Journal of Energy Resources Technology. Received August 21, 2014;

Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
SIMULATION OF OXY- FUEL COMBUSTION
OF HEAVY OIL FUEL IN A MODEL FURNACE
ed
Co
py
ot
Ma
nu
sc
rip
tN
M. A. Habib, third author 1
King Fahd University of Petroleum & Minerals
Professor,
Mechanical Engineering Department
KFUPM, Dhahran-31261, Saudi Arabia
Office: Bldg-22, Room-121-B
Phone: 966-13-860 4467
Fax:
966-13-860 2949
e-mail: [email protected]
ite
d
Rached Ben-Manosur, first author
King Fahd University of Petroleum & Minerals
Assistant Professor,
Mechanical Engineering Department
KFUPM, Dhahran-31261, Saudi Arabia
Office: Bldg-22, Room-121.1
Phone: 966-13-860 2014
Fax:
966-13-860 2949
e-mail: [email protected]
Ac
ce
pt
ed
Pervez Ahmed, second author
King Fahd University of Petroleum & Minerals
Research Engineer,
KACST-Technology Innovation Center on Carbon Capture and Sequestration
KFUPM, Dhahran-31261, Saudi Arabia
Office: Bldg-15, Room-5114
Phone: 966-13-860 7869
Fax:
966-13-860 2266
e-mail: [email protected]
1
Corresponding Author/ Prof. M. A. Habib, Tel: +966 13 860 4467,
Email: [email protected]
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
ABSTRACT
The present study aims at investigating the characteristics of oxy-combustion of heavy oil liquid fuel in a
down-fired model furnace. Non-premixed PDF mixture model was used to simulate the combustion
characteristics and turbulence chemistry. The validation of the present model was performed against the
experimental data and is found to be in good agreement. The results depict that the oxy-combustion of
ite
d
liquid fuels results in lower soot. It is observed that the soot formation is reduced when N2 in air-
ed
combustion is replaced by O2 in oxy-combustion. However, it increases as the amount of oxygen in oxycombustion increases. Replacing nitrogen in the air combustion by carbon dioxide in oxyfuel combustion
tN
ot
Keywords: Model furnace, liquid fuels, oxy-combustion, CFD
Co
py
tends to reduce the temperature levels in the upstream sections of the combustion chamber.
rip
1. INTRODUCTION
nu
sc
Significant efforts in the combustion process have been made in the past to
Ma
obtain high thermal efficiencies along with a reduction in NOx emissions [1]. Many
researchers [2-6] have reviewed the aspects of atomization and vaporization of liquid
pt
ed
fuel droplets along with spray combustion with emphasis on combustion engines and
ce
gas-turbine combustor applications. These aspects provide a method of instant
Ac
vaporization and mixing of liquid fuels with oxidizer that significantly affects the
combustion rate. Spraying liquid fuels during the combustion process greatly affects the
ignition, heat release rate, exhaust emissions and pollutant formations [7]. Many [8-11]
modeling efforts of liquid atomization and spray combustion have been attempted
during the past few years and satisfactory level of agreement between the experimental
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
data and numerical simulations in flame structure and soot predictions are successfully
obtained.
Givler and Abraham [5] reviewed the vaporization and combustion of paraffin
fuel droplets from n-pentane to n-dodecane under supercritical pressures and
ite
d
temperatures. Assumptions in modeling the vaporization and combustion and their
ed
relative importance were also discussed. Their work indicates that a vaporizing droplet
Co
py
reach the critical state for ambient pressures and ambient temperatures greater than
approximately twice the fuel’s critical pressure and critical temperature respectively
tN
ot
where as combusting droplets reach this state when ambient pressure is approximately
rip
2.5 times that of the fuel’s critical pressure. Aoki et al [12] studied numerically the spray
nu
sc
combustion of petroleum coke and heavy oil in a swirling fired-type furnace for soot
emissions in exhaust gas and formation of H2S at the bottom of the furnace. They found
Ma
satisfactory agreement between the numerical simulations and the experimental data.
ed
They also found that with increase in the ratio of tangential air to the secondary air, the
pt
unburned ratio decreases. They also showed that the tangential air introduced from the
ce
bottom of the furnace results in a reduction in the H2S formation rate. Watanabe et al
Ac
[10] studied numerically spray combustion of liquid kerosene oil including soot models.
They found that the temperature without soot radiation model was higher than that of
the experimental data. However, including soot model reduced the differences
considerably as soot has significant effect on radiative heat transfer. Moreover, mole
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
fraction of NO in the exhaust including soot model is in good agreement with the
experimental data.
Syngas combustion under lean conditions in a supercharged dual engine using a
constructed syngas chemical kinetic mechanism was studied by Azimov et al [13].
ite
d
Various syngas combinations of initial H2 concentration were examined to determine
ed
the total heat of reaction during the reaction propagation. Optimum results were
Co
py
obtained in below 0.8 equivalence ratio for almost all types of syngas under study. In
their experiments conducted by Roby et al [14], they demonstrated that the
tN
ot
vaporization of lean, pre-vaporized premixed liquid fuel in atmospheric and high
rip
pressure combustor rigs incorporated with swirl-stablilized burners present similar
nu
sc
operation to that obtained for burning natural gas. Extended lean operation was
observed for the liquid fuels as a result of the wider lean flammability range for liquid
Ma
fuel compared with natural gas. Moreover, LPP systems were able to produce low NOx
ed
and low CO emissions without auto-ignition and flashback. Derudi and Rota [15, 16]
pt
carried out mild combustion experiments of different liquid fuels. They found that
ce
combustion characteristics are more influenced by the physical state of the fuel than by
Ac
its chemical composition indicating that the mild combustion burners can lower soot
formation allowing the use of wide range of liquid fuels. Saario et al [17] simulated a
cylindrical laboratory furnace for heavy oil fuel combustion using two different
turbulence models, k-Ɛ and Reynolds stress model (RSM). It was found that the standard
k-Ɛ model was not able to predict the swirling flow field satisfactorily whereas the RSM
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
model significantly improved the predictions with some discrepancies near the vicinity
of the furnace. Combustion and characteristics of liquid fuel in engines were reported
[18, 19]. In this work, the effect of different parameters on combustion emissions were
analyzed.
ite
d
In their work of large eddy simulations of an evaporating two-phase flow in a
ed
burner, Sanjose et al [20] demonstrated that simplified injection methods are
Co
py
appropriate for the simulations of real-time combustion geometries. The extent of spray
formation of fuels such as gasoline, n-pentane, n-butanol, ethanol and iso-octane at
tN
ot
different temperatures and pressures were investigated. The influence of hydrogen
rip
additions on hydrocarbon emissions were reported [21]. Their results indicated that the
nu
sc
lean burn limit of methane–air turbulent combustion was improved with hydrogen
addition. The extent of spray formation of fuels such as gasoline, n-pentane, n-butanol,
Ma
ethanol and iso-octane at different temperatures and pressures were investigated by
ed
The extent of spray formation of fuels such as gasoline, n-pentane, n-butanol, ethanol
pt
and iso-octane at different temperatures and pressures were investigated by Aleiferis et
ce
al [22] using phase Doppler and laser diffraction techniques. Jager and Kohne [23]
Ac
showed that developing burner systems for light fuel oil can reduce NOx emissions
compared to conventional burners. They also modeled the phenomenon using CFX
software for optimization using RNG k-epsilon turbulence model. Cerea et al [24]
investigated the combustion of liquid bio-fuels using a dual nozzle laboratory scale
burner. Their investigations showed that dual nozzle burners can lower NOx and soot
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
concentrations indicating the possible use of wide range of low-BTU liquid fuels. For
practical applications in computation fluid dynamics (CFD) codes, the simplified model
for radiative heating, describing the average droplet absorption efficiency factor,
appears to be the most useful both from the point of view of accuracy and CPU
efficiency [4]. The computation of spray dynamics in terms of the equations of fluid
ite
d
mechanics and droplet ballistics is now possible by use of CFD codes incorporating spray
ed
modeling [2]. Van Blarigan et al [25] conducted an experimental investigation of
Co
py
methane fuel oxy-combustion in a variable compression ratio, spark-ignited piston
engine. Their results indicated that an optimum oxygen concentration at which fuel-
nu
sc
rip
conditions with 29% oxygen content.
tN
ot
conversion efficiency is maximized exists when operating under oxy-combustion
There is no enough work conducted in the past on the combustion
Ma
characteristics of liquid fuels using CFD packages and there is no previous work for oxy-
ed
fuel combustion of heavy oil fuels. Therefore, the present study aims at investigating the
pt
oxy-combustion characteristics of heavy oil fuel in a down fired furnace in order to
ce
better understand the fluid dynamics of the process. Moreover, the best operating
Ac
O2/CO2 compositions for oxy-combustion of heavy oil fuel that are required to obtain
the characteristics similar to air-fuel combustion are also investigated.
2. FURNACE MODEL
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
The furnace model used for the present study is shown in Figure 1. It is a down
fired cylindrical furnace. Due to symmetry the present furnace is modeled as 2D axissymmetric to reduce computational and maintain accurate predictions. The height and
diameter of the furnace are 2.4m and 0.6m respectively. The first half of the furnace is
blanketed with a refractory lining around which a thick ceramic layer is laid. An air
ite
d
assisted atomizer burner gun is used to inject fuel into the furnace from the top while
ed
the secondary air is supplied through a double concentric configuration around the
Co
py
burner gun.
3. MATHEMATICAL MODELING
tN
ot
Turbulent combustion modeling involves a wide range of coupled problems such
rip
as chemical reaction scheme, two or three phase system and radiative heat transfer in
nu
sc
addition to fluid mechanical properties. The model is governed by Navier-Stokes
equations, species and energy transport equations with incorporated Reynolds and
Ma
Favre averaging [26]. Probability Density Function (PDF) approach is recommended for
ed
better accuracy compared to Eddy Dissipation Model (EDM) [27]. Mixture Fraction ‘f’ is
pt
defined as mass fraction of primary (fuel) stream:
ce
sY fu − Yox + Yox ,0
sY fu ,1 + Yox ,0
Ac
f =
where,
Yfu  mass fraction of fuel
Yox  mass fraction of oxidizer
subscript ‘0’  oxidant stream
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
(1)
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
subscript ‘1’  fuel stream
and s is the stoichiometric oxygen to fuel ratio.
Probability Density function, which can be defined as the fraction of time
p ( f )∆f =lim
1
∑τ i
T i
(2)
ed
T →∞
ite
d
spent by the fluid in the vicinity of state ‘f’, is given by equation (2) as follows:
Co
py
where, ‘T’ is the time scale and τi is the amount of time that ‘f’ spends in the Δf
band. Favre (density weighted) average of any other scalar quantities, like temperature
ot
and species, can be calculated by integrating the product of the scalar & PDF [28]. A PDF
tN
table for a desired composition of fuel and oxidizer stream is generated by performing
nu
sc
rip
the chemistry calculations and co-relating the gas variables with the mixture fraction.
Ma
Discrete Ordinate (DO) model is used to solve radiative heat transfer in the
reactor, with 5 flow iterations for each iteration of radiation. Absorption coefficient of
pt
ed
the gas mixture is determined by domain based weighted sum-of-gray-gas model. The
ce
Euler-Lagrange approach is utilized to solve the discreet phase model. The fluid phase is
Ac
treated as continuum, while the dispersed phase is solved by tracking the
particles/droplets through the calculated flow field, where both the phases can
exchange mass, momentum and energy. The trajectory of discrete particle is predicted
by integrating the force balance on the particle in the Lagrangian reference frame.
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
du p
dt
= FD (u − u p ) +
gx (ρ p − ρ )
+ Fx
ρp
(3)
where, FD(u – up) is the drag force and Fx is the force arising due to the pressure
gradient along the fluid. The dispersion of droplets is calculated through Stochastic
tracking (Discrete Random Walk) model, which solves the interaction of the droplets
ed
ite
d
with a succession of discrete stylized fluid phase turbulent eddies.
Co
py
The heat and mass transfer for discrete phase are solved by incorporating
inert heating Law is implied:
rip
tN
dTp
m p c p = hAp (T∞ − Tp )+ ε p Apσ (θ R4 − Tp4 )
dt
ot
three Laws. When droplet temperature is less than the vaporization temperature, the
(4)
nu
sc
where, the heat transfer coefficient is calculated using the correlation of Ranz &
k∞
 2 + 0.6 Re1/d 2 Pr1/ 3 
dD 
(5)
pt
ed
h
=
Ma
Marshall [29].
ce
Droplet Vaporization Law is applied when droplet temperature is above the
Ac
vaporization temperature but below the boiling point.
=
N v κ c ( Cv , D − Cv , ∞ )
(6)
Droplet Boiling Law is applied to predict the convective boiling of the
droplet when the temperature of the droplet has reached the boiling point.
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
−r
dmD
= hAD (T∞ − TD ) + ε D ADσ (TR4 − TD4 )
dt
(7)
The effect of soot concentration on the radiation absorption coefficient has been
taken into account by determining the absorption coefficient for soot. The overall
ite
d
absorption coefficient was defined as the sum of the absorption coefficients of the
ed
radiating gas and soot. The soot formation and consumption are governed by the
Co
py
following transport equation which was solved for the soot mass fraction.
(8)
ot
→
µ
∂
( ρY s ) + ∇.( ρν Y s ) = ∇.( t ∇Y s ) + R s , net
∂t
σs
tN
where Ys is the soot mass fraction, σx is the turbulent Prandtl number for soot
rip
transport and Rs,net is the net rate of soot generation [kg/m3s]. Rs,net is calculated
nu
sc
according to
R=
R s , gen − R s ,ox
s , net
(9)
pt
ed
R s=
C s Pf Φ r e − ERT
, gen
Ma
The rate of soot formation was based on a simple empirical rate [17]
ce
where Cs and r are constants, Pf is the fuel partial pressure, and φ is the fuel
Ac
equivalence ratio. The constants Cs and r were set to 1.5 /(kg.m.s) and 3, respectively.
Soot formation occurs only when φ is greater than the incipient soot limit and a value
above which soot formation becomes negligible. The soot oxidation can be limited
either by soot concentration or oxygen concentration, and thereby the effective
oxidation rate is the minimum of the following equations
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
R s ,ox ,1 = A ρY s
ε
(10)
k
Y
Y sν s
ε
)
R s ,ox ,2 = A ρ ( ox )(
ν s Y sν s +Y f ν f k
(11)
ite
d
where A is a constant, Yx is the stoichiometric oxygen to soot mass ratio and νf is
the stoichiometric oxygen to fuel mass ratio. The definition of the turbulent mixing time
Co
py
ed
scale k/3 and the model constant A proposed in the above soot combustion rate
expression are somewhat speculative.
ot
3.1 Numerical solution
tN
A finite-volume based commercial CFD-code was used to solve the governing
nu
sc
rip
equations. The time-averaged equations for the conservation of mass, momentum, fuel
mixture fraction and its variance, and enthalpy were solved along with transport
Ma
equations for k and Ɛ. The pressure field was calculated from the continuity equation
ed
using the SIMPLE algorithm. These partial differential equations are discretized and
pt
approximated by algebraic equations for finite number of volumes in the domain [28].
ce
Grid independence test has been carried out on three different grids. The mass fraction
Ac
of O2 at a radial distance of 20 mm from the entrance of the boiler has been presented
in. It is found that the grid with 9526 nodes is unable to capture the distribution
accurately. Grids with 15889 and 22795 nodes present results that are in close
proximity. Grids with more than 25000 nodes are also tested but resulted in no obvious
advantage but rather increased the computational effort and time. Therefore, an
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
optimal grid of 22795 nodes has been used throughout the study. The difference
between grid size 15889 and 22795 is less than 1 %.
4 RESULTS AND DISCUSSIONS
ite
d
The results of air-combustion modeling of heavy oil fuel are discussed in this
ed
section along with the validation and comparison of air-combustion and oxy-combustion
Co
py
characteristics of heavy oil liquid fuel. Combustion characteristics of oxy-combustion of
heavy oil liquid fuel in a model furnace having a vertical cylindrical combustion chamber
ot
down fired were calculated and are presented in this section. The fuel has 85% Carbon,
tN
11.2 % hydrogen in addition to small amounts of nitrogen sulfur and ash. Excess air of
rip
15 % was used. The influence of replacing nitrogen by carbon dioxide on temperature
nu
sc
distributions, evaporation rates and species concentrations are presented. As well, the
Ma
influence of the oxygen content in the oxidizer (the mixture of the oxygen and carbon
dioxide) on these parameters are presented and discussed.
ed
4.1 Model Validation
ce
pt
The validation for the present combustion model was performed against the
Ac
experimental results presented by Saario et al [17]. In this work, experimental data for a
combustor fired by an industry-type swirl burner for which the initial conditions of the
spray have been characterized. The combustion data include measurements of gas
species concentrations (O2, CO, CO2 and NOx) at several locations along the furnace. The
combustion chamber is cylindrical in shape. Its axis is vertical and it is down-fired. The
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
burner consists of a central gun and a secondary air supply in a conventional doubleconcentric configuration, terminating in a refractory quarl. Swirl is imparted to the
secondary stream. Figure 2 shows the molar distribution of O2 and CO2 at three different
radial positions of 20mm, 300mm and 600mm from the entrance of the furnace. The
figure indicates that the numerical results obtained using the current developed model
ite
d
is in good agreement with the experimental data. Due to lack of experimental data for
ed
temperature profiles of liquid fuels combustion the present temperature contours are
Co
py
compared with the work done by Saario et al [17] and are found to match well with
their work.
tN
ot
4.2 General features of characteristics of oxy-fuel combustion
rip
Figure 3 presents the contours of the velocity vector in the boiler furnace for
nu
sc
oxy-fuel combustion of heavy oil fuel. The velocity vector contour reveals that there are
Ma
two recirculation regions created downstream of the sudden expansion as a result of
created adverse pressure gradient and are expected to improve mixing and enhance
ed
combustion. Due to the expansion of the swirled flow in the furnace regions, reversed
ce
pt
flows are formed. When the jet of gases eject out of the nozzle into the furnace, two
Ac
different regions are distinguished. One is due to the sudden expansion of the gases that
creates recirculation in the corners due to entrainment of oxygen. These are called
corner recirculation zones (CRZ). The CRZ occupies almost the whole width of the
furnace. The second is when the secondary flow hits the walls, thus, a second
recirculation zone (SRZ) below the secondary flow zone is created downstream. A
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
results also indicate that similar evaporation rates to that of air combustion can be
ite
d
achieved at increased oxygen content in the oxidizer gas.
ed
ACKNOWLEDGMENT
Co
py
The authors wish to acknowledge the support received from King Fahd University of
Petroleum and Minerals (KFUPM) and SABIC for funding this work through project No.
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
SB 121002.
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
REFERENCES
[1]
G.-M. Choi and M. Katsuki, "Chemical kinetic study on the reduction of nitric
oxide in highly preheated air combustion," Proceedings of the Combustion Institute, vol.
29, pp. 1165-1171, 2002.
[2]
X. Jiang, G. Siamas, K. Jagus, and T. Karayiannis, "Physical modelling and
advanced simulations of gas–liquid two-phase jet flows in atomization and sprays,"
Progress in Energy and combustion Science, vol. 36, pp. 131-167, 2010.
ed
ite
d
[3]
W. A. Sirignano, "Fuel droplet vaporization and spray combustion theory,"
Progress in Energy and combustion Science, vol. 9, pp. 291-322, 1983.
Co
py
[4]
S. S. Sazhin, "Advanced models of fuel droplet heating and evaporation,"
Progress in Energy and combustion Science, vol. 32, pp. 162-214, 2006.
ot
[5]
S. D. Givler and J. Abraham, "Supercritical droplet vaporization and combustion
studies," Progress in Energy and combustion Science, vol. 22, pp. 1-28, 1996.
nu
sc
rip
tN
[6]
O. Askari, H. Metghalchi, S. K. Hannani, A. Moghaddas, R. Ebrahimi and H.
Hemmati, Fundamental Study of Spray and Partially Premixed Combustion of
Methane/Air Mixture, J. Energy Resour. Technol. 135, 021001 (2012) (9 pages); Paper
No: JERT-12-1093; doi:10.1115/1.4007911
Ma
[7]
Y. Wei, K. Wang. W. Wang, S. Liu and Y. Yang, Contribution Ratio Study of Fuel
Alcohol and Gasoline on the Alcohol and Hydrocarbon Emissions of a Gasohol Engine, J.
Energy Resour. Technol. 136, 022201 (2013) (7 pages); Paper No: JERT-12-1219;
doi:10.1115/1.4024716
ce
pt
ed
[8]
L. Zhang and S.-C. Kong, "Modeling of multi-component fuel vaporization and
combustion for gasoline and diesel spray," Chemical Engineering Science, vol. 64, pp.
3688-3696, 2009.
Ac
[9]
J. Broukal and J. Hájek, "Validation of an effervescent spray model with
secondary atomization and its application to modeling of a large-scale furnace," Applied
Thermal Engineering, vol. 31, pp. 2153-2164, 2011.
[10] H. Watanabe, Y. Suwa, Y. Matsushita, Y. Morozumi, H. Aoki, S. Tanno, and T.
Miura, "Spray combustion simulation including soot and NO formation," Energy
conversion and management, vol. 48, pp. 2077-2089, 2007.
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
the furnace. It can be seen that the extent of flame is shortened dramatically due to
high swirl number of 1.1 otherwise with low swirl number the flame would have
propagated to an extended length. This can be verified by the velocity contour
presented in Fig. 7a.
4.3 Oxy-fuel combustion characteristics
ite
d
Figure 8 presents the temperature contours of oxy-heavy oil and oxy-heavy oil
ed
fuel combustion. Three cases of oxy-heavy oil fuel combustion namely OF_23, OF_30
Co
py
and OF_35 are presented. OF_23 is similar to the air-heavy oil fuel case except that the
N2 in air is replaced by CO2. OF_30 and OF_35 represent the oxy-heavy oil fuel cases
tN
ot
with enriched oxygen i.e. 30% and 35% O2 respectively. It is important to mention that
rip
the contours are presented on different scales and the corresponding colors at the
nu
sc
bottom highlights the temperature variation. It is observed that the replacement of N2
(in air) by CO2 (in oxy-heavy oil fuel) in case of OF_23 reduces the maximum
Ma
temperature obtained from 1836K in air-case to 1400K in OF_23 case. This may be
ed
attributed to the higher specific heat of CO2 compared to N2. Specific heat of CO2 is
pt
around 7 % higher than that of N2. At higher temperatures, the specific heat capacity of
ce
CO2 increases resulting in a reduced combustion temperature. It can also be observed
Ac
that by replacing CO2 with N2 in case of OF_23 shifts the maximum temperature zone
towards the walls of the furnace. Moreover, the zone of maximum temperature is also
increased. In order to explain this phenomenon, the velocities at three different radial
sections are plotted in Fig. 9. It is clear from these figures that the velocity is higher in
case of air-fuel case compared to oxy-combustion cases. This may be attributed to the
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
higher molecular weight of CO2 (44) compared to N2 (28). The higher densities in oxycombustion cases compared to air-combustion case, shown in Fig. 10a, along the axial
distance also adds to the justification of reduced velocities. Furthermore, the high
velocities increases the turbulent viscosity, as shown in Fig. 10b, that helps in better
mixing of fuel with the oxidizer. Therefore, due to lower turbulent viscosity in OF_23
ite
d
case the maximum flame temperature zone is dispersed and is moved towards the side
ed
walls. As we increase the oxygen content at the inlet as in case of OF_30 and OF_35 the
Co
py
zone of high temperature region is reduced. However, the maximum temperature
obtained increases as we move from OF_23 case to OF_35 case. About 1440K, 1800K
tN
ot
and 1988K temperatures are obtained in OF_23, OF_30 and OF_35 case respectively. An
rip
increase of approximately 27% is seen in the maximum temperature obtained when the
nu
sc
oxygen percentage is increased from 23% to 35%. Due to increase oxygen content it is
expected that the fuel is completed burned with high heat release rates thereby
Ma
increasing the maximum temperature. On the other hand, the increase in oxygen
ed
content tends to reduce the exhaust gas temperatures. The specific heat is shown in Fig.
pt
10c and indicate that the multiple ρCp is higher for the oxy-fuel combustion cases. This is
Ac
combustion.
ce
expected to result in delay in ignition in the oxy-fuel cases in comparison to air-fuel
Figure 11a presents comparison of temperature profiles along the axis of the
furnace for the air and oxy-combustion cases. It is observed that the temperature is
reduced when N2 in air combustion is replaced by CO2 in oxy-combustion case.
However, the temperature levels increase with increased O2 in the oxy-combustion
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
cases. These results confirm the temperature contours explained above. Moreover,
from these temperature profiles one can depict that in order to maintain the
temperature levels of air-combustion case, the amount of oxygen has to be slightly
higher in oxy-combustion cases,. It is clear from these plots that for the case of OF_30
i.e. 30% O2 (remaining CO2) similar temperature profiles of air-combustion are obtained.
ite
d
It is also revealed from these plots the region of maximum temperature is shifted away
ed
from the inlet when N2 in air is replaced by O2 in oxy-combustion cases. This may be
Co
py
attributed to the higher molecular weight of CO2 (44) compared to N2 (28) that tends to
reduce the velocities. In order to further explain the delay in combustion, the turbulent
tN
ot
intensity profiles are presented in Fig. 11b and 8c. It can be seen that air-combustion
rip
case has the highest turbulent intensity compared to oxy-combustion cases. The higher
nu
sc
turbulent intensity ensures better mixing whereas lower turbulent intensity causes a
delay in mixing that leads to a shift in the maximum temperature zones.
Ma
Figure 12 presents the mass fraction of products of combustion along the center
ed
line of the furnace. High mass fraction of CO is observed near to the inlet region of the
pt
furnace where the availability of oxygen is low. However, it becomes almost negligible
ce
as one moves downstream. At downstream sections, the CO formed in the furnace gets
Ac
converted to CO2 due to enough supply of oxygen that reaches through the secondary
recirculation zone where CO is completely oxidized to CO2 and its mass fraction along
the center line increases. Since 15% excess air is supplied to the furnace traces of
oxygen can still be found even after combustion. This excess amount is supplied in order
to make sure that enough oxygen is supplied for the complete combustion of the fuel. It
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
is obvious that the formation of H2O starts once the fuel is burnt in the reaction. In
addition, as there was a small amount of sulfur already present in the heavy oil fuel,
very small traces of H2S formed during the reaction can be observed along the center
line near to the inlet region. However, it tends to become negligible as one moves to the
downstream sections.
ite
d
The contours of mass fractions for the air combustion case are shown in Fig. 13
ed
and confirm those of the axial distributions. The figure also indicates absence of O2 in
Co
py
the vicinity of the centerline at the upstream region where CO and sulfur exist. As one
moves downstream, CO and sulfur are burnt with diffused oxygen and they almost
tN
ot
vanish close to exit of the furnace.
rip
Contours of mass fraction of soot are shown in Fig. 14 for air-combustion and
nu
sc
three different oxy-combustion cases. It is important to mention that the contours are
presented on same scales and the corresponding colors at the bottom highlights the
Ma
soot mass fraction variation. It is seen from these contours that the soot concentration
ed
starts with a low value close to the inlet in the vicinity of the fuel injector. Downstream
pt
of the location of the flame lift-off, it increases and reaches a maximum value where the
ce
mixture is rich and small amount of oxygen is available. As one moves towards the
Ac
leading edge of the flame, the soot concentration decreases as it is oxidized by the
secondary air that diffuse into the flame. It is known that there are competing processes
between soot formation and soot oxidation depending on the local equivalence ratio
and temperature (Kitamura et al, [30]). It is observed that the air-combustion predicts
the highest soot formation. It is reduced by approximately 1000 times when N2 is
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
replaced by CO2 in oxy-combustion. However, as we increase the oxygen content in the
oxy-combustion cases, soot mass fraction also increases. It can be seen that, in air
combustion and OF_23 cases, the soot mass fraction region is bigger but its size
decreases as the amount of oxygen in OF_30 and OF_35 is increased.
The maximum local soot concentration was found to be 7.2E-02 g/Nm3 air,
ite
d
1.06E-04 for OF-23 case, 1.55E-03 for OF-30 case, 6.8E-03 for OF-35 case in case of air-
ed
fuel combustion where as in oxy-fuel combustion. Soot formation starts increasing from
Co
py
the inlet section where there is no availability of enough oxygen as can be seen from the
O2 contours. Downstream of this point, it becomes zero since all the soot formed is
tN
ot
oxidized due to oxygen recirculating. It is believed that soot formation and soot
rip
oxidation occur simultaneously depending upon the amount of available oxygen and the
nu
sc
local temperature. The main disadvantage of sulfur is the risk of corrosion by sulfuric
acid formed during and after combustion, and condensation in cool parts of the
Ma
chimney or stack, air pre-heater and economizer.
ed
Figure 15 presents CO concentrations for the case of air and three cases of
pt
oxyfuel combustion. It is believed that the emissions of CO are influenced by the
ce
temperature in the furnace reactions and post-flame zones. CO increases with increase
Ac
in temperature. This can be verified by the CO mass fraction contours presented in Fig.
15. It is observed that the temperature also increases as the amount of oxygen in the
oxy-combustion cases increases and so the concentration of CO. As the flame
temperature decreases, NO, decrease while CO and hydrocarbons tend to increase.
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Since the latter two emissions are indicative of combustion efficiency and stability,
decreasing flame temperature can lead to both decreased efficiency and blowout limits.
Radiation from a non-luminous flame is due primarily to carbon dioxide and
water vapor, i.e., the products of complete combustion. The amount of heat radiated to
the walls of the combustor is the product of flame emissivity and temperature raised to
ite
d
the fourth power. As flame temperature increases, the cooling requirements for the
ed
liner also increase. This is not a problem for gaseous fuels since emissivity is usually
Co
py
considerably low, and regardless of gaseous flame temperature, the radiant loading to
the combustor wall will not be as great as is the case with distillate. Nevertheless, it may
tN
ot
be desirable to alter the liner cooling to maintain acceptable metal temperature
rip
gradients on the liner. It may be important to note that molar concentrations of CO 8.9
nu
sc
mol/m3 OF-35, 9.1 mol/m3 OF-30, 8.1 mol/m3 OF-23, 3.04 mol/m3 AF-23. In the
hydrocarbon fuel combustion, intermediate reaction species such as O, CO, OH, H, etc.,
Ma
are normally formed. The oxidation of CO is one of the slowest reactions. Since the
ed
combustion process inside a gas turbine combustor must take place rapidly, any
pt
decrease in reaction rate may lead to decreased combustion efficiency, i.e., CO in the
ce
exhaust and the potential for high CO emissions exists. As this ratio H/O increases,
Ac
flammability limits and reaction rates are enhanced and the propensity for CO emissions
decreases.
4.4 Fuel Evaporation
Evaporation rates are presented in Figure 16a and Figure 16b at two radial
locations of 20mm and 320mm for the case of air combustion and three cases of oxy-
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
fuel combustion. As shown from Fig. 16a, the oxy-fuel case exhibits the highest rate of
evaporation. As O2% is increased, the evaporation rates are reduced and become close
to the air case. At the downstream section, fuel is completely evaporated for the air
case and for the O2 enriched cases. The delay in evaporation rate is only shown for the
oxy-fuel case with 21% oxygen.
ed
ite
d
5 CONCLUSIONS
Co
py
The characteristics of oxy-combustion of liquid fuels such as heavy oil fuel in a
down-fired model furnace have been studied. Non-premixed PDF mixture model was
ot
used to simulate the combustion characteristics and turbulence chemistry. The
tN
validation of the present model was performed against the experimental data and is
rip
found to be in good agreement. The results indicate that the oxy-combustion of liquid
nu
sc
fuels results in lower soot. Evaporation rates are shown to be high in the oxy-fuel
Ma
combustion case. Oxygen enriching the oxy-fuel cases provide similar evaporation rates
as the air case. It is observed that the soot formation is decreased when N2 in air-
ed
combustion is replaced by O2 in oxy-combustion. However, it increases as the amount of
ce
pt
oxygen in oxy-combustion increases. The results also show that production of carbon
Ac
monoxide and sulfur depends significantly on the oxygen content in the oxidizer gas.
The results also shows that replacing nitrogen in the air combustion by carbon dioxide in
oxy-fuel combustion tends to reduce the temperature levels in the upstream sections of
the combustion chamber. This was attributed to the high thermal capacity of
combustion gases in the case of oxy-fuel combustion leading to delay in ignition. The
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
ed
Ma
nu
sc
rip
tN
ot
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Ac
ce
pt
Fig. 9 Radial velocity distributions at two axial locations (a) X = 20mm (b) X = 600mm for
air and oxy-fuel combustion
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
REFERENCES
[1]
G.-M. Choi and M. Katsuki, "Chemical kinetic study on the reduction of nitric
oxide in highly preheated air combustion," Proceedings of the Combustion Institute, vol.
29, pp. 1165-1171, 2002.
[2]
X. Jiang, G. Siamas, K. Jagus, and T. Karayiannis, "Physical modelling and
advanced simulations of gas–liquid two-phase jet flows in atomization and sprays,"
Progress in Energy and combustion Science, vol. 36, pp. 131-167, 2010.
ed
ite
d
[3]
W. A. Sirignano, "Fuel droplet vaporization and spray combustion theory,"
Progress in Energy and combustion Science, vol. 9, pp. 291-322, 1983.
Co
py
[4]
S. S. Sazhin, "Advanced models of fuel droplet heating and evaporation,"
Progress in Energy and combustion Science, vol. 32, pp. 162-214, 2006.
ot
[5]
S. D. Givler and J. Abraham, "Supercritical droplet vaporization and combustion
studies," Progress in Energy and combustion Science, vol. 22, pp. 1-28, 1996.
nu
sc
rip
tN
[6]
O. Askari, H. Metghalchi, S. K. Hannani, A. Moghaddas, R. Ebrahimi and H.
Hemmati, Fundamental Study of Spray and Partially Premixed Combustion of
Methane/Air Mixture, J. Energy Resour. Technol. 135, 021001 (2012) (9 pages); Paper
No: JERT-12-1093; doi:10.1115/1.4007911
Ma
[7]
Y. Wei, K. Wang. W. Wang, S. Liu and Y. Yang, Contribution Ratio Study of Fuel
Alcohol and Gasoline on the Alcohol and Hydrocarbon Emissions of a Gasohol Engine, J.
Energy Resour. Technol. 136, 022201 (2013) (7 pages); Paper No: JERT-12-1219;
doi:10.1115/1.4024716
ce
pt
ed
[8]
L. Zhang and S.-C. Kong, "Modeling of multi-component fuel vaporization and
combustion for gasoline and diesel spray," Chemical Engineering Science, vol. 64, pp.
3688-3696, 2009.
Ac
[9]
J. Broukal and J. Hájek, "Validation of an effervescent spray model with
secondary atomization and its application to modeling of a large-scale furnace," Applied
Thermal Engineering, vol. 31, pp. 2153-2164, 2011.
[10] H. Watanabe, Y. Suwa, Y. Matsushita, Y. Morozumi, H. Aoki, S. Tanno, and T.
Miura, "Spray combustion simulation including soot and NO formation," Energy
conversion and management, vol. 48, pp. 2077-2089, 2007.
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
[11] H. Watanabe, Y. Suwa, Y. Matsushita, Y. Morozumi, H. Aoki, S. Tanno, and T.
Miura, "Numerical investigation of spray combustion in jet mixing type combustor for
low NOx emission," Energy conversion and management, vol. 49, pp. 1530-1537, 2008.
[12] H. Aoki, S. Tanno, T. Miura, and S. Ohnishi, "Three Dimensional Spray
Combustion Simulation in a Practical Boiler," Transactions of the Japan Society of
Mechanical Engineers Series B, vol. 57, pp. 2152-2158, 1991.
ed
ite
d
[13] U. Azimov, M. Okuno, K. Tsuboi, N. Kawahara, and E. Tomita, "Multidimensional
CFD simulation of syngas combustion in a micro-pilot-ignited dual-fuel engine using a
constructed chemical kinetics mechanism," International Journal of Hydrogen Energy,
vol. 36, pp. 13793-13807, 2011.
Co
py
[14] R. J. Roby, M. S. Klassen, L. D. Eskin, M. Ramotowski, and G. Gaines,
"Development of a system for lean, prevaporized, premixed combustion," in
Proceedings of the Thirty-sixth turbomachinery symposium. TX: Texas A&M University,
College Station, 2007, pp. 55-60.
rip
tN
ot
[15] M. Derudi and R. Rota, "NOx Emissions of A Mild Combustion Burner Operated
with Jet Fuel Surrogates." The thirty fourth Meeting of the Italian Section of the
Combustion Institute, 2011, pp. 1-6, Rome.
nu
sc
[16] M. Derudi and R. Rota, "Experimental study of the mild combustion of liquid
hydrocarbons," Proceedings of the Combustion Institute, vol. 33, pp. 3325-3332, 2011.
Ma
[17] A. Saario, A. Rebola, P. Coelho, M. Costa, and A. Oksanen, "Heavy fuel oil
combustion in a cylindrical laboratory furnace: measurements and modeling," Fuel, vol.
84, pp. 359-369, 2005.
ce
pt
ed
[18] R. K. Maurya and A. K. Agrawal, Combustion and Emission Characterization of nButanol Fueled HCCI Engine, J. Energy Resour. Technol. 137, 011101 (2014) (12 pages);
Paper No: JERT-13-1318; doi:10.1115/1.4027898
Ac
[19] D. Romero, R. N. Parthasaratky and S. R. Gollahalli, Laminar Flame Characteristics
of Partially Premixed Prevaporized Palm Mthyl Ester and Diesel Flames, J. Energy
Resour. Technol. 136, 032204 (2014) (11 pages); Paper No: JERT-13-1338;
doi:10.1115/1.4027406
[20] M. Sanjosé, J. Senoner, F. Jaegle, B. Cuenot, S. Moreau, and T. Poinsot, "Fuel
injection model for Euler–Euler and Euler–Lagrange large-eddy simulations of an
evaporating spray inside an aeronautical combustor," International Journal of
Multiphase Flow, vol. 37, pp. 514-529, 2011.
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
[21] O. Askari, H. Mtghalchi, S. K. Hannani, H. Hemmati and r. Ebrahimi, Lean Partially
premixed Combustion Investigation of Methane Direct-Injection under Different
Characteristic Parameters, J. Energy Resour. Technol. 136, 022202 (2014) (7 pages);
Paper No: JERT-13-1184; doi:10.1115/1.4026204
[22] P. G. Aleiferis and Z. R. van Romunde, "An analysis of spray development with
iso-octane, n-pentane, gasoline, ethanol and n-butanol from a multi-hole injector under
hot fuel conditions," Fuel, vol. 105, pp. 143-168, 2013.
ed
ite
d
[23] F. K. Jäger and H. Köhne, "CFD modelling in process burner development:
Combustion of light residual fuel oils," in Computational Fluid and Solid Mechanics 2003,
K. J. Bathe, Ed., ed Oxford: Elsevier Science Ltd, 2003, pp. 966-970.
Co
py
[24] M. Cerea, M. Derudi, and R. Rota, "Preliminary Study On Mild Combustion
Characteristics of a Liquid Biofuel." The thirty fifth Meeting of the Italian Section of the
Combustion Institute, 2012, pp VIII.11.1 - VIII.11.6, Milano.
rip
tN
ot
[25] A. Van Blarigan, D. Kozarac, R. Seiser, R. Cattolica, J. Y. Chen and R. Dibble,
Experimental Study of Methane Fuel Oxycombustion in a Spark-Ignited Engine, J. Energy
Resour.
Technol.
136,
012203
(2013);
Paper
No: JERT-13-1090;
doi:10.1115/1.4024974
Ma
nu
sc
[26] B. Rohani, M. A. Wahid, M. M. Sies and K. M. Saqr, 2012, “Comparison of eddy
dissipation model and presumed probability density function model for temperature
prediction in a non-premixed turbulent methane flame,” AIP Conf. Proc., 1440, pp. 384–
391.
ed
[27] D. Veynante and L. Vervisch, 2002, “Turbulent combustion modeling,” Prog.
Energy Combust. Sci., 28(3), pp. 193–266.
ce
pt
[28] H. K. Versteeg and W. Malasekara, 1995, An introduction to computational fluid
dynamics, the finite volume method, Longman Group Ltd, Essex.
Ac
[29] W. Ranz and W. Marshall, 1952, “Evaporation from drops,” Chem. Eng. Prog,
48(3), pp. 141–146.
[30] T. Kitamura, T. Ito, J. Senda and H. Fujimoto, Mechanism of smokeless diesel
combustion with oxygenated fuels based on the dependence of the equivalence ration
and temperature on soot particle formation, Int. J Engine Research, Vol. 3, No 4, pp 223248.
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Figure Captions List
Line diagram of the model furnace in the present study
Fig. 2
Comparison of present calculations and experimental data [17]
Fig. 3
Contour of velocity (m/s) vector for oxy-fuel combustion of heavy oil fuel
Fig. 4
Contour of turbulent intensity (%) for oxy-combustion of heavy oil fuel
Fig. 5
The instantaneous view of the droplet diameter (m) distribution inside
the furnace
Fig. 6
Contours of Evaporation of heavy oil fuel droplet diameter (m) in oxyfuel combustion
Fig. 7
(a) Velocity and (b) temperature contours for oxy-heavy oil fuel
combustion
Fig. 8
Temperature contours of air and oxy-fuel combustion
Fig. 9
Radial velocity distributions at two axial locations (a) X = 20mm (b) X =
600mm for air and oxy-fuel combustion
Fig. 10
Axial distributions of (a) Density (b) Turbulent viscosity and (c) Specific
heat for air and oxy-fuel combustion
Fig. 11
Axial distributions of (a) temperature and (b) Turbulent intensity for air
and oxy-fuel combustion
Fig. 12
Axial distributions of species concentrations of (a) CO (b) CO2 (c) H2S (d)
S (e) O2 for air and oxy-fuel combustion
ce
Ac
Fig. 13
pt
ed
Ma
nu
sc
rip
tN
ot
Co
py
ed
ite
d
Fig. 1
Contours of species concentrations of (a) O2 (b) CO2 (c) CO and (d) Soot
for air combustion
Fig. 14
Contours of soot concentrations for air and oxy-fuel combustion
Fig. 15
Contours of CO concentrations for air and oxy-fuel combustion
Fig. 16
Radial distributions of evaporation rate for air and oxy-fuel combustion
at different axial locations (a) X=20mm (b) X=300mm from the inlet
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Co
py
ed
ite
d
FIGURES
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Fig. 1 Line diagram of the model furnace in the present study
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Fig. 2 Comparison of present calculations and experimental data [17]
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Fig. 3 Contour of velocity (m/s) vector for oxy-fuel combustion of heavy oil fuel
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Fig. 4 Contour of turbulent intensity (%) for oxy-combustion of heavy oil fuel
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Fig. 5 The instantaneous view of the droplet diameter (m) distribution inside the furnace
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Fig. 6 Contours of Evaporation of heavy oil fuel droplet diameter (m) in oxy-fuel
combustion
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
ot
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
Fig 7 (a) Velocity and (b) temperature contours for oxy-heavy oil fuel combustion
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Fig. 8 Temperature contours of air and oxy-fuel combustion
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
ed
Ma
nu
sc
rip
tN
ot
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Ac
ce
pt
Fig. 9 Radial velocity distributions at two axial locations (a) X = 20mm (b) X = 600mm for
air and oxy-fuel combustion
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Co
py
ed
ite
d
Fig 10 Axial distributions of (a) Density (b) Turbulent viscosity and (c) Specific heat
for air and oxy-fuel combustion
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
ed
Ma
nu
sc
rip
tN
ot
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Ac
ce
pt
Fig 11 Axial distributions of (a) Temperature and (b) Turbulent intensity for air and oxyfuel combustion
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Fig 12 Axial distributions of species concentrations of (a) CO (b) CO2 (c) H2S (d) S (e) O2
for air and oxy-fuel combustion
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Fig. 13 Contours of species concentrations of (a) O2 (b) CO2 (c) CO and (d) Soot for air
combustion
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Fig. 14 Contours of soot concentrations for air and oxy-fuel combustion
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Fig. 15 Contours of CO concentrations for air and oxy-fuel combustion
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms
Ac
ce
pt
ed
Ma
nu
sc
rip
tN
ot
Co
py
ed
ite
d
Journal of Energy Resources Technology. Received August 21, 2014;
Accepted manuscript posted November 7, 2014. doi:10.1115/1.4029007
Copyright (c) 2014 by ASME
Fig. 16 Radial distributions of evaporation rate for air and oxy-fuel combustion at
different axial locations (a) X=20mm (b) X=300mm from the inlet
Downloaded From: http://asmedigitalcollection.asme.org/ on 11/14/2014 Terms of Use: http://asme.org/terms