You are on page 1of 19

Original Research Article

International Journal of Spray and


Combustion Dynamics
0(0) 1–19
Numerical modeling of soot formation in ! The Author(s) 2016
Reprints and permissions:
a turbulent C2H4/air diffusion flame sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/1756827716638814
scd.sagepub.com

Manedhar Reddy Busupally and Ashoke De

Abstract
Soot formation in a lifted C2H4-Air turbulent diffusion flame is studied using two different paths for soot nucleation and
oxidation; by a 2D axisymmetric RANS simulation using ANSYS FLUENT 15.0. The turbulence-chemistry interactions are
modeled using two different approaches: steady laminar flamelet approach and flamelet-generated manifold. Chemical
mechanism is represented by POLIMI to study the effect of species concentration on soot formation. P1 approximation is
employed to approximate the radiative transfer equation into truncated series expansion in spherical harmonics while
the weighted sum of gray gases is invoked to model the absorption coefficient while the soot model accounts for
nucleation, coagulation, surface growth, and oxidation. The first route for nucleation considers acetylene concentration
as a linear function of soot nucleation rate, whereas the second route considers two and three ring aromatic species as
function of nucleation rate. Equilibrium-based and instantaneous approach has been used to estimate the OH concen-
tration for soot oxidation. Lee and Fenimore-Jones soot oxidation models are studied to shed light on the effect of OH
on soot oxidation. Moreover, the soot-radiation interactions are also included in terms of absorption coefficient of soot.
Furthermore, the soot-turbulence interactions have been invoked using a temperature/mixture fraction-based single
variable PDF. Both the turbulence-chemistry interaction models are able to accurately predict the flame liftoff height, and
for accurate prediction of flame length, radiative heat loss should be accounted in an accurate way. The soot-turbulence
interactions are found sensitive to the PDF used in present study.

Keywords
Soot, radiation, soot–turbulence interaction
Date received: 11 October 2014; accepted: 20 June 2015

chemical reactions by detailed mechanisms to resolve


Introduction the phenomenon such as extinction and ignition, emis-
Since the beginning of human history, combustion has sions like NOx, CO and soot.3 Accurate inclusion of
a significant role in our daily life. Modern society is radiative heat transfer is also important; as its fourth
heavily dependent on the transformation of chemical power dependence on temperature makes it a dominant
energy in fossil fuels to thermal energy by combustion mode for heat transfer and effects NOx emissions
process. However, generation of pollutants such as through mechanism of thermal NO. The incomplete
nitrogen oxides (NOx) and soot results in adverse combustion of hydrocarbons results in the formation
effect on human health. It also results in change in of condensed carbon particles known as soot and an
global environmental pattern due to greenhouse effects. indicator of poor fuel utilization. In diffusion flames,
The design of gas turbine and reciprocating engines are
to be made more efficient and low in pollutant emis-
sions.1,2 Combustion is a complex phenomenon and Department of Aerospace Engineering, Indian Institute of Technology
generally involves complex chemical reactions along Kanpur, Kanpur, India
with heat and mass transfer. Turbulent combustion
Corresponding author:
involves complex interactions between turbulent trans- Ashoke De, Department of Aerospace Engineering, Indian Institute of
ports, kinetic rate of reactions, radiative heat transfer. Technology Kanpur, Kanpur 208016, India.
Combustion simulations require representation of Email: ashoke@iitk.ac.in.

Creative Commons CC-BY-NC: This article is distributed under the terms of the Creative Commons Attribution-NonCommercial 3.0 License (http://www.
creativecommons.org/licenses/by-nc/3.0/) which permits non-commercial use, reproduction and distribution of the work without further permission provided the
original work is attributed as specified on the SAGE and Open Access pages (https://us.sagepub.com/en-us/nam/open-access-at-sage).
2 International Journal of Spray and Combustion Dynamics 0(0)

soot formation occurs in the fuel rich side due to low field. The temperature field is effect by through radi-
concentration of oxidizer species. ation and concentration of gaseous species is effected
The soot evolution process has been reviewed by by the formation of intermediate products. The effect of
Haynes and Wagner,4 Bockhorn,5 Kennedy6 and has detailed chemical mechanisms and radiative heat trans-
been classified into four major sub processes that fer models on soot and incorporation of different tur-
includes the formation and conglomeration of polycyc- bulence-chemistry interaction models is a subject of
lic aromatic hydrocarbons (PAH) (particle inception or research. The complex interactions of soot with turbu-
nucleation), surface reaction of particles, coagulation lence, chemistry and particle dynamics complicates
and oxidation of particles. The formation of particle modeling. Models with varying degree of sophistication
like structure by coagulation of PAH is known as were used in practical systems6 for predicting soot for-
soot particle inception or nucleation. Chemistry of mation rate.
combustion and soot particle dynamics are linked by Empirical or semi-empirical soot models were
soot particle inception, it controls the number of soot applied for most of the studies. Brookes and Moss9
particles emerging from flames. Three different precur- and Kronenburg et al.10 have modeled soot formation
sors are proposed to act as the pathway of inception: in a methane/air flame using an extended flamelet and
polyacetylenes, ionic species and PAHs. Numerous conditional moment closure approach, respectively.
experimental and modeling studies have concluded The soot predictions were in reasonable agreement
PAH as the principal precursor. These first ring aro- with the experimental measurements. A unsteady
matics grow by H-Abstraction-Carbon-Addition Lagrangian flamelet model was used by Pitsch et al.11
(HACA) mechanism to form large PAHs. The growth to model soot formation in an ethylene flame. Detailed
of these particles is limited by oxidation of aromatic chemical kinetics was augmented to describe the first
species. Oxygen plays a dual role, carbon mass is aromatic species and evolution of soot was incorpo-
reduced by oxidation limiting growth and H-radical rated using the method of moment approach. The
formation required for HACA is promoted. The soot predictions were in good agreement with the
number of particles formed depends on the balance of experimental data and improved with the consideration
formation and oxidation. Surface growth is similar to of differential diffusion, the model could not be applied
nucleation but is a heterogeneous process and compli- to complicated geometries as the computation were
cates modeling. Majority of soot mass is accumulated restricted by the unsteady Lagrangian flamelet model.
by soot surface growth. The surface growth of particles A laminar flamelet model was used by Ma et al.12 to
is still uncertain, in general acetylene is considered as model soot formation in two ethylene/air flames. Soot
the principal reactant at the surface of gas phase spe- formation was modeled using a two equation model of
cies. Frenklach et al.7 and Harris et al.8 described the Moss.13 The work focused on the implementation
effect of acetylene on growth. They have parameterized of naphthalene-based soot inception route and effect of
surface growth as function of acetylene concentration soot surface area on growth rate. They have observed
and density of active area of soot surface. Coagulation that naphthalene approach improved soot predictions
is the process of collision between initial soot particles and uncertainties reduced with the consideration of
to form larger particles and occurs along with surface growth as a function of square root of surface area.
growth. Particle coagulation only changes the soot par- A sectional approach was used by Kohler et al.14 to
ticle size distribution. Soot oxidation converts solid model soot formation in a unconfined C2H4/air jet
soot particles into gas phase species, acts as a counter- flame, the turbulent-chemistry interactions were imple-
balance of surface growth. Surface reactions with mented by a multivariate assumed PDF. The model was
molecular oxygen and OH radial lead in the formation able to capture the magnitude of maximum soot
of CO and CO2. Oxidation due to OH is less rigorous volume fraction but discrepancy was observed in the
than molecular oxygen as OH exists in super equilib- location and shape. The discrepancies were attributed
rium concentrations. to the turbulence model. The objective of the present
Parameters that control soot yield can be identified study is (a) to study the effect of different turbulence-
by accurate prediction of soot formation and oxidation chemistry interaction approaches and radiation model-
can save time and cost in design practical systems. ing approaches on flow field; (b) to study the effect of
Because of low molecular diffusivity, soot transport is different soot inception routes and oxidation
governed by convection and gradient of temperature. approaches on soot volume fraction predictions; and
The transport of soot results in a two-way coupling (c) to study the effect of different soot-turbulence inter-
between concentration of species, temperature and action approaches on soot evolution process. The com-
soot. The complex formation process of soot results parison of flow field predictions with experimental data
in large time scales, longer than heat release rate in clearly elucidates the effect of turbulence-chemistry
flames. Thus, soot cannot be decoupled from the flow interactions and radiation on temperature and soot
Busupally and De 3

evolution. The study also examines the ability of an for the unburnt reactants and one for the burnt
empirical approach for prediction of OH radical con- reactants.
centration and its effect soot volume fraction.  
@ @ @ t @~c
ð~cÞ þ ~
ð~vcÞ ¼ þ Sc ð3Þ
Numerical methods @t @xk @xj Sct @xk

The numerical approach using for solving the turbu- The progress variable is defined as equation (4).
lence-chemistry interactions, radiation modeling and  
P
the soot model are described in the following k ak Yk  Yuk Yc
section. c¼ P eq ¼ eq ð4Þ
k ak Y k Yc

The mean reaction rate is modeled as equation (5).


Turbulence-chemistry interactions
X
The fuel and oxidizer can be considered as partially Sc ¼ 0 Ul ð5Þ
premixed at the base, for a lifted turbulent diffusion
flame. The rich and lean regions of the flame can be The flame front
P convolution is defined by the flame
separated by the surface of instantaneous stoichio- surface density .16 The flame surface is approach con-
metric mixture fraction. The burnt and unburnt siders the laminar flame thickness to be larger than the
gases can be separated by an instantaneous flame Kolmogorov eddies, the internal laminar flow is not
front when the mixture of fuel and oxidizer are distorted and turbulence only wrinkles the flame
inhomogeneous and fluctuating. Thus, a partially sheet. The transport equation for flame surface area17
premixed approach is used for this purpose. The is given as equation (6).
position of flame front is determined by solving P
transport equation for the mean reaction progress @ @ X
þ ð~ Þ
variable. The mean mixture fraction and mixture @t @xk
0 P 1
fraction variance transport equations are solved in
@ @ t @  A X
the flow field. Two approaches were used to describe ¼ þ ðP1 þ P2 þ P3 Þ þP4  D
the turbulence-chemistry interactions and explained @xj Sct @xk
in detailed below.
ð6Þ
Steady laminar flamelet approach. The turbulent flame is P
represented as an ensemble of one dimensional laminar where is the flame surface area density, the source
structures in laminar flamelet model.15 The iso-surface term due to turbulence interaction is given by P1, the
of the mixture fraction is used to prescribe the flame source term due to dilatation in the flame is given by
surface within the turbulent flow field. The transport P2,18 the source term due to expansion of burned
equations of the mean mixture fraction and variance gases is given by P3, the source term due to
are given as: normal propagation is given by P4,18 and D is the
! flame area dissipation. The source terms are closed as
@ ~ @  ~ @ t @f~ shown in Table 1. c calculated from c~ as given in
ðf Þ þ uk f ¼ ð1Þ equation (7).
@t @xk @xj t @xk
c~ ub
@ ~ 002 @  ~ 002 c ¼   ð7Þ
ðf Þ þ uk f
@t @xk 1  c~ 1  ub
002 ! 002 !2
@ t @f~ @f~ "
¼ þ Cg t  Cd  f~ 002 ð2Þ
@xj t @xk @xk k

where t , Cg , and Cd are 0.85, 2.86, and 2.0, respect-


Table 1. Source terms for extended coherent flamelet model
ively. The premixed combustion is described by the (by Veynante et al.18).
C-equation approach; the scalar represents the pro-
gress of reaction. In a turbulent flow field, the tem- P1 P2 P3 P4 D
poral and spatial evolution of reaction is described by P P2
21 @ @2 c
a transport equation of C as given in equation (3). C 1:6Kt ~
ðÞ 0:2Ul Ul 0:4Ul
3  @xk c ð1  c Þ @xk 2 ð1  c Þ
varies between zero and one within the flame, zero
4 International Journal of Spray and Combustion Dynamics 0(0)

where c~ is favre averaged progress variable. The turbu- transport equation in terms of un-normalized pro-
lent time scale as is given as equation (8). gress variable (equation (13)) is solved as closure for
FGM.
"
Kt ¼ ½ð1  0 Þ þ 0 k  ð8Þ !
k @Y~ c @  @ @ ~c
Y
 þ  i Y~ c Þ ¼
ðu  eff
D þ S c ð13Þ
@t @xi @xi @xi
where K is the turbulent kinetic energy, " is the
turbulent dissipation, and k intermediate turbulent
net flame stretch (ITNFS). k is given by where S c is modeled from the finite rate kinetic rate as
equation (9). equation (14). is determined from the flamelet library.
Z Z
1
log10 ðk Þ ¼  expððs þ 0:4ÞÞ S c ¼  SFR ðc, fÞPðc, fÞdcdf ¼ S FR ð14Þ
ðs þ 0:4Þ
  0 
u
þ ð1  expððs þ 0:4ÞÞÞ  s  0:11 The thermochemical properties are determined as
Ul
equation (15).
ð9Þ
Z 1 Z 1

where S is defined as equation (10). ~ ¼ c~ ~


b ð f, Þ pð f, Þdfd þ ð1  cÞ u ð f Þ pð f Þdf
0 0
  ð15Þ
lt
s ¼ log10 ð10Þ
0 l

where 0 l is the laminar flame thickness and is modeled


Radiation modeling
as equation (11) In the present work, the medium is considered to be
optically thick and the radiative heat transfer equation
 has been assumed as a truncated series expansion in
0 l ¼ ð11Þ
Ul spherical harmonics (P1 approximation)20 given by
equation (16).
 
The averaged species mass fraction and temperature 1
r  rGv ¼ Kav ðKbv  Gv Þ ð16Þ
for a turbulent flame can be determined as 3ðKav  Ksv Þ  Aksv
equation (12).
where the right side of equation (16) is the source of
Z 1 Z 1 radiative heat added to energy equation. The absorp-
~ ¼ c~ ~
b ð f, Þ pð f, Þfd þ ð1  cÞ u ð f Þ pð f Þdf tion coefficient of the product gasses and soot is
0 0
ð12Þ given as Kav . In the present work, the absorptive
coefficient is modeled by the weighted sum of gray
The non-adiabatic steady flamelets for burnt react- gases (WSGG) model. The non-gray medium of is
ants have an extra dimension of mean enthalpy for represented by four fictions gases and details of the
temperature and mean density. In the presumed PDF approach can be found in Rakesh et al.21 The effect
of equation (12), f and are assumed to be statistically on radiative heat transfer by soot formation is
independent and the PDF pð f, Þ can be assumed as included by the absorption coefficient of soot and
pf ð f Þ p ð Þ. A b-PDF is assumed for pf ð f Þ and double given as equation (17).
delta PDF is assumed for p ð Þ.

av ¼
i þ
soot ð17Þ
Flamelet-generated manifold (FGM) approach. In FGM19
model, a turbulent flames scalar evolution is estimated The absorption coefficient due to soot particulates is
from the scalar evolution in a laminar flame. A diffu- given as:
sion FGM is used in the present study, they are con-
structed by converting the steady laminar flamelets

soot ¼ b1 soot ½1 þ bT ðT  2000Þ ð18Þ
species fields to progress variable c. ~ With the increase
in strain rate, the chemistry of flamelets departs from
equilibrium and progress variable decrease from unity where b1 ¼ 1232.4m2/kg and bT ¼ 4.8  104 K1 are
to extinction, c~extinction . Instead of the C-equation, a given by Taylor et al.22 and Smith et al.23
Busupally and De 5

Soot modeling The default parameters were used in the present


In the present work, two transport equations in terms study and can be found in Hall et al.1 and Brookes
of volume fraction of soot and concentration of nuclei and Moss.13 The mass fraction of soot precursor has
are used to describe the evolution of soot. been calculated as function of mixture fraction. Two
  approaches were used to describe soot oxidation phe-
@Ysoot t dM nomenon, Lee model considered both OH and O224,25
ÞYsoot ¼ r 
þ r  ð~ rYsoot þ ð19Þ
@t soot dt and Fenimore-Jones considers only OH radical25 for
oxidation of soot. The effect of rate of oxidation on
 
@bnuc  t  1 dN net soot production was studied based on collisional
Þbnuc ¼ r 
þ r  ð~ rb þ efficiency parameter and surface kinetics.
@t nuc nuc Nnorm dt
ð20Þ
OH radical concentration determination. In the present
where N is the number density of soot particle and M is
study, two approaches were used to determine the
the soot mass density; as defined by Brookes and
OH radical concentration. In the first approach, the
Moss13 model to predict soot formation in turbulent
OH radical concentration is determined, as given in
diffusion flames. The source terms of instantaneous
equations (24)26,27 and concentration of O radical is
soot particles and net set productions are given equa-
tions (21) and (22), respectively.

    
dN Xprec P T 24RT 12 12 2
¼ C  NA exp   C  dp N ð21Þ
dt RT T soot NA
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl
ffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
Nucleation Coagulation

      "  2=3 #
dM Xprec P T Xsgs P T 1=3 6M
¼ C  MP  exp  þ C  exp  ðNÞ
dt RT T RT T soot
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
Nucleation SurfaceGrowth
"  2=3        #
pffiffiffiffi 1=3 6M XOH P T!,1 XO2 P T!,2
 Coxid  TðNÞ  C!,1  coll exp  þ C!,2  exp 
soot RT T RT T
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
Oxidation
ð22Þ

In the present study, two models are used to study determined by equation (25),28 known as the equili-
the effect of inception rate on soot formation. In the brium approach.
first approach, the soot precursor was consider as a
linear function of acetylene concentration as given in ½OH ¼ 2:129  102 T0:57 e4595=T ½O1=2 ½H2O1=2 ð24Þ
equation (21). In the second, the soot precursor was
considered as a function of two and three ringed aro- ½O ¼ 3:97  105 T1=2 e31090=T ½O2 1=2 ð25Þ
matic species, given in equation (23) (proposed by Hall
et al.1). In the second approach, the OH radical concentra-
"  # tion is determined from the combustion model.
  
dN NA 2 YC2 H2 2 YC6 H5 WH2
¼ 8  C,1   Soot-turbulence interactions. The effect of turbulence on
dt Inc: MP WC2 H2 WC6 H5 YH2
 soot is included by a single variable PDF in terms of
T,1 temperature or mixture fraction to describe the effect of
 exp  þ 8  C,2
T temporal fluctuations on net soot production.
"   #
NA 2 YC2 H2 2 YC6 H6 YC6 H5 WH2 Z
 
MP WC2 H2 WC6 H6 WC6 H5 YH2 S~soot ¼ Ssoot ðxÞPðxÞdx ð26Þ

T,2
 exp  ð23Þ
T where x is instantaneous temperature or mixture frac-
tion. The solution of soot transport is used to obtain
6 International Journal of Spray and Combustion Dynamics 0(0)

the mean value of independent variables required for velocity profiles are slightly adjusted iteratively to
constructing the PDF. match the experimental flow rates.
In the present work, ANSYS FLUENT 15.031 is used
for all the calculations, the non-gray radiative transfer
Description of test case
model is incorporated using an user defined function
A C2H4-air lifted diffusion flame is considered in the (UDF).
present study. The details of the burner are provided by The turbulence-chemistry interactions are modeled
Kohler et al.29,30 The fuel jet has a diameter of 2 mm using steady laminar flamelet approach (SF) and
and the outer diameter of the pipe is 6 mm. The diam- FGM. The flame thickness using ITNFS treatment is
eter of the coflow is 140 mm, the global equivalence calculated to be 0.00026 for SLF model. Chemistry
ratio is 0.48. The schematic of the burner is shown in has been represented by a detailed POLIMI mechanism
Figure 1(a). The soot measurements for the present containing 82 species and 1450 reactions.32 The absorp-
burner were performed by Kohler et al.14,30 For the tion coefficient of the non-gray medium is calculated by
present computational work, an axisymmetric 2D grid the WSGG method, considering four fictitious gases
has been used. Figure 1(b) shows the computational and the weight functions are calculated from Smiths
domain. Initially, three grids such as 720  220, tables23 as a function of H2O and CO2 partial pressures.
360  110, and 180  55 points are used to study the The pressure-velocity coupling is done using the
grid independence and Figure 2 shows the predictions
using three different grids. As the variation of axial
velocity with the first two grids is negligible, the third
grid is unable to capture the flame lift-off properly and Table 2. Experimental data of C2H4/Air flame.
results in the over prediction of velocity in the axial
direction. That’s why the rest of the computations Variable Jet inlet Coflow inlet
have been performed using the second grid Velocity (m/s) 44 0.29
(360  110). The experimental boundary conditions Temperature (K) 298 312
are shown in Table 1, a fully developed turbulent pro- Re 10,000 –
file was given as inlet condition for velocity, turbulent
Flow rate (g/min) 10.4 320
kinetic energy, and turbulent dissipation. The inlet

Figure 1. (a) Schematic of C2H4/air flame burner and (b) illustrative view of the mesh used (D: diameter of fuel jet).

Figure 2. Centerline plot of axial velocity using FGM model: mesh 1720  220, mesh 2 is 360  110, mesh 3 is 180  55, and symbols
are measurements (Kohler et al.14).
Busupally and De 7

SIMPLE algorithm. A second-order upwind scheme is experimental radial profiles of mean axial velocity are
used to discretize all the convective fluxes. presented. The turbulence model constants of k-e
model are modified to study the influence on jet spread-
ing. The computations have been performed using
Results and discussion
standard (Ce1 ¼ 1.44 and Ce2 ¼ 1.92), modified 1
In this section, the flow field computations with SLF (Ce1 ¼ 1.6 and Ce2 ¼ 1.92) and modified 2 (Ce1 ¼ 1.44
and FGM approaches are presented and compared with and Ce2 ¼ 2.0) constants. The SLF model over predicts
experimental measurements. The effect of turbulence- centerline velocity near the burner, this can be attribu-
chemistry interactions and radiative heat transfer on ted to the over prediction of mean temperature. Both
the flame lift off height and length are studied. Then, the SLF and FGM model predictions indicate over pre-
the soot predictions with different inception and oxida- diction of jet width, this can be ascribed to the over
tion approaches are presented and compared with the prediction of turbulent kinetic energy by the turbulence
experimental measurements; followed by a qualitative model. The excessive spreading of the jet by standard
study of the soot source terms. The effect of different k-e model is confirmed by Figure 4, the central jet
soot-turbulence approaches on evolution of soot has decelerates rapidly compared with experimental
been comprehensively studied. measurements due to prediction of excessive eddy
viscosity. This leads to under prediction of mean
axial velocity in the downstream direction.
Velocity and scalar fields Adjusting the Ce1 from 1.44 to 1.6 has improved
The contour plot of predicted mean axial velocity is the spreading rate prediction. Modifying the Ce2
shown in Figure 3 using Ce1 ¼ 1.6 and Ce2 ¼ 1.92. from 1.92 to 2.0 has resulted in further increase of
The shape and magnitude of the computes results are over prediction of spreading rate and under predic-
in good agreement with the experimental measurement. tion of mean axial velocity near the burner. The
The SLF model exhibits slight broadening of velocity inability of the model in accurately predicting the
profile after flame ignition, this has discussed in the spreading rate may be attributed to the round-jet
subsequent section. In Figure 4, the computed and anomaly, which is quite well known and has been

Figure 3. Contour of mean axial velocity (a) experimental, (b) SLF model, and (c) FGM model using non-gray radiation model using
Ce1 ¼ 1.6 and Ce2 ¼ 1.92.
8 International Journal of Spray and Combustion Dynamics 0(0)

Figure 4. Radial profile of mean axial velocity, (I) SLF model and (II) FGM model: solid lines are with gray radiation, dashed lines are
with non-gray radiation, and symbols are measurements.

reported in plenty of literatures.33 The modified k-e tend to improve along the downstream directions.
model (Ce1 ¼ 1.6 and Ce2 ¼ 1.92) is able to capture Considering the experimental uncertainties, the pre-
the axial velocity profiles reasonably well in com- dictions can be considered to be in good agreement
parison with the experimental measurements and with the experimental data. The velocity profiles are
has been used for the rest of the computations. not effect by the non-gray radiation approach.
Kohler et al. 14 have observed an under prediction Figure 5 shows the contour of calculated mean tem-
close to the centerline of the burner, which was pos- perature. The liftoff height is calculated based on the
sibly due to the usage of bulk velocity as velocity gradient of temperature along the radial direction.
inlet instead of a turbulent profile, and there results Notably, the lift-off height is observed to be 25 mm for
Busupally and De 9

Figure 5. Contour of mean temperature: (a) SLF with gray radiation model, (b) SLF with non-gray radiation model, (c) FGM with gray
radiation model, and (d) FGM with non-gray radiation model.

SLF model, while it is 24 mm for FGM model. Both the centerline maximum temperature reduced to 1850K
models are able to accurately capture the experimental from 2100K. The global maximum temperature
lift off height of 26 mm,30 indicating the ability of these shifted upstream by 35 D in the upstream direction
models to capture turbulence-chemistry interactions and 4 D away from the centerline; the magnitude
accurately. The non-gray radiation approach does not reduced from 2150 K to 1950 K.
significantly affect the liftoff height. The radial profiles For the FGM model, with the inclusion of non-gray
of mean temperature are presented in Figure 6. Figure 6 radiation approach, the flame length decreased by
indicates an over prediction in the centerline tempera- 90 mm to 490 mm and the centerline maximum tem-
ture near the burner by SLF model and improves along perature reduced to 1950 K from 2150 K. The
the downstream direction. The SLF model neglects local global maximum temperature shifted upstream by
flame extinction, leading to over prediction of flame 20 D in the upstream direction and 2 D away
temperature near the burner; this has resulted in the from the centerline; the magnitude reduced from
over prediction of velocity about the centerline near 2250 K to 2000 K. The SLF model predictions are
the burner inlet. For the present study, burning flame- closer to the experimental measurements of 410 mm.
lets are only used, this doesn’t capture the local flame The computed and experimental fuel distributions
extinction and as the flamelet model neglects the curva- are compared in Figure 7, in the region before to com-
ture effects on the flame structure, this results in a bustion. The experimental data indicate a decrease in
sudden change in scalar dissipation resulting in rapid fuel fraction at 7 mm, the computed fuel fraction with
rise in temperature near the regions of ignition. The both SLF and FGM indicate a decrease at 17 mm
over prediction in temperature also results in excessive about the centerline. The over prediction can be attri-
thermal expansion, as observed in Figure 3(b). The bute to the inability of turbulence model in capturing
FGM model has been able to accurately predict the cen- the spreading rate. The observations are similar to that
terline temperature. Both the models indicate over pre- of Kohler et al.14 The radial and axial profiles of fuel
diction of flame width as already mentioned; predicted fraction are shown in Figure 8. At X ¼ 5 mm, the mag-
temperature profiles rise slower compared with the nitude and shape are in good agreement with the
experimental data. The non-gray radiation approach experimental data. At X ¼ 15 mm, the magnitude of
does not indicate significant effect on temperature near fuel fraction is over predicted indicating under predic-
the burner inlet, the effect of radiation increases along tion of turbulent shear stress and acceleration of central
the downstream direction. For the SLF model, with the jet. The over prediction reduces along the downstream,
inclusion of non-gray radiation approach, the flame as observed at X ¼ 20 mm. A similar comment can be
length decreased by 130 mm to 430 mm and the made on the central line profile, it tends to improve in
10 International Journal of Spray and Combustion Dynamics 0(0)

Figure 6. Radial and axial profile of mean temperature, (I) SLF model and (II) FGM model: solid lines are with gray radiation, dashed
lines are with non-gray radiation, and symbols are measurements.

the downstream direction due to increase in turbulent experimental data. The prediction with FGM indicate
shear stress. The contours of predicted and experimen- a slight under prediction in terms of magnitude and a
tal OH concentration are presented in Figure 9. Similar similar comment can be made for the distribution of the
to Kohler et al.,14 the experimental OH concentration is OH concentration in the radial direction as in case of
normalized to 2500 ppm for easier comparison with SLF model. Figure 10 displays the radials profiles of OH
predicted data. concentration. The SLF model indicates slight under
The prediction with SLF model indicates a slight over predictions at x ¼ 25 mm, the predictions improve
prediction in magnitude of OH concentration and along the downstream direction and indicate slight
increase in periphery distribution compared with over prediction at x ¼ 100 mm. OH concentration at
Busupally and De 11

Figure 7. Contour of C2H4 mole fraction (a) experimental, (b) SLF model, and (c) FGM model using non-gray radiation model.

Figure 8. Radial and axial profiles of C2H4 mole fraction using non-gray radiation approach: symbols are measurements.

x ¼ 25 mm was considerable under predicted by FGM prediction by SLF and FGM model can be attributed
model and the under prediction continued in the down- to the over prediction of scalar dissipation and slower
stream direction, with a broader profile than the experi- finite rate kinetics, respectively. The location of radial
mental prediction. The disparities in OH radical maximum was accurately captured by both the models.
12 International Journal of Spray and Combustion Dynamics 0(0)

Figure 9. Contour of OH mole fraction (a) experimental, (b) SLF model, and (c) FGM model using non-gray radiation model.

Figure 10. Radial profile of OH mole fraction using non-gray radiation approach: symbols are measurements.

fraction using Moss–Brookes and Moss–Brookes–Hall


Soot predictions approaches increases by a factor of 2 for the SLF
The axial profile of soot volume fraction using different model and decrease by a factor of 8 with the FGM
inception, oxidation and turbulence-chemistry inter- model. The decrease in soot with FGM model can be
action effects are shown in Figure 11. From attributed to the models inability in accurately predict-
Figure 11(a) and (b), it can inferred that soot volume ing the concentration of two and three ring aromatics.
Busupally and De 13

Figure 11. Axial profile of soot volume fraction along centerline: solid lines are with OH equilibrium approach and solid symbols are
measurements.

It can be deduced that there is considerable increase in factor of 1000 with first aromatic-based approach.
soot volume fraction with the two-three ringed aro- The SLF and FGM models do not indicate considerable
matic inception route when compared with that of acet- effect on nucleation with acetylene-based route. The
ylene-based route for the SLF model. The Moss– nucleation with two-three ringed aromatic-based
Brookes and Moss–Brookes–Hall approaches indicate approach is relatively wider in the mixture space with
decrease in soot volume fraction by a factor of 1.5 SLF model than that of FGM model, resulting in the
with the instantaneous OH approach, indicating that substantial under prediction in soot volume fraction.
equilibrium approach is under predicting the OH con- The narrow band in mixture fraction space with FGM
centration. The location of maximum soot volume frac- model is due to the inability of the model in accurately
tion occurs downstream by 50 D when compared with estimating the mass fractions of higher order hydrocar-
experimental data, this can be ascribed to the over pre- bons. From Figure 12, it can be observed that different
diction of flame length. Figure 11(c) and (d) displays oxidation and OH concentration approaches have not
decrease in soot volume fraction with the Fenimore- significantly effected nucleation, the net effect of oxida-
Jones oxidation model and are considerable under pre- tion on nucleation is observed to be negligible.
dicted when compared with the experimental data for
the FGM approach. The decrease in soot volume frac- Soot surface growth: Surface growth increases in magni-
tion with the Fenimore-Jones model is due to higher tude and widens with two-three ringed aromatic
oxidation scaling than when compared with Lee approach for the SLF model. For the FGM model, sur-
model. The SLF model along with Moss–Brookes– face growth decreases with Moss–Brookes–Hall model,
Hall and instantaneous approach, gives a reasonable this can be attributed to the lack of active sites of higher
good estimate and the location of maximum soot order actives required for the heterogeneous reactions.
volume fraction shifts in the upstream direction. Furthermore, it can also be observed that for the FGM
model and OH instantaneous approach significantly
Soot nucleation: The soot source terms axial profiles are reduction in soot surface growth, this is due to the
presented in Figure 12. It can be inferred that soot under prediction of active sites for heterogeneous reac-
nucleation occurs at lower mixture fraction for the acet- tions. The slight decrease due to OH instantaneous
ylene-based inception route and occurs at a higher mix- approach is due to over prediction of oxidizer species.
ture fraction with the two-three ringed aromatic-based The oxidation scaling parameter is 0.015 and 1; the
route (Moss–Brookes–Hall model), for the FGM, the collision efficiency is 0.13 and 0.04 for the Lee and
maximum nucleation occurs at 0.27 in the mixture Fenimore-Jones models, respectively. The decrease in
fraction space. The magnitude also increases by a surface growth with Fenimore-Jones model is due to
14 International Journal of Spray and Combustion Dynamics 0(0)

Figure 12. Axial profile of soot source terms along centerline: (I) with SLF model and (II) with FGM model.

the high oxidation scaling parameter compared rate is considerably under-predicted by the FGM
with Lee, the high collision efficiency of Lee model model. Similar to surface growth, oxidation also
could not compensate for the low scaling parameter decreases with the first aromatic species route with
as the diameter of the soot particles are large. the FGM model.
Thus, the parameters require additional tuning the The total soot source term, the balance of soot
accurately capture the oxidation of active sites of sur- nucleation, surface growth and oxidation are presented
face growth. in Figure 13. It can be noted that majority of the sur-
face growth occurs in the vicinity of the shear layer and
Soot oxidation: From Figure 12, it can be inferred that slightly away from the center line. The net oxidation
oxidation increases in magnitude with the Fenimore- shifts upstream and increases in magnitude with the
Jones model, it also shifts to the space of high mixture Fenimore-Jones oxidation model. Figure 13(a) and (c)
fraction. The width of oxidation rate in mixture frac- indicates shift in surface growth in upstream direction,
tion space is not significantly affected. The oxidation a similar comment can be made on Figure 13(b) and
Busupally and De 15

Figure 13. Contour of total soot source term with Moss–Brookes–Hall model (a) SLF with Lee oxidation model, (b) SLF with
Fenimore-Jones oxidation model, (c) FGM with Lee oxidation, and (d) FGM with Fenimore-Jones oxidation model.

Figure 14. Axial profile of soot volume fraction along centerline using SLF and Moss–Brookes–Hall model: solid symbols are
measurements (a) Equilibrium approach and (b) Instantaneous approach.

(d). The total growth reduces significantly for the FGM early evolution of soot by 100 D, as observed by
with Fenimore-Jones oxidation model. Kohler et al.14 and Blacha et al.34 They were also
able to capture the magnitude of maximum soot
Soot-turbulence interactions: The effect of soot-turbulence volume fraction. They have attributed the early evo-
interaction on axial profile soot volume fraction is pre- lution of soot to the avoiding of soot fluctuations, a
sented in Figure 14. The net soot produced slightly similar comment can be made for the present results
increases with temperature and mixture fraction-based as we have also utilized a presumed shape PDF for
soot-turbulence interaction approach. The equilibrium modeling the turbulence chemistry interactions. Thus,
approach is able to predict the location of maximum the turbulence-chemistry interactions require add-
soot volume fraction but the magnitude has been over- itional considerations. The radial profiles of soot
predicted by a factor of 3. However, the instantan- volume fraction are presented in Figure 15. A similar
eous approach is able to predict the magnitude of soot tend as predicted in the axial profiles can be observed.
volume fraction about the centerline but the location At x ¼ 63 mm, the profiles indicate slight over predic-
predicted to be upstream by 50 D when compared tion of maximum soot volume fraction. The over pre-
with the experimental predictions. diction increases at x ¼ 163 mm and 213 mm due to
The shift in the upstream direction is due to over over predication of nucleation. At x ¼ 263 mm, there
prediction of OH concentration. The over prediction is under prediction in the radial maximum soot
of soot inception rate about the centerline results in volume fraction.
16 International Journal of Spray and Combustion Dynamics 0(0)

Figure 15. Radial profiles of soot volume fraction using SLF and Moss–Brookes–Hall model: solid symbols are measurements.

Figure 16. Contour of soot volume fraction using SLF and Moss–Brookes–Hall model (a) experimental, (b) without STI, (c) with
temperature-based STI, and (d) with mixture fraction-based STI.

It can be observed that soot formation occurs in a axial and radial locations of maximum soot volume
thinner region than the experimental measurements, fraction are not properly reproduced. The contours of
while the predictions increase considerable with the soot source terms are shown in Figure 17. From Figure
mixture fraction-based approach than when compared 17(a) and (b), the mixture fraction-based approach has
with temperature-based approach. The computed predicted as increase in nuclei concentration and coagu-
results are in reasonable agreement with the experimen- lation compared to without STI and temperature-based
tal data. The computed and experimental soot volume STI. This can be due to the mixture fraction PDF abil-
fractions are compared in Figure 16. Formation of soot ity in accurately capturing the increase in number of
occurs slightly earlier than experimental data. With the nuclei, due to increase in flame surface area. The tem-
inclusion of soot-turbulence interactions, the soot perature PDF does not capture this effect, might be due
volume fraction increases in the downstream direction to the under prediction of rms temperature, leading to
and also widens in the radial direction. But both the decrease in net soot production rate. The soot surface
Busupally and De

Figure 17. Contour of soot source terms using SLF and Moss–Brookes–Hall model (a) without STI, (b) with temperature-based STI, and (c) with mixture fraction-based STI.
17
18 International Journal of Spray and Combustion Dynamics 0(0)

growth and oxidation source terms by both the in turbulence-chemistry interactions to accurately pre-
approaches indicate increase when compared with the dict percusses and oxidizing species. The infrequent
prediction without turbulence. The increase with tem- soot events also require additional considerations to
perature-based PDF approach is less than the mixture capture the soot-turbulence interactions.
fraction PDF, as discussed earlier. The RANS-based
approaches cannot capture temporal fluctuations Acknowledgments
accurately, thus the STI require additional consider-
The authors would like to acknowledge the IITK computer
ations. Over the past couple of decades, soot modeling
center (www.iitk.ac.in/cc) for providing the resources to per-
has undergone several strides in predicting the evolu- form the computation work, data analysis and article prep-
tion of soot in combustion systems. The present models aration. The authors would like to thank Dr. Klaus Peter of
are able to accurately provide a qualitative description, DLR, Germany for providing the experimental data of C2H4/
sometimes quantitative as well, for soot evolution. The air flame.
interactions between turbulence, chemistry and soot
need to be accurately modeled for improving quantita-
Declaration of conflicting interests
tive predictions. Higher order PAH (Poly Aromatic
Hydrocarbons) may be used for modeling soot nucle- The author(s) declared no potential conflicts of interest with
respect to the research, authorship, and/or publication of this
ation. The parameters effecting oxidation of soot have
article.
to be studied and role of oxyl radical on oxidation has
to be further investigated.
Funding
The author(s) received no financial support for the research,
Conclusions authorship, and/or publication of this article.
In this work, soot formation in a C2H4/air lifted turbu-
lent diffusion flame has been studied using two turbu- References
lence-chemistry interactions (SLF and FGM)
1. Hall RJ, Smooke MD and Colket MB. Physical and
approaches. The velocity predictions are found to be
chemical aspects of combustion: A tribute to Irvine glass-
in good agreement with the experimental data. The lift- man. Amsterdams: Gordon & Breach, 1997, p.189.
off height and flame length are reasonably predicted by 2. Mehta R. Detailed modeling of soot formation and turbu-
both SLF and FGM models, while the radiative loss is lence-radiation interactions in turbulent jet flames. PhD
accurately accounted by the non-gray radiation model. dissertation, Pennsylvania, USA: Pennsylvania State
Soot production has been examined using two nucle- University, 2008.
ation approaches, two OH concentration approaches 3. Wang L. Detailed chemistry, soot, and radiation calcula-
and two oxidation approaches with SLF and FGM tions in turbulent reacting flows. 2004.
models. With SLF model, the two and three ring aro- 4. Haynes B and Wagner H. Soot formation. Prog Energy
matic approach produce a good representation of Combust Sci 1981; 7: 229–273.
nucleation, whereas surface growth is significantly 5. Bockhorn H (ed.) Soot formation in combustion. Berlin:
Springer-Verlag, 1994.
affected by OH concentration approaches, as it turns
6. Kennedy I. Models of soot formation and oxidation.
out that equilibrium approach under-predict the OH Prog Energy Combust Sci 1997; 23: 95–132.
concentration. Fenimore-Jones oxidation model 7. Frenklach M, Clary DW, Gardiner WC, et al. Detailed
reduces soot volume fraction considerably, indicating kinetic modeling of soot formation in shock-tube pyroly-
the dominance of OH radicals on soot oxidation. sis of acetylene. Symp Combust 1985; 20: 887–901.
With FGM, the soot predictions are considerably over 8. Harris SJ, Weiner AM and Blint RJ. Formation of small
predicted, due to the inability of the model in capturing aromatic molecules in a sooting ethylene flame. Combust
the mass fractions of higher hydrocarbons accurately. Flame 1988; 72: 91–109.
The mixture fraction-based STI approach indicate 9. Brookes S and Moss J. Predictions of soot and thermal
better estimate of soot-turbulence interactions when radiation properties in confined turbulent jet diffusion
compared with the temperature-based approach. The flames. Combust Flame 1999; 116: 486–503.
10. Kronenburg A, Bilger R and Kent J. Modeling soot for-
SLF model with Moss–Brookes–Hall model and
mation in turbulent methane–air jet diffusion flames.
Fenimore-Jones oxidation model and mixture fraction
Combust Flame 2000; 121: 24–40.
STI approach is able to capture the magnitude of max- 11. Pitsch H, Riesmeier E and Peters N. Unsteady flamelet
imum soot volume fraction and location and is esti- modeling of soot formation in turbulent diffusion flames.
mated to be 50 D in the upstream direction. Combust Sci Technol 2000; 158: 389–406.
Considerable improvement is found in location of max- 12. Ma G, Wen JZ, Lightstone MF, et al. Optimization of
imum soot volume fraction when compared with pre- soot modeling in turbulent nonpremixed ethylene/air jet
vious predictions. Further considerations are required flames. Combust Sci Technol 2005; 177: 1567–1602.
Busupally and De 19

13. Brookes S and Moss J. Predictions of soot and thermal 32. Ranzi E, Frassoldati A, Grana R, et al. Hierarchical and
radiation properties in confined turbulent jet diffusion comparative kinetic modeling of laminar flame speeds of
flames. Combust Flame 1999; 116: 486–503. hydrocarbon and oxygenated fuels. Prog Energy Combust
14. Köhler M, Geigle K-P, Blacha T, et al. Experimental Sci 2012; 38: 468–501.
characterization and numerical simulation of a sooting 33. Pope SB. An explanation of the round jet/plane jet anom-
lifted turbulent jet diffusion flame. Combust Flame aly. AIAA J 1978; 16: 279–281.
2012; 159: 2620–2635. 34. Blacha T, Di Domenico M, Köhler M, et al. Soot model-
15. Peters N. Turbulent combustion. United Kingdom: ing in a turbulent unconfined C2H4/air jet flame. 49th
Cambridge University Press, 2000. AIAA aerospace sciences meeting including the new hori-
16. Poinsot T and Veynante D. Theoretical and numerical zons forum and aerospace exposition. January 2011,
combustion, 2nd ed. Philadelphia, PA: Edwards, 2005. pp.1–10 ,USA.
17. Candel SM and Poinsot TJ. Flame stretch and the bal-
ance equation for the flame area. Combust Sci Technol
1990; 70: 1–15.
18. Veynante D and Vervisch L. Turbulent combustion mod- Appendix 1
eling. Prog Energy Combust Sci 2002; 28: 193–266. Notation
19. van Oijen JA and de Goey LPH. Modelling of premixed
laminar flames using flamelet-generated manifolds. a2,i emissivity weighting factor
Combust Sci Technol 2000; 161: 113–137. as characteristic strain rate
20. Siegel R and Howel Jr. Thermal radiation heat transfer. bnuc normalized radical nuclei concentration
4th ed. New York, NY: CRC press, 2002. c~ progress variable
21. Yadav R, Kushari A, Verma AK, et al. Weighted sum of ki absorption coefficient of gas
gray gas modeling for nongray radiation in combusting
kT turbulent time scale
environment using the hybrid solution methodology.
ksoot absorption coefficient of soot
Numer Heat Transf Part B Fundam 2013; 64: 174–197.
22. Taylor PB and Foster PJ. Some gray gas weighting coef- f mixture fraction
ficients for CO2-H2O-soot mixtures. Int J Heat Mass Sc reaction progress source term
Transfer 1975; 18: 1331–1332. T Temperature
23. Smith TF, Shen ZF and Friedman JN. Evaluation of Ul flame speed along the surface
coefficients for the weighted sum of gray gases model. Yk mass fraction of kth species
J Heat Transfer 1982; 104: 602–608. Ysoot soot mass fraction
24. Lee KB, Thring MW and Beér JM. On the rate of com-
bustion of soot in a laminar soot flame. Combust Flame
1962; 6: 137–145.
25. Fenimore CP and Jones GW. Oxidation of soot by Greek symbols
hydroxyl radicals. J Phys Chem 1967; 71: 593–597.
 thermal diffusivity of unburnt reactants
26. Baulch DL, Cobos CJ, Cox RA, et al. Evaluated kinetic
data for combustion modelling. J Phys Chem Ref Data
i absorption coefficient of ith gray gas
1992; 21: 411. 0 density of fresh gas
27. Westbrook CK and Dryer FL. Chemical kinetic model- soot turbulent Prandtl number of soot
ing of hydrocarbon combustion. Prog Energy Combust transport
Sci 1984; 10: 1–57.  representative scalar
28. Westeenberg AA. Kinetics of NO and CO in lean, pre- scalar dissipation rate
mixed hydrocarbon-air flames. Combust Sci Technol st scalar dissipation rate at f ¼ fst
1971; 4: 59–64.
29. Köhler M, Boxx I, Geigle KP, et al. Simultaneous planar
measurements of soot structure and velocity fields in a
turbulent lifted jet flame at 3 kHz. Appl Phys B 2011; 103: Subscripts
271–279.
30. Köhler M, Geigle KP, Meier W, et al. Sooting turbulent b burnt reactants
jet flame: characterization and quantitative soot measure- eq chemical equilibrium
ments. Appl Phys B 2011; 104: 409–425. u unburnt reactants
31. ANSYS Inc, Ansys fluent 15.0 user’s guide. Csnosburg,
PA, USA.

You might also like