You are on page 1of 18

Combustion and Flame 259 (2024) 113128

Contents lists available at ScienceDirect

Combustion and Flame


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

LES investigation of soot formation in a turbulent non-premixed jet


flame with sectional method and FGM chemistry
Abhijit Kalbhor a,∗, Daniel Mira b, Ambrus Both b, Jeroen van Oijen a,c
a
Department of Mechanical Engineering, Eindhoven University of Technology, Eindhoven, 5600 MB, the Netherlands
b
Barcelona Supercomputing Center (BSC), Plaça Eusebi Güell, Barcelona, 1-3 08034, Spain
c
Eindhoven Institute for Renewable Energy Systems (EIRES), Eindhoven, 5600 MB, the Netherlands

a r t i c l e i n f o a b s t r a c t

Article history: This study proposes a detailed soot modeling framework for large-eddy simulation (LES) to accurately
Received 18 May 2023 predict soot formation and particle size distributions (PSD) in turbulent reacting flows. The framework
Revised 7 October 2023
incorporates Flamelet Generated Manifold (FGM) chemistry and a soot model based on the discrete sec-
Accepted 9 October 2023
tional method (DSM) to predict both qualitative and quantitative sooting behavior while keeping the
computational cost affordable. Two elementary modeling strategies are considered in the LES formalism
Keywords: for describing soot formation rates. These strategies rely on an a-priori tabulation of soot formation rates
Discrete sectional method and their run-time computation. The LES formalism is applied to the simulations of a well-characterized,
Flamelet generated manifold non-premixed, turbulent jet flame. A comparative analysis of strategies employed for filtered soot source
Large eddy simulation
term treatment is conducted to investigate their impact on the prediction of soot quantities and the evo-
Turbulent non-premixed flames
lution of soot PSDs. The LES results for the gas phase and soot phase are compared against the available
experimental data. A good prediction of soot evolution is achieved with the two methodologies. The tab-
ulation of soot formation rates leads to a significant reduction in computational cost compared to the
model based on their explicit runtime computation. The LES results reveal that the modeling of filtered
soot source terms has a significant impact on the quantitative prediction of soot formation. The possible
reasons for the observed differences in the soot prediction are discussed. The run-time computation-
based model provides a more consistent treatment of the non-linear interactions between the gas and
soot phases in soot source terms compared to the tabulated soot chemistry approach. On the other hand,
the tabulated soot chemistry model is an interesting and efficient modeling approach for predicting soot
formation in turbulent conditions. Overall, both approaches have their strengths and limitations, and the
choice of approach may depend on the specific needs of the application.
© 2023 The Author(s). Published by Elsevier Inc. on behalf of The Combustion Institute.
This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/)

1. Introduction CPU intensive [4]. Several approaches facilitate the description of


the PBE, including stochastic methods [5], moment-based meth-
Due to the negative effects of soot on the environment and ods [6,7], kinetic models [8], and discrete sectional methods [9–11].
human health, the legislative regulations on soot emissions However, due to limitations in terms of computational overhead,
from combustion devices have become stricter. The new regula- primarily moment-based [7,12,13] and, to a lesser extent, sectional
tions [1,2] will require monitoring the number of particles, particu- methods [11,14] are more commonly used in large-scale flame sim-
larly smaller ones, in addition to the total soot volume. Hence, pre- ulations. Moment-based approaches allow an adequate description
dicting the evolution of soot particle size distribution is essential to of the spatial and temporal evolution of soot particles, however,
develop appropriate mitigation strategies for their emissions. Pre- they provide only limited access to particle size distribution and
dicting particle size distribution with soot models requires solving often rely on approximations and closure models to estimate
the population balance equation (PBE), which describes the dy- the unknown soot particle number density function (NDF). In
namics of a polydispersed particulate system coupled with reactive contrast, discrete sectional method-based approaches offer direct
flow equations [3]. As a result, sooting flame computations become information on the local size distribution of soot particles but are
less commonly used as they are more computationally demanding.
The state of the art in soot modeling research suggests that de-

Corresponding author.
tailed soot models coupled with detailed kinetics, transport, and
E-mail address: a.j.kalbhor@tue.nl (A. Kalbhor).

https://doi.org/10.1016/j.combustflame.2023.113128
0010-2180/© 2023 The Author(s). Published by Elsevier Inc. on behalf of The Combustion Institute. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/)
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

radiation models provide a fairly good prediction of soot forma- pling of the soot model with the turbulent flow. These include:
tion in laminar flames [15] after careful calibration of the models. a) the correlation between reactive scalars and the soot scalar
The progression toward soot prediction in turbulent flames, how- (number density) concerning growth and oxidation terms, and b)
ever, remains a great challenge owing to its complex multi-physical the correlation between number densities in relation to coagula-
nature, characterized by an intricate coupling between flow pa- tion terms. Different strategies have been proposed to model soot-
rameters, gas-phase chemistry, and soot properties. Numerical pre- turbulence interactions in mono-dispersed and moment-based
diction of soot in turbulent conditions necessitates adequate char- models [7,30] and more recently in sectional models [4,11]. Mueller
acterization of the flame behavior as well as different physico- and Pitsch [34] introduced a subfilter PDF model with double-delta
chemical phenomena involved in soot formation. PDF, considering both sooting and non-sooting modes for subgrid-
Turbulent flows are characterized by a wide range of flow scales scale turbulence-soot interaction. This approach was employed in
from energy-containing large scales to the smallest Kolmogorov several other studies [11,35] and subsequently reformulated by
scales [16]. In comparison, the time scales of soot formation are Berger et al. [36], by defining the sooting mode sub-structure as
typically even larger than the integral flow time scales. This dis- a log-normal distribution. Furthermore, Yang et al. [30] proposed
parity makes the prediction of soot formation in turbulent flames a presumed subfilter PDF that explicitly considers soot distribution
a challenging task. Hence, although significant efforts have been in mixture fraction space. Recently, Colmán et al. [37] extended the
made to describe turbulent-chemistry interaction, understanding presumed subfilter PDF model to account for finite-rate soot oxi-
the effects of turbulence on soot formation and their modeling are dation by analyzing the local relative motion of diffusionless soot
still open subjects to be addressed. particles relative to mixture fraction iso-contours. Although closure
The prediction of soot formation in turbulent flames using models for soot-turbulence interactions are important in the de-
the aforementioned soot modeling approaches is preferable to be tailed description of the soot formation process, their impact on
conducted in the context of Direct Numerical Simulations (DNS), the overall prediction of global soot quantities appears less promi-
where a full description of all the temporal and spatial scales as- nent compared to the sensitivity of soot production to model pa-
sociated with turbulence and chemistry are resolved. However, the rameters involved in soot subprocesses [15]. As a result, in some
high computational cost required to resolve the coupled phenom- cases [13,20], interactions between turbulence and soot are ne-
ena associated with the scales of turbulence, chemistry, and parti- glected.
cle dynamics limits the DNS investigation to selected cases [17–19]. Earlier studies on the LES of turbulent sooting flames have been
On the other end, Reynolds Averaged Navier-Stokes (RANS)-based primarily focused on moment-based soot models. El-Asrag and
methods, are more affordable and mainly used at industrially- Menon [38] employed the Linear Eddy Model (LEM) in combina-
relevant conditions [20]. However, RANS approaches hardly provide tion with the moment method for soot prediction with LES. Later,
information on the transient phenomena of the flow, flame dynam- Mueller and Pitsch [7,34] combined the moment-based soot model
ics, and soot formation. Therefore, methods like Large Eddy Simu- with an extended Flamelet/Progress Variable (FPV) [39] model by
lation (LES), which facilitates information on transient features of including lumped-PAH inception kinetics. In this LES approach, the
the flow are more relevant despite their higher computational cost. presumed-PDF method was used to account for subgrid-scale inter-
In fact, LES has been applied very successfully for a wide range actions between turbulence and soot. Xuan and Blanquart [12] pro-
of problems including jet, swirl, or spray flames [4,11,21–26] and posed an alternative approach to the lumped-PAH inception ki-
it generally shows a good trade-off between accuracy and compu- netics, where the filtered transport equations for aromatic species
tational cost. Therefore, the development of reliable soot models were solved with closure based on a PAH relaxation model. This
for LES is essential for the study of soot formation in both funda- aromatic chemistry-turbulence interaction was found to be impor-
mental and industrial configurations. In recent years, LES has been tant in accurately reproducing soot yield in turbulent flames. To
adopted to investigate soot formation in turbulent flames, depart- obtain the information on particle size distribution in LES, the FPV
ing from semi-empirical methods [27–29] and evolving towards presumed-PDF model of Mueller and Pitsch [34] was extended to
more detailed soot modeling approaches [4,7,11,13,30,31]. a sectional method by Rodrigues et al. [11]. The qualitative trends
LES of turbulent sooting flames relies on three aspects: an ac- in soot formation were reproduced in their work while providing
curate description of gas-phase chemistry, modeling of the dynam- detailed information on the evolution of local soot PSD. The decou-
ics of soot particle population, and integration of models for cap- pling of soot and gas-phase species in the LES presumed-PDF ap-
turing interactions between soot-turbulence-chemistry. For reliable proach leads to narrow profiles of the mean soot volume fraction.
prediction of soot formation in turbulent combustion, a detailed Hence, the transported joint PDF (JPDF) method, which allows for
understanding of gas-phase kinetics is necessary to properly cap- a more accurate representation of the turbulence-chemistry inter-
ture the flame structure and formation of gas-phase species partic- actions by directly coupling the full scalar space, is applied to the
ipating in the soot processes (such as polycyclic aromatic hydrocar- sectional model by Tian et al. [31]. Encouraging agreement was ob-
bons (PAHs) or unsaturated hydrocarbons such as acetylene). How- tained with experimental measurements for soot quantities in tur-
ever, using detailed kinetic schemes involving hundreds of species bulent non-premixed flames.
is impractical in LES employing finite-rate chemistry-based meth- Besides flamelet-based methods, several other combustion
ods. As a result, tabulated chemistry methods [7,11], reduced ki- models have been investigated in the LES framework to model soot
netic schemes [32], and globally optimized chemistry [33] are of- formation in turbulent flames. The Conditional Moment Closure
ten used instead. Concerning the soot-phase description, although (CMC) approach was applied to a sectional soot model by Gkan-
quantitative modeling of soot formation in turbulent combustion tonas et al. [24] to study soot formation and predict their size dis-
is still beyond the frontier of current modeling advancements [15], tribution in a lab-scale swirl Rich-Quench-Lean (RQL) combustor.
the state-of-the-art soot models can provide partial solutions to In LES of lifted non-premixed turbulent flames with the sectional
practical problems of interest when used judiciously along with method, Grader et al. [14] used a finite-rate chemistry model,
appropriate predictions of the turbulent flow structures. that requires no assumptions concerning the combustion regime. A
Another challenge in LES of turbulent sooting flames is mod- good prediction of soot evolution was achieved in their study en-
eling the subgrid-scale interactions between soot and turbulence, abling detailed investigations of soot formation and oxidation pro-
as soot chemistry involves non-linear interactions between the cesses. Sewerin and Rigopoulos [40] adapted the stochastic fields
gas and solid phases. In the context of LES, subgrid turbulence- method to the LES of sooting flames, which recently extended to
soot interactions refer to the correlations that arise from the cou- the sectional method by Sun and Rigopoulos [4] to predict the soot

2
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

particle size distribution in turbulent non-premixed flames. Con-   


∂ (ρ̄
u) 2
cerning sectional method-based approaches, simplifying the gas- + ∇ · (ρ̄u ) = −∇ p + ∇ · ρ̄ (
u ν + νt ) 2S − (∇ · 
u )I
phase kinetics by the use of various combustion models in the LES ∂t 3
can reduce CPU time, but the calculation of the soot source term (2)
remains computationally expensive. Recent research has explored    
∂ (ρ̄
h)  ν
different approaches to improve computational efficiency, includ- + ∇ · (ρ̄ 
uh ) = ∇ · ρ̄ D +
t
∇
h (3)
ing a three-equation model of Franzelli et al. [27] based on mono- ∂t P rt
disperse closure of source terms. Machine learning models and vir-
where ρ̄ ,  u,   and P rt represent the density, ve-
ν , νt , S, I , D
h, p, 
tual chemistry have also been proposed [33] for predicting soot
locity vector, total enthalpy, pressure, kinematic viscosity, subgrid-
formation, but they do not provide information on size distribu-
scale (SGS) turbulent viscosity, resolved strain tensor (S = 12 [∇ u+
tions. Overall, developing detailed, and at the same time compu-
(∇ 
u )T ]), identity tensor, thermal diffusion coefficient, and turbu-
tationally efficient methods for predicting soot formation is crucial
lent Prandtl number, respectively. In the conservation equations,
for their application in industrially relevant conditions.
the overline operator denotes the LES filtering, and the tilde rep-
Given the aforementioned challenges in accurately predicting
resents the Favre-filtering operations. The subgrid stresses in the
soot formation in turbulent conditions, we propose an LES frame-
momentum equation are closed using the Boussinesq approxima-
work that combines a discrete sectional soot model with FGM tab-
tion [16]. The Vreman [45] model with modeling constant ck = 0.1
ulated chemistry [41,42] to efficiently predict soot formation and
is adopted for the SGS turbulent viscosity. The viscous heating ef-
provide information on soot particle size distribution in turbulent
fects are considered to be relatively small and neglected in the en-
flames. The complete tabulation of soot chemistry, similar to the
thalpy equation, while the unresolved heat flux is modeled using a
treatment of gas-phase combustion chemistry in FGM, presents an
gradient diffusion approach. Guided by the previous studies of the
interesting and efficient framework for its application to the simu-
authors using the same numerical methods [22,46], the turbulent
lations of turbulent sooting flames. While tabulated soot chemistry
Prandtl number P rt set constant with a value of 0.7.
has been investigated with semi-empirical soot models [43], de-
tailed soot models [44] in laminar flames, accurately modeling soot
2.2. Combustion modeling
formation rates in turbulent conditions remains a challenge due
to the nonlinear interaction between gas-phase thermochemistry,
Considering the complexity of the combustion process and
soot particle dynamics, and turbulence. In this study, we present
the need for a regime definition to find a suitable combustion
a detailed formalism of the tabulated soot chemistry model for
model, the combustion process is assumed to occur in the flamelet
LES based on the discrete sectional method. This new methodol-
regime [47]. Therefore, turbulent combustion is modeled with the
ogy is compared to a more detailed approach based on the local
FGM [41,42,48] approach. In FGM, the thermochemical states of the
evaluation of the formation rates of the soot quantities. While the
flame can be described using a pre-computed flamelet database
first strategy relies on an a-priori tabulation of soot formation rates
parameterized by a reduced set of control variables. In the con-
for computationally efficient application of the sectional model,
text of non-premixed combustion, mixture fraction Z and reaction
the second involves run-time computation of soot formation rates
progress variable Y are selected as the control variables to repre-
based on primitive tabulated quantities. The comparative assess-
sent mixing and combustion progress, respectively.
ment of these two strategies is conducted for soot formation in a
In the current work, the FGM database is constructed using a
laboratory-scale turbulent non-premixed sooting jet flame. The fo-
combination of non-premixed counterflow steady flamelets, which
cus is given to the assessment of the tabulated approach to predict
cover a range of different strain rates from lower values up to
the complex interactions between the gas phase and the soot par-
the extinction strain rate, along with unsteady quenching flamelets
ticles in turbulent conditions.
at the extinction strain rate. These flamelet solutions are subse-
The paper is organized as follows. In Section 2, the LES for-
quently mapped onto a cartesian grid with high resolution in both
malism is introduced, including a summary of the combustion
mixture fraction Z and progress variable Y spaces. After these
and soot models. In Section 3, the details of the target flame
manifolds are constructed for all the tabulated quantities, a non-
and numerical setup are presented. Subsequently, in Section 4, the
uniform lookup table is created by interpolating the solutions from
proposed LES formalism is applied to a turbulent non-premixed
this database to the number of entries specified for the flamelet
jet flame. Numerical results are first compared with experimental
database.
data, followed by the discussion and assessment of modeling ap-
The mixture fraction is determined from element mass fractions
proaches. Conclusions are presented at the end in Section 5.
following the definition of Bilger [49], while the progress variable
is chosen as a linear combination of the mass fractions of selected
2. LES formalism with FGM-DSM methodology species such that it preserves the unique mapping of flamelet in
composition space. The progress variable is expressed as:
This section outlines the LES-FGM-DSM formalism for model- 
ing the formation and evolution of soot in turbulent non-premixed Y= α jY j , (4)
combustion with FGM chemistry. In LES, the filtered governing
where α j and Y j are the weight factor and mass fraction for the
equations for reacting flows are solved along with the equation of
species j respectively. The specific definition applied in this work
state. Details of the modeling approach and the numerical methods
is detailed in Section 3.2. To describe the combustion process dur-
are given in the next subsections.
ing LES calculations, transport equations are solved for the filtered
mixture fraction  , which take the
Z and filtered progress variable Y
2.1. Reactive flow equations form:
   
∂ (ρ̄ 
Z)  + νt ∇ 
Under the low-Mach number approximation, the Favre-filtered u
+ ∇ · (ρ̄Z ) = ∇ · ρ̄ D Z , (5)
conservation equations for mass, momentum, and total enthalpy ∂t Sct
are obtained as follows [16,22]:    
∂ ρ̄ ∂ (ρ̄ Y)  + νt ∇ Y
) = ∇ · ρ̄ D  + ω˙ Y ,
+ ∇ · (ρ̄
u) = 0 + ∇ · (ρ̄
uY (6)
∂t
(1) ∂t Sct

3
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

 denotes the diffusion coefficient. The unresolved tur- ⎡ ⎤


where D   −1    
bulent flux terms appearing in control variable transport equa- ⎣ 1
⎦ ψ (Z, C )  
ψ= P (Z )P (C )dZdC , (10)
tions after the LES filtering are closed using a gradient diffusion ρ ρ (Z, C )
approach with a turbulent Schmidt number Sct set equal to 0.7.
The term ω˙ Y in Eq. (6) refers to the filtered progress variable where ψ refers to any quantity defined in the flamelet solu-
source term, which is also obtained from the flamelet database. tions. Several LES studies [52–56] have demonstrated β -function
In Eqs. (5) and (6), the diffusion coefficient D is evaluated un-
marginal PDFs as a suitable choice for the convolution of flamelet
der the assumption of a unity Lewis number and included as a solutions. Hence, a β -function PDF is used to characterize the
lookup quantity in the manifold. It is worth highlighting that pre- distribution of both mixture fraction and progress variable. For the
vious studies [50,51] have indicated that non-unity Lewis num- application in LES, the thermochemical parameters of the mixture
bers can indeed have a significant impact on the transport of large are mapped on four non-dimensional coordinates in the FGM
species like pyrene (A4). Therefore, neglecting this effect by assum- database as:
ing unity Lewis numbers may result in quantitative underestima-
=
ψ (
Z , SZ , C, SC ), (11)
tions of PAH, and thus soot formation in LES. In this context, more
consistent models that account for the transport of control vari- where Sφ , called the segregation factor, represents the normalized
ables with non-unity Lewis numbers, as demonstrated in studies variance of control variable φ , required in the β -PDF model. For
such as [21], can be valuable for capturing preferential diffusion Sφ = 0 the variance φ 2 is zero, which leads to ψ (φ
) = ψ (φ ). The
effects in sooting flame simulations. However, we recognize this segregation factors for variables Z and C, are given as:
aspect as an area for potential improvement and a direction for
future investigations. Z 2 C 2
SZ = and SC = . (12)
The normalized progress variable C is introduced for simplifying  
Z (1 − Z ) 
C ( 1 − C)
the construction and look-up of quantities from the FGM database.
It is defined as: The subgrid variance of scaled progress variable C 2 is related
to its unscaled variance Y 2 through [57]:
Y − Ymin (Z )  
C= , (7)
Ymax (Z ) − Ymin (Z ) )2 − 2C (Y 
)2 − (Y
Y 2 + (Y 2
min min Ymax ) − (Ymin )
For the normalization of the progress variable Y during the con- C 2 = − C2 .
struction of the flamelet database, Ymax (Z ) and Ymin (Z ) are used (Ymax
− Ymin )2
at each mixture fraction Z level of the flamelets spanning different (13)
strain rates. In Eq. (7), the maximum value denoted as Ymax (Z ) at Note that when the flamelet manifold is extended to account
each mixture fraction Z level corresponds to the progress variable for subgrid-scale effects by the PDF integration, the scaled progress
in the lowest strain rate flamelet, as the applied progress variable variable C also depends on the subgrid variance of mixture frac-
definition increases with the advancement of reactions and the sta- tion:
ble flamelet solution at the lowest strain rate is the one closest to
− Y
Y   2
min (Z , Z )
chemical equilibrium. Meanwhile Ymin (Z ) represents the value of
C = . (14)
the progress variable in the unburnt (mixing) state where the in-  
Ymax (Z , Z ) − Y
 2   2
min (Z , Z )
termediate and product species used in the progress variable defi-
In this work, transport equations are solved for the closure of
nition are all absent, hence Ymin (Z ) = 0 for all Z. It is important to
subgrid-scale variances of mixture fraction and progress variable.
note that in the CFD calculation, Y is an independent control vari-
The transport equations for the sub-grid variances Z 2 and Y 2 are
able that is being transported and does not depend on the trans-
given by:
ported mixture fraction.
   
∂ (ρ̄ Z 2 )  + νt ∇ Z 2
2.3. Turbulence-chemistry interactions + ∇ · (ρ̄
uZ 2 ) − ∇ · ρ̄ D
∂t Sct
νt  2
In turbulent conditions, the scalar variables are exposed to = 2ρ̄ |∇ Z | − ρ̄ χZ , (15)
spatio-temporal fluctuations that can not be resolved with the LES Sct
filter length, given in this case by the mesh size. Therefore, the    
flamelet database incorporating laminar flamelet solutions must be ∂ (ρ̄ Y 2 )  + νt ∇ Y 2
+ ∇ · (ρ̄
uY 2 ) − ∇ · ρ̄ D
extended to account for the effect of turbulence on subgrid levels. ∂t Sct
Different methods for the modeling of subgrid-scale turbulence- νt  2
chemistry interactions exist in the literature. Here, the stochas- = 2ρ̄ |∇ Y | − ρ̄ χY + 2(Y ω˙ Y − Yω˙ Y ), (16)
Sct
tic approach using presumed-shape Probability Density Functions
(PDF) is used to model these interactions. Only fluctuations in Z where χZ and χY respectively denote the subgrid-scale scalar dis-
and C are considered in the LES to describe the statistical impact sipation rates for mixture fraction and progress variable, modeled
of turbulence on the flame. The filtered thermochemical variable as:
ψ, convoluted over the joint PDF of Z and C, can be expressed as: ε ε
  χ
Z = Z 2 and χY = Y 2 , (17)
k k
=
ψ ψ (Z, C )P(Z, C )dZdC . (8)
where k is the unresolved turbulent kinetic energy and ε is its dis-
Commonly, statistical independence between Z and C is assumed. sipation rate. To close χ
Z and χ
Y , the ratio of ε and k is defined
Thus, the joint PDF P (Z, C ) can be represented by individual following the expression of Both et al. [58]:
presumed PDF distributions of Z and C. The Favre-filtered variable  1/3
 and the filtered variable ψ can then be obtained by integrating ε Cε2 μt |S|2
ψ = , (18)
the marginal PDFs of Z and C: k ρ̄ 2

 
= where Cε = 1.8 is a model parameter, and is the LES filter size
ψ ψ (Z, C )P(Z )P(C )dZdC , (9)
(taken as the cube root of the volume of the grid element).

4
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

2.4. Soot modeling gas-phase consumption of precursor species due to soot for-
mation is accounted for during the flamelet calculations step,
2.4.1. Sectional soot model and no additional transport equation is solved for their descrip-
The chemistry and dynamics of soot particles are described us- tion in the LES. Nevertheless, there is also the possibility of ex-
ing the discrete sectional method [59]. In the present model, soot tending the current model to explicitly solve transport equa-
particle volume ranges are divided into a finite number of sections. tions for the precursor species, a technique adopted in several
In each section, i, the governing equation for the soot mass frac- other studies [11,19], allowing for the evolution of slower pre-
tion Ys,i is solved by considering flow convection, diffusion, ther- cursor species in FGM chemistry. For brevity, this approach, in
mophoresis, and chemical rates. The sectional soot transport equa- which the soot source terms are computed on-the-fly is referred
tion can be formulated as: to as FGM-C here.
∂ (ρYs,i ) Additionally, an alternative variant of the FGM-C approach with
+ ∇ · (ρ [u + vT ]Ys,i ) = ∇ · (ρ Ds,i ∇ Ys,i ) + ω˙ s,i
∂t a slightly different formulation for filtered soot source term is
∀ i ∈ [1, nsec ] (19) evaluated. This variant involves tabulating the gas-phase rate
of the filtered chemical soot source term instead of calculat-
where ρ , ρs , u, vT , Ds,i , ω˙ s,i denote gas density, soot density (as- ing it from local gas-phase species concentrations. This partic-
sumed to be equivalent to the density of solid carbon), velocity, ular variant, referred to as FGM-CR, is aimed at understand-
thermophoretic velocity (calculated with Frienlander’s [60] expres- ing the implications of using filtered concentrations of species
sion), soot diffusion coefficient (assumed to be constant for all par- to compute rates. Given that this method is a variant of the
ticle sizes for numerical stability), and sectional source term, re- FGM-C model, it will be addressed separately to enhance clar-
spectively. The soot source terms account for the chemical and ity. A comprehensive analysis of this approach is provided in
physical processes involved in soot formation, including nucleation, Section 4.2.5.
condensation of Polycyclic Aromatic Hydrocarbons (PAHs), surface • The second model concerns the closure of the filtered soot
growth, oxidation, and coagulation. Nucleation is assumed to occur source term through the presumed PDF approach. In a general
through the dimerization of two pyrene (A4) molecules, and PAH form, the filtered soot source term convoluted with presumed
condensation onto soot particles is modeled using the method de- PDF is written as:
scribed by Roy [61]. Surface reactions, including growth and oxida-  
1
tion, are modeled using the standard hydrogen-abstraction-C2 H2 - ω˙ s,i = ρ̄ ω˙ s,i (φg , φs )P(φg , φs )dφg dφs . (23)
addition (HACA) mechanism by Appel et al. [62]. Coagulation of ρ
soot particles is described using the model proposed by Kumar Using Bayes’ theorem, the joint PDF can be split into two
and Ramkrishna [63], and the morphological properties (e.g. fractal marginal PDFs for the thermochemical variable and for the
dimension) of soot are not considered for simplicity. As a conse- (φg , φs ) = P
soot such that P (φg )P (φs |φg ). In this model, the soot
quence, the present soot model does have limitations related to its source terms are parameterized through control variables and
omission of aggregation and the potential influence this may have tabulated in the manifold. The marginal PDF P (φg ) is assumed
on both the growth and size distribution of soot. Further details to have β -PDF function for Z and C, while the conditional PDF
about the soot model and its validation can be found in previous of soot variable P (φs |φg ) is treated as a δ -function. This approx-
work [59,64,65]. imation facilitates the partial inclusion of turbulence-soot inter-
actions by accounting for the effect of subgrid-scale fluctuations
2.4.2. Filtered sectional soot equations in the mixture-fraction and progress variables on soot source
For modeling soot formation and evolution in LES, the filtered terms. Besides, the assumption of δ -PDF function (neglecting
equations for sectional soot mass fractions are obtained as: subgrid-scale fluctuations) for P (φg ) is also examined to under-
   
∂ (ρ̄Ys,i )  νt 
stand the impact of PDF integration on the performance of the
+ ∇ · (ρ [u + vT ]Ys,i ) = ∇ · ρ̄ D̄s,i +
  ∇ Ys,i + ω˙ s,i , tabulated soot chemistry, which will be explained later.
∂t Sct
The approach employed in other studies [7,66] for the transport
∀ i ∈ [1, nsec ]. (20) of slowly evolving species such as PAH and NO is applied for
The thermophoretic velocity 
vT is modeled following [11], as: soot. Accordingly, the soot source term for the soot section is
split into production (ω˙ s,i
+
or ω˙ s,i ) and consumption parts (ω˙ s,i

prod
∇
T or ω˙ s,i
vT = −0.554ν̄


. (21)
cons ). The consumption part is linearized by soot mass frac-

T tion to avoid the un-physical consumption of soot when Y  = 0.


s,i
Since the soot source term ω˙ s,i includes contributions from the var- The filtered soot source term is modeled as:
ious sub-processes of soot formation, it depends on both gas-phase  tab
thermo-chemical variables (φg ) and soot variables (φs ). Therefore, +
 + tab
− ω −
˙ s,i
ω˙ s,i = ω˙ + ω˙ ≈ ω˙ s,i s,i
+Y , (24)
the filtered source term ω˙ s,i poses closure problems requiring s,i s,i
Ys,i
models to account for subgrid-scale soot-chemistry-turbulence in-
teractions. Here we consider two different strategies to treat the where the production and linearized consumption terms are
filtered soot source terms. parameterized through gas-phase thermochemical variables and
• In the first approach, the chemical source term for the soot sec- tabulated (denoted by the superscript tab) in the manifold. This
tion is evaluated during runtime using tabulated filtered ther- approach does not consider the separation of gas and soot parts
mochemical parameters and local concentrations of gas-phase for the soot production term. However, with the linearization of
species relevant to the soot model. The filtered soot source consumption terms, the partial separation of gas and soot time
term is approximated by neglecting the subgrid-scale interac- scales can be achieved.
tion between soot and turbulence, such that: In general, soot concentrations are negligible in lean mixture
 ,φ ). fraction (
Z < Zst ) regions, as soot particles are oxidized through
ω˙ s,i = ω˙ s,i (φ g s (22) reactions with OH, whose concentration is high near the sto-
In the current model, the transient behavior of soot precur- ichiometric mixture fraction. In this model, since soot source
sors (e.g., A4) is partially considered, primarily during the 1- terms are obtained from the flamelet solutions, the filtered con-

D flamelet calculation phase (using unsteady flamelets). The sumption term ω˙ s,i ≈ 0 in lean regions. This can lead to the

5
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

Table 1
Modeling approaches for filtered soot source terms.

Approach Model Description

Computation FGM-C Runtime evaluation using tabulated primitive variables (gas-phase species)
Computation FGM-CR Runtime evaluation using tabulated gas-phase rates for soot subprocesses
Tabulation FGM-T Direct a-priori tabulation

spurious existence of soot (during transport) in lean mixture


factions if soot oxidation is not properly accounted for dur-
ing the soot chemistry tabulation. To avoid this issue, the soot
oxidation contribution, in the linearized consumption term in
Eq. (24) is further approximated as:
⎡   ⎤tab
− ω˙ s,i

− ω˙ s,i
−,ox
dω˙ s,i
−,ox
ω˙ ≈ Ys,i ⎣
s,i + ⎦ , (25)
Ys,i dYs,i

−,ox
where ω˙ s,i denotes the sectional soot consumption rate by ox-
idation subprocesses. For computational efficiency, the coagu-
lation process of soot particles is not explicitly solved at run-
time, but the inter-sectional mass transfer due to the coagula-
tion process (or other subprocesses) is included in the flamelet
computation. This approach, in which the soot source term is
tabulated, is referred to as FGM-T from hereon.

The key differences between models are summarized in Table 1.


In brief, the FGM-T approach is based on the direct tabulation
of soot source terms, and the application of these source terms
in soot transport equations. Meanwhile, the FGM-C approach re-
lies on the tabulation of primitive variables (e.g. concentrations of
gas-phase species involved in surface growth and oxidation) and
combines them with the local CFD solution of temperature, den-
Fig. 1. An overview of the LES-FGM-DSM implementation. Symbols are explained
sity, and fields representing soot to evaluate the soot source terms. in the main text.
The FGM-CR model, a variant of FGM-C, incorporates the tabula-
tion of gas-phase rates related to various soot processes. The tab- 3.1. Experimental configuration
ulated gas-phase rates are subsequently integrated with the local
solutions of soot fields and gas-phase properties, akin to the FGM- The selected non-premixed jet flame features a burner with a
C approach. central jet of pure C2 H4 and low-speed pilot injection of fully re-
The schematic illustration of the LES-FGM-DSM implementation acted C2 H4 -air pre-mixture at an equivalence ratio of 0.9. The pilot
is shown in Fig. 1. The LES formalism with the aforementioned tube (with an inner diameter of 15.2mm, and an outer diameter of
soot source term models is applied to the simulation of soot pro- 19.1mm) comprises an insert, which provides 64 small pilot flames
duction in a turbulent piloted jet flame. The piloted jet flame offers for the stabilization of the main flame. The spacing and number
an ideal combustion environment to analyze the evolution of soot of pilot flames have been designed to produce a uniform flow rate
formation sub-processes, such as nucleation (near to burner) fol- of burned products near the burner exit plane. The global mass
lowed by soot growth (within flame mid-height), and soot oxida- flow rate of the pilot corresponds to 1.77 X 10−4 kg/s. Fuel (C2 H4 )
tion (near the flame tip). The experimental details and numerical is injected at 294K with a bulk velocity of 54.7m/s, resulting in a
setup for the LES of this turbulent jet flame are presented in the Reynolds number of 20,0 0 0 (based on the diameter of the main
following section. jet d j = 3.2mm) at the fuel nozzle exit. The pilot is surrounded by
a co-flow of air at a bulk velocity of 0.6m/s. Additional details on
the experimental conditions can be found in Ref. [67].
3. Turbulent sooting flame configuration
3.2. Numerical setup
The turbulent non-premixed ethylene-air piloted jet flame, ex-
perimentally investigated at Sandia National Laboratories [67] is Based on the experimental observation of the visible flame
considered here. This flame is also one of the target flames from height, the computational domain is designed to be 1m×0.3m×2π
the International Sooting Flame (ISF) workshop [68]. The selected in axial (z), radial (r) and circumferential directions, respectively.
flame presents well-characterized exit conditions for various noz- The computational domain is discretized via 384×192×64 cells
zles. It provides an extensive experimental database of soot mea- in the axial, radial, and circumferential directions, respectively. A
surements, which justifies the appropriateness of this flame for schematic illustration of the computational domain and grid is
soot model assessment and validation. Unfortunately, measure- given in Fig. 2a. The computational grid is selected such that the
ments for the velocity field and/or mixing are not available for the typical cell size near the fuel exit is ≈ 0.1mm. The grid is stretched
selected flame, which presents an additional challenge in the vali- in the axial and radial directions while the circumferential direc-
dation of the turbulent flow field in simulations. The details of the tion is uniformly spaced. In the radial direction, the grid is non-
experimental configuration and simulation setup are presented be- uniformly concentrated in the shear layers between the different
low. injection streams. The structured grid is made of approximately 5

6
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

Fig. 2. Illustration of the computational domain and grid (a), and the probability density function of the M parameter in Pope’s criteria for the LES grid (b).

million hexahedral elements. The mesh has been also verified for istry including soot kinetics using the code CHEM1D [71]. In LES
Pope’s criteria [69] in Fig. 2b. More than 90% of the grid cells show simulations, 30 soot sections are transported to describe particle
a ratio of resolved to total kinetic energy (M) larger than 0.8, indi- size distribution, hence the flamelets are also computed with 30
cating the kinetic energy is sufficiently resolved with the selected sections. To cover the composition space from chemical equilib-
mesh. rium to mixing in the flamelet database, first, a series of strained
Synthetic turbulence derived from a Laplacian filter following steady counterflow flamelets are computed by varying the applied
the method proposed by Kempf et al. [70] is employed to define stain rate from lower values (close to chemical equilibrium) until
the inflow condition for the fuel. The mean axial velocity of the the extinction limit. Subsequently, the composition space between
turbulent database is set to a power-law profile with an exponent the extinction limit and the mixing solution is covered by simulat-
of 1/7. Considering the Reynolds number of a pipe flow at the con- ing unsteady quenching flamelets at the extinguishing strain rate.
ditions of the fuel inlet, the turbulent intensity is around 5%. It is The detailed kinetic scheme KM2 of Wang et al. [72], involving 202
worth noting that there is no consensus in the literature regard- species and 1351 reactions, is used for the gas phase chemistry
ing the selection of inlet boundary conditions in this case. Other during the computation of flamelets. This mechanism has been
numerical investigations [4,31] of similar flames have used turbu- extensively validated for soot formation prediction in ethylene-
lence intensity values ranging from 6% to 10%. Unfortunately, due fueled laminar [65,73,74] and turbulent flames [11]. Considering
to the unavailability of velocity measurements, the velocity profile the large Reynolds number of the turbulent jet, turbulent diffusivi-
and Reynolds stresses can not be compared and it was estimated ties are expected to be higher than molecular diffusivities. Hence a
with 5% in this study. The velocity inlet profiles for the pilot and unity-Lewis diffusion model is considered for the species transport
coflow air streams are treated as uniform. The velocity of the air in the flamelet computation. For the mapping of thermochemical
stream is specified as 0.6m/s while the bulk velocity of the pilot variables, the progress variable (Eq. (4)) is defined based on H2 O,
is adjusted to impose the mass flow rate satisfying the experimen- CO2 , CO, O2 , H2 , and A4 species mass fractions with their corre-
tal condition. No-slip adiabatic wall conditions are imposed at the sponding weight factors αH2 O = 0.0555, αCO2 = 0.0228, αCO = 0.0357,
injector boundaries. The pilot inlet is assumed to have a compo- αO2 = −3.13 × 10−4 , αH2 = 0.173, αA4 = 0.0988. The progress vari-
sition close to the equilibrium state of an ethylene-air mixture at able definition is determined using a guess-and-check approach,
an equivalence ratio of 0.9 ( Z = 0.0577, C= 1) and temperature of and shown to preserve the unique mapping of Y in composition
2256K. The fuel and coflow inlet temperatures are maintained at space [75]. Including pyrene (A4) in the progress variable defini-
294K. tion ensures a unique mapping of the tabulated quantities under
For the creation of the FGM database, a series of 1-D non- conditions where the soot models are especially sensitive to the
premixed counterflow flamelets are calculated with detailed chem- local gas phase composition. However, since the A4 species is not

7
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

transported in either of the approaches used for soot source terms tions, the flame reaction zones (indicated by OH contours) show
in the present study, the accuracy of the current FGM-DSM frame- local extinction events within the shear layers. The formation of
work is not significantly impacted by the inclusion of A4 in the PAH (A4) and C2 H2 is predominant in the fuel-rich region. How-
progress variable Y. The thermochemical variables are stored in the ever, a systematic lag can be noticed in the spatial locations of
FGM database with a non-uniform (refined near stoichiometric and A4 formation compared to C2 H2 as the relatively large time scales
equilibrium regions) resolution of 101×11×101×11 grid points in  Z, governing A4 formation reflect in its downstream spatial evolution.
SZ , C, and SC space respectively. The production of A4 is observed to initiate mainly after z/d j  50,
Previous studies [11,76] conducted on similar flames have while C2 H2 , a precursor species responsible for surface growth is
demonstrated the importance of considering thermal radiation ef- found to occur closer to the burner exit (z/d j  20). Furthermore,
fects due to the temperature sensitivity of soot formation. The cur- the regions of higher OH concentration, in which soot oxidation
rent LES-FGM-DSM framework can be expanded to include radia- is dominant, are shown to be prominent in the region beyond
tive heat transfer by further augmentation of the flamelet database z/d j  170.
for non-adiabatic conditions and additional parameterization in en-
thalpy space. However, the primary focus of this study is the com- 4.1.1. Gas-phase validation
parative evaluation of two different methods for modeling soot for- For the preliminary validation of gas-phase, computed time-
mation chemistry. Hence, the effects of thermal radiation from gas averaged radial profiles of mean OH mass fractions at different
and soot are neglected for simplicity. downstream heights are shown in Fig. 4 along with the experi-
The simulations are carried out in a Cartesian coordinate sys- mentally measured OH signal. In addition, a comparison of the PAH
tem using the multi-physics code Alya [77], developed at the signal against computed profiles of mean C2 H2 and A4 mass frac-
Barcelona Supercomputing Center (BSC). In the Alya code, a tions at several axial locations is presented in Fig. 4. Overall, the
second-order conservative finite element scheme is used for spatial computed profiles show fair qualitative agreement with the exper-
discretization, while an explicit third-order Runge-Kutta scheme is iments regarding downstream evolution, suggesting that the tur-
employed for the time integration. A low-dissipation scheme based bulent combustion models applied can favorably capture the main
on the fractional step algorithm proposed by Both et al. [46] is features of the flame structure in the gas phase. A slight overpre-
used for continuity and momentum equations under a low-Mach diction of the jet spreading rate is observed in the radial direction
number approximation of reacting flows. The simulations are per- at downstream positions leading to a wider spread in the species
formed using the Hawk cluster equipped with AMD EPYC 7742 profile. This however is found to have a minor effect on the quan-
processors at the High-Performance Computing Center Stuttgart. titative prediction of soot formation.
The temporal statistics for the quantities are performed for a pe-
riod of approximately 250ms after the simulations have reached a 4.2. Characterization of soot formation
statistically steady state.
4.2.1. Instantaneous fields
4. Results and discussion To investigate the effects of tabulating the source terms and in-
clude the PDF integration on the evolution and distribution of soot,
4.1. Gas-phase characteristics the instantaneous fields of soot volume fraction and soot forma-
+ −
tion rates (split into production ω˙ s and consumption ω˙ s parts) ob-
To describe the main combustion features of the piloted tur- tained with FGM-C and FGM-T approaches are compared in Fig. 5.
bulent jet flame, the instantaneous fields of temperature, and key From the soot volume fraction fields, it is evident that soot is
species involved in soot chemistry are presented in Fig. 3. An iso- mainly formed in the fuel-rich zones beyond z/d j > 50, character-
contour of stoichiometric mixture fraction (
Z = Zst ) characterizing ized by high A4 mass fractions (see Fig. 3). Soot inception is pre-
the flame front is also shown. The flame is found to be stabilized dominant near the flame base, leading to the formation of small-
by the pilot and attached to the burner. Under turbulent condi- sized particles and, consequently, low values of soot volume frac-

Fig. 3. Instantaneous 2-D fields of temperature, OH mass fraction, C2 H2 mass fraction, A4 mass fraction. The stoichiometric mixture fraction is shown with dashed iso-lines.
The instantaneous results correspond to the FGM-C case.

8
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

Fig. 4. Comparison of radial profiles for computed (normalized with maximum) mean mass fractions against measured signals for OH (left panel), PAH (right panel) at
different axial locations.

tion. From Fig. 5, it can be observed that the soot production rate bution of soot consumption rates is found to be qualitatively and
(dominated by surface growth primarily) is concentrated in the quantitatively somewhat similar in both the FGM-C and FGM-T ap-
middle of the flame. Therefore, a high amount of soot volume frac- proaches. Therefore, the noticed discrepancies in the quantitative
tion is noticed in the middle region of the flame, where incepted prediction of soot with FGM-C and FGM-T methods are primarily
particles grow. On the other hand, consumption rates (oxidation) the consequence of differences in the soot production rates.
are prominent at the tip region of the flame, and near the stoi-
chiometric iso-contour where particles are oxidized due to a high 4.2.2. Soot volume fraction fluctuations
concentration of OH. Similar trends are noticed in the previous nu- Scatter plots of soot volume fraction (colored by temperature)
merical studies [11,78]. In the presence of turbulent fluctuations, against mixture fraction at different positions along the axial direc-
the soot formation is highly intermittent, forming sporadic pockets tion are compared in Fig. 7. The scatter represents instantaneous
that detach from the fuel-rich zone and are convected downstream values collected at several time instants. The conditional means of
where soot oxidation takes place. soot volume fraction are included for reference. It can be observed
The FGM-T method results in a higher soot volume fraction in that a large extent of soot is present in fuel-rich mixture fractions
the middle region of the flame compared to FGM-C. Additionally, while below the stoichiometric values, no soot is observed, as it
the FGM-T model predicted soot formation further upstream than is rapidly oxidized on the lean side. At lower heights, soot vol-
the FGM-C model. This is because the soot production rate in FGM- ume fraction samples span a large range of mixture fractions and
T is higher near the flame base (z/d j  50) and in the middle region temperatures. With an increase in axial location, higher values of
(z/d j  100) than compared to FGM-C, causing higher soot concen- soot volume fraction are detected, however, the spread in mixture
tration. For a more quantitative illustration of soot formation char- fraction space tends to gradually decrease. The decreasing branch
acteristics between FGM-C and FGM-T, the scatter plots of the soot of soot volume fraction in mixture fractions beyond 0.2, vanishes
volume fraction and soot formation rates fields are compared in at the downstream position, due to enhanced mixing leading to
Fig. 6 (for the results from Fig. 5). The scatter plots are colored lower mixture fraction values. The qualitative trends of soot vol-
with the temperature. It is clear that soot formation is predom- ume fraction scatter are similar in FGM-C and FGM-T simulations.
inant in the fuel-rich regions spanning 0.1 <  Z < 0.4. Within this However, more soot production is evident for FGM-T, especially at
sooting region of composition space, higher values of soot produc- higher mixture fractions than their FGM-C counterparts. Far down-
tion rates are evident for FGM-T as compared to FGM-C, which ex- stream of the flame (z/d j  180), a lower soot volume fraction is
plains the higher fv values obtained by FGM-T. Below the stoichio- found in FGM-T compared to FGM-C.
metric mixture fraction (Zst ), almost all the soot is consumed for The correlation between mixture fraction and soot volume frac-
both methods since soot oxidation is predominant in the compo- tion can be further analyzed through conditional probability den-
sition space close to the stoichiometric value. However, the distri- sity functions (P ( fv |
Z )) of soot volume fractions in Fig. 8. The PDFs

9
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

prod cons
Fig. 5. Instantaneous fields of soot volume fraction f v , soot production rate (ω˙ s ), soot consumption rate (ω˙ s ) for the FGM-C (top panel) and FGM-T (bottom panel)
method. The dashed iso-lines denote Zst .

are examined for three fuel-rich intervals. The P ( fv |Z ) at differ- tion results compared to the experiment are more pronounced at
ent axial positions indicate that fluctuations in soot volume frac- lower heights for the FGM-T approach where soot volume fraction
tion are mainly concentrated within rich regions, identified by values are higher as evident in Fig. 5. Nevertheless, the upstream
2Zst < 
Z < 3Zst . Moreover, a substantial amount of soot also exists translation with an underestimation is also noticed in several other
in highly rich regions 
Z ≥ 3Zst . The mean values of fv are lower for numerical works [4,79], hence the overall performance of the cur-
FGM-C as compared to FGM-T. For mixture fractions below 2Zst , rent LES-FGM-DSM approaches is quite reasonable in the context
the peak value of P ( fv |
Z ) approaches zero, confirming a minimal of the current state of the art in modeling turbulent sooting flames.
amount of soot in these regions, as soot oxidation is prominent Nevertheless, it is important to note that while subprocesses like
within lean regions. In the FGM-C method, fluctuations in soot soot growth and condensation contribute to the lower soot inter-
volume fraction primarily arise from resolved fluctuations in the mittency in downstream regions, it is also crucial to emphasize the
flamelet independent variables (Z, C) and turbulent transport of impact of the LES resolution and the flapping of the jet on the in-
soot. In FGM-T, besides fluctuations in flamelet independent vari- termittency.
ables, and turbulent transport of soot, fluctuations in the chemical
source term of soot sections are taken into account through the
presumed-PDF integration during tabulation. 4.2.3. Mean soot profiles
The fluctuations in soot volume fraction are often characterized The time-averaged fields of soot volume fraction obtained from
by soot intermittency (Is ). The soot intermittency is defined exper- LES with FGM-C and FGM-T models are compared in Fig. 10a. As
imentally as the probability of observing an instantaneous value can be observed, soot is predominantly restricted to fuel-rich re-
of soot volume fraction lower than 0.03ppm [11]. In Fig. 9 the gions within the stoichiometric mean mixture fraction iso-contour,
experimental probe-resolved and numerical intermittency profiles while the peak soot volume fraction locations are found at approx-
are shown for sampled data at several time instants. The intermit- imately z/d j  125. Because of the soot oxidation (predominantly
tency profiles are favorably captured beyond z/d j > 120 confirming through OH species near stoichiometric conditions) and flow fluc-
the good prediction of soot particle oxidation and turbulent fluctu- tuations, soot particles do not exist over the complete mixture
ations (resolved). The simulation results tend to underestimate the fraction space. For the FGM-T model, substantial soot concentra-
intermittency close to the burner. The discrepancies in the simula- tion is observed at locations close to the burner exit (z/d j  30),

10
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

Fig. 6. Scatter plots of instantaneous soot volume fraction (top panel) and (bottom panel) soot formation rates for FGM-T and FGM-C methods.

Fig. 7. Comparison of scatter plots of soot volume fraction colored by temperature along with conditional means of f v |
Z (dashed lines) at different axial locations for
FGM-C (a) and FGM-T (b) approaches.

11
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

Fig. 8. Comparison of soot volume fraction PDF conditioned on the mixture fraction at different axial locations along the flame for FGM-C (a), and FGM-T (b).

are compared against the measurements in Fig. 10b. The RMS for
computed soot volume fraction is calculated as:

2
fvRMS = fv − fv 2 . (26)
The qualitative trends in the experimental data are reasonably re-
produced in the simulations. The normalized RMS profiles of soot
volume fraction show very good agreement with measurements
for both approaches, however, the magnitude of the soot volume
fraction is underpredicted. The peaks observed in the soot volume
fraction in the lean region away from the jet (r/d j  5) are not well
captured in the simulations. Contrary to measurements, soot al-
most ceases to exist beyond z/d j  180 in the computed results.
Fig. 9. Comparison of soot intermittency Is profiles along the centerline for FGM-C
The underprediction of soot volume fraction in far downstream re-
and FGM-T approaches (lines) with measurements (symbols). gions could be attributed to either the overprediction of OH oxi-
dation rates or the underprediction of the overall flame length in
while the soot formation is somewhat delayed in FGM-C simu- simulations.
lations. For a more quantitative illustration, the radial profiles of The axial profiles of mean and RMS soot volume fraction along
computed mean and RMS soot volume fraction at several heights the centerline axis are compared in Fig. 11. In the present turbulent

Fig. 10. Time-averaged fields of soot volume fraction for FGM-T, and FGM-C closure models (a), and a comparison between experimental (symbols) and numerical (line) data
for mean and normalized RMS of soot volume fraction profiles at several axial heights (b).

12
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

Fig. 11. Comparison between experimental (symbols) and numerical (line) profiles of mean and RMS of soot volume fraction at the centerline for FGM-T and FGM-C ap-
proaches.

jet flame, the overall prediction of soot formation along the center- high computational efficiency and low computational cost. This as-
line is mainly controlled by surface growth (in the middle region) pect will be addressed in Section 4.3.
and oxidation (due to OH) as shown in [4]. A reasonable agreement As previously mentioned, the FGM-T model accounts for the
between simulated and measured soot volume fraction profiles is subgrid-scale chemistry-soot-turbulence interactions, as the tabu-
obtained, but the peak value of fv is under-predicted by a factor lated source terms are integrated with the presumed β -PDF func-
of two in the simulations. The prediction of RMS fluctuations of tion. To investigate the impact of the presumed-PDF model on soot
soot volume fraction is similar in magnitude to the mean with ap- prediction with FGM-T, additional simulation is carried out by ne-
proximately a factor 2 lower. This underprediction is aligned with glecting the influence of subgrid-scale fluctuations on tabulated
the state-of-the-art results from this flame, where a comparable soot source terms by using a δ -PDF. In Fig. 12, the axial profiles of
under/over prediction is found with other approaches [4,11,12,31]. mean and RMS soot volume fraction along the centerline are com-
From the present LES results, it can be inferred that the FGM- pared for the two PDF functions. The soot volume fraction profiles
DSM formalism can fairly capture the soot volume fraction distri- are only marginally influenced by the exclusion of the presumed-
bution under turbulent conditions. Besides, for both approaches, PDF treatment accounting for subgrid-scale interactions on soot
the position of the peak soot volume fraction is slightly shifted source terms. Such a response of soot formation can be elucidated
upstream compared to experiments, indicating that the soot con- by analyzing the spatial distribution of soot volume fraction and
sumption is predicted early in simulations. Nevertheless, an early mixture fraction variance (which characterize the influence of tur-
formation of soot is evident for the FGM-T approach compared to bulence fluctuations on scalar mixing). The time-averaged fields of
FGM-C. soot volume fraction (with δ -PDF) and mixture fraction variance
In the current LES, soot mass fractions slowly increase while are shown in Fig. 13. As expected, the high variance of mixture
going downstream from the inlet toward a statistically steady state fraction is primarily found in the shear layers generated by strong
from the initial no-soot condition. Therefore, capturing the chemi- velocity gradients between the main jet and coflow. In the base re-
cal trajectories of soot evolution (from the gas phase to the steady gion of the flame, close to the burner, the mixture fraction variance
state) becomes crucial in the accurate prediction of soot formation. is significant. However, soot formation is not predominant in this
In the FGM-C approach, soot subprocesses are explicitly computed region. On the contrary, the soot formation zone is mainly spanned
using the local soot mass fractions. This yields a better qualitative in the mid-flame region, where mixture fraction gradients are low
description of unsteady soot evolution for the LES. On the contrary, (leading to low variance). As a result, the subgrid-scale fluctuations
in FGM-T soot source terms are calculated and stored for soot mass only slightly affect the soot source terms. Consequently, overall
fractions in a steady-state flamelet, therefore, the chemical trajec- soot formation is found to be only marginally affected by subgrid-
tories concerning the formation of soot from the gas phase to the scale turbulent fluctuations in the present flame.
steady state are not explicitly retained in the FGM tabulation strat-
egy. Naturally, an extension of FGM to incorporate the chemical 4.2.4. Time-averaged particle size distributions
trajectories of soot formation would require augmentation of the The coupling of LES with the sectional method provides infor-
database with unsteady flamelets (at every level of scalar dissipa- mation on the spatio-temporal evolution of the soot PSD. Hence,
tion rate), and an additional controlling variable to entirely param- the calculated time-averaged PSDs at different axial locations along
eterize the reaction progress of soot, making manifold generation the flame centerline are plotted in Fig. 14 for FGM-T and FGM-
more complex. Therefore, this aspect is left out of the scope of the C. The evolution of the PSD along its trajectory in the flame is
present work. Moreover, the non-linear dependency of soot pro- strongly correlated to the particle history characterized by a suc-
duction rates on soot variables is not included in the present FGM- cession of chemical and collisional processes associated with soot
T formulation. Hence, the direct look-up of soot production rates formation. The time-averaged soot PSDs feature mainly unimodal
may lead to their overestimation. As a result, higher soot volume shapes for FGM-C and FGM-T approaches. The number density of
fractions are noticed at lower axial positions for FGM-T compared larger-sized particles increases during this process as nucleated
to FGM-C. Nevertheless, the results for both approaches are overall soot particles grow primarily through surface reactions. For FGM-T,
in good agreement with the current state-of-the-art of this flame the shift of PSD towards large diameters is observed as compared
and the FGM-T approach shows high potential for LES due to the to FGM-C, which translates into higher soot volume fractions pre-

13
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

Fig. 12. Comparison between experimental (symbols) and numerical (line) profiles of mean and RMS of soot volume fraction at the centerline for FGM-T case with β and δ
PDF integration applied to tabulated soot rates.

ferences in the formation and unsteady evolution of soot in both


approaches. For instance, in FGM-C, the coagulation of soot parti-
cles is explicitly computed, contrary to the FGM-T model.

4.2.5. Extension of the FGM-C model


Besides the two methods mentioned in Section 2 for soot
source term closure, an additional model is considered as an ex-
tension of the FGM-C approach. Similar to FGM-C, this model re-
lies on the run-time computation of soot source terms while tab-
ulating the pre-computed gas-phase rate contribution in the FGM
database. However, the filtered soot source term formulation dif-
fers slightly from the FGM-C approach, as discussed below. In a
general form, the sectional soot source term can be expressed as:


ω˙ s,i = ω˙ s,ij (φg , φs ), (27)
j


= kgj (φg )ζi j (φg )ij (φs ) ∀ i ∈ [1, nsec ], (28)
j

where j refers to the soot sub-process (e.g. nucleation), kg is the


gas-phase factor independent of i, the ζi term depends on the sec-
tion (e.g. collision frequency factor), while i is a function of soot
scalars (e.g. number density). The closure of filtered soot source
term is achieved using the presumed-PDF approach:
  
1 (φg , φs )dφg dφs .
ω˙ s,i = ρ̄ kgj (φg )ζi j (φg )ij (φs )P (29)
j
ρ
Since the time scales associated with the evolution of the thermo-
Fig. 13. Time averaged fields of mean soot volume fraction and mixture fraction chemical state are typically smaller than the time scales of soot
variance (range adjusted for better visualization) for FGM-T case with δ -PDF inte-
production [7], the independence between gas and soot phases is
gration applied to tabulated soot rates.
assumed, as discussed in Refs. [7,11], which gives:
      j 
 1 j
ω˙ s,i = ρ̄ (φg )dφg
k g ( φg ) P   j ( φ )P ( φ |φ )d φ ,
ζi φ
dicted by the FGM-T method. Beyond z/d j ≈ 150, the number den- ρ g i s s g s
j
    
sity starts to drop due to the combined effect of soot oxidation and  ,
kgj ij (φg φs )
fuel-lean conditions. As noticed earlier in Fig. 11 for both FGM-T
(30)
and FGM-C, the soot concentration is almost negligible at around
z/d j ≈ 200 since large soot particles are oxidized. This leads to a re- j
where kg represents the filtered rates for the soot subprocesses,
duction in the number density of larger-sized soot particles down-
j
stream of the flame (as can be observed for z/d j ≈ 195 in Fig. 14). and depend only on the gas phase. The kg terms are tabulated in
In general, both FGM-T and FGM-C demonstrate similar qualita- the manifold as a function of control variables 
Z and C. The second
tive features of soot PSD evolution in this turbulent jet flame. The term i in Eq. (30) depends on the soot variable (φ
j  =Y here).
s s,i
quantitative discrepancies in PSD profiles can be attributed to dif- This term is computed at runtime. For model simplicity, the ζi

14
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

a-priori from flamelets and directly tabulated in the manifold.


j 
Since k (φ ) = k (φ
j
g g g ), the FGM-CR model essentially facilitates a
g
more physically consistent treatment of subgrid-scale fluctuations
of gas-phase variables, as compared to the FGM-C approach. At
the same time, similar to FGM-C, subgrid-scale turbulence-soot
interactions are not accounted for in the FGM-CR approach. This
strategy, therefore, can be regarded as an extension of the FGM-C
model.
The fields of mean soot volume fraction are compared in
Fig. 15a for the FGM-CR and FGM-C models. In general, the soot
formation zones in FGM-CR and FGM-C show similarities in the
spatial distribution as well as the magnitude of the soot volume
fraction. A slight reduction in soot volume fraction is noticed near
the flame tip. For more quantitative illustration, the radial profiles
of mean and RMS (normalized with peak) soot volume fraction at
several heights are compared against the measurements and FGM-
C model in Fig. 15b. The results obtained with the FGM-T method
are also presented for reference. It can be observed the radial dis-
tribution of soot volume profiles for FGM-CR case exhibit quali-
tative similarities with FGM-C results. In the far downstream re-
gion, (close to the flame tip) where soot oxidation is prevalent,
the slightly lower soot volume fraction is predicted by the FGM-
CR strategy, as compared to FGM-C. The overall shape and peak
location of the soot volume fraction obtained with FGM-CR agree
closely with FGM-C. However, slight deviations between FGM-C
and FGM-CR are evident in the downstream regions. The strong
intermittency is evident (Fig. 5) in downstream regions where soot
pockets detach near the flame tip. This intermittent behavior is
captured differently during soot source term filtering in the FGM-
C and FGM-CR models. As a result, deviations are also seen in the
soot volume fraction.

4.3. Remark on computational performance

Besides the gas-phase chemistry, in sooting flame simulations


with the sectional method, the greatest CPU overhead is due
to the computation of soot source terms. Especially the calcula-
tion of particle dynamics (coagulation) takes almost half of the
total CPU time. Employing FGM chemistry leads to a signifi-
cant reduction (up to factor 2) in CPU cost for gas-phase reac-
tions. However, computation of soot source terms at runtime re-
mains a CPU-intensive task, which is alleviated here with tab-
ulation of the sectional source term in the FGM-T approach.
In the current LES with 30 soot sections, the FGM-T approach
yields a factor 3 reduction in CPU time per time-step as com-
pared to FGM-C. The computational speedup with FGM-T for a
higher number of sections can be even more significant since
the computational time scales non-linearly with nsec . For in-
Fig. 14. Time-averaged soot PSDs at different axial locations along the centerline
for FGM-T and FGM-C approaches.
stance, the FGM-T approach provides a computational speed-up of
about a factor 7 compared to FGM-C when 60 soot sections are
considered.
In summary, it is evident from the LES results that the model-
term, which depends on the sectional soot properties (such as soot ing of filtered soot source terms strongly influences the soot for-
volume) and gas-phase variables (such as temperature) is com- mation prediction in turbulent flames. The complete tabulation of
puted during simulation. To account for subgrid-scale turbulence- soot chemistry (FGM-T) tends to effectively capture the experimen-
j
chemistry interactions, the gas-phase contribution, kg , is modeled tally observed features of soot distribution in turbulent conditions.
with the presumed-PDF approach. The marginal PDF P (φg ) is as- Compared to the FGM-C method, in which soot chemistry is cal-
sumed to have a β function. The conditional PDF of solid phase culated during runtime, the FGM-T method tends to overpredict
contribution P (φs |φg ) is treated as a δ function. For brevity, this the soot concentration. However, considering the state-of-the-art
method is referred to as FGM-CR hereon. in the numerical prediction of soot formation in turbulent flames
It is important to note that FGM-C and FGM-CR methods pri- and the experimental uncertainty in the measured data, this dis-
marily differ in the treatment of the gas-phase contribution in crepancy can be considered acceptable for practical applications.
j It is also important to emphasize the good computational effi-
the closure of soot source terms. In the FGM-C case, kg is com-
ciency of FGM-T within the context of the sectional method which
.
puted on-the-fly using tabulated filtered gas-phase quantities φ g makes it an attractive alternative to the FGM-C. However, as al-
j
On the other hand in the FGM-CR, the filtered kg is a-computed ready highlighted, the complete tabulation of the soot chemistry

15
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

Fig. 15. Time-averaged fields of soot volume fraction for FGM-C, and FGM-CR closure models (a), and a comparison between experimental (symbols) and numerical (line)
data for mean and normalized RMS of soot volume fraction profiles at several axial heights (b).

has certain limitations, for example, soot-independent treatment of with a presumed PDF (probability density function) approach tends
the soot production term, and a lack of information on the gas-to- to yield good quantitative soot prediction, making it a promising
soot history effects in flamelets. Therefore, in simulation applica- tool to study soot production in light of industrial applications
tions where the unsteady evolution of soot quantities is not of in- with LES. However, it was found that tabulating soot chemistry has
terest, the FGM-T approach is more suitable for predicting the soot limitations in capturing the transient evolution of soot. Therefore,
formation. Especially, in simulations of practical combustion sys- the tabulated soot source term approach requires further improve-
tems, the FGM-T is an excellent strategy to gain an understanding ments to meet limitations regarding capturing soot history effects.
of soot formation and information on size distribution, at afford- On the other end, the model that solves for the complete set of
able computational cost. On the other hand, for more fundamental soot formation subprocesses provides a better qualitative descrip-
and parametric studies, the FGM-C method is recommended as it tion of soot formation and evolution. It is further observed that
accounts for the non-linear interactions between the soot and gas the separation of soot and gas phase terms only marginally im-
phase, without any modeling assumptions, and can be more reli- pacts soot formation prediction for the run-time soot source term
able for highly transient cases. computation model in turbulent conditions.
In summary, the present study demonstrates that the sectional
5. Concluding remarks soot model coupled with FGM chemistry is a feasible and efficient
approach to predict soot production in turbulent flames. The LES
This study presents two strategies based on the discrete sec- framework developed in this work for the FGM-DSM coupling does
tional method coupled with FGM tabulated chemistry in the not consider subgrid-scale interactions between soot and turbu-
context of LES for the prediction of soot formation in turbulent lence. Including subgrid-scale soot-turbulence interaction models
non-premixed flames. The performance of the two strategies (e.g. Ref. [7,11]) in the LES would be an important extension to in-
for modeling soot source terms is assessed on a turbulent vestigate their impact on overall soot prediction. In addition, future
non-premixed jet flame with a focus on the prediction of soot for- work could consider incorporating developing new models or ex-
mation and particle size distributions. The LES results for the gas tending the linearization model used in the tabulated soot chem-
phase and soot phase are compared against the available experi- istry approach to account for the non-linear effects of the soot pro-
mental data. Despite some discrepancies, the results for the soot duction rate. Moreover, the authors acknowledge that the modeling
phase show reasonable agreement with the experimental results. aspects related to thermal radiation, preferential diffusion, and soot
The LES study suggests that the soot source term closure in agglomeration hold significant potential for enhancing the accuracy
FGM-DSM coupling substantially influences the quantitative and of the proposed soot modeling framework and are thus considered
qualitative prediction of soot formation in turbulent conditions. essential components for future research work.
The observed differences in soot prediction for the investigated
soot source-term models can be attributed to their capabilities in Novelty and significance
capturing unsteady chemical trajectories of soot formation, which
evolve at a slower time scale than flame propagation. The compu- Applications of detailed soot models, such as the discrete sec-
tationally efficient model based on tabulated soot formation rates tional method, have become increasingly important with emission

16
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

regulations on soot particle size. However, several studies that em- [10] K. Netzell, H. Lehtiniemi, F. Mauss, Calculating the soot particle size distri-
ploy sectional methods for soot prediction in turbulent conditions bution function in turbulent diffusion flames using a sectional method, Proc.
Combust. Inst. 31 (1) (2007) 667–674.
rely on the explicit computation of soot reaction rates, which suf- [11] P. Rodrigues, B. Franzelli, R. Vicquelin, O. Gicquel, N. Darabiha, Coupling an
fer from high computational costs with an increase in the num- LES approach and a soot sectional model for the study of sooting turbulent
ber of soot sections. To address this, the present work offers a non-premixed flames, Combust. Flame 190 (2018) 477–499.
[12] Y. Xuan, G. Blanquart, Effects of aromatic chemistry-turbulence interactions on
computationally efficient tabulated soot chemistry-based sectional soot formation in a turbulent non-premixed flame, Proc. Combust. Inst. 35 (2)
soot modeling framework for the LES approach. The compara- (2015) 1911–1919.
tive analysis presented in this study demonstrates the good pre- [13] F. Ferraro, S. Gierth, S. Salenbauch, W. Han, C. Hasse, Soot particle size distribu-
tion reconstruction in a turbulent sooting flame with the split-based extended
dictive capabilities of the proposed method despite several mod-
quadrature method of moments, Phys. Fluids 34 (7) (2022) 075121.
eling approximations. Overall, the proposed LES framework of- [14] M. Grader, C. Eberle, P. Gerlinger, Large-eddy simulation and analysis of a soot-
fers an attractive modeling choice for reducing the computational ing lifted turbulent jet flame, Combust. Flame 215 (2020) 458–470.
[15] S. Rigopoulos, Modelling of soot aerosol dynamics in turbulent flow, Flow Tur-
cost of sooting flame calculations in industry-relevant conditions
bul. Combust. 103 (3) (2019) 565–604.
while maintaining good predictive capabilities. The proposed tabu- [16] T. Poinsot, D. Veynante, Theoretical and numerical combustion, RT Edwards,
lated soot chemistry model, therefore, has the potential to signif- Inc., 2005.
icantly advance the design and development of clean combustion [17] F. Bisetti, G. Blanquart, M.E. Mueller, H. Pitsch, On the formation and early
evolution of soot in turbulent nonpremixed flames, Combust. Flame 159 (1)
systems. (2012) 317–335.
[18] A. Attili, F. Bisetti, M.E. Mueller, H. Pitsch, Formation, growth, and transport of
Declaration of Competing Interest soot in a three-dimensional turbulent non-premixed jet flame, Combust. Flame
161 (7) (2014) 1849–1865.
[19] A. Wick, A. Attili, F. Bisetti, H. Pitsch, DNS-driven analysis of the
The authors declare that they have no known competing finan- flamelet/progress variable model assumptions on soot inception, growth, and
cial interests or personal relationships that could have appeared to oxidation in turbulent flames, Combust. Flame 214 (2020) 437–449.
[20] D. Aubagnac-Karkar, J. Michel, O. Colin, P. Vervisch-Kljakic, N. Darabiha, Sec-
influence the work reported in this paper. tional soot model coupled to tabulated chemistry for diesel RANS simulations,
Combust. Flame 162 (8) (2015) 3081–3099.
CRediT authorship contribution statement [21] S. Yang, J.K. Lew, M.E. Mueller, Large eddy simulation of soot evolution in tur-
bulent reacting flows: strain-sensitive transport approach for polycyclic aro-
matic hydrocarbons, Combust. Flame 220 (2020) 219–234.
Abhijit Kalbhor: Conceptualization, Methodology, Visualization, [22] D. Mira, O. Lehmkuhl, A. Both, P. Stathopoulos, T. Tanneberger, T.G. Reichel,
Investigation, Formal analysis, Writing – original draft. Daniel C.O. Paschereit, M. Vázquez, G. Houzeaux, Numerical characterization of a pre-
mixed hydrogen flame under conditions close to flashback, Flow Turbul. Com-
Mira: Conceptualization, Visualization, Resources, Writing – review
bust. 104 (2) (2020) 479–507.
& editing. Ambrus Both: Validation, Writing – review & editing. [23] B. Franzelli, L. Tardelli, M. Stöhr, K. Geigle, P. Domingo, Assessment of LES of
Jeroen van Oijen: Conceptualization, Funding acquisition, Supervi- intermittent soot production in an aero-engine model combustor using high-
sion, Writing – review & editing. -speed measurements, Proc. Combust. Inst. (2022).
[24] S. Gkantonas, M. Sirignano, A. Giusti, A. D’Anna, E. Mastorakos, Comprehensive
soot particle size distribution modelling of a model Rich-Quench-Lean burner,
Acknowledgments Fuel 270 (2020) 117483.
[25] J. Benajes, J.M. García-Oliver, J.M. Pastor, I. Olmeda, A. Both, D. Mira, Analysis of
local extinction of a n-heptane spray flame using large-eddy simulation with
The research leading to these results has received fund- tabulated chemistry, Combust. Flame 235 (2022) 111730.
ing from the European Union’s Horizon 2020 Programme un- [26] H. Bao, A. Kalbhor, N. Maes, B. Somers, J. van Oijen, Investigation of soot for-
der the ESTiMatE project, grant agreement No. 821418, and the mation in n-dodecane spray flames using LES and a discrete sectional method,
Proc. Combust. Inst. (2022).
AHEAD PID2020-118387RB-C33 and SAFLOW TED2021-131618B- [27] B. Franzelli, E. Riber, B. Cuenot, M. Ihme, Numerical modeling of soot pro-
C21 projects from the Ministerio de Ciencia e Innovación. DM duction in aero-engine combustors using large eddy simulations, Turbo Expo:
acknowledges the grant from the Ministerio de Ciencia e Inno- Power for Land, Sea, and Air, vol. 56697, American Society of Mechanical En-
gineers, 2015. p. V04BT04A049
vación, Ayudas para contratos Ramón y Cajal (RYC) 2021: RYC2021-
[28] J. Consalvi, F. Nmira, Transported scalar PDF modeling of oxygen-enriched tur-
034654-I. The authors gratefully acknowledge the Partnership for bulent jet diffusion flames: soot production and radiative heat transfer, Fuel
Advanced Computing in Europe (PRACE) grant Soot Aero for com- 178 (2016) 37–48.
[29] L. Pachano, C. Xu, J.M. García-Oliver, J.M. Pastor, R. Novella, P. Kundu, A two-e-
putational resources and technical support in the numerical simu-
quation soot-in-flamelet modeling approach applied under spray a conditions,
lations. Combust. Flame 231 (2021) 111488.
[30] S. Yang, J. Lew, M. Mueller, Large eddy simulation of soot evolution in turbu-
References lent reacting flows: presumed subfilter PDF model for soot–turbulence–chem-
istry interactions, Combust. Flame 209 (2019) 200–213.
[1] EEA, Regulation (EC) no 715/2007 of the European parliament and of the coun- [31] L. Tian, M. Schiener, R. Lindstedt, Fully coupled sectional modelling of soot
cil of 20 June 2007 on type approval of motor vehicles with respect to emis- particle dynamics in a turbulent diffusion flame, Proc. Combust. Inst. 38 (1)
sions from light passenger and commercial vehicles (Euro 5 and Euro 6), 2022. (2021) 1365–1373.
[2] Committee on aviation environmental protection report (doc 10126). ed. [32] A. Felden, E. Riber, B. Cuenot, Impact of direct integration of analytically re-
by international civil aviation organization, 2023, (https://store.icao.int/en/ duced chemistry in LES of a sooting swirled non-premixed combustor, Com-
committee- on- aviation- environmental- protection- report- doc- 10126). bust. Flame 191 (2018) 270–286.
[3] S. Rigopoulos, Population balance modelling of polydispersed particles in reac- [33] H.M. Colmán, N. Darabiha, D. Veynante, B. Fiorina, A turbulent combustion
tive flows, Prog. Energy Combust. Sci. 36 (4) (2010) 412–443. model for soot formation at the LES subgrid-scale using virtual chemistry ap-
[4] B. Sun, S. Rigopoulos, Modelling of soot formation and aggregation in turbulent proach, Combust. Flame 247 (2023) 112496.
flows with the LES-PBE-PDF approach and a conservative sectional method, [34] M.E. Mueller, H. Pitsch, Large eddy simulation subfilter modeling of soot-tur-
Combust. Flame 242 (2022) 112152. bulence interactions, Phys. Fluids 23 (11) (2011) 115104.
[5] M. Balthasar, M. Kraft, A stochastic approach to calculate the particle size [35] S.T. Chong, V. Raman, M.E. Mueller, P. Selvaraj, H.G. Im, Effect of soot model,
distribution function of soot particles in laminar premixed flames, Combust. moment method, and chemical kinetics on soot formation in a model aircraft
Flame 133 (3) (2003) 289–298. combustor, Proc. Combust. Inst. 37 (1) (2019) 1065–1074.
[6] M. Mueller, G. Blanquart, H. Pitsch, Hybrid method of moments for mod- [36] L. Berger, A. Wick, A. Attili, M.E. Mueller, H. Pitsch, Modeling subfilter soot-tur-
eling soot formation and growth, Combust. Flame 156 (6) (2009) 1143– bulence interactions in large eddy simulation: an a priori study, Proc. Combust.
1155. Inst. 38 (2) (2021) 2783–2790.
[7] M. Mueller, H. Pitsch, LES model for sooting turbulent nonpremixed flames, [37] H.M. Colmán, A. Attili, M.E. Mueller, Large eddy simulation of turbulent non-
Combust. Flame 159 (6) (2012) 2166–2180. premixed sooting flames: presumed subfilter PDF model for finite-rate oxida-
[8] A. D’anna, J. Kent, A model of particulate and species formation applied to tion of soot, Combust. Flame (2023) 112602.
laminar, nonpremixed flames for three aliphatic-hydrocarbon fuels, Combust. [38] H. El-Asrag, S. Menon, Large eddy simulation of soot formation in a turbulent
Flame 152 (4) (2008) 573–587. non-premixed jet flame, Combust. Flame 156 (2) (2009) 385–395.
[9] F. Gelbard, Y. Tambour, J. Seinfeld, Sectional representations for simulating [39] C.D. Pierce, P. Moin, Progress-variable approach for large-eddy simulation of
aerosol dynamics, J. Colloid Interface Sci. 76 (2) (1980) 541–556. non-premixed turbulent combustion, J. Fluid Mech. 504 (2004) 73–97.

17
A. Kalbhor, D. Mira, A. Both et al. Combustion and Flame 259 (2024) 113128

[40] F. Sewerin, S. Rigopoulos, An LES-PBE-PDF approach for predicting the soot [61] S.P. Roy, Aerosol-dynamics-based soot modeling of flames, Pennsylvania State
particle size distribution in turbulent flames, Combust. Flame 189 (2018) University, 2014 Ph.D. thesis.
62–76. [62] J. Appel, H. Bockhorn, M. Frenklach, Kinetic modeling of soot formation with
[41] J. van Oijen, L. de Goey, Modelling of premixed laminar flames using flamelet– detailed chemistry and physics: laminar premixed flames of C2 hydrocarbons,
generated manifolds, Combust. Sci. Technol. 161 (1) (20 0 0) 113–137. Combust. Flame 121 (1–2) (20 0 0) 122–136.
[42] J. van Oijen, A. Donini, R. Bastiaans, J. ten Thije Boonkkamp, L. de Goey, [63] S. Kumar, D. Ramkrishna, On the solution of population balance equa-
State-of-the-art in premixed combustion modeling using flamelet generated tions by discretization–I. A fixed pivot technique, Chem. Eng. Sci. 51 (8) (1996)
manifolds, Prog. Energy Combust. Sci. 57 (2016) 30–74. 1311–1332.
[43] D. Carbonell, A. Oliva, C.D. Perez-Segarra, Implementation of two-equation soot [64] C. Hoerlle, Modelling of soot formation based on the Discrete Sectional
flamelet models for laminar diffusion flames, Combust. Flame 156 (3) (2009) Method: CO2 effects and coupling with the FGM technique, Universidade Fed-
621–632. eral do Rio Grande do Sul, Porto Alegre, Brazil, 2020 Ph.D. thesis.
[44] A. Kalbhor, D. Mira, J. van Oijen, A computationally efficient approach for soot [65] A. Kalbhor, J. van Oijen, Effects of hydrogen enrichment and water vapour dilu-
modeling with discrete sectional method and FGM chemistry, Combust. Flame tion on soot formation in laminar ethylene counterflow flames, Int. J. Hydrogen
255 (2023) 112868. Energy 45 (43) (2020) 23653–23673.
[45] A. Vreman, An eddy-viscosity subgrid-scale model for turbulent shear flow: [66] M. Ihme, H. Pitsch, Modeling of radiation and nitric oxide formation in turbu-
algebraic theory and applications, Phys. Fluids 16 (10) (2004) 3670–3681. lent nonpremixed flames using a flamelet/progress variable formulation, Phys.
[46] A. Both, O. Lehmkuhl, D. Mira, M. Ortega, Low-dissipation finite element strat- Fluids 20 (5) (2008) 055110.
egy for low mach number reacting flows, Comput. Fluids 200 (2020) 104436. [67] J. Zhang, C.R. Shaddix, R.W. Schefer, Design of “model-friendly” turbulent
[47] N. Peters, Laminar flamelet concepts in turbulent combustion, Proc. Combust. non-premixed jet burners for C2+ hydrocarbon fuels, Rev. Sci. Instrum. 82 (7)
Inst. 21 (1) (1988) 1231–1250. (2011) 074101.
[48] J. van Oijen, Flamelet-generated manifolds: development and application to [68] International sooting flame (ISF) workshop, 2022, (https://www.adelaide.edu.
premixed laminar flames, Eindhoven University of Technology, Eindhoven, the au/cet/isfworkshop/).
Netherlands, 2002 Ph.D. thesis. [69] S.B. Pope, Ten questions concerning the large-eddy simulation of turbulent
[49] R. Bilger, S. Stårner, R. Kee, On reduced mechanisms for methane-air combus- flows, New J. Phys. 6 (1) (2004) 35.
tion in nonpremixed flames, Combust. Flame 80 (2) (1990) 135–149. [70] A. Kempf, M. Klein, J. Janicka, Efficient generation of initial-and inflow-condi-
[50] N. Malik, T. Løvås, F. Mauss, The effect of preferential diffusion on the soot tions for transient turbulent flows in arbitrary geometries, Flow Turbul. Com-
initiation process in ethylene diffusion flames, Flow Turbul. Combust. 87 (2) bust. 74 (1) (2005) 67–84.
(2011) 293–312. [71] L. Somers, The simulation of flat flames with detailed and reduced chemical
[51] A. Attili, F. Bisetti, M.E. Mueller, H. Pitsch, Effects of non-unity lewis number models, Eindhoven University of Technology, Eindhoven, the Netherlands, 1994
of gas-phase species in turbulent nonpremixed sooting flames, Combust. Flame Ph.D. thesis.
166 (2016) 192–202. [72] Y. Wang, A. Raj, S. Chung, A PAH growth mechanism and synergistic effect on
[52] P. Domingo, L. Vervisch, D. Veynante, Large-eddy simulation of a lifted PAH formation in counterflow diffusion flames, Combust. Flame 160 (9) (2013)
methane jet flame in a vitiated coflow, Combust. Flame 152 (3) (2008) 1667–1676.
415–432. [73] Y. Wang, A. Raj, S. Chung, Soot modeling of counterflow diffusion flames of
[53] A. Vreman, B. Albrecht, J. Van Oijen, L. De Goey, R. Bastiaans, Premixed and ethylene-based binary mixture fuels, Combust. Flame 162 (3) (2015) 586–596.
nonpremixed generated manifolds in large-eddy simulation of Sandia flame D [74] Y. Wang, S. Chung, Strain rate effect on sooting characteristics in laminar coun-
and F, Combust. Flame 153 (3) (2008) 394–416. terflow diffusion flames, Combust. Flame 165 (2016) 433–444.
[54] L. Ma, X. Huang, D. Roekaerts, Large eddy simulation of CO2 diluted oxy-fuel [75] A. Kalbhor, J. van Oijen, An assessment of the sectional soot model and FGM
spray flames, Fuel 201 (2017) 165–175. tabulated chemistry coupling in laminar flame simulations, Combust. Flame
[55] Y. Hu, R. Kurose, Large-eddy simulation of turbulent autoigniting hydrogen 229 (2021) 111381.
lifted jet flame with a multi-regime flamelet approach, Int. J. Hydrogen Energy [76] F. Liu, J.-L. Consalvi, F. Nmira, The importance of accurately modelling soot and
44 (12) (2019) 6313–6324. radiation coupling in laminar and laboratory-scale turbulent diffusion flames,
[56] J.C. Massey, Z.X. Chen, N. Swaminathan, Modelling heat loss effects in the large Combust. Flame (2022) 112573.
eddy simulation of a lean swirl-stabilised flame, Flow Turbul. Combust. 106 (4) [77] M. Vázquez, G. Houzeaux, S. Koric, A. Artigues, J. Aguado-Sierra, R. Arís,
(2021) 1355–1378. D. Mira, H. Calmet, F. Cucchietti, H. Owen, et al., Alya: multiphysics engineer-
[57] A. Both, High-fidelity numerical simulations of reacting flows with tabulated ing simulation toward exascale, J. Comput. Sci. 14 (2016) 15–27.
chemistry, Universitat Politèchnica de Catalunya, Spain, 2023 Ph.D. thesis. [78] Y. Xuan, G. Blanquart, A flamelet-based a priori analysis on the chemistry tab-
[58] A. Both, O. Lehmkuhl, D. Mira, M. Ortega, Low-dissipation finite element ulation of polycyclic aromatic hydrocarbons in non-premixed flames, Combust.
strategy for low mach number reacting flows, Comput. Fluids 200 (2020) Flame 161 (6) (2014) 1516–1525.
104436. [79] H. Colmán, A. Cuoci, N. Darabiha, B. Fiorina, A virtual chemistry model for
[59] C. Hoerlle, F. Pereira, Effects of CO2 addition on soot formation of ethylene soot prediction in flames including radiative heat transfer, Combust. Flame 238
non-premixed flames under oxygen enriched atmospheres, Combust. Flame (2022) 111879.
203 (2019) 407–423.
[60] S. Friendlander, Smoke, dust and haze: fundamentals of aerosol dynamics, Ox-
ford University Press, New York, USA, 20 0 0.

18

You might also like