Вы находитесь на странице: 1из 15

Energy Conversion and Management 81 (2014) 83–97

Contents lists available at ScienceDirect

Energy Conversion and Management


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

Evaluation of gas radiation models in CFD modeling of oxy-combustion


M.A. Rajhi a, R. Ben-Mansour a, M.A. Habib a, M.A. Nemitallah a,b,⇑, K. Andersson c
a
Mechanical Engineering Departments, KFUPM, Dhahran 31261, Saudi Arabia
b
Mechanical Engineering Department, Massachusetts Institute of Technology, 77 Mass Avenue, Cambridge, USA
c
Department of Energy and Environment, Energy Technology, Chalmers University of Technology, SE-412 96 Goteborg, Sweden

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

Article history: Proper determination of the radiation energy is very important for proper predictions of the combustion
Received 4 October 2013 characteristics inside combustion devices using CFD modeling. For this purpose, different gas radiation
Accepted 7 February 2014 models were developed and applied in the present work. These radiation models vary in their accuracy
Available online 3 March 2014
and complexity according to the application. In this work, a CFD model for a typical industrial water tube
boiler was developed, considering three different combustion environments. The combustion environ-
Keywords: ments are air–fuel combustion (21% O2 and 79% N2), oxy-fuel combustion (21% O2 and 79% CO2) and
Gas radiation models
oxy-fuel combustion (27% O2 and 73% CO2). Simple grey gas (SGG), exponential wide band model
Oxy-fuel combustion
Air combustion
(EWBM), Leckner, Perry and weighted sum of grey gases (WSGG) radiation models were examined and
Computational fluid dynamics (CFD) their influences on the combustion characteristics were evaluated. Among those radiation models, the
EWBM was found to provide close results to the experimental data for the present boiler combustion
application. The oxy-fuel combustion characteristics were analyzed and compared with those of air–fuel
combustion.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction effect, it can be neglected in the RTE. By applying this simplifica-


tion, the radiative transfer equation becomes a differential equa-
Energy production from fossil fuel combustion results in the tion, which is much easier to be solved than the integral–
emission of carbon dioxide (CO2) and nitrogen oxides (NOx). differential equation. Viskanta [1] indicated that, in cases where
Reduction of these pollutants becomes an international commu- soot particles are small, such as those cases of methane and pro-
nity interest and many investigators have been adopted for this pane combustion, the scattering is a function of the soot particle
purpose. One of the most promising options to capture carbon diameter raised to the power four. Accordingly, it is justified to ne-
dioxide (CO2) and reduce NOx emission is oxy-fuel combustion glect the scattering effect in the present calculations. However, this
in which the combustion utilizes pure oxygen instead of air. Ther- assumption cannot be generalized to other combustion applica-
mal radiation is the principal mode of heat transfer in the combus- tions. Thus, care should be taken during the treatment of the soot
tion chamber and the radiative properties of combustion products modeling as it directly affects the radiation heat transfer. Soot par-
such as CO2 and H2O affect the heat transfer to the furnace wall ticle size effects on the radiation heat transfer were reported in
tubes. many studies for different combustion applications [2–4] and dif-
In the present work, the transfer of radiant energy was obtained ferent flame types [5,6].
from the solution of the radiative transfer equation (RTE) as de- The generation rate of soot particles inside many combustion
scribed in the coming sections. The radiative transfer equation is applications has been described in various models. Those models
of the integral–differential type. This equation is complex and, have shown good agreement with experimental data for different
therefore, it is difficult to be exactly solved in 3-D geometries. flame types and different combustion applications [7]. The influ-
The following simplifications to this equation were, therefore, ence of soot particles on the radiation heat transfer in a multi com-
incorporated. Since the scattering in the combustion of gas fuel partment system has been demonstrated by Yeoh et al. [8]. Their
(such as Natural Gas, CH4 and Propane, C3H8) has an insignificant numerical results have shown good agreement when compared
with the experimental data of Luo and Beck [9], considering the
soot model proposed by Moss et al. [10] and Syed et al. [11]. Yeoh
⇑ Corresponding author at: Mechanical Engineering Department, Massachusetts et al. [8] concluded that the accurate calculation of the rate of soot
Institute of Technology, 77 Mass Avenue, Cambridge, USA. Tel.: +966 598212483.
generation is very essential for accurate CFD modeling of
E-mail addresses: medhatahmed@kfupm.edu.sa, mahmed@mit.edu (M.A. Nemi-
combustion.
tallah).

http://dx.doi.org/10.1016/j.enconman.2014.02.019
0196-8904/Ó 2014 Elsevier Ltd. All rights reserved.
84 M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97

Nomenclature

Iv spectral radiation intensity (W m2 sr1 Hz1) r Stefan Boltzmann constant (W m2 K4)
I total intensity (W m2 sr1) s transmissivity
Ib blackbody intensity (W m2 sr1) sH optical depth at band head
Ib,m spectral black body intensity (W m2 sr1 Hz1) sk transmissivity of band k
L path length (m) sk,i?n transmissivity of band k and path i ? n
c velocity of radiation (m/s) x bandwidth parameter (cm1)
Ak total band absorptance of band k (Hz if frequency is eo zero partial pressure emissivity
used in the integration, lm if wavelength is used, jv spectral absorption coefficient
cm1 if wavenumber is used) rv scattering coefficient
Pe equivalent broadening pressure Uv phase function
X density path-length (g/m2) a integrated band intensity (cm1/(g/m2))
PE effective pressure (bar) x band width parameter (cm1)
ae,i emissivity weighting factors for the grey gas i Dec+w overlap between bands of CO2 and H2O
be,i,j emissivity gas temperature polynomial coefficients g, l, n direction cosine
aa,i absorptivity weighting factors
ca,i,j,k absorptivity polynomial coefficients Abbreviations
Eov C;k spectral blackbody emissive power (W/(m2 Hz), or other DOM discrete ordinates method
units if wavelength or wavenumber is used) RTE radiative transfer equation
LBL line-by-line model
Greek symbols SNB statistical narrow band model
b mean line-width-to-spacing parameter WBM wide band model
e emissivity CFD computational fluid dynamics
ec emissivity of carbon dioxide WSGG weighted sum of grey gases model
eg total gas emissivity EWBM exponential wide band model
ew emissivity of water vapour SLW spectral line-based weighted-sum-of-grey-gases model
ek spectral emissivity SGG simple grey gas model
f partial pressure ratio BA block approximation method
g pressure corrected line-width-to-spacing parameter BEA band energy approximation
k stoichiometric ratio
kmax pressure correction parameter

The Soot generation rate should be accurately predicted in the During oxy-fuel combustion, the flue gas tri-atomic molecules
flame zone in order to correctly predict the combustion character- concentration increases drastically and this leads to changes in
istics inside any combustion system, especially for the case of oxy- the gas emissivity. In order to assess the suitability of retrofitting
fuel combustion [12]. Changing the energy carrying medium from an air-fired boiler to oxy-fuel combustion, Zheng et al. [19,20] cal-
nitrogen in case of conventional air combustion to carbon dioxide culated the gas emissivity from the correlations suggested by Leck-
in case of oxy-fuel combustion is expected to result in modifica- ner [21] for total gas emissivity for the water vapour and carbon
tions in the soot generation rate [13]. It has been confirmed, in dioxide.
many recent publications, that oxy-fuel combustion results in less Andersson and Johansson [22] studied experimentally the com-
amount of soot [14,15]. bustion of propane fuel for three test cases: air–fuel combustion
Predictions of nine total emissivity models against the calcula- (referenced one) and other two oxy-combustion cases (OF21 @
tions conducted using the EWBM were compared by Lallemant 21% O2 and OF27 @ 27% O2). The difference in total radiation inten-
et al. [16]. Hottel et al. [17] showed that the thermal radiation from sity between these cases was determined and compared. In addi-
water vapour, carbon dioxide, fly ash, soot and carbon monoxide tion, the difference in combustion environment was described.
presents the major contributor of heat transfer from a flame pro- They reported that, in case of the OF 21 combustion, the tempera-
duced by conventional fuels. Any change of CO2 and H2O concen- ture levels were found to be lower than those of air-fired condition.
trations results in a change in radiative heat transfer. For air This was attributed to two reasons; (1) the higher specific heat
combustion with conventional partial pressures of CO2 and H2O, capacity of CO2 compared to N2, and (2) increase in radiation losses
the heat transfer calculations are carried out by using a three caused by the CO2 with increase the gas emissivity. In a recent
grey-one clear gas model to estimate flame emissivity. study, Johansson et al. [23] evaluated several approximate gas
The concentrations of CO2 and H2O gases in oxy-fuel combus- radiative property models for oxy-fuel environments in large
tion are high in comparison to those of air-fired combustion. Both atmospheric boiler. These models included statistical narrow band
CO2 and H2O, unlike N2 which is transparent to thermal radiation, (SNB) models, WSGG model, spectral line-based weighted-sum-of-
emit and absorb thermal radiation. Accordingly, the emissivity of grey-gases model (SLW) and two grey-gas approximations model.
the flue gas increase in the case of oxy-fuel combustion due to The accuracy of the radiation models is determined by the
the high concentrations of absorbing/emitting gases. These have importance of the radiative source term. Lallemant et al. [16] indi-
subsequent effects on the radiative heat transfer. Wall [18] stated cated that the heat released by combustion dominates the radia-
that the heat transfer prediction is critical to oxy-fuel technology, tive release in regions of intensive combustion. They compared
since there are changes in gas properties due to CO2 recycling from gas radiation models with measurements of total radiation inten-
furnace outlet back to the furnace inlet. These changes are attrib- sity in a 300 kW non-sooting natural-gas flame. The agreement be-
uted to alteration of gas radiative properties and gas heat capacity. tween measurements and computed data by the EWBM was
M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97 85

excellent for measurements against cold furnace walls. The WSGG spectral transmissivity averaged over a narrow band. This model
model is recommended when the influence of the radiative source is suitable for the radiation heat transfer prediction in high tem-
term is significant. Porter et al. [24] used the spherical harmonic perature mediums. The wide band model (WBM) is a simplification
(P1) and the discrete ordinates method (DO) to solve the RTE in of the SNB model; it yields wide band absorptance and requires the
multi-dimensional problem. The spectral nature of radiation has knowledge of the path length in the model as well as the spectral
been treated by using non-grey gas full spectrum k-distribution parameters associated with the path length. In the following sub-
method (FSCK) and a grey method. The simulation results of air– sections, the details of the considered gas radiation models in the
fuel and oxy-fuel combustion were compared with the statistical present work are presented.
narrow band model (SNB). They concluded that the non-grey full
spectrum k-distribution method was in good agreement with the 2.1. Simple grey gas model (SGG)
statistical narrow band model (SNB). Significant errors were intro-
duced in the wall heat flux when a grey model was used. The simple grey gas model (SGG), in addition to being simple, it
The WSGG model is based on the fitting of air–fuel combustion requires low execution time. The model considers the effective
characteristics. This model was modified by Johansson et al. [25] to absorption coefficient as the main parameter controlling the radi-
become suitable to cover the temperature range of 500–2500 K in ative properties of a gas mixture and assumes that radiant absorp-
order to be used in case of cover oxy-fuel combustion. They have tion and emission by gas molecules to be independent of the
reported that the modified WSGG is a computationally efficient op- frequency of the radiation. Under the grey gas assumption, and
tion for CFD simulations. Hjärtstam et al. [26] utilized the CFD neglecting scattering of radiation, the RTE for the radiation inten-
method to investigate the influence of gas and gas-soot radiation sity (integrated over the entire spectrum) in 3-D Cartesian coordi-
mechanisms in air and oxy-fuel flames when propane fuel is burnt nates is expressed as follows:
in a 100 kW test furnace. They showed that in the case of Oxy-fuel
@I @I @I
combustion both grey and non-grey weighted sum of grey gases n þ g þ l ¼ je I þ je Ib ð2Þ
@x @y @z
models overestimate the peak temperature. In addition, they sta-
ted that, the inclusion of soot modeling in both grey and non-grey In the full modeling of a gas-fired furnace, the RTE is solved for
gas radiation models has a significant impact in the CFD calcula- known temperature field and species concentration that are deter-
tions. In the present work, a CFD model was developed and used mined by conservations equations of motion, energy and mass of
to simulate three different combustion environments based on dif- species. At the start of the solution, an initial guess is made for
ferent gas radiation models. all of the parameters. Then, iterative methodology is used in order
to get the final converged solution. The RTE is solved based on the
2. Gas radiation models calculations of temperature and species concentrations from the
previous step. A good estimation of the effective absorption coeffi-
In the present study, the radiant energy was calculated from the cient from the known properties (a mean beam length Lm and a
RTE solution. This equation is based on the conservation principle characteristic gas temperature) of emission by the gas can be ob-
applied to a monochromatic bundle of radiation. The RTE is given tained from interpretation of the total emissivity of the gas upon
by Viskanta [1] as follows: the bounding surface. The characteristic temperature can be the
Z Z volume–average gas temperature Tm, as given in the following
1 dIv rm equation:
¼ ðjv þ rv ÞIv þ jv n2v Ibv þ s0 ! ~
Um ð~ s; m0
c dt 4p Dm; X¼4p je ¼ ð1=Lm ÞIn½1  eg ðT m ; Lm Þ ð3Þ
s0 ÞdX0 dm0
! mÞIv 0 ð~ ð1Þ
Or the local temperature, T, giving a locally varying value,
where the different coefficients and terms of the above equation
je ¼ ð1=Lm ÞIn½1  eg ðT; Lm Þ ð4Þ
are defined as: ‘‘Iv’’ is the spectral radiation intensity, ‘‘c’’ is the
speed of electromagnetic wave in vacuum, ‘‘jv’’ is the spectral The mean beam length is estimated as
absorption coefficient, ‘‘rv’’ is the scattering coefficient, ‘‘Ibv’’ is
Lm ¼ 3:6V=S ð5Þ
the Planck’s spectral blackbody intensity of radiation, ‘‘nv’’ is the
spectral index of refraction of the medium and ‘‘Uv’’ is the phase where V represents volume of furnace and S is the wall surface area.
function. The spectral radiation intensity is a function of three spa-
tial coordinates, time and two angles, in addition to the radiation 2.2. Exponential wide band model (EWBM)
frequency or the wavelength. Solution of radiative transfer requires
models to account for the directional and spectral natures of radi- The exponential wide band model is based on a physical analy-
ation. One of the popular methods to treat the directional nature of sis of gas absorption. This model provides a set of semi-empirical
radiation is the discrete ordinates method (DO). The method is expressions to predict the total band absorptance of infrared active
originally suggested by Chandrasekhar [27] for astrophysical appli- molecules. This model can be used to predict radiative properties
cations. This method was applied in the present calculations and in a wide range of temperature, total pressure range, volumetric
its solution was obtained through the solution of Eq. (1) for a set fraction and path length [29]. In this model, it is assumed that
of discrete directions spanning the total solid angle range of 4p. the total absorption of a vibration–rotation band can be approxi-
The spectral nature of radiation is very important aspect in the mated with correlations dependent on three parameters: the inte-
gas radiation treatment. In general, there are three different mod- grated band intensity, a, the band width parameter, x, and the
els used to define the radiative properties of combustion gases mean line-width-to-spacing parameter, b. These parameters de-
[28]. Those models are the Spectral line-by-line models, the Spec- pend on temperature, but pressure effects are also accounted for
tral band models and the Global model. The complexity of the through the equivalent broadening pressure, Pe. The parameters
models decreases from the line model to the global model. In the yield asymptotic relations for the total band absorptance Ak. These
line-by-line (LBL) model, the RTE is integrated over detailed molec- relations are known as the four-region expression, which is defined
ular spectrum for the gases. This model is used only for benchmark as linear, square root, log-root and logarithmic regions.
solutions due to the enormous amount of computational require- The expressions of the band transmissivity, sk, are derived from
ments. The statistical narrow band model (SNB) provides the the relation:
86 M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97

 
sH @Ak p pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sk ¼ ð6Þ PE ¼ PT 1 þ 4:9 w 273=T ð16Þ
Ak @ sH pT
The optical depth at the band head sH and the pressure correc-  
p
tion parameter g are calculated from: PE ¼ PT 1 þ 0:28 c ð17Þ
pT
aX
sH ¼ ð7Þ The total emissivity of the gas mixture of H2O and CO2 for the
x
desired partial and total pressures is equal to the sum of emissiv-
g ¼ bPe ð8Þ ities of the two gases minus a correction term Decw due to overlap
in some spectral regions. The overlap between the two gases is
where X is the density path-length. The parameters a and b are cal- approximated by the following equation:
culated according to the simplified relations presented by Lallemant 
and Weber [29], while x is given by a correlation dependent on f
Decw ¼  0:0089f10:4  k2:76 ð18Þ
temperature. In order to calculate the total emissivity of H2O–CO2 10:7 þ 101f
mixtures, band energy approximation (BEA) is used. In this method,
In which the partial pressure ratio f is given by
it is assumed that the blackbody emissive power is constant over
each absorption band. The total emissivity is given by the following pw
f¼ ð19Þ
equation [30]: pw þ pc

X o
k¼N E and the parameter k is defined as
v C;k
eg ffi  Ak  Decþw ð9Þ
k ¼ log10 ððpw þ pc Þ  LÞ ð20Þ
k¼1 r  T4

The correction term, Dec+w, accounts for the overlap between


the 2.7 and 15 lm bands for mixtures of CO2 and H2O. The simpli- 2.4. Perry model
fied procedure proposed by Modak [31] is used in the present
study to calculate this overlapping. Empirical correlations for the emissivities of water vapour, car-
bon dioxide, and four mixtures of the two gases were developed in
2.3. Leckner model the 8th edition of the Perry’s Chemical Engineering Handbook [32].
They are valid to be used for pressure-path-length of range .005–
The model was developed by Leckner in 1972 [21]. The CO2 and 10 m-atm. The emissivities at three temperatures of 1000, 1500
H2O emissivities are treated separately. Then, the total emissivity and 2000 K at six values of partial pressure ratios of H2O/CO2,
of the mixture is determined by taking a total band overlap correc- namely, 0, 0.5, 1.0, 2.0, 3.0 and 1 can be calculated by Eq. (21).
tion term into consideration. The advantage of this model is that it Empirical constants for different partial pressure ratios of H2O/
can be applied for any arbitrary partial pressures of CO2 and H2O. CO2 are given in [32]. These correlations were developed based
The total emissivity for a mixture of CO2 and H2O is calculated from on the data in Hottel emissivity charts [17] and were adjusted to
the following equation: the more recent data from cross-handler [33]. Linear interpolation
or extrapolation of the emissivities determined at 1000, 1500 and
eg ¼ ec þ ew  Decw ð10Þ 2000 K is used in order to obtain the emissivity at different tem-
In this model, a zero partial pressure emissivity for either CO2 or perature and it is given by the following equation:
H2O is given by: eg T H ðT g  T L Þ þ eg T L ðT H  T g Þ
eg T g ¼ ð21Þ
X
i¼M 500
i
ln eo ¼ ao þ ai k ð11Þ where TH and TL are the higher and the lower temperature,
i¼1
respectively.
In which:
2.5. Weighted sum of grey gas model (WSGG)
X
j¼N
ai ¼ C oi þ C ji sj ð12Þ
j¼1 There are several gas radiative properties models, called global
models; they are based on the concept of weighted sum of grey
k ¼ log 10ðpi LÞ ð13Þ gases. Example of these models is Hottel and Sarofim’s gas radia-
tive properties model [17]. Eq. (22) is used for the evaluation of to-
where pi L is the pressure path length of either H2O or CO2 (given in
tal emissivity in terms of the weighted sum of grey gases and it is
bar.cm) and s = T/1000 (the gas temperature T is given in K). The
useful especially for the zonal method of analysis of radiative
coefficients Cji are found in Ref. [21]. The pressure correction term
transfer.
is determined using the relation:
   X
i¼I
ei ei e¼ ae;i ðTÞ½1  eji PL  ð22Þ
1 1 ¼ expfnðkmax  kÞ2 g ð14Þ
eo eo max i¼0

where ei is the emissivity of either CO2 or H2O. At high pressures the where ae,i is the emissivity weighting factors for the grey gas i. The
emissivity follows an asymptotic behavior according to: weighting factors are dependent on gas temperature T. ji, is the
  absorption coefficient, P is the sum of partial pressures of absorbing
ei A  PE þ B gases and L is the thickness of gas layer, or path length. The weight-
¼ ð15Þ
eo max PE þ A þ B  1 ing factor for clear gas, i.e. for i = 0, is defined as:
The parameters n, kmax, A and B are given in Ref. [21]. The effec-
X
i¼I
tive pressure PE is defined in Eqs. (16) and (17) for H2O and CO2 ae;0 ðTÞ ¼ 1  ae;i ð23Þ
respectively. i¼1
M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97 87

A common representation of the temperature dependency of


the weighting factors is a polynomial of order J  1 given as:

X
j¼J
ae;i ðTÞ ¼ be;i;j T j1 ð24Þ
j¼1

where be,i,j is referred to as the emissivity gas temperature polyno-


mial coefficients. The absorption coefficients ji and the polynomial
coefficients be,i,j are obtained by fitting Eq. (22) to a table of total
emissivities previously calculated from the EWBM. The total emis-
sivities for various gas mixtures can be obtained utilizing experi-
mental measurements, by calculation based on spectral lines or
bands based on Hottel’s charts, etc. The total absorptivity is calcu-
lated as follows:

X
i¼I
a¼ aa;i ðT; T s Þ½1  eji PL  ð25Þ
i¼1

where aa,i is the absorptivity weighting factor, T is the gas temper-


ature and Ts is the surface irradiation temperature. The weighting Fig. 1. Layout of the furnace of the present boiler.
factors for a CO2/H2O/clear gas mixture with PCO2 = 0.1P, PH2O = 0.2-
P, and P = 101.3 kPa, have been reported by Truelove [34].
All the WSGG model parameters, i.e. weighting factors and excessive air). It is assumed that all fuel is converted to combustion
absorption coefficients, are intended for air-combustion condi- products of CO2 and H2O and the total heat input into the furnace is
tions, and their validity to be used in oxy-fuel combustion is lim- 295 MW. The fuel mass flow rate through each burner is about
ited. Pressure path-lengths and ratios of H2O to CO2 in oxy-fuel 0.98 kg/s. Primary and secondary air passing through each burner
combustion are drastically different from that of conventional are equal to 8.87 kg/s and 10.7 kg/s respectively. The combustion
combustion. Therefore, there is a need to derive new parameters process takes place in environment of atmospheric pressure. The
to be used in these models. Johansson et al. [23] have developed volume of the furnace was discretized into 1,335,180 control vol-
two WSGG models, the first consist of three grey one clear gas, umes. This analysis covers three different combustion environ-
and the second of four grey one clear gas. ments; (1) AF21 (21% O2 and 79% N2), (2) OF21 (21% O2 and 79%
CO2) and (3) OF27 (27% O2 and 73% CO2).
3. Boiler description and CFD modeling
3.2. CFD modeling
3.1. Boiler description
The set of governing differential equations, i.e. continuity,
CFD modeling of a full scale fossil-fuel fired furnace is con- momentum, energy and conservation of chemical species, together
ducted in this study. The objective of this modeling is to investigate with the boundary conditions were numerically solved. The stea-
the influence of radiation models on combustion characteristics dy-state mass, momentum, energy and species conservation equa-
and to analyze the impact of replacing air with a mixture of O2– tions for Newtonian fluids were considered:
CO2 in the combustion. The boiler is manufactured by Mitsubishi r  ðqUÞ ¼ 0 ð26Þ
Heavy Industries (MHI) and is operated in Saudi Aramco refineries.
A brief summary of design data is given in Table 1. The layout of
r  ðqUUÞ ¼ rP þ lr2 U ð27Þ
this boiler is depicted in Fig. 1. It consists of D-type tubes that ex-
tend from the mud drum (lower drum) through the bottom wall,
the side wall, top of furnace and ends at the upper drum. The tubes
ðqC P ÞU  rT ¼ r  ðkrTÞ ð28Þ
forming the bottom wall are thermally insulated. Tubes on the
front wall extend from lower to upper drum while the tubes form- r  ðqUY i Þ  r  ðqDi rY i Þ ¼ 0 ð29Þ
ing the side walls extend from a lower header to an upper header. where u is the velocity vector, q is the fluid density, p is the pres-
The water-steam mixture inside the tube is, under steady state sure, l is the dynamic viscosity, k is the thermal conductivity, Di
conditions, normally saturated having a saturation temperature is the diffusion coefficient, and Yi is the scalar species mass fraction.
corresponding to the drum pressure. Accordingly, the furnace walls The CFD Fluent package [35] utilizing the finite volume method is
provide a constant temperature (538.6 K) boundary condition. As used for representing and evaluating partial differential equations
shown in Fig. 1, the boiler consists of 6 burners located in two lev- in the form of algebraic equations. The CFD calculations normally
els. Each burner has eight fuel nozzles with 24 fuel jets in each noz- provide detailed results including velocity, temperature, species
zle. The burner construction is shown in Fig. 2. Methane fuel (CH4) concentrations and heat flux that are not easily obtained through
is burnt in the furnace with stoichiometric ratio of about 1.16 (16% experimental measurements. The models that used in the solution
are [36–40]; (1) Incompressible ideal gas assumption, (2) Standard
k–e turbulent model, (3) Temperature-dependent properties, (4)
Table 1
Design data of the boiler. Two-step oxy-combustion volumetric reactions for species trans-
port, (5) Finite-rate/Eddy-dissipation turbulence-chemistry interac-
Mass flow rate of Drum pressure, Steam Number of
tion, and (6) Discrete Ordinate Radiation Model (DO). The
steam, lb/h psig temperature, F burners
discretization of the governing equations was conducted through
750,000 750 750 6 on two
the use of a segregated solver (each equation is solved separately.
levels
In order to couple the calculations of pressure and velocity, a
Semi-Implicit Method for Pressure-Linked Equations (SIMPLE)
88 M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97

Fig. 2. The burner construction.

Table 2
Modified two-step methane–oxygen combustion mechanisms with kinetic rate data [41].

Reaction number Reactions A b Ea (J/kmole) Reaction orders


Reaction 1 CH4 + 1.5O2 ? CO + 2H2O 1.59  1013 0 1.998  108 [CH4]0.7[O2]0.8
Reaction 2 CO + 0.5O2 ? CO2 3.98  108 0 4.18  107 [CO][O2]0.25[H2O]0.5
Reaction 3 CO2 ? CO + 0.5O2 6.16  1013 0.97 3.277  108 [CO2][H2O]0.5[O2]0.25

scheme was applied. The solution was converged when the summa- Table 3
Different grids for furnace modeling.
tion of the residual in the governing equation at all the domain
nodes was less than 0.01%. The ideal gas law was used to determine Grid name Number of cells Percentage of increasing
the gas density and the pressure staggered scheme (PRESTO) was 1 Grid-1 935,805 –
applied. Thermal conductivity, viscosity, and specific heat of the 2 Grid-2 1,095,555 +17
gas were determined as a mass fraction-weighted average of all 3 Grid-3 1,335,180 +43
the species. For each species, the specific heat was determined
through the application of a temperature piecewise polynomial
fit. To obtain more accurate results, the second-order upwind dis- 2000
cretization scheme was applied for the solution of momentum, spe-
cies and energy equations. A constant temperature boundary 1800 Grid-1
Grid-2
condition was applied to all furnace walls, 538.6 K, and the combus- Grid-3
1600
tion is conducted under atmospheric conditions. Methane (CH4)
Temperature (K)

was used as the operating fuel with a rate of 0.98 kg/s and an excess 1400
air factor of 1.16% was applied during the combustion process. For
1200
each burner, the primary and secondary air flows were set to
8.87 kg/s and 10.7 kg/s, respectively. 1000
The solution of the RTE cannot be obtained unless the absorp-
tion coefficient is known. It can be determined from the computed 800
emissivity at each computational cell by employing the Beer–Lam-
600
bert’s law and taking the mean beam length of the geometry as the
path length. SGG model, WSGG model, the EWBM, Leckner model 400
and Perry model are applied in the solution. They are coded and
200
compiled as User-defined functions (UDFs). Among these models,
0 1 2 3 4 5 6 7
the EWBM provided close results to the experimental data for
Distance along Z-axis (m)
the present boiler combustion application. The standard k–e model
is a semi-empirical model based on model transport equations for Fig. 3. Temperature profile at line (x = 4.539, y = 2.135) along Z-axis.
the turbulence kinetic energy (k) and its dissipation rate (e). The
model transport equation for k is derived from the exact equation, wall treatment. However, in the present combustion modeling,
while the model transport equation for e was obtained using phys- the most important zone is the close region to the burner not that
ical reasoning and bears little resemblance to its mathematically zone close to the wall. Based on accurate calculations of the turbu-
exact counterpart. On the other hand, for the realizable model, lence interaction and the reaction rates in this zone, the whole
the term ‘‘realizable’’ means that the model satisfies certain math- combustion characteristics can be properly calculated. Based on
ematical constraints on the normal stresses. The RNG model ac- that, the standard k–e model was applied in all of the present
counts for the flow curvature and considers additional terms for calculations.
M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97 89

-6.0E+04 present work for the calculations of the reaction kinetics. The mod-
-8.0E+04 el was modified to handle the increased CO2 concentration under
Grid-1 oxy-fuel conditions. The two step reaction kinetics model has been
-1.0E+05 Grid-2
considered in this work for simplification of the reactions kinetics
Total Heat Flux (W/m2)

-1.2E+05 Grid-3
calculations with reasonable results regarding the species concen-
-1.4E+05 trations as compared to the detailed model [41]. For this model, the
-1.6E+05 modified reactions rates data by Andersen et al. [41] are listed in
-1.8E+05 Table 2.
-2.0E+05
3.4. Grid independence tests
-2.2E+05
-2.4E+05 It is a very important issue in the numerical solution to ensure
-2.6E+05 that the obtained results are grid independent. For this purpose,
-2.8E+05 three different discretized geometries were examined in this study.
-3.0E+05 They are listed in Table 3. It is found that the deviation of the re-
1 2 3 4 5 6 sults with increasing the level of discretization above 1,335,180
Distance along Z-axis (m) cells did not exceed 0.01%. The temperature profiles along Z-axis
for the three grids were plotted in Fig. 3. The maximum variation
Fig. 4. Total heat flux profile at line (x = 0, y = 3) along Z-axis. in temperature between Grid-1 and Grid-3 is calculated to be
1.6% however; it is reduced to 0.2% in case of Grid-3 as compared
3.3. Reaction kinetics model to Grid-2. Furthermore, total heat flux has been checked in the
analysis. The profiles of the total heat flux along Z-axis at front wall
CFD modeling in combustion application using the global mech- and y = 3 m are plotted in Fig. 4. The difference between Grid-3 and
anisms can be found in the literature [41,6,42]. The simplest oxida- Grid-1 is found to be 0.9%. The deviation of 0.4% is computed for
tion mechanism assumes that products of the chemical reactions Grid-2 related to Grid-3. It should be mentioned from the above
consist only of CO2 and H2O and is used in case of air combustion. discussion that the variation in the results if Grid-3 is used in the
The physical situation that this assumption reasonably approxi- solution will be insignificant.
mates is a diffusion flame suspended in the flow field at relatively
high temperature. Here, the oxygen diffuses across the boundary 4. Results and discussions
layer to meet methane in a small well-defined reaction zone. The
high temperature of the flame accelerates the local homogeneous 4.1. Models validation
kinetics, and mostly produces CO2 and H2O. The replacement of in-
ert N2 with a chemically reactive compound, CO2, in case of oxy- Validation of the computational models applied in this study is
fuel combustion influences the importance of some of the elemen- obtained through comparisons with the measurement of Anders-
tary reactions governing the combustion and requires a modifica- son experimental work for air–fuel and oxy-fuel combustion
tion of the global multistep reaction mechanisms to make them [22,44]. The experiment was conducted on the100 kW oxy-fuel
valid under oxy-fuel conditions [43]. The modified two-step hydro- combustion test unit that connected to the Chalmers 12 MW re-
carbon oxidation mechanism by Andersen et al. [41] is used in the search boiler. The combustor is down-fired with a cylindrical

Fig. 5. Combustion chamber of the Chalmers 100 kW oxy-fuel unit with details of the burner [26].
90 M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97

2000
EXP
1900 EWBM
SGG(0.2 1/m)
1800 SGG(0.4 1/m)
WSGG[Smith(1982)]
1700
Temperature (K)

1600
1500
1400
1300
1200
1100
1000
900
0 0.1 0.2 0.3 0.4
Radial distance from center line (m)
Fig. 9. CO2 concentration at distance 0.384 m from the burner, OF21 combustion.
Fig. 6. Temperature profile at distance 0.384 m from the burner with different
radiation models, air combustion.

2000

0.2
EXP 1900
0.175 EWBM
Temperature (K)
0.15
CO2 concentration

1800
0.125

0.1 1700
EWBM
SGG(0.1 1/m)
0.075 SGG(0.18 1/m)
1600 SGG(0.25 1/m)
WSGG
0.05

0.025 1500
0 1 2 3 4 5 6 7
0
0 0.1 0.2 0.3 0.4 Distance along Z-axis (m)
Radial distance from center line (m) Fig. 10. Temperature profile at line (x = 3.264, y = 1.0675) along Z-axis, air
combustion.
Fig. 7. CO2 concentration at distance 1.4 m from the burner, air combustion.

1450 1700
EXP
1400 EWBM
SGG(0.2 1/m)
SGG(0.4 1/m)
1350 WSGG[Andersson(3+1)] 1600
WSGG[Andersson(4+1)]
Temperature (K)

1300 WSGG[Smith(1982)]
Temperature (K)

SGG(0.31 1/m)

1250 1500

1200
EWBM
1150 1400 Leckner
Perry
1100 SGG(0.1 1/m)
SGG(0.18 1/m)
1050 1300
SGG(0.25 1/m)
WSGG4
1000

950 0 1 2 3 4 5 6 7
0 0.1 0.2 0.3 0.4 Distance along Z-axis (m)
Radial distance from center line (m)
Fig. 11. Temperature profile at line (x = 3.264, y = 1.0675) along Z-axis, OF21
Fig. 8. Temperature profile at distance 0.553 m from the burner with different combustion.
radiation models, OF21 combustion.

combustion test unit with details of the burner. Three test condi-
refractory lined reactor which measures 0.8 m in inner diameter tions were considered in their work. The parameters of these three
and 2.4 m in inner height. Fig. 5 illustrates the schematic of the cases are: (1) Air-firing (O2 @ 21% Vol and N2 @79% Vol), (2) OF21
M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97 91

1800
EWBM
Leckner
1700 Perry
SGG(0.1 1/m)
SGG(0.18 1/m)
Temperature (K)

SGG(0.25 1/m)
1600 WSGG4

1500

1400

1300

1200
0 1 2 3 4 5 6
Distance along Z-axis (m)

Fig. 12. Temperature profile at line (x = 4.539, y = 4) along Z-axis, OF27 combustion.

Fig. 13. Temperature profiles at line (x = 3.264, y = 1.0675) along Z-axis for
different combustion cases, EWBM.

Fig. 14. Specific thermal capacity profiles at line (x = 3.264, y = 1.0675) along Z-
axis for different combustion cases, EWBM.

(O2 @ 21% Vol and CO2 @79% Vol) and (3) OF27 (O2 @ 27% Vol and
CO2 @73% Vol). All tests were performed for propane gas fuel
(C3H8) burnet at an excess air factor of 1.15 (15% excessive air).
Fig. 15. Contours for temperature of vertical plane passing through the middle
The feed gas temperature is between 25 and 30 °C. Heat input is burners 2 and 5 (x = 4.539 m), EWBM.
92 M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97

Fig. 17. Contours for CH4 mass fraction of vertical plane passing through the middle
burners 2 and 5 (x = 4.539 m), EWBM.

adjusted to be 80 kW h. The fuel mass flow rate through the fuel


lance is 1.74 g/s. flows having volume rates of 37 m3/h and
Fig. 16. Contours for temperature of horizontal plane passing through the upper 54 m3/h are fed through primary and secondary air lances, respec-
burners (y = 0 m), EWBM. tively. The total heat absorption is 48.7 kW in the air case and
M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97 93

Fig. 18. Turbulent viscosity profiles at line (x = 3.264, y = 1.0675) along Z-axis for
different combustion cases, EWBM.

45.8 kW for the OF21 case. Fig. 6 illustrates radial temperature dis-
tribution at distance 0.384 m from the burner. It is found that the
temperatures are over-estimated by all models. The maximum dis-
crepancy is about 20%. The effective absorption coefficient of 0.4
(1/m) predicts higher temperature by 17% compared to the mea-
surements whereas the absorption coefficient of 0.2 (1/m). Those
differences in the results between experimental and numerical
model using the SGG radiation model may be attributed to the cal-
culations of the absorption coefficient as shown in the figure. Basi-
cally, an optimal value of SGG model effective absorption
coefficient for a given situation can be found by trial and error by
comparing the numerical results obtained using the SGG model
with experimental data. Thus, the SGG radiation model absorption
coefficient can be optimized using the experimental data. The com-
parison in Fig. 6 supports this concept. The experimental data are
in good agreement with the EWBM except in the region near the
center of the furnace. The difference between the EWBM and the
WSGG model is less than 1%. In addition, CO2 concentrations are
considered in the validation, they are under-predicted in numerical
solution. However, the difference is about 3% as shown in Fig. 7. In
general, temperature, velocity and species concentration can be
predicted using CFD models with error of about 20% in the air–fuel
combustion case.
In case of OF21 combustion, Fig. 8 shows the temperature pro-
file at distance 0.553 m from the burner. It is found that the tem-
perature predicted by the EWBM is above the measured values
by 3.8%. In addition, three grey one clear gas WSGG model, which
is introduced by Andersson and Johansson [22], leads to tempera-
ture results with error of 4%. The absorption coefficient of 0.2 (1/m)
over-predicts the temperature by 7.5%. The more accurate results
are obtained if the absorption coefficient of 0.4 (1/m) is used in
the solution, the difference will not exceed +4%. As discussed in
description part for the basic assumptions for the SGG radiation
model, the optimal value of the absorption coefficient for any
one situation is not generally optimal for another. That’s why there
are fewer differences in comparison with the experimental data in
Fig. 3 as compared to those large differences in Fig. 6. The WSGG
model over-predicts the temperature by 5.5%. It is noted in Fig. 9
that the computed CO2 concentrations are in good agreement with
experimental data and the error less than 5%.
In addition, a comparison between experimental and numerical
results was made using the same radiation model, EWBM, in our
previous work (see Ref. [36]). In this work [36], an atmospheric dif-
fusion oxy-combustion flame in a gas turbine model combustor Fig. 19. Contours for CO2 mass fraction of vertical plane passing through the middle
has been investigated experimentally and numerically. burners 2 and 5 (x = 4.539 m), EWBM.
94 M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97

Oxy-combustion and emission characterization, flame stabilization


and oxy-combustion model validation analyses were the main
goals of this research work. The combustor was fuelled with CH4
and a mixture of CO2 and O2 as oxidizer. The same modified two-
step oxy-combustion reaction kinetics model for methane-oxygen
combustion has been used like the present study and the same
radiation model, EWBM, was applied in order to predict accurately
the oxy-combustion characteristics. Both experimental and
numerical results were in a very good agreement.

4.2. Performance of gas radiation models

It was found, in the air–fuel combustion as shown in Fig. 10,


that the weighted sum of grey gas model (WSGG) introduced by
Smith et al. [45] predicts higher temperatures by 0.39% compared
to the EWBM. The SGG model with absorption coefficient equal to
0.25 (1/m) under-predicts the temperature by 4.5% while the
absorption coefficient of 0.18 (1/m) leads to lower results by
3.3%. The deviation in the temperature is reduced to 0.37% in case
of the absorption coefficient has a value of 0.1 (1/m). In the case of
OF21 combustion, Fig. 11 represents the temperature profiles at
line located between first and second columns of burners in the
mid-way between the rows of burners along Z-axis (i.e.
x = 3.264 m, y = 1.0675 m), the results of four grey one WSGG
model developed by Andersson and Johansson et al. [22] are in
good agreement with those obtained in case of the EWBM. It
over-predicts the temperature by 1.2%. Leckner model results in in-
creased temperature values by about 7.9% whereas Perry model
over-estimates the temperature by 2.8%. Simple grey gas model
with absorption coefficient of 0.18 (1/m) exhibits more accurate
results with difference of less than 0.89%. The results of simple grey
gas model with the absorption coefficient equal to 0.25 (1/m) are
under-predicted by 2.1%. In addition, temperature distributions
at x = 4.539 m and y = 4 m along Z-axis are shown in Fig. 12. In
the same trend, Leckner model, Perry model, and WSGG model
over-predict the temperature by 15%, 5% and 1.2% respectively.
Simple grey gas model leads to error of 2.8% when absorption
coefficient of 0.25 (1/m) is used. This error is reduced to 1.35%
in case the absorption coefficient equal to 0.18 (1/m). It should
be emphasized that the maximum difference in the result when
the absorption coefficient is selected from the range of 0.25 (1/
m) to 0.1 (1/m) will not exceed 4.3%.

4.3. Characteristics of oxy-fuel and air–fuel combustion processes

The effects of replacing N2 by CO2 as in the oxy-fuel combustion


are discussed for this boiler in this section. It is clear from the tem-
perature profiles in Fig. 13 that the temperatures are lower in the
case of oxy-fuel combustion compared to air–fuel combustion;
reduction reaches up to 21% for OF21 and 18% for OF27. It is
thought that the reason of this reduction is attributed to the higher
thermal capacity of CO2 compared to N2. Fig. 14 demonstrates the
specific thermal capacity profiles at line (x = 3.264, y = 1.0675)
along Z-axis for the air–fuel combustion and the oxy-fuel combus-
tion cases, it indicates higher specific thermal capacity in the case
of the oxy-fuel combustion in comparison to that of air–fuel com-
bustion. This explains the lower temperature levels in oxy-fuel
combustion. Increasing the CO2 recirculation percentage results
in higher specific thermal capacity in the combustion products
and, thus, resulting in lower temperature levels. In addition to this
reason, the radiative emissivity and absorptivity of the CO2 is high-
er when it is compared to N2 which is transparent to the radiation
energy and because of this, flame cooling effects exhibited in the
oxy-fuel combustion cases.
Figs. 15 and 16 show a comparison between the temperature Fig. 20. Contours of radiation heat flux for air–fuel and oxyfuel combustion at the
contours of the air–fuel case and the two oxy-fuel combustion front, top and left walls of the furnace, EWBM.
M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97 95

cases. The contours are taken in a vertical plane passing through


the middle burners (x = 4.539 m) and horizontal plane passing
through the upper burners (y = 0 m). It is clear from the figures
that the temperature levels in the oxy-fuel combustion cases are
lower in comparison to the air–fuel combustion case. As ex-
plained above, the specific heat capacity in the oxy-fuel com-
bustion is higher than that of air–fuel combustion. However,
as CO2 mass percentage increases, the temperature levels are re-
duced, as indicated by Figs. 15b and 16b. It is noted that the
combustion rate is slower in the case of oxy-fuel combustion
as compared to that of air-fired combustion, and this leading
to extending the burning zone causing the flame length to in-
crease. It is also shown that the temperature rise is slower in
the oxy-fuel combustion cases in comparison to air–fuel com-
bustion and this is attributed to the higher specific heat of
CO2. In Fig. 16 and for burner number 3, left burner, the flame
is extended due to the effects of the exit flow. The left burner is
the closest burner to the exit section of the boiler and that’s
why the flame is extended and the combustion is delayed. In
the same figure and for the right burner, burner number 2,
it’s far away from the exit section and the flow velocity around
this burner is small. This may justify the short and wide flame
for the right burner.
The fuel burning rate is also affected by replacing N2 by CO2.
Fig. 17 shows the contours of CH4 concentration in the air–fuel
combustion and oxy-fuel combustion. It is noted that the fuel
consumption rate is slower in the oxy-fuel combustion cases.
Fig. 17b indicates further reduction in fuel burning rate as the
CO2 percentage is increased. In order to clarify the influence
of oxy-fuel combustion on the burning rate, the turbulent vis-
cosity for the cases of air–fuel and oxy-fuel combustion were
computed and are presented. Fig. 18 shows the turbulent viscos-
ity profiles at line (x = 3.264, y = 1.0675) along Z-axis for the
three combustion cases. It is noted that the higher turbulent
viscosity and hence higher mixing rates in the case of the
oxy-fuel combustion in comparison to those of air–fuel combus-
tion accelerates the reduction in flame temperature and there-
fore decreases the fuel consumption rate (i.e. CO2 acts as a
cooling gas). This explains the lower levels of temperature in
the oxy-fuel combustion. In our previous work [36], an oxy-
combustion gas turbine combustor flame has been characterized
in details through the measurement of by measuring the ex-
haust gas temperatures and species concentrations and compar-
ing them with those from the numerical model. In this work, it
is found that the combustion stability is improved with increas-
ing the percentage of O2 at inlet however there is a limitation in
temperature (see Ref. [35] for more details).
Fig. 19 shows the contours of the CO2 concentrations of vertical
plane passing through the middle burners 2 and 5 (x = 4.539 m).
The CO2 concentrations rise from around 14% in the case of air–fuel
combustion to above 80% in the oxy-fuel combustion in most of the
furnace volume. Figs. 20 and 21 show the contours of total heat
flux (radiative + convective) and radiation heat flux for air–fuel
and oxy-fuel combustion at the front, top and left walls of the fur-
nace. The negative sign of the heat flux means the flux from the
walls to the adjacent gases. It is found that the radiation heat flux
represents more than 80% of the total heat flux in the furnace in
case of air–fuel combustion and about 70% in oxy-fuel combustion
and this is justifying the importance of selecting the appropriate
gas radiation models in order to quantify its value. Finally, it is
found that the total heat flux on the walls in case of air–fuel com-
bustion is higher than that of oxy-fuel combustion (OF21 and
OF27) and this is due to the higher temperature in the air–fuel
combustion. Total heat flux profiles on the front wall (x = 0) and
y = 1.0675 along Z-axis for air–fuel and oxy-fuel combustion are Fig. 21. Contours of total heat flux for air–fuel and oxyfuel combustion at the front,
plotted in Fig. 22. top and left walls of the furnace, EWBM.
96 M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97

[3] Hirotatsu W, Yoshikazu S, Yohsuke M, Yoshio M, Hideyuki O, Shoji T, et al.


Spray combustion simulation including soot and NO formation. Energy
Convers Manage 2007;48:2077–89.
[4] Shahad HAK. An experimental investigation of soot particle size inside the
combustion chamber of a diesel engine. Energy Convers Manage
1989;29:141–9.
[5] Haroun AK, Yassar KA. Investigation of soot formation and temperature field in
laminar diffusion flames of LPG–air mixture. Energy Convers Manage
2000;41:1897–916.
[6] Bidi M, Hosseini R, Nobari MRH. Numerical analysis of methane–air
combustion considering radiation effect. Energy Convers Manage
2008;49:3634–47.
[7] Kennedy IM. Models of soot formation and oxidation. Prog Energy Combust Sci
1997;23:95–132.
[8] Yeoh GH, Yuen RKK, Lo SM, Chen DH. On numerical comparison of enclosure
fire in a multi-compartment building. Fire Safety 2003;38:85–94.
[9] Luo M, Beck V. Fire environment in a multi-room building. Fire Safety
1994;23:413–38.
[10] Moss JB, Stewart CD, Young KJ. Modeling soot formation and burnout in a high
temperature laminar diffusion flame burning under oxygen-enriched
conditions. Combust Flame 1995;101:491–500.
[11] Syed KJ, Stewart CD, Moss JB. Modelling soot formation and thermal radiation
in buoyant turbulent diffusion flames. In: Proceedings of the 23th symposium
on combustion. The Combustion Institute; 1990. p. 1533–41.
[12] Wall T, Liu Y, Spero C, et al. An overview on oxy-fuel coal combustion State of
Fig. 22. Total heat flux profiles at line (x = 0, y = 1.0675) along Z-axis for different the art research and technology development. Chem Eng Res Des
combustion cases, EWBM. 2009;87:1003–16.
[13] Kitto JB, Stultz SC. Steam: Its Generation and Use. The Babcock and Wilcox
Company; 2005. p. 4–14.
5. Conclusion
[14] Morris WJ, Yu D, Wendt JOL. Soot, unburned carbon and ultrafine particle
emissions from air- and oxy-coal flames. Proc Combust Inst 2011;33:3415–21.
Different gas radiation models were used in CFD modeling of [15] Schiemann M, Scherer V, Wirtz S. Optical coal particle temperature
the typical industrial water tube boiler and their influence on the measurement under oxy-fuel conditions: measurement methodology and
initial results. Chem Eng Technol 2009;32:2000–4.
results was investigated. It is found that Leckner model and Perry [16] Lallemant N, Sayre A, Weber R. Evaluation of emissivity correlations for H2O–
model over-predict the temperature compared to the exponential CO2–N2/air mixtures and coupling with solution methods of the radiative
wide band model. Weighted-sum-of-grey-gases model can predict transfer equation. Progr Energy Combust Sci 1996;22:543–74.
[17] Hottel H, Sarofim A. Radiative transfer. New York: McGraw-Hill, Inc.; 1967.
accurate results compared to the benchmark model, however, it re- [18] Wall TF. Combustion processes for carbon capture. Proc Combust Inst
quires new parameters for different ratios of H2O and CO2. It is a 2007;31:31–47.
computationally efficient option for CFD simulation where there [19] Zheng L, Clements B, Runstedtler A. A generic simulation method for the lower
and upper furnace of coal-fired utility boilers using both air firing and oxy-fuel
is a need for a simple and accurate gas radiation model. One of combustion with CO2 recirculation. In: 27th international technical conference
the limitations of Weighted-sum-of-grey-gases model and Perry on coal utilization and fuel systems, Clearwater, Florida; 2002.
model is that they are not applicable for pressures other than [20] Zheng L, Clements B, Douglas MA. Simulation of an oxy-fuel retrofit to a typical
400 MWe utility boiler for CO2 capture. In: 26th international technical
one atmosphere. The exponential wide band model and Leckner
conference on coal utilization and fuel systems, Clearwater, Florida; 2001.
model are valid for use for various pressure and ratios of CO2 and [21] Leckner B. Spectral and total emissivity of water vapor and carbon dioxide.
H2O. The influence of these models on the prediction of gases con- Combust Flame 1972;19:33–48.
[22] Andersson K, Johnsson F. Flame and radiation characteristics of gas-fired O2/
centrations is very limited. An optimal value of the absorption
CO2 combustion. Fuel 2007;86:656–68.
coefficient can be determined by comparing the results to the more [23] Johansson R, Andersson K, Leckner B, Thunman H. Models for gaseous radiative
accurate model (i.e. EWBM). The influence of the CO2 circulation on heat transfer applied to oxy-fuel conditions in boilers. Int J Heat Mass Transfer
the oxy-fuel combustion characteristics was investigated. It is 2010;53:220–30.
[24] Porter R, Liu F, Pourkashanian M, Williams A, Smith D. Evaluation of solution
found that the temperature levels are reduced in the case of oxy- methods for radiative heat transfer in gaseous oxy-fuel combustion
fuel in comparison to the air–fuel combustion case. It is also found environments. J Quant Spectrosc Radiative Transfer 2010;111:2084–94.
that higher specific thermal capacity in the combustion products [25] Johansson R, Leckner B, Andersson K, Johnsson F. Account for variations in the
H2O to CO2 molar ratio when modeling gaseous radiative heat transfer with
are obtained as the CO2 recirculation is increased. Thus, lower tem- the weighted-sum-of-grey-gases model. Combust Flame 2011;158:893–901.
perature levels are exhibited in the case of oxy-fuel combustion. [26] Hjärtstam S, Johansson R, Andersson K, Johnsson F. Computational fluid
The fuel consumption rates are slower in the oxy-fuel combustion dynamics modeling of oxy-fuel flames: the role of soot and gas radiation.
Energy Fuels 2009. ef200983e.
case in comparison to the air–fuel combustion case. It is also [27] Chandrasekhar S. Radiative transfer. New York, NY: Dover Publications; 1960.
shown that the energy absorbed is much higher in the case of [28] Modest MF, Zhang H. The full-spectrum correlated-k distribution and its
the air–fuel combustion along the surfaces. relationship to the weighted-sum-of gray- gas method. In: Proceedings of the
2000 IMECE, vol. HTD-366-1, ASME, Orlando, FL; 2000. p. 75–84.
[29] Lallemant N, Weber R. A computationally efficient procedure for calculating
gas radiative properties using the exponential wide band model. Int J Heat
Acknowledgements Mass Transfer 1996;39:3273–86.
[30] Siegel R, Howell J. Thermal radiation heat transfer. 4th ed. New York: Taylor
The authors wish to acknowledge the support received from and Francis; 2002.
[31] Modak AT. Radiation from products of combustion. Fire Res 1979;1:339–61.
King Abdulaziz City for Science and Technology (KACST) through
[32] Hottel HC, Noble JJ, Sarofim AF. Heat and mass transfer. In: Perry’s chemical
The Vice-Rectorship of Applied and Scientific Research at KFUPM engineers’ handbook. New York: McGraw-Hill; 2007 [chapter 5].
under KACST project# AT-29-89. [33] Grosshandler WL. RADCAL: a narrow-band model for radiation calculations in
a combustion environment. NIST Technical Note 1402; 1993.
[34] Truelove JS. A mixed grey gas model for flame radiation. United Kingdom
Atomic Energy Authority, Report AERER-8494. Harwell; 1976.
References [35] Fluent 6.2 User’s Guide. Fluent Inc. Center Research Park, 10 CavendishCourt,
Lebanon, NH 03766, USA; 2003.
[36] Nemitallah MA, Habib MA. Experimental and numerical investigations of an
[1] Viskanta R. Radiative transfer in combustion systems: fundamentals &
atmospheric diffusion oxy-combustion flame in a gas turbine model
applications. Purdue University, USA.
combustor: oxy-combustion and emission characterization, flame
[2] Xiaobei C, Liang C, Fangqin Y, Shijun D. Study on soot formation characteristics
stabilization and model validation. Appl Energy 2013;111:401–15.
in the diesel combustion process based on an improved detailed soot model.
Energy Convers Manage 2013;75:1–10.
M.A. Rajhi et al. / Energy Conversion and Management 81 (2014) 83–97 97

[37] Ben Mansour R, Nemitallah MA, Habib MA. Numerical investigations of oxygen [41] Andersen J, Rasmussen CL, Giselsson T, Glarborg P. Global combustion
permeation and methane-oxycombustion in a stagnation flow ion transport mechanisms for use in CFD modeling under oxy-fuel conditions. Energy
membrane reactor. Energy 2013;1:1–11. Fuels 2009;23:1379–89.
[38] Nemitallah MA, Habib MA, Ben Mansour R. Investigations of oxy-fuel _
[42] Ilker _ Cevahir T. Effect of turbulence and radiation models
Y, Murat T, Mustafa I,
combustion and oxygen permeation in an ITM reactor using a two-step oxy- on combustion characteristics in propane–hydrogen diffusion flames. Energy
combustion reaction kinetics model. J Membr Sci 2013;432:1–12. Convers Manage 2013;72:179–86.
[39] Ben-Mansour R, Habib M, Badr H, Uddin A, Nemitallah MA. Characteristics of [43] Glarborg P, Bentzen LL. Chemical effects of a high CO2 concentration in oxy-
oxy-fuel combustion in an oxygen transport reactor. Energy Fuels fuel combustion of methane. Energy Fuels 2007;22:291–6.
2012;26:4599–606. [44] Andersson K. Characterization of oxy-fuel flames – their composition,
[40] Habib MA, Nemitallah MA, Ben-Mansour R. Recent development in oxy- temperature and radiation. PhD. thesis, Chalmers University of Technology,
combustion technology and its applications to gas turbine combustors and Sweden; 2007.
ITM reactors. Energy Fuels 2013;27:2–19. [45] Smith TF, Shen ZF, Friedman JN. Evaluation of coefficients for the weighted
sum of gray gases model. J Heat Transfer 1982;04:133–50.

Вам также может понравиться