Você está na página 1de 13

International Journal of Heat and Mass Transfer 74 (2014) 143155

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Effect of air ow distribution on soot formation and radiative heat


transfer in a model liquid fuel spray combustor ring kerosene
Prakash Ghose a, Jitendra Patra a, Amitava Datta a,, Achintya Mukhopadhyay b
a
b

Department of Power Engineering, Jadavpur University, Salt Lake Campus, Kolkata 700 098, India
Department of Mechanical Engineering, Jadavpur University, Kolkata 700 032, India

a r t i c l e

i n f o

Article history:
Received 12 July 2013
Received in revised form 6 February 2014
Accepted 2 March 2014
Available online 1 April 2014
Keywords:
Spray combustion
Soot
Radiation
Discrete droplet model
Air ow distribution

a b s t r a c t
In the present paper, a numerical model has been developed for spray combustion in a model gas turbine
combustor admitting air as co-axial primary and secondary streams. The model incorporates soot formation and radiative heat transfer and has been validated with experiments conducted on a combustor of
identical geometry. The paper investigates the effect of air ow distribution, between primary and secondary streams, on ame structure, soot formation and radiative heat transfer in the combustor ring
kerosene as fuel. Turbulence is modeled using realizable ke model and radiation is modeled using discrete ordinate method with weighted sum of gray gases model. The combustion is modeled using equilibrium presumed probability density model. The results show that an increase in the proportion of
primary air ow, from 30% to 50% of the total air, results in a more compact ame with lower soot production and a better pattern factor at the combustor exit. However, the corresponding reduction in secondary air ow rate increases the combustor wall temperature. The decrease in soot in ame at higher
primary air fraction reduces the incident radiative heat ux on the injector body while, the injector surface temperature remains almost unaffected due to increased convective heat transfer rate from the gas.
2014 Elsevier Ltd. All rights reserved.

1. Introduction
Numerical modeling of liquid fuel spray combustion is widely
used as a predictive tool for the performance analysis of gas turbine combustors and liquid fuel furnaces. Spray combustion is a
complex phenomenon consisting of various physical and chemical
processes, like atomization of liquid jet and movement of droplets
in a gaseous eld, vaporization of droplets, turbulent transport and
mixing, chemical reaction, thermal radiation and pollutant formation. The prediction of the entire process depends on the accuracy
of the component models, which have been employed in the whole
scheme. Some of the models are quite well established in the literature. For example, the uid ow is commonly solved with RANS
based models for their computational economy, though it has been
found that LES [1] and DNS [2] based models can provide valuable
insight of ow in the combustors. However, the latter models are
very expensive particularly in three dimensional geometries. In
the RANS based models, the turbulence quantities are usually
solved using two equation models with eddy viscosity concept.
Different forms of the ke models, like standard, RNG and
realizable ke models are commonly employed in the literature
Corresponding author. Tel.: +91 33 23355813; fax: +91 33 23357254.
E-mail address: amdatta_ju@yahoo.com (A. Datta).
http://dx.doi.org/10.1016/j.ijheatmasstransfer.2014.03.001
0017-9310/ 2014 Elsevier Ltd. All rights reserved.

[35]. Hsiao and Mongia [6] and Joung and Huh [7] used standard
and realizable ke models to predict swirling ows in conned
geometries and found reasonable prediction of ow parameters.
Karim et al. [8] showed that standard ke model is over-diffusive
in highly swirling ows in comparison with realizable ke model.
In spray combustion calculation, a suitable model is also required
to predict the initial spray characteristics following breakup of
the liquid fuel jet. Thereafter, the interactions between the continuous and dispersed phases are often captured using the discrete
droplet model (DDM) in the EulerianLagrangian formulation [9
11]. In the discrete droplet model, the liquid spray is considered
to consist of a nite number of droplet classes, whose trajectories
in the gas phase are tracked using suitable governing equations.
The mass, momentum and energy exchanges between the two
phases are computed as source terms and accounted in the gas
phase governing equations.
Soot formation in combustion is important in liquid fuel spray
ames. Soot particles present in the ames result in a highly luminous radiation and thereby inuence the heat transfer phenomenon from the ame. As a consequence, soot in ame augments
the wall and burner heating considerably. It is therefore important
to precisely model the soot formation process in spray ames and
study its inuence on radiative heat transfer from the ame.
Kerosene (or jet fuel) is widely used as fuel in aero gas turbine

144

P. Ghose et al. / International Journal of Heat and Mass Transfer 74 (2014) 143155

Nomenclature
A
ak
Cd
Cdrag
Cf
C fs
CL
Cp
Cl
D
d
dL
do
dpsoot
DHv
h
hD
I
k
kw
M
m
_
m
N
NA
Oh
P
P(n)
DP
p
Q
q
R
Re
S
T
t
U
u

surface area, m2
weighting factor
coefcient of discharge
drag coefcient
vapor concentration in the continuous phase
vapor concentration at the droplet surface
ligament constant
specic heat, J/kg K
a variable, function of mean strain rate
diameter of the combustor, m
diameter, m
ligament diameter, m
mean droplet diameter, m
mean diameter of soot particle, m
latent heat of vaporization, J/kg
convective heat transfer coefcient
mass transfer coefcient
radiation intensity, W/m2 sr
turbulent kinetic energy, m2/s2
wave number
soot mass concentration, kg/m3
mass, kg
mass ow rate, kg/s
particle number density, 1/m3
Avogadro number 6.022045e+26 kmol/l
Ohnesorge number
pressure, N/m2
probability density function
pressure differential, N/m2
total partial pressure, N/m2
ratio between gas and liquid density
radiative heat ux, W/m2
universal gas constant
Reynolds number
source
temperature, K
liquid sheet thickness, m
resultant (total) velocity of fuel jet, m/s
velocity, m/s

combustors, where a high overall airfuel ratio is maintained in order to keep the exit gas temperature within the allowable limit for
the turbine blade material. The total air is distributed in different
zones, so that a stable ame can be established on the burner.
The air ow distribution inuences the stoichiometry in the ame
zone and affects the soot formation there. The cooling of the combustor wall and the temperature uniformity of the exit gas also depend on the air ow distribution in the combustor.
Soot formation in hydrocarbon combustion is a very complex
process, which initiates with the formation of precursor molecules
and completes through the growth of poly aromatic hydrocarbons
[12,13]. Detailed formulations using elementary reactions for the
gas phase and soot [14,15] are often found to be unfeasible in
the real combustor congurations (e.g. in gas turbine combustor)
because of their complexities. Therefore, different semi-empirical
models had been proposed by Kennedy et al. [16], Leung et al.
[17], Moss et al. [18], and Brookes and Moss [19] for the prediction
of soot in hydrocarbon ames, like methane or ethylene. The models compute the soot nucleation and surface growth rates based on
the concentration of precursor species, which is commonly considered as acetylene in these works. Oxidation models, proposed by
Lee et al. [20], Nagle and Strickland-Constable [21], and Fennimore
and Jones [22] were also adopted in the soot models. Wen et al.

We
wi
X
z

Weber number
quadrature weight
mole fraction
path length

Greek symbols
C
diffusivity
q
density
/
scalar variable
l
dynamic viscosity
lt
dynamic eddy viscosity
r
Prandtl/Schmidt number
rs
surface tension of the liquid
e
rate of dissipation of turbulent K.E.
h
spray cone half angle
q
density
xr
growth rate of sinuous wave
m
kinematic viscosity
n
mixture fraction
j
absorption coefcient
jk
gray gas absorption coefcient
Subscript
b
crit
d
eff
f
f
g
i
inj
j
k
l
or
p
rad
/

black body
critical
droplet
effective
liquid
fuel
gas
ith coordinate direction
injection
jth coordinate direction
kth coordinate direction
liquid
orice
particle
radiation
scalar variable

[23] modeled soot formation in a turbulent jet diffusion ame of


kerosene vapor and air. They considered two different nucleation
models of soot, viz. (i) acetylene nucleation model considering
acetylene as the precursor and (ii) PAH nucleation mode taking
two and three ring aromatics as precursor. However, surface
growth is modeled using acetylene concentration and the number
of active sites on the soot particles. It was found that the PAH
nucleation model contributes signicant improvement in the model prediction in comparison to the experimental data. On the other
hand, Moss and Aksit [24] applied the soot model, proposed by
Brookes and Moss (for a methane non-premixed ame), in the laminar non-premixed ame of a surrogate kerosene fuel. They found
that adjustments in the model parameters, from the values in
methaneair ame, are necessary to satisfactorily reproduce the
experimental measurements of soot under the change of fuel.
Accordingly, they proposed two alternate models. In one of them,
acetylene is considered as the precursor in nucleation and surface
growth and the model parameters are changed to match the experimental results. In the other case, acetylene is considered as the
precursor of soot nucleation only, while the precursor for surface
growth is taken as acetylene and benzene. New model constants
are evaluated for this case as well. Both the models are found to
predict the soot concentration nearly equally, though the latter

145

P. Ghose et al. / International Journal of Heat and Mass Transfer 74 (2014) 143155

one comes closer to the measurement at some locations. These


works show that uncertainties still exist in considering the soot
model with kerosene fuels.
Soot particles in combustion signicantly contribute to radiation heat transfer due to their high emissivity [25,26]. Byun and
Baek [27] investigated numerically the combustion of liquid kerosene in a rocket engine considering soot formation and radiation.
The radiation model considered the contributions of non-gray
gases using weighted sum of gray gas model and gray soot particles. They reported that radiation from the ame makes the high
temperature zone smaller and affects the wall temperature considerably. Tesse et al. [28] reported that in a sooty turbulent non-premixed ame, in addition to soot particles, gaseous species such as
CO2 and H2O play important role in global radiative heat loss.
In the present work, we have presented numerical results on
kerosene (C12H23) spray combustion in a model combustion chamber (Fig. 1) and predict the soot formation and radiation heat transfer from the ame. The combustion chamber has two co-axial air
entries, considered as primary and secondary air streams. The primary air is admitted through a constant vane angle swirler, at the
center of which a pressure swirl atomizer is tted to inject the fuel.
Therefore, the primary air ow regulates the stoichiometry of fuel
and oxidizer in the ame zone while, the secondary air contributes
in cooling the combustor wall and diluting the product gas mixture. The two-equation soot formation model following Brookes
and Moss [19] has been adopted along with the soot oxidation
model due to Fennimore and Jones [22]. The model parameters
have been suitably chosen by matching the predicted soot results
with the experimental data for kerosene ame. Radiation heat
transfer has been modeled using the discrete ordinate (DO) model
[29]. Experiments have been conducted to nd out the initial spray
conditions, like spray cone angle and injection pressure differential, which are required as input to the model. Further experiments
have also been conducted in a combustor of the same geometry to
get results for model validation. Finally, soot distribution in the
combustor and distributions of incident radiation heat ux and
the temperature on the combustor wall and fuel injector are compared for different primary to secondary air ow splits using
numerical predictions.
2. Model
A numerical model of the spray combustion process in the combustion chamber is developed in three dimensions taking into account the geometry of the swirler. The combustion chamber is
0.1 m in diameter (D) and 0.5 m in length from the plane of the
swirler (Fig. 1). The inlet plane to the computation domain is

considered at 40 mm upstream to the swirler plane, both in the


primary and secondary air streams. The Favre-averaged governing
equations in the gas phase are solved and the turbulence parameters are computed using realizable ke model [30]. Favre averaging
helps to avoid extra terms in the governing equations due to the
uctuating and variable density in the ow. The injected spray
with a particular cone angle from the pressure swirl atomizer is
considered to break up following the linearized instability sheet
atomization model (LISA) of Schmidt et al. [31]. This model assumes little knowledge of the atomizer hydrodynamics and helps
to nd out an average droplet size following break up using the
available observations of external spray and atomizer characteristics. The LISA model is simple, requires less computational cost and
reduces the empirical constants and geometrical data required for
predicting lm formation and sheet break up following spray from
a pressure swirl atomizer. This model has been subsequently used
by various researchers for computing dispersion and combustion
of spray from swirl type atomizers [32,33]. The motion of the evaporating droplets in the continuous gas phase is tracked stochastically using an EulerianLagrangian approach and the inter-phase
transport terms are suitably accounted for use in the gas phase
conservation equations. The gas phase combustion reactions are
assumed to be innitely fast in comparison to the transport processes so that local chemical equilibrium among the species is
reached. An assumed b-pdf is considered to statistically correlate
the average concentrations of the species with the mean mixture
fraction computed from the model. The entire computation has
been conducted using the Ansys Fluent commercial software
(version 13.0).
2.1. Gas phase ow models
The gas phase conservation equations are solved in the Eulerian
frame of reference taking into consideration the source terms for
the inter-phase transport from the liquid phase. The general form
~ in
of the Favre-averaged conservation equation for variables (/)
the gas phase can be written as.

~
@
~ @ q
~ @ C/;eff @ /
 /
u
~ j /
q
@xj
@t
@xj
@xj

!
S g Sl

The terms on the left hand side of the above equation represent
the temporal variation and advective transport of the variables,
respectively, while the terms on the right hand side are the diffusive transport, gas phase source (Sg) and inter-phase source (Sl).
~ for the different gas phase conservation equations
The variables /
along with the respective gas phase and inter-phase source terms

Fig. 1. Physical geometry of the combustor under study.

146

P. Ghose et al. / International Journal of Heat and Mass Transfer 74 (2014) 143155

Table 1
Gas and liquid phase source terms for the gas phase governing equations.
Conservation of

~
/

Sg

Mass

1
~i
u

~
h
~
k
~e

S_ rad
 ~e
Gk  q

Momentum
Energy
Turbulent kinetic energy
Dissipation rate of turbulent kinetic energy
Mixture fraction
Mixture fraction variance


~ 2l
k
 @x@ i p 23 q
3 eff

C g1 lt

The liquid fuel is injected into the combustor through a pressure swirl atomizer tted at the center of the swirler. The mass
_ f ), injector orice diameter (dor), injection
ow rate of the liquid (m
pressure differential (Dpinj) and spray cone angle (2h) have been given as input to the model. These quantities are obtained from separate experiments performed in a test rig for spray with a pressure
swirl atomizer. The transition from the internal hydrodynamics
within the atomizer to the fully developed spray occurs in three
steps: sheet formation at the atomizer exit, sheet breakup and
atomization. These steps have been modeled using the linearized
instability sheet atomization (LISA) model [31], considering primary atomization following breakup to form the spray. Secondary
atomization, coalescence and collision of the droplets in the spray
have been neglected.
In the model, the injection velocity of the liquid at the atomizer
exit is calculated using the injection pressure differential and the
coefcient of discharge (Cd) of the nozzle in use as,

qf

A mass balance of the liquid fuel at the atomizer exit evaluates


the thickness of the liquid sheet (t) emanating from the nozzle as,

_ f pqf U inj cos htdor  t


m

@x@ j

leff

me

~i
@u
@xj

~
@u

@xij

S_
~ i S_
S_ M u

i

C g2 ~~e
k

~
n
~
n002

2.2. Spray atomization model

s
2Dpinj

~j
@u
@xj

p
2
 S~
C1 q
e  C 2 q ~ ~ep~, where S = 2Sij Sij and Sij = @~u 1 @~u 

are listed in Table 1. The term C/;eff leff =r/;eff in Eq. (1) charac~ in turbulent ow.
terizes the effective diffusivity for the variable /
The effective dynamic viscosity (leff = l + lt) accounts the molecular diffusivity (l) and the eddy diffusivity (lt), where the latter
 k~~e2 . The term r/,eff represents the effective
term is given by lt C l q
Prandtl/Schmidt number for the uid.
~ ~e) are
In the present model, the turbulent ow variables (k;
solved using realizable ke model [30]. In this model, a variable
Cl is considered in the expression of eddy viscosity to ensure realizability even at large mean strain rate avoiding negative normal
stress and satisfying Schwarzs inquality for shear stresses. Moreover, the source terms of the dissipation (~e) equation are different
in realizable ke model, compared to the standard model, in order
to make the equation more robust. The energy equation is solved to
nd out the average enthalpy at each grid point. Radiative heat exchange occurs as the gas phase source term (Srad) in the energy
equation.

U inj C d

Sl

From the knowledge of the liquid sheet thickness and sheet velocity, the breakup of the liquid sheet leading to the formation of droplets is modeled. The liquid sheet rst breaks up into ligaments as a
result of growth of the instabilities developing on the liquid surface.
The ligaments, in turn, further break up into drops. The linear stability analysis of Senecal et al. [34] is used for investigating the instability of the liquid sheet. Since the liquid sheet thickness is much

@~
n002
@xj

2

j
i
@xj @xi

~ S_
S_ E h
~S_
k
~
eS_
~
nS_
~002 S_
n

 ~002

qn

smaller than the mean radius of the sheet, for the purpose of stability analysis, the curvature of the liquid sheet is neglected and the
results for planar liquid sheet moving with a prescribed velocity Uinj
in a stationary gas phase is used. The sinuous mode of instability
dominates over the varicose mode at low velocities and low gasto-liquid density ratios while the two modes become indistinguishable at large velocities. Hence instability of only sinuous mode is
used for the instability analysis. Following Senecal et al. [34], the
dispersion relation is given by


2
2ml kw tanh kw 2t
 t
tanh kw 2 Q
q





4
2
2
2
3
4m2l kw tanh kw 2t  Q 2 U 2 kw  tanh kw 2t Q QU 2 kw rkw =ql
 t

tanh kw 2 Q

xr 

In deriving the above equation, second order terms in viscosity have


been neglected as they are very small in value. For Weber number
Weg > 27
, a condition satised by most modern pressure-swirl
16
atomizers, the most unstable waves are short waves. For short
waves, the ligament diameter is assumed to be proportional to
the most unstable wavelength that breaks up the sheet. Thus the
ligament diameter is given by

dL

2p C L
kwcrit

In the above equation, kwcrit is the most unstable wavenumber giving


the highest growth rate as obtained from the dispersion relation. CL
is a ligament constant, considered as 0.5. The breakup of the ligament into drops is obtained from Webers analysis for capillary jets.
The resulting drop size is given by

do 1:88dL 1 3Oh

1=6

In the above equation, Oh is the Ohnesorge number dened as


Oh = ll/(qlrsdL)1/2.
The initial size distribution of the droplets in the spray following atomization is considered using the RosinRammler distribution function. The droplet diameter following breakup (do) is
considered as the size parameter of the distribution function, while
the dispersion parameter is taken as 3.5 [35].
2.3. Liquid phase ow model
In order to compute the inter-phase source terms over the life
time of the droplets, the spray is considered to comprise of a nite
number of droplet classes distributed over an initial dispersion angle. The velocity, mass and temperature histories of each of the
droplet classes are obtained along their trajectories using the
respective conservation equations in a Lagrangian frame.
The trajectory of a droplet of the kth class is computed by evaluating the velocity and position along its motion. The velocity of

P. Ghose et al. / International Journal of Heat and Mass Transfer 74 (2014) 143155

the droplet is found out from the conservation of momentum


equation considering only inertia and drag forces to be signicant.
The equation can then be written as,

md k

dupi k p
 dk2 jui  upi kjui  upi kC drag
q
dt
8

dmp
2
pdp hD C fs  C f
dt

where, C fs and Cf are the mass fractions of the fuel vapor on the
droplet surface and in the surrounding gas.
In order to nd out the variation of temperature of the droplet
an energy balance across the droplet surface is considered as,

mp cpp

dT p
dmp
hAp T  T p 
DH v
dt
dt

with source terms listed in Table 1. The density-weighted average


values of species concentration, temperature and density are obe
tained assuming local beta-probability density function Pn
in
turbulent reacting ow as,

where, Cdrag is the drag coefcient on the droplet, which is evaluated following the spherical drag law [36]. The effect of gas phase
turbulence on the droplet dispersion is simulated using a stochastic
approach. Instantaneous gas phase velocity (ui) around the droplet
is obtained in the above equation by computing the uctuating
velocity following a discrete random walk model. The position of
the droplet is obtained by integrating the velocity over short time
range. If any droplet, in course of its motion, strikes the combustor
wall, it is assumed to reect from the wall following elastic
collision.
Evaporation of the liquid from the surface of the droplets takes
place considering the vapor pressure on the droplet surface to be
equal to the saturation pressure at the droplet temperature. A
piecewise linear variation of the saturation pressure for the liquid
fuel with temperature is considered for evaluation. The mass transfer coefcient (hD) is calculated from the Sherwood number correlation of Ranz and Marshall [37]. The change in droplet mass can
therefore be accounted as,

The heat transfer coefcient (h) is found out from the Nusselt number correlation of Ranz and Marshall [37] and the radiation exchange with the gas phase is neglected.
The liquid phase conservation equations are solved for each of
the droplet classes along their trajectory till the class gets evaporated. The inter-phase source terms for mass, momentum and energy are accordingly computed at different grid points depending
upon the positions of the droplets and are used in the gas phase
equations.
2.4. Equilibrium presumed probability density function model
The equilibrium presumed probability density function model
is used to obtain the mean values of species concentration, temperature and density in the gas phase from the mixture fraction values
[38]. Mixture fraction is dened as the elemental mass fraction
originated from the fuel, which enters the gas phase within the
computational domain due to evaporation of the liquid droplets.
The combustion reactions take place in the gas phase and the
kinetics of reactions are assumed to be fast enough to reach equilibrium provided the local fuel air ratio remains within a rich ammability limit (Equivalence ratio at RFL = 1.5). Beyond this limit,
the mixing of species in the gas phase is only considered without
any chemical reaction. The instantaneous species concentration,
temperature and density in the gas phase have been computed
as functions of mixture fraction considering chemical equilibrium
and heat transfer from the system. Sixteen chemical species (O2,
N2, C12H23, CO2, CO, H2O, H2, OH, H, O, HO2, H2O2, HCO, CHO,
HONO, HCOOH) have been considered in the equilibrium product
mixture following chemical reaction.
The Favre averaged mixture fraction and its variance are obtained within the combustor by solving their respective equations

147

~
/

e
/n Pndn

10

~
are constructed in terms of mean mixture fracLocal beta pdfs, Pn,
tion and its variance as,

na1 1  nb1
e
Pn
R 1 a1
n 1  nb1 dn
0

11

h~ ~
i
n
n n1
n a~n
where, a ~
 1 and b 1  ~
~
n002
2.5. Soot model
Soot formation chemistry in hydrocarbon ame is much slower
compared to the combustion reactions and is separately modeled.
We have adopted the BrookesMoss model [19] of soot formation
in a turbulent ame. The model solves two soot quantities, soot
number density (N) and soot mass concentration (M) within the
combustion chamber. These two quantities are evaluated considering separate model expressions for nucleation, coagulation, surface
growth and oxidation. While the nucleation and coagulation inuence the number density of the soot particles, the soot mass concentration depends on nucleation, surface growth and oxidation.
Acetylene is considered as the precursor species both for nucleation and surface growth. The equilibrium chemistry model
adopted in the gas phase does not solve acetylene as a product species. Moss and Aksit [24] presented a variation of acetylene concentration against mixture fraction as a amelet relation for
kerosene surrogate fuel. We have used these data to form a correlation which computes the acetylene concentration from the mixture fraction value in a kerosene ame. A mixture fraction based
probability density function approach is adopted to compute the
soot formation accounting the turbulence-chemistry interaction.
The soot model of Brookes and Moss calculates the instantaneous production rate of soot particles number density (N) as,

1=2
X C 2 H2 P
dN
Ta
24 RT
1=2
exp 
 Cb
C a NA
dpsoot N2
dt
T
qsoot NA
RT

12

The instantaneous production rate of soot mass per unit volume


(M) is expressed as,

dM
X C 2 H2 P
Ta
exp 
C a Mp
dt
RT
T

"

2=3 #n
Tc
X C 2 H2 P
6M
exp 
pN1=3
Cc
RT
T
qsoot

2=3
X OH P
6M
T pN1=3
 C oxid C x gcoll
RT
qsoot

13

The model constants Ca and Cc for nucleation and surface


growth have been given in the work of Brookes and Moss as
54 s1 and 11700 kg m kmol1 s1, respectively, for methane
ame. The model constants Ca and Cc are modied, considering
the fact that these two kinetic dominated processes will have different rates in a kerosene ame from the rates in a methane ame.
The modied constants are xed by comparing the predicted results with the experiments conducted in a turbulent non-premixed
ame of kerosene vapor and air [23].

148

P. Ghose et al. / International Journal of Heat and Mass Transfer 74 (2014) 143155

2.6. Radiation model


The radiation calculation within the combustor has been performed assuming the medium to consist of participating gases
and soot in which scattering is neglected (considering the particles
to be extremely ne and dispersed). The solution for radiative exchange has been done using discrete ordinate model [26,39,40] and
the non-gray behavior of the gases has been accounted by
weighted sum gray gas model (wsggm) [39].
In the discrete ordinate method, a discrete representation of the
directional variation of the radiative intensity is considered. The
radiative transfer equation is solved for a set of n discrete directions ^si ; i 1; 2; 3 . . . n spanning over the total range of solid angle 4p. The equation for a particular direction is given as,

^si :rIi jIb  Ii

14

where, Ii is the radiation intensity in the ith direction, j is the


absorption coefcient and Ib is the blackbody radiation intensity.
The absorption coefcient has both gas phase and particle (soot)
phase contributions expressed as,

j jgas jsoot

15

The gas phase contribution of the absorption coefcient (jgas)


has been modeled using weighted sum of gray gas model (wsggm)
with the constant gray gas absorption coefcients (jk) for the participating gases (k = 1K) along with suitable weighting factors (ak)
as,

jgas 

h
i
P
ln 1  Kk0 ak 1  ejk pz
z

16

where, p is the total partial pressure of all the absorbing gases and z
is the path length, which is considered as the mean beam length
corresponding to the combustor geometry [39]. The model considers the contributions of carbon dioxide and water vapor in the gas
phase for the absorption coefcient. The weighting factors are taken
as temperature dependent polynomial functions [41] as,

ak

bk;j T j

17

The absorption coefcient contributed by soot (jsoot) is found


from the equation,

jsoot 1232:4qsoot b1 4:8  104 T  2000c

18

The radiative source term (Srad) in the energy equation is computed as the divergence of the radiative heat ux vector as,

rad j 4rT 4 
Srad r:q

N
X
wi Ii

!
19

i1

where, wi is the quadrature weight associated with the direction i in


the discrete ordinate method [39].
3. Numerical model, operating parameters and boundary
conditions
The governing equations have been solved using the pressure
based, steady solver in Ansys Fluent 13.0. The pressure velocity
coupling in the gas phase has been accounted with the SIMPLE
algorithm. The terms in the governing equations have been discretized by the power law scheme while, the radiation model equation
is discretized using second order upwinding scheme. The Eulerian
and Lagrangian phase calculations have been performed in an iterative way with 200 continuous phase iterations per particulate
phase iteration. Twenty discrete classes of droplets have been injected initially within a dispersion angle of 6. The discrete ordinate model of radiative transfer considers 5 divisions in the polar

direction and 5 divisions in the azimuthal direction of every octant


around the control volume for radiation computation. Thus a total
of 200 angular directions over the solid angle 4p have been taken
in the model.
_ air into the combustor is set at
The total mass ow rate of air m
0.04 kg/s, which is entering the combustor at 300 K. The Reynolds
_ air =pDl) based on the air ow inlet conditions
number (Re 4m
and combustor diameter is 26,300. The air ow is split between
the primary and secondary streams at the entry to the combustor.
Three different air ow splits, with primary:secondary as 30:70,
40:60 and 50:50, have been considered in the analysis. A constant
angle (60) vane swirler is tted at the entry of the primary stream
to the combustor (Fig. 1). However, the inlet plane of computation
is considered at 40 mm upstream to the swirler plane. The fuel, at
300 K temperature, is injected through a 0.25 mm diameter orice
of the pressure swirl atomizer at a rate of 0.00036 kg/s in all the
cases. The spray cone angle and the injection pressure differential
corresponding to the liquid ow rate through the atomizer are
experimentally obtained to feed to the model.
The plug ow velocity boundary condition is considered at the
inlet planes of both primary and secondary streams, depending on
their respective ow rates. A 4% turbulent kinetic energy and a turbulent length scale of 0.007 m are set at the inlet boundaries.
Atmospheric pressure boundary condition is considered at the outlet plane of the combustor.
The inner surface of the peripheral wall of the combustor is considered as opaque and diffuse with an internal emissivity dened
for stainless steel [39]. A mixed heat transfer boundary condition,
considering both radiation and convection from the surface, is applied on the outer peripheral wall of the combustor. The wall body
is of stainless steel with 5 mm thickness. The emissivity of the outer wall surface is taken as equal to the inner surface and a convective heat transfer coefcient is assumed, which results in a balance
of energy owing out from the combustor. However, the injector
and the other solid walls are considered as adiabatic.

4. Experimental
Sample experiments have been performed in test rigs to generate the spray data required for the model and also for the validation of combustion model predictions. The spray data are
obtained by conducting experiments in a spray test rig. The details
of the spray test rig are given in a separate publication [42]. In the
test rig, kerosene is injected through the pressure swirl atomizer
under consideration to generate sprays in open atmosphere. The
injection pressure is measured by tting a calibrated pressure
gauge just before the nozzle and the corresponding volume ow
rate is measured by collecting the liquid in a measuring ask over
denite time. The coefcient of discharge of the nozzle is found out
from the measured values of volume ow rate and injection pressure differential. The spray cone angles are measured by obtaining
the spray images using a light sheet and a camera.
In another experiment conducted in the model combustor, having the same geometry and shape as used in the computation, a
kerosene spray ame is established with the same pressure swirl
atomizer as used in the spray test rig. The primary and secondary
air ow supplies to the combustor and the ow rates are measured
by orice meters. The fuel ow rate is measured using a calibrated
rotameter. In order to measure the combustor wall temperature,
eight thermocouples (K-type) are tted with silicon heat sink compound close to the inner wall surface at an interval of 50 mm in the
axial direction. A traversing thermocouple near the exit plane measures the variation in exit gas temperature from the centerline to
the wall. Radiation correction has been done for these thermocouples to get the measured gas temperatures at different radial loca-

149

P. Ghose et al. / International Journal of Heat and Mass Transfer 74 (2014) 143155

tions. The temperature variations along the wall and in the exit gas
are used for the validation of the model predictions.
5. Results and discussion
5.1. Validation of soot model
It has been pointed out in the introduction that most of the soot
models have been formulated for predictions in gaseous hydrocarbon ames, like methane and ethylene ames. The works of Moss
and Aksit [24] and Wen et al. [23] showed that it is possible to use
the existing soot models, developed for gaseous fuels, in kerosene
ames by suitably adjusting the empirical model parameters.
Accordingly, we have used the soot model proposed by Brookes
and Moss [19] for the kerosene spray ame along with the soot oxidation model of Fennimore and Jones [22]. The model parameters
are adjusted to match the experimental data of kerosene non-premixed ame. In order to x the model parameters we have simulated the turbulent, non-premixed ame of Wen et al. [23], with
kerosene vapor as fuel in a co-axial burner. However, the other
models like the realizable ke model, discrete ordinate radiation
model and the equilibrium presumed probability density function
model, as described earlier have all been incorporated in the
computation.
Fig. 2(a) shows the predicted centerline distribution of soot volume fraction above the burner for the experimental conditions
considered by Wen et al. The experimental data points have also
been plotted in the gure. It is shown in the gure that too little

soot is predicted in the combustion zone with the values of model


constants proposed by Brookes and Moss for methane ame. In order to predict the soot according to the experiment, we have varied
the model constants Ca and Cc and compared the results with the
experiments. With a Ca = 324 s1 and Cc = 70200 kg m kmol1 s1
the predicted results agree reasonably well with the experiments,
particularly in the upstream region, closer to the burner. We have,
therefore, considered the Brookes and Moss soot model with acetylene as precursor species and modied empirical constants in the
prediction of kerosene spray ame.
Fig. 2(b) shows the variation of predicted soot volume fraction
using the modied empirical constants as a function of radial distance at a height of 100 mm from the inlet plane. The corresponding experimental data measured by Wen et al. [23] is also plotted.
The results indicate qualitative agreement showing the peak volume fraction away from the axis. The predicted peak volume fraction value comes quite close to the experimental peak. In fact the
agreement between the peak volume fractions in the distribution
is much better than those in the work of Wen et al. Thus the model
of soot for the kerosene ame justies its applicability.
5.2. Validation of numerical model
Subsequent to the selection of the soot model parameters, we
have computed the spray combustion of kerosene fuel in the
model combustor considering the soot formation in ame. Experiments have been performed in the model combustor under the
same ow and spray conditions as in computation and with one

1.0E-04

Soot Volume Fraction

1.0E-05
1.0E-06
1.0E-07
1.0E-08
Experimental, Wen et al.

1.0E-09

Precursor corr. Of methane

1.0E-10

Precursor corr. Of kerosene with modifed const.

1.0E-11
1.0E-12
50

100

150

200

250

300

350

400

450

(a)

Axial Distance (mm)

(a)

0.05

3.0E-06

0.04

Experimental, Wen et al., 2003

0.03

Precursor corr. Kerosene with modified const.

2.0E-06
1.5E-06
1.0E-06

Radial distance (m)

Soot Volume Fraction

2.5E-06

No. of elements = 254720

0.02

No. of elements = 324394

0.01

No. of elements = 462875

0
-0.01
-0.02
-0.03
-0.04

5.0E-07

-0.05

0.0E+00
0

10

15

20

25

30

Radial distance (mm)

(b)
Fig. 2. Comparison of predicted and measured soot volume fractions along the (a)
centerline and (b) radial direction at 100 mm above inlet of Wen et al. ame.

200

400

600

800

1000

Temperature (K)

(b)
Fig. 3. Comparison of predicted values of (a) wall temperature along the length of
the combustor, (b) exit gas temperature across the radial direction for three
different grid congurations.

150

P. Ghose et al. / International Journal of Heat and Mass Transfer 74 (2014) 143155

air ow split (50:50 between primary and secondary). Validation


of the model predictions has been made by comparing the predicted temperature distributions along the wall and across the
exit plane against the experimental values. A grid independence
test is rst done by rening the grid over the computational domain and by observing the variation in predicted temperatures
along the wall (Fig. 3a) and over the exit plane (Fig. 3b). An
unstructured, quadrilateral mesh conguration is chosen with
254,720, 324,394 and 462,875 elements in the domain. It is found
from the gures that with the rst renement of the grid (from
254,720 to 324,394 elements) the maximum changes in the wall
temperature and the exit gas temperature are found to be 5.6%
and 8.2%, respectively. While with further renement (from
324,394 to 462,875 elements), the above two peak variations
come down to 2.2% and 1.2%, respectively. Considering these,
we have nally chosen the grid conguration with 324,394 elements in the mesh for further computation.
The variations in temperature along the combustor wall
(Fig. 4a) and in the exit gas (Fig. 4b) with the chosen grid conguration agree quite well with the measured values. A discrepancy in
the temperature prediction is noticed on the wall around the ame
zone, and may be attributed to the variation in radiative heat
transfer from the ame. The predicted temperature distribution
in the exhaust gas agrees very well with the experiments over
the entire cross-section of the combustor. Overall considering all
the compared variations of temperature, it can be concluded that
the adopted spray combustion model predicts the parameters reasonably well in the combustor.

1000

Computational

Temperature (K)

800

Experimental
600

400

200

0
0

0.1

0.2

0.3

0.4

0.5

800

1000

Axial distance (m)

(a)
0.05

Computational

Radial distance (m)

0.04

Experimental
0.03

0.02

0.01

0
0

200

400

600

Temperature (K)

(b)
Fig. 4. Comparison of predicted and measured values of (a) wall temperature along
the length of the combustor, (b) exit gas temperature across the radial direction.

5.3. Velocity and temperature distributions in the combustor


We have applied the spray combustion model to predict results
with different air ow splits between the primary and secondary
streams. The total air and fuel ow rates are maintained same
while, three different primary air to secondary air ow ratios
(30:70, 40:60 and 50:50) have been considered. As the primary
air fraction is increased the stoichiometry in the ame region becomes leaner. This alters the structure of the ame and the velocity
and temperature distributions in the ame region. However, the
corresponding reduction in the secondary air ow rate reduces
the momentum in the ow adjacent to the combustor wall, affecting the ow mixing and the convective cooling of the combustor
wall.
Fig. 5(ac) shows the mean velocity and temperature distributions in the vertical plane passing through the axis of the combustor for the three different air ow splits. The axial distance in the
gures has been measured from the plane of the swirler/atomizer,
where x = 0 is considered. The highest temperature zones in the
gures, adjacent to the fuel injector, depict the ame regions. It
is clearly evident from the gures that the ame becomes shorter
in size with the increase in the primary air. The primary air is
admitted in the combustor with a swirling motion casuing a toroidal recirculating zone about the axis adjacent to the fuel injector.
When the primary air ow rate increases the tangential momentum in the inlet stream also increases, and it generates a stronger
central recirculation zone. Under the inuence of the strong swirling motion, the mixing process intensies in the ame zone. The
kinetics of the reactions is considered to be very fast and therefore
the overall combustion rate is controlled by the rates of physical
processes. At higher primary air ow, the increased rates of the
physical processes, like vaporization and mixing, increase the overall reaction rate in the combustor. As a result, the ame becomes
shorter with increase in the primary air. When the primary air fraction is 50% of the total ow rate, the central recirculation bubble is
so strong that it breaks the ame bubble on the axis and the highest temperature ame zone is conned within an annulus close to
the inlet (refer Fig. 5(c)). The ame is short and intense in this case
due to the increased rates of the physical processes. All the temperature contours further show that there is only a little deviation
from axi-symmetry within the combustor. A closer look reveals
that the deviation somewhat increases with the increase in primary air fraction in the combustor.
Fig. 6 compares the centerline temperature variations in the
combustor for the three different air ow splits. It is seen that in
all the cases, the temperature at the plane of the atomizer (x = 0)
is somewhat high. It rst decreases over a very short length in
the downstream direction and then increases to reach a peak value.
Subsequently, the temperature decreases again till the combustor
exit plane is reached. The temperature on the atomizer surface is
high because of the incident radiation from the ame. The peak
centerline temperature is the maximum for the 40:60 ow split
case, though in the 30:70 ow split case the maximum temperature is reached at a further downstream location. In case of
50:50 air ow split, the maximum centerline temperature is much
lower and occurs closer to the inlet plane. This variation in the centerline distributions can be clearly explained from the respective
temperature contour plots. For the 30:70 and 40:60 air ow splits,
the maximum temperature zones are located on the centerline. On
the other hand, in the 50:50 case, the maximum temperature zone
occurs in an annulus, which is away from the center, and the peak
centerline temperature is much less.
The variation in gas temperature at the combustor exit often
has signicance. In case of gas turbine, the exit gas temperature
distribution depicts the combustor pattern factor. A low pattern
factor, signifying more uniform exit gas temperature, is desirable

151

P. Ghose et al. / International Journal of Heat and Mass Transfer 74 (2014) 143155

Fig. 5. Velocity vector and Temperature distributions across the vertical plane through the combustor axis for three different air ow splits between primary and secondary
streams (a) 30:70, (b) 40:60 and (c) 50:50.

2100

0.05

50:50

0.04

40:60

0.03

30:70

1500

Radial distance (m)

Temperature (K)

1800

1200
900
600

50:50

0.02

40:60

0.01

30:70

0
-0.01
-0.02
-0.03

300

-0.04

-0.05

0.1

0.2

0.3

0.4

0.5

Axial distance (m)

200

400

600

800

1000

Temperature (K)

Fig. 6. Variation of centerline temperature along the length of the combustor for
different air ow splits between primary and secondary streams.

Fig. 7. Variation of exit gas temperature from the combustor at three different air
ow splits between primary and secondary streams.

for the health of the turbine. Fig. 7 shows the radial variation of the
gas temperature at the exit to the combustor for the three different
air ow splits. More uniform temperature variation is obtained
when the air ow split is 50:50. As the primary air fraction is less,
the peak temperature at the exit plane, occurring at the axis of the
combustor, increases and the non-uniformity in the temperature
distribution becomes more. This is because of the fact that in the
case of 30:70 air ow split, the ame is longer and the maximum
temperature in the combustor occurs closer to the exit plane.
Therefore, the distance available to transport the energy in the lateral direction becomes considerably shorter. As a result, greater
non-uniformity in the temperature distribution prevails at the exit
plane. When the primary air fraction increases to 40% of the total
air ow, the ame shortens in length and the maximum tempera-

ture on the centerline occurs earlier along the combustor. The temperature variation attens at the exit due to increased transport of
energy in the radial direction. In the third case of 50% primary air,
the maximum temperature is reached even earlier and at an offaxis location. Therefore, not only the axial length available for energy transport is more but also the radial distance over which energy has to be transported becomes less. As a result, the most
uniform exit temperature distribution among the three cases is obtained with the 50:50 air ow split.
5.4. Soot distribution in the combustor
The soot distributions in the combustor are plotted in Fig. 8(a
c) for the three different air ow splits. The soot laden zone is pro-

152

P. Ghose et al. / International Journal of Heat and Mass Transfer 74 (2014) 143155

Fig. 8. Soot volume fraction distributions across the vertical plane through the combustor axis for three different air ow splits between primary and secondary streams (a)
30:70, (b) 40:60 and (c) 50:50.

longed and the peak soot volume fraction is more when the primary air fraction is less. This is due to the fact that with the lower
primary air, the soot precursor concentration in the ame region
increases. The higher precursor concentration along with the extended high temperature zone results in increased formation of
soot over the combustor. It is further to be noted from the soot distribution patterns that, under all the three cases, the maximum
soot volume fraction occurs on the centerline of the combustor.
Fig. 9 shows the variation of soot volume fraction on the centerline of the combustor for three different air ow splits. The peak
soot volume fraction on the centerline is about 9 times higher in
the 30:70 split case and more than 5 times higher in the 40:60 split
case, compared to the 50:50 split case. Furthermore, it is seen from
the soot contours that the concentration of soot near the fuel

3.0E-06

Soot volume fraction

2.4E-06

50:50
40:60
30:70

1.8E-06

1.2E-06

6.0E-07

0.0E+00
0

0.1

0.2

0.3

0.4

0.5

Axial distance (m)

Fig. 9. Variation of soot volume fraction along the combustor centerline for three
different air ow splits between primary and secondary streams.

injector (x = 0) remains quite high. This results in the deposition


of considerable soot on the atomizer body after continuous operation, which is also evident in the experiments. The higher soot volume fraction near the atomizer with lower primary air results in
faster build up of soot on the atomizer surface. When the soot build
up becomes large, the atomization quality of the fuel suffers and
the combustion gets affected.

5.5. Temperature and incident radiative heat ux on combustor wall


and fuel injector
The soot laden gas at high temperature causes increased radiation from the ame zone. The concentration of soot in the ame
zone is dependent on the quantity of primary air supplied to the
combustor. On the other hand, the secondary air, which enters
along the outer wall of the combustor, helps to keep the wall surface cool. Therefore, the split between the primary and secondary
air ow into the combustor will have an effect on the combustor
wall and fuel injector surface temperatures.
This is evident in Fig. 10(a), which plots the variation of incident
radiative heat ux on the combustor peripheral wall for the three
different air ow splits. Taking into account that there is not much
deviation from symmetry in the temperature distribution, the plot
has been made only along a line in the axial direction. The corresponding wall surface temperatures are plotted in Fig. 10(b). It is
clearly evident from Fig. 10(a) that the highest radiative ux on
the wall is incident around the ame close to the inlet to the combustor. This is caused by the high temperature of the ame and the
high luminous radiation from the soot present in the ame zone. At
the downstream location, the radiative ux on the wall is mostly
from the participating gases in the ow. The maximum incident
heat ux due to radiation on the peripheral wall is achieved when
the primary air ow is 30% of the total air ow. The soot volume

P. Ghose et al. / International Journal of Heat and Mass Transfer 74 (2014) 143155

Incident radiation (W/m )

100000
50:50
40:60
30:70

80000

60000

40000

20000

0
0

0.1

0.2

0.3

0.4

0.5

Axial distance (m)

(a)

1000
50:50

Temperature (K)

800

40:60
30:70

600

400

200

0
0

0.1

0.2

0.3

0.4

0.5

Axial distance (m)

(b)
Fig. 10. Variation of (a) incident radiation, (b) outer wall temperature along the
combustor length for three different air ow splits.

153

fraction in the ame is much higher in this case, which is the prime
reason of the increased radiative ux. The maximum incident radiation decreases by more than 50% when the primary air fraction is
increased to 50% of the total ow. However, the peripheral wall of
the combustor is cooled convectively by the ow of secondary air
adjacent to the wall. The secondary air enters the combustor coaxially with the primary air ow and grazes along the wall, while
exchanging energy with the high temperature core as well as with
the combustor wall. When the primary air fraction is more, the
fraction of the secondary air is less and it gives less convective
cooling of the wall. Fig. 10(b) shows the distribution of wall temperature along the length of the combustor. The wall temperature
is seen to increase along the combustor length in all the three
cases. However, the highest wall temperature is attained with
the 50:50 air ow split and the lowest with 30:70 split. This is
attributed to the fact that even with a much higher incident radiation in the 30:70 ow split case, the higher convective cooling
due to increased secondary air ow keeps the wall at a lower temperature. On the contrary, though in the 50:50 case, the maximum
radiative ux incident on the wall is low, but the wall temperature
reaches a higher value as the secondary air ow adjacent to the
wall is less.
It is also signicant to study the incident heat ux and the surface temperature on the fuel injector considering the life of the
injector. Figs. 11 and 12 show the incident radiation ux and surface temperature, respectively, on the swirlerinjector assembly at
the inlet to the combustor for the three different cases of air ow
splits. It is clearly seen that the fuel injector is the more critical part
than the swirler as it receives more radiative heat ux and attains a
higher temperature. On the other hand, the primary air, which directly ows over the swirler keeps the surface of the swirler cool.
The incident radiation on the fuel injector is the highest for the
case of lowest primary air fraction (Fig. 11(a)) due to increased
radiation from the ame. This is attributed to the higher soot con-

Fig. 11. Distributions of incident radiation on the swirlerinjector planes for three different air ow splits between primary and secondary streams (a) 30:70, (b) 40:60 and
(c) 50:50.

Fig. 12. Temperature distributions across the swirlerinjector plane for three different air ow splits between primary and secondary streams (a) 30:70, (b) 40:60 and (c)
50:50.

154

P. Ghose et al. / International Journal of Heat and Mass Transfer 74 (2014) 143155

centration in the ame in this case. The high incident radiation


causes the fuel injector surface temperature to reach a value above
1100 K in this case (Fig. 12(a)). The incident radiation ux on the
injector surface reduces as the primary air fraction is increased
(Fig. 11(b and c)). However, the surface temperature distributions
on the injector (Fig. 12(b and c)) do not show a decrease in value
for the corresponding cases. This may be attributed to the stronger
convective heat transfer with the increase in primary air fraction.
When the primary air ow rate increases, the central recirculation
zone established on the combustor axis becomes more intensied.
As a result, the high temperature gas from the downstream ows
back with a higher velocity towards the injector. The resulting
higher convective heat transfer offsets the lower incident radiative
heat ux and maintains the injector nearly at the same high temperature for all the three cases. Thus the fuel injector remains as
the more critical component of the combustor and its material
has to be selected properly to save it from failure.

6. Conclusions
In the present work, a numerical model has been developed for
simulating spray combustion in a model gas turbine combustor.
The model includes sub-models for soot formation and radiation
heat transfer both from the ame and the soot particles. A Favreaveraged transport model is used for the turbulent ow with turbulence modeled by realizable ke model. Turbulent combustion
is represented using equilibrium presumed probability density
function and BrookesMoss model was adopted for the soot formation with model parameters suited for kerosene ame. The solution
for radiative exchange has been done using discrete ordinate model and the non-gray behavior of the gases has been accounted by
weighted sum gray gas model (wsggm). Lagrangian approach has
been used for modeling the liquid phase transport using expressions pertinent to isolated droplets for drag, heat transfer and
evaporation. The spray formation from a pressure-swirl atomizer
was modeled using Linearized Instability Sheet Atomization (LISA)
model for primary atomization. The initial size distribution of the
droplets in the spray following atomization is considered using
the RosinRammler distribution function. The droplet diameter
following breakup (do) is considered as the size parameter of the
distribution function, while a xed value (3.5) is used for the dispersion parameter. The model was validated with experimental results both from the literature and obtained from a rig in our
laboratory.
The model combustor under consideration has an air ow split
between primary and secondary streams entering co-axially. The
effect of air ow distribution on different combustor parameters
has been investigated by considering three primary to secondary
ow rates as 30:70, 40:60 and 50:50. The results show that as
the proportion of the primary air increases from 30% to 50% of
the total air ow, the ame becomes more compact. This leads to
less soot production in the ame zone and a more uniform temperature pattern factor at the combustor exit. With the reduction in
soot formation, the incident radiative heat ux decreases with
the increase in primary air fraction. The increase in primary air
from 30% to 50% of the total air ow reduces the maximum incident heat ux on the peripheral wall by more than 50%. However,
reduction in the air ow rate near the combustor wall leads to
higher wall temperature, particularly close to the inlet. In addition,
the lower soot formation in the ame at higher primary air fraction
decreases the radiative heat ux from the ame on the injector surface by more than 25%. However, the convective heat transfer to
the injector surface counterbalances the variation in radiative heat
ux and the injector surface temperature remains nearly the same
under all the three air ow splits.

Conict of interest
I, on behalf of my co-authors, certify that there is no conict of
interest with any organization regarding the material discussed in
this manuscript.
Acknowledgment
This work has been supported by the Gas Turbine Research
Establishment, Govt. of India under the GATET scheme (Grant No.
GTRE/GATET/CA07/1012/026/11/001).
References
[1] P. Moin, S.V. Apte, Large-eddy simulation for realistic gas turbine combustors,
AIAA J. 44 (4) (2006) 698708.
[2] K. Luo, H. Pitsch, M.G. Pai, O. Desjardins, Direct numerical simulations and
analysis of three-dimensional n-heptane spray ames in a model swirl
combustor, Proc. Combust. Inst. 33 (2011) 21432152.
[3] C. Hollmann, E. Gutheil, Modelling of turbulent spray diffusion ames
including detailed chemistry, Combust. Inst. (1996) 17311738. Twenty
Sixth Symp. (Int) on Combustion.
[4] A. Datta, S.K. Som, Combustion and emission characteristics in a gas turbine
combustor at different pressure and swirl conditions, Appl. Therm. Eng. 19
(1999) 949967.
[5] H. Watanabe, Y. Matsushita, H. Aoki, T. Miura, Numerical simulation of
emulsied fuel spray combustion with pufng and micro-explosion, Combust.
Flame 157 (2010) 839852.
[6] G. Hsiao, H.C. Mongia, Swirl cup modeling part 3: grid independent solution
with different turbulence models, in: 41st Aerospace Sciences Meeting and
Exhibit, AIAA Paper 2003-1349, 2003.
[7] D. Joung, K.Y. Huh, Numerical simulation of non-reacting and reacting ows in
a 5 MW commercial gas turbine combustor, ASME Paper No. GT 2009-59987,
2009.
[8] V.M Karim, M. Bart, D. Erik, Comparative study of k-e turbulence models in
inert and reacting swirling ows, in: 33rd AIAA Fluid Dynamics Conference
and Exhibit, Paper No. AIAA 2003-3744, 2003.
[9] G.M. Faeth, Mixing, transport and combustion in sprays, Prog. Energy Combust.
Sci. 13 (1987) 293345.
[10] Kenneth K. Kuo, Ragini Acharya, Fundamentals of Turbulent and Multiphase
Combustion, John Wiley & Sons Inc., 2012, pp. 509575.
[11] M. Hallmann, M. Scheurlen, S. Wittig, Computation of turbulent evaporating
sprays: Eulerian versus Lagrangin approach, ASME J. Eng. Gas Turbines Power
117 (1) (1995) 112119.
[12] H. Richter, J.B. Howard, Formation of polycyclic aromatic hydrocarbos and
their growth to soot a review of chemical reaction pathways, Prog. Energy
Combust. Sci. 26 (2000) 565608.
[13] Z.A. Mansurov, Soot formation in combustion processes, Combust. Explo.
Shock Waves 41 (6) (2005) 727744.
[14] H. Richter, S. Granata, W.H. Green, J.B. Howard, Detailed modeling of PAH and
soot formation in laminar preliminary mixture benzene/oxygen/argon at low
pressure ame, Proc. Combust. Inst. 30 (2004) 13971405.
[15] C.S. McEnally, L.D. Pfefferle, B. Atakan, K. Kohse-Hoinghaus, Studies of
aromatic hydrocarbon formation mechanisms in ames: progress towards
closing the fuel gap, Prog. Energy Combust. Sci. 32 (2006) 247294.
[16] I.M. Kennedy, W. Kollmann, J.Y. Chen, A model for the soot formation in
laminar diffusion ame, Combust. Flame 81 (1990) 7385.
[17] K.M. Leung, R.P. Lindstedtand, W.P. Jones, A simplied reaction mechanism for
soot formation in nonpremixed ames, Combust. Flame 87 (1991) 289305.
[18] J. B Moss, C.D. Stewart, K.J. Young, Modeling soot formation and burnout in a
high temperature laminar diffusion ame burning under oxygen-enriched
conditions, Combust. Flame 101 (1995) 491500.
[19] S.J. Brookes, J.B. Moss, Predictions of soot and thermal radiation properties in
conned turbulent jet diffusion ames, Combust. Flame 116 (1999) 486503.
[20] K.B. Lee, M.W. Thring, J.M. Beer, On the rate of combustion of soot in a laminar
soot ame, Combust. Flame 6 (1962) 137145.
[21] J. Nagle, R.F. Strickland-Constable, Fifth Carbon Conference 1 (1962) 154164.
[22] C.P. Fenimore, G.W. Jones, Oxidation of soot by hydroxyl radicals, J. Phys.
Chem. 71 (1967) 593597.
[23] Z. Wen, S. Yun, M.J. Thomson, M.F. Lightstone, Modeling soot formation in
turbulent kerosene/air jet diffusion ames, Combust. Flame 135 (2003) 323
340.
[24] J.B. Moss, I.M. Aksit, Modelling soot formation in a laminar diffusion ame
burning a surrogate kerosene fuel, Proc. Combust. Inst. 31 (2007) 31393146.
[25] A.B. Al-Omari, K. Kawajiri, T. Yonesawa, Soot processes in a methane-fueled
furnace and their impact on radiation heat transfer to furnace walls, Int. J. Heat
Mass Transfer 44 (2001) 25672581.
[26] S.C. Paul, M.C. Paul, Radiative heat transfer during turbulent combustion
process, Int. Commun. Heat Mass Transfer 37 (2010) 16.
[27] D. Byun, S.W. Baek, Numerical investigation of combustion with non-gray
thermal radiation and soot formation effect in a liquid rocket engine, Int. J.
Heat Mass Transfer 50 (2007) 412422.

P. Ghose et al. / International Journal of Heat and Mass Transfer 74 (2014) 143155
[28] L. Tesse, F. Dupoirieux, J. Taine, Monte Carlo modeling of radiative transfer in a
turbulent sooty ame, Int. J. Heat Mass Transfer 47 (2004) 555572.
[29] G.D. Raithby, E.H. Chul, A nite volume method for predicting radiant heat
transfer in enclosures with participating media, J. Heat Transfer 112 (1990)
415423.
[30] T.H. Shih, W.W. Liou, A. Shabbir, Z. Yang, J. Zhu, A new k-e eddy viscosity
model for high Reynolds number turbulent ows, Comput. Fluids 24 (3) (1995)
227238.
[31] D.P. Schmidt, I. Nouar, P.K. Senecal, C.J. Rutland, J.K. Martin, R.D. Reitz, Pressure
swirl atomization in the near eld, SAE Paper No. 1999-01-0496, SAE, 1999.
[32] S.H. Bafekr, M. Shams, R. Ebrahimi, A. Shadaram, Numerical simulation of
pressure-swirl spray dispersion by using EulerianLagrangian method, J.
Dispersion Sci. Technol. 32 (2011) 4755.
[33] S.H. Park, H.J. Kim, H.K. Suh, C.S. Lee, Atomization and spray characteristics of
bioethanol and bioethanol blended gasoline fuel injected through a direct
injection gasoline injector, Int. J. Heat Fluid Flow 30 (2009) 11831192.
[34] P.K. Senecal, D.P. Schmidt, I. Nouar, C.J. Rutland, R.D. Reitz, M.L. Corradini,
Modeling high-speed viscous liquid sheet atomization, Int. J. Multiph. Flow 25
(1999) 10731097.

155

[35] A.H. Lefebvre, X.F. Wang, Mean drop sizes from pressure-swirl nozzles, J.
Propul. Power 3 (1) (1987) 1118.
[36] S.A. Morsi, A.J. Alexander, An investigation of particle trajectories in two phase
ow system, J. Fluid Mech. 55 (2) (1972) 193208.
[37] W.E. Ranz, W.R. Marshall Jr., Evaporation from drops, part I and part II, Chem.
Eng. Prog. 48 (4) (1952) 173180.
[38] D. Joung, K.Y. Huh, 3D RANS simulation of turbulent ow and combustion in a
5 MW reverse-ow type gas turbine combustor, J. Eng. Gas Turbines Power
132 (11) (2010) 111504.
[39] M.F. Modest, Radiative Heat Transfer, McGraw-Hill, 1993.
[40] H. Watanabe, R. Kurose, S. Komori, H. Pitsch, Effects of radiation on spray ame
characteristics and soot formation, Combust. Flame 152 (2008) 213.
[41] T.F. Smith, Z.F. Shen, J.N. Friedman, Evaluation of coefcients for the weighted
sum of gray gases model, J. Heat Transfer 104 (1982) 602608.
[42] A. Basak, J. Patra, R. Ganguly, A. Datta, Effect of transesterication of vegetable
oil on liquid ow number and spray cone angle for pressure and twin uid
atomizers, accepted for publication in Fuel.

Você também pode gostar