxidation removes soot mass.
DNS studies of metal-oxide particles are less common and mostly focused on titania particles
for which the chemistry is better known. Wang & Garrick (2005) and Garrick & Wang (2010)
conducted DNS of titania formation from titanium tetrachloride in methane-air flames using
one-step global chemical mechanisms for both fuel and TiCl4 oxidation. They showed that particle
generation is not collocated with fuel oxidation, signifying the differences in the oxidation chem-
istry of the metal-oxide precursor and the fuel. It should be noted that the coupling between the
two oxidation processes is only through temperature, with higher temperature through exothermic
reactions of methane combustion driving the precursor oxidation process. Singh & Raman (2012)

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 175

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

Soot Naphthalene Soot Soot


mass fraction mass fraction number density mass fraction

a b c d
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

x
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

y
Ys × 10 4 YA2 × 5,000 Ns × 10 –13 (cm 3) Ys × 10 3

0 2 4 6 0 0.5 1.0 0 0.5 1.0 0 0.5 1.0

Figure 4
Direct numerical simulations with detailed chemistry showing the spatial intermittency of soot formation. (a) Contour plot of the soot
mass fraction (first-order moment in volume). (b–d ) Contour plots of the (b) naphthalene mass fraction, (c) soot number density
(zero-order moment), and (d ) soot mass fraction in the domain within the yellow box in panel a. The soot number density is highest in
regions of strong nucleation, but there the soot particles are small. The highest soot mass fraction occurs in regions in which the
number density is relatively low, indicating that the soot particles in these regions are large and hence more difficult to eliminate
through surface oxidation. Figure adapted with permission from Bisetti et al. (2014).

conducted two-dimensional DNS computations using detailed chemical kinetics that account for
the interaction between methane combustion and intermediates of TiCl4 oxidation. Interestingly,
the simulations showed that a stable intermediate is the most preferred product, and the oxidation
to titania is severely limited by the formation of this unwanted product. In addition, the simulations
demonstrated that the formation of titania occurs at locations farther away from the flame surface,
confirming assumptions used in other studies. Extensions of such detailed chemistry studies to
the characteristics of the titania particles and for other metal oxides have yet to be reported.

4.2. Large-Eddy Simulations


The use of LES for reactive flow modeling has become increasingly common as the computing
power available increases. Relatedly, the range of flows accessible to LES has also increased,
starting from simple jet flames (Elasrag & Menon 2009, Elasrag et al. 2007, Mueller et al. 2013,
Xuan & Blanquart 2015) to realistic combustor models (Koo et al. 2015, Lecocq et al. 2014,
Mueller & Pitsch 2013). As an example of the realistic configurations that are currently accessible,
Figure 5 shows the simulation of an aircraft-engine combustor (Mueller & Pitsch 2013). It is
clear that a spatial intermittency of soot is present in these LES calculations. The time-averaged
statistics further indicate that the high soot content visible in the near-injector region of the
instantaneous contour is not present very often, leading to a much lower time-averaged soot
volume fraction. In this sense, the level of soot intermittency itself is a spatially varying quantity.

176 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

a b

Figure 5
(a) Instantaneous and (b) time-averaged contours of the soot volume fraction inside an aircraft-engine combustor. The isoline indicates
the stoichiometric mixture fraction. The intermittency of the soot spatial distribution can clearly be seen in the instantaneous contours.
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

The spatial and temporal intermittency of soot makes the accurate prediction of soot emissions from aircraft engines using numerical
simulation quite challenging. Figure reproduced with permission from Mueller & Pitsch (2013).
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

Although LES can show intermittency, these fluctuations are necessarily quantitatively reliable.
LES, by definition, should capture multipoint but single time-resolved-scale statistics (Langford
& Moser 1999, Pope 2010). Even if LES captures the mean soot volume fraction correctly while
exhibiting sporadic generation of soot particles, it is not guaranteed that this leads to the correct
temporal structure of soot formation. Because particle evolution occurs over long timescales,
comparable to the integral scales of the motion, it is essential that the history of particle-gas
interactions be predicted correctly. In all one-point formulations, this time history is not directly
considered. LES models impose a local equilibrium assumption, which assumes that at the small
scales the flow reaches statistical equilibrium corresponding to isotropic flow at each time step.
These assumptions, although necessary to make LES tractable, may also lead to inaccuracies in
the description of the time-dependent behavior. Nevertheless, the use of LES marks a critical
evolution in the mathematical description of such complex physics.
Consistent with the engineering approach to nanoparticle synthesis, the use of LES and other
detailed flow models has emerged only recently (see, e.g., Loeffler et al. 2011; Sung et al. 2011,
2013). Figure 6 shows LES of titania synthesis from a titanium tetrachloride precursor in a
methane-air flame. For this simulation, detailed chemical kinetics (West et al. 2009) combined
with a flamelet model were used. The simulations show that much of the precursor oxidation
occurs very close to the injection region and away from the main combustion zone. Previous
simulations using global kinetic models demonstrated that particle generation occurs at much
lower temperatures compared to the flame temperature ( Johannessen et al. 2000). However, the
effective use of LES for nanoparticles is emerging, limited in part by the availability of rigorous
validation data.

4.3. Reynolds-Averaged Navier-Stokes Simulations


In the area of turbulent combustion modeling, it has long been recognized that the use of RANS as
compared to LES leads to a nearly insurmountable modeling challenge. Because neither large-scale
mixing nor small-scale combustion is explicitly represented, the simulation results are highly sensi-
tive to the many model parameters. Despite this deficiency, RANS is computationally inexpensive
as compared to LES. Consequently, it still remains the most commonly used tool for engineering
design and for simulating complex geometries (see, e.g., Blacha et al. 2011, Zhong et al. 2012).
Several studies of sooting flames using RANS have been reported. Consistent with the DNS
findings, Wen et al. (2003) showed that the use of a PAH-based inception model improves soot

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 177

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

Temperature, T (K) [TiCl4 ] (mol/m3) Particle size (nm)

300 775 1,250 1,725 2,200 0 0.34 0.68 1.0 2.0 0 100 200 300
40
a b c

30
x/D

20
Access provided by Stockholm University - Library on 08/24/15. For personal use only.
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

10

0
–10 0 10 –10 0 10 –10 0 10
r/D r/D r/D

Figure 6
Instantaneous snapshots of (a) temperature with an isocontour of the stoichiometric mixture fraction, (b) precursor concentration, and
(c) average particle size from a simulation of titania synthesis in a nonpremixed methane-air flame from Sung et al. (2011) using the
detailed mechanism of West et al. (2009). The precursor TiCl4 is injected in the center stream and is largely consumed before the flame
front. Most of the nucleation occurs near the base of the flame, after which primary particles rapidly aggregate into larger particles.
Aggregates that pass through the flame front can undergo sintering owing to the high temperature. Above the flame front, the principal
growth mechanism is aggregation.

volume fraction predictions compared to an acetylene-based description. The details of the


treatment of turbulence-chemistry interaction modeling were found to be critical in diesel-
engine-relevant simulations, in which changes to the way soot source terms are computed led
to large changes in the location and volume fraction of the soot formed (Bolla et al. 2014).
Overall, some of the simulations show good agreement with experimental data (Chittipotula
et al. 2011, Kronenburg et al. 2000, Lindstedt & Louloudi 2005, Said et al. 1997, Woolley
et al. 2009), especially for jet-like diffusion flames, for which RANS methods are generally more
reliable even for gas-phase species predictions. Conversely, studies with different models show
larger discrepancies (Köhler et al. 2012, Reddy et al. 2015, Zucca et al. 2006). These simulations
indicate that even when the statistics of the flow and major gas-phase species are predicted well,
soot predictions are not equally good. Much of this discrepancy has been attributed to the lack
of unsteady flow description in the RANS approach.
Given its low computational cost, RANS is a viable approach for computational design and
optimization, in which an ensemble of simulations is needed to assess the reactor behavior. The
engineering approach here is to use a small number of experiments to calibrate the model and
then to scan the parameter space computationally to identify ideal regions of operation. Such
calibration has been used for both oxide and soot calculations. Johannessen et al. (2001) employed

178 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

HAB = 27

SiO2 shell
thickness
(nm)
3

1
0.4
Access provided by Stockholm University - Library on 08/24/15. For personal use only.
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

HAB = 19

INC R E A S I N G LE V E LS O F S I LI CA P R E CU R S O R I N J E CT I O N R A T E S

Figure 7
Evolution of silica coating thickness for increasing levels of silica precursor injection rates ( from left to right). The streamlines are
colored based on the coating thickness at that location, and the red isosurface corresponds to 1% concentration of the silica precursor
as compared to the injection value. Figure adapted with permission from Buesser & Pratsinis (2011). Abbreviation: HAB, height above
burner.

a one-step precursor-oxidation model in which a single oxidation temperature is used to describe


the conversion to the solid phase. In that study, the sintering-rate parameters were calibrated
such that the simulations matched a subset of the measured particle properties. The resulting
calibrated model was found to predict other measured properties with reasonable accuracy. A
similar optimization technique has been employed for soot modeling, with genetic algorithms
along with experimental data used to optimize model parameters (Chittipotula et al. 2012). Ma et al.
(2005) employed an optimal combination of models by testing the submodels against experimental
data. The reduced cost of RANS allows the computation of complex systems with either complex
fluid mechanics or secondary processes such as coating. Furthermore, such tools can be easily
used to design reactors (Buesser & Pratsinis 2011). Figure 7 shows a model reactor simulation for
coating titania particles with SiO2 particles to reduce the photocatalytic activity for NOx storage.
RANS allows the optimization of the mixing design and provides qualitative trends in the mixing
performance with changes in the geometry.

5. MODEL VALIDATION USING EXPERIMENTAL DATA


Validation is an assessment of the ability of a model to capture reality, or true physics. Experimental
measurements, performed at conditions that are overall similar to those in which the models will
be used, are often the basis for such validation exercises. Computational models that describe
multiscale, multiphysics systems are invariably complex and involve a number of submodels, as
discussed above. To understand the performance of such complex models, one must focus not only
on the final outcome, the particle characteristics in a turbulent flame, but also on the accuracy of
the individual submodels. Here, validation will involve a hierarchy of experiments that test the
component physics as well as their interactions.

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 179

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

5.1. Sooting Flames


Although laminar flame (e.g., Karataş & Gülder 2012) and turbulent flame (e.g., Coppalle &
Joyeux 1994, Kent & Honnery 1987) experiments had been conducted in the past, the tremendous
advancement of laser diagnostics has allowed much more detailed measurements recently (e.g.,
Desgroux et al. 2013, Franzelli et al. 2015, Mahmoud et al. 2015, Narayanaswamy & Clemens
2013, Qamar et al. 2009, Zhang et al. 2011). Of particular interest are simultaneous measurements,
in which two or more flow quantities are measured to provide joint statistics. Examples include the
velocity-soot volume fraction (Narayanaswamy & Clemens 2013), velocity-OH (Geigle et al. 2013,
2015), PAH-OH (Franzelli et al. 2015), and temperature-soot volume fraction (Chan et al. 2011).
Apart from the macroscopic comparison of soot or composition fields, quantitative simultaneous
measurements allow specific assumptions about the turbulence and its influence on chemistry
to be validated. In addition, the diagnostic tools are increasingly being applied to complex flow
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

configurations that are more representative of practical combustors (see, e.g., Chatterjee et al.
2014; Geigle et al. 2013, 2015).
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

The availability of well-characterized experimental data has enabled detailed validation as well.
Figure 8 shows a comparison of LES calculations of a sooting bluff-body stabilized ethylene flame
and the corresponding experimental data (Mueller et al. 2013). The LES calculations are able to
predict the experimental soot volume fraction with reasonable accuracy. Here, the submodels that
describe the chemistry, precursor evolution, and other processes have been validated separately
using other experimental data. Although this result is encouraging, other such comparisons indicate
far less accuracy. In comparisons of natural gas flames (Donde et al. 2013, Mueller & Pitsch
2012), the simulations predict a much higher soot volume fraction compared to experiments,
whereas in other cases, such as the ethylene/hydrogen configuration (Mahmoud et al. 2015),
the simulations predict far less soot [see the International Sooting Flames Workshop website
(http://www.adelaide.edu.au/cet/isfworkshop/)]. Part of these wild fluctuations in predictive
accuracy results from the extreme sensitivity of soot formation to the gas-phase composition.
Consequently, even small errors in the prediction of gas-phase properties can lead to large errors
in soot predictions. For example, Mueller & Raman (2014) showed that less than a 10% error in
temperature predictions upstream of the soot-forming region in a jet flame can lead to a nearly
30% error in the prediction of soot farther downstream. However, the more critical issue is the
lack of a framework to identify the key modeling errors.

5.2. Metal-Oxide Particles


Given the large variety of metal-oxide particles (for which characteristics beyond the volume and
surface area may be necessary in certain cases), soot-like detailed modeling has yet to become an
integral part of the aerosol design process for metallic oxides. Consistent with this diversity in
aerosol synthesis, detailed experiments that measure multiple quantities with rigorously defined
flow configurations are missing. In the aerosol synthesis of fine particles, the use and validation
of computational models follow an approach driven by engineering objectives, with the tools
employed predominantly for the qualitative assessment of the flow and particle physics. Because
there is much less information on the basic precursor chemistry for almost all systems but titanium
tetrachloride, global models that approximately capture the location of precursor oxidation are
used. In addition, very few measurements of gas-phase or fine-particle properties are available
from experiments (Buesser & Pratsinis 2011, Pratsinis et al. 1996), such as the temperature or
fine-particle properties at a few select locations. In this low-data environment, the prudent engi-
neering approach is to use some basic measurements to calibrate the unknown parameters and then

180 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

Soot volume fraction, fv (ppb)

0 20 40 60 80 100
200
a b c x/DB = 0.576
150
10 1.0
Experimental data
Boundary conditions
100
8

50

〈fv 〉 (ppb)
6
x/DB

x/DB

0
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

200
0.5
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

x/DB = 1.008
4
150

2 100

50
0
0

0
–2 –1 0 1 2 0.6 0.4 0.2 0 0 0.1 0.2 0.3 0.4 0.5
r/DB r/DB r/DB

Figure 8
Large-eddy predictions for a sooting bluff-body ethylene flame. (a) Photograph of an ethylene bluff-body stabilized flame. (b) Time-
averaged soot volume fraction contours, with the magenta line showing the isoline of the stoichiometric composition. (c) Time-
averaged soot volume fraction profiles in the radial direction plotted for two different axial locations. The dark yellow squares denote
the experimental data. The blue and red lines correspond to two different boundary conditions used for the simulations, indicating the
strong sensitivity of the results to small changes in the numerical treatment of the flow. Figure adapted with permission from Mueller
et al. (2013).

to employ this calibrated model for further analysis or design. However, such calibrated models
can be trusted only in flow conditions close to the calibration conditions. In other words, the
models have an interpolative character, reliable only at flow conditions that lie in between calibra-
tion conditions, and cannot be trusted at other conditions without corroborating data. However,
despite this limitation, the detailed flow picture obtained from computational tools can be used to
understand the behavior of the system and to reduce the cost of experiment-based design.

5.3. Model Validation and Dependability: A Path Forward


The issue of validation and model dependability has received considerable attention in the past
decade, mainly driven by advances in uncertainty quantification (UQ) science (see, e.g., Babuška
et al. 2008, Braman et al. 2013, Frenklach 2007, Huan & Marzouk 2013, Najm 2009, Oberkampf UQ: uncertainty
& Trucano 2002). Here, in addition to the development of models, the characterization of the quantification
uncertainty introduced by the choice of model also becomes important. In the context of validation,
three essential components are needed: (a) the determination of the uncertainty associated with

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 181

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

each model and model parameters, (b) the determination of the models that affect the outcome of a
particular validation exercise, and (c) the relevance of the validation problem to the final operational
use of the models. As can be inferred, this is a multistage process that employs appropriate data
(experimental or otherwise) to build confidence in models at each stage of the modeling exercise.
The first issue of model uncertainty is handled through a probabilistic description of models
(see, e.g., Huan & Marzouk 2013, Prager et al. 2013). For example, model parameters are treated
as random variables with an associated PDF, which is inferred using experimental data. This ap-
proach has been used in the context of chemistry modeling quite widely (see, e.g., Braman et al.
2013, Frenklach 2007, Sheen & Wang 2011). Such a probabilistic description can be propagated
through a turbulent combustion or flow model to obtain a distribution of results. In this sense,
the uncertainty in model parameters is translated into uncertainty in the simulation predictions.
Currently, such an approach has been applied mainly to chemistry models and only for model
parameters. In reality, the model itself introduces an error, which is the main source of issues
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

in fine-particle descriptions. For instance, the use of the volume and surface area to parameter-
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

ize the complex morphology of fine particles, the use of particular condensation and oxidation
mechanisms, and the model for turbulence-chemistry interactions are all subject to errors. This
model-form error or model inadequacy description is not yet available for any practical problem,
but it is expected that in the future, modeling will refer not only to the description of a physical
process, but also to an estimate of the error introduced by the underlying assumptions.
The second issue concerns the importance of specific models in the simulation of a particular ex-
perimental configuration. One example is the bluff-body stabilized flame (Dally et al. 1998) used for
validating turbulent combustion models. Although this flame is geometrically complex and exhibits
intricate recirculation patterns that are unsteady, the combustion process itself is nominally in the
flamelet limit. Consequently, this flame provides a better test case for turbulent flow modeling with
heat release described by the flamelet approach. Unsurprisingly, the RANS approach was found
to be inaccurate for this flame [see the International Workshop on Measurement and Computa-
tion of Turbulent Nonpremixed Flames website (http://www.sandia.gov/TNF/abstract.html)],
whereas the LES approach with consistent boundary conditions reproduced the experimental data
accurately regardless of the combustion model used (Raman & Pitsch 2005, Raman et al. 2005).
Because the relative impact of models depends on the flow regime the experiment operates in, the
cause of failure or success of a simulation has to be assessed with regard to the relative impact of the
models. This requires the evaluation of sensitivities to target quantities. In complex configurations
that have a large number of parameters, forward sensitivity approaches are not tractable. Instead,
adjoint-based techniques are better suited (see, e.g., Braman et al. 2015). Extension of such tools
to LES and chaotic flow solvers will be of tremendous utility (see, e.g., Wang 2013).
The third issue of the relevance of experiments is the acceptable regime of usage for the model.
For instance, if the processes that dictate fine-particle formation in a high-pressure flame are
not important at low-pressure conditions, then validation experiments performed solely at low
pressures will not be sufficient. Similarly, the use of premixed flames alone to characterize soot
precursor and particle-phase models might be insufficient for use in aircraft combustors, where
considerable partial premixing and diffusion-like flame structures coexist with premixed flames
(see, e.g., Heye et al. 2013, Luo et al. 2011). Because laminar flames only access a limited range of
compositions that occur in a turbulent flame, the validity of the models for states encountered in
applications is questionable. Identifying laboratory-scale experiments that mimic end applications,
and the trajectories in composition space seen by the particles, is currently based on experience
and macroscopic physical arguments. However, combining UQ approaches with sensitivity tools
will provide a way forward in the rigorous design of experiments for fine-particle formation in
turbulent flames.

182 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

SUMMARY POINTS
1. Fine-particle formation in turbulent flames is heavily influenced by the flow physics.
2. The complex physical and chemical processes that lead to particle formation and growth
are modeled by a PBE.
3. Given its high dimensionality and integro-differential form, solving the PBE in turbulent
flow simulations is challenging and most often done using moment methods.
4. The interactions among turbulence, chemistry, and particle formation in turbulent flames
occur over a wide range of timescales and length scales, making such interactions more
difficult to model than in the absence of fine particles.
5. Soot particles in turbulent flames exhibit strong spatial and temporal intermittency, which
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

must be correctly accounted for to predict emissions.


Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

6. Metal-oxide particles can have very different physical properties depending on their time
history and spatial trajectories through the flame.
7. Choosing the appropriate level of flow modeling depends on many factors, such as the
availability of high-fidelity experimental data, the quality of the submodels appearing
in the PBE, and the ultimate end use of the simulations (e.g., testing novel precursor
formulations in a known burner versus predicting soot emission levels with a specified
degree of certainty).

FUTURE ISSUES
1. Further research is needed to improve the chemical/physical submodels for the nucle-
ation/growth/oxidation of fine particles in flames. Nonetheless, emphasis should be given
to improving submodels that strongly affect the predicted particle properties.
2. Further research is needed to understand and model turbulence-chemistry interactions
and differential diffusion in relation to fine-particle formation in turbulent flames.
3. Realizability-preserving numerical schemes are needed for all the terms appearing in the
moment transport equation. The computational cost for such schemes must be balanced
by their accuracy and robustness over a wide range of operating conditions.
4. Combining UQ approaches with sensitivity-analysis tools provides a promising path
toward the rigorous design and optimization of fine-particle production processes.

DISCLOSURE STATEMENT
The authors are not aware of any biases that might be perceived as affecting the objectivity of this
review.

ACKNOWLEDGMENTS
The financial support of the Strategic Environmental Research and Development Program
(SERDP WP-2151) and the National Science Foundation (CBET-0730369) is gratefully
acknowledged.

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 183

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

LITERATURE CITED
Akhtar MK, Pratsinis SE, Mastrangelo SVR. 1992. Dopants in vapor-phase synthesis of titania powders.
J. Am. Ceram. Soc. 75:3408–16
Akroyd J, Smith AJ, Shirley R, McGlashan LR, Kraft M. 2011. A coupled CFD-population balance approach
for nanoparticle synthesis in turbulent reacting flows. Chem. Eng. Sci. 66:3792–805
Appel J, Bockhorn H, Frenklach M. 2000. Kinetic modeling of soot formation with detailed chemistry and
physics: laminar premixed flames of C2 hydrocarbons. Combust. Flame 121:122–36
Arabi-Katbi OI, Pratsinis SE, Morrison PW Jr, Megaridis CM. 2001. Monitoring the flame synthesis of TiO2
particles by in-situ FTIR spectroscopy and thermophoretic sampling. Combust. Flame 124:560–72
Attili A, Bisetti F. 2013. Application of a robust and efficient Lagrangian particle scheme to soot transport in
turbulent flames. Comput. Fluids 84:164–75
Attili A, Bisetti F, Mueller ME, Pitsch H. 2014. Formation, growth, and transport of soot in a three-dimensional
turbulent non-premixed jet flame. Combust. Flame 161:1849–65
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

Attili A, Bisetti F, Mueller ME, Pitsch H. 2015. Damköhler number effects on soot formation and growth in
turbulent nonpremixed flames. Proc. Combust. Inst. 35:1215–23
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

Axelbaum RL, DuFaux DP, Frey CA, Kelton KF, Lawton SA, et al. 1996. Gas-phase combustion synthesis of
titanium boride (TiB2 ) nanocrystallites. J. Mater. Res. 11:948–54
Babuška I, Nobile F, Tempone R. 2008. A systematic approach to model validation based on Bayesian updates
and prediction related rejection criteria. Comput. Methods Appl. Mech. Eng. 197:2517–39
Balachandar S, Eaton JK. 2010. Turbulent dispersed multiphase flow. Annu. Rev. Fluid Mech. 42:111–33
Bhatt J, Lindstedt R. 2009. Analysis of the impact of agglomeration and surface chemistry models on soot
formation and oxidation. Proc. Combust. Inst. 32:713–20
Bisetti F, Attili A, Pitsch H. 2014. Advancing predictive models for particulate formation in turbulent flames
via massively parallel direct numerical simulations. Philos. Trans. R. Soc. A 372:20130324
Bisetti F, Blanquart G, Mueller ME, Pitsch H. 2012. On the formation and early evolution of soot in turbulent
nonpremixed flames. Combust. Flame 159:317–35
Blacha T, Di Domenico M, Rachner M, Gerlinger P, Aigner M. 2011. Modeling of soot and NOx in a full
scale turbine engine combustor with detailed chemistry. In ASME 2011 Turbo Expo, Vol. 2: Combustion,
Fuels and Emissions, Parts A and B, pp. 33–42. New York: ASME
Blanquart G, Pepiot-Desjardins P, Pitsch H. 2009. Chemical mechanism for high temperature combustion
of engine relevant fuels with emphasis on soot precursors. Combust. Flame 156:588–607
Blanquart G, Pitsch H. 2009. Analyzing the effects of temperature on soot formation with a joint volume-
surface-hydrogen model. Combust. Flame 156:1614–26
Böhm H, Kohse-Höinghaus K, Lacas F, Rolon C, Darabiha N, Candel S. 2001. On PAH formation in strained
counterflow diffusion flames. Combust. Flame 124:127–36
Bolla M, Farrace D, Wright YM, Boulouchos K, Mastorakos E. 2014. Influence of turbulence–chemistry in-
teraction for n-heptane spray combustion under diesel engine conditions with emphasis on soot formation
and oxidation. Combust. Theory Model. 18:330–60
Bose S, Moin P, You D. 2010. Grid-independent large-eddy simulation using explicit filtering. Phys. Fluids
22:105103
Braman K, Oliver TA, Raman V. 2013. Bayesian analysis of syngas chemistry models. Combust. Theory Model.
17:858–87
Braman K, Oliver TA, Raman V. 2015. Adjoint-based sensitivity analysis of flames. Combust. Theory Model.
19:29–56
Buesser B, Gröhn AJ. 2012. Multiscale aspects of modeling gas-phase nanoparticle synthesis. Chem. Eng.
Technol. 35:1133–43
Buesser B, Pratsinis SE. 2011. Design of gas-phase synthesis of core-shell particles by computational fluid-
aerosol dynamics. AIChE J. 57:3132–42
Buesser B, Pratsinis SE. 2012. Design of nanomaterial synthesis by aerosol processes. Annu. Rev. Chem. Biomol.
Eng. 3:103–27
Calcote H, Felder W, Keil D, Olson D. 1991. A new flame process for synthesis of Si3 N4 powders for advanced
ceramics. Symp. (Int.) Combust. 23:1739–44

184 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

Chan QN, Medwell PR, Kalt PA, Alwahabi ZT, Dally BB, Nathan GJ. 2011. Simultaneous imaging of
temperature and soot volume fraction. Proc. Combust. Inst. 33:791–98
Chatterjee S, Halmo C, Gülder ÖL. 2014. Structure of the velocity and soot concentration fields of a swirl
stabilized turbulent non-premixed flame in a gas turbine model combustor. In ASME 2014 Gas Turbine
India Conference, Pap. GTINDIA2014-8114. New York: ASME
Cheng JC, Vigil RD, Fox RO. 2010. A competitive aggregation model for Flash NanoPrecipitation. J. Colloid
Interface Sci. 351:330–42
Chittipotula T, Janiga G, Thévenin D. 2011. Improved soot prediction models for turbulent non-premixed
ethylene/air flames. Proc. Combust. Inst. 33:559–67
Chittipotula T, Janiga G, Thévenin D. 2012. Optimizing soot prediction models for turbulent non-premixed
ethylene/air flames. Chem. Eng. Sci. 70:67–76
Colucci PJ, Jaberi FA, Givi P. 1998. Filtered density function for large eddy simulation of turbulent reacting
flows. Phys. Fluids 10:499–515
Coppalle A, Joyeux D. 1994. Temperature and soot volume fraction in turbulent diffusion flames: measure-
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

ments of mean and fluctuating values. Combust. Flame 96:275–85


Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

Dally BB, Fletcher DF, Masri AR. 1998. Flow and mixing fields of turbulent bluff-body jets and flames.
Combust. Theory Model. 2:193–219
D’Anna A. 2009. Combustion-formed nanoparticles. Proc. Combust. Inst. 32:593–613
Desgroux P, Mercier X, Thomson KA. 2013. Study of the formation of soot and its precursors in flames using
optical diagnostics. Proc. Combust. Inst. 34:1713–38
Dette H, Studden WJ. 1997. The Theory of Canonical Moments with Applications in Statistics, Probability, and
Analysis. New York: Wiley
Donde P, Koo H, Raman V. 2012. A multivariate quadrature based moment method for LES based modeling
of supersonic combustion. J. Comput. Phys. 231:5805–21
Donde P, Raman V, Mueller ME, Pitsch H. 2013. LES/PDF based modeling of soot-turbulence interactions
in turbulent flames. Proc. Combust. Inst. 34:1183–92
Duclos J, Veynante D, Poinsot T. 1993. A comparison of flamelet models for premixed turbulent combustion.
Combust. Flame 95:101–17
Einstein A. 1905. Über die von der molekularkinetischen Theorie der Wärme geforderte Bewegung von in
ruhenden Flüssigkeiten suspendierten Teilchen. Ann. Phys. 17:549–60
Elasrag H, Lu T, Law C, Menon S. 2007. Simulation of soot formation in turbulent premixed flames. Combust.
Flame 150:108–26
Elasrag H, Menon S. 2009. Large eddy simulation of soot formation in a turbulent non-premixed jet flame.
Combust. Flame 156:385–95
Fiorina B, Vicquelin R, Auzillon P, Darabiha N, Gicquel O, Veynante D. 2010. A filtered tabulated chemistry
model for LES of premixed combustion. Combust. Flame 157:465–75
Fox RO. 1999. The Lagrangian spectral relaxation model for differential diffusion in homogeneous turbulence.
Phys. Fluids 11:1550–71
Fox RO. 2003. Computational Models for Turbulent Reacting Flows. Cambridge, UK: Cambridge Univ. Press
Fox RO. 2012. Large-eddy-simulation tools for multiphase flows. Annu. Rev. Fluid Mech. 44:47–76
Fox RO, Raman V. 2004. A multienvironment conditional probability density function model for turbulent
reacting flows. Phys. Fluids 16:4551–65
Franzelli B, Scouflaire P, Candel S. 2015. Time-resolved spatial patterns and interactions of soot, PAH and
OH in a turbulent diffusion flame. Proc. Combust. Inst. 35:1921–29
Frenklach M. 2002a. Method of moments with interpolative closure. Chem. Eng. Sci. 57:2229–39
Frenklach M. 2002b. Reaction mechanism of soot formation in flames. Phys. Chem. Chem. Phys. 4:2028–37
Frenklach M. 2007. Transforming data into knowledge: process informatics for combustion chemistry. Proc.
Combust. Inst. 31:125–40
Frenklach M, Harris SJ. 1987. Aerosol dynamics modeling using the method of moments. J. Colloid Interface
Sci. 118:252–61
Friedlander SK. 2000. Smoke, Dust, and Haze: Fundamentals of Aerosol Dynamics. New York: Oxford Univ.
Press

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 185

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

Fuchs N. 1964. Mechanics of Aerosols. New York: Pergamon


Garrick SC, Wang G. 2010. Modeling and simulation of titanium dioxide nanoparticle synthesis with finite-
rate sintering in planar jets. J. Nanopart. Res. 13:973–84
Geigle KP, Hadef R, Meier W. 2013. Soot formation and flame characterization of an aero-engine model
combustor burning ethylene at elevated pressure. J. Eng. Gas Turbines Power 136:021505
Geigle KP, Köhler M, O’Loughlin W, Meier W. 2015. Investigation of soot formation in pressurized swirl
flames by laser measurements of temperature, flame structures and soot concentrations. Proc. Combust.
Inst. 35:3373–80
Ghoshtagore RN. 1970. Mechanism of heterogeneous deposition of thin film rutile. J. Electrochem. Soc.
117:529–34
Gicquel O, Darabiha N, Thévenin D. 2000. Liminar premixed hydrogen/air counterflow flame simulations
using flame prolongation of ILDM with differential diffusion. Proc. Combust. Inst. 28:1901–8
Glassman I. 1989. Soot formation in combustion processes. Symp. (Int.) Combust. 22:295–311
Glassman I, Davis KA, Brezinsky K. 1992. A gas-phase combustion synthesis process for non-oxide ceramics.
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

Symp. (Int.) Combust. 24:1877–82


Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

Gröhn AJ, Buesser B, Jokiniemi JK, Pratsinis SE. 2011. Design of turbulent flame aerosol reactors by mixing-
limited fluid dynamics. Ind. Eng. Chem. Res. 50:3159–68
Gullbrand J, Chow F. 2003. The effect of numerical errors and turbulence models in large-eddy simulations
of channel flow, with and without explicit filtering. J. Fluid Mech. 495:323–41
Hawkes E, Cant R. 2000. A flame surface density approach to large-eddy simulation of premixed turbulent
combustion. Proc. Combust. Inst. 28:51–58
Haworth D. 2010. Progress in probability density function methods for turbulent reacting flows. Prog. Energy
Combust. Sci. 36:168–259
Haynes B, Wagner H. 1981. Soot formation. Prog. Energy Combust. Sci. 7:229–73
Heine MC, Mädler L, Jossen R, Pratsinis SE. 2006. Direct measurement of entrainment during nanoparticle
synthesis in spray flames. Combust. Flame 144:809–20
Henriksen TL, Nathan GJ, Alwahabi ZT, Qamar N, Ring TA, Eddings EG. 2009. Planar measurements of
soot volume fraction and OH in a JP-8 pool fire. Combust. Flame 156:1480–92
Herrmann M, Blanquart G, Raman V. 2006. A bounded QUICK scheme for preserving scalar bounds in
large-eddy simulations. AIAA J. 44:2879–80
Heye C, Raman V, Masri AR. 2013. LES/probability density function approach for the simulation of an
ethanol spray flame. Proc. Combust. Inst. 34:1633–41
Hu B, Yang B, Koylu UO. 2003. Soot measurements at the axis of an ethylene/air non-premixed turbulent jet
flame. Combust. Flame 134:93–106
Hu Y, Ding H, Li C. 2011. Preparation of hollow alumina nanospheres via surfactant-assisted flame spray
pyrolysis. Particuology 9:528–32
Huan X, Marzouk YM. 2013. Simulation-based optimal Bayesian experimental design for nonlinear systems.
J. Comput. Phys. 232:288–317
Ifeacho P, Wiggers H, Schulz C, Schneider L, Bacher G. 2007. Ga2 O3 nanoparticles synthesized in a low-
pressure flame reactor. J. Nanopart. Res. 10:121–27
Jiang GS, Peng D. 2000. Weighted ENO schemes for Hamilton-Jacobi equations. SIAM J. Sci. Comput.
21:2126–43
Johannessen T, Pratsinis SE, Livbjerg H. 2000. Computational fluid-particle dynamics for the flame synthesis
of alumina particles. Chem. Eng. Sci. 55:177–91
Johannessen T, Pratsinis SE, Livbjerg H. 2001. Computational analysis of coagulation and coalescence in the
flame synthesis of titania particles. Powder Technol. 118:242–50
Jossen R, Mueller R, Pratsinis SE, Watson M, Akhtar MK. 2005. Morphology and composition of spray-
flame-made yttria-stabilized zirconia nanoparticles. Nanotechnology 16:S609–17
Karataş AE, Gülder ÖL. 2012. Soot formation in high pressure laminar diffusion flames. Prog. Energy Combust.
Sci. 38:818–45
Kaul CM, Raman V. 2011. A posteriori analysis of numerical errors in subfilter scalar variance modeling for
large eddy simulation. Phys. Fluids 23:035102

186 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

Kaul CM, Raman V, Balarac G, Pitsch H. 2009. Numerical errors in the computation of subfilter scalar
variance in large eddy simulations. Phys. Fluids 21:055102
Kennedy IM. 1997. Models of soot formation and oxidation. Prog. Energy Combust. Sci. 23:95–132
Kent JH, Honnery D. 1987. Soot and mixture fraction in turbulent diffusion flames. Combust. Sci. Technol.
54:383–98
Kholghy M, Saffaripour M, Yip C, Thomson MJ. 2013. The evolution of soot morphology in a laminar coflow
diffusion flame of a surrogate for jet A-1. Combust. Flame 160:2119–30
Klimenko AY, Bilger RW. 1999. Conditional moment closure for turbulent combustion. Prog. Energy Combust.
Sci. 25:595–687
Kobata A, Kusakabe K, Morooka S. 1991. Growth and transformation of TiO2 crystallites in aerosol reactor.
AIChE J. 37:347–59
Köhler M, Geigle KP, Blacha T, Gerlinger P, Meier W. 2012. Experimental characterization and numerical
simulation of a sooting lifted turbulent jet diffusion flame. Combust. Flame 159:2620–35
Koo H, Raman V, Mueller ME, Geigle KP. 2015. Large-eddy simulation of a turbulent sooting flame in a swirling
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

combustor. Presented at AIAA Aerospace Sci. Meet., 53rd, Kissimmee, FL, AIAA Pap. 2015-0167
Kraft M. 2005. Modelling of particulate processes. KONA 23:18–35
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

Kronenburg A, Bilger R, Kent J. 2000. Modeling soot formation in turbulent methane–air jet diffusion flames.
Combust. Flame 121:24–40
Langford JA, Moser RD. 1999. Optimal LES formulations for isotropic turbulence. J. Fluid Mech. 398:321–46
Lecocq G, Poitou D, Hernández I, Duchaine F, Riber E, Cuenot B. 2014. A methodology for soot prediction
including thermal radiation in complex industrial burners. Flow Turbul. Combust. 92:947–70
Leonard BP. 1979. Stable and accurate convective modeling procedure based on quadratic upstream interpo-
lation. Comput. Methods Appl. Mech. Eng. 19:59–98
Leung K, Lindstedt R, Jones W. 1991. A simplified reaction mechanism for soot formation in nonpremixed
flames. Combust. Flame 87:289–305
Lignell DO, Chen JH, Smith PJ. 2008. Three-dimensional direct numerical simulation of soot formation and
transport in a temporally evolving nonpremixed ethylene jet flame. Combust. Flame 155:316–33
Lignell DO, Chen JH, Smith PJ, Lu T, Law CK. 2007. The effect of flame structure on soot formation and
transport in turbulent nonpremixed flames using direct numerical simulation. Combust. Flame 151:2–28
Lindstedt R, Louloudi S. 2005. Joint-scalar transported PDF modeling of soot formation and oxidation. Proc.
Combust. Inst. 30:775–83
Liu J, Hu Y, Gu F, Li C. 2011. Large-scale synthesis of hollow titania spheres via flame combustion. Particuology
9:632–36
Loeffler J, Das S, Garrick SC. 2011. Large eddy simulation of titanium dioxide nanoparticle formation and
growth in turbulent jets. Aerosol Sci. Technol. 45:616–28
Luo K, Pitsch H, Pai M, Desjardins O. 2011. Direct numerical simulations and analysis of three-dimensional
n-heptane spray flames in a model swirl combustor. Proc. Combust. Inst. 33:2143–52
Ma G, Wen JZ, Lightstone MF, Thomson MJ. 2005. Optimization of soot modeling in turbulent nonpremixed
ethylene/air jet flames. Combust. Sci. Technol. 177:1567–602
Mädler L, Kammler H, Mueller R, Pratsinis S. 2002. Controlled synthesis of nanostructured particles by flame
spray pyrolysis. J. Aerosol Sci. 33:369–89
Mahmoud S, Nathan G, Medwell P, Dally B, Alwahabi Z. 2015. Simultaneous planar measurements of tem-
perature and soot volume fraction in a turbulent non-premixed jet flame. Proc. Combust. Inst. 35:1931–38
Marchisio DL, Fox RO. 2005. Solution of population balance equations using the direct quadrature method
of moments. J. Aerosol Sci. 36:43–73
Marchisio DL, Fox RO. 2013. Computational Models for Polydisperse Particulate and Multiphase Systems.
Cambridge, UK: Cambridge Univ. Press
Marchisio DL, Fox RO, Barresi AA. 2001. Simulation of turbulent precipitation in a semi-batch Taylor–
Couette reactor using computational fluid dynamics. AIChE J. 47:664–76
Marchisio DL, Pikturna JT, Fox RO, Vigil RD, Barresi AA. 2003. Quadrature method of moments for
population balances with nucleation, growth and aggregation. AIChE J. 49:1266–76
McDermott R, Pope SB. 2007. A particle formulation for treating differential diffusion in filtered density
function methods. J. Comput. Phys. 226:947–93

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 187

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

McGraw R. 2007. Numerical advection of correlated tracers: preserving particle size/composition moment
sequences during transport of aerosol mixtures. J. Phys. Conf. Ser. 78:012045
Menz WJ, Brownbridge GP, Kraft M. 2014. Global sensitivity analysis of a model for silicon nanoparticle
synthesis. J. Aerosol Sci. 76:188–99
Mueller ME, Blanquart G, Pitsch H. 2009. Hybrid method of moments for modeling soot formation and
growth. Combust. Flame 156:1143–55
Mueller ME, Chan QN, Qamar NH, Dally BB, Pitsch H, et al. 2013. Experimental and computational study
of soot evolution in a turbulent nonpremixed bluff body ethylene flame. Combust. Flame 160:1298–309
Mueller ME, Pitsch H. 2011. Large eddy simulation subfilter modeling of soot-turbulence interactions. Phys.
Fluids 23:115104
Mueller ME, Pitsch H. 2012. LES model for sooting turbulent nonpremixed flames. Combust. Flame 159:2166–
80
Mueller ME, Pitsch H. 2013. Large eddy simulation of soot evolution in an aircraft combustor. Phys. Fluids
25:110812
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

Mueller ME, Raman V. 2014. Effects of turbulent combustion modeling errors on soot evolution in a turbulent
nonpremixed jet flame. Combust. Flame 161:1842–48
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

Najm HN. 2009. Uncertainty quantification and polynomial chaos techniques in computational fluid dynamics.
Annu. Rev. Fluid Mech. 41:35–52
Narayanaswamy V, Clemens N. 2013. Simultaneous LII and PIV measurements in the soot formation region
of turbulent non-premixed jet flames. Proc. Combust. Inst. 34:1455–63
Oberkampf WL, Trucano TG. 2002. Verification and validation in computational fluid dynamics. Prog.
Aerospace Sci. 38:209–72
Peters N. 2000. Turbulent Combustion. Cambridge, UK: Cambridge Univ. Press
Pierce CD, Moin P. 2004. Progress-variable approach for large-eddy simulation of non-premixed turbulent
combustion. J. Fluid Mech. 504:73–97
Piton D, Fox RO, Marcant B. 2000. Simulation of fine particle formation by precipitation using computational
fluid dynamics. Can. J. Chem. Eng. 78:983–93
Pitsch H. 2006. Large-eddy simulation of turbulent combustion. Annu. Rev. Fluid Mech. 38:453–82
Pitsch H, Peters N. 1998. A consistent flamelet formulation for non-premixed combustion considering dif-
ferential diffusion effects. Combust. Flame 144:26–40
Pitsch H, Steiner H. 2000. Large-eddy simulation of a turbulent piloted methane/air diffusion flame (Sandia
flame D). Phys. Fluids 12:2541–54
Poinsot T, Veynante D. 2001. Theoretical and Numerical Combustion. Philadelphia: R.T. Edwards
Pope SB. 1981. A Monte-Carlo method for the PDF equations of turbulent reactive flow. Combust. Sci. Technol.
25:159–74
Pope SB. 1985. PDF methods for turbulent reactive flows. Prog. Energy Combust. Sci. 11:119–92
Pope SB. 2000. Turbulent Flows. Cambridge, UK: Cambridge Univ. Press
Pope SB. 2010. Self-conditioned fields for large-eddy simulations of turbulent flows. J. Fluid Mech. 652:139–69
Prager J, Najm HN, Sargsyan K, Safta C, Pitz WJ. 2013. Uncertainty quantification of reaction mechanisms
accounting for correlations introduced by rate rules and fitted Arrhenius parameters. Combust. Flame
160:1583–93
Pratsinis SE. 1998. Flame aerosol synthesis of ceramic powders. Prog. Energy Combust. Sci. 24:197–219
Pratsinis SE, Bai H, Biswas P, Frenklach M, Mastrangelo SVR. 1990. Kinetics of titanium(IV) chloride
oxidation. J. Am. Ceram. Soc. 73:2158–62
Pratsinis SE, Vemury S. 1996. Particle formation in gases: a review. Powder Technol. 88:267–73
Pratsinis SE, Zhu W, Vemury S. 1996. The role of gas mixing in flame synthesis of titania powders. Powder
Technol. 86:87–93
Qamar N, Alwahabi Z, Chan Q, Nathan G, Roekaerts D, King K. 2009. Soot volume fraction in a piloted
turbulent jet non-premixed flame of natural gas. Combust. Flame 156:1339–47
Raman V, Fox RO, Harvey AD. 2004. Hybrid finite-volume/transported PDF simulations of a partially
premixed methane-air flame. Combust. Flame 136:327–50
Raman V, Pitsch H. 2005. Large-eddy simulation of bluff-body stabilized non-premixed flame using a
recursive-refinement procedure. Combust. Flame 142:329–47

188 Raman · Fox

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

Raman V, Pitsch H. 2006. A consistent LES/filtered-density function formulation for the simulation of tur-
bulent flames with detailed chemistry. Proc. Combust. Inst. 31:1711–19
Raman V, Pitsch H, Fox RO. 2005. A consistent hybrid LES-FDF scheme for the simulation of turbulent
reactive flows. Combust. Flame 143:56–78
Raman V, Pitsch H, Fox RO. 2006. Eulerian transported probability density function sub-filter model for
large-eddy simulation of turbulent combustion. Combust. Theory Model. 10:439–58
Ramkrishna D. 2000. Population Balances. San Diego: Academic
Reade WC, Collins LR. 2000. Effect of preferential concentration on turbulent collision rates. Phys. Fluids
12:2530–40
Reddy M, De A, Yadav R. 2015. Effect of precursors and radiation on soot formation in turbulent diffusion
flame. Fuel 148:58–72
Reveillon J, Vervisch L. 2005. Analysis of weakly turbulent dilute-spray flames and spray combustion regimes.
J. Fluid Mech. 537:317–47
Richter H, Howard J. 2000. Formation of polycyclic aromatic hydrocarbons and their growth to soot: a review
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

of chemical reaction pathways. Prog. Energy Combust. Sci. 26:565–608


Rosner DE. 2005. Flame synthesis of valuable nanoparticles: recent progress/current needs in areas of rate
Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

laws, population dynamics, and characterization. Ind. Eng. Chem. Res. 44:6045–55
Roth P. 2007. Particle synthesis in flames. Proc. Combust. Inst. 31:1773–88
Saggese C, Sánchez NE, Frassoldati A, Cuoci A, Faravelli T, et al. 2014. Kinetic modeling study of polycyclic
aromatic hydrocarbons and soot formation in acetylene pyrolysis. Energy Fuels 28:1489–501
Said R, Garo A, Borghi R. 1997. Soot formation modeling for turbulent flames. Combust. Flame 108:71–86
Sheen DA, Wang H. 2011. The method of uncertainty quantification and minimization using polynomial
chaos expansions. Combust. Flame 158:2358–74
Shirley R, Akroyd J, Miller LA, Inderwildi OR, Riedel U, Kraft M. 2011. Theoretical insights into the surface
growth of rutile TiO2 . Combust. Flame 158:1868–76
Shohat JA, Tamarkin JD. 1943. The Problem of Moments, Vol. 1. Providence, RI: Am. Math. Soc.
Singh R, Raman V. 2012. Two-dimensional direct numerical simulation of nanoparticle precursor evolution
in turbulent flames using detailed chemistry. Chem. Eng. J. 207–208:794–802
Stanmore BR, Brilhac JF, Gilot P. 2001. The oxidation of soot: a review of experiments, mechanisms and
models. Carbon 39:2247–68
Steiner H, Bushe WK. 2001. Large eddy simulation of a turbulent reacting jet with conditional source-term
estimation. Phys. Fluids 13:754–69
Sundaram S, Collins LR. 1997. Collision statistics in an isotropic particle-laden turbulent suspension. 1. Direct
numerical simulations. J. Fluid Mech. 335:75–109
Sung Y, Raman V, Fox RO. 2011. Large-eddy-simulation-based multiscale modeling of TiO2 nanoparticle
synthesis in a turbulent flame reactor using detailed nucleation chemistry. Chem. Eng. Sci. 66:4370–81
Sung Y, Raman V, Koo H, Mehta M, Fox RO. 2013. Large-eddy simulation modeling of turbulent flame
synthesis of titania nanoparticles using a bivariate particle description. AIChE J. 60:459–72
Talbot L. 1981. Thermophoresis: a review. In Rarefied Gas Dynamics: Int. Symp., 12th, Charlottesville, VA, July
7–11, 1980, Tech. Pap. Part 1, pp. 467–88. New York: AIAA
van Oijen JA, de Goey LPH. 2000. Modelling of premixed laminar flames using flamelet-generated manifolds.
Combust. Sci. Technol. 161:113–37
van Oijen JA, Lammers FA, de Goey LPH. 2001. Modeling of complex premixed burner systems by using
flamelet-generated manifolds. Combust. Flame 127:2124–34
Vasilyev O, Lund T, Moin P. 1998. A general class of commutative filters for LES in complex geometries.
J. Comput. Phys. 146:82–104
Vemury S, Pratsinis SE. 1995. Dopants in flame synthesis of titania. J. Am. Ceram. Soc. 78:2984–92
Vikas V, Wang ZJ, Passalacqua A, Fox RO. 2011. Realizable high-order finite-volume schemes for quadrature-
based moment methods. J. Comput. Phys. 230:5328–52
Wang CS, Friedlander SK, Mädler L. 2005. Nanoparticle aerosol science and technology: an overview. China
Particuol. 3:243–54
Wang G, Garrick SC. 2005. Modeling and simulation of titania synthesis in two-dimensional methane–air
flames. J. Nanopart. Res. 7:621–32

www.annualreviews.org • Modeling of Fine-Particle Formation in Turbulent Flames 189

Changes may still occur before final publication online and in print
FL48CH07-Raman ARI 11 August 2015 15:22

Wang H. 2011. Formation of nascent soot and other condensed-phase materials in flames. Proc. Combust. Inst.
33:41–67
Wang H, Frenklach M. 1997. A detailed kinetic modeling study of aromatics formation in laminar premixed
acetylene and ethylene flames. Combust. Flame 110:173–221
Wang Q. 2013. Forward and adjoint sensitivity computation of chaotic dynamical systems. J. Comput. Phys.
235:1–13
Wen Z, Yun S, Thomson M, Lightstone M. 2003. Modeling soot formation in turbulent kerosene/air jet
diffusion flames. Combust. Flame 135:323–40
West RH, Shirley RA, Kraft M, Goldsmith CF, Green WH. 2009. A detailed kinetic model for combustion
synthesis of titania from TiCl4 . Combust. Flame 156:1764–70
Woolley RM, Fairweather M, Yunardi. 2009. Conditional moment closure modelling of soot formation in
turbulent, non-premixed methane and propane flames. Fuel 88:393–407
Wright DL. 2007. Numerical advection of moments of the particle size distribution in Eulerian models.
J. Aerosol Sci. 38:352–69
Access provided by Stockholm University - Library on 08/24/15. For personal use only.

Xuan Y, Blanquart G. 2015. Effects of aromatic chemistry-turbulence interactions on soot formation in a


Annu. Rev. Fluid Mech. 2016.48. Downloaded from www.annualreviews.org

turbulent non-premixed flame. Proc. Combust. Inst. 35:1911–19


Yeh C, Yeh S, Ma H. 2004. Flame synthesis of titania particles from titanium tetraisopropoxide in premixed
flames. Powder Technol. 145:1–9
Yoo CS, Im HG. 2007. Transient soot dynamics in turbulent nonpremixed ethylene–air counterflow flames.
Proc. Combust. Inst. 31:701–8
Yu M, Lin J. 2010. Nanoparticle-laden flows via moment method: a review. Int. J. Multiphase Flow 36:144–51
Yuan C, Laurent F, Fox RO. 2012. An extended quadrature method of moments for population balance
equations. J. Aerosol Sci. 51:1–23
Zhang J, Shaddix CR, Schefer RW. 2011. Design of “model-friendly” turbulent non-premixed jet burners for
C2 + hydrocarbon fuels. Rev. Sci. Instrum. 82:074101
Zhong BJ, Dang S, Song YN, Gong JS. 2012. 3-D simulation of soot formation in a direct-injection diesel
engine based on a comprehensive chemical mechanism and method of moments. Combust. Theory Model.
16:143–71
Zucca A, Marchisio DL, Barresi AA, Fox RO. 2006. Implementation of the population balance equation in
CFD codes for modelling soot formation in turbulent flames. Chem. Eng. Sci. 61:87–95

190 Raman · Fox

Changes may still occur before final publication online and in print

You might also like