176
Numerical Modelling of Soot Formation Robert Iain Arthur Patterson Trinity College A dissertation submitted for the degree of Doctor of Philosophy at the University of Cambridge April 2007

Numerical Modelling of Soot Formation

Embed Size (px)

Citation preview

Page 1: Numerical Modelling of Soot Formation

Numerical Modelling of SootFormation

Robert Iain Arthur Patterson

Trinity College

A dissertation submitted for the degree of Doctor of Philosophy at the

University of Cambridge

April 2007

Page 2: Numerical Modelling of Soot Formation

Preface

This dissertation is the result of my own work and includes nothing which is theoutcome of work done in collaboration, except where specifically indicated in thetext. The work presented was undertaken at the Department of Chemical Engi-neering, University of Cambridge, between October 2003 and March 2007. Chap-ters 1, 2 & 3 of this dissertation include work from the dissertation I submitted inJune 2004 for a Certificate of Postgraduate Study. No other part of this thesishas been submitted for a degree to this or any other university. This dissertationcontains approximately 35,000 words and 27 figures.

Some of the work in this dissertation has been published:

1. R. I. A. Patterson, J. Singh, M. Balthasar, M. Kraft, and J. R. Norris, “TheLinear Process Deferment Algorithm: A new technique for solving popula-tion balance equations”, SIAM Journal on Scientific Computing, 28, 303-320, (2006).

2. R. I. A. Patterson, J. Singh, M. Balthasar, M. Kraft, and W. Wagner, “Ex-tending stochastic soot simulation to higher pressures”, Combustion andFlame, 145, 638-642, (2006).

3. R. I. A. Patterson, and M. Kraft, “A simple model for the aggregate structureof soot particles”, Combustion and Flame, in press (2007).

R I A Patterson

11 April 2007 (approved corrections 2 June 2007)

1

Page 3: Numerical Modelling of Soot Formation

Summary

This thesis presents developments and applications of Monte Carlo algorithms forthe simulation of soot formation in premixed laminar flames. The thesis beginswith a brief review of experimental data on soot formation, models for the growthof soot and possible simulation techniques. The introduction concludes with amathematical formulation of the soot growth problems that the thesis addresses.

The development of a new algorithm, to incorporate a model of soot particlecoagulation at intermediate pressures into Monte Carlo simulations, marks thebeginning of the numerical work. Simulations using this algorithm are validatedin detail and the method thus established forms the base for the calculations in themajority of the thesis.

Investigations of the performance of the Monte Carlo simulations are thendescribed. To address the problems highlighted by these investigations, a newprobabilistic method for the simulation of chemical reactions on the surface ofsoot particles is introduced and tested. It is shown to reduce computational timesby a factor of up to a thousand, so that high quality calculations can be performedin very few minutes. A range of deterministic alternatives to this method are alsoinvestigated.

The power and flexibility of the accelerated algorithm is then used for modeldevelopment. Simple new models for the aggregate structure of soot particles areanalysed, a range of different flames are considered and it is found that numericalresults are only mildly sensitive to the details of particle shape models.

Finally, a new weighted particle Monte Carlo simulation algorithm is derived.This generalises ideas from work by other authors, which have been found tooffer computational savings and increased precision. The algorithm offers extraflexibility for use in problems with particle inflow and outflow and allows forcomputational effort to be concentrated on important classes of particles. It isapplied to a range of problems considered earlier in the thesis and is found to beof similar efficiency to the fully developed version of the original Monte Carloalgorithm.

2

Page 4: Numerical Modelling of Soot Formation

Contents

1 Introduction 71.1 Experimental Investigations of Soot . . . . . . . . . . . . . . . . 8

1.1.1 Collection and Examination of Soot Particles . . . . . . . 9

1.1.2 In-Situ Light Scattering and Absorption . . . . . . . . . . 10

1.1.3 Other scattering . . . . . . . . . . . . . . . . . . . . . . . 12

1.1.4 On-line Particle Sampling . . . . . . . . . . . . . . . . . 12

1.2 Combustion Chemistry . . . . . . . . . . . . . . . . . . . . . . . 14

1.2.1 Polyaromatic Hydrocarbon Growth . . . . . . . . . . . . 15

1.3 Population Balance Modelling . . . . . . . . . . . . . . . . . . . 17

1.3.1 Explicit Discretisation of Type Space . . . . . . . . . . . 18

1.3.2 Integral Methods . . . . . . . . . . . . . . . . . . . . . . 25

1.3.3 Stochastic Particle Methods . . . . . . . . . . . . . . . . 28

1.4 Mathematical Framework . . . . . . . . . . . . . . . . . . . . . . 30

1.4.1 Models for individual soot particles . . . . . . . . . . . . 32

2 Simulation at Higher Pressures 352.1 Coagulation Kernel . . . . . . . . . . . . . . . . . . . . . . . . . 36

2.1.1 Majorant Kernels . . . . . . . . . . . . . . . . . . . . . . 37

2.1.2 Validation . . . . . . . . . . . . . . . . . . . . . . . . . . 40

2.2 Application to Flames . . . . . . . . . . . . . . . . . . . . . . . . 42

2.3 Implementation of DSA . . . . . . . . . . . . . . . . . . . . . . . 45

2.3.1 Variation in Number of Particles . . . . . . . . . . . . . . 45

2.3.2 Binary Tree . . . . . . . . . . . . . . . . . . . . . . . . . 46

3

Page 5: Numerical Modelling of Soot Formation

2.3.3 Complexity of the implementation . . . . . . . . . . . . . 49

2.3.4 DSA Applied to Flames . . . . . . . . . . . . . . . . . . 50

2.3.5 Profiling of DSA Simulation . . . . . . . . . . . . . . . . 51

3 Accelerating the Surface Processes 553.1 Operator Splitting . . . . . . . . . . . . . . . . . . . . . . . . . . 55

3.1.1 Mathematical Outline of Splitting . . . . . . . . . . . . . 56

3.1.2 Implementation of Operator Splitting . . . . . . . . . . . 57

3.2 Numerical Results with Splitting . . . . . . . . . . . . . . . . . . 59

3.2.1 Comparison of JW1.69 Moments . . . . . . . . . . . . . 59

3.2.2 Particle Distribution Accuracy . . . . . . . . . . . . . . . 62

3.2.3 Profiling of Operator Splitting . . . . . . . . . . . . . . . 63

3.2.4 Conclusions on Operator Splitting . . . . . . . . . . . . . 65

3.3 Deferment of Surface Reactions . . . . . . . . . . . . . . . . . . 65

3.3.1 Measure Theoretic Formulation . . . . . . . . . . . . . . 66

3.3.2 Deferred Surface Process Operator . . . . . . . . . . . . . 67

3.3.3 Rate Kernel . . . . . . . . . . . . . . . . . . . . . . . . . 70

3.3.4 Implementation of Deferment . . . . . . . . . . . . . . . 72

3.3.5 Numerical Results . . . . . . . . . . . . . . . . . . . . . 73

3.3.6 Conclusions on LPDA . . . . . . . . . . . . . . . . . . . 75

3.4 Comparison of Simulation Methods . . . . . . . . . . . . . . . . 75

3.4.1 Equal Numbers of Computational Particles . . . . . . . . 75

3.4.2 Equal Precision for JW1.69 . . . . . . . . . . . . . . . . 77

3.4.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . 78

4 Deterministic Simulation of the Surface Processes 794.1 Operator Splitting . . . . . . . . . . . . . . . . . . . . . . . . . . 80

4.1.1 Modified Choice of Top Level Time Step . . . . . . . . . 80

4.1.2 Elimination of second level of splitting . . . . . . . . . . 81

4.1.3 Initial Trial of Deterministic Method . . . . . . . . . . . . 81

4.1.4 Adaptive splitting . . . . . . . . . . . . . . . . . . . . . . 84

4.1.5 Testing with a second flame . . . . . . . . . . . . . . . . 85

4

Page 6: Numerical Modelling of Soot Formation

4.2 Deterministic Simulation of Deferred Events . . . . . . . . . . . . 86

4.2.1 Results with LPDA . . . . . . . . . . . . . . . . . . . . . 88

4.3 Recommended Algorithm . . . . . . . . . . . . . . . . . . . . . . 89

5 Models for Particle Shape 905.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

5.2 Framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

5.2.1 Flame chemistry model . . . . . . . . . . . . . . . . . . . 93

5.2.2 Numerical method . . . . . . . . . . . . . . . . . . . . . 93

5.2.3 Particle shape variable . . . . . . . . . . . . . . . . . . . 94

5.2.4 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . 94

5.3 Models for Lengths . . . . . . . . . . . . . . . . . . . . . . . . . 95

5.3.1 Collision Diameter . . . . . . . . . . . . . . . . . . . . . 95

5.3.2 Radius of Curvature . . . . . . . . . . . . . . . . . . . . 99

5.4 Model Comparison . . . . . . . . . . . . . . . . . . . . . . . . . 102

5.4.1 Bulk Properties . . . . . . . . . . . . . . . . . . . . . . . 102

5.4.2 Particle Size Distributions . . . . . . . . . . . . . . . . . 108

5.4.3 Individual particle behaviour . . . . . . . . . . . . . . . . 111

5.4.4 Future Experimental Validation . . . . . . . . . . . . . . 116

5.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

6 Explicit Statistical Weights 1186.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

6.2 General Approach . . . . . . . . . . . . . . . . . . . . . . . . . . 121

6.3 Weighting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

6.3.1 Dynamics of the New Measure . . . . . . . . . . . . . . . 123

6.3.2 Simulation Algorithms . . . . . . . . . . . . . . . . . . . 126

6.4 Numerical Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

6.4.1 Initial Validation . . . . . . . . . . . . . . . . . . . . . . 127

6.4.2 LPDA and real flames . . . . . . . . . . . . . . . . . . . 129

6.4.3 Performance . . . . . . . . . . . . . . . . . . . . . . . . 135

6.4.4 Further Comparison . . . . . . . . . . . . . . . . . . . . 136

5

Page 7: Numerical Modelling of Soot Formation

6.4.5 Potential Applications . . . . . . . . . . . . . . . . . . . 138

6.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

7 Conclusion 140

6

Page 8: Numerical Modelling of Soot Formation

Chapter 1

Introduction

Soot is generally regarded as a pollutant. There are many reasons for this; one is

the way that it collects as a black or grey deposit on buildings, an issue discussed

in [25]. This discolouring is especially noticeable on stonework that dates back to

times when coal fires were in widespread use and which has not been shotblasted

since. However, the adverse effects of soot particles on those who inhale them,

which are a cause of concern on a continental scale [3], are more serious. There

is evidence of a causal link between soot inhalation and respiratory disorders [46]

suggesting that epidemiological observations are more than correlations. Arden

Pope III and Dockery [10] undertook an extensive consideration of statistical data

reported by many authors, which broadly supports the existence of causal links.

They also give some indication of the rather limited and imprecise nature of the

current understanding of the health effects of soot on populations, a subject previ-

ously addressed by Mauderly [124], and which must temper political discussion

of the topic. Interestingly, soot is also reported to have the beneficial effect of

absorbing unwanted chemicals from the environment [96], rather like activated

carbon.

Carbon black, which is basically soot produced deliberately in controlled con-

ditions, is widely used as a filler to give rubber better mechanical properties

[56, 115], for example, in automotive tyres. In addition to adding it to rubber,

carbon black is used in other composite materials [210] and to make inks.

7

Page 9: Numerical Modelling of Soot Formation

These roles of soot, both as an industrial feedstock and as a pollutant, gener-

ate desire for models of soot formation that have significant predictive properties.

Manufacturers of carbon black want insight into its formation, to facilitate optimi-

sation of manufacturing processes for quality of product (measured, for example,

by particle size) and production cost. Others, typically designers of internal com-

bustion engines, want to understand the causes of soot formation to guide the

design of products that minimise emissions.

There is a third motivation for modelling soot formation: that of intellectual

curiosity. Fire is ubiquitous and detailed understanding of the chemistry of hydro-

carbon combustion has only developed in the last ten to twenty years, for example

compare the work of Maas and Warnatz [112] from 1988, in which only hydrogen

was considered, to that of Hoyermann, Mauß, and Zeuch [81] from 2004 where a

C4 hydrocarbon was successfully treated. While much remains to be done to un-

derstand hydrocarbon combustion, soot has the attraction of being even less well

understood.

1.1 Experimental Investigations of Soot

The main part of this thesis will deal with computational techniques for modelling

soot formation and growth. However, science is about models that describe the

real world and engineering is about changing that reality, so brief mention will

be made of some types of experimental work, which motivate the development

of the models considered later in this thesis and which might be used to validate

computational results.

As background to the experimental work discussed below, it may be helpful

to bear in mind that there is debate over the exact nature of soot and no definitive

model of how it forms. All one can say with confidence is that

• soot is mainly carbon with a modest amount of hydrogen and traces of any

metals present during its formation.

• soot particles form during fuel rich combustion via a process that includes

the formation of polyaromatic hydrocarbon molecules.

8

Page 10: Numerical Modelling of Soot Formation

• soot particles contain many large two dimensional structures, which resem-

ble graphite.

1.1.1 Collection and Examination of Soot Particles

Careful study of samples of soot particles provides significant information about

them. A very early example dates from 1901 when Hartley and Ramage [78]

analysed the metal content of soot particles and were able to distinguish between

soot formed in two different combustion environments.

Analysis of the chemicals on the surface of soot particles, by washing samples

in an organic solvent followed by chromatography [27], has shown the presence of

multi-ring aromatic compounds and interestingly hinted, nearly forty years in ad-

vance [205], at a current topic in mechanism development [219]. The same kind of

collection and washing technique is still used in work to understand the structure

of the large molecules in soot [7, 233]. Chromatography and mass spectrometry

were used for the analysis reported in [7], while Yan et al. [233] also used nuclear

magnetic resonance and analysed the insoluble soot residue with Raman spec-

troscopy to measure its structural resemblance to graphite. Recently, laser des-

orption has been coupled to mass spectrometry [23, 149] to measure the relative

abundance of organic compounds on soot particles and thus, to probe the chemi-

cal mechanisms leading to soot formation from different fuels. In this way Oktem

et al. [149] identified a significant quantity of non-aromatic species on particles.

These straight chain compounds are not part of the standard soot growth pathways

in the modelling literature so extended models, perhaps involving alkynes [144],

will be required to explain this data.

On a slightly larger scale, transmission electron microscopy (TEM1) has been

used to investigate the aggregate structure of soot particles. An outline of the

fractal analysis usually performed can be found in [101] and the procedure has

reached the point at which Tian et al. [195] could automate it. TEM analysis is

important for testing light scattering methods [168], which are mentioned below.

1In this thesis TEM will be used to refer to both the microscopy technique and, mostly in theplural form TEMs, to the micrographs thus produced.

9

Page 11: Numerical Modelling of Soot Formation

Another feature of TEMs is that they show a quite irregular arrangement of the

chemical structures at the surface of soot particles, in this regard the high reso-

lution images in El-Leathy et al. [52, figure 1] seem typical of those collected in

many different situations.

TEM is not the only electron microscopy method used to image soot particles;

Scanning Electron Microscopy (SEM) was used by di Stasio [37] to help validate

a light scattering technique. Atomic force microscopy has also been used to pro-

duce very finely resolved measurements of the size distribution of young particles

collected by thermophoretic deposition, before the aggregate structures usually

analysed by TEM had formed [18].

These two classes of data indicate opportunities for model development and

validation in the areas of soot precursor and surface chemistry, and particle struc-

ture, collision and transport. Models that explain either class of data will require

multivariate descriptions of soot particles. This thesis will present the develop-

ment of an efficient numerical technique for such models and look at an applica-

tion to the modelling of the aggregate structure of soot. Surface chemistry mea-

surements are generally more recent than those of particle configuration; mod-

elling of surface chemistry is a work that is just getting under way and will not be

addressed in any detail in this thesis.

1.1.2 In-Situ Light Scattering and Absorption

An early example of the use of light scattering to estimate particle sizes and to

infer a multi-modal particle size distribution can be found in Erickson et al. [53].

The quantitative analysis of the observed data depends on the uncertain value used

for the complex refractive index of soot. This problem is discussed in [53] and

later in work on the fractal structure of soot particles [187]; it is still the subject of

research [13]. To avoid this problem, a light scattering method to measure the size

of the primary particles or sub-units in soot aggregates, which does not depend on

the complex refractive index of the soot, has been suggested by di Stasio [37]. A

detailed review of fractal structure investigation by light scattering is presented in

[185].

10

Page 12: Numerical Modelling of Soot Formation

Fractal structure light scattering measurements are usually accompanied by

soot volume fraction measurements based on the amount of light absorbed [70,

234] since this is a rather important quantity. These measurements again have

the disadvantage of depending on the somewhat uncertain refractive index of the

soot being measured. This is a significant issue because comparisons of predicted

and measured soot volume fractions are a standard model validation technique.

Different optical techniques can give very similar results using a common value

for the refractive index [33], but other values lead to inconsistencies. Values from

supposedly precise laser measurements must therefore be treated with care.

Laser Induced Incandescence (LII) [129, 161] detects black-body radiation

emitted by soot particles heated by an incident laser beam; soot volume frac-

tion can be inferred from the initial intensity of the incandescence and particle

size from its decay. Because it uses re-emission LII can be used as a point sens-

ing technique for spatially resolved data [161]. However LII systems have to be

calibrated, for example, by comparison to an absorption method. A technique

has been presented by de Iuliis et al. [32], which takes measurements from two

incident wavelengths to avoid the need for calibration. This introduces some de-

pendence on the refractive index, but only on the ratio of the index at the two

wavelengths used and one should note that many papers assume a constant value

for all wavelengths. An uncertainty of 13% is quoted for the soot volume fraction

measurements reported in [32]. The laser excitation stage of LII can also excite

fluorescence in large soot precursor molecules [31] offering some extra informa-

tion on the problematic soot inception process.

These optical techniques allow for quite large amounts of data on the bulk

properties of soot to be collected at a range of locations in a combustion system.

This means that models of soot particle population dynamics may be tested by

comparing to a set of data points from the same system. Extremely high numerical

precision is superfluous and this thesis will show ways to simulate the predictions

of a soot model at many points with modest computational resources, provided

one is willing to accept a small amount of numerical imprecision.

11

Page 13: Numerical Modelling of Soot Formation

1.1.3 Other scattering

Naturally, wave scattering experiments have not been confined to visible or near-

visible light. Small Angle X-ray Scattering (SAXS) has been used to detect very

small soot or precursor particles (< 3 nm) in diffusion flames and to draw in-

ferences about the shape of some of these particles [40, 79]. Similar results are

reported for laminar premixed flames by Gardner et al. [63]. SAXS is a line of

sight technique, but it is also possible to look at x-rays which are scattered much

further from their initial direction. Detection of radiation scattered by more than

15–20 is known as Wide Angle X-ray Scattering (WAXS) [151]. A comparison

of WAXS and SAXS is given by Ossler and Larsson [152] in their introduction;

WAXS provides more information about the chemical substructure of the soot but

less about the sizes of the particles.

Moving away from electromagnetic radiation, scattering experiments have

also been performed using neutrons [132, 208]. Interpreting neutron scattering

data for soot presents a difficulty analogous to the refractive index problem for

light scattering—the carbon-hydrogen ratio [241]. However, Small Angle Neu-

tron Scattering (SANS) provides an alternative way to measure mean particle size

and soot volume fraction so that two independent methods may be compared.

1.1.4 On-line Particle Sampling

The final class of methods that will be mentioned in this brief review of exper-

imental techniques involve continuous sampling from a flame via a small tube.

The procedures for doing this require multiple dilution stages to try to quench

all reactions and stop evolution of the soot particle size distribution, see Maricq

et al. [122] and the more detailed account of Zhao et al. [239]. The gas stream

thus collected with entrained soot particles is then fed into various types of anal-

ysis equipment to determine what particles are present, generally to produce an

estimate of the Particle Size Distribution (PSD).

One way to proceed is to use Differential Mobility Analysis (DMA) [28],

which selects particles with a particular charge to mobility ratio. Diffusion of

12

Page 14: Numerical Modelling of Soot Formation

small soot particles limits the precision of this technique for very small soot parti-

cles, but it has been used coupled with Condensation Particle Counting (CPC) to

measure particle size distributions down to sizes of 3 nm [122, 238]. The combi-

nation of DMA, with the selection voltage scanned to successively select particles

of different mobilities and CPC is known as Scanning Mobility Particle Sizing

(SMPS). The multiple stages of dilution involved in this process make it quite

hard to obtain absolute values of the PSD, instead Zhao et al. [238, 240] report

values scaled to make the PSD a probability distribution, however, absolute values

are reported by Stipe et al. [189]. More detailed information on the method can

be found in the introduction to [192].

DMA selects one size/mobility of particle at a time, but it is also possible

to sort a particle stream with similar electrical mobility techniques and detect

the output in all the size/mobility classes at once [192]. This recent method is

known as Differential Mobility Spectrometry (DMS), it offers a practical way

of collecting large numbers of particle size distributions from flames [54]. The

ability to collect successive spectra at intervals of less than 1 s is also interesting,

particularly for automotive applications, which have been the main use of the

technique so far [192].

A sampled and diluted gas stream may also be fed into a mass spectrometer

[72]. This method cannot detect large soot particles but gives very detailed in-

formation on the sizes of small soot particles and smaller structures containing

only tens of C atoms [180]. This is very similar to the use of mass spectrometry

referred to in §1.1.1

The techniques mentioned in this section are most interesting for their abil-

ity to measure particle size distributions. Their main limitations are problems in

relating different mobility properties to particle structure. Data collected in this

way offer great scope for testing models of particle coagulation and growth and in

particular models which can predict particle mobility as well as mass.

13

Page 15: Numerical Modelling of Soot Formation

1.2 Combustion Chemistry

Having given some indication of the data that is available for testing models, it

is worth considering the main input to a soot model—the chemical environment.

The work in this thesis will concentrate on numerical methods for treating soot

particle inception and growth, that is, stages 4 and 5 of figure 1 in McEnally

et al. [125]. Consequently the results of combustion chemistry calculations will

be taken largely for granted. The mechanism for the calculations underlying parts

of this work is that of Wang and Frenklach [207]. Chemical kinetic data in [207]

was taken from experimental studies and quantum mechanical calculations. As is

inevitable for any large, new reaction scheme, values were not available for some

constants and so had to be estimated from data for similar reactions. This kinetic

model was then incorporated in soot calculations and compared to a range of

experimental data for soot volume fractions and, in some cases species concentra-

tions [8]. On the basis of comparisons with the multiple sets of experimental data

available to them, the authors adjusted some constants, within plausible ranges,

to produce a mechanism that became a widely used standard. This mechanism is

known as the ABF mechanism after the authors of the paper Appel, Bockhorn,

and Frenklach [8].

McEnally et al. [125] provide an extensive review of work that has been done

on mechanism development and some of the problems involved, in particular, that

much of this work involves under-determined optimisation problems. It is there-

fore not surprising that they give an example of two mutually exclusive but inter-

nally consistent sets of results. The general development of combustion modelling

is much too broad a topic to address here. Readers may be interested to consult

[130], which also challenges the route of Wang and Frenklach for the formation of

the first aromatic rings. Different routes for this step are discussed by Frenklach

[58]. The growth of aromatic structures seems to be the key connection between

combustion chemistry and soot formation and so will be considered a little more.

14

Page 16: Numerical Modelling of Soot Formation

1.2.1 Polyaromatic Hydrocarbon Growth

Because polyaromatic hydrocarbon (PAH) molecules are the initial building

blocks for soot particles [8, 40, 207] and soot particles have a large graphitic

component [38, 149, 151], models for the growth of multiple ring structures have

attracted a lot of attention. Being able to quantitatively predict the formation of

the PAH species that are responsible for soot particle inception is very impor-

tant if soot models are to reproduce all the features of particle size distributions

[179]. Singh, Patterson, Kraft, and Wang [179] also highlight the significant con-

sequences of the model for soot particle growth by addition of new aromatic units

to the existing PAH / graphite structures in particles.

Some insight into these processes is offered by detailed experimental work

to characterise the chemical structure [152, 176] and physical shape [40] of the

components of soot particles. The first conclusion to draw from these studies is

that soot is complicated—a wide range of graphitic structures are observed, to say

nothing of the aliphatic species detected in [149]. Along with these observations

goes a large amount of analysis seeking to elucidate a detailed growth mecha-

nism, since the Hydrogen Abstraction Carbon Addition (HACA) [60] used in the

ABF mechanism cannot be a complete model, because ring closure by acetylene

addition fills in ‘bays’ on the edges of PAH molecules without generating more

such structures [176]. More detailed mechanisms involving five member rings

have been developed by Frenklach and co-workers to provide a more complete

description than simple HACA [61, 219]. They performed quantum mechanical

calculations using density functional theory to identify elementary reaction steps

and their activation energies as well as product structure. The note in [61], that

calculations were carried out on a “Xeon cluster”, confirms that a number of years

of Moore’s law2 growth in computing power will be required before these calcu-

lations can be made for individual particles in flame simulations.

A similar approach has been followed by Violi and co-workers [30, 204, 206].

2Moore’s law refers to an observation, in 1965, by Gordon Moore, the founder of Intel, thatimplies computing power on chips roughly doubles every 18 months [137]. With 40 years moredata he could still say the same thing!

15

Page 17: Numerical Modelling of Soot Formation

Unlike the ABF model, they conclude [30] that acetylene surface growth, which

they also include, is not an important means for mass to enter the nanoparticle

and soot phase, because most of the soot mass is already in PAH molecules (al-

though not in soot particles) quite early in lightly sooting flames. By implement-

ing their models into a hybrid kinetic Monte Carlo-molecular dynamics simulation

[201, 203] they were able to simulate the growth of PAH molecules up to a few

thousand C atoms including details of all the atoms and bonds. Monte Carlo step-

ping was used to grow the particle according to the rates determined by the con-

figuration and chemical environment. The configuration of the molecule was then

relaxed with a few picoseconds worth of molecular dynamics simulation. These

calculations support observations of high aspect ratio soot precursor/sub-primary

particles [40]. The authors of these papers do not comment on the computational

requirements of their methods, however, the title of a related paper by some of

the same people [103] begins with the words “massively parallel”! Nevertheless,

by means of ‘coarse grained’ molecular dynamics [87], in which potentials are

matched to groups of atoms, this work has been extended to include coagulation

from an initial population of 10,000 PAH molecules each of 200 C atoms [86].

The coarse grain potentials were chosen to match atomic scale molecular dynam-

ics calculations before the simulations were started. This means that, while the

method gives some new information about the way soot particles might form, it

cannot yet be incorporated in flame simulations. The problem is that atomistic

molecular dynamics calculations would be required for every new particle that

formed, in order to calculate the potentials for use in the coarse grained part of

the simulation. When the computational demands of these calculations have been

reduced and the available computing resources increased, the present author looks

forward to seeing the natural merging of the work just discussed and the much less

detailed Monte Carlo simulations that are the main topic of this thesis.

Examples of the scale of model that is currently feasible for PAH growth in

simulations of soot formation are given in [167, 218]. Further progress should

be possible as adaptive methods [107, 165, 175], which simplify the chemical

mechanism according to the local conditions, release resources for more detailed

16

Page 18: Numerical Modelling of Soot Formation

PAH modelling.

1.3 Population Balance Modelling

The above sections attempt to provide a little background to the work on soot

modelling in this thesis. A general, spatially homogeneous population balance

equation [164] is

∂tf (x, t) +∇. [A (x, t) f (x, t)] = h (x, t) (1.1)

where f is the density of the particle distribution, x is the particle type and the

model is specified by A and h. Continuous changes to particles, effectively con-

vection in particle type space, are encoded in A. In certain situations there is

some cross-over between A and h. Processes which lead to small discrete change

in particle type are naturally incorporated in h by two terms; one for the death

of the original particles and one for the birth of the changed particles. However,

such processes may be approximated by a continuous change of particle type and

moved into A. This approximation is made in all the deterministic calculation

methods considered in this chapter and in chapter 4. In situations where there

is only particle growth and no coagulation Matsoukas and Lin [123] have shown

that a simple continuous approximation to discrete growth loses the diffusive / dis-

persive character of the solution to the population balance problem. The Monte

Carlo methods used in this thesis remove any need for continuous approximations

to discrete processes. The errors introduced in the alternative methods considered

in this chapter are likely to be small since the overall structure of the solution to a

population balance problem will normally be dominated by the effects of coagu-

lation when it is present.

In this thesis, unlike in Ramkrishna [164], populations are treated as homo-

geneous in physical space so terms in spatial position do not appear in (1.1). A

review of approaches to (1.1) is given in [164], which is a hard-back book, so

the next paragraphs will try to concentrate mainly on techniques that have been

17

Page 19: Numerical Modelling of Soot Formation

applied to solve the population balance equations for soot and other combustion

generated particles.

1.3.1 Explicit Discretisation of Type Space

The simplest approach to the population balance problem is to reduce the number

of particle types to a manageable level, that is, to impose a finite grid on the

particle type space. The dynamics of the population size within each grid cell are

then approximated in a computationally tractable way. In some situations there

may be a natural discretisation. For the soot formation problems considered in

this thesis a natural discretisation is given by counting the number of carbon atoms

in a particle. When such a discretisation exists, one may express the population

balance problem as a countably infinite sequence of coupled ordinary differential

equations [198] (assuming spatial homogeneity), truncate this series and solve

the differential equations by standard methods while monitoring the truncation

error. This is occasionally useful as a validation technique (see chapter 2) for

other algorithms on specially constructed test problems. However, the number of

equations that have to be considered to keep the truncation error within acceptable

limits means the method is rarely useful for application to physically realistic

systems.

Fixed Sectional Methods

Methods in which a grid is placed on the particle type or state space, and an a pri-

ori assumption made about the shape of the population density within each section

or grid cell, are known as fixed sectional methods. The idea appears to have been

introduced by Gelbard and Seinfeld [65]. Hounslow et al. [80] first adapted the

equations to conserve both particle number and mass, a subject covered in more

depth in [104], where it is noted that an arbitrary number of moments could be

conserved. Conservation of a larger number of (optionally centered) moments has

been addressed recently by Alopaeus et al. [6] who demonstrate increased accu-

racy and suggest that the extra CPU time required per section can be recouped

18

Page 20: Numerical Modelling of Soot Formation

by reducing the number of sections. The method was used to simulate soot under

the assumption of spherical particles in an axisymmetric laminar diffusion flame

[182] without any apparent difficulties due to the stability problems at low par-

ticle concentrations mentioned [80, 105]. The diffusion flame soot and detailed

chemistry model were solved in a fully coupled calculation, using time marching

and grid refinement to get the full spatial solution, and so fixed sectional methods

have proved fruitful tools for understanding the evolution of the gas and particu-

late phase in different flame regions [183, 184].

Gridding may also be applied to multi-dimensional particle type spaces [227]

(that is, where particles are described by more than one independent co-ordinate

such as particle mass and surface area) and was of particular relevance in the in-

vestigation of sintering in inorganic nano-particles. Simplifications to the bivari-

ate sectional technique have been developed [142] but it remains computationally

demanding. The method seems unattractive for models with a particle type of

dimension greater than two because the number of sections grows exponentially

with the number of dimensions. However, using a fixed sectional method and

some analytic expressions for the coagulation source terms, Immanuel and Doyle

III [85] were able to perform computations for a trivariate model with computa-

tion times on unspecified hardware of around 1 hour. A paper by the same authors

published [84] two years earlier (for a 1-dimensional model) reported computa-

tion times on machines with twin CPUs running at 800 MHz so, even if the same

hardware was used to produce the 1 hour figure, the demands of the trivariate case

are seen to be substantial.

A partial solution to this problem is actually presented in the original paper

of Gelbard and Seinfeld [65] by allowing additional particle properties to vary as

functions of the main property, which was directly discretised into sections. This

has the effect of restricting the support of the number density to a 1-dimensional

manifold but allowing that manifold some freedom to move over the higher di-

mensional type space. These models can be described as ‘1.5-dimensional’ al-

though the phrase ‘embedded 1-dimensional’ might be more mathematically pre-

cise. They have been successfully validated against more detailed 2-dimensional

19

Page 21: Numerical Modelling of Soot Formation

sectional calculations for inorganic nanoparticles [89] and found to reduce com-

putation times by over 3 orders of magnitude. The accuracy of these ‘embedded

1-dimensional’ methods is not surprising since the distributions reported in [227–

229] are approximately ridge shaped. Embedded 1-dimensional methods have

been used to provide a computationally efficient representation of particle shape

during the development of improved soot chemistry models [216, 218]. They are

well suited to this application, because low computational cost is needed to leave

resources free for the chemistry solver, and information on the amount of particle

surface available for chemical reactions is important for physical accuracy.

Modelling the data from the experimental methods mentioned in §1.1 will re-

quire multivariate descriptions of soot particles. As mentioned previously, the

applicability of sectional methods to multi-dimensional situations is rather lim-

ited, because, in the absence of special problem structure, the number of sections

grows exponentially in the number of variables. In addition to this, fixed sectional

methods also face problems with numerical diffusion because small increments

of mass growth on particle surfaces are treated via the divergence term on the left

hand side of (1.1). Some progress on this issue has been made [155] and used to

model soot structure in plug flow reactors [156]. The linear approximation to the

distribution within each section suggested in [226] might also be expected to help

in this regard. A review with extensive numerical comparisons between fixed sec-

tional methods is given by Vanni [198] who recommends the form of the method

described by Kumar and Ramkrishna [104]. Fixed sectional methods, at least in

their more basic forms, are a special case of the finite element methods considered

below [84].

Moving Sectional Methods

In the simple case of a plug flow reactor (nothing more complex will be studied in

this thesis) the sectional model can be developed further to address the problem of

numerical diffusion between sections. This arises, as mentioned above, because

surface growth causes small changes in size to all particles within a section. If the

section boundaries are fixed, some particles will therefore cross into neighbouring

20

Page 22: Numerical Modelling of Soot Formation

sections. However, without knowing the distribution of the particles within the

sections it is not possible to account for this accurately.

The solution proposed in [105] is to move the section boundaries to match the

deterministic size changes due to surface reactions. The method, with some addi-

tional development, has been applied to TiO2 synthesis [197] where it was used

to explore particle growth pathways and the transition from coalescent to aggre-

gate particle development, both current issues in soot research. Wen et al. [215]

demonstrate the use of a moving sectional technique with an approximation for

particle shape, coupled to a chemical kinetics solver for plug flow reactors. This

raises the possibility of the first fully coupled premixed laminar flame simulations,

neglecting streamwise diffusion, that include arbitrary soot particle size distribu-

tions. However, as shown in a follow-up paper, in which the authors compare

moving and fixed sectional techniques for a plug flow reactor, the computational

demands are considerable—over 12 hours for their recommended method on a

machine with a 1.8 GHz Intel Xeon CPU [217].

The above work assumes, that within each section, the particle distribution is

concentrated at a ‘pivot’ point [105]. Analogously to the work of Wynn [226]

for the fixed sectional case, finite elements may be used to model the distribution

within sections instead of delta-functions. This allows for more accurate calcu-

lation of the source terms for coagulation and any other effects, which cannot be

represented as continuous particle growth. Hu et al. [82] used cubic spline ele-

ments, and this seems to address the dispersion issues that can arise due to the

artificial distribution of coagulation products between sections [105]. The cost of

this extension is that the inter-sectional coagulation rates have to be calculated by

a more complex quadrature method rather than with the simple formulæ of the

pivot case.

One alternative to the traditional quadrature formula is to use Monte Carlo

methods, which exploit the stochastic nature of coagulation, for the coagulation

integrals. Sun et al. [191] implement this approach with a time-drive Monte Carlo

method but do not consider general distributions within sections.

21

Page 23: Numerical Modelling of Soot Formation

Sectional Methods and CFD

Fixed sectional methods are well suited to incorporation in Computational Fluid

Dynamics (CFD) calculations [156, 160] because transport simply moves par-

ticles between identical sections at different spatial locations. Moving sections

are more troublesome because the correspondence between particle sections in

adjacent fluid cells is not known. The work of Pyykonen and Jokiniemi [160] ad-

dresses this problem by mapping between the two sectional approaches: they used

a moving sectional technique to simulate growth processes in Lagrangian stream

tubes, but transformed their population onto fixed sections to include cross stream

tube transport (and coagulation) at intervals of a few steps. The same approach

is used in aerosol modelling for air quality purposes [98] where it is attributed to

other papers published at the same time as [160].

Examples of the use of fixed sectional techniques to simulate flame aerosols in

a laminar flow field include [237] and, more recently, the study of unwanted par-

ticle formation during a continuous vapour deposition process similar to that used

in computer chip manufacture [190]. Turbulent flow fields require more com-

putational power. For example, Miller and Garrick [131] coupled only 10 fixed

sections to DNS of a planar jet (Reynolds number 4000), but their computation

took 105 CPU hours, which they spread over 250 CPUs in a parallel computa-

tion. In their conclusion Miller and Garrick suggest that using turbulence models

should bring calculation times below 1 CPU day. This is of limited practical use in

the short term, but advances in algorithms and computer power should eventually

enable coupling to turbulent systems with sectional methods that produce detailed

particle size distributions. In the view of the present author, other particle based

techniques are likely to enable this goal to be reached more quickly.

Finite Differences

The sectional methods discussed above arise from the structure of the problem as

a population balance. Methods derived for general differential equation problems

are also applied, for example finite differences. The author is not aware of any

papers in which finite differences have been used for soot population modelling,

22

Page 24: Numerical Modelling of Soot Formation

but good agreement between computations and measurements are reported for

the growth of silver bromide crystals in [143]. Some authors report problems

with the stability of the method [113, 221], but Braatz and co-workers seem to

have found the method tractable for crystallisation problems [75] with splitting to

allow for a bivariate population balance problem. Coupling to CFD is reported by

the same group [222] suggesting there are no major numerical problems, although

quantitative comparisons to experimental data have not yet been published!

Finite Elements

Moving further into the standard tool kit of differential equation techniques, one

comes to finite elements. The principle of these methods is to project the solu-

tion within a grid cell onto a pre-determined finite dimensional vector space of

functions with finite support [174]. The projection will not, in general, exactly

solve the population balance problem and the projection is chosen to minimise

the weighted integral of the difference or residual. There is thus a large overlap

between finite element techniques and the method of weighted residuals [164],

although neither class of methods contains all examples of the other. In particular

weighted residuals can be used with ‘infinite elements’, such as Hermite [77] and

Laguerre [83] functions to give spectral methods for the distribution, these will be

mentioned briefly in the section on integral methods below §1.3.2.

Collocation methods use Dirac delta functions as the weighting for the resid-

ual integral and require that the residual be 0 at a particular set of points within

each interval. Gelbard and Seinfeld [64] used collocation with cubic basis func-

tions to solve problems for which analytic solutions were available for compari-

son. Because collocation is only concerned with the residue at a discrete set of

points much less quadrature is required for the coagulation source terms than other

weighted residual methods [64, 145]. However, there are reports that collocation

methods are less stable than Galerkin methods, both when solving steady state

[146] and dynamic [113] problems.

Galerkin methods use the basis functions themselves as the weights with

which to integrate the residual. As mentioned above, this makes the calculation

23

Page 25: Numerical Modelling of Soot Formation

of the coagulation source terms rather expensive, although the level of precision

in the calculation of these integrals is not too significant for the final results [173].

Roussos et al. [173] present results using a range of different order polynomials

as the basis functions and show quadratic basis functions perform better than cu-

bics and quartics, contradicting Mantzaris et al. [117] who recommend the use of

quartics. Non-polynomial basis functions have also been successfully used, for

example, sawtooth functions [158]. Unlike finite difference and collocation tech-

niques, Galerkin methods have been coupled to chemical kinetics and used to test

extended soot formation models [2, 9, 144].

Least square methods [42, 164] minimise the integral of the square of the

residual over the (truncated) particle domain. Dorao and Jakobsen [44] extend the

method to handle particle type, physical space and time in a unified way. They

report numerical tests but, understandably, do not have an equally sophisticated

alternative method available for comparison. The direct quadrature method of

moments discussed below should provide some opportunities for such compar-

isons in the future. Arbitrary flow fields can be incorporated in the method of

Dorao and Jakobsen [44], but have to be precalculated.

For obvious reasons many authors repeat their calculations using a range of

numerical parameters, for example polynomial order or grid spacing. At the cost

of increased computational complexity, adaptive methods [45, 225] may also be

used, as for differential equations arising in other areas of science.

Mono-Disperse Assumption

The most extreme case of discretisation is to use a single point grid, that is, to

assume a mono-disperse population. Such a method will yield very little infor-

mation about a particle population and risks major systematic error. On the other

hand it is very simple to implement, fast to compute and is sufficient for providing

estimates of particle number and total mass as part of more complicated calcula-

tions [111]. With this method on a two-dimensional particle type space, that is

by tracking total particle number, volume and surface area Tsantilis and Pratsinis

[196] are even able to model the onset of aggregation in the formation of silicon

24

Page 26: Numerical Modelling of Soot Formation

nanoparticles.

The continued use of mono-disperse models, which no attempt is made to

review here, is an important reminder of what can be achieved with simplicity.

Since this thesis is not about what can be achieved with the minimum computa-

tional effort, but focuses more on what is the maximum that can be achieved with

desktop computer power it seems that a detailed review of the literature on this

topic would be beyond its scope.

1.3.2 Integral Methods

This class of methods includes the mono-disperse assumption as a degenerate

case and is also motivated by the observation that, for many purposes, the full

particle size distribution is not necessary. Instead it may be sufficient to consider

the moments, or other integrals, of the distribution. For a k variable population

balance with population density f the moments are defined by

Mi1,i2,...,ik :=

∫ (∏j

xijj

)f (x1, x2, . . . , xk) dx1dx2, . . . dxk, (1.2)

where the ij will be called the exponents of the moment. Only in very limited

situations [83] is it possible to construct a closed set of moment equations, for an

example see Terry et al. [193].

Functional Closures

This category comprises the methods in which closure is achieved by making as-

sumptions about the shape of the population distribution or the form of the depen-

dence of the moments on the exponents. One approach is to use series expansions,

for example using Hermite [77], Laguerre [20, 83] or other [116] functions. The

underlying principle of such approaches is to estimate the size distribution from a

finite number of its moments and then use the estimated distribution to calculate

additional quantities. One of the more simple applications of the technique is the

assumption of a log-normal distribution [159, 220]. The ‘moment problem’ (how

25

Page 27: Numerical Modelling of Soot Formation

to estimate a distribution from a set of moments) has been extensively considered

in many settings. Some recent information may be found in [12].

Many flame simulations have simply fitted a polynomial to the moments or,

more usually, their logarithms, at the exponents for which they are known and then

interpolated and extrapolated to find moments with other exponents. Different

piece-wise polynomial interpolants have been investigated [41] and small amounts

of extrapolation were also found to work. It is also observed in the same paper

that assuming a log-normal shaped particle size distribution is equivalent to using

a quadratic interpolant on the first three moments [41].

Frenklach and Harris [59] introduced a method in which polynomials were

fitted to (the logarithms of) the moments for simulating soot aerosols in the free

molecular regime. The collision rates in the free molecular regime could not be

expressed as power laws so the authors introduced ‘grid functions’ to enable an

approximate closure of their equation system. To reduce the error associated with

extrapolation to moments with negative exponents they also introduced a moment,

µ−∞, for the single variable case they were dealing with defined by

µ−∞ = limi→−∞

Mi

xi. (1.3)

This was possible because the soot model used defined a smallest particle size, x.

Consequently, µ−∞ could be interpreted as the number of particles of this size.

Introducing µ−∞ meant moments with negative exponents could be found by in-

terpolation much more accurately than would be possible by extrapolating from

the moments with positive exponents. A small amount of extrapolation was re-

quired to reach moments with exponents greater than 3, but the method became

known as the ‘Method of Moments with Interpolative Closure’. It will be abbre-

viated MoMIC in this thesis, except when the extrapolation is to be emphasized,

when MoMIEC will be used.

To extend coagulation calculations to soot particles in the transition regime

between the free molecular and slip flow cases Kazakov and Frenklach [92] used

a harmonic mean interpolation for the moment change rates. This was based on

work in [159] and has been very widely used despite the fact that the moment

26

Page 28: Numerical Modelling of Soot Formation

change rates cannot be derived from any particle-particle interaction rule. In a key

paper for the development of soot models [8] the method was further extended

to allow particle surface growth rates to depend on the mean particle size. This

introduced another point where the method driven requirement for the model to

be expressed in terms of moments led to a model that could no longer be de-

scribed in terms of particle dynamics. By fitting two parameters in the surface

growth expression, MoMIC calculations of soot volume fraction were able to get

within one order of magnitude of experimentally observed values for a range of

flames and better results were possible by optimising the parameters for individual

flames. MoMIC is reviewed in [57] and has continued to be used for soot model

development [14].

Because of its small size (less than ten scalars) the MoMIC is highly suitable

for incorporation in reacting flow calculations to account for the effects of soot.

The development described above was carried out in the context of premixed lam-

inar flames. Moody and Collins [136] used a simple chemistry model with the

method of moments for titania particles in DNS calculations and showed that in-

creased levels of mixing resulted in narrower particle size distributions. Wang

et al. [209] used the soot model and method from [8] with a k − ε turbulent flow

calculation, but found their results to be significantly affected by the radiation

models they used. An alternative to polynomial interpolation is the Weyl trans-

form fractional derivative method of Alexiadis et al. [5].

Quadrature Methods of Moments

These methods can be described in several different ways. Their name arises be-

cause the particle population is represented by a series of weighted quadrature

points which are chosen to give precise values for certain moments. This makes

them like presumed distribution shape methods, the presumed shape being a fi-

nite combination of weighted Dirac-delta functions. The quadrature points may

be chosen on the basis of the leading terms of a series expansion of the distribu-

tion in its moments [20], which is a return to the ‘functional closures’ discussed

above. Alternatively, one can view quadrature methods of moments as determinis-

27

Page 29: Numerical Modelling of Soot Formation

tic weighted particle methods, although the maximum number of particles that can

be handled is limited in some situations [43, 172] by ill-conditioning problems.

There are a number of different ways to calculate the evolution of the quadrature

points: one may take a statistical perspective on representing the unknown distri-

bution [235] or a more explicitly quadrature driven approach [119, 126]. Grosch

et al. [71] include a detailed account of the development of QMoM in their in-

troduction, which readers should consult for more detailed information on the

method. Their primary purpose though is to present a generalised framework for

QMoM, from which some new variants can be derived. QMoM can also be con-

structed as a method of weighted residuals [43].

The original form of QMoM was readily extended to bivariate particle types

and a twelve point scheme found to be highly accurate, while smaller schemes

also produced reasonable results [223]. A simulation of the structure of alumina

particles in a precomputed laminar diffusion flame [172] predicted aggregate and

primary particle diameters that compared well with experimental data, showing

the method could be very useful for models of soot particle structure. The statis-

tical principal component approach can be extended to multivariate models [236]

as can the direct method [119]. The direct method (DQMoM) has been applied

to a range of complex real systems by its designers, most immediately relevant is

the fully coupled simulation of soot formation in a turbulent flame [244]. It seems

very clear that DQMoM, or possibly a slight variation of it, is suitable for use with

quite detailed soot models in reacting flow calculations, which attempt to predict

the kind of measurements outlined in §1.1. The main limitation would appear to

be in the area of reconstructing the underlying particle distributions.

1.3.3 Stochastic Particle Methods

The Direct Simulation Monte Carlo (DSMC) technique that underlies most of the

work in this thesis was first used by Bird [22]. In a short paper on the relaxation of

the momentum distribution in a rarified gas he outlines the basic idea: represent

the physical population of gas particles by a collection of computational particles,

but track only the particle property one is interested in—momentum in the original

28

Page 30: Numerical Modelling of Soot Formation

application. An assumption of chaos within a homogenous volume then justifies a

random generation of collision events. Also in the 1960s, the same idea was used

to simulate collision based droplet mixing [188], for which droplet composition

and not momentum was tracked. The standard reference for the DSMC in gas

dynamics is the book by the same author [21] and its later edition. An overview

of the method and some of its applications to gas dynamics, that is solving the

Boltzmann equation, is given by Oran et al. [150].

Approaches to Particle Growth Problems

In particle formation problems the particle property of interest is mass (and pos-

sibly structure) not momentum, so DSMC means specifying collision and coag-

ulation rates as expectations based on the assumption that all particles perform

independent random walks. A comprehensive mathematical review of stochastic

approaches to coagulation problems [4] is also notable for the number of numer-

ically motivated conjectures it contains. Mathematical results about stochastic

particle algorithms in the probability literature typically say, that in an appropri-

ate limit, a sequence of Markov chains converges to one or more weak solutions

of a population balance equation [34, 48, 49]. Sequences with multiple limits

have been constructed [147] but never reported by those using the technique in

scientific rather than mathematical situations.

The convergence approach mentioned in the preceding paragraph depends on

first finding the population balance equation, generally by taking a limit of an un-

derlying stochastic model. Direct Monte Carlo simulation of particle formation

problems did not begin with this relatively sophisticated approach via stochastic

calculus; the stochastic calculus formalised what was already being done. Gille-

spie [67] derived and precisely described a continuous time Markov chain Monte

Carlo simulation algorithm for coagulation based on a model of the underlying

physical system as a stochastic process. One of the key steps is the derivation of

exponential inter-event waiting times [55], which relies on a model of the under-

lying physical process that is equivalent to a continuous time Markov chain. Other

algorithms for the simulated time step are possible [110, 181] but still depend on

29

Page 31: Numerical Modelling of Soot Formation

the same stochastic model for the underlying physical system.

The stochastic approach of constructing Markov processes that solve the pop-

ulation balance equation in some limit has given rise to algorithms other than

direct simulation [35, 48, 97]. That is to say, processes have been constructed

where computational particles and events can no longer be regarded directly as

the history of a small sample of the physical system. Two examples and further

discussion are given in chapter 6.

Applications of Stochastic Particle Methods to Particle Formation

A form of direct simulation was used by Wu and Friedlander [224] and Rosner and

Yu [171] to find self preserving size distributions for a bivariate particle sintering

model. Solving for a self preserving distribution removed the need to track the

relationship between simulation steps and time. While the method could have

been used to follow the dynamics, this was not done because it was not part of

the authors’ immediate goal. After Goodson and Kraft [68] 3 addressed a key

numerical issue, Balthasar and Kraft [16] first outlined DSMC for soot formation

in laminar premixed flames and further application is found in [238]. Most of the

further development of this work is reported in this thesis along with references to

additional applications and so is not discussed further here. However, applications

to inorganic nanoparticles should be noted [138, 211, 212].

1.4 Mathematical Framework

This section provides an overview of the mathematical framework in which soot

formation and growth will be considered in this thesis. The paradigm is that:

• Individual soot particles may be completely described by elements of some

type space E on which addition, corresponding to coagulation, is defined.

• The soot population is described at time t by the number n (t, x) per cm3 of

particles of type x ∈ E.

3In [68] DSMC is used in a much narrower sense than it is used in this thesis.

30

Page 32: Numerical Modelling of Soot Formation

• n evolves according to the discrete Smoluchowski coagulation equation:

d

dtn (t, x) =

(Kt (x) +

∑l∈U

S(l)t (x)

)(n (t, ·)) + I (t, x) . (1.4)

In these definitions is an implicit assumption that E is countable so that sum-

mations are meaningful. This assumption is common in the literature, but not

essential, and later in this thesis it will be removed, by replacing the sums with

integrals. The coagulation operator K, which is time dependent and non-linear, is

then defined by

Kt (x) (n (t, ·)) =1

2

∑y,z∈E:y+z=x

Kt (y, z)n (t, y)n (t, z)

−∑y∈E

Kt (x, y)n (t, x)n (t, y) .(1.5)

The first sum represents coagulation to form particles of type x, the second loss

of particles of type x due to coagulation. Kt (x, y) defines a map from the con-

centrations of particles of types x and y to their coagulation rate at time t given

by Kt (x, y)n (t, x)n (t, y). K is known as the coagulation kernel and details are

given in [92, 177].

Surface reactions which only involve one physical particle at a time are de-

scribed by the linear operator S defined by

S(l)t (x) (n (t, ·)) =

∑y∈E

βt (y)P(g(l) (y) = x

)n (t, y)− β(l)

t (x)n (t, x) . (1.6)

Where l ∈ U , an index set for the surface reactions. In the context of the soot

model used in this thesis [8, 207]

U = C2H2 addition,OH oxidation,O2 oxidation, pyrene condensation . (1.7)

β(l)t (x) is the rate at which a particle of type x undergoes surface reaction l at

time t.

31

Page 33: Numerical Modelling of Soot Formation

g(l) (x) is the type of the resultant particle when a particle of type x undergoes

surface reaction l. If the reaction removes the particle from the soot population a

special value 0 ∈ E will be used so that g(l) is everywhere defined. It is possible

for g(l) to be a random function, in the deterministic case the probability in (1.6) is

to be understood as an indicator function. This is an area in which the framework

is slightly more realistic than much earlier work, for example [217], in which

surface growth is modelled as a continuous process of infinitesimal changes.

I (t, x) is the rate at which particles of type x enter the system at time t. It is

defined by the soot inception model (since no transport problems are considered).

In the context of the soot model used in this thesis [8, 207], this means I is always

zero except when x represents a particle of 32 C atoms.

Particles of less than 32 C atoms are assumed to belong to the gas phase which

is simulated separately using the 1-dimensional laminar flame code, PREMIX and

a Method of Moments approximation to allow for the removal of gas phase species

into the soot population, see [16, 177] for details. This precalculation of the gas

phase flame environment is very important because it yields the time dependence

of β, the K and I . Because soot particle diffusion is small and may be neglected,

one can then transform the problems considered in this thesis from boundary value

problems (a flame with two known ends) to initial value problems (a small parcel

of gas moving through a known flame). Two way coupling of stochastic particle

simulations to chemical reaction calculations is reported by Celnik, Patterson,

Kraft, and Wagner [26].

Throughout this thesis the initial condition of no particles will be used,

n (0, x) = 0 for all x ∈ E, since unburned gases entering a flame are assumed

not to contain soot.

1.4.1 Models for individual soot particles

Many different descriptions of soot particles are possible and provide varying lev-

els of detail. The three models most relevant to this work are set out here in de-

creasing order of detail. The molecular dynamics work of Violi and Venkatnathan

[202] and Violi [201] mentioned above would represent a higher level of detail,

32

Page 34: Numerical Modelling of Soot Formation

as would coarse grained approximations to that work [86]. For a brief discussion

see §1.2.1.

Primary particle model

HereE is the subset of(E × R3

)Nwith only finitely many non-zero components.

A soot particle is described as a sequence (order is not generally significant) of

‘primary particles’ each described by an element of E and some location e.g. dis-

placement from the first particle in the list. As far as the author is aware, all work

so far has modelled primary particles as spheres of constant density described by

their mass or volume and so used E = [0,∞).

Balthasar et al. [17], building on the work of Mitchell and Frenklach [134]

and Mitchell [135], modelled coagulation by assuming two particles stick rigidly

at the first point of contact and without realignment. They performed surface

growth on each primary particle by integration - with infinitesimal steps—losing

the discrete nature of the surface events. The computational effort required to

simulate this model is great because of the numerical calculations required to find

collision radii for the aggregate particles.

Surface and Volume Model

This model takes E = R+ × R+ and says that a particle is described by x ∈ E if,

and only if, it has volume x1 and surface area x2. Addition is defined on E by

z = x+ y ⇐⇒ z1 = x1 + y1 and z2 = x2 + y2

and hence coagulation is modelled without any coalescence. This model is consid-

ered in chapter 5. It is an intermediate stage between the Primary Particle Model

and the Coalescent Sphere Model. The definition of the g(l) is non-trivial even

with a clear picture of the underlying chemical processes. As discussed in §1.3,

one can use an infinitesimal surface growth approximation [17, 135]. However,

even infinitesimal growth approximations do not avoid the need for a physical

model of the change of particle shape during surface growth and so some possible

33

Page 35: Numerical Modelling of Soot Formation

models are considered in chapter 5.

Coalescent Sphere Model

The simplest and standard case is when all particles are assumed to be spheres

with a common density. For this take E = R+ and let x be the mass or the

volume of the particle it describes. It is then convenient in the case of soot particles

(provided H and other species may be neglected) to take E = N where x ∈ E is

the number of C atoms in a particle it describes. Addition is defined just as for the

natural numbers: the result of the coagulation of two spheres is a new sphere with

volume equal to the sum of the volumes of the initial spheres, that is, coagulation

is completely coalescent. Except for chapter 5 the coalescent sphere model for

soot particles is used in this thesis.

Different coagulation kernels are appropriate in different conditions. The co-

agulation kernels are also used to model the pyrene condensation and soot incep-

tion processes, for details see chapter 2 and [177]. All the work reported in this

thesis was carried out using the transition regime kernel defined in equation (2.1),

the use of more specific kernels was not found to affect the results.

34

Page 36: Numerical Modelling of Soot Formation

Chapter 2

Simulation at Higher Pressures

In the first paper on DSMC methods (also known as the Direct Simulation Al-

gorithm, DSA) for simulating soot particle populations in flames Balthasar and

Kraft [16] only considered flames at atmospheric pressure. This was so that the

Knudsen number (the mean free path divided by the diameter) of the soot parti-

cles would satisfy Kn > 10, that is, the particles would be small compared to the

mean free path. In this chapter a physical model for coagulation at lower Knudsen

numbers, taken from published literature, is implemented within the DSA. Inclu-

sion of the low Knudsen number model in the algorithm required a generalisation

of the stochastic approach used in the original paper [16], and that generalisation

is presented and tested here. The result of the work reported in this chapter was

to establish a numerical method and a particular computer implementation of it,

which could then be used as the starting point for the work in the rest of the thesis.

The main reason for the use of DSMC methods in the simulation of soot

growth was to handle the non-linear coagulation term defined in (1.5), which

represents the process by which soot particles collide and stick together. As in

the initial paper [16], coagulation is treated as completely coalescent throughout

this thesis with the exception of chapter 5. This means that when two spherical

particles coagulate all the material from both incoming particles is assumed to be

converted into one new spherical particle. Coagulation is sometimes used to refer

to a non-coalescent process, in which the incoming particles initially remain in

35

Page 37: Numerical Modelling of Soot Formation

point contact; this is the focus of chapter 5.

2.1 Coagulation Kernel

In the physical model used in [16, 92] the probability of coagulation between two

particles in a unit of time is proportional to the coagulation kernel (a symmetric

function of both particles). In [92] the standard coagulation kernels, Kfm for

the free molecular regime1 (Kn > 10) and Ksf for the slip flow and continuum

regimes2 (Kn < 1) are used and the Fuchs coagulation kernel [62], which is

applicable at all Knudsen numbers, is quoted. In the same paper it is shown that

the Fuchs kernel differs only slightly from one half3 of the harmonic mean of

the slip flow and free molecular regime kernels. The Fuchs and harmonic mean

kernels are considered in more detail in [153] where both are shown to be part of

a more general family of ‘flux matching’ kernels.

The calculations in [92] use the Method of Moments with Interpolative Clo-

sure (MoMIC) [57], so neither the harmonic mean nor the Fuchs kernel can be

used directly because of their relatively complicated form. In [92] Kazakov and

Frenklach adapt the harmonic mean idea to their very fast MoMIC numerical tech-

nique by calculating two coagulation rates, one with the free molecular kernel and

one with the slip flow kernel and taking the harmonic mean of these two rates.

This is different from using the harmonic mean kernel, but they found it worked

well. An important reason for this success is that, as can be seen in the figures

of [92], the free molecular and slip flow kernels become very large outside their

domains of applicability. In a harmonic mean a very large coagulation rate calcu-

lated from an inapplicable kernel has very little effect and so the correct behaviour

is recovered for the slip flow and free molecular regimes.

In this work the harmonic mean approach is applied to the coagulation kernels

1The form of the free molecular kernel is given in (5.4)2The form of the continuum kernel, which is very similar to that of the slip flow kernel, is given

in (5.3)3This factor of 2 is not mentioned in the text in either [92] or [62], but the equations in those

papers are correct. The factor will not be referred to again here.

36

Page 38: Numerical Modelling of Soot Formation

and so a transition kernel is defined by

Ktr =

(1

Ksf+

1

Kfm

)−1

(2.1)

for all Knudsen numbers, in particular the previously inaccessible range

1 ≤ Kn ≤ 10.

2.1.1 Majorant Kernels

At the heart of the DSMC approach to simulating coagulation is the repeated se-

lection of pairs of particles xi, xj i 6= j from a computational population

x1, x2, . . . , xn, which statistically represents the physical system. The selec-

tion has to be done so that the probability of a pair being selected is

K (xi, xj)

R, R =

∑i6=j

K (xi, xj) , (2.2)

where K is the coagulation kernel and the inverse of the mean time step length is

sum R.

A naıve approach to this selection will have a run time proportional to N2

whereN is the size of the computational population. To avoid this rather crippling

time cost majorant kernels [49] can be used. A majorant is a function K ≥ K for

which the computation of

R =∑i6=j

K (xi, xj) (2.3)

and the inversion for particle selection purposes of the distribution

K (xi, xj)

R(2.4)

is relatively fast (run timeO (N) orO (N logN)). It is then possible to recover the

distribution and rate defined by K by rejecting the selection of a pair of particles

37

Page 39: Numerical Modelling of Soot Formation

xi, xj with probability

1− K (xi, xj)

K (xi, xj)(2.5)

which is O (1).

Balthasar and Kraft [16], who only considered the free molecular regime, used

the majorant derived in [68] and which will be referred to as Kfm. The very sim-

ple form of Ksf means no majorant is needed for the slip flow regime. However,

the slightly opaque form for the transition regime coagulation kernel given by

(2.1) makes it difficult to construct an efficient majorant for the transition regime.

In the problem considered in [68] the kernel could be factorised into two sep-

arate parts for which tight upper bounds were obtainable. No such ‘divide and

conquer’ strategy seems possible for the transition kernel defined in (2.1) so a

more general mathematical approach was adopted. The new approach, of which

majorant kernels are a special case, will be termed ‘majorant rates’ since it is de-

rived by considering the overall coagulation rates rather than the kernel which is

evaluated for pairs of particles.

Define

Rsf =∑i6=j

Ksf (xi, xj) , Rfm =

∑i6=j

Kfm (xi, xj) , (2.6)

and note that for any a, b > 0,(1

a+

1

b

)−1

≤ min (a, b) (2.7)

and so using (2.1),

Ktr (xi, xj) ≤ min(Ksf (xi, xj) , K

fm (xi, xj)). (2.8)

Since Kfm ≤ Kfm the latter may replace the former in (2.8) to give

Ktr (xi, xj) ≤ min(Ksf (xi, xj) , K

fm (xi, xj)), (2.9)

38

Page 40: Numerical Modelling of Soot Formation

hence ∑i6=j

Ktr (xi, xj) ≤∑i6=j

min(Ksf (xi, xj) , K

fm (xi, xj))

≤ min

(∑i6=j

Ksf (xi, xj) ,∑i6=j

Kfm (xi, xj)

)= min

(Rsf , Rfm

).

(2.10)

Therefore, it is convenient to define the ‘majorant rate’

Rtr = min(Rsf , Rfm

). (2.11)

This is the rate at which potential coagulation events are generated and, because

it is the maximum of two quantities, which are easy to calculate using basic ma-

jorant kernels, it is suitable for use in speed critical parts of simulations. With

this majorant rate the preliminary selection of a particle pair is according to the

distribution

Ksf (xi, xj)

Rtrwhen Rtr = Rsf and

Kfm (xi, xj)

Rtrotherwise. (2.12)

In either case the selection only depends on Ksf and Kfm, to which standard

majorant kernel techniques can be applied, see [68] for details.

The preliminary selection is accepted with probability

Ktr (xi, xj)

Ksf (xi, xj)when Rtr = Rsf and

Ktr (xi, xj)

Kfm (xi, xj)otherwise (2.13)

and since Ktr only has to be evaluated for a single pair of particles the computa-

tional cost is not significant.

The case when the preliminary selection is rejected is known as a fictitious

jump (as referred to in [16])—time is advanced but no change is made to the com-

putational population. With this extension of the majorant concept the algorithm

set out in [16] is applicable at the higher pressures considered here. Neither ma-

39

Page 41: Numerical Modelling of Soot Formation

jorant kernels nor the extended concept of majorant rates introduced here involve

any approximation in the solution of the model system being solved. The results

of Eibeck and Wagner [51] show, that subject to limits associated with machine

precision, any desired precision may be achieved by an appropriate choice of nu-

merical parameters. For very high precisions the appropriate choice of numerical

parameters will lead to very large demands, but they will be finite.

An implementation of these techniques using the Fortran 90 programming lan-

guage is available for download and use under the Gnu General Public Licence

[29].

2.1.2 Validation

A range of problems were devised for which direct solution using an ODE solver

was possible (see §1.3.1). These problems were used to test the implementation

of the stochastic method by comparing its solutions with those obtained from

the direct ODE solver. Results are presented for two of these test problems—

a simulation of pure coagulation from a monodisperse initial particle population

(6.56×1011 particles per cm3 of diameter 0.88 nm) in air and one with no particles

at time 0, but with 0.88 nm particles being incepted at a rate per unit volume of

2.63× 1013 ×(

0.05− t0.05

)2

cm−3s−1

at time t. The other details of the test problems were:

• Particles coalesced immediately to form spheres upon collision;

• Particle density of 1.8 g cm−3;

• Constant temperature and pressure of 500 K and 600 bar respectively;

• Approximate range of Knudsen numbers: 5–20;

• No chemical reactions on particle surfaces considered.

40

Page 42: Numerical Modelling of Soot Formation

105

106

107

108

109

1010

1011

0 1 2 3 4 5

Coagulation only

Stochastic 0.02 sLSODE 0.02 sStochastic 0.05 sLSODE 0.05 s

num

ber

dens

ity /

cm-3

particle diameter / nm

105

106

107

108

109

1010

1011

1012

0 1 2 3 4

Nucleation and coagulation

Stochastic 0.02 sLSODE 0.02 sStochastic 0.05 sLSODE 0.05 s

num

ber

dens

ity /

cm-3

particle diameter / nm

Figure 2.1: Size distributions calculated with ODE solver and stochastically.

41

Page 43: Numerical Modelling of Soot Formation

Figure 2.1 shows a very encouraging degree of agreement between the re-

sults produced by the stochastic code using the method outlined above and direct

solutions calculated with the Livermore ODE solver (LSODE) [163]. Since the

number of particles in each size class is predicted correctly this is strong numer-

ical evidence for the validity of the new majorant approach and indeed the com-

plete simulation algorithm. The mathematical properties of the simulation results

are unaffected by the extension of the majorant approach. While majorants are

practically essential, they are no more than computational conveniences and the

simulated coagulation rate does not depend on the majorant.

ODE solvers such as LSODE cannot be used to obtain particle distributions

such as those plotted in figure 2.1 for real flames because they have run times that

scale quadratically with the number of particle sizes considered. In the test prob-

lems shown the number of particle sizes was restricted to 2000, for real flames

millions of different sizes would have to be considered and run times on the desk-

top computers (CPU speeds around 2 GHz) available for this study would have

been measured in years!

2.2 Application to Flames

Results for two 10 bar laminar, premixed ethylene flames JW10.68 and JW10.60

[1, 93] are shown in figure 2.24 along with results from the MoMIC calculations

reported in [8] and experimental data for the soot volume fraction fv also from

[8]. The same data for the 1 bar laminar, premixed ethylene flame XSF1.78 [230]

is also shown. All three sets of stochastic calculations made use of the harmonic

mean kernel (2.1) and the majorant approach of §2.1.1. The data in figure 2.2

shows the stochastic algorithm produced values within a factor of 2 of those gen-

erated by the established MoMIC technique. The soot model parameters used in

the simulations are the values reported in [8] as optimal (for the MoMIC).

From figure 2.2 one sees that the physical model leads to results that differ no-

ticeably from the observed data and that this is true whichever numerical method

4Figure 2.2 is courtesy of Dr J Singh

42

Page 44: Numerical Modelling of Soot Formation

109

1010

1011

10-9

10-8

10-7

0 0.05 0.1 0.15

JW 10.60

Stochastic

MoM

Experimental

num

ber

dens

ity /

cm

-3 soot volume fraction

time / s

109

1010

1011

1012

10-9

10-8

10-7

10-6

0 0.05 0.1 0.15

JW 10.68

Stochastic

MoM

Experimental

num

ber

dens

ity /

cm-3 soot volum

e fraction

time / s

109

1010

1011

10-9

10-8

10-7

0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1

XSF 1.78

Stochastic

Experimental

MoM

num

ber

dens

ity /

cm-3 soot volum

e fraction

time / s

Figure 2.2: Comparison of calculations and observations for flames.

43

Page 45: Numerical Modelling of Soot Formation

is used for the calculations. This deviation illustrates very clearly the limitations

of the current understanding of soot formation mechanisms. As discussed in the

introduction, kinetic data for gas phase reactions are not precise and the processes

on the surface of soot particles are only just beginning to be modelled in detail. In

particular, the surface activity of soot particles is known to decrease as particles

move through flames, but no physical model is available and some sort of corre-

lation with size [8] or age [178] has to be used instead. Also, while the MoMIC

and the stochastic data are qualitatively the same and close on a log scale, there

are some differences. Some divergence was to be expected because DSA requires

rates to be expressed as functions of the properties of individual particles, whereas

MoMIC ultimately requires all rates to be expressed in terms of the lower integer

order moments of the particle mass distribution, and these two requirements are

not entirely compatible.

Computationally MoMIC is extremely cheap and only the very recent emer-

gence, after this work was completed, of the DQMoM (see §1.3.2) has raised

questions about its position as the method of choice for incorporation in larger

simulations. However MoMIC, by its nature, cannot produce the full size distri-

butions of particle populations such as are plotted in figure 2.1, and which are

likely to be of considerable interest to manufacturers of carbon black and to those

seeking to understand pollution from diesel engines.

The scope for improvements in the physical model of the soot growth pro-

cess is also clear and the stochastic Direct Simulation techniques outlined here

and by Balthasar and Kraft [16] (the paper this work extends) offer considerable

advantages for those working on such models. The most important of these is

that arbitrarily precise solutions of the model equations are possible [51]. In ad-

dition the convergence result in [51] shows that by varying the parameters of the

numerical method (not the physical model) one can estimate the numerical error

and then control it. Secondly the representation of the particle distribution by a

sample of particles means that simulation of processes at the individual particle

level is possible. Interactions between particles and the surrounding system may

be simulated using the kinetics for those particular particles rather than average

44

Page 46: Numerical Modelling of Soot Formation

quantities as in moment based methods, see for example [178]. Within the DSA

framework extremely complex descriptions of the internal particle structure can

also be used e.g. [17]. However, this power of the DSMC framework comes at a

cost which can be very noticeable—if a flame has many physical events involving

soot particles then a simulation of the flame will be slow. Details of the problem

and strategies for dealing with it form a considerable part of this thesis, but first

some more detail about the basic computational implementation is given.

2.3 Implementation of DSA

The key program features introduced here are used in all the Monte Carlo simula-

tions reported in this thesis. Some detail about the binary tree is given because of

differences from the implementation outlined by Wells and Kraft [213].

2.3.1 Variation in Number of Particles

In a direct simulation on a computer of a chemical system where the number of

computational particles may increase and decrease, one has to bound the number

of these particles being tracked. In soot formation problems, decreases in parti-

cle count are mainly due to coagulation. Increases in particle count are the result

of particle inception. One approach to controlling the number of computational

particles is given by Smith and Matsoukas [181], in which a constant number is

maintained by resampling the population every time the number of computational

particles changes. Alternatively one can allow the number of computational parti-

cles to drop below 50% of the program capacity and then duplicate all the particles

[110, 114]. A discussion of this issue may be found in the introduction of Zheng

[243].

Particle doubling has the attraction of preserving all the statistical properties

of the computational population, which is not generally possible when resampling

multiplies the number of computational particles by a non-integer factor. During

the work reported here a constant computational particle number algorithm was

tried and found to have the expected, accurate mean behaviour. However, the

45

Page 47: Numerical Modelling of Soot Formation

constant computational particle number method led to a much greater statistical

uncertainty (by around 1 order of magnitude in the first few moments of the sim-

ulated distribution) in the results compared with using particle doubling. This is

consistent with the results of Maisels et al. [114], therefore only particle doubling

was used for the remainder of the work. In the case where the computational

particle number rises beyond the upper limit set for a computation any rescaling

has to be done by a non-integer factor (since it must be strictly between 0 and 1).

For handling this case it is convenient to resample as soon as the computational

particle limit is exceeded by one since there are no benefits in allowing a larger

increase.

2.3.2 Binary Tree

The basic algorithm for the simulation of particle formation and growth is de-

scribed in [68, 177]. Unlike in those papers, a binary tree method was used here

for particle selection according to the distributions defined in (2.12), as this en-

abled the use of more complex soot models.

The ensemble of stochastic particles was stored in a list connected to a binary

tree. The tree provided a convenient method for selecting particles according

to arbitrary probability distributions. In an earlier version, described briefly by

Wells and Kraft [213], accumulated error problems had to be explicitly addressed.

The version presented here avoids these problems at the cost of using double the

amount of memory, but since the program has only modest memory requirements

this does not cause any difficulties.

Figure 2.3 shows a binary tree along with the numbering scheme used for the

nodes. Each node is shown as a pair of adjoining boxes and, except for nodes on

the bottom level, each has two ‘children’ beneath it to which it is joined by lines.

The figure should also make clear the concept of levels. The tree shown has 4—

the maximum number of nodes on the path from the top of the tree (also called

the root) to any point on the bottom (called a leaf). In this section the number of

levels in a tree will be denoted by l. The numbering scheme is not fundamental

to the operation of the tree but having some simple scheme is necessary for im-

46

Page 48: Numerical Modelling of Soot Formation

1F 1

,left

F 1,rig

ht

3F 3

,left

F 3,rig

ht

2F 2

,left

F 2,rig

ht 5

F 5,l

F 5,r

11

F 11

,lF 1

1,r

10

F 10

,lF 1

0,r

4F 4

,lF 4

,r

9F 9

,lF 9

,r

8F 8

,lF 8

,r

7F 7

,lF 7

,r

15

F 15

,lF 1

5,r

14

F 14

,lF 1

4,r

6F 6

,lF 6

,r

13

F 13

,lF 1

3,r

12

F 12

,lF 1

2,r

Tre

e

f 1 1

f 2 2

f 3 3

f 4 4

f 5 5

f 6 6

f 7 7

f 8 8

f 9 9

f 10

10

f 11

11

f 12

12

f 13

13

f 14

14

f 15

15

f 16

16

List

Figure 2.3: 4 level binary tree

47

Page 49: Numerical Modelling of Soot Formation

plementation purposes. The advantage of this scheme is that for a node numbered

i > 1 it is easy to calculate the number of its parent as bi/2c, the integer part of

i/2. The children (if they exist) of a node numbered i are therefore numbered 2i

and 2i+ 1.

Suppose one has a tree with l > 1 and a list of less than 2l stochastic particles

numbered 1, 2, ...,m and associated with the particle numbered i is a weight fi.

Suppose further one wishes to select a particle at random from the list in such a

way that the probability of selecting particle i is

fi∑mi=1 fi

. (2.14)

The tree, as shown in figure 2.3, should be initialised from the bottom up in the

following way, starting with the bottom level (nodes numbered 2(l−1) to 2l − 1).

F2(l−1)+b i−12c,left =fi, i ≤ m, i odd

F2(l−1)+b i−12c,right =fi, i ≤ m, i even

F2(l−1)+b i−12c,left =0 m < i ≤ 2l, i odd

F2(l−1)+b i−12c,right =0 m < i ≤ 2l, i even

and then working upwards level by level setting

Fj,left =F2j,left + F2j,right

Fj,right =F2j+1,left + F2j+1,right.

A particle i is said to be under a leaf node j and to its left if fi is assigned to Fj,left

in the above procedure; ‘under to the right’ is to be understood analogously.

A tree constructed in this way has the property that, for j < 2l−1, Fj,left

and Fj,right are the relative probabilities according to the distribution in (2.14) of

selecting a particle under one of the descendants of j via its left or right hand child

48

Page 50: Numerical Modelling of Soot Formation

respectively. The algorithm for particle selection is therefore simple:

1. j ← 1

2. Generate a U (0, Fj,left + Fj,right) random variable, r.

3. If node j is a leaf, break out of the loop by going to stage 6

Else continue

4. If r < Fj,left choose the left child and so

j ← 2j

Else choose the right child so that

j ← 2j + 1

5. Loop back to stage 2

6. If r < Fj,left select the particle under node j and to the left

Else select the particle under node j and to the right

Generating one random number is sufficient; if at stage 4 one selects the right

hand branch one can simply update r by r ← r − Fj,left before updating j and

use the new value of r in place of the random variable that would be generated in

stage 2.

Obviously having to completely rebuild the tree each time one wants to select

a particle would make this a very inefficient algorithm. However, once a tree

has been initialised, if one of the fi changes or an extra particle is added without

exceeding the capacity of the tree (2l particles), the tree can be updated simply by

recalculating the values of F on the path from the leaf under which the change has

occurred to the root. It is generally necessary to maintain binary trees for several

different distributions simultaneously. This was done by using vector valued f

and F .

2.3.3 Complexity of the implementation

For a binary tree with l levels (capacity 2l particles) each event requires O (l)

operations. One has to descend the tree to select a particle then ascend updating

49

Page 51: Numerical Modelling of Soot Formation

with the results of the event.

The number of events scale linearly with the sample volume, V , as do the

number of particles. Hence for reasonable choices of V and l (those for which the

full capacity of the tree is used) a simulation will need a tree withO (log V ) levels

and will have a run time that is O (V log V ).

The performance of a simulation program, which used a binary tree as de-

scribed above, was measured for different sample volumes and computational

particle numbers. In figure 2.4 run times can be seen to scale as n log n in the

Table 2.1: Scaling of run times with tree depth

log2 Max Particles run time / s12 11313 24014 54415 122416 273017 5724

maximum number of computational particles and, since the sample volumes were

chosen to be proportional to n, these results are consistent with the complexity

derived above.

2.3.4 DSA Applied to Flames

This section focuses on computations for two premixed laminar ethylene flames

studied in [93]—JW1.69 and JW10.68. The first is a 1 bar5 flame with C/O ratio

0.69 and a peak soot volume fraction of approximately 2 × 10−8, the second a

10 bar flame with C/O ratio 0.68 and a peak soot volume fraction of approximately

2× 10−6. These soot volume fractions are at the extremes of the range studied in

[93].5All pressures reported in this thesis are absolute pressures.

50

Page 52: Numerical Modelling of Soot Formation

Simulations were performed using different numbers of computational parti-

cles and varying numbers of repetitions. These gave some idea of the (generally

small) systematic error in moments 0 to 3 of the final size distribution due to using

a low number of computational particles and the number of repetitions needed to

control the statistical noise. Some sample results are given in table 2.2. The initial

sample volume used in the simulations is denoted by “vol”, “m2” refers to the

second moment of the mass distribution, the uncertainty is the half width, using a

central limit theorem estimate, of the 95% confidence interval for the mean of the

distribution being sampled (that of m2).

Table 2.2: Illustrative results

flame vol / cm3 treerepetitions

m2 timedepth uncertainty / min

JW1.69 3.5× 10−8 13 10 2.3% 50JW1.69 4.375× 10−7 10 30 6.8% 16.5

JW10.68 2.5× 10−7 11 20 1.5% 4020JW10.68 1.25× 10−7 10 20 2.9% 1534

The run times for JW10.68 indicate that DSA, as implemented, is not a very

practical tool for investigating sooty flames on a desktop computer. In order to

inform attempts to reduce the run times, execution time profiling was carried out

to identify the bottlenecks in the program.

2.3.5 Profiling of DSA Simulation

Table 2.3 shows the number of each type of event performed during 10 runs of

a simulation of the premixed flame JW1.69 (see [93]) with 16 tree levels and an

initial sample volume of 2.8 × 10−7 cm3. JW1.69 was used rather than JW10.68

because profiling leads to a very large increase in run times and so was not fea-

sible for the JW10.68 case. Unlike for other surface reactions, the rate of pyrene

condensation does not have a simple dependence on particle surface area. Pyrene

condensation was therefore handled separately in the simulations so that the rel-

51

Page 53: Numerical Modelling of Soot Formation

0

50

100

150

200

250

300

350

0 20 40 60 80 100 120 140

run

time

/ log

(m

ax c

ompu

tatio

nal p

artic

les)

max computational particles / 1000

Figure 2.4: Simulation run time scaling with binary tree

Table 2.3: Relative frequency of stochastic events in JW1.69

time steps 1.25× 109

surface events (not pyrene) 1.20× 109

pyrene condensation 4× 107

coagulation 5× 106

inception 4× 106

52

Page 54: Numerical Modelling of Soot Formation

evant rate models could be properly implemented. The main conclusion to be

drawn from table 2.3 is that essentially all the events are surface events and that

there are not many pyrene events compared to the other events. One may also note

that coagulation events are even more rare than pyrene events.

Figure 2.5 shows how this breaks down in computational time. The data were

obtained over 5 runs simulating JW1.69 with a tree depth of 12 and an initial

sample volume of 1.75 × 10−8 cm3. From figure 2.5 one sees that 64% of the

selection of event7%

other time step overheads

17%

performing surface events31%

calculating stochastic step

length40%

unclassified3%

pyrene events2%

Figure 2.5: % of execution time spent on different tasks—DSA

computational effort is consumed in calculating the random time steps. These

parts of the program are key to the simulation of a non-linear coagulation process,

but are not fundamental for the surface processes. The surface growth of a single

particle is simply a non-homogeneous Poisson process until the next time that

particle is involved in a coagulation and which, conditional on the coagulation

time, does not depend on any other particles. However, most of the time steps

53

Page 55: Numerical Modelling of Soot Formation

reported in table 2.3 are concerned with the surface processes and not coagulation.

A considerable amount of the 31% of time spent on “performing surface events” is

also attributable to the binary tree structure necessary for simulating coagulation.

Alternative ways of incorporating the surface area processes into simulations are

therefore investigated in the following chapters.

54

Page 56: Numerical Modelling of Soot Formation

Chapter 3

Accelerating the Surface Processes

The previous chapter outlined the extension of the DSMC method known as DSA

into a general soot simulator. This chapter and the next describe work done to

make the DSA a practical tool for use on a desktop PC. The focus in this chap-

ter is on two ways in which the natural random nature of the surface processes,

which take up so much of the computation time in the basic DSA as introduced

in chapter 2, can be preserved while reducing their computational demands. The

possibility of a deterministic approximation to these processes will be investigated

in chapter 4.

3.1 Operator Splitting

In (1.4) the evolution of the particle distribution is seen to be driven by three kinds

of process:

1. Coagulation—two particle events

2. Single particle events

(a) Surface area processes

(b) Pyrene condensation

3. Particle inception—involves no pre-existing particles

55

Page 57: Numerical Modelling of Soot Formation

From the analysis in chapter 2 it is clear that most of the time taken in a simulation

is spent updating the binary tree after events under 2(a). Operator splitting [69]

is a numerical approach to solving differential equations that allows for the intro-

duction of different methods to account for the effects of single particle events.

3.1.1 Mathematical Outline of Splitting

Consider an ordinary differential equation (ODE) with a source term that can be

expressed as the sum of two operators A and B , for example,

d

dtf = (A+ B) f

f (0) =f0.

(3.1)

Assume further that (3.1) has a unique, well behaved solution. In this case the

simplest way to compute the solution over time is based on the idea that for small

h

f (t+ h)− f (t) ≈ h (A+ B) f (t) .

This is the basis of DSA where all the different processes are simulated simultan-

eously—the jump rate of every process depends on the current state of the particle

ensemble.

However, at the cost of introducing a further approximation and by assuming

that the resulting ODEs have well behaved solutions, one can define auxiliary

functions f1, f2 and advance the solution over a small time interval [t0, t0 + δt]

in two stages, using whatever numerical schemes are appropriate for each stage.

First solve

d

dtf1 =Af1

f1 (t0) =f (t0)

(3.2)

56

Page 58: Numerical Modelling of Soot Formation

and then

d

dtf2 =Bf2

f2 (t0) =f1 (t0 + δt) .

(3.3)

The key point is that completely different numerical methods may be used to solve

(3.2) and (3.3).

3.1.2 Implementation of Operator Splitting

Equation (1.4) was split into two parts by handling OH oxidation and acetylene ad-

dition (sometimes also the insignificant process of O2 oxidation) separately from

all the other processes, which are simulated in exactly the same way as in the

DSA. Typically this splitting was performed over time intervals of a length such

that each particle would, on average, undergo a few surface reactions during a

splitting step. The choice of this average number was the means of determining

the length of the splitting step and thus the accuracy and computational speed of

the approximation.

The steps in advancing a simulation over times t from t0 to t1 are 1:

1. t← t0, t2 ← t0

2. WHILE(t < t1)

(a) Choose a step length 4t and set t2 = t + 4t. In this work 4t was

chosen as the time in which n × r events due to the split processes

would be expected under DSA. The number of computational parti-

cles is n and r is an arbitrary positive parameter, which is chosen to

optimise accuracy and computational speed. Various conditions may

lead to large values of4t being calculated; these must be reduced.

(b) Perform the unsplit processes including coagulation from t to t2 ex-

actly as for DSA using the binary tree.

1t2 and t3 are auxiliary times

57

Page 59: Numerical Modelling of Soot Formation

(c) Extract a list of particles from the binary tree.

(d) WHILE(t < t2)

i. Choose a secondary time step length δt such that the chemical

conditions vary only slightly over the time interval [t, t+ δt] and

t+ δt ≤ t2.

ii. t3 ← t+ δt

iii. Simulate the split processes up to t3 for each particle. Methods

for doing this are discussed below.

iv. t← t3

v. Return to top of the loop—2d

(e) Put the updated computational particles back into the binary tree. (One

should now have t = t2.)

(f) Return to the top of the loop—2

3. Stop

Simulating Split Events for One Particle

As a first step, the method used in DSA was applied to individual particles. The

inhomogeneous Poisson process referred to in §2.3.5 was generated by repeatedly

calculating the exponentially distributed waiting times and performing events at

their end points. This method is referred to as ‘pp’ (Poisson Process) in later

sections.

In order to reduce further the computational effort a second level of splitting

was introduced. By dealing with the split processes one by one and making a

further approximation, it is only necessary to generate one random variable per

process per particle as follows:

1. Calculate the rate for the selected process at the start of the interval.

58

Page 60: Numerical Modelling of Soot Formation

2. Assume this rate applies throughout the interval (e.g. ignoring the effects of

the particle size changing) and so calculate the expected number of events

λ.

3. Generate a random variable with the Poisson (λ) distribution as the num-

ber of events which actually occur and update the computational particle

accordingly.

This method is referred to as ‘pv’ (Poisson Variable) in later sections.

Complexity

The split surface processes were performed without using the binary tree; this

avoids the O (log n) update operations (n is the number of stochastic particles,

which should be roughly the same as the capacity of the binary tree) after each

event that were slowing the simulations down. Instead, after performing surface

reactions, an O (1) operation on each of the n ∼ O (V ) particles individually for

the full length of the splitting time step, the binary tree was completely rebuilt

which takes O (n) ∼ O (V ) time.

Therefore the surface processes, which previously had overall complexity

O (V log V ) in the sample volume V , now only had complexity O (V ), which

is an improvement over the complexity reported in §2.3.3. Since the surface pro-

cesses dominate the simulation (see table 2.3), this should reduce the apparent

complexity of the simulation.

3.2 Numerical Results with Splitting

3.2.1 Comparison of JW1.69 Moments

Some results, which were generated using the implementations of operator split-

ting described above, are presented for JW1.69 (one of the laminar premixed

flames studied in [93]). They are compared to DSA results generated as described

in chapter 2. JW1.69 was used because it provided a realistic test case that could

be simulated quickly and also for historic reasons.

59

Page 61: Numerical Modelling of Soot Formation

Table 3.1 gives a summary of the simulations performed and the computer

time required for their calculation. All timed simulations were performed using a

binary tree of depth 13 and with an initial sample volume of 3.5× 10−8cm3. The

computer used had a 2.4 GHz Pentium 4 CPU, except in the cases marked with ‘†’which indicates a 2 GHz Athlon CPU, which gave almost identical performance

on a test case performed on both machines.

Table 3.1: Detail of operator splitting runs for JW1.69

algorithm r time for 10 repetitions / minDSA 0 50pv 1 30pv 2 16pv 5 8pv 5 8†pv 10 5.5†pp 5 12.5†pp 10 10†

With the ‘pv’ algorithm one sees a considerable speed increase moving from

r = 0 (DSA) to r = 5 with speed being almost linear in r, but changing r from

5 to 10 makes much less difference to the execution time. One therefore expects

the bottleneck to be the inner loop of the algorithm, see §3.1.2. The ‘pp’ splitting

leads to longer run times than the ‘pv’ because each surface event is handled

individually.

However, speed is not everything; one has also to consider the error introduced

by the splitting approximation, so in figure 3.1 the time evolution is shown for

the second moment of the particle distribution in a small volume of gas passing

through the flame. To show the error introduced by the splitting results obtained

using DSA with 30 repetitions, a binary tree depth of 17 and an initial sample

volume of 5.6 × 10−7cm3 are also plotted. The numerical parameters used to

generate the DSA results were chosen to make computational errors negligible on

the scale shown. Figure 3.1 shows that the splitting only introduces a very modest

60

Page 62: Numerical Modelling of Soot Formation

0

0.5

1

1.5

2

2.5

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 0.02 0.04 0.06 0.08 0.1 0.12

dsa 17

pv 2

pv 10

pp 10

parti

cle

num

ber d

ensi

ty /

1011

cm

-3second m

ass mom

ent / 10-24 g

2cm-3

time / s

Figure 3.1: Number density and second mass moment for JW1.69

61

Page 63: Numerical Modelling of Soot Formation

error in the second moment of the distribution, at least for r ≤ 5, compared to the

high precision DSA. For lower order moments the error is smaller.

3.2.2 Particle Distribution Accuracy

An advantage of direct simulation is that one gets a complete description of the

particle size distribution. Particle size distributions are presented here for the

flame JW10.68 (see [93]). For comparison with table 2.3 an event count for

JW10.68 is given in table 3.2. These are the number of events in 20 repeti-

tions of the simulation, with a tree depth of 10 and an initial sample volume of

1.25× 10−9 cm3.

Table 3.2: Relative frequency of stochastic events in JW10.68

time steps 2.06× 109

surface events (not pyrene) 2.05× 109

pyrene condensation 1× 107

coagulation 4× 105

inception 2× 105

JW10.68 is a much sootier flame than JW1.69 so it provides a test of the

method in a situation where soot is really significant and, as can been seen, it is

also far more computationally demanding. Table 3.3 gives details of simulations

performed using a binary tree depth of 11 and an initial sample volume of 2.5 ×10−9cm3 on the PCs described above.

The acceleration due to splitting is much larger for JW10.68 than for JW1.69

(see table 3.1)—a factor of 30 compared to one of 6 for DSA to ‘pv’ r = 5.

Further increases in r were found to offer much smaller speed gains with the ‘pv’

algorithm, while the pp algorithm times appear to have only a weak dependence

on r, suggesting the time limiting step is the event by event simulation of the split

processes.

Figure 3.2 shows the particle distributions obtained from some of the simu-

lations described in table 3.3. Also shown is the average from a set of 10 high

62

Page 64: Numerical Modelling of Soot Formation

Table 3.3: Detail of operator splitting runs for JW10.68

algorithm r time for 10 repetitions / minDSA 0 4020pv 5 134pv 10 86pp 20 250pp 50 235

precision DSA runs using a binary tree depth of 14, these took 13 days of PC

time to generate! This last set of data is shown so that any error introduced by the

splitting procedure can be seen.

3.2.3 Profiling of Operator Splitting

Operator splitting yields a considerable acceleration from DSA—see table 3.3,

but the computation times are still substantial and would be noticeable even with

1000 stochastic particles (tree depth of 10). Therefore it makes sense to repeat

the analysis of §2.3.5 and see if further improvements can be made. Figure 3.3

shows how the computation time is used during a ‘pv’ algorithm run with the

parameters used for the DSA profiling in §2.3.5. “Poisson processes” refers to the

approximation of the surface event Poisson processes by Poisson random variables

for each particle during the splitting steps.

From these data one can see that simulating the surface reactions is still by

far the most computationally intensive task in the simulation even for JW1.69,

a flame with relatively little surface growth compared to a sooty flame such as

JW10.68. Note also that a substantial amount of time is spent approximating the

Poisson processes that represent the surface reactions during splitting. A further

approximation is therefore developed in §3.3.

63

Page 65: Numerical Modelling of Soot Formation

20 40 60 80 100 120 140 16010

5

106

107

108

sphere equivalent diameter / nm

num

ber

dens

ity /

cm−

3

referenceDSApv r=5pv r=10

Figure 3.2: Particle size distribution at end of flame JW10.68

64

Page 66: Numerical Modelling of Soot Formation

rebuilding binary tree30%

pyrene events5%

other (splitting related)

20%

unclassified5%

Poisson processses

40%

Figure 3.3: Percentage of execution time spent on different tasks with ‘pv’ method

3.2.4 Conclusions on Operator Splitting

Operator splitting using the ‘pp’ and ‘pv’ methods provides a substantial perfor-

mance gain from DSA, in the most striking case reducing a run time by a factor

of almost 10. The approximations made to achieve this acceleration have a small

effect on the results and a particular consequence of this is that the approximation

error does not depend very strongly on the size of the time splitting steps (within

the range of values considered). Compared with the uncertainties in the simulation

of the gas phase chemistry (see for example Smooke et al. [182]), the systematic

error should not be a cause for concern.

3.3 Deferment of Surface Reactions

A new approach2 was to tag each stochastic particle with the time when it is added

to the ensemble so (stochastic or computational) particles are described by el-

ements of the space E = E × [0,∞). Some surface reactions (possibly now

2Due to a suggestion of Prof. J. Norris of the Statistical Laboratory, University of Cambridge.

65

Page 67: Numerical Modelling of Soot Formation

including pyrene condensation) can be selected for ‘deferment’, such reactions

are completely neglected as the simulation calculates time steps. Once a particle

has been selected for an event, the deferred events are simulated from the time

recorded in the tag up to the current time; only then does the non-deferred event

take place. The tag is then reset to the current time so that the procedure can be

repeated.

3.3.1 Measure Theoretic Formulation

In order to derive the generator for a stochastic jump process implementing this

idea it is helpful to work in terms of measures, replacing

n : [0,∞)× E → [0,∞)

by a family of measures giving the computational particle distribution on E at

each time.

To do this, endow E with a σ-algebra E that makes the g(l), the β(l)t , Kt, +

and I (t, ·), as defined below, measurable. The R valued functions Kt and β(l)t

are extended to E by saying they ignore the time stamp component; g(l) and +

(strictly g(l)t and +t) by saying the time stamp of their result is the current time.

Let M be the space of measures on E with the weak topology and the associ-

ated Borel σ-algebraM. The particle distribution is described by a mapping from

time to M

λ : [0,∞)→M ; t 7→ λt = λ (t, ·) .

So that for A ∈ E the number of particles per unit volume described by ele-

ments of A at time t will be λt (A).

The particle inflow rate will be given by a similar map

I : [0,∞)→M ; t 7→ It = I (t, ·) .

The rate of inflow of particles with descriptions in A at time t is It (A).

66

Page 68: Numerical Modelling of Soot Formation

Equation (1.4) can now be written for A ∈ E as

d

dtλt (A) =

∫x∈A

It (dx)

+∑l∈U

[∫x∈E:g(l)(x)∈A

β(l)t (x)λt (dx)−

∫x∈A

β(l)t (x)λt (dx)

]+

1

2

∫y,z∈E:y+z∈A

Kt (y, z)λt (dy)λt (dz)

−∫

x∈A,y∈E

Kt (x, y)λt (dx)λt (dy) .

(3.4)

Hence for φ in a suitable class of test functions C one has a weak form, (see

e.g. [51])

d

dt

∫x∈E

φ (x)λt (dx) =

∫x∈E

φ (x) It (dx)

+∑l∈U

∫x∈E

β(l)t (x)

[φ(g(l) (x)

)− φ (x)

]λt (dx)

+1

2

∫x,y∈E

[φ (x+ y)− φ (x)− φ (y)]

Kt (x, y)λt (dx)λt (dy) .

(3.5)

This includes the time stamp without really making use of it in λ, the repre-

sentation of the particle population.

3.3.2 Deferred Surface Process Operator

Now define a family of mappings Rt : t ∈ [0,∞) which map x ∈ E to a random

variable Rtx ∈ E × t ⊆ E with the distribution that a particle with description

x would have at time t in the absence of deferment. This only makes sense if t is

later than the time stamp in x.

Also define an ‘update operator’

Pt : M →M

67

Page 69: Numerical Modelling of Soot Formation

by

Pt (λ) (A) =

∫x∈E

P (Rtx ∈ A)λ (dx) . (3.6)

Let U ′ ⊆ U3 be the possibly empty set of surface reactions that are not de-

ferred. One now wants to define a ‘deferred solution’ λ via an equation like (3.5)

such that, for any φ ∈ C

Φ (λt) = Φ(Pt

(λt

))(3.7)

where

Φ (ν) :=

∫x∈E

φ (x) ν (dx) (3.8)

and λ solves (3.5).

A fairly obvious approach is

d

dt

∫x∈E

φ (x) λt (dx) =

∫x∈E

φ (x) It (dx)

+∑l∈U ′

∫x,ξ∈E

[φ(g(l) (ξ)

)− φ (x)

(l)t (ξ)P (Rtx = dξ) λt (dx)

+1

2

∫x,y,ξ,ζ∈E

[φ (ξ + ζ)− φ (x)− φ (y)]Kt (ξ, ζ)

P (Rtx = dξ)P (Rty = dζ) λt (dx) λt (dy) .

(3.9)

This defines a (or possibly some, since uniqueness is hard to establish) determin-

istic map from time to the soot population. However, while the resulting λ may

satisfy (3.7), there are real problems using it as a basis for a simulation. Firstly

there is the usual need to incorporate a majorant coagulation kernel (see e.g. [68]).

Secondly, even given a majorant, one still cannot handle individually the rates of

the infinity of possible jumps given by Rt.

Therefore an equation was developed that allows for the use of surface growth

generated for events that prove to be fictitious. The basic idea is to find β and K

3For the definition of U see (1.7)

68

Page 70: Numerical Modelling of Soot Formation

such that

β(l)t

(x; λt

)≥ β

(l)t (Rtx) (3.10a)

and

Kt

(x, y; λt

)≥ Kt (Rtx,Rty) (3.10b)

as it is then sufficient to calculate rates for a particle using a majorant without

knowing anything about deferred events. The deferred events can then be incor-

porated when another event is actually performed. In addition, if a particle is

selected for a potential event that eventually turns out to be fictitious, postponed

processes can be simulated on it anyway and the updated particle returned to the

computational ensemble thus reducing the deferment error.

Given the random nature ofRtx it is likely to be hard to find β and K to satisfy

(3.10) with probability 1. Therefore the equation 4

4Note: for a, b ∈ R a+ ≡ max (a, 0) and a ∧ b ≡ min (a, b)

69

Page 71: Numerical Modelling of Soot Formation

d

dt

∫x∈E

φ (x) λt (dx) =

∫x∈E

φ (x) It (dx)

+∑l∈U ′

∫x,ξ∈E

[φ(g(l) (ξ)

)− φ (x)

] [β

(l)t (ξ) ∧ β(l)

t

(x; λt

)]P (Rtx = dξ) λt (dx)

+∑l∈U ′

∫x,ξ∈E

[φ (ξ)− φ (x)][β

(l)t

(x; λt

)− β(l)

t (ξ)]

+

P (Rtx = dξ) λt (dx)

+1

2

∫x,y,ξ,ζ∈E

[φ (ξ + ζ)− φ (x)− φ (y)][Kt (ξ, ζ) ∧ Kt

(x, y; λt

)]P (Rtx = dξ)P (Rty = dζ) λt (dx) λt (dy)

+1

2

∫x,y,ξ,ζ∈E

[φ (ξ) + φ (ζ)− φ (x)− φ (y)][Kt

(x, y; λt

)−Kt (ξ, ζ)

]+

P (Rtx = dξ)P (Rty = dζ) λt (dx) λt (dy)

(3.11)

is formulated to make sense, even if the conditions in (3.10) are sometimes vi-

olated. Clearly (3.7) will not be satisfied if (3.10) do not hold, but provided this

violation only happens with very small probability one can hope to recover a close

approximation to (3.7).

3.3.3 Rate Kernel

Define jump operators on the space of signed Borel measures on E by

JVx,y,z,w; µ 7→ µ− V −1 (δx + δy − δz − δw) x, y, z, w ∈ E

and

JVx,y,z; µ 7→ µ− V −1 (δx + δy − δz) x, y, z ∈ E

70

Page 72: Numerical Modelling of Soot Formation

and

JVx,y; µ 7→ µ− V −1 (δx − δy) x, y ∈ E

and

JVx ; µ 7→ µ+ V −1δx x ∈ E,

where V is the sample volume for which events are to be simulated.

For t ∈ [0,∞) and V ∈ (0,∞) define kernels

αVt : M ×M→ [0,∞)

by specifying the value for µ ∈M, C ∈M

αVt (µ,C) =

∫x∈E

1JV

x µ ∈ CV It (dx)

+∑l∈U ′

∫x,ξ∈E

1

JV

x,g(l)(ξ)µ ∈ CV[β

(l)t (ξ) ∧ β(l)

t (x;µ)]

P (Rtx = dξ)µ (dx)

+∑l∈U ′

∫x,ξ∈E

1JV

x,ξµ ∈ CV[β

(l)t (x;µ)− β(l)

t (ξ)]

+

P (Rtx = dξ)µ (dx)

+1

2

∫x,y,ξ,ζ∈E

1JV

x,y,ξ+ζµ ∈ CV[Kt (ξ, ζ) ∧ Kt (x, y;µ)

]P (Rtx = dξ)P (Rty = dζ)µ (dx)µ (dy)

+1

2

∫x,y,ξ,ζ∈E

1JV

x,y,ξ,ζµ ∈ CV[Kt (x, y;µ)−Kt (ξ, ζ)

]+

P (Rtx = dξ)P (Rty = dζ)µ(2) (dx, dy)

(3.12)

where, for A,B ∈ E, µ(2) (A×B) = µ (A)µ (B)− µ (A ∩B) .

Using the kernels define generators for stochastic processes on M by

GVt (Ψ) (µ) =

∫ν∈M

[Ψ (ν)−Ψ (µ)]αVt (µ, dν) (3.13)

71

Page 73: Numerical Modelling of Soot Formation

for Ψ : M → R in a suitable class of test functions (not necessarily of the form

given in (3.8)) and assuming (3.10) one can approximate (3.11) as

d

dtΦ(λt

)= GV

t (Φ)(λt

)(3.14)

which has an error that is O (V −1) as V → ∞. By analogy with the work of

Eibeck and Wagner (see [48] for a basic introduction and [51] for their latest

generalisation), the stochastic jump processes µVt given by the generators GV

t con-

verge in distribution as V →∞ to a deterministic solution of (3.11) provided the

initial conditions also converge. One then expects Pt

(λt

)to converge in distribu-

tion to a deterministic solution, λt, of (3.5) in the same limit, that is, to be a weak

solution of the extended form of Smoluchowski’s coagulation equation (1.4).

3.3.4 Implementation of Deferment

The implementation of this Linear Process Deferment Algorithm (LPDA) is es-

sentially an application of DSA to the processes given in (3.13).

It is worth noting the procedure at the times (generally every millisecond)

when statistics on the size distribution of the particles are collected and occasion-

ally the entire distribution is recorded. At these times, all deferred growth due to

all particles is simulated so that the size distribution is fully up to date. This has

the convenient effect of ensuring no stochastic particle is left for too long before

its deferred surface growth is performed. One can therefore think of LPDA as a

form of splitting but with a ‘just in time’ feature to remove the main source of ap-

proximation error. However, with LPDA the number of times the entire ensemble

has to be updated is orders of magnitude less than for ordinary operator splitting

and so for practical purposes it is a completely different algorithm.

Complexity

The (random) number of deferred events due to any particular particle is O (1) in

the sample volume V . There are O (V ) non-deferred events in a run and each of

these requires use of the binary tree which takes O (log V ) per event (see §2.3.3).

72

Page 74: Numerical Modelling of Soot Formation

Hence the overall complexity is still O (V log V ) in the (initial) sample volume

for reasonable choices of tree depth. However in §3.1.2 it was noted that most

computational effort seemed to go on surface reactions and so if all these were

deferred the complexity might not be observed for realistic values of V .

3.3.5 Numerical Results

Accuracy of Moments

The same JW1.69 test case was used, with a tree depth of 13, as in §3.2.1. For the

calculations reported in this chapter and chapter 4 all surface reactions apart from

pyrene condensation were deferred, that is,

U ′ = pyrene condensation . (3.15)

Pyrene condensation was not deferred because of the relative complexity of β(pyr)

and the potentially large change in particle size caused by a single pyrene conden-

sation event5. Ten runs took 250 seconds, an appreciable acceleration compared

to the operator splitting cases. Figure 3.4 compares the zeroth and second mo-

ments of the mass distribution obtained with LPDA to the DSA precision run and

to the ‘pv’ r = 10 run. One can see that LPDA slightly overstates the number

of particles compared to the other two methods at least until 0.08 s. Interestingly,

with the second moment, LPDA tracks the operator splitting solution until about

0.1 s and then moves much closer to the DSA solution. The first moment is not

shown, but here LPDA is virtually indistinguishable from DSA on the scale of the

figures and noticeably better than operator splitting.

Particle Distribution Accuracy

As with operator splitting this topic is investigated for JW10.68; a summary of the

simulations is given in table 3.4. The deferred surface growth computations were

performed on the same machines used to generate results for §3.2.1. As before ‘†’5The work in chapter 5, carried out at a time when LPDA was better understood, ultimately

showed that it is acceptable to take U ′ = , the empty set, by deferring pyrene events.

73

Page 75: Numerical Modelling of Soot Formation

0

0.5

1

1.5

2

2.5

0 0.02 0.04 0.06 0.08 0.1 0.12

dsa 17pv 10lpda

parti

cle

num

ber d

ensi

ty /

1011

cm

-3

time / s

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 0.02 0.04 0.06 0.08 0.1 0.12

dsa17pv 10lpda

seco

nd m

ass

mom

ent /

10-2

4 g c

m-3

time / s

Figure 3.4: LPDA applied to JW1.69

74

Page 76: Numerical Modelling of Soot Formation

denotes the Athlon machine. One observes very considerable speed gains from

using LPDA.

Table 3.4: Detail of deferred surface growth runs for JW10.68

number of tree levels time for 10 repetitions / minDSA 13 8497

LPDA 15 15 22†LPDA 14 14 11†LPDA 13 13 5.25†LPDA 12 12 2.5†

The LPDA run time with 13 tree levels is more than 3 orders of magnitude

less than that for DSA which is a very positive result. Figure 3.5 shows that the

particle distribution is well reproduced using LPDA.

3.3.6 Conclusions on LPDA

This method is extremely attractive for further use in the simulation of laminar

premixed flames because of the large speed gains possible without introducing

large errors. The comments on the errors made in §3.2.4 apply here as well.

The initial sample volume was scaled in direct proportion to the tree capacity

for all the runs, for 13 tree levels an initial sample volume of 1 × 10−8 cm3 was

used. It is therefore interesting to see the run times showingO (V ) behaviour, not

O (V log V ) as discussed in §3.3.4.

3.4 Comparison of Simulation Methods

3.4.1 Equal Numbers of Computational Particles

A clear ordering in terms of speed can be seen in table 3.5. The differences in the

particle distributions produced using the different methods are not material.

75

Page 77: Numerical Modelling of Soot Formation

0 20 40 60 80 100 120 140 16010

5

106

107

108

JW10.68 PSDF

sphere equivalent diameter / nm

num

ber

dens

ity /

cm-3

dsa 14lpda 13lpda 15

Figure 3.5: Particle distribution for JW10.68

76

Page 78: Numerical Modelling of Soot Formation

Table 3.5: Comparison of simulation methods

algorithm flame number of tree levels time for 10 repetitions / minDSA JW10.68 13 8497

pv r = 5 JW10.68 13 844LPDA JW10.68 13 6.33DSA JW1.69 13 50

pv r = 5 JW1.69 13 8LPDA JW1.69 13 4.2

The ratios of run times between methods are very different for the two flames

considered, as are the total run times. Unsurprisingly, the greatest accelerations

were achieved for the problem that takes the most CPU time.

3.4.2 Equal Precision for JW1.69

Finally, the time required to obtain results of comparable quality using the three

main algorithms discussed in this chapter is considered. Result quality is mea-

sured using the second moment of the particle size distribution on exit from the

flame. The reference value is the mean obtained from 30 runs using basic DSA, 17

tree levels and an initial sample volume of 5.6× 10−7 cm3. Table 3.6 gives details

of parameters which lead to 99.9% confidence intervals for the second moment

being entirely contained in an interval of ±10% around the reference value.

Table 3.6: Time to achieve tolerances for JW1.69

algorithm number of tree levels repetitions total time / sDSA 10 30 990DSA 11 15 645

LPDA 11 15 98LPDA 12 10 133

pv r = 5 12 20 500pv r = 5 13 10 480

77

Page 79: Numerical Modelling of Soot Formation

The parameters shown were selected as the least multiples of 5 runs that led

to adequate results for the second moment. More runs or more tree levels were

required for LPDA and operator splitting because these methods slightly bias the

mean value of the second moment and so tighter confidence intervals are needed

to stay within the ±10% range.

3.4.3 Summary

The ‘pv’ method and LPDA both seem to provide a good level of accuracy and

offer significant savings over the DSA in computation times. For a given level of

accuracy LPDA required the least amount of computer time and so is the method

of choice. LPDA was implemented in a way which preserved the stochastic nature

of reactions of the surface of soot particles. This raises the possibility of achieving

further computational savings by using a simpler, deterministic approximation for

the surface reactions.

78

Page 80: Numerical Modelling of Soot Formation

Chapter 4

Deterministic Simulation of theSurface Processes

The previous chapter focussed on the acceleration of the basic DSA while pre-

serving the random nature of the chemical reactions with the surfaces of soot

particles. However, the high rates of these reactions suggest, that on all time

scales of interest the variability in the number of reactions might be small. In this

case, deterministic approximations should offer simpler (hence faster) simulation

techniques without significant loss of accuracy.

The exploration was carried out within the frameworks of the basic operator

splitting algorithm set out in chapter 3 and of the LPDA presented in the same

chapter. The coalescent sphere particle model (§1.4.1) was used throughout so

that all particles were always spherical. Deterministic approximations to random

variables have been successfully used as an acceleration technique for cell simu-

lation in biology [94]. The mathematics of the deterministic approximation and

some intermediate stages are set out in a little more detail in [66], where it is

applied with success to the simulation of chemical reactions.

79

Page 81: Numerical Modelling of Soot Formation

4.1 Operator Splitting

Operator splitting provides the simplest situation in which to test numerical meth-

ods for including the effects of surface reactions in simulations. The basic algo-

rithm is detailed in §3.1.1 & §3.1.2, however, the benefits of using a more complex

splitting system, such as a Strang splitting, remain to be investigated. To cope with

the wider variety of algorithms studied in this chapter, the names used in chapter 3

have been expanded to form a more descriptive scheme. The ‘pv’ algorithm from

chapter 3 is denoted pv2s in this chapter; ‘pv’ for Poisson variable as previously,

‘2’ because there are 2 levels of splitting and ‘s’ because the top level splitting

time step is specified in terms of the rates of the split processes.

4.1.1 Modified Choice of Top Level Time Step

For the principal test case, the flame JW10.68 considered previously, it was found

that considerable performance benefits accrued if the length of the splitting time

steps was chosen in terms of the rates of the unsplit processes, rather than the

split processes as in chapter 3. The length of the time over which one part of the

split operator was applied before the other part of the operator was applied was

chosen so that the expected number of non-split events per computational parti-

cle during this interval was r. This constitutes a small amendment to point 2a)

of the algorithm in §3.1.2. Using this new choice of time step, r could be set to

0.1 which reduced run times by a factor of 30 compared with the old method in

the case r = 20. This acceleration was achieved while still providing the same

accuracy obtained with the old method for r = 5. Using a value of r = 0.02 with

the new method offered an acceleration by a factor of 4 over the old method with

r = 5, at the same time as a gain in accuracy. Accuracy was quantified as the

deviation of the lower order moments from the high precision DSA results used

as the reference solution in this chapter and in chapter 3. Times are not directly

comparable to those reported in chapter 3 because the work reported in this chap-

ter was done with a different compiler running on slightly different hardware. The

modified algorithm is denoted pv2j; for the meaning of ‘pv2’ see the start of §4.1

80

Page 82: Numerical Modelling of Soot Formation

above and ‘j’ is used to indicate that the top level splitting time step is specified

in terms of the rate of the non-split processes, which are simulated as a compos-

ite jump process. The same splitting strategy is applied, using a different Monte

Carlo technique, to a slightly different physical system in [47].

4.1.2 Elimination of second level of splitting

For the flame used as the main test case in the first part of this chapter, JW10.68, it

was found that the second level of splitting in the ‘pv’ algorithm of chapter 3 was

redundant, since the top level splitting steps were already of the order of the time

over which the chemical conditions varied. Reducing to the 1 level algorithms

pv1s and pv1j had a negligible effect on the results produced by the simulations

and, in most cases, made little difference to the run times. In other systems the

chemistry could vary over time scales much shorter than the 1 level splitting step

length, so the more general applicability of these methods will be discussed fur-

ther. Most of the work in this chapter will use JW10.68 as a test case, since

it enables splitting strategies to be considered without additional complications

arising from a second level of splitting.

4.1.3 Initial Trial of Deterministic Method

In each place in the pv1j algorithm, where a Poisson random variable was gener-

ated to give the number of occurrences of a particular kind of surface event, the

(pseudo-)random value was replaced with its mean. This algorithm is denoted

e1j; ‘e’ for Euler, ‘1’ because there is still 1 level of splitting and ‘j’ for the rea-

son given above in connection with the pv algorithms. To provide a check on the

results obtained by implementing the e1j algorithm, the existence of an analytic

solution to the spherical particle growth problem was exploited.

Under the assumption of deterministic, continuous surface growth propor-

tional to particle surface area and equating the type of a particle, x, with its mass,

81

Page 83: Numerical Modelling of Soot Formation

one has, for some function k of the chemical environment

d

dtx = k (t)x

23 . (4.1)

This may be solved for individual particles, giving

x (t1)13 − x (t0)

13 =

1

3

∫ t1

t0

k (t) dt. (4.2)

Since k (t) is known in advance (§1.4), the numerical calculation of the right hand

side of (4.2) and hence the change in particle mass is simple.

Therefore (4.2) was used to calculate the total size change due to all split

processes as an alternative to the Euler method just described. This alternative

method is denoted by as1j with ‘as’ standing for analytic splitting, and ‘1j’ having

the same significance as above. The deviations from the reference solution for the

first three moments of the soot particle distribution are given for the various algo-

rithms and two values of r in table 4.1. It is interesting to note that the moments

are never underestimated by the 1 level splitting algorithms; the explanation of

this would seem to be that splitting prevents small particles from being oxidised

out of the distribution, because it treats surface oxidation and growth together. The

e1j and as1j algorithms show close agreement, offering some assurance that they

were implemented correctly. One sees that these deterministic splitting methods

overstate m11 and, to a greater extent, m2, even in relation to the pv1j method,

indicating that e1j and as1j over-predict soot particle sizes. Unlike for pv1j, the

choice of r seems to have little effect on the errors for e1j and as1j.

Times in seconds for 10 runs on single AMD Athlon XP3000+ Linux PCs are

given in table 4.2 for simulations using each algorithm. The same settings—initial

sample volume of 2.5× 109cm3 and binary tree depth of 11 were used in all cases

so that the number of computational particles varied between 1024 and 2047 in

all the simulations after the initial inception peak.

The slight speed advantage of as1j over e1j will be connected to the fact that in

the as1j algorithm the effects of all split processes are calculated together, whereas1m1 is used here for M1 as defined in (1.2), m0 and m2 should be understood in the same way.

82

Page 84: Numerical Modelling of Soot Formation

Table 4.1: Numerical errors in JW10.68 distribution moments

time / s r (see text) algorithm m0 m1 m20.08 0.02 pv1j 0% 0% 1%

e1j 0% 4% 9%as1j 1% 5% 8%

0.1 pv1j 1% 2% 3%e1j 1% 4% 8%as1j 1% 5% 7%

0.16 (exit) 0.02 pv1j 1% 1% 1%e1j 1% 4% 7%as1j 1% 5% 9%

0.1 pv1j 1% 2% 5%e1j 2% 5% 8%as1j 2% 5% 7%

Table 4.2: Run times in seconds for different algorithms on the same hardware

r pv1j e1j as1j ad1j0.02 560 397 259 3100.1 138 92 76 81

83

Page 85: Numerical Modelling of Soot Formation

for e1j this step has to be repeated for each split process. On the basis of this data,

either pv1j or as1j is to be preferred for further use, but not e1j since it is slower

than as1j, while producing similar results. The choice between pv1j and as1j will

depend on the relative emphasis placed on accuracy and computational speed.

4.1.4 Adaptive splitting

In this section an algorithm denoted ad1j is introduced where ‘ad’ stands for adap-

tive in a sense that will be explained, and ‘1j’ has the same significance as above.

As mentioned in §4.1.3, splitting seems to introduce an error by preventing small

particles being oxidised out of the soot population, because oxidation quickly fol-

lowed by surface growth can lead to a zero or positive net size change. To test this

explanation of the observed error and to try to reduce the overall error caused by

assuming deterministic surface growth in the as1j and es1j algorithms, even when

a very small number of events was probable, the as1j algorithm was modified to

use a stochastic method for particles when the expected number of events within a

second level splitting step was less than 5. The stochastic method called the ‘pp’

algorithm in chapter 3 (which corresponds to direct simulation of the split pro-

cesses for single particles) was used in this case, as it addresses the small particle

oxidation issue. Table 4.3 contains the same data for the ad1j algorithm as that

given in table 4.1 for the earlier algorithms.

Table 4.3: Numerical errors in JW10.68 distribution moments

time / s r algorithm m0 m1 m20.08 0.02 ad1j 0% 0% 0%

0.1 ad1j -1% 0% 2%0.16 (exit) 0.02 ad1j 0% 0% 1%

0.1 ad1j -1% 1% 3%

These results show the ad1j algorithm can reproduce the reference solution

very successfully, and is at least as accurate as the pv1j algorithm. An implemen-

tation of the ad1j algorithm is naturally slower than a comparable implementation

84

Page 86: Numerical Modelling of Soot Formation

of as1j, but it is still faster than e1j (see table 4.2). The negative deviation of m0

from the reference value with r = 0.1 is consistent with the claims made above

about the oxidation of very small particles, but the difference is too small to allow

any firm conclusions.

4.1.5 Testing with a second flame

The above algorithms were checked for a second flame—JW1.69. The pv2s al-

gorithm with r = 5, 20 produced similar results to pv1s for the same r. With

r = 0.02, 0.1 pv2j produced similar results to pv1j for the corresponding r, which

is not surprising given that the top level time steps were already on the time scale

of the variation in the chemistry. The time differences between the 1 and 2 level

algorithms were negligible. However, unlike for JW10.68, the ‘s’ versions of the

pv1 and pv2 algorithms were faster than the ‘j’ versions. (The most likely cause

of the difference in performance between the two flames is that the relative fre-

quencies of the split and unsplit events are quite different, for details see table 2.3

and table 3.2.) Implementations of pv1j and pv2j with r = 0.1 took 16 minutes

for 20 repetitions of JW1.69 with the number of computational particles varying

between 4096 and 8191, whereas pv1s and pv2s with r = 5 only took about 10

minutes. The results from pv1j and pv2j in this test were at least as accurate as

those from pv1s and pv2s.

The e1s, e2s and e1j algorithms with the r values just mentioned all produced

similar output to each other when applied to JW1.69. However, the moments cal-

culated with these three algorithms were very different from those of the reference

solution—the error reached about 100% for m2. The analytic splitting method

as1j with r = 0.1 produced moment values close to those of as1s with r = 5.

Both ‘as’ methods were less inaccurate than the ‘e’ algorithms. It is not clear

why as1j should lead to different moments from those obtained with e1j and e2j.

Algorithm ad1j, the variant of as1j, produced the same high accuracy for JW1.69

that it had produced for JW10.68, but was still slower than pv2j with r = 5 (13.5

minutes compared to 10.5).

The best algorithm for the two flames investigated would seem to be ad1j with

85

Page 87: Numerical Modelling of Soot Formation

r = 0.1. It works well for the flame JW10.68, which led to particularly slow DSA

calculations, and is also satisfactory when applied to JW1.69.

4.2 Deterministic Simulation of Deferred Events

Results from the profiling of an LPDA simulation of JW1.69 are reported in fig-

ure 4.1. They show that 30% of the computation time was spent on the top level

stochastic stepping for the processes simulated as in the DSA—particle incep-

tion, pyrene condensation and particle coagulation. Simulating individual coag-

ulation events took a further 13% of the time, some of which was attributable

to the need to perform deferred events on coagulation candidates. About 50%

of the computation time was required for the simulation of pyrene condensation

events. Approximately one third of this 50% was spent simulating deferred events

on particles selected to undergo pyrene condensation, and another third on updat-

ing the binary tree structure with the results of the event. Given the similarities

in the implementations of coagulation and pyrene condensation, it seems reason-

able to expect a similar breakdown of the 13% of computation time spent on the

simulation of coagulation events. Most of the computation time not accounted for

above was spent in the simulation of deferred events for all particles at times when

data was recorded. The simulation of particle inception did not require significant

amounts of time.

The profiling reported in figure 4.1 shows that pyrene condensation is the most

time consuming part of the simulation after the other surface reactions have been

removed from the DSA style simulation process according to (3.15). This is con-

sistent with the event counts presented in table 3.2. If the simulation of deferred

events could be accelerated so that the time required was a small fraction of that

in the profiled program one could look for a reduction of up to 20% in the total

simulation time. Given that a considerable amount of work had already been done

to accelerate the binary tree operations and the DSA simulation code, there were

only two sensible routes to try to achieve further acceleration. Firstly, one could

try to incorporate the pyrene condensation in the framework developed for the

86

Page 88: Numerical Modelling of Soot Formation

other time step overheads

15%

pyrene (deferred surface growth)

20%

pyrene (updating ensemble)

14%

coagulation13%

other (pyrene event related)

17%

unclassified1%

updating entire ensemble

6%

calculating stochastic step

length14%

Figure 4.1: % of execution time spent on different tasks—LPDA with ‘pv’

87

Page 89: Numerical Modelling of Soot Formation

other, more simple surface reactions. (This was successfully done in the bivariate

simulations of chapter 5.) The second approach, described here, is to look for

ways to accelerate the simulation of deferred processes for particles involved in

non-deferred events.

4.2.1 Results with LPDA

On the basis of the work on operator splitting above, the effects of using the ad1

algorithm to perform the deferred processes within LPDA was investigated. A

summary of the output for JW10.68 is given in table 4.4 and suggests that incor-

porating an ‘ad’ algorithm into LPDA slightly increased the accuracy from that

achieved in the initial version (using pv2).

Table 4.4: Numerical errors in JW10.68 distribution moments

time / s algorithm m0 m1 m20.08 ad2 0% 0% 0%

ad1 1% 3% 5%pv2 2% 4% 7%pv1 0% -2% -4%

0.16 (exit) ad2 1% 3% 7%ad1 2% 4% 7%pv2 -1% -2% -3%pv1 0% -8% -14%

It can also be seen that pv1 and pv2 lead to significantly different results (un-

like for the splitting situation). Tests on JW1.69 show the same relative perfor-

mance from the sub-algorithms used for the deferred processes. Run times (in

seconds) for LPDA simulations with the 4 different sub-algorithms for the de-

ferred process are given in table 4.5 for the same settings used when measuring

the run times quoted in previous sections of this chapter. The ad1 treatment of the

deferred processes reduces the computation time by about 20% for both flames

considered, which is the maximum suggested by the profiling discussed above.

88

Page 90: Numerical Modelling of Soot Formation

Table 4.5: Run times in seconds for different sub-algorithms on the same hard-ware

sub-algorithm pv2 pv1 ad2 ad1JW10.68 52 43 47 41JW1.69 354 308 337 298

The times using the pv1 sub-algorithm show much of the reduction is due to the

removal of the time stepping within the simulation of deferred processes.

4.3 Recommended Algorithm

When using the ad1j r = 0.1 splitting algorithm (the algorithm recommended as

the best of the splitting methods, see §4.1.5) run times of 13.3 minutes and 81

seconds were achieved for simulations of the flames JW1.69 and JW10.68 respec-

tively. Using LPDA with a pv2 treatment of the deferred surface processes took

5.9 minutes and 52 seconds respectively for the same two test cases. Therefore,

for the two flames tested, LPDA seems to be the best algorithm to use and is to be

recommended as the default option for other flames.

89

Page 91: Numerical Modelling of Soot Formation

Chapter 5

Models for Particle Shape

The focus in this chapter shifts, from the numerical development that has occupied

the thesis so far, to an application of the LPDA, which seems to be the most

suitable algorithm for simulating soot formation in laminar premixed flames. The

work reported in this chapter can be viewed both as a demonstration of the power

of the Monte Carlo simulation approach established in the preceding chapters and

as a contribution to physical model development.

5.1 Background

The fractal nature of large soot particles has been known for some time [102,

186, 232]. The extensive review paper [185] gives an indication of how much

work has been done on experimental techniques, to measure and to account for

the properties of these fractal structures. Detailed consideration of how to account

for the fractal nature of soot particles from road vehicles has even been included

in atmospheric modelling [88].

Although initial results [11, 245, 246] for inorganic nanoparticles show the

potential importance of good models for the aggregate, simulation of the fractal

structure of soot particles is in its infancy. Mitchell and Frenklach investigated

soot aggregation [133–135] by representing a single aggregate particle as a union

of intersecting spheres. The natural extension of that approach to a population

90

Page 92: Numerical Modelling of Soot Formation

of particles is discussed by Balthasar et al. [17] and Morgan et al. [140], but

this is very computationally expensive. A simplification of the model includes

an assumption that particle size and structure are independent, and incorporates

a numerical fit for the functional form of the particle collision diameter. It en-

ables very fast calculations to be made for the lower moments of the particle mass

distribution and mean shape [14]. These ‘method of moments’ calculations of-

fer considerable insight into the development of particle structure, although it is

difficult to assess the validity of various modelling approximations.

Park and Rogak [155] added a partial representation of aerosol particle struc-

ture to a one dimensional sectional technique to simulate particle formation in a

plug flow reactor [156]. Their work differed from the approach of Frenklach and

his co-workers by representing particle structure with a physical quantity—the

average number of primary particles per aggregate, with a separate average be-

ing taken within each size section. To calculate particle collision diameters from

this information, an assumed fractal dimension of 1.8 was imposed [215, 216],

which implies a very open particle structure. This value of 1.8 has been used by

other authors, for example, in [196] and reported from diffusion flame soot [101].

Slightly lower values were reported from light scattering experiments on premixed

methane-oxygen flames in [187].

However, support for values around 1.8 is not universal as 2.2, 2.1 and 2.2

are reported by Maricq et al. [121] at increasing heights in a premixed ethylene

flame. They also report some remarkable comparisons of experimental and mod-

elling data [120], and show fractal dimension decreasing with increasing height

above burner in a laminar premixed flame. A fractal dimension of approximately

1.8 has been reported for randomly diffusing aggregates in a wide range of simu-

lation work, for example, the Brownian trajectory cluster-cluster case of Meakin

[127]. However, it is shown in [19] that small units of surface growth lead to

higher fractal dimensions than pure aggregate—aggregate coagulation. Meakin

also showed that aggregate restructuring can lead to more compact particles, that

is, cause an increase in fractal dimension [128] and it appears that this would even-

tually lead to a fractal dimension of 2.5 [39, 100]. Changes in fractal dimensions

91

Page 93: Numerical Modelling of Soot Formation

are therefore to be expected as soot particles move from chemically rich environ-

ments where they become dense to regions dominated by coagulation. Very rich

flames will emit dense particles with high fractal dimensions [149]. In flames

with a significant coagulation-only post reaction zone, particles will develop a

very open structure with a fractal dimension around 1.8 before leaving the flame.

The influence of fractal dimension on coagulation rates is reported to be small

in the continuum regime (particles large compared to the mean free path) [99] but

to be significant in the free molecular regime [99, 245]. This means, that even in

high pressure flames, fractal structure should be considered in order to calculate

accurate collision rates, because the free molecular regime will apply to small soot

particles. For example, a 5 nm particle in gas at a pressure of 10 bar is in the free

molecular regime. Therefore despite the successful application of imposed fractal

dimension models something more general is also required.

This chapter introduces and reports tests of models for the aggregate structure

of soot particles with fully bivariate (that is, where each particle is described by

two internal co-ordinates, which are not functionally dependent on each other)

simulations. Development of these models is a critical first step in building simu-

lations with realistic models for soot surface chemistry. Without such simulations

it will be impossible to progress from descriptive fits [8, 95, 231] to physical mod-

els for surface growth rates. Simulation of particle structure will also enable more

direct comparisons with experimental data based on particle mobility [109, 240]

and scattering techniques such as those discussed above and, for example, in [208]

and [24]. Readers may also be interested in the work of Muhlenweg et al. [142],

which addresses the question of how to go beyond the spherical particle model for

inorganic particles using sectional methods.

92

Page 94: Numerical Modelling of Soot Formation

5.2 Framework

5.2.1 Flame chemistry model

As stated in §1.4, the basic soot model, an extension to which is described below,

is that of Appel et al. [8]. It includes an assumption of spherical particles and is

based on the gas-phase chemistry from Wang and Frenklach [207]. Soot particles

are assumed to form exclusively through the coalescence of two pyrene molecules

to make a (spherical) particle of 32 C atoms. Surface growth is described by

surface deposition of further pyrene molecules and by the Hydrogen-Abstraction-

Carbon-Addition (HACA) mechanism. Surface oxidation by oxygen (O2) and

hydroxyl radicals (OH) is included. For all the flames examined it was found that

acetylene addition was the main route of mass growth.

5.2.2 Numerical method

The simulations reported in this chapter were performed using the LPDA accel-

eration of DSA with the ‘pv2’ treatment of deferred processes recommended in

§4.3. Further tests, alluded to in §4.2, showed that pyrene growth could be de-

ferred despite leading to much larger jumps in particle size than the other surface

processes. Accordingly, in order to achieve additional savings and to simplify the

software, pyrene condensation events were deferred in all the simulations reported

in this chapter, that is, (3.15) was replaced by

U ′ = . (5.1)

The stochastic particle simulation method developed in the preceding chapters

is particularly well suited to the model exploration work of this chapter, because

the complexity of the individual particle dynamics can be increased without in-

creasing the computational complexity of the calculations. Since the program

complexity was unchanged the amount of computer time required for the bivari-

ate simulations reported in this work was only slightly greater than for univariate

simulations reported earlier in this thesis, which only tracked particle masses; the

93

Page 95: Numerical Modelling of Soot Formation

small increase was due to the doubling of the amount of data to process. The

programs used in this chapter were written by making small extensions to code

previously used for univariate particle simulations.

5.2.3 Particle shape variable

Soot modelling has generally concentrated on particle mass or volume. The equiv-

alence of mass and volume follows from the assumption that all soot material has

the same density. The work reported here uses the value supplied with the soot

mechanism [8] as implemented by its authors [166]: 1.8 g cm−3. The apparent or

effective density [121] of aggregate structures may take a lower value, which is

reflected in the shape modelling in the present work.

In this chapter mass and surface area are used as the independent variables

in the particle descriptions or types, hence the term “bivariate model” used exten-

sively hereafter. Surface area is used rather than the number of primary particles as

in [155] because it gives a description immediately equivalent to that of the shape

descriptor by Mitchell and Frenklach [133–135]. Surface area has also been used

in bivariate simulations of inorganic nanoparticles [172, 223].

5.2.4 Notation

For convenience the following dimensionless functions are defined; the quantities

are divided by their respective values for newly incepted particles (type x0) so they

all take the value 1 at x0. (x0 describes a spherical particle of 32 C atoms—the

assumed form of all newly incepted particles, see §1.4.)

• c(x) the collision diameter of the particle

• m(x) the mass of the particle

• s(x) the surface area of the particle

• v(x) the volume of the particle (= m(x))

94

Page 96: Numerical Modelling of Soot Formation

A shape descriptor d is defined as [14, 135]

d (x) =log (s(x))

log (v(x))(5.2)

and hence setting d (x0) = 23

(the value would otherwise be undefined at x0) yields23≤ d ≤ 1. Note that for a sphere, given x0, any two of d, v and s define the value

of the third, and that d depends on x0 via the normalisation of v and s. A detailed

discussion of the shape descriptor is given by Mitchell and Frenklach [133]; a

sphere has a descriptor of 23, a chain aggregate of particles (any length > 1) of

type x0 has shape descriptor of 1, and, for example, a chain of 100 spheres of

diameter 13 nm has a shape descriptor of 0.79.

5.3 Models for Lengths

5.3.1 Collision Diameter

As mentioned in the introduction, the coagulation rate is much more sensitive

to collision diameter in the free molecular regime than in the continuum regime.

This is not especially surprising given the form of the coagulation kernel between

two particles [99] in the continuum regime:

Kcn(x, y) ∝ (c(x) + c(y))

(1

c(x)+

1

c(y)

). (5.3)

This only depends on the ratio of collision diameters of the two particles not their

absolute value and so some kind of cancellation is expected if both are increased.

However, the free molecular kernel [99]:

Kfm(x, y) ∝(

1

m(x)+

1

m(y)

) 12

(c(x) + c(y))2 (5.4)

scales with the square of the absolute value of the collision diameters.

In the spherical particle model the collision diameters of the particles are sim-

ply taken to be the ordinary diameters of the spheres, and so with the scaling of

95

Page 97: Numerical Modelling of Soot Formation

§5.2.4 c(x) = v(x)13 . In this section ways to proceed for non-spherical particles

are set out.

Mitchell & Frenklach Collision Diameter

In Mitchell and Frenklach [134, 135], the collision cross-section was expressed

in terms of the radius of gyration, which was calculated using a time consuming

Monte Carlo integration. Mitchell [135] also proposed the following model for

the collision diameter, cagg, of an aggregate x comprising spherical particles in

point contact, as a practical approximation for use in further work:

cagg (x) = kv(x)13 + 2 (1− k)×

(av(x)

13

)b

+ v(x)b(a2

13

)b

+ 2b

1b

(5.5)

a = 1.53311

b = −1.35419

k = 0.43074.

The above expression was extended to general particles by interpolation using the

shape descriptor d to give

c(x) = 3

[(d (x)− 2

3

)cagg (x) + (1− d (x)) v(x)

13

]. (5.6)

Balthasar & Frenklach Collision Diameter

Balthasar and Frenklach [14] went further in simplifying the work mentioned

above. They introduced a correlation derived from the more detailed work which

expressed collision cross-section as a function of aggregate volume and shape de-

scriptor:

c(x) = (2.7375d (x)− 0.825) v(x)13 . (5.7)

By making several further approximations, in particular, that at any time the shape

descriptor was the same for all particles, they were able to use the method of mo-

96

Page 98: Numerical Modelling of Soot Formation

ments with interpolative closure (MoMIC)[57] to achieve notable computational

savings.

Arithmetic Mean Collision Diameter

To test the sensitivity of results to the exact collision cross-section model, the

arithmetic mean of the volume and surface equivalent diameters is introduced:

c(x) =1

2

(v(x)

13 + s(x)

12

). (5.8)

It represents the simplest expression that uses the information contained in the

surface area and the volume.

Weighted Geometric Average Collision Diameters

A second alternative introduced was the geometric mean of the equivalent volume

and surface diameters

c(x) = v(x)16 s(x)

14 . (5.9)

This is a member of the family of weighted geometric averages

c(x) = v(x)a s(x)b , (5.10)

which satisfy

3a+ 2b = 1. (5.11)

These turn out to have an interesting connection to assumed fractal dimensions:

consider an idealised aggregate soot particle of volume V and surface area S con-

sisting of np spherical primary particles of diameter dp in point contact with each

other (the model used in [154]). Solving for np and dp gives

dp =6V

S(5.12)

np =S3

36πV 2(5.13)

97

Page 99: Numerical Modelling of Soot Formation

The standard fractal relationship, in this case for collision diameter C and

aggregate volume is

np = k

(C

dp

)D

(5.14)

where k is the fractal prefactor and D the fractal dimension. One then finds [196,

equation 4]

C = 6× (36πk)−1D V 1− 2

D S3D−1. (5.15)

For a general particle x this becomes

c(x) = k−1D v(x)1− 2

D s(x)3D−1 , (5.16)

which, for k = 1, is in the form of (5.10) and satisfies (5.11).

Note, that under the k = 1 assumption (also made by Zucca et al. [244])

using the equally weighted geometric mean collision diameter (5.9) is the same

as assuming a fractal dimension of 2.4 and that a fractal dimension of 1.8 implies

a = −19

and b = 23. The negative value of a is a somewhat counter-intuitive but

cannot easily be dismissed because of the extensive reports of soot particles with

a fractal dimension of 1.8, some of which were referred to in §5.1.

Fractal Prefactor

Setting the fractal prefactor equal to 1 makes (5.14) consistent for a single sphere,

but experimental results [195], simulated aggregates [148] and other theoretical

considerations [106] lead to larger values of k. Oh and Sorensen refer to a number

of different values for k in their introduction to [148], their analysis then shows

that values of k will be higher for aggregates with overlap between the primary

particles.

The fractal relationship (5.14) is often expressed using twice the radius of

gyration, 2Rg, rather than the collision diameter C. For a sphere

C =

(5

3

) 12

2Rg (5.17)

98

Page 100: Numerical Modelling of Soot Formation

suggesting that reported values of k (which vary from near 1 [148] to at least 7

[118]) should be increased by 30% for use here. A simple test for the influence

of k on simulation results will be reported below using k = 1 and k1D = 2 with

a linear interpolation to preserve the correct value for spherical particles. The

interpolation is between k1D = 2 for np ≥ 10 and k

1D = 1 for np = 1 giving

k1D =

np + 8

91 ≤ np ≤ 10 and k

1D = 2 np > 10. (5.18)

More complex approaches have also been used [172], where the fractal dimension

was interpolated as well as the prefactor to reach the correct limit for a single

spherical particle.

5.3.2 Radius of Curvature

Physical Considerations

The soot model [8] implies that particles form as hard spheres when two pyrene

molecules coalesce to form a ball of diameter 0.88 nm. Electron micrographs

[18, 38, 149, 194] show soot particles are composed of far fewer primary particles

than would be the case if all newly incepted particles remained distinct and have

a much lower surface to volume ratio than 0.88 nm diameter spheres. Therefore

any model for surface reactions must lead to a decrease in the surface to volume

ratio of particles and some kind of merging of primary particles [15]. A model of

uniform surface growth on the free surface has been proposed by Balthasar and

Frenklach [15], leading to aggregates of intersecting spheres. This model may be

varied by concentrating reactions around the points of contact between primary

particles [141], but both variations treat the interior of particles as homogeneous.

Analysis of high resolution TEMs suggests that the assumption of homogeneity

will have to be removed when more precise data and realistic models become

available.

99

Page 101: Numerical Modelling of Soot Formation

Bivariate Modelling

In the context of the bivariate model for soot particles such surface reactions (and

any that remove mass) must be defined by the changes they cause in particle mass

and surface area. The particle mass change is obvious—it is the mass of the

deposited molecule—and the volume change is simply this value divided by the

density. For a sphere (initial radius R, surface area S viewed as functions of

volume V ) the change in surface area is

d

dVS =

2

R. (5.19)

The relationship between the volume and surface is controlled by the surface cur-

vature. For a general particle shape the value of R which makes (5.19) true at

a point on the particle surface is known as the radius of curvature (for a general

surface in a three-dimensional space there are two radii of curvature at each point,

for simplicity they are assumed to be equal in this work).

The correct expression for the radius of curvature of non-spherical particles

in the bivariate model considered here is not at all obvious. As discussed above,

the model for curvature should cause aggregate particles to become rounder. The

assumption of rounding was made more precise by dividing it into the following

assumptions:

1. spherical particles must remain spherical in the limit of small volume incre-

ments;

2. all particles must be geometrically possible, in particular the surface to vol-

ume ratio must be achievable;

3. non-spherical particles must become rounder during surface growth and ox-

idation.

Note that (3) can be viewed in a statistical sense; it is not necessary that every

instance of every surface reaction causes a positive increase in roundness.

One consequence of these conditions is that at least two radii of curvature

(these are not the principal radii of curvature from differential geometry) are re-

100

Page 102: Numerical Modelling of Soot Formation

quired for non-spherical particles, one to use when ∆V > 0 and the other when

∆V < 0. The following curvature radii were used in this work, subscript init

denotes the state of the particle immediately before the mass change:

Rgr =

(Sinit

) 12

(5.20)

for growth processes and

Rox =3Vinit

Sinit

(5.21)

for oxidation processes.

In the scaled, dimensionless style of §5.2.4, by (5.19) one has

rgr (x) = s(x)12 (5.22)

and

rox (x) =v(x)

s(x). (5.23)

So for ∆v > 0

∆s =2

3∆v s(xinit)

− 12 (5.24)

and for ∆v < 0

∆s =2

3∆v

s(xinit)

v(xinit). (5.25)

Other definitions are possible and numerical tests of model 4, the simpler case

of all length scales equal, are reported below.

The general ordering of the length scales introduced is

rox(x) ≤ v(x)13 ≤ c(x) ≤ rgr(x) . (5.26)

Values for a couple of typical particle sizes, a smaller particle from low in the

flame and a larger particle from high in the flame, (see figures 5.5 & 5.7) are

given in table 5.1. As expected the smaller particle, of a type found in a region

with high surface growth rates, has a much smaller range of length scales than the

particle that has experienced prolonged coagulation with little surface growth in

101

Page 103: Numerical Modelling of Soot Formation

the post-reaction zone of the flame.

Table 5.1: Length scales for typical JW10.673 particle sizes

volume surf 2Rox vol equiv collision diam 2Rgr

/ area / diam model 2 model 3 /cm3 /cm2 nm / nm / nm / nm nm

1× 10−17 3× 10−11 20.0 26.8 28.8 28.8 30.91× 10−15 1.3× 10−9 46 124 147 159 203

5.4 Model Comparison

5.4.1 Bulk Properties

To investigate the significance of the modelling assumptions, bulk properties of

the particle populations predicted using the different models were compared for

the 10 bar laminar premixed ethylene flame JW10.673 [93], which was also the

focus of [14]. Results are shown in figures 5.1 & 5.2 as functions of Height Above

the Burner (HAB) and a summary of the different models, along with the numbers

used to refer to them in the legends of the figures, is given in table 5.2.

Monte Carlo methods introduce some random noise into the results and in

figures 5.1–5.4 the values shown are averaged over 20 realisations of the Markov

processes. The estimated 95% confidence intervals for the volume fraction and

surface area concentration are less than ±1% of the plotted average values.

The results shown in figures 5.1 & 5.2 indicate that the differences between

the various models for particle shape are generally smaller than the difference

between the spherical particle model and the closest non-spherical model, this is

confirmed by inspecting other moments of the mass and surface distributions (not

plotted). Two pairs of models lead to particularly close results: models 2 & 7, both

of which are omitted from the plots for clarity as their curves were consistently

5% below those for model 3 in both plots. The second almost indistinguishable

pair were models 3 & 6; only data for model 3 are plotted in this chapter. If these

102

Page 104: Numerical Modelling of Soot Formation

0

1

2

3

4

5

6

7

0 0.5 1 1.5 2 2.5 3 3.5

model 1model 3model 4model 5experimental

soot

vol

ume

frac

tion

f v / 10

-6

HAB / cm

Figure 5.1: JW10.673 soot volume fraction

0

0.5

1

1.5

2

2.5

3

3.5

4

0 0.5 1 1.5 2 2.5 3 3.5

model 1model 3model 4model 5

surf

ace

area

per

uni

t vol

ume

/ cm

2 cm

-3

HAB / cm

Figure 5.2: JW10.673 particle surface area concentration

103

Page 105: Numerical Modelling of Soot Formation

Table 5.2: Summary of simulated models

model name c(x) = rgr (x) = rox (x) =

1 spherical v(x)13 v(x)

13 v(x)

13

2 Balthasar(2.7375× d (x)−

s(x)12

v(x)s(x)

0.825)× v(x)13

3 geometric mean v(x)16 s(x)

14 s(x)

12

v(x)s(x)

4 — v(x)16 s(x)

14 c(x) c(x)

5 Df = 1.8 v(x)−19 s(x)

23 s(x)

12

v(x)s(x)

6 arithmetic 12

(v(x)

13 + s(x)

12

)s(x)

12

v(x)s(x)

7 Mitchell equation (5.6) s(x)12

v(x)s(x)

8 —model 5 with

s(x)12

v(x)s(x)prefactor from (5.18)

were the only models available one would conclude that the exact choice of col-

lision diameter model was not significant provided one did not assume particles

were spherical. In [157] it was noted that C2H2 addition was the dominant process

in soot growth in two flames similar to the one studied here. According to the soot

chemistry model [8, 207], the rate of C2H2 addition is proportional to particle sur-

face area, hence all the non-spherical models predict larger soot volume fractions

than the spherical particle model.

However, the results for the assumed fractal dimension, model 5, show a con-

siderable deviation from all the other collision diameter models. Significant differ-

ences are also seen for model 4 and these will be discussed below. Model 5 does

not have a noticeable effect on number density compared with the other models

(results for number density are not plotted but, after the initial spike, they differ

from the results for the geometric mean collision diameter model by less than 5%),

however, it leads to a significant reduction in surface area and in the soot volume

fraction. The reduction in soot volume fraction predicted by model 5 compared

with the other non-spherical models appears to be due to enhanced OH oxida-

tion rates. The soot chemistry model treats OH oxidation as a collision controlled

104

Page 106: Numerical Modelling of Soot Formation

process with rate proportional to the square of the collision diameter of particles.

Particle collision diameters are generally much larger under model 2 than under

the other models so the OH oxidation rate is also higher. This explanation was

verified by performing simulations with model 2 modified to calculate the OH

oxidation rate with the equivalent volume sphere diameter in place of the particle

collision diameter. The soot volume fraction calculated for the modified model

was within 1% of that calculated using model 3. This illustrates the importance

of good models for particle shape in order to predict correctly reactions between

particles and the surrounding gas phase.

To test the significance of the fractal prefactor discussed in §5.3.1 model 8

was simulated. For JW10.673 the values of soot volume fraction and surface area

were found to lie between those for models 3 and 4, an increase of one third over

model 5. However, the effect on the predicted number of particles was negligible.

Additional calculations were carried out for the 10 bar laminar premixed ethy-

lene flame JW10.60 [93] and soot volume fraction results are plotted in figure 5.3.

Coupled with the very small differences in particle numbers predicted (not plotted)

by models 5 and 8, this data confirms the observation made above that the main

effect of the collision diameter model is to control the OH oxidation rate, which is

particularly high in JW10.60. Since coagulation rates are more sensitive to colli-

sion diameter in the free molecular regime [99, 245] the 1 bar flame JW1.69 [93]

from the same data set as JW10.60 and JW10.673 was simulated. This flame has

significant nucleation at all heights above the burner and therefore a continuous

supply of small particles for coagulation. For this flame particle number, shown in

figure 5.4, was much more sensitive to the change from model 5 to model 8 than

the soot volume fraction. Therefore at lower pressures and when small particles

are present, the effect of collision diameter on coagulation becomes significant.

As mentioned in §5.3.2 it is necessary to investigate the importance of the

models for particle curvature and hence results for model 4 are included in fig-

ures 5.1, 5.2 & 5.3. One sees that as the radii of curvature used are brought closer

to the collision diameter (thus making particles less smooth) the surface reaction

rates increase. In this context it is significant that the soot model [8, 207] treats the

105

Page 107: Numerical Modelling of Soot Formation

0

2

4

6

8

10

12

0 1 2 3 4 5 6

model 1model 3model 4model 5model 8experimental

soot

vol

ume

frac

tion

f v / 10

-8

HAB / cm

Figure 5.3: JW10.60 soot volume fraction

106

Page 108: Numerical Modelling of Soot Formation

109

1010

1011

0 0.5 1 1.5 2 2.5 3 3.5 4

model 1model 5model 8pa

rtic

le n

umbe

r de

nsity

/ cm

-3

HAB / cm

Figure 5.4: JW1.69 number density

107

Page 109: Numerical Modelling of Soot Formation

rate of oxidation by OH radicals, which is the main oxidation process, as depen-

dent on the collision diameter of particles not their surface area. Since oxidation

by O2 is not a very significant process in the simulated systems [157] this means

that the main effect of increased particle surface area is an increase in the rate of

acetylene addition, which can be seen in soot volume fraction in figures 5.1&5.3.

The parallel increase in surface area as the particles get larger can be noted from

figure 5.2. Predicted soot volume fractions were computed using models 1, 3 and

4 at the top of the flames JW10.673, JW10.60, JW1.69 introduced above and also

for the flames A1 and A3 from [240]. With the exception of JW10.60 the dif-

ference between the values predicted by models 1 and 3 was at least three times

greater than the difference between the values predicted by models 3 and 4, the

same trend was observed at all distances from the burners. JW10.60 is unique

among the flames considered in having a fall in soot volume due to oxidation after

the reaction zone, but it is not entirely clear why this should make the results so

sensitive to the surface curvature model. In general, while the uncertainty about

surface curvature is a concern, it does not present a fundamental obstacle to the ap-

plication of the models proposed here to flames with weak oxidation. More work

is clearly needed to understand what happens in strongly oxidising situations.

Observed soot volume fraction results are included in figures 5.1 & 5.3 to

give some idea of the accuracy of the results. One of the main purposes of the

current work was to investigate how much of the disagreement between simulated

and observed values might be due to inadequate models of soot particle structure.

The results in this section show that, where they exist, significant disagreements

between simulated and observed results will not be resolved by new models for

particle structure alone. Such issues will have to be addressed with improved

chemistry models [156, 217], which are an active topic of research [95, 183, 219],

along with better models for particle structure.

5.4.2 Particle Size Distributions

The particle distributions simulated with model 3 at 0.4 cm, 0.6 cm and 4.2 cm

above the burner surface are shown in figures 5.5, 5.6 & 5.7. The distributions are

108

Page 110: Numerical Modelling of Soot Formation

shown in two ways—as a scatter plot in surface-volume space and as a density on

the volume axis. The densities are estimates for

d

dVN(V ) (5.27)

where N (V ) is the number of particles per cm3 with volume ≤ V . They were

calculated from the discrete simulation data using a Gaussian blurring technique;

specifically, with the ‘density’ function of the computer statistics package R [162].

More details can be found in the online documentation and in [199, §5.6]. The

distributions for the other models are qualitatively the same, but with quantitative

differences at large particle sizes.

0 2 4 6 8 10

0.0

0.5

1.0

1.5

2.0

2.5

particle volume / 10-17

cm3

pa

rtic

le s

urf

ace

are

a /

10

-10cm

2

00

.51

1.5

2

vo

lum

e d

istr

ibu

tio

n d

en

sity /

10

27cm

-6

Figure 5.5: JW10.673 distribution simulated with model 3, 0.4 cm above burner

The most important feature of figures 5.5, 5.6 & 5.7 are that, in all three cases,

the distribution is concentrated on a line. From these slopes (the intercept terms

were negligible) one can infer the mean primary particle diameter by assuming

all soot particles are composed of primary particles in point contact. Diameters

calculated in this way are given in table 5.3.

109

Page 111: Numerical Modelling of Soot Formation

0 2 4 6 8 10

05

10

15

particle volume / 10-16

cm3

pa

rtic

le s

urf

ace

are

a /

10

-10cm

2

01

23

45

vo

lum

e d

istr

ibu

tio

n d

en

sity /

10

25cm

-6

Figure 5.6: JW10.673 distribution simulated with model 3, 0.6 cm above burner

0.0 0.5 1.0 1.5 2.0 2.5

01

23

particle volume / 10-15

cm3

pa

rtic

le s

urf

ace

are

a /

10

-9cm

2

02

46

8

vo

lum

e d

istr

ibu

tio

n d

en

sity /

10

24cm

-6

Figure 5.7: JW10.673 distribution simulated with model 3, 4.2 cm above burner

110

Page 112: Numerical Modelling of Soot Formation

Table 5.3: Primary particle diameters calculated by linear regression forJW10.673

Diameter / nmheight above burner 0.4 cm 0.6 cm 1.4 cm 3.4 cm

model 2 24 31 38 39model 3 24 31 38 40model 4 22 30 37 38model 5 24 30 37 38model 8 23 30 36 37

For each height above the burner the primary particle diameters given in ta-

ble 5.3 are quite similar (within 5% of the value for model 2), even for model 4

which leads to a noticeably different soot volume fraction (see figure 5.1). The

closeness of the values would make testing the models by comparing primary

particle diameters to those observed by electron microscopy very difficult. This

supports the view that any reasonable approximation for the aggregate structure

of soot particles is sufficient given the current relative imprecision of the chemical

mechanisms involved in soot formation and growth.

5.4.3 Individual particle behaviour

To gain an understanding of why different models might lead to similar results,

histories of particles in a small volume of gas, moving through the laminar pre-

mixed ethylene flame JW10.673 ([93]) used previously, were extracted from the

simulations. In figure 5.8 the time evolution of the shape descriptor (5.2) is plotted

for two of these particles. Figure 5.10, tracks the size of the same two particles

in the simulation, which used model 2. Figure 5.9, gives a more detailed view

of the very active early life of the particles until just after the point at which they

coagulate with each other (this point is circled on the plot).

On the more detailed plot one sees abrupt upward jumps in the ratio when the

particle undergoes coagulation followed by periods of smooth decrease as surface

growth makes the particle more round according to §5.3.2. The features seen in

111

Page 113: Numerical Modelling of Soot Formation

0.65

0.7

0.75

0.8

0.85

0.1 1 10

particle Aparticle B

shap

e de

scrip

tor

height above burner / cm

x-axis region shown inmore detail in next figure.

Figure 5.8: Evolution of mean shape descriptor for JW10.673, simulated withmodel 2.

0.65

0.7

0.75

0.8

0.85

0.16 0.17 0.18 0.19 0.2 0.21 0.22

particle Aparticle B

shap

e de

scrip

tor

height above burner / cm

Figure 5.9: Detail of mean shape descriptor for JW10.673, simulated withmodel 2.

112

Page 114: Numerical Modelling of Soot Formation

0

20

40

60

80

100

120

0.1 1 10

particle Aparticle B

volu

me

equi

vale

nt d

iam

/ nm

height above burner / cm

Figure 5.10: JW10.673, size history for particles from figure 5.8.

figure 5.8 are also seen in similar plots for particles simulated with other models.

All three plots show an early period of rapid activity in which the shape de-

scriptor undergoes damped oscillations and size grows steadily. The final shape

and size of the particles however, are largely determined by a few coagulation

events after the initial activity has subsided. Figure 5.10 gives a very clear view of

how chemical reactions with the surface of the soot particles all occur relatively

close to the burner face as shown by the smooth growth in particle size. Figure 5.9

illustrates the way that these surface reactions largely negate the shape changing

effects of the coagulation that occurs up to 0.5 cm. Because of this cancelling out,

a detailed model of the rapid processes may not be necessary for many purposes.

For flames similar to JW10.673 it may be sufficient to predict the outcome of the

highly active early phase of the flame, even if the details within that zone are not

resolved correctly. A modest initial peak in the mean number of primary particles

per aggregate has been found in simulations of titania nanoparticles [197], which

is consistent with a situation where some particles have just formed by aggrega-

tive collisions and others have had time for rapid surface growth and restructuring

113

Page 115: Numerical Modelling of Soot Formation

to take effect. Some parametric studies of the competition between particle re-

structuring and coagulation have been reported [100], they differ from the present

work in assuming particle restructuring is a constant mass process. Sadly, these

parametric studies suggest that predicting the location of equilibrium between co-

agulation may be difficult, because the results are very sensitive to the parameters.

That the fine structure of particles remains constant after the initial period of

activity can also be seen from the primary particle diameters in table 5.3. The

mean primary particle diameter changes noticeably over the first 1.5 cm of the

flame but then hardly changes at all for the remainder of the flame. This is in ac-

cordance with figures 5.8 & 5.10 which show that surface processes which round

out particles are only significant early in the flame and further from the burner all

that happens is coagulation in which both surface and volume are conserved.

Shape Descriptor

Figure 5.11 shows the shape descriptor of the first particle to be incepted and an

average taken over the first nineteen particles to be incepted in a single simula-

tion of JW10.673 using model 2. The population average calculated over twenty

simulations is also shown. Unsurprisingly, the first particles to form undergo more

coagulation and hence develop a less spherical structure than particles which form

later.

The difference between the shape descriptor averaged over the first nineteen

particles and that averaged over the complete particle ensemble (between one and

two thousand computational particles) explains some of the features of [14, fig-

ure 3]. In that paper Balthasar and Frenklach observe that for the flame JW10.673

the average shape descriptor of their “collector particle” (one of the first particles

to be incepted, see [135] for details) is much larger than the ensemble average

calculated with their MoMIC approximation which is about 0.7 after 0.1 s and

reaches about 0.72 after 0.2 s. These results tend to confirm the accuracy of their

MoMIC approximation even though it assumes all particles have the same shape

descriptor (which changes with time) and show that their Monte Carlo results are

only applicable for the first few particles to be incepted. It should, however, be

114

Page 116: Numerical Modelling of Soot Formation

0.66

0.68

0.7

0.72

0.74

0.76

0.78

0.8

0.82

0.02 0.04 0.06 0.08 0.1

first particle, 1 runfirst 19 particle average, 1 runensemble average, 10 runs

shap

e de

scrip

tor

gas residence time / s

Figure 5.11: Mean shape descriptors for JW10.673, simulated with model 2

115

Page 117: Numerical Modelling of Soot Formation

noted that the flame JW10.673, for which this comparison has been performed,

has a unimodal particle size distribution and that the MoMIC tends to perform

well in such cases, but struggles with bimodal flames.

5.4.4 Future Experimental Validation

The ultimate test of the models presented in this work, is the foundational ques-

tion of modern science—does the model offer a predictive description of reality?

Three ways in which the simple models advanced in the present work might even-

tually be tested are:

• by calculating the volume, area and primary particle diameters of soot ag-

gregates using TEM [194] or SEM [39] methods to compare with the dis-

tributions generated in the present work. This would provide a very direct

test of the models, and since it could be done at different points in flames

should be able to identify the onset of any divergence between the models

and the experiments.

• by using the particle collision diameter models introduced in the present

work to calculate particle mobility according to the parameterisation of

[108, 109] and comparing the results with SMPS [122, 238] measurements.

One should be aware that with such an approach it might prove rather diffi-

cult to identify the effects of inaccuracies in flame chemistry.

• by using the particle length scale models to estimate the distribution of radii

of gyration and thus to predict scattering results [185], whether of visible

band light, x-rays [63] or neutrons [132, 241].

5.5 Summary

Predictions of simple, bivariate models of aggregate soot formation, which reflect

the non-spherical nature of soot particles, have been found to be quantitatively

different from those of a single variable model, which assumes particle sphericity.

116

Page 118: Numerical Modelling of Soot Formation

Numerical investigations indicate that several collision diameter models derived

in different ways lead to very similar bivariate distributions and, in particular, the

model presented in [14] is consistent with some simple models suggested here.

Therefore, the simple particle shape models considered offer a useful starting

point for more detailed modelling of soot surface chemistry. The importance of in-

cluding particle shape is clearly illustrated by the difference between the spherical

particle model and all the extensions which conserve surface area during coagula-

tion. More work on surface curvature, to understand how particles change shape

as they undergo surface reactions and whether either the detailed aggregate model

or a bivariate model can capture this, is needed.

117

Page 119: Numerical Modelling of Soot Formation

Chapter 6

Explicit Statistical Weights

This chapter introduces an alternative to the treatment of coagulation that is at

the core of the DSA and its LPDA acceleration. In chapter 4 it was seen that

the performance of LPDA was limited by the need to simulate the non-deferred

processes. In chapter 5 it was noted that all surface reaction processes could be

deferred so that only particle inception and coagulation had to be treated accord-

ing to the basic DSA. Therefore, statistically weighted particle methods, which

have been successfully used to accelerate other simulations, are developed. Two

weighted particle algorithms are implemented and validated. They are compared

with each other and with DSA/LPDA simulations that use particle doubling as a

variance reduction technique. One weighted algorithm is found consistently to

outperform the other by a significant amount. The better weighted algorithm is

found to offer broadly similar performance and accuracy to direct simulation with

particle doubling.

6.1 Background

In the basic Direct Simulation Monte Carlo (DSMC) approach to the Smolu-

chowski and Boltzmann equations, every computational particle represents the

same number of physical particles. The accuracy with which a quantity is com-

puted depends on the number of computational particles used. In spatially re-

118

Page 120: Numerical Modelling of Soot Formation

solved gas simulation, regions with few physical particles are not accurately sim-

ulated. To deal with this problem Rjasanow and Wagner [169] introduced a

Stochastic Weighted Particle Method (SWPM) in which computational particles

were tagged with a statistical weight—the number of physical particles they repre-

sent. Using this technique a larger number of computational particles with smaller

statistical weights could be used in regions of low density than was the case with

basic DSMC, without increasing the time spent simulating regions of higher den-

sity. As a result, computational accuracy could be controlled separately for each

spatial cell (for example kept approximately equal at all locations) and indepen-

dently of the values of physical properties being simulated.

A similar kind of difficulty can occur in spatially homogeneous particle coag-

ulation problems. For these problems it is found that the resolution of the high end

of the particle size distribution in a DSMC simulation can be inadequate because

there are very few computational particles and thus the statistical noise is great.

A particle weighting designed for such coagulation problems was proposed by

Eibeck and Wagner [48], in which the statistical weight of a computational parti-

cle was a function of the physical particle it described and so there was no need to

explicitly tag the computational particles with their statistical weights. The par-

ticular statistical weight used was the inverse of the physical particle mass, which

has the useful consequence that each computational particle represents the same

mass of physical particles per unit volume. In systems with constant total mass

one would therefore expect the number of computational particles to remain con-

stant in mean. By some judicious algebra the Mass Flow Algorithm (MFA) was

derived so that the number of computational particles was constant in an absolute

rather than just mean sense [48].

The MFA effectively redistributes computational particles over the size space,

placing more computational particles at larger particle sizes and fewer at smaller

sizes as compared with the constant weighting DSMC method. Consequently

the high end of the particle size distribution is calculated more accurately at the

expense of the low end calculations. A more general consideration of statisti-

cal weights as functions of the physical particle properties was undertaken by

119

Page 121: Numerical Modelling of Soot Formation

Kolodko and Sabelfeld [97], whose comparisons included the special case of the

Mass Flow Algorithm (MFA) [48], but the greater generality came at the price of

coagulation events that sometimes increased and sometimes decreased the number

of computational particles.

Wells and Kraft [214] applied the MFA to a coagulation (and sintering) prob-

lem in nanoparticle dynamics. Consideration of systems of engineering interest

inevitably led to a desire to simulate systems which exchange mass with their

surroundings and thus a MFA cannot be expected to maintain a constant num-

ber of computational particles, even in mean, without continual resampling of

the distribution. Nevertheless in [138, 139] MFA was successfully extended to

systems with particle inception and where individual particle masses changed by

interaction with the environment. The treatment of this last class of processes—

surface growth—is set out in [139]. In essence, surface growth was simulated as

two separate processes. In one, the mass of physical particles referred to by a

computational particle is changed and in the other, new computational particles

are introduced to the system to account for the increase in mass. Analogous pro-

cesses which remove mass from the system would be treated by having the second

process remove computational particles.

In [36], where the physical motivation was atmospheric aerosols, computa-

tional particles were assigned weights that were not simply functions of the phys-

ical particles they represented. Using operator splitting and deterministic integra-

tion, surface processes were treated by updating the statistical weight of a compu-

tational particle so that it represented the same number of physical particles both

before and after the surface events. Coupled with the MFA approach to coagu-

lation this meant that the only change in the number of computational particles

during simulations was due to processes of particle inception.

Maintaining a constant number of computational particles or at least main-

taining a lower bound, is essential for controlling the variance of the Monte Carlo

sample solutions [97]. This issue is discussed in §2.3.1, where the procedure of

ensemble doubling, used in the previous chapters of this thesis to avoid catas-

trophic losses of precision, is introduced. The work of Eibeck and Wagner [48],

120

Page 122: Numerical Modelling of Soot Formation

referred to above, took the more indirect and mathematical approach of formu-

lating a new stochastic process that could solve the same coagulation problems

as before, but without the drop in computational particle numbers associated with

basic direct simulation. Variance control was also addressed by Haibo et al. [76],

who introduced statistical weights to arrive by a slightly more informal route at

the ‘w2’ method (independently) derived in this chapter. The same authors have

introduced a further, similar method for the same purpose [242].

6.2 General Approach

The work in this chapter is aimed, like the rest of this thesis, at solving soot for-

mation problems in laminar premixed flames according to the model in §1.4. Pre-

vious chapters have established the power of the LPDA for the surface reactions

in the soot model. This chapter therefore starts from the basic LPDA formulation

of the soot population problem (1.4), first given in (3.9) and repeated here—find

measures λt to solve

d

dt

∫x∈E

φ (x) λt (dx) =

∫x∈E

φ (x) It (dx)

+∑l∈U ′

∫x,ξ∈E

[φ(g(l) (ξ)

)− φ (x)

(l)t (ξ)P (Rtx = dξ) λt (dx)

+1

2

∫x,y,ξ,ζ∈E

[φ (ξ + ζ)− φ (x)− φ (y)]Kt (ξ, ζ)

P (Rtx = dξ)P (Rty = dζ) λt (dx) λt (dy) .

(6.1)

Progress is then made by exploiting weighted particle methods within this frame-

work. The use of the update operator from equation (3.6) in order to make mea-

sures such as λ physically interpretable is implicit in the following sections.

121

Page 123: Numerical Modelling of Soot Formation

6.3 Weighting

Let W be a set and Ω : W → R+ a map that defines how many physical particles

an element of W indicates. In practice W = R+ and Ω = id would seem to be the

natural choices. LetW be a σ−algebra on W which makes Ω measurable. Note

that it is not possible to talk about statistical weights at this stage since there are

no statistics—(3.9) is a deterministic equation. However, one can look for maps ν

analogous to λ but taking values that are measures on W × E and satisfying∫W×E

Ω (w)φ (x) νt (dw, dx) =

∫E

φ (x)λt (dx) (6.2)

for all t ∈ [0, T ].

It is convenient to extend the Rt to act on W × E and to extend the It to be

measures on the product σ−algebra on W × E. The extensions, Rt and It must

have the same physical interpretation as the original forms, so one requires∫E

φ (x) It (dx) =

∫W×E

φ (x) Ω (u) It (du, dx) (6.3)

and

Ω (u)

∫E

φ (ξ)P (Rtx = dξ) =

∫W×E

Ω (w)φ (ξ)P(Rt (u, x) = (dw, dξ)

)∀ (u, x) ∈ W × E. (6.4)

The simplest form for Rt is

Rt (u, x) = (u,Rtx) (6.5)

and this form will be used throughout this work. It is clearly the most natural

extension—R represents the effects of processes that act on individual particles

independently and therefore would not be expected to alter the number of physical

particles being modelled.

122

Page 124: Numerical Modelling of Soot Formation

Extensions of1 β and K so that their final arguments can be measures on W ×E rather than just E are also needed. If ν is a measure on W × E then let λ be the

measure on E given by

λ (A) =

∫W×A

Ω (u, x) ν (du, dx) ∀A ∈ E (6.6)

and

β (x; ν) := β (x;λ) , (6.7)

K (x, y; ν) := K (x, y;λ) . (6.8)

A simple extension of It is also possible:

It (du, dx) :=δw (du)

Ω (w)It (dx) . (6.9)

This approach will be used in this chapter but it should be noted that w may be

chosen in arbitrary ways, for example varying with time and with x [170].

6.3.1 Dynamics of the New Measure

To proceed to a stochastic particle algorithm one requires the dynamics of ν from

(6.2). Consider

ψ : W × E → R; (w, x) 7→ Ω (w)φ (x) (6.10)

so ∫W×E

ψ (w, x) νt (dw, dx) =

∫E

φ (x)λt (dx) ∀ t (6.11)

and in particular

d

dt

∫W×E

ψ (w, x) νt (dw, dx) =d

dt

∫E

φ (x)λt (dx) . (6.12)

By expanding the right hand side of (6.12) according to (3.9), repeatedly sub-

1The labels (l) on β and g will be omitted from the following presentation to avoid the notationbecoming too cluttered.

123

Page 125: Numerical Modelling of Soot Formation

stituting (6.11) and assuming νt ((u, x) : Ω (u) = 0) = 0 ∀ t one has

d

dt

∫W×E

ψ (u, x) νt (du, dx) = J1 + J2 + J3 (6.13)

where

J1 :=

∫(u,x)∈W×E

ψ (u, x) It (du, dx) , (6.14)

J2 :=

∫(W×E)

2

[ψ (w, g (ξ))

Ω (w)− ψ (u, x)

Ω (u)

]βt (ξ; νt) Ω (u)

P

(Rt (u, x) = (du′, dξ)

)νt (du, dx)

(6.15)

and

J3 :=1

2

∫(W×E)

4

[ψ (w, ξ + ζ)

Ω (w)− ψ (u, x)

Ω (u)− ψ (v, y)

Ω (v)

]K (ξ, ζ, νt) Ω (u) Ω (v)P

(Rt (u, x) = (du′, dξ)

)P

(Rt (v, y) = (dv′, dζ)

)νt (du, dx) νt (dv, dy) .

(6.16)

In (6.15) & (6.16) the variables w appear as free parameters so (6.13) holds

whatever values of w (provided Ω (w) > 0) are substituted into J2 and J3. There-

fore one may choose the rules for calculating the w from the u′ and v′ to opti-

mise the stochastic particle algorithm that is being derived. To derive such an

algorithm one wants to express all the integrands in (6.14)–(6.16) in the form∑i ψ (ui, xi)−

∑j ψ (vj, yj) multiplied by some rate expression and to integrate

over all the (vj, yj). The differences of sums become the definitions of the jumps

of the particle algorithm and the rest of the integral becomes the jump rate. The

variables u′ and v′ found in (6.14)–(6.16) are dummy variables, which represent

the weight component of Rt (u, x) and Rt (v, y) respectively.

124

Page 126: Numerical Modelling of Soot Formation

As stated in (6.5) u′ = u and v′ = v with probability 1 so (6.15) can be

simplified by choosing w = u′ = u to give

J2 =

∫E×(W×E)

[ψ (u, g (ξ))− ψ (u, x)] βt (ξ, νt)P (Rtx = dξ) νt (du, dx) .

(6.17)

There is more than one way to proceed from (6.16), two approaches that lead

to different simulation procedures are given here. Both, by exploiting symmetry

in (u, x) and (v, y), yield, for their respective definitions of w,

J3 =

∫E2×(W×E)

2[ψ (w, ξ + ζ)− ψ (u, x)] K (ξ, ζ, νt) Ω (v)

P (Rtx = dξ)P (Rty = dζ) νt (du, dx) νt (dv, dy) .

(6.18)

The first approach is to choose w such that

1

Ω (w)=

1

Ω (u)+

1

Ω (v). (6.19)

(This is a possible choice for at least the simplest choices of W and Ω—consider

W = R+ with Ω a positive multiple of the identity then w = (u−1 + v−1)

−1.) The

second approach is to choose w such that

Ω (w) =Ω (u)

2, (6.20)

which is also possible if W = R+ and Ω is a positive multiple of the identity. The

idea of halving the statistical weight has previously been used by Haibo et al. [76]

as mentioned in §6.1.

125

Page 127: Numerical Modelling of Soot Formation

6.3.2 Simulation Algorithms

All the equations considered above can be thought of as deterministic, mean field

equations [4]. Stochastic algorithms are derived (see, for example, [164, §4.6])

by taking (6.13) as defining the generator of a Markov process [90, ch 19], after

introducing a scaling parameter to control the level of discretisation. The relation-

ship of pure jump Markov processes to the mean field equations has been exten-

sively studied in papers such as [35, 49–51, 74, 147] and their references. These

and other papers typically prove that the trajectories of a sequence of pure jump

Markov processes converge in some sense to a solution of the deterministic mean

field equation. The work reported in the present chapter was conceived as a more

practical attempt to improve a computer program already in use for solving soot

problems in chemistry and engineering. As such, investigation of convergence is

confined to numerical tests.

In this work the discretisation level or scaling may be controlled through It as

follows: Choose a sequence (wN) in W such that Ω (wN) > 0 and Ω (wN) 0;

define

INt (du, dx) :=

δwN(du)

Ω (wN)It (dx) . (6.21)

Then replacing I with IN and substituting (6.18)&(6.17) for J2 and J3 in (6.13)

leads to a sequence of equations

d

dt

∫W×E

ψ (u, x) νt (du, dx) =

∫(u,x)∈W×E

ψ (u, x) INt (du, dx)

+

∫(W×E)

2[ψ (u, g (ξ))− ψ (u, x)] βt (ξ, νt)

P

(Rt (u, x) = (du′, dξ)

)νt (du, dx)

+

∫(W×E)

4[ψ (w, ξ + ζ)− ψ (u, x)] K (ξ, ζ, νt) Ω (v)P

(Rt (u, x) = (du′, dξ)

)P

(Rt (v, y) = (dv′, dζ)

)νt (du, dx) νt (dv, dy) .

(6.22)

126

Page 128: Numerical Modelling of Soot Formation

A sequence of generators for pure jump Markov processes can then be derived

from (6.22) in the standard way. The rates of the events which make up the

Markov processes are given in table 6.1. In table 6.1, and for the rest of the

chapter, W = R+ and Ω is the identity mapping. Under reasonable conditions,

the trajectories of these processes can be expected, as N → ∞, to converge to

solutions of the mean field problem. Throughout the remainder of the chapter

Table 6.1: Process summary

−→ (wN , x) INt ((wN , x))

(u, x) −→ (u, g (ξ)) P

(Rt (u, x) = ξ

)β (ξ) ν ((u, x))

(u, x) −→ (w, ξ + ζ)K (ξ, ζ, ν)P (Rtx = ξ)P (Rty = ζ)

×ν ((u, x)) ν ((v, y))

the two rules for calculating w for coagulation products will be denoted ‘w1’ and

‘w2’ as specified in table 6.2.

Table 6.2: Coagulation weight rules

w1 w−1 = u−1 + v−1

w2 w = u/2

6.4 Numerical Tests

6.4.1 Initial Validation

Validation of the new weighted algorithms began using problems from chapter 2,

for which the entire particle size distribution can be calculated using an ODE

solver. Direct solution of the population balance equations is possible because all

but the first few thousand size classes can be neglected. Initial tests simulated only

inception and coagulation processes, specifically the conditions were

127

Page 129: Numerical Modelling of Soot Formation

100

104

106

108

1010

0 500 1000 1500 2000 2500 3000

ODE solutionw1w1 modified infloww2

part

icle

num

ber

dens

ity /

cm-3

particle size / # C atoms

Figure 6.1: PSD for coagulation and inception test problem

1. physical particle inception rate It = 2.63× 1013 ×(

0.05−t0.05

)2 cm−3 s−1,

2. constant temperature of 500 K

3. constant pressure of 600 bar.

The particle size distributions are presented in figure 6.1 and show good agreement

between the ODE solver results and all the stochastic weighted particle meth-

ods. Both rules for the weight of the coagulation product (see §6.3.2, table 6.2)

were used with computational particle inflow proportional to the inception rate

128

Page 130: Numerical Modelling of Soot Formation

for physical particles, that is, with w from (6.9) constant. These two data sets

are labelled ‘w1’ and ‘w2’ in the legend of figure 6.1. In the nucleation and co-

agulation only case considered here, w1 corresponds to the mass flow algorithm

of [49]. The w1 coagulation rule was also tested with a modified inflow rule; in

this case the rate of inflow of computational particles was constant throughout the

simulation, but the statistical weight of the new computational particles was au-

tomatically adjusted to simulate the physical particle inception rate. Adjustment

of the statistical weight of particles entering a system has previously been used in

simulations of the Boltzmann equation [170].

The statistical noise associated with the methods was also considered. Fig-

ure 6.2 summarises the results. The 95% confidence interval sizes were calcu-

lated from a central limit theorem estimate based on 30 realisations of the Markov

chain for each simulation method, each realisation used just under 216 computa-

tional particles. The results in figure 6.2 indicate that w2 is generally more noisy

than w1 and that there is little difference between the two inflow methods used

with w1. The data is an initial indication that w1 is to be preferred to w2 since

fewer realisations of the w1 Markov chain than of the w2 Markov chain would be

needed to obtain the same size of confidence interval.

Further comparisons to the particle size distribution produced by the ODE

solver were performed for a test problem including a surface reaction (pyrene

condensation). Good agreement was found between the weighted particle meth-

ods and the deterministic solution of the size distribution for the limited case for

which this was possible.

6.4.2 LPDA and real flames

Having established that the algorithms and their implementations worked cor-

rectly, in limited cases for which the population balance equations could be solved

directly, testing moved on to premixed laminar flames. For these tests the LPDA

as described in chapter 3 was employed throughout, including for the DSA calcu-

lation used to provide a reference solution. The first flame based test compared

the accuracy of the moments of the soot particle size distribution calculated with

129

Page 131: Numerical Modelling of Soot Formation

0.01

0.1

1

10

0 500 1000 1500 2000 2500 3000

w1w1 modified infloww295

% c

onfid

ence

inte

rval

hal

f wid

th /

%

particle size / # C atoms

Figure 6.2: Statistical noise for coagulation and inception test problem

130

Page 132: Numerical Modelling of Soot Formation

the w1 and w2 weightings for the flame JW10.68 [93]. The second moment of the

mass distribution is shown in figure 6.3, 95% confidence intervals for the DSA

and w1 data are within ±2% of the plotted values and so confidence intervals are

only shown for w2. The calculations for the weighted algorithms were performed

0

5

10

15

20

0 1 2 3 4 5 6 7

DSAw1w2 meanw2 confidence bounds

seco

nd m

omen

t of m

ass

dist

ribut

ion

/ 10-2

2 g2 c

m-3

height above burner / cm

Figure 6.3: Second moment of JW10.68 mass distribution

with 30 runs using around 2000 computational particles from the end of the incep-

tion peak until the end of each simulation. To ensure no error was introduced by

the deferral of processes all computational particles were updated every time the

simulation covered 5 × 10−4 s of real time. The large difference in the statistical

variability generated by the two weighting methods should be observed. The w2

method leads to a variance for the second moment of the size distribution that is

more than 10 times larger than that obtained with the w1 method. The situation

with the zeroth and first moments is similar.

131

Page 133: Numerical Modelling of Soot Formation

The real attraction of stochastic particle methods is that they provide an ex-

plicit estimate of the particle size distribution. As a test case for the distribution

the flame JW1.69 [93] was used. This flame is known to have a bimodal particle

size distribution [16] and therefore to present an interesting test case for the way

in which the w1 weighting method transfers computational effort to larger particle

sizes. The w2 method was not used for this comparison, because the results above

suggest that far more realisations of the w2 Markov chain would be required than

of the w1 Markov chain in order to achieve the same precision. Therefore, to meet

any particular error tolerance, less computer time would be required using the w1

method.

Densities were calculated using the statistical computation package R [162]

by performing Gaussian blurring of the observations with a bandwidth of 0.0245.

The densities presented here were calculated in logarithmic size space, that is,

they are (estimates of)d

dlog10 xN (log10 x) (6.23)

where N (log10 x) is the number of particles per cubic centimetre comprising no

more than x Carbon atoms. Data calculated from 50 repetitions of the w1 Markov

chain with just under 213 computational particles are compared to data from high

precision DSA (without LPDA) calculations which used 30 repetitions with be-

tween 216 and 216 computational particles.The results in figure 6.4 show a very

high degree of agreement between the two algorithms, these particular data apply

to the top of the flame—about 4.2 cm above the burner.

In figure 6.4 large oscillations in the density generated from the w1 data can

be seen for particle sizes between 300 and 10,000 carbon atoms. These oscilla-

tions are a symptom of the way the weighted algorithm transfers computational

resolution to large particle sizes as discussed in the next paragraph and illustrated

in figure 6.5. The weighted method clearly would be rather unsuitable for this

flame if the number of particles containing 300–10,000 carbon atoms was the

main quantity of interest. In such a case a DSA method with particle doubling

[110, 114] as a variance reduction technique should be used if possible. However,

as will be seen for other measures of solution accuracy, the weighted method can

132

Page 134: Numerical Modelling of Soot Formation

105

106

107

108

109

1010

1011

100 1000 104 105 106 107 108

dsaw1

dens

ity o

f par

ticle

num

ber

dist

ribut

ion

/ cm

-3

particle size / # C atoms

Figure 6.4: Physical particle size distribution for JW1.69

133

Page 135: Numerical Modelling of Soot Formation

offer as good or better performance than the un-weighted alternative.

It is also interesting to look at the distribution of computational particles on the

size spectrum, since the number of particles is what controls the precision of the

calculation. In figure 6.5 the normalised densities of the computational particle

distribution for the calculations used for figure 6.4 are plotted. The normalisation

ensures that the area under both the w1 and the DSA curve is 1 (when integrated

against d (log10 x)) and so there are no effects due to the different numbers of

computational particles used with the two algorithms. Figure 6.5 gives a very

0

1

2

3

4

5

6

7

10 100 1000 104 105 106 107 108

dsaw1

norm

alis

ed d

ensi

ty

particle size / # C atoms

Figure 6.5: Relative computational particle distribution for JW1.69

clear view of the way in which w1 and DSA concentrate computational effort on

different parts of the size range. This shows that the choice of algorithm will de-

pend, to some extent, on the problem that is being solved. Problems that mainly

concern the largest particles are likely to be addressed best using a weighted algo-

rithm, problems concerning the smallest particles should be tackled with a DSA.

The remainder of this chapter attempts to investigate this choice in a quantitative

134

Page 136: Numerical Modelling of Soot Formation

way.

6.4.3 Performance

Simulations of the flame HWA3 [240] were performed using DSA and the w1

weighting. The first set of tests reported ignored the acetylene, OH and O2 sur-

face reactions since simulation of these processes takes a significant amount of

time and is the same whether or not weighted particles are used. The results

obtained in this way give no information about the soot produced by the flame

but provide a comparison that should focus a little more on the properties of the

weighted algorithms. Simulation size is described by the maximum number of

computational particles in the simulation, settings were chosen so that most of

this capacity was used. The initial sample volume (DSA) and the weighting in Itwere chosen to use almost all the capacity of the binary tree, in which the com-

putational particles were stored, at the point when the physical particle number

peaked. For DSA particle doubling [157] was used so that the tree was never less

than 50% full. For the weighted methods doubling was not needed as the number

of computational particles did not decrease, this being one of the attractions of the

weighted method, see (6.18).

Table 6.3 summarises performance on the simplified flame; run times were

measured on the same desktop PC which has a 2 GHz Athlon XP CPU (2400+).

The memory requirement of the simulations are low, only a few MB are required,

even for the largest simulations. In table 6.3 the standard deviation of the popula-

tion of samples for certain functionals of the solution are given for 1.34 cm above

the burner, which is approximately the end of the flame. The functionals used are

the zeroth, first, second, third moments of the mass distribution (denoted m0, m1,

m2, m3 respectively) and the number of particles containing 5000–6000 Carbon

atoms.

From table 6.3 one sees that, for a given tree size, DSA simulations take

around 75% of the computational time. For the zeroth moment DSA yields an

estimate that is only half as variable as the w1 approach. However, the advantage

of DSA drops as one moves to higher moments and by the third moment DSA

135

Page 137: Numerical Modelling of Soot Formation

Table 6.3: Variability of algorithms for reduced HWA3

tree method time per population std. dev. /%size run / s m0 m1 m2 m3 5–6×103

212 w1 0.6 9.7 6.4 15.0 25.4 11.1214 w1 2.5 4.4 3.6 7.7 12.0 6.5216 w1 10.4 3.0 1.8 3.9 6.0 2.9212 DSA 0.5 5.4 4.2 14.6 34.9 24.4214 DSA 1.9 2.7 2.3 7.6 19.5 11.7216 DSA 7.6 1.5 1.1 4.0 9.6 7.4

is significantly inferior to w1. The value of the higher moments is primarily de-

termined by the larger particles in the distribution and it is not surprising the w1

offers an advantage in this case since it increases the computational resolution for

this part of the distribution. The variance in the estimates of the number of par-

ticles containing 5000–6000 Carbon atoms is an even more extreme example of

the way in which w1 resolves the larger sizes better (at the expense of the smaller

ones) compared with DSA.

For all the functionals both algorithms appeared to have converged in mean to

the true value for the largest tree sizes reported in table 6.3. This was verified by

performing larger simulations on other hardware, which were not timed, and so

are not reported in detail. These results suggest that, for some functionals, which

heavily emphasize the distribution of larger particles w1 offers a faster way of

getting good estimates than DSA.

6.4.4 Further Comparison

The same tests were carried out, with the same flame, HWA3, but including all re-

actions on the surfaces of soot particles by means of the LPDA. A couple of results

for w2 are included for interest but fit the pattern discussed above and will not re-

ceive any further attention. Stochastic simulation of this flame is of considerable

interest because measured particle size distributions have been published in [240].

Computation times are not comparable with those from table 6.3 because different

136

Page 138: Numerical Modelling of Soot Formation

hardware was used. The simulations with the full flame model take much longer

because of the high rates of surface reactions and so Opteron 252 processors run-

ning at 2.6 GHz in 64 bit mode were used for the computations. The results are

summarised in table 6.4.

Table 6.4: Variability of algorithms for physical system

tree method time per population std. dev. /%size run / s m0 m1 m2 m3 9–10×105

212 w1 4.4 9.6 1.9 3.8 6.2 18.9214 w1 18.0 5.2 1.1 2.3 3.6 9.8216 w1 73.7 2.4 0.6 1.3 2.1 4.9218 w1 274 1.1 0.2 0.4 0.7 2.1214 w2 18.9 5.3 3.3 2.3 7.5 36.1216 w2 75.1 2.2 1.5 3.4 5.7 16.8212 DSA 3.0 4.6 1.1 2.9 7.1 39.1214 DSA 12.2 3.0 0.5 1.4 3.3 19.6216 DSA 49.5 1.3 0.3 0.6 1.8 9.3218 DSA 213 0.7 0.1 0.4 1.0 5.3

In common with the results for the simplified problem DSA is seen to be

around one third faster than the w1 approach for a given tree size. All the sets of

simulations produced reasonably accurate estimates of the quantities considered

in table 6.4: For the moments, the mean from 80 repetitions with each tree size

was within 1% of the values from extremely high precision calculations. For the

number of particles containing 9–10×105 C atoms, the difference between the

mean and the high precision solution just reached 4% in some cases, which is

not statistically significant. As in the simplified case DSA generates less statis-

tically noisy estimates for the first few moments of the size distribution but w1

becomes more attractive for functionals that place a greater stress on the largest

sizes of particles. However, for the reduced case, w1 was significantly less noisy

for the third moment (m3) than DSA for a given tree size, but for the full flame

the crossover is only just beginning at m3. It can be seen that for the number of

particles in the size range 9–10×105 the w1 algorithm produces an estimate with

137

Page 139: Numerical Modelling of Soot Formation

roughly the same variance as the DSA with 4 times the number of particles. Ex-

amination of the ‘time per run’ column of table 6.4 shows that w1 can therefore

provide an estimate, of any given precision, of the number of particles in the size

range 9–10×105 in roughly one third of the time of DSA.

6.4.5 Potential Applications

While the weighted particle algorithm presents an alternative to the DSA with

particle doubling for simulations of Smoluchowski’s coagulation equation for spa-

tially homogeneous systems, it also has other potential applications where it might

clearly distinguish itself from the DSA. One application, which has already re-

ceived considerable attention for the purposes of gas dynamics simulation, is the

simulation of particle populations in a grid of cells. For such problems that ability

to control the weighting explicitly is important to capture effects in regions with

low particle densities [91], when a small proportion of the population has very im-

portant effects [170]. Weighted particle methods would therefore seem attractive

for simulations of spatially resolved coagulating systems, a purpose for which the

MFA has already been used [73].

Explicit weights also simplify the implementation of particle transport be-

tween cells by accounting automatically for differences in the statistical weight

assigned to computational particles in different cells and facilitating conservative

resampling of particle populations [200]. It is also possible to exploit weighting

to adjust computational resolution independently of the main kinetic simulation

process by resampling the computational particle population. An application of

resampling would be to construct a different computational resolution profile from

the one shown in figure 6.5 in order to move the statistical noise seen at particle

sizes in the range 300–10,000 carbon atoms in figure 6.4 to a different size range.

6.5 Summary

A general weighted form of the Smoluchowski coagulation equation with addi-

tional linear terms has been formulated. From this equation two new weighted

138

Page 140: Numerical Modelling of Soot Formation

particle simulation algorithms have been derived. Implementations of these algo-

rithms has been successfully tested on a range of problems and one of the algo-

rithms has been found to be up to three times faster than direct simulation.

139

Page 141: Numerical Modelling of Soot Formation

Chapter 7

Conclusion

Finally, a short review of the work in this thesis is presented and a suggestion

made regarding how to make further progress in understanding soot formation

and growth.

In chapter 2 a general Direct Simulation Monte Carlo (DSMC) algorithm for

the simulation of soot formation at all pressures was developed and shown to

produce accurate results by comparisons with results from a standard ODE solver.

Detailed investigations have been carried out into the computational demands of

this DSMC algorithm, which showed that most of the computation time was spent

using complicated techniques, designed for the non-linear coagulation process, to

simulate linear surface growth. To solve the problem thus revealed, a modified

stochastic algorithm (LPDA) was formulated, in which surface reaction processes

were deferred. The generator of the corresponding Markov chain was presented

for the first time in chapter 3. The LPDA was found to be superior to operator

splitting methods and tests show it accelerates computations for flames of physical

interest by a factors of up to one thousand, while causing little loss of accuracy.

Attempts to achieve further accelerations by deterministic approximations to

the surface reaction processes in chapter 4 yielded little additional progress. As

shown in that chapter, the computational bottleneck in LPDA is the un-accelerated

un-deferred processes. Therefore, further accelerations will only be possible with

a new approach to the processes which the LPDA cannot accelerate. An exam-

140

Page 142: Numerical Modelling of Soot Formation

ple of a possible approach using statistically weighted computational particles is

given in chapter 6. The weighted algorithm was seen to have some potentially

useful properties and found to offer similar performance to the un-weighted direct

simulation algorithm with the particle doubling variance reduction technique as

used in this thesis.

The LPDA is well suited to model development, because the addition of new

features to particle models requires no change to the basic programme structure

and has little effect on computational cost. These properties have been exploited

by colleagues [179] and a detailed demonstration was given in chapter 5. In that

chapter it was shown that simple models for soot particle shape are quantitatively

similar and offer qualitative improvements over older models which treat all soot

particles as spherical. The chapter also highlights the need for a better understand-

ing of the details of the chemical reactions on the surfaces of soot particles and

work on this topic was briefly reviewed in §1.2.1.

Progress in the future would be greatly aided by detailed and extensive ex-

perimental studies of a small number of systems. Quality rather than quantity of

experimental data is needed for thorough calibration and validation of the detailed

models that can now be tested with Monte Carlo simulations.

141

Page 143: Numerical Modelling of Soot Formation

Acknowledgements

The author is grateful for the support of his supervisor, Dr M Kraft, and other

colleagues, who provided the foundations on which this thesis is built. He also

wishes to thank his parents for their dedication in correcting the drafts of the

thesis, a task in which Messrs M S Celnik and M H Sankey also kindly assisted.

Chapter 5 is based on an article, which completed the peer review process

as this thesis was being completed. A considerable amount of valuable advice

was received from the anonymous reviewers. Discussions with Mr N M Morgan

were very helpful in formulating the connection between the weighted geometric

average collision diameters and assumed fractal dimension models in §5.3.1.

Discussions with Dr W Wagner were very helpful in developing the work

reported in chapter 6.

Finally, the author thanks God for making such an interesting world for him

to investigate.

142

Page 144: Numerical Modelling of Soot Formation

List of Figures

2.1 Size distributions calculated with ODE solver and stochastically. . 41

2.2 Comparison of calculations and observations for flames. . . . . . 43

2.3 4 level binary tree . . . . . . . . . . . . . . . . . . . . . . . . . . 47

2.4 Simulation run time scaling with binary tree . . . . . . . . . . . . 52

2.5 % of execution time spent on different tasks—DSA . . . . . . . . 53

3.1 Number density and second mass moment for JW1.69 . . . . . . . 61

3.2 Particle size distribution at end of flame JW10.68 . . . . . . . . . 64

3.3 Percentage of execution time spent on different tasks with ‘pv’

method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

3.4 LPDA applied to JW1.69 . . . . . . . . . . . . . . . . . . . . . . 74

3.5 Particle distribution for JW10.68 . . . . . . . . . . . . . . . . . . 76

4.1 % of execution time spent on different tasks—LPDA with ‘pv’ . . 87

5.1 JW10.673 soot volume fraction . . . . . . . . . . . . . . . . . . . 103

5.2 JW10.673 particle surface area concentration . . . . . . . . . . . 103

5.3 JW10.60 soot volume fraction . . . . . . . . . . . . . . . . . . . 106

5.4 JW1.69 number density . . . . . . . . . . . . . . . . . . . . . . . 107

5.5 JW10.673 distribution simulated with model 3, 0.4 cm above burner109

5.6 JW10.673 distribution simulated with model 3, 0.6 cm above burner110

5.7 JW10.673 distribution simulated with model 3, 4.2 cm above burner110

5.8 Evolution of mean shape descriptor for JW10.673, simulated with

model 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

143

Page 145: Numerical Modelling of Soot Formation

5.9 Detail of mean shape descriptor for JW10.673, simulated with

model 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

5.10 JW10.673, size history for particles from figure 5.8. . . . . . . . . 113

5.11 Mean shape descriptors for JW10.673, simulated with model 2 . . 115

6.1 PSD for coagulation and inception test problem . . . . . . . . . . 128

6.2 Statistical noise for coagulation and inception test problem . . . . 130

6.3 Second moment of JW10.68 mass distribution . . . . . . . . . . . 131

6.4 Physical particle size distribution for JW1.69 . . . . . . . . . . . 133

6.5 Relative computational particle distribution for JW1.69 . . . . . . 134

144

Page 146: Numerical Modelling of Soot Formation

List of Tables

2.1 Scaling of run times with tree depth . . . . . . . . . . . . . . . . 50

2.2 Illustrative results . . . . . . . . . . . . . . . . . . . . . . . . . . 51

2.3 Relative frequency of stochastic events in JW1.69 . . . . . . . . . 52

3.1 Detail of operator splitting runs for JW1.69 . . . . . . . . . . . . 60

3.2 Relative frequency of stochastic events in JW10.68 . . . . . . . . 62

3.3 Detail of operator splitting runs for JW10.68 . . . . . . . . . . . . 63

3.4 Detail of deferred surface growth runs for JW10.68 . . . . . . . . 75

3.5 Comparison of simulation methods . . . . . . . . . . . . . . . . . 77

3.6 Time to achieve tolerances for JW1.69 . . . . . . . . . . . . . . . 77

4.1 Numerical errors in JW10.68 distribution moments . . . . . . . . 83

4.2 Run times in seconds for different algorithms on the same hardware 83

4.3 Numerical errors in JW10.68 distribution moments . . . . . . . . 84

4.4 Numerical errors in JW10.68 distribution moments . . . . . . . . 88

4.5 Run times in seconds for different sub-algorithms on the same

hardware . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

5.1 Length scales for typical JW10.673 particle sizes . . . . . . . . . 102

5.2 Summary of simulated models . . . . . . . . . . . . . . . . . . . 104

5.3 Primary particle diameters calculated by linear regression for JW10.673111

6.1 Process summary . . . . . . . . . . . . . . . . . . . . . . . . . . 127

6.2 Coagulation weight rules . . . . . . . . . . . . . . . . . . . . . . 127

6.3 Variability of algorithms for reduced HWA3 . . . . . . . . . . . . 136

145

Page 147: Numerical Modelling of Soot Formation

6.4 Variability of algorithms for physical system . . . . . . . . . . . . 137

146

Page 148: Numerical Modelling of Soot Formation

Bibliography

[1] Computational Modelling Group. URL http://como.cheng.

cam.ac.uk/index.php?Page=Resources\&Section=

SootDatabase. Cambridge Soot Database.

[2] G. L. Agafonov, M. Nullmeier, P. A. Vlasov, J. Warnatz, and I. S. Zaslonko.

Kinetic modeling of solid carbon particle formation and thermal decompo-

sition during carbon suboxide pyrolysis behind shock waves. Combust. Sci.

Tech., 174:185–213, 2002.

[3] K. Ahmad. Pollution cloud over south Asia is increasing ill health. Lancet,

360:549, 2002.

[4] D. J. Aldous. Deterministic and stochastic models for coalescence (aggre-

gation and coagulation) : a review of the mean-field theory for probabilists.

Bernoulli, 5, 1999.

[5] A. Alexiadis, M. Vanni, and P. Gardin. Extension of the method of moments

for population balances involving fractional moments and application to a

typical agglomeration problem. J. Colloid Interf. Sci., 276:106–112, 2004.

doi:10.1016/j.jcis.2004.03.052.

[6] V. Alopaeus, M. Laakkonen, and J. Aittamaa. Solution of popula-

tion balances with breakage and agglomeration by high-order moment-

conserving method of classes. Chem. Eng. Sci., 61:6732–6752, 2006.

doi:10.1016/j.ces.2006.07.010.

147

Page 149: Numerical Modelling of Soot Formation

[7] B. Apicella, A. Carpentieri, M. Alfe, R. Barbella, A. Tregrossi, P. Pucci,

and A. Ciajolo. Mass spectrometric analysis of large PAH in a fuel-rich

ethylene flame. Proc. Combust. Inst., 31(1):547–553, 2001.

[8] J. Appel, H. Bockhorn, and M. Frenklach. Kinetic modeling of soot forma-

tion with detailed chemistry and physics: Laminar premixed flames of C2

hydrocarbons. Combust. Flame, 121:122–136, 2000.

[9] J. Appel, H. Bockhorn, and M. Wulkow. A detailed numerical study of

the evolution of soot particle size distributions in laminar premixed flames.

Chemosphere, 42, 2001. doi:10.1016/S0045-6535(00)00237-X.

[10] C. Arden Pope III and D. W. Dockery. Health effects of fine particulate

air pollution: Lines that connect. J. Air Waste Mgmt. Assoc., 56:709–742,

2006.

[11] C. Artelt, H. J. Schmid, and W. Peukert. On the relevance of accounting

for the evolution of the fractal dimension in aerosol process simulations. J.

Aerosol Sci., 34:511–534, 2003.

[12] G. A. Athanassoulis and P. N. Gavriliadis. The truncated Hausdorff mo-

ment problem solved by using kernel density functions. Probablist. Eng.

Mech., 17(3):273–291, 2002. doi:10.1016/S0266-8920(02)00012-7.

[13] I. Ayrancıa, R. Vaillon, N. Selsuka, F. Andre, and D. Escudie. Determina-

tion of soot temperature, volume fraction and refractive index from flame

emission spectrometry. J. Quant. Spectrosc. Ra., 104(2):266–276, 2007.

[14] M. Balthasar and M. Frenklach. Detailed kinetic modeling of soot aggre-

gate formation in laminar premixed flames. Combust. Flame, 140:130–145,

2005.

[15] M. Balthasar and M. Frenklach. Monte-Carlo simulation of soot particle

coagulation and aggregation: the effect of a realistic size distribution. Proc.

Combust. Inst., (30):1467–1475, 2005.

148

Page 150: Numerical Modelling of Soot Formation

[16] M. Balthasar and M. Kraft. A stochastic approach to solve the particle size

distribution function of soot particles in laminar premixed flames. Combust.

Flame, 133:289–298, 2003.

[17] M. Balthasar, M. Kraft, and M. Frenklach. Kinetic Monte-Carlo simula-

tions of soot particle aggregation. American Chemical Society Preprint

Papers, 50(1):135–136, 2005.

[18] A. C. Barone, A. D’Alessio, and A. D’Anna. Morphological characteri-

zation of the early process of soot formation by atomic force microscopy.

Combust. Flame, 132:181–187, 2003.

[19] F. Barra, B. Davidovitch, A. Levermann, and I. Procaccia.

Laplacian growth and diffusion limited aggregation: Differ-

ent universality classes. Phys. Rev. Lett., 87(13):134501, 2001.

doi:10.1103/PhysRevLett.87.134501.

[20] J. C. Barrett and N. A. Webb. A comparison of some approximate methods

for solving the aerosol general dynamic equation. J. Aerosol Sci., 29(1/2):

31–39, 1998. doi:10.1016/S0021-8502(97)00455-2.

[21] G. A. Bird. Molecular gas dynamics. Clarendon press, 1976.

[22] G. A. Bird. Approach to translational equilibrium in a rigid sphere gas.

Phys. Fluids, 6(10):1518–1519, 1963.

[23] Y. Bouvier, C. Mihesan, M. Ziskind, E. Therssen, C. Focsa, J. F. Pauwels,

and P. Desgroux. Molecular species adsorbed on soot particles issued from

low sooting methane and acetylene laminar flames: A laser-based experi-

ment. Proc. Combust. Inst., 31(1):841–849, 2007.

[24] A. Braun, N. Shah, F. E. Huggins, K. E. Kelly, A. Sarofim, C. Jacobsen,

S. Wirick, H. Francis, J. Ilavsky, G. E. Thomas, and G. P. Huffman. X-ray

scattering and spectroscopy studies on diesel soot from oxygenated fuel

under various engine load conditions. Carbon, 43:2588–2599, 2005.

149

Page 151: Numerical Modelling of Soot Formation

[25] P. Brimblecombe and C. M. Grossi. Aesthetic thresholds and blackening of

stone buildings. Sci. Total Environ., 349:175–189, 2005.

[26] M. Celnik, R. Patterson, M. Kraft, and W. Wagner. Cou-

pling a stochastic soot population balance to gas-phase chemistry

using operator splitting. Combust. Flame, 148(3):158–176, 2007.

doi:10.1016/j.combustflame.2006.10.007.

[27] B. B. Chakraborty and R. Long. Gas chromotagraphic analysis of poly-

cyclic aromatic hydrocarbons in soot samples. Environ. Sci. Tech., 1:828,

1967.

[28] D. R. Chen, D. Y. H. Pui, D. Hummes, H. Fissan, F. R. Quant, and G. J.

Sem. Design and evaluation of a nanometer aerosol differential mobility

analyzer (nano-DMA). J. Aerosol Sci., 29(5-6):497–509, 1998.

[29] Computational Modelling Group. sweep. URL http://como.cheng.

cam.ac.uk/index.php?Page=Resources\&Section=

Software.

[30] A. D’Anna and A. Violi. Detailed modeling of the molecular growth pro-

cess in aromatic and aliphatic premixed flames. Energ. Fuel., 19:79–86,

2005. doi:10.1021/ef0499675.

[31] A. D’Anna, M. Commodo, S. Violi, C. Allouis, and J. Kent. Nano organic

carbon and soot in turbulent non-premixed ethylene flames. Proc. Combust.

Inst., 31(1):621–629, 2007.

[32] S. de Iuliis, F. Cignoli, and G. Zizak. Two-color laser-induced incandes-

cence (2C-LII) technique for absolute soot volume fraction measurements

in flames. Appl. Optics, 44(34):7414–7423, 2005.

[33] S. de Iuliis, F. Migliorini, F. Cignoli, and G. Zizak. 2d soot volume frac-

tion imaging in an ethylene diffusion flame by two-color laser-induced in-

candescence (2C-LII) technique and comparison with results from other

optical diagnostics. Proc. Combust. Inst., 31(1):869–876, 2007.

150

Page 152: Numerical Modelling of Soot Formation

[34] M. Deaconu, N. Fournier, and E. Tanre. A pure jump markov process

associated with Smoluchowski’s coagulation equation. Ann. Prob., 30(4):

1763–1796, 2002.

[35] M. Deaconu, N. Fournier, and E. Tanre. Rate of Convergence of a Stochas-

tic Particle System for the Smoluchowski Coagulation Equation. Meth.

Comput. App. Prob., 5(2):131–158, 2003.

[36] E. Debry, B. Sportisse, and B. Jourdain. A stochastic approach for the

numerical simulation of the general dynamics equation for aerosols. J.

Comput. Phys., 184:649–669, 2003.

[37] S. di Stasio. Feasibility of an optical experimental method for the sizing of

primary spherules in sub-micron agglomerates by polarized light scattering.

Appl. Phys. B, 70:635–643, 2000.

[38] S. di Stasio. Electron microscopy evidence of aggregation under three dif-

ferent size scales for soot nanoparticles in flame. Carbon, 39:109–118,

2001.

[39] S. di Stasio, A. G. Konstandopoulos, and M. Kostoglou. Clustercluster

aggregation kinetics and primary particle growth of soot nanoparticles in

flame by light scattering and numerical simulations. J. Colloid Interf. Sci.,

247:33–46, 2002. doi:10.1006/jcis.2001.8095.

[40] S. di Stasio, J. B. A. Mitchell, J. L. LeGarrec, L. Biennier, and M. Wulff.

Synchrotron SAXS in situ identification of three different size modes for

soot nanoparticles in a diffusion flame. Carbon, 44:1267–1279, 2006.

doi:10.1016/j.carbon.2005.10.042.

[41] R. B. Diemer and J. H. Olson. A moment methodology for coagulation and

breakage problems: Part 2–moment models and distribution reconstruction.

Chem. Eng. Sci., 57:2211–2228, 2002.

151

Page 153: Numerical Modelling of Soot Formation

[42] C. A. Dorao and H. A. Jakobsen. A least squares method for the solution

of population balance problems. Comput. Chem. Eng., 30:535–547, 2006.

doi:10.1016/j.compchemeng.2005.10.012.

[43] C. A. Dorao and H. A. Jakobsen. The quadrature method of moments and

its relationship with the method of weighted residuals. Chem. Eng. Sci., 61:

7795–7804, 2006. doi:10.1016/j.ces.2006.09.014.

[44] C. A. Dorao and H. A. Jakobsen. Timespace-property least squares spectral

method for population balance problems. Chem. Eng. Sci., 62:1323–1333,

2007. doi:10.1016/j.ces.2006.11.016.

[45] C. A. Dorao and H. A. Jakobsen. hp-adaptive least squares spectral ele-

ment method for population balance equations. Appl. Numer. Math., 2007.

doi:10.1016/j.apnum.2006.12.005. in press.

[46] K. Drumma, D. I. Attiab, S. Kannta, P. Mickea, R. Buhla, and K. Kienasta.

Soot-exposed mononuclear cells increase inflammatory cytokine mRNA

expression and protein secretion in cocultured bronchial epithelial cells.

Respiration, 67:291–297, 2000.

[47] Y. Efendiev and M. R. Zachariah. Hierarchical hybrid Monte-Carlo method

for simulation of two-component aerosol nucleation, coagulation and phase

segregation. J. Aerosol Sci., 34:169–188, 2003.

[48] A. Eibeck and W. Wagner. Stochastic particle approximation for Smolu-

chowski’s coagulation equation. Ann. Appl. Probab., 11(4):1137–1165,

2001.

[49] A. Eibeck and W. Wagner. An efficient stochastic algorithm for studying

coagulation dynamics and gelation phenomena. SIAM J. Sci. Comput., 22:

802–821, 2000.

[50] A. Eibeck and W. Wagner. Approximative solution of the coagulation-

fragmentation equation by stochastic particle systems. Stoch. Anal. Appl.,

18(6):921–948, 2000.

152

Page 154: Numerical Modelling of Soot Formation

[51] A. Eibeck and W. Wagner. Stochastic interacting particle systems and non-

linear kinetic equations. Ann. Appl. Probab., 13(3):845–889, 2003.

[52] A. M. El-Leathy, C. H. Kim, G. M. Faeth, and F. Xu. Soot surface reactions

in high-temperature laminar diffusion flames. AIAA Journal, 42:988–996,

2004.

[53] W. D. Erickson, G. C. Williams, and H. C. Hottel. Light scattering mea-

surements on soot in a benzene-air flame. Combust. Flame, 8(2):89–170,

1964.

[54] P. S. Fennel, J. S. Dennis, and A. N. Hayhurst. The size distributions of

nanoparticles of the oxides of Mg, Ba and Al in flames: Their measurement

and dependence on the concentrations of free radicals in the flame. Proc.

Combust. Inst., 31(2):1939–1945, 2007. doi:10.1016/j.proci.2006.07.137.

[55] K. A. Fichthorn and W. H. Weinberg. Theoretical foundations of dynamical

Monte Carlo simulations. J. Chem. Phys., 95(2):1090–1096, 1991.

[56] F. Findik, R. Yilmaz, and T. Koksal. Investigation of mechanical and phys-

ical properties of several industrial rubbers. Materials and Design, 25:

269–276, 2004.

[57] M. Frenklach. Method of moments with interpolative closure. Chem. Eng.

Sci., 57:2229–2239, 2002.

[58] M. Frenklach. Reaction mechanism of soot formation in flames. Phys.

Chem. Chem. Phys., 4:2028–2037, 2002. doi:10.1039/b110045a.

[59] M. Frenklach and S. J. Harris. Aerosol dynamics modeling using the

method of moments. J. Colloid Inter. Sci., 118(1):252–261, 1987.

[60] M. Frenklach and H. Wang. Soot Formation in Combustion: Mechanisms

and Models, pages 165–192. Springer Verlag, 1994.

153

Page 155: Numerical Modelling of Soot Formation

[61] M. Frenklach, C. A. Schuetz, and J. Ping. Migration mechanism

of aromatic-edge growth. Proc. Combust. Inst., 30:1389–1396, 2005.

doi:10.1016/j.proci.2004.07.048.

[62] N. A. Fuchs. The Mechanics of Aerosols. Pergammon Press, Oxford, 1964.

Original edition in Russian published in 1955.

[63] C. Gardner, G. N. Greaves, G. K. Hargrave, S. Jarvis, P. Wildman, F. Me-

neau, W. Bras, and G. Thomas. In situ measurements of soot formation in

simple flames using small angle X-ray scattering. Nucl. Instrum. Meth. B,

238:334–339, 2005. doi:10.1016/j.nimb.2005.06.072.

[64] F. Gelbard and J. H. Seinfeld. Numerical solution of the dynamic equa-

tion for particulate systems. J. Comput. Phys., 28(3):357–375, 1978.

doi:10.1016/0021-9991(78)90058-X.

[65] F. Gelbard and J. H. Seinfeld. Simulation of multicomponent aerosol dy-

namics. J. Coll. Interf. Sci., 78(2):485–501, 1980.

[66] D. T. Gillespie. Approximate accelerated stochastic simulation of chemi-

cally reacting systems. J. Chem. Phys., 114(4):1716–1733, 2001.

[67] D. T. Gillespie. An exact method for numerically simulating the stochastic

coalescence process in a cloud. J. Atmos. Sci., 32:1977–1989, 1975.

[68] M. Goodson and M. Kraft. An efficient stochastic algorithm for simulating

nano-particle dynamics. J. Comput. Phys., 183:210–232, 2002.

[69] A. R. Gourlay. Splitting methods for time dependent partial differential

equations. In D. Jacobs, editor, The State of the Art in Numerical Analysis,

pages 757–796, London, 1977. Academic Press.

[70] S. C. Graham, J. B. Homer, and J. L. J. Rosenfeld. The formation and

coagulation of soot aerosols generated by the pyrolysis of aromatic hydro-

carbons. Proc. Royal Soc., 344:259–285, 1975.

154

Page 156: Numerical Modelling of Soot Formation

[71] R. Grosch, H. Briesen, W. Marquardt, and M. Wulkow. Generalization

and numerical investigation of QMOM. AIChE J., 53(1):207–227, 2007.

doi:10.1002/aic.11041.

[72] H.-H. Grotheer, H. Pokorny, K.-L. Barth, M. Thierley, and M. Aigner. Mass

spectrometry up to 1 million mass units for the simultaneous detection of

primary soot and of soot precursors (nanoparticles) in flames. Chemo-

sphere, 57:1335–1342, 2004.

[73] F. Guias. A stochastic numerical method for diffusion equations and appli-

cations to spatially inhomogeneous coagulation processes. In H. Niederre-

iter and D. Talay, editors, Monte Carlo and Quasi-Monte Carlo Methods

2004, pages 147–161. Springer, 2006. doi:10.1007/3-540-31186-6 10.

[74] F. Guias. A Monte Carlo approach to the Smoluchowski equations. MCMA,

3(4):313–326, 1997.

[75] R. Gunawan, I. Fusman, and R. D. Braatz. High resolution algorithms for

multidimensional population balance equations. AIChE J., 50:2738–2749,

2004.

[76] Z. Haibo, Z. Chuguang, and X. Minghou. Multi-Monte Carlo ap-

proach for general dynamic equation considering simultaneous parti-

cle coagulation and breakage. Powder Tech., 154:164–178, 2005.

doi:10.1016/j.powtec.2005.04.042.

[77] R. A. Hamilton, J. S. Curtis, and D. Ramkrishna. Beyond log-normal dis-

tributions: Hermite spectra for solving population balances. AIChE J., 49:

2328–2343, 2003.

[78] W. N. Hartley and H. Ramage. The mineral constituents of dust and soot

from various sources. Proc. Royal Soc., 68:97–109, 1901.

[79] J. P. Hessler, S. Seifert, R. E. Winans, and T. H. Fletcher. Small-angle x-

ray studies of soot inception and growth. Faraday Discuss., 119:395–407,

2001. doi:10.1039/b102822g.

155

Page 157: Numerical Modelling of Soot Formation

[80] M. J. Hounslow, R. L. Ryall, and V. R. Marshall. A discretized population

balance for nucleation, growth, and aggregation. AIChE Journal, 34(11):

1821–1832, 1988.

[81] K. Hoyermann, F. Mauß, and T. Zeuch. A detailed chemical reaction

mechanism for the oxidation of hydrocarbons and its application to the

analysis of benzene formation in fuel-rich premixed laminar acetylene

and propene flames. Phys. Chem. Chem. Phys., 6:3824–3835, 2004.

doi:10.1039/b404632c.

[82] Q. Hu, S. Rohani, and A. Jutan. New numerical method for solving the

dynamic population balance equations. AIChE J., 51(11):3000–3006, 2005.

[83] H. M. Hulburt and S. Katz. Some problems in particle technology: A

statistical mechanical formulation. Chem. Eng. Sci., 19:555–574, 1964.

[84] C. D. Immanuel and F. J. Doyle III. Computationally efficient solution of

population balance models incorporating nucleation, growth and coagula-

tion: application to emulsion polymerization. Chem. Eng. Sci., 58:3681–

3698, 2003. doi:10.1016/S0009-2509(03)00216-1.

[85] C. D. Immanuel and F. J. Doyle III. Solution technique for a multi-

dimensional population balance model describing granulation processes.

Powder Tech., 156:213–225, 2005. doi:10.1016/j.powtec.2005.04.013.

[86] S. Izvekov and A. Violi. A coarse-grained molecular dynamics study of

carbon nanoparticle aggregation. J. Chem. Theory Comput., 2:504–512,

2005. doi:10.1021/ct060030d.

[87] S. Izvekov, A. Violi, and G. A. Voth. Systematic coarse-graining of

nanoparticle interactions in molecular dynamics simulation. J. Phys. Chem.

B, 109:17019–17024, 2005. doi:10.1021/jp0530496.

[88] M. Z. Jacobson and J. H. Seinfeld. Evolution of nanoparticle size and

mixing state near the point of emission. Atmospheric Environment, 38:

1839–1850, 2004.

156

Page 158: Numerical Modelling of Soot Formation

[89] J. I. Jeong and M. Choi. A sectional method for the analysis of growth

of polydisperse non-spherical particles undergoing coagulation and co-

alescence. J. Aerosol Sci., 32(5):565–582, 2001. doi:10.1016/S0021-

8502(00)00103-8.

[90] O. Kallenberg. Foundations of modern probability. Springer, New York, 2

edition, 2001.

[91] K. C. Kannenberg and I. D. Boyd. Strategies for efficient particle resolution

in the direct simulation Monte Carlo method. J. Comput. Phys., 157:727–

745, 2000. doi:10.1006/jcph.1999.6397.

[92] A. Kazakov and M. Frenklach. Dynamic modeling of soot particle coagu-

lation and aggregation: Implementation with the method of moments and

application to high-pressure laminar premixed flames. Combust. Flame,

114:484–501, 1998.

[93] A. Kazakov, H. Wang, and M. Frenklach. Detailed modeling of soot forma-

tion in laminar premixed ethylene flames at a pressure of 10 bar. Combust.

Flame, 100:111–120, 1995.

[94] T. R. Kiehl, R. M. Mattheyses, and M. K. Simmons. Hybrid simulation of

cellular behaviour. Bioinformatics, 20(3):316–322, 2004.

[95] C. H. Kim, A. M. El-Leathy, F. Xu, and G. M. Faeth. Soot surface growth

and oxidation in laminar diffusion flames at pressures of 0.1-1.0 atm. Com-

bust. Flame, 136:191–207, 2004.

[96] A. A. Koelmans, M. T. O. Jonker, G. Cornelissen, T. D. Bucheli, P. C.

M. V. Noort, and O. Gustafsson. Black carbon: The reverse of its dark side.

Chemosphere, 63:365–377, 2006.

[97] A. Kolodko and K. Sabelfeld. Stochastic particle methods for Smolu-

chowski coagulation equation: variance reduction and error estimations.

MCMA, 9(4):315–339, 2003.

157

Page 159: Numerical Modelling of Soot Formation

[98] B. Koo, T. M. Gaydos, and S. N. Pandis. Evaluation of the equilibrium,

dynamic, and hybrid aerosol modeling approaches. Aerosol Sci. Tech., 37

(1):53–64, 2003. doi:10.1080/02786820300893.

[99] M. Kostoglou and A. G. Konstandopoulos. Evolution of aggregate size and

fractal dimension during Brownian coagulation. J. Aerosol Sci, 32:1399–

1420, 2001.

[100] M. Kostoglou, A. G. Konstandopoulos, and S. K. Friedlander. Bi-

variate population dynamics simulation of fractal aerosol aggregate co-

agulation and restructuring. J. Aerosol Sci., 37:1102–1115, 2006.

doi:10.1016/j.jaerosci.2005.11.009.

[101] U. O. Koylu, G. M. Faeth, T. L. Farias, and M. G. Carvalho. Fractal and

projected structure properties of soot aggregates. Combust. Flame, 100:

621–633, 1995.

[102] U. O. Koylu, C. S. McEnally, D. E. Rosner, and L. D. Pfefferle. Simulta-

neous measurements of soot volume fraction and particle size / microstruc-

ture using a thermophoretic sampling technique. Combust. Flame, 109:

488–500, 1996.

[103] A. Kubota, C. J. Mundy, W. J. Pitz, C. Melius, C. K. Westbrook, and M.-J.

Caturla. Massively parallel combined Monte Carlo and molecular dynam-

ics methods to study the long-time-scale evolution of particulate matter and

molecular structures under reactive flow conditions. In Proceedings of the

Third Joint Meeting of the U.S. Sections of The Combustion Institute, 2004.

[104] S. Kumar and D. Ramkrishna. On the solution of population balance equa-

tions by discretization—I. A fixed pivot technique. Chem. Eng. Sci., 51(8):

1311–1332, 1996.

[105] S. Kumar and D. Ramkrishna. On the solution of population balance equa-

tions by discretization—III. Nucleation, growth and aggregation of parti-

cles. Chem. Eng. Sci., 52(24):4659–4679, 1996.

158

Page 160: Numerical Modelling of Soot Formation

[106] M. Lapuerta, R. Ballesteros, and F. J. Martos. A method to determine the

fractal dimension of diesel soot agglomerates. J. Colloid Interf. Sci., 303:

149–158, 2006. doi:10.1016/j.jcis.2006.07.066.

[107] J. C. Lee, H. N. Najm, S. Lefantzi, J. Ray, M. Frenklach, M. Val-

orani, and D. A. Goussis. A CSP and tabulation-based adap-

tive chemistry model. Combust. Theor. Model., 11(1):73–102, 2007.

doi:10.1080/13647830600763595.

[108] Z. Li and H. Wang. Drag force, diffusion coefficient, and electric mobility

of small particles. I. Theory applicable to the free-molecule regime. Phys.

Rev. E, 68:061206, 2003.

[109] Z. Li and H. Wang. Drag force, diffusion coefficient, and electric mobility

of small particles. II. Application. Phys. Rev. E, 68:061207, 2003.

[110] K. Liffman. A direct simulation Monte-Carlo method for cluster coagula-

tion. J. Comput. Phys., 100(1):116–127, 1992.

[111] G. Ma, J. Z. Wen, M. F. Lightstone, and M. J. Thomson. Optimization of

soot modeling in turbulent nonpremixed ethylene/air jet flames. Combust.

Sci. Tech., 177:1567–1602, 2005. doi:10.1080/00102200590956786.

[112] U. Maas and J. Warnatz. Ignition processes in hydrogen-oxygen mixtures.

Combust. Flame, 74:53–69, 1988.

[113] A. W. Mahoney and D. Ramkrishna. Effcient solution of population bal-

ance equations with discontinuities by finite elements. Chem. Eng. Sci., 57:

1107–1119, 2002.

[114] A. Maisels, F. E. Kruis, and H. Fissan. Direct simulation Monte Carlo

for simultaneous nucleation, coagulation and surface growth in dispersed

systems. Chem. Eng. Sci, 59:2231–2239, 2004.

159

Page 161: Numerical Modelling of Soot Formation

[115] U. K. Mandal and S. Aggarwal. Studies of rubber-filler interaction in

carboxylated nitrile rubber through microhardness measurement. Polymer

Testing, 20:305–311, 2001.

[116] N. V. Mantzaris, P. Daoutidis, and F. Srienc. Numerical solution of multi-

variable cell population balance models. II. Spectral methods. Comput.

Chem. Eng., 25:1441–1462, 2001. doi:10.1016/S0098-1354(01)00710-4.

[117] N. V. Mantzaris, P. Daoutidis, and F. Srienc. Numerical solution of

multi-variable cell population balance models. III. Finite element meth-

ods. Comput. Chem. Eng., 25:1463–1481, 2001. doi:10.1016/S0098-

1354(01)00711-6.

[118] S. L. Manzello and M. Y. Choi. Morphology of soot collected in micro-

gravity droplet flames. Int. J. Heat Mass Tran., 45(5):1109–1116, 2002.

doi:10.1016/S0017-9310(01)00164-8.

[119] D. L. Marchisio and R. O. Fox. Solution of population balance equations

using the direct quadrature method of moments. J. Aerosol Sci., 36:43–73,

2005. doi:10.1016/j.jaerosci.2004.07.009.

[120] M. M. Maricq. Coagulation dynamics of fractal-like soot aggregates. J.

Aerosol Sci, 38:141–156, 2007.

[121] M. M. Maricq and N. Xu. The effective density and fractal dimension of

soot particles from premixed flames and motor vehicle exhaust. J. Aerosol

Sci, 35:1251–1274, 2004.

[122] M. M. Maricq, S. J. Harris, and J. J. Szente. Soot size distributions in rich

premixed ethylene flames. Combust. Flame, 132:328–342, 2003.

[123] T. Matsoukas and Y. Lin. Fokker-planck equation for particle

growth by monomer attachment. Phys. Rev. E., 74:031122, 2006.

doi:10.1103/PhysRevE.74.031122.

160

Page 162: Numerical Modelling of Soot Formation

[124] J. L. Mauderly. Diesel emissions: Is more health research still needed?

Toxicol. Sci., 62(1):6–9, 2001.

[125] C. S. McEnally, L. D. Pfefferle, B. Atakan, and K. Kohse-Hoinghaus. Stud-

ies of aromatic hydrocarbon formation mechanisms in flames: Progress to-

wards closing the fuel gap. Prog. Energ. Combust., 32:247–294, 2006.

[126] R. McGraw and D. L. Wright. Chemically resolved aerosol dynamics for

internal mixtures by the quadrature method of moments. J. Aerosol Sci.,

34:189–209, 2003.

[127] P. Meakin. Effects of cluster trajectories on cluster—cluster aggrega-

tion: A comparison of linear and Brownian trajectories in two- and three-

dimensional simulations. Phys. Rev. A, 29(2):997–999, 1984.

[128] P. Meakin and R. Jullien. The effects of restructuring on the geome-

try of clusters formed by diffusion-limited, ballistic, and reaction-limited

cluster—cluster aggregation. J. Chem. Phys., 89(1):246–250, 1988.

[129] L. A. Melton. Soot diagnostics based on laser heating. Appl. Optics, 23

(13):2201–2208, 1984.

[130] J. A. Miller, M. J. Pilling, and J. Troe. Unravelling combustion mecha-

nisms through a quantitative understanding of elementary reactions. Proc.

Combust. Inst., 30(1):43–88, 2005. doi:10.1016/j.proci.2004.08.281.

[131] S. E. Miller and S. C. Garrick. Nanoparticle coagulation in a planar jet.

Aerosol Sci. Tech., 38:79–89, 2004. doi:10.1080/02786820490247669.

[132] J. B. A. Mitchell, J. L. L. Garrec, A. I. Florescu-Mitchell, and

S. di Stasio. Small-angle neutron scattering study of soot particles in

an ethylene–air diffusion flame. Combust. Flame, 145:80–87, 2006.

doi:10.1016/j.combustflame.2005.12.003.

[133] P. Mitchell and M. Frenklach. Particle aggregation with simultaneous sur-

face growth. Phys. Rev. E, 67:061407, 2003.

161

Page 163: Numerical Modelling of Soot Formation

[134] P. Mitchell and M. Frenklach. Monte Carlo simulation of soot aggregation

with simultaneous surface growth - why primary particles appear spheri-

cal. In 27th Symposium (International) on Combustion, pages 1507–1514.

Combustion Institute, 1998.

[135] P. A. Mitchell. Monte Carlo Simulation of Soot Aggregation with Simul-

taneous Surface Growth. PhD thesis, University of California, Berkeley,

2001.

[136] E. G. Moody and L. R. Collins. Effect of mixing on the nucleation

and growth of titania particles. Aerosol Sci. Tech., 37:403–424, 2003.

doi:10.1080/02786820390125179.

[137] G. E. Moore. Cramming more components onto integrated circuits. Elec-

tronics, 38, 1965. URL ftp://download.intel.com/museum/

Moores_Law/Articles-Press_Releases/Gordon_Moore_

1965_Article.pdf.

[138] N. Morgan, C. Wells, M. Kraft, and W. Wagner. Modelling nanoparticle

dynamics: coagulation, sintering, particle inception and surface growth.

Combust. Theor. Model., 9(3):449–461, 2005.

[139] N. Morgan, C. Wells, M. Goodson, M. Kraft, and W. Wagner. A new

numerical approach for the simulation of the growth of inorganic nanopar-

ticles. J. Comput. Phys., 211(2):638–658, 2006.

[140] N. Morgan, M. Kraft, M. Balthasar, D. Wong, M. Frenklach, and

P. Mitchell. Numerical nimulations of soot aggregation in premixed lami-

nar flames. Proc. Combust. Inst., 31(1):693–700, 2007.

[141] N. M. Morgan, R. I. A. Patterson, A. Raj, and M. Kraft. Modes of neck

growth in nanoparticle aggregates. Technical Report 45, c4e Preprint-

Series, Cambridge, 2007.

162

Page 164: Numerical Modelling of Soot Formation

[142] H. Muhlenweg, A. Gutscha, A. Schilda, and S. E. Pratsinis. Process simu-

lation of gas-to-particle-synthesis via population balances: Investigation of

three models. Chem. Eng. Sci., 57:2305–2322, 2002.

[143] H. Muhr, R. David, J. Villermaux, and P. H. Jezequel. Crystallization and

precipitation engineering-VI. Solving population balance in the case of the

precipitation of silver bromide crystals with high primary nucleation rates

by using the first order upwind differentiation. Chem. Eng. Sci., 51(2):

309–319, 1996. doi:10.1016/0009-2509(95)00257-X.

[144] I. Naydenova, M. Nullmeier, J. Warnatz, and P. A. Vlasov. Detailed kinetic

modelling of soot formation during shock-tube pyrolysis of C6H6: Direct

comparison with the results of time-resolved laser-induced incandescence

(LII) and cw-laser extinction measurements. Combust. Sci. Tech., 176(10):

1667–1703, 2004. doi:10.1080/00102200490487544.

[145] M. Nicmanis and M. J. Hounslow. A finite element analysis of the steady

state population balance equation for particulate systems: Aggregation and

growth. Comput. Chem. Engng., 29(Suppl.):S261–S266, 1996.

[146] M. Nicmanis and M. J. Hounslow. Finite-element methods for steady-state

population balance equations. AIChE J., 44:2258–2272, 1998.

[147] J. R. Norris. Smouchowski’s coagulation equation: uniqueness, nonunique-

ness and a hydrodynamic limit for the stochastic coalescent. Ann. Appl.

Probab., 9(1):78–109, 1999.

[148] C. Oh and C. M. Sorensen. The effect of overlap between monomers on

the determination of fractal cluster morphology. J. Colloid Interf. Sci., 193:

17–25, 1997.

[149] B. Oktem, M. P. Tolocka, B. Zhao, H. Wang, and M. V. Johnston. Chemical

species associated with the early stage of soot growth in a laminar premixed

ethylene-oxygen-argon flame. Combust. Flame, 142:364–373, 2005.

163

Page 165: Numerical Modelling of Soot Formation

[150] E. S. Oran, C. K. Oh, and B. Z. Cybyk. Direct simulation Monte Carlo:

Recent advances and applications. Annu. Rev. Fluid Mech., 30:403–441,

1998.

[151] F. Ossler and J. Larsson. Exploring the formation of carbon-

based molecules, clusters and particles by in situ detection of scat-

tered x-ray radiation. Chem. Phys. Lett., 387:367–371, 2004.

doi:10.1016/j.cplett.2004.02.046.

[152] F. Ossler and J. Larsson. Measurements of the structures of nanoparticles

in flames by in situ detection of scattered x-ray radiation. J. Appl. Phys.,

98:114317, 2005. doi:10.1063/1.2140080.

[153] E. Otto, H. Fissan, S. H. Park, and K. W. Lee. The log-normal size distri-

bution theory of Brownian aerosol coagulation for the entire particle size

range: Part II Analytical solution using Dahneke’s coagulation kernel. J.

Aerosol Sci., 30(1):17–34, 1999.

[154] S. H. Park and S. N. Rogak. A one-dimensional model for coagulation,

sintering, and surface growth of aerosol agglomerates. Aerosol Sci. Tech.,

37(12):947–960, 2003.

[155] S. H. Park and S. N. Rogak. A novel fixed-sectional model for the formation

and growth of aerosol agglomerates. J. Aerosol Sci., 35(11):1385–1404,

2004.

[156] S. H. Park, S. N. Rogak, W. K. Bushe, J. Z. Wen, and M. J. Thomson.

An aerosol model to predict size and structure of soot particles. Combust.

Theory Model., 9(3):499–513, 2005.

[157] R. I. A. Patterson, J. Singh, M. Balthasar, M. Kraft, and J. R. Norris. The

linear process deferment algorithm: A new technique for solving popula-

tion balance equations. SIAM J. Sci. Comput., 145(1):303–320, 2006.

164

Page 166: Numerical Modelling of Soot Formation

[158] C. Pilinis. Derivation and numerical solution of the species mass distribu-

tion equations for multicomponent particulate systems. Atmos. Environ.,

24(7):1923–1928, 1990.

[159] S. E. Pratsinis. Simultaneous nucleation, condensation, and coagulation in

aerosol reactors. J. Colloid Interf. Sci., 124(2):416–427, 1988.

[160] J. Pyykonen and J. Jokiniemi. Computational fluid dynamics based sec-

tional aerosol modelling schemes. J. Aerosol Sci., 31(5):531–550, 200.

[161] B. Quay, T. W. Lee, T. Ni, and R. J. Santoro. Spatially resolved measure-

ments of soot volume fraction using laser-induced incandescence. Com-

bust. Flame, 97(3-4):384–392, 1994.

[162] R Development Core Team. R: A language and environment for statistical

computing. R Foundation for Statistical Computing, Vienna, Austria, 2005.

URL http://www.R-project.org. ISBN 3-900051-07-0.

[163] K. Radhakrishnana and A. C. Hindmarsh. Description and use of LSODE,

the Livermore solver for ordinary differential equations. Technical Report

UCRL-ID-113855, LLNL Report, 1994.

[164] D. Ramkrishna. Population Balances. Academic Press, 2000.

[165] Z. Ren and S. B. Pope. The geometry of reaction trajectories and attracting

manifolds in composition space. Combust. Theor. Model., 10(3):361–388,

2006.

[166] K. L. Revzan, N. J. Brown, and M. Frenklach.

http://www.me.berkeley.edu/soot/.

[167] H. Richter, S. Granata, W. H. Green, and J. B. Howard. Detailed

modeling of PAH and soot formation in a laminar premixed ben-

zene/oxygen/argon low-pressure flame. Proc. Combust. Inst., 30:1397–

1405, 2005. doi:10.1016/j.proci.2004.08.088.

165

Page 167: Numerical Modelling of Soot Formation

[168] N. Rieflera, S. di Stasio, and T. Wriedt. Structural analysis of clusters using

configurational and orientational averaging in light scattering analysis. J.

Quant. Spectrosc. Ra., 89:323–342, 2004.

[169] S. Rjasanow and W. Wagner. A stochastic weighted particle method for the

Boltzmann equation. J. Comput. Phys., 124:243–253, 1996.

[170] S. Rjasanow and W. Wagner. Simulation of rare events by the stochastic

weighted particle method for the Boltzmann equation. Math. and Comput.

Modelling, 33:926–907, 2001.

[171] D. E. Rosner and S. Yu. MC simulation of aerosol aggregation and simula-

taneous spheroidization. AIChE J., 47(3):545–561, 2001.

[172] D. H. Rosner and J. J. Pyykonen. Bivariate moment simulation of coag-

ulating and sintering nanoparticles in flames. AIChE J., 48(3):476–491,

2002.

[173] A. I. Roussos, A. H. Alexopoulos, and C. Kiparissides. Part III: Dynamic

evolution of the particle size distribution in batch and continuous particulate

processes: A Galerkin on finite elements approach. Chem. Eng. Sci., 60:

6998–7010, 2005.

[174] A. Sandu and C. Borden. A framework for the numerical treat-

ment of aerosol dynamics. Appl. Numer. Math., 45:475–497, 2003.

doi:10.1016/S0168-9274(02)00251-9.

[175] D. A. Schwer, P. Lu, and W. H. G. Jr. An adaptive chemistry approach to

modeling complex kinetics in reacting flows. Combust. Flame, 133:451–

465, 2003. doi:10.1016/S0010-2180(03)00045-2.

[176] K. Siegmann and K. Sattler. Formation mechanism for polycyclic aromatic

hydrocarbons in methane flames. J. Chem. Phys., 112(2):698–709, 2000.

166

Page 168: Numerical Modelling of Soot Formation

[177] J. Singh, R. Patterson, M. Balthasar, M. Kraft, and W. Wagner. Modelling

soot particle size distribution: Dynamics of pressure regimes. Technical

Report 25, c4e Preprint-Series, Cambridge, 2004.

[178] J. Singh, M. Balthasar, M. Kraft, and W. Wagner. Stochastic modelling of

soot particle size and age distribution in laminar premixed flames. Proc.

Combust. Inst., 30:1457–1465, 2005.

[179] J. Singh, R. I. A. Patterson, M. Kraft, and H. Wang. Numerical simulation

and sensitivity analysis of detailed soot particle size distribution in laminar

premixed ethylene flames. Combust. Flame, 145:117–127, 2006.

[180] J. Slowik, K. Stainken, P. Davidovits, L. Williams, J. Jayne, C. Kolb,

D. Worsnop, Y. Rudich, P. DeCarlo, and J. Jimenez. Particle morphol-

ogy and density characterization by combined mobility and aerodynamic

diameter measurements. part 2: Application to combustion-generated soot

aerosols as a function of fuel equivalence ratio. Aerosol Sci. Tech., 38(12):

1206–1222, 2004. doi:10.1080/027868290903916.

[181] M. Smith and T. Matsoukas. Constant-number Monte Carlo simulation of

population balances. Chem. Eng. Sci., 53(9):1777–1786, 1998.

[182] M. D. Smooke, C. S. McEnally, L. Pfefferle, R. J. Hall, and M. B. Colket.

Computational and experimental study of soot formation in a coflow, lami-

nar diffusion flame. Combust. Flame, 117:117–139, 1999.

[183] M. D. Smooke, R. J. H. M. B. Colket, J. Fielding, M. B. Long, C. S. McE-

nally, and L. D. Pfefferle. Investigation of the transition from lightly soot-

ing towards heavily sooting co-flow ethylene diffusion flames. Combust.

Theor. Model., 8(3):593–606, 2004. doi:10.1088/1364-7830/8/3/009.

[184] M. D. Smooke, B. C. C. M B Long, M. B. Colket, and R. J. Hall. Soot

formation in laminar diffusion flames. Combust. Flame, 143(4):613–628,

2005. doi:10.1016/j.combustflame.2005.08.028.

167

Page 169: Numerical Modelling of Soot Formation

[185] C. M. Sorensen. Light scattering by fractal aggregates: A review. Aerosol

Sci. Tech., 35:648–687, 2001.

[186] C. M. Sorensen and G. D. Feke. Post flame soot. In D. P. Lund and E. A.

Angell, editors, Proceedings of the International Conference on Fire Re-

search and Engineering, pages 280–285, Boston, MA, 1995. Society of

Fire Protection Engineers.

[187] C. M. Sorensen, J. Cai, and N. Lu. Light-scattering measurements of

monomer size, monomers per aggregate, and fractal dimension for soot

aggregates in flames. Applied Optics, 31(30):6547–6557, 1992.

[188] L. A. Spielman and O. Levenspiel. A Monte Carlo treatment for reacting

and coalescing dispersed phase systems. Chem. Eng. Sci., 20:247–254,

1965.

[189] C. B. Stipe, B. S. Higgins, D. Lucas, C. P. Koshland, and R. F. Sawyer.

Inverted co-flow diffusion flame for producing soot. Rev. Sci. Instrum., 76:

023908, 2005. doi:10.1063/1.1851492.

[190] Z. Sun, R. L. Axelbaum, and R. W. Davis. A sectional model for inves-

tigating microcontamination in a rotating disk CVD reactor. Aerosol Sci.

Tech., 38:1161–1170, 2004. doi:10.1080/027868290896799.

[191] Z. Sun, R. L. Axelbaum, and J. I. Huertas. Monte Carlo sim-

ulation of multicomponent aerosols undergoing simultaneous coagu-

lation and condensation. Aerosol Sci. Tech., 38:963–971, 2004.

doi:10.1080/027868290513847.

[192] J. P. R. Symonds, K. S. J. Reavell, J. S. Olfert, B. W. Campbell,

and S. J. Swift. Diesel soot mass calculation in real-time with a dif-

ferential mobility spectrometer. J. Aerosol Sci., 38(1):52–68, 2007.

doi:10.1016/j.jaerosci.2006.10.001.

168

Page 170: Numerical Modelling of Soot Formation

[193] D. A. Terry, R. McGraw, and R. H. Rangel. Method of moments solutions

for a laminar flow aerosol reactor model. Aerosol Sci. Tech., 34:353–362,

2001.

[194] K. Tian, K. A. Thomson, F. Liu, D. R. Snelling, G. J. Smallwood, and

D. Wang. Determination of the morphology of soot aggregates using the

relative optical density method for the analysis of TEM images. Combust.

Flame, 144:782–791, 2006. doi:10.1016/j.combustflame.2005.06.017.

[195] K. Tian, F. Liu, M. Yang, K. A. Thomson, D. R. Snelling, and G. J. Small-

wood. Numerical simulation aided relative optical density analysis of TEM

images for soot morphology determination. Proc. Combust. Inst., 31(1):

861–868, 2007.

[196] S. Tsantilis and S. E. Pratsinis. Soft- and hard-agglomerate aerosols made

at high temperatures. Langmuir, 20(14):5933–5939, 2004.

[197] S. Tsantilis, H. K. Kammler, and S. E. Pratsinis. Population balance mod-

eling of flame synthesis of titania nanoparticles. Chem. Eng. Sci., 57:2139–

2156, 2002.

[198] M. Vanni. Approximate population balance equations for aggregation-

breakage processes. J. Colloid Interf. Sci, 221(2):143–160, 2000.

doi:10.1006/jcis.1999.6571.

[199] W. N. Venables and B. D. Ripley. Modern Applied Statistics with S.

Springer, 4th edition, 2002. ISBN 0-387-95457-0.

[200] A. Vikhansky and M. Kraft. Conservative method for the re-

duction of the number of particles in the Monte Carlo simulation

method for kinetic equations. J. Comput. Phys., 203:371–378, 2005.

doi:10.1016/j.jcp.2004.09.007.

[201] A. Violi. Modeling of soot particle inception in aromatic and

aliphatic premixed flames. Combust. Flame, 139:279–287, 2004.

doi:10.1016/j.combustflame.2004.08.013.

169

Page 171: Numerical Modelling of Soot Formation

[202] A. Violi and A. Venkatnathan. Combustion-generated nanoparticles pro-

duced in a benzene flame: A multiscale approach. J. Chem. Phys., 125:

054302, 2006. doi:10.1063/1.2234481.

[203] A. Violi, A. F. Sarofim, and G. A. Voth. Kinetic Monte-Carlo—molecular

dynamics approach to model soot inception. Combust. Sci. Tech., 176:991–

1005, 2004. doi:10.1080/00102200490428594.

[204] A. Violi, G. A. Voth, and A. F. Sarofim. The relative roles of acetylene and

aromatic precursors during soot particle inception. Proc. Combust. Inst.,

30(1):1343–1351, 2005. doi:10.1016/j.proci.2004.08.226.

[205] L. Wallcave. Gas chromatographic analysis of polycyclic aromatic hydro-

carbons in soot samples. Environ. Sci. Tech., 3(10):948, 1969.

[206] D. Wang, A. Violi, D. H. Kim, and J. A. Mulholland. Formation of Naph-

thalene, Indene, and Benzene from Cyclopentadiene pyrolysis: A DFT

study. J. Phys. Chem. A, 110:4719–4725, 2006. doi:10.1021/jp053628a.

[207] H. Wang and M. Frenklach. A detailed kinetic modeling study of aromatics

formation in laminar premixed acetylene and ethylene flames. Combust.

Flame, 110:173–221, 1997.

[208] H. Wang, B. Shao, B. Wyslouzil, and K. Streletzky. Small-angle neutron

scattering of soot formed in laminar premixed ethylene flames. Proc. Com-

bust. Inst., 29:2749–2757, 2002.

[209] L. Wang, M. F. Modest, D. C. Haworth, and S. R. Turns. Mod-

elling nongrey gas-phase and soot radiation in luminous turbulent non-

premixed jet flames. Combust. Theor. Model., 9(3):479–498, 2005.

doi:10.1080/13647830500194834.

[210] Y. H. Wei, L. X. Zhang, and W. Ke. Evaluation of corrosion protection

of carbon black filled fusion-bonded epoxy coatings on mild steel during

exposure to a quiescent 3% NaCl solution. Corrosion Science, 49:287–302,

2007.

170

Page 172: Numerical Modelling of Soot Formation

[211] C. Wells, N. Morgan, M. Kraft, and W. Wagner. A new method for calculat-

ing the diameters of partially-sintered nanoparticles and its effect on simu-

lated particle properties. Chemical Engineering Sci, 61:158–166, 2006.

[212] C. Wells, N. Morgan, M. Kraft, and W. Wagner. A new method for cal-

culating the diameters of partially-sintered nanoparticles and its effect on

simulated particle properties. Chem. Eng. Sci., 61(1):158–166, 2006.

[213] C. G. Wells and M. Kraft. Direct simulations and mass flow algorithms to

solve a sintering-coagulation equation. Technical Report 10, c4e Preprint-

Series, Cambridge, 2003. Material from this preprint has been published as

[214].

[214] C. G. Wells and M. Kraft. Direct simulation and mass flow stochastic al-

gorithms to solve a sintering-coagulation equation. MCMA, 11:175–197,

2005. doi:10.1163/156939605777585980.

[215] J. Z. Wen, M. J. Thompson, S. H. Park, S. N. Rogak, and M. F. Lightstone.

Study of soot growth in a plug flow reactor using a moving sectional model.

Proc. Combust. Inst., 30:1477–1484, 2005.

[216] J. Z. Wen, M. J. Thomson, and M. F. Lightstone. Numerical study of car-

bonaceous nanoparticle formation behind shock waves. Combust. Theory

Model., 10(2):257–272, 2006.

[217] J. Z. Wen, M. J. Thomson, M. F. Lightstone, S. H. Park, and S. N. Rogak.

An improved moving sectional aerosol model of soot formation in a plug

flow reactor. Combust. Sci. Tech., 178(5):921–951, 2006.

[218] J. Z. Wen, M. J. Thomson, M. F. Lightstone, and S. N. Rogak. Detailed ki-

netic modeling of carbonaceous nanoparticle inception and surface growth

during the pyrolysis of C6H6 behind shock. Energ. Fuel., 20:547–559,

2006. doi:10.1021/ef050081q.

171

Page 173: Numerical Modelling of Soot Formation

[219] R. Whitesides, A. C. Kollias, D. Domin, W. A. Lester, jr, and M. Frenklach.

Graphene layer growth: Collision of migrating five-member rings. Proc.

Combust. Inst., 31(1):539–546, 2007.

[220] M. Wilck and F. Stratmann. A 2-d multicomponent modal aerosol model

and its application to laminar flow reactors. J. Aerosol Sci., 28(6):959–972,

1997.

[221] J. A. Wojcik and A. G. Jones. Dynamics and stability of continuous

MSMPR agglomerative precipitation: numerical analysis of the dual parti-

cle coordinate model. Computers Chem. Engng., 22(4/5):535–545, 1998.

[222] X. Y. Woo, R. B. H. Tan, P. S. Chow, and R. D. Braatz. Simula-

tion of mixing effects in antisolvent crystallization using a coupled CFD-

PDF-PBE approach. Crystal Growth & Design, 6:1291–1303, 2006.

doi:10.1021/cg0503090.

[223] D. L. Wright, R. McGraw, and D. E. Rosner. Bivariate extension of the

quadrature method of moments for modeling simultaneous coagulation and

sintering of particle populations. J. Colloid Interf. Sci., 236:242–251, 2001.

doi:10.1006/jcis.2000.7409.

[224] M. K. Wu and S. K. Friedlander. Enhanced power law agglomerate growth

in the free molecular regime. J. Aerosol Sci., 24(3):273–282, 1993.

[225] M. Wulkow and U. N. A Gerstlauer. Modeling and simulation of crystal-

lization processes using parsival. Chem. Eng. Sci., 56:2575–2588, 2001.

doi:10.1016/S0009-2509(00)00432-2.

[226] E. J. W. Wynn. Simulating aggregation and reaction: New Hounslow DPB

and four-parameter summary. AIChE J., 50(3):578–588, 2004.

[227] Y. Xiong and S. Pratsinis. Formation of irregular particles by coagulation

and sintering: A two dimensional solution of the population balance equa-

tion. J. Aerosol Sci., 22(S1):S199–S202, 1991.

172

Page 174: Numerical Modelling of Soot Formation

[228] Y. Xiong and S. E. Pratsinis. Formation of agglomerate particles by coagu-

lation and sintering—part I.: A two dimensional solution of the population

balance equation. J. Aerosol Sci., 24(3):283–300, 1993.

[229] Y. Xiong, M. K. Akhtar, and S. E. Pratsinis. Formation of agglomerate

particles by coagulation and sintering—part I.: A. J. Aerosol Sci., 24(3):

301–313, 1993.

[230] F. Xu, P. B. Sunderland, and G. M. Faeth. Soot formation in laminar pre-

mixed ethylene/air flames at atmospheric pressure. Combust. Flame, 108:

471–493, 1997.

[231] F. Xu, P. B. Sunderland, and G. M. Faeth. Soot formation in laminar pre-

mixed ethylene/air flames at atmospheric pressure. Combust. Flame, 108:

471–493, 1997.

[232] F. Xu, K. C. Lin, and G. M. Faeth. Soot formation in laminar premixed

methane/oxygen flames at atmospheric pressure. Combust. Flame, 115:

195–209, 1998.

[233] S. Yan, Y.-J. Jiang, N. D. Marsh, E. G. Eddings, A. F. Sarofim, and R. J.

Pugmire. Study of the evolution of soot from various fuels. Energ. Fuel.,

19:1804–1811, 2005.

[234] B. Yang and U. O. Koylu. Detailed soot field in a turbulent non-premixed

ethylene/air flame from laser scattering and extinction experiments. Com-

bust. Flame, 141(1-2):55–65, 2005.

[235] C. Yoon and R. McGraw. Representation of generally mixed multivariate

aerosols by the quadrature method of moments: I. Statistical foundation. J.

Aerosol Sci., 35:561–576, 2004. doi:10.1016/j.jaerosci.2003.11.003.

[236] C. Yoon and R. McGraw. Representation of generally mixed multivariate

aerosols by the quadrature method of moments: II. Aerosol dynamics. J.

Aerosol Sci., 35:577–598, 2004. doi:10.1016/j.jaerosci.2003.11.012.

173

Page 175: Numerical Modelling of Soot Formation

[237] S. Yu, Y. Yoon, M. Muller-Roosen, and I. M. Kennedy. A two-dimensional

discrete-sectional model for metal aerosol dynamics in a flame. Aerosol

Sci. Tech., 28:185–196, 1998.

[238] B. Zhao, Z. Yang, M. V. Johnston, H. Wang, A. S. Wexler, M. Balthasar,

and M. Kraft. Measurement and numerical simulation of soot particle size

distribution functions in a laminar premixed ethylene-oxygen-argon flame.

Combust. Flame, 133:173–188, 2003.

[239] B. Zhao, Z. Yang, J. Wang, M. V. Johnston, and H. Wang. Analy-

sis of soot nanoparticles in a laminar premixed ethylene flame by scan-

ning mobility particle sizer. Aerosol Sci. Tech., 37(8):611–620, 2003.

doi:10.1080/02786820300908.

[240] B. Zhao, Z. Yang, Z. Li, M. V. Johnston, and H. Wang. Particle size dis-

tribution function of incipient soot in laminar premixed ethylene flames:

effect of flame temperature. Proc. Combust. Inst., 30:1441–1448, 2005.

[241] B. Zhao, K. Uchikawa, and H. Wang. A comparative study of nanoparticles

in premixed flames by scanning mobility particle sizer, small angle neutron

scattering, and transmission electron microscopy. Proc. Combust. Inst., 31

(1):851–860, 2007.

[242] H. Zhao, C. Zheng, and M. Xu. Multi-Monte Carlo approach for particle

coagulation: description and validation. Appl. Math. Comput., 167:1383–

1399, 2005. doi:10.1016/j.amc.2004.08.014. This paper is by the same

authors as [76]; the two published manuscripts just reverse the orders of

the parts of the authors’ names.

[243] H. Z. A. M. T. M. C. Zheng. Analysis of four Monte Carlo methods for the

solution of population balances in dispersed systems. Powder Tech., 173:

38–50, 2007. doi:10.1016/j.powtec.2006.12.010.

[244] A. Zucca, D. L. Marchisio, A. A. Barresi, and R. O. Fox. Implemen-

tation of the population balance equation in CFD codes for modelling

174

Page 176: Numerical Modelling of Soot Formation

soot formation in turbulent flames. Chem. Eng. Sci., 61:87–95, 2006.

doi:10.1016/j.ces.2004.11.061.

[245] M. Zurita-Gotor and D. E. Rosner. Effective diameters for collisions of

fractal-like aggregates: Recommendations for improved aerosol coagula-

tion frequency predictions. J. Colloid Interf. Sci., 255:10–26, 2002.

[246] M. Zurita-Gotor and D. E. Rosner. Aggregate size distribution evolution for

brownian coagulation–sensitivity to an improved rate constant. J. Colloid

Interf. Sci., 274:502–514, 2004.

175