Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Fuel Processing Technology 92 (2011) 1489–1497

Contents lists available at ScienceDirect

Fuel Processing Technology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / f u p r o c

CFD modelling of oxy-coal combustion in an entrained flow reactor


L. Álvarez a, M. Gharebaghi b, M. Pourkashanian b, A. Williams b,⁎, J. Riaza a, C. Pevida a, J.J. Pis a, F. Rubiera a
a
Instituto Nacional del Carbón, CSIC, Apartado 73, 33080 Oviedo, Spain
b
Centre for Computational Fluid Dynamics, School of Process, Environmental and Materials Engineering, University of Leeds, Leeds LS2 9JT, UK

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

Article history: Oxy-fuel combustion is seen as one of the major options for CO2 capture for both new and existing coal fired
Received 21 December 2010 power stations. Coal is burned with a mixture of oxygen and recycled flue gas to obtain a rich CO2 stream
Received in revised form 18 March 2011 ready for sequestration. Computational fluid dynamics (CFD) tests for coal combustion under different O2/CO2
Accepted 22 March 2011
(21–35% vol O2) atmospheres in an entrained flow reactor (EFR) were carried out using three coals of different
Available online 16 April 2011
volatile matter content. The temperature profiles, burning rates, burnout and concentration of major species,
Keywords:
such as O2, CO2, CO, were predicted and compared with an air reference case. A decrease in gas temperature
Computational fluid dynamics and burning rate was observed for 21% O2/79% CO2 environment in comparison to the air reference case due to
Oxy-fuel combustion the difference in gas properties between N2 and CO2. Experimental coal burnouts obtained in the EFR, were
Entrained flow reactor used to test the accuracy of the CFD model. The numerical results showed a decrease in coal burnout when N2
was replaced by CO2 for the same oxygen concentration (21%), but an improvement in the O2/CO2 atmosphere
for an oxygen concentration higher than 30%. The numerical results for oxy-coal combustion were in good
agreement with the experimental results.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction that the reaction rate of char with oxygen is much higher than that of
char with CO2 and with H2O at flame temperatures. They are therefore
Oxy-fuel combustion is one of the most promising CO2 capture often ignored in the simulations for pulverised fuel combustion, and
technologies as it could be adapted to both new and existing there is some evidence for this even under flame-staging condi-
pulverised coal-fired power stations [1–4]. During oxy-fuel combus- tions [11], since under these conditions some regions have low
tion a mixture of O2 and recycled flue gas (mainly CO2 and H2O) is oxygen concentrations where char gasification reactions might take
used for fuel combustion. The gas composition inside the boiler differs place in addition to normal combustion. The objective of this work is
greatly from that employed in conventional air combustion. Oxy-fuel to evaluate the effect of replacing N2 by CO2 on char burnout and
combustion differs from air combustion in several ways, including gaseous species during coal combustion by means of a CFD model.
heat transfer, flame ignition, coal burnout and pollutant formation Experimental results obtained in an entrained flow reactor (EFR), a
[2,3]. Several studies have been carried out in order to obtain data for a heated vertical tube reactor with a laminar flow of hot gas into which
fundamental interpretation of the oxy-fuel combustion process or for a stream of pulverised coal is injected, were employed to test the CFD
the purpose of studying the entire oxy-fuel process in an industrial model.
relevant size [3–6].
Computational Fluid Dynamics (CFD) models are powerful 2. Experimental cases considered
predictive tools in combustion research and they have been widely
used to simulate combustion in coal-fired power stations. A CFD Three coals of different rank, HVN, (Hullera Vasco Leonesa, Spain),
model could be an aid for understanding combustion characteristics DAB (Datong, China) and BA (Batán, Spain), were used in the
when retrofitting a pulverised fuel boiler designed for air firing to oxy- combustion experiments. The proximate and ultimate analyses and
fuel firing [6]. Thus, for example, modelling studies on flame the high heating values of the coals are presented in Table 1. The
aerodynamics [7] and heat transfer [8] in oxy-fuel combustion proximate analysis was conducted on a LECO TGA-601, and the
processes have been carried out. There is still debate on whether ultimate analysis on a LECO CHNS-932. The coals were ground and
high concentrations of CO2 and H2O in the combustion gases would sieved to obtain a 75–150 μm particle size fraction.
increase or decrease burnout. There is experimental evidence [9,10] The experimental data were obtained using an EFR, whose details
have been previously reported [12,13]. The EFR has an internal
diameter of 4 cm and a length of 200 cm, where the reaction zone is
⁎ Corresponding author. Tel.: +44 113 3432507; fax: +44 113 2467310. assumed up to a length of 1.42 m. Three binary gas mixtures of O2, and
E-mail address: fueaw@leeds.ac.uk (A. Williams). CO2 (21% O2/79% CO2, 30% O2/70% CO2 and 35% O2/65% CO2) were

0378-3820/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.fuproc.2011.03.010
1490 L. Álvarez et al. / Fuel Processing Technology 92 (2011) 1489–1497

Table 1
Nomenclature Proximate and ultimate analyses, high heating value of the coals.

a absorption coefficient Sample HVN DAB BA

Ai intrinsic pre-exponential factor (kg/m2 s) Origin Spain China Spain


Ap surface of the particle (m2) Rank sa hvb hvb

Ap, i specific internal surface area of the char (m2/g)


Proximate analysis
cd drag coefficient Moisture content (wt.%) 1.1 2.9 1.2
cp heat capacity of the particle (J/kg K) Ash (wt.%, db) 10.7 10.9 6.9
C1 mass diffusion limited rate constant V.M. (wt.%, db) 9.2 28.8 33.9
F.C. (wt.%, db)a 80.1 60.3 59.2
dp,0 initial diameter of the particle (m)
dp diameter of the particle (m) Ultimate analysis (wt.%, daf)a
D0 bulk molecular diffusion rate coefficient (m2/s) C 91.7 81.9 88.5
De effective diffusion coefficient in the particle pores (kg/ H 3.5 5.0 5.5
m2 Pa s) N 1.9 0.8 1.9
S 1.6 1.2 1.1
DKn Knudsen diffusion (m2/s)
Oa 1.3 11.1 3.0
Ei intrinsic activation energy (J/kg mol) High heating value (MJ/kg, db) 31.8 28.8 33.1
h convective heat transfer coefficient (W/m2 K)
sa: semi-anthracite; hvb: high-volatile bituminous coal.
I radiation intensity (W/m2) db: dry basis; daf: dry and ash free bases; mmf: mineral matter free basis.
ki intrinsic reactivity of the char (kg/m2 s) a
Calculated by difference.
mp,0 initial mass of the particle (kg)
mp mass of the particle (kg)
Mw, ox molecular weight of the oxidant (kg/kg mol)
n refractive index
rp mean pore radius of the char particles (m) removed by means of a cyclone and a filter, and the exhaust gases
Sb stoichiometry of the reaction were monitored using a battery of analysers (O2, CO, CO2 SO2, NOx).
SФ source term for the scalar k Burnout was determined using the ash tracer method [14].
Tp temperature of the particle (K)
T∞ local temperature of the continuous phase (K) 3. Modelling approach
uf fluid velocity (m/s)
up particle velocity (m/s) The reaction zone of the EFR has been constructed and meshed
U fractional degree of burnout using Gambit pre-processor. A three-dimensional structured grid
R chemical rate of the char (kg/m2 Pa s) consisting of ~ 75,000 cells is employed to describe a quarter of the
Yox local mass fraction of the oxidant in the gas total volume, an isotropic view of the geometry and grid system is
α mode of burning parameter shown in Fig. 2(a). The mesh is locally refined in the near injection
zone in order to enhance the prediction of the devolatilisation and
εp particle emissivity
η effectiveness factor initial combustion steps of the coal particles. A detail of the grid at the
injection zone is shown in Fig. 2(b). The grid density is found to be
θ particle porosity
θR radiation temperature (K) sufficient to obtain grid independent solutions.
A commercial CFD program, ANSYS Fluent version 12 [15] was
ρox density of the oxidant in the bulk gas (kg/m3)
ρp,0 initial density of the particle (kg/m3) used to simulate the combustion process in the EFR. The CFD model
was formulated using the Eulerian–Lagrangian frame of reference
ρp density of the particle (kg/m3)
σ Stefan-Boltzmann constant (5.67·10−8 W/m2 K4) meaning transport equations were solved for the continuous phase
and trajectories of particles were computed through the calculated
σs scattering coefficient
τ tortuosity gas field. There is constant exchange of momentum, mass and heat
between the combusting gas and the particles, and the exchange
Ґ diffusion coefficient
φ Thiele modulus between the two systems appear as source/sink terms in the
governing equations. For gas flow, the governing mass, momentum,
Ф phase function
Фk scalar k and energy equations can be typically represented by the following
scalar transport Eq. (1):
Ω′ solid angle
 
∂ρ Φk ∂ ρ ui Φk −Γk ∂Φ
∂xi
k

+ = SΦk k = 1; …; N ð1Þ
∂x ∂xi
used, and the air case (21% O2/79% N2) was taken as reference. The EFR
is electrically heated and preheated gases were introduced into the Ґ and SФ are the diffusion coefficient and relevant source terms
reactor through flow straighteners to ensure laminar flow conditions. for the N scalars (Фk) in the system. In the case of pulverised fuel
Coal was introduced through a cooled injector to guarantee a combustion, the source term SФ should include the contributions
temperature of less than 373 K before entering the EFR reaction from the solid particles, representing the exchange of mass,
zone. These coal particles were injected into the centre of the momentum, and heat between the two phases. The RNG k–ε
preheated gases stream. The experiments were performed at 1273 K turbulence model with default parameters is used to model the
and the gas flow rate was adjusted to 22.3 L/min (1273 K, 1 atm) in dynamic of the flow [16]. Although this model was developed for
order to obtain a particle residence time of 2.5 s in the EFR and the turbulent flows, some authors have employed it successfully for low
oxygen excess was set at 25%. Reynolds-number flows [17,18].
Reaction products were quenched by aspiration into a stream of In a pulverised fuel chamber, radiation is the dominant mode of
nitrogen using a water-cooled probe. This probe was inserted into the heat transfer. The gray radiative transfer equation (RTE) at a position r
reaction chamber from below as can be seen in Fig. 1. Particles were in the direction s for an absorbing, emitting, and scattering medium,
L. Álvarez et al. / Fuel Processing Technology 92 (2011) 1489–1497 1491

Feeding system

Pre-heater

Cooled injector

Mass flow
controllers
Reaction tube

Gas analysers

O2
NO
N2 O2 CO2 CO
SO2
CO2

Gas cylinders

Sampling probe To vent

Filter
Cyclone

Fig. 1. Schematic diagram of the entrained flow reactor (EFR) used in the experiments.

such as those in combustion systems, can be expressed as follows in gas mixture absorption coefficient is calculated by using the weighted
Eq. (2): sum of gray gases model (WSGGM). In the light of previous studies
[22], a value of 0.9 was taken for the emissivity of the combusting
dI ðr; sÞ an2 σT 4 σ   char. The scattering factor indicates the scattering properties of a
+ ða + σs ÞIðr; sÞ = + s ∫ Iðr; sÞ Φ s; s′ dΩ′ ð2Þ particle; 1 indicates total forward scattering (effectively treated as
ds π π 4π
transmission). In this work this factor was set to 0.6 in accordance
with previous work [23]. This value is still subject to discussion, some
where a and σs are the absorption and the scattering coefficient authors [24] suggest a value of 0.9 is more appropriate.
respectively, I is the radiation intensity and s′ is the scattering It is well-known that soot formation has impact on the radiative
direction. A review of the various radiation models used in heat flux in a furnace. To model soot formation, one-step Khan and
combustion systems for CFD applications in order to solve the RTE Greeves prediction model [25] based on the total volatile concentra-
can be found in [19]. The Discrete Ordinate (DO) and the P1 radiation tion was employed. This model is empirically-based and represents an
models have generally been chosen for pulverised fuel combustion for approximate model of soot formation in combustion systems. The
CFD applications [20,21]. In this work, Discrete Ordinate model is used combustion of soot is assumed to be governed by the Magnussen
because of higher accuracy and smaller optical length of the EFR. The combustion rates [26].
The coal particles in the furnace were tracked in the Lagrangian
frame of reference with the particle momentum equation. The
equation of motion for a representative coal particle is given by
Eq. (3). The particle temperature in the furnace was calculated taking
into account the heat transfer due to convection and radiation and the
heat generated by the char reaction. The heat balance equation is
expressed by Eq. (4):

duP cd ρp Ap  
mP = juf −up j uf −up ð3Þ
dt 2

dTp    
4 4
mp cp = hAp T∞ −Tp + εp Ap σ θR −Tp ð4Þ
dt

The initial density of the particles is assumed to be 1550 kg/m3 for


the semi-anthracite and 1400 kg/m3 for the high volatile coals, and a
specific heat of 1.68 J/kg K is assumed for all of the coals involved.
These values are included in ANSYS Fluent materials database and
they have been determined for a wide variety of coals. The particles
are assumed to be spherical and their size distribution is fitted to a
Rosin–Rammler distribution ranging from 75 to 150 μm, with an
average size of 115 μm.
The devolatilisation rate of each coal was modelled using a single
step first-order Arrhenius reaction. The kinetic parameters (A, Ea)
were obtained from previous work [27,28] using the FG-DVC code
Fig. 2. (a) Computational mesh used in the CFD model and (b) detail for the injection at [29]. A mixture fraction/PDF chemical equilibrium modelling method
the top of the vertical reactor. [30] was employed for volatiles combustion. In this approach,
1492 L. Álvarez et al. / Fuel Processing Technology 92 (2011) 1489–1497

individual component concentrations for the species of interest are Assuming that the pore size distribution is unimodal and the bulk
derived from the predicted mixture fraction distribution. PDF tables and Knudsen diffusion proceed in parallel, the effective diffusion, De,
for both air and oxy-fuel conditions using the pre-PDF pre-processor is given by Eq. (13) where D0 is the bulk molecular diffusion
were calculated for each coal. Twenty species including chemical coefficient and θ the porosity of the char particle.
species (CO2, O2, CO, H2O, H2, CH4, C2H2, and SO2) and radicals and  
intermediate species (C, H, O, N, S, OH, CS …) have been included. θ 1 1 −1
De = + ð13Þ
For char combustion, the Smith's intrinsic model [31] was adopted. τ2 D
Kn D0
This model assumes an order unity for the surface reaction with
oxygen and the effects of both bulk diffusion and chemical reaction For computational ease a value of 0.5 for porosity [23] was
are considered. The char combusting rate can be expressed as Eq. (5), assumed here. A value of 1.41 for the tortuosity of the pores was used;
and using Eq. (6) to compute the bulk molecular diffusion rate this value corresponds to an average intersecting angle between the
coefficient, D0: pores and the external surface of 45° [34]. DKn is the Knudsen diffusion
" # coefficient, and it is defined by Eq. (14).

dmP 2 ρ RT Y D0 R sffiffiffiffiffiffiffiffiffiffiffiffi
= πdp P ∞ ox ð5Þ
dt Mw;ox D0 + R P Tp
DKn = 97:0 rp ð14Þ
Mw;ox
n o
0:75
D0 = C1 ½ðTP + T∞ Þ= 2 = dP ð6Þ P
where r p is the mean pore radius of the char particles. Here a value of
6·10−8 m was chosen based on [32].
where C1 is the mass diffusion rate constant that was set at the value The boundary conditions, mass flow inlets and wall temperatures
of 5·10−12. are established using measurements made during the experimental
The chemical rate is expressed in terms of an intrinsic chemical runs. The coal feed rates and inlet conditions considered for each case
and pore diffusion rates as can be seen in Eq. (7) where η is the are shown in Table 2. Inlet mass flow rates were calculated through
effectiveness factor, or the ratio of the actual combustion rate to the experimental flow rates in a quarter of the total volume of the EFR,
rate attainable if no pore diffusion resistance existed [32]. and coal feed rates were calculated from the stoichometric values to
ensure an oxygen excess of 25%.
dp
R=η ρ A k ð7Þ
6 p p;i i
4. Results and discussion

An intrinsic activation energy (Ei) of 155 kJ/mol (±10 kJ/mol)


Simulations for three coals (HVN, DAB and BA) in air and defined
based on a previous study [27] was used for the reaction of carbon
oxyfuel environments have been performed. The objective of these
with molecular oxygen in the intrinsic burnout rate, in Eq. (8).
simulations is to gain an understanding of the combustion behaviour
  of different coals under oxy-fuel conditions for future implementation
ki = Ai exp −Ei = RTp ð8Þ [36].
Fig. 3(a) presents the temperature contours in the mid plane of the
The char particle size and density variation during combustion is EFR and Fig. 4 shows the area-averaged temperature profile during
specified through the burning mode that relates the diameter of the HVN combustion for all the cases.
char particle to the fractional degree of burnout [33,34] as can be seen As can be seen in Fig. 3(a) significant differences appear when N2 is
in Eqs. (9) and (10). replaced by CO2 for the same oxygen concentration (cases I and II) and
there is a temperature reduction due to the higher specific heat of CO2.
dp α This decrease can also be seen in Fig. 4, in the air reference case the
= ð1−U Þ ð9Þ
dp;0 predicted peak area-average temperature is about 1348 K whereas in
the oxy-fuel case it is about 1286 K. A slight increase is observed in
mp temperature contours in O2/CO2 conditions when the O2 concentra-
U = 1− ð10Þ tion is increased (cases III and IV). Regarding the peak area-average
mp;0
temperature, a value of 1300 K is obtained for 30% O2/70% CO2 and
1308 K for 35% O2/65% CO2.
A value of 0.25 for the burning model is used. This corresponds to a
Differences related to combustion behaviour between air and oxy-
decrease in both char particle size and density during coal burnout.
fuel conditions are also observed. Fig. 3(b), which shows the O2
The coal char combustion rate is related to the measured surface of the
concentration for all cases, and suggests that under oxy-fuel
char, there are correlations between the initial surface area of the
conditions combustion takes place closer to the injection zone. The
pyrolysed char and the fixed content of the parent coal [27]. However,
burning rates due to char combustion are shown in Fig. 3(c) for all
the specific internal surface area of the char defined as Ap, i is not a
cases. The predicted rates confirm that char combustion in O2/CO2
constant value during char combustion. The mean value of Ap, i is
starts earlier than in air, i.e., it is enhanced by the higher CO2
higher than that of the initial value of the char formed from pyrolysis
concentrations [37]. But the values of these burning rates, are lower
[32]. Here an estimated value of 300 m2/g was used for bituminous
for 21% O2/79% CO2 (case II) than for 21% O2/79% N2 (case I), which
coals, and 40 m2/g for semi-anthracite coals [35].
The effectiveness factor is determined by Eq. (11), where φ is the
Thiele modulus (Eq. (12)). Table 2
CFD inputs for the gases and coal feed rates.

3 Atmosphere Gas inlet Mass flow rate of coal (g/min)


η= ðϕ cothϕ−1Þ ð11Þ
ϕ (g/min)
HVN DAB BA

21% O2/79% N2 1.548 0.110 0.123 0.106


  21% O2/79% CO2 2.118 0.110 0.123 0.106
dp Sb ρp Ap;i ki ρox 1 = 2 30% O2/70% CO2 2.058 0.157 0.176 0.151
ϕ= ð12Þ 35% O2/65% CO2 2.016 0.182 – 0.175
2 De ρox
L. Álvarez et al. / Fuel Processing Technology 92 (2011) 1489–1497 1493

1500
HVN

1400

Temperature (K)
1300

21%O2/79%N2

1200 21%O2/79%CO2
30%O2770%CO2
35%O2/65%CO2
1100
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Distance (m)

Fig. 4. Predicted area-averaged temperature variation with axial location for HVN
combustion.

III and IV), the burning rate increases with the oxygen concentration,
resulting in an increase in temperature and oxygen consumption rate,
as can be seen in Fig. 3(a) and (b).
Fig. 5(a) shows the temperature contours in the mid plane of the
EFR during DAB and BA combustion and Fig. 6 show the area-averaged
temperature profile for all test conditions. The mid-plane gas
temperature profiles for DAB and BA are similar as they have similar
volatile matter contents. For DAB and BA, the predicted temperature
contours are higher than for HVN, as is the predicted peak
temperature. These differences can be explained due to the lower
volatile matter content of HVN, which releases less volatiles during
devolatilisation, so the temperature contours in the upper part of the
reactor, where volatiles combustion occurs, are lower. In the air
reference case, this temperature reaches a value of 1348 K for HVN
whereas for DAB it is about 1457 K and 1480 K for BA. Similar trends
are found in the oxy-fuel cases, with BA and DAB reaching higher
temperature than HVN under O2/CO2 conditions.
The oxygen concentration profiles and burnout rates of DAB and
BA also differ from HVN. These differences are due to the higher
amount of volatiles released by DAB and BA. The oxygen profiles for
HVN combustion are more uniform than those corresponding to DAB
and BA combustion, which are represented in Fig. 5(b). For DAB and
BA the oxygen is quickly consumed during the combustion of
volatiles, which is why their burning rate is also higher as seen in
Fig. 5(c).
For DAB and BA the predicted peak area-averaged temperatures
during air combustion are 1480 K and 1457 K respectively, whereas
for oxy-fuel combustion the peak temperatures are about 1312 K
and 1317 K. An increase in temperature is observed with oxygen
concentration under O2/CO2 conditions for both coals as shown in
Fig. 6. This increase is more pronounced than that observed for
HVN.
The differences in oxygen profiles and burnout rate as shown in
Fig. 5(b) and (c), suggest that devolatilisation under oxy-fuel
conditions also takes place before it does in air. After devolatilisation
occurs, the burnout rate is lower in 21% O2/79% CO2 than in 21% O2/
79% N2, indicating that no improvement occurs due to CO2 char
gasification in high volatile coals combustion either.
Residence times and particle heating rates in entrained flow
Fig. 3. (a) Predicted temperature (K), (b) O2 concentration (%) and (c) char burnout reactors are close to those occurring in industrial coal combustors; as
rate (kg/s) inside the entrained flow reactor during HVN (semianthracite) combustion
for each atmosphere: 21% O2/79% N2 (I), 21% O2/79% CO2 (II), 30% O2/70% CO2 (III) and
can be seen in [38], the numerically predicted particle residence time
35% O2/65% CO2 (IV). Length scale for (a), (b) and (c) is 40 cm. for most of the particles inside an industrial coal combustor is not
higher than 3–4 s. The entrained flow reactors enable more detailed
CFD models to be tested and improved. Fig. 7 shows the experimental
suggests that the higher CO2 concentrations do not improve the char burnouts for the combustion of the three coals (HVN, DAB and BA)
reaction rate in the temperature range of this study. Moreover, the studied in the EFR under both air and O2/CO2 conditions. A value of
predicted CO concentration is nearly zero, confirming that CO2 char 25% excess oxygen, a particle residence time of 2.5 s and a
gasification is extremely small. For the rest of the oxy-fuel cases (cases temperature of 1273 K were used in all the experiments.
1494 L. Álvarez et al. / Fuel Processing Technology 92 (2011) 1489–1497

Fig. 5. (a) Predicted temperature (K), (b) O2 concentration (%) and (c) char burnout rate (kg/s) inside the entrained flow reactor during DAB and BA (high volatile bituminous)
combustion for each atmosphere: 21% O2/79% N2 (I), 21% O2/79% CO2 (II), 30% O2/70% CO2 (III) and 35% O2/65% CO2 (IV). Length scale for (a), (b) and (c) is 40 cm.

For HVN the burnout obtained with the mixture 21% O2/79% CO2 is combustion decreases [39,40]. As a consequence the combustion rate
lower than that obtained in air. When N2 is replaced by CO2, the heat of the particles decreases and the coal burnout becomes worse. When
capacity of the gases increases and the particle temperature during the oxygen concentration in the O2/CO2 mixtures is increased to a
L. Álvarez et al. / Fuel Processing Technology 92 (2011) 1489–1497 1495

a 1500 rank coals like BA and DAB reach high burnout in air, so there is lesser
DAB
margin for improvement by increasing the oxygen concentration.
Under the experimental conditions of this work, an improvement
1400
Temperature (K)

in char burnout due to char-CO2 gasification was not expected.


Varhegyi et al. [43] found that the reaction rate of char-CO2 is slower
1300 than char-O2 at temperatures up to 1223 K. Other researchers [37,44]
have found that the reactivity of char in O2/CO2 conditions may be
21%O2/79%N2 increased due to the high CO2 concentrations at high temperatures
1200 (up to 1600 K).
21%O2/79%CO2
To validate the CFD model the experimental burnouts and also the
30%O2/70%CO2
gases concentration were employed. Fig. 8 presents a comparison
1100 between the experimental and predicted coal burnout results. In all
0 0.2 0.4 0.6 0.8 1 1.2 1.4 cases the predicted coal burnout follows the same trend as the
Distance (m) experimental results. Burnout in a mixture of 21% O2/79% CO2 is lower
than in air, but there is improvement when the oxygen concentration
b is 30% or higher for both experimental and numerical cases. The
1500
burnouts predicted with the commercial CFD model were higher due
BA
to the default input values that this code employs. However, when the
1400 char burnout sub-model was run using as input values the model and
Temperature (K)

devolatilisation kinetics parameters described in Section 3 and

1300
a
HVN
21%O2/79%N2 98.7 98.0 99.5 99.9
100
1200 21%O2/79%CO2 87.9
90.9
82.3 81 82.9
30%O2/70%CO2 79.5 79.7
80 77.1

35%O2/65%CO2
Burnout (%)
1100
60
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Distance (m) 40
Experimental
Fig. 6. Predicted area-averaged temperature variation with axial location for (a) DAB
and (b) BA combustion. 20 CFD predictions
Model predictions

0
21%O2/79%N2 21%O2/79%CO2 30%O2/70%CO2 35%O2/65%CO2
value of 30% or 35%, the burnout achieved is higher than that in air due
to an enhancement of the char combustion rate [41,42]. As can be seen
in Fig. 3(c), the decrease in char burning rate in 21% O2/79% CO2 with
b
DAB
respect to air was confirmed with CFD tests, the improvement in 30% 100 96.3
99.2 98.6
95.8
99.3 98.7 97.3
99.9 99.8

O2/70% CO2 and 35% O2/65% CO2 atmospheres was confirmed as well.
DAB and BA present higher burnouts than HVN for all the 80
atmospheres studied because they are more reactive coals. Coal
Burnout (%)

burnout for DAB and BA seems to be less affected by combustion 60


atmosphere than HVN. Although, for DAB a slight decrease is observed
for the 21% O2/79% CO2 environment in comparison with the air 40
reference case. There is also an improvement for the 30% O2/70% CO2
Experimental
environment, but this improvement is less marked than for HVN. Low 20 CFD predictions
Model predictions

0
21%O2/79%N2 21%O2/79%CO2 30%O2/70%CO2

100
96.3 95.8 97.3 95.9 96.5 96.5 96.1 c
BA
99.3 98.6 99.3 98.7 100 99.7 100 100
100 95.9 96.5 96.5 96.1
90

81.7 82.4
Burnout (%)

80
79.4
80
Burnout (%)

77.2
60
70
40
21%O2/79%N2
21%O2/79%CO2 Experimental
60 CFD predictions
30%O2/70%CO2 20
Model predictions
35%O2/65%CO2

50 0
HVN DAB BA 21%O2/79%N2 21%O2/79%CO2 30%O2/70%CO2 35%O2/65%CO2

Fig. 7. Experimental burnouts of the three coals studied at both air and O2/CO2 Fig. 8. Comparison between experimental and numerical coal burnouts for (a) HVN
conditions. (b) DAB and (c) BA combustion.
1496 L. Álvarez et al. / Fuel Processing Technology 92 (2011) 1489–1497

summarised in Table 3, a better agreement between the experimental a


HVN
and predicted values was obtained. 14 13.4

A comparison between experimental and numerical O2 concen- 12


trations at furnace exit is shown in Fig. 9. When the commercial CFD 10.0 10.3
10 9.5
model was employed, O2 concentrations were under predicted, just 8.5
9.0
7.8 8.0 7.9
the opposite case than for coal burnout. With default inputs the 8

O2 (%)
6.8
degree of coal conversion during combustion is higher, and so is the
6 5.1 5.1
amount of oxygen required in coal combustion. When the devolati-
lisation and combustion parameters described in Table 3 are 4 Experimental
CFD predictions
employed, burnout shows a better agreement with the experimental
2 Model predictions
values, and so do the O2 concentrations.
0
21%O2/79%N2 21%O2/79%CO2 30%O2/70%CO2 35%O2/65%CO2
5. Conclusions
b
DAB
A CFD model for an entrained flow reactor was employed to study 14
overall combustion behaviour under O2/CO2 conditions. The comput- 12 11.0
ed results show changes in the temperature profiles inside the reactor
and differences in coal combustion behaviour when N2 is replaced by 10
CO2. A decrease in the predicted peak temperature is observed,

O2 (%)
8 7.0
associated with differences in gas density, heat capacity and radiative 6.3 6.3
5.9
properties of N2 and CO2. A reduction of about 60 K was found for the 6
4.5 4.6 4.6 4.7
semi-anthracite coal, whereas reductions of 145–160 K were ob- 4
Experimental
served for the high volatile bituminous coals, when N2 is replaced by CFD predictions
CO2 at comparable oxygen mole fractions. In this work, the CO2 2 Model predictions

absorption coefficient was predicted by using the WSGGM model. 0


Further improvements to radiative properties at oxy-fuel conditions 21%O2/79%N2 21%O2/79%CO2 30%O2/70%CO2
are being examined in order to improve the accuracy of heat transfer
prediction.
c
BA
14
High CO2 concentrations seem to enhance coal devolatilisation, but
do not appear to affect subsequent char combustion. A decrease in 12 11.5
10.5
coal burnout rate is observed for 21% O2/79% CO2 with respect to the 10 10
10
air reference case, for the three coals studied. Similar burnout rates to 9 8.8
O2 (%)

those for air are attained with an oxygen content of 30% in the CO2 8 7.0
6.4
mixture. A worsening on coal burnout values is observed in both 6
experimental and predicted values in 21% O2/79% CO2 if compared 4.5 4.6 4.6 4.7

with air conditions. This reduction in char burnout is particularly 4 Experimental


CFD predictions
marked in the case of the semi-anthracite. Under O2/CO2 conditions, 2 Model predictions
the oxygen concentrations must be higher than 21% to obtain similar
or higher burnout as in air. Improvements in the model may lead to 0
21%O2/79%N2 21%O2/79%CO2 30%O2/70%CO2 35%O2/65%CO2
better coal burnout predictions, species concentration and tempera-
ture distributions. Although the computed burnout and O2 concen- Fig. 9. Comparison between experimental and numerical O2 concentrations at furnace
tration results are in good agreement with the experimental values, a exit for (a) HVN (b) DAB and (c) BA combustion.
more accurate description of the combustion kinetics by including
gasification reactions or updating the rate coefficient parameters
would lead to better predictions. European Regional Development Fund. L.A. and J.R. acknowledge
funding from the CSIC JAE program, which was co-financed by the
Acknowledgements European Social Fund, and the Asturias Regional Government (PCTI
program), respectively. M.G. acknowledges financial support from E.ON
The present study was carried out with financial support from the UK, EPSRC, and the Dorothy Hodgkin Postgraduate Awards. We also
Spanish MICINN (Project PS-120000-2005-2) co-financed by the thank Dr. L Ma and Professor J M Jones for helpful advice.

Table 3
Devolatilisation and combustion reactivity data inputs for Fluent mode.

Parameter/case Default HVN DAB BA

Devolatilisation model Single rate Single rate Single rate Single rate
Pre-exponential factor (1/s) 38200 3.6E14 4.68E11 2.01E11
Activation energy (kJ/mol) 74 229.7 155.9 148.6
Combustion model Intrinsic Intrinsic Intrinsic Intrinsic
Mass diffusion–limited rate constant 5E−12 5E−12 5E−12 5E−12
Kinetic-limited rate pre-exponential factor (g/cm s) 0.030198 0.030198 0.030198 0.030198
Kinetic-limited rate activation energy (kJ/mol) 179 155 ± 10 155 ± 10 155 ± 10
Char porosity 0.5 0.5 0.5 0.5
Mean porous radio (m) 6E−8 6E−8 6E−8 6E−8
Specific internal area (m2/g) 300 40 300 300
Tortuosity 1.41 1.41 1.41 1.41
Burning mode, alpha 0 0.25 0.25 0.25
L. Álvarez et al. / Fuel Processing Technology 92 (2011) 1489–1497 1497

References [22] E. Hampartsoumian, D. Hainsworth, J.M. Taylor, A. Williams, The radiant


emissivity of some materials at high temperatures-review, Journal of the Institute
[1] B.J.P. Buhre, L.J. Elliot, C.D. Sheng, R.P. Gupta, T.F. Wall, Oxy-fuel combustion of Energy 74 (2001) 91–99.
technology for coal-fired power generation, Progress in Energy and Combustion [23] R.I. Backreedy, L.M. Fletcher, L. Ma, M. Pourkashanian, A. Williams, Modelling
Science 31 (2005) 285–307. pulverised coal combustion using a detailed coal combustion model, Combustion
[2] T.F. Wall, Y. Liu, C. Spero, L. Elliot, S. Khare, R. Rathman, F. Zeenathal, B. Science and Technology 178 (2006) 763–787.
Moghtaderi, B.J.P. Buhre, C.D. Sheng, R.P. Gupta, T. Yamada, K. Makino, J. Yu, An [24] F.C. Lockwood, S.M.A. Rizvi, N.G. Shah, Comparative predictive experience of coal
overview on oxyfuel coal combustion — state of the art research and technology firing, Proceedings of the Institution of Mechanical Engineers — Part C:
development, Chemical Engineering Research and Design 87 (2009) 1003–1016. Mechanical Engineering Science 200 (1986) 79–87.
[3] M.B. Toftegaard, J. Brix, P.A. Jensen, P. Glarborg, A.D. Jensen, Oxy-fuel combustion [25] I.M. Khan, G.A. Greeves, A method for calculating the formation and combustion of
of solid fuels, Progress in Energy and Combustion Science 36 (2010) 581–625. soot in diesel engines, in: N.H. Afgan, J.M. Beer (Eds.), Heat Transfer in Flames 25,
[4] P. Edge, M. Gharebaghi, R. Irons, R. Porter, R.T. Porter, M. Pourkashanian, D. Smith, Scripta, Washington DC, 1973.
P. Stephenson, A. Williams, Combustion modelling opportunities and challenges [26] B.F. Magnussen, B.W. Hjertager, On mathematical modelling of turbulent
for oxy-coal carbon capture technology, Chemical Engineering Research and combustion with special emphasis on soot formation and combustion, Sympo-
Design, 2011, doi:10.1016/j cherd. 2010.11.010. sium (International) on Combustion, 16, 1977, pp. 719–729.
[5] L. Strömberg, G. Lindgren, J. Jacoby, R. Giering, M. Anhenden, U. Burchardt, H. [27] A. Williams, R.I. Backreedy, R. Habib, J.M. Jones, M. Pourkashanian, Modelling coal
Altmann, F. Kluger, G.-N. Stamatelopoulos, Update on Vattenfall's 30 MWth combustion: the current position, Fuel 81 (2002) 605–618.
Oxyfuel pilot plant in Schwarze pump, Energy Procedia 1 (2009) 581–589. [28] J.M. Jones, P.M. Patterson, M. Pourkashanian, A. Williams, A. Arenillas, F. Rubiera,
[6] N. Nikolopoulos, A. Nikolopoulos, E. Karampinis, P. Grammelis, E. Kakaras, J.J. Pis, Modelling NOx formation in coal particle at high temperature: an
Numerical investigation of the oxy-fuel combustion in large scale boilers adopting investigation of the devolatilisation kinetic factors, Fuel 78 (1999) 1171–1179.
the ECO-scrub technology, Fuel 90 (2011) 198–214. [29] Y. Zhao, M.A. Serio, R. Bassilakis, P.R. Solomon, A method of predicting coal
[7] S.P. Khare, T.F. Wall, A.Z. Farida, Y. Liu, B. Moghtaderi, R.P. Gupta, Factors devolatilization based on the elemental composition, Symposium (International)
influencing the ignition of flames from air-fired swirl pf burners retrofitted to oxy- on Combustion/The Combustion Institute, 25, 1994, pp. 553–560.
fuel, Fuel 87 (2008) 1042–1049. [30] Y.R. Sivathanu, G.M. Faeth, Generalized state relationships for scalar properties in
[8] R. Porter, F. Liu, M. Pourkashanian, A. Williams, D. Smith, Evaluation of solution non-premixed hydrocarbons/air flames, Combustion and Flame 82 (1990)
methods for radiative heat transfer in gaseous oxy-fuel combustion environ- 211–230.
ments, Journal of Quantitative Spectroscopy and Radiative Transfer 111 (2010) [31] I.W. Smith, The combustion rates of coal chars: a review, Symposium
2084–2094. (International) on Combustion, 19, 1982, pp. 1045–1065.
[9] W.F. DeGroot, G.N. Richards, Relatives rates of carbon gasification in oxygen, [32] N.M. Laurendeau, Heterogeneous kinetics of coal char gasification and combus-
steam and carbon dioxide, Carbon 27 (1989) 247–252. tion, Progress in Energy Combustion and Science 4 (1978) 221–270.
[10] H. Watanabe, M. Otaka, Numerical simulation of coal gasification in entrained [33] I.W. Smith, Kinetics of combustion of size-graded pulverized fuels in the
flow coal gasifier, Fuel 85 (2006) 1935–1943. temperature range 1200–2270 K, Combustion and Flame 17 (1971) 303–314.
[11] J.M. Jones, D. Waldron, M. Pourkashanian, A. Williams, Prediction of NOx and [34] I.W. Smith, The kinetics of combustion of pulverised semi-anthracite in the
unburned carbon in ash in a highly staged pulverised coal furnace using over-fire temperature range of 1400–2200 K, Combustion and Flame 17 (1971) 421–428.
air, Journal of the Energy Institute 83 (2010) 144–150. [35] S. Charpenay, M.A. Serio, P.R. Solomon, The prediction of coal char reactivity under
[12] A. Arenillas, R.I. Backreedy, J.M. Jones, J.J. Pis, M. Pourkashanian, F. Rubiera, A. combustion conditions, Symposium (International) on Combustion, 24, 1992,
Williams, Modelling of NO formation in the combustion of coal blends, Fuel 81 pp. 1189–1197.
(2002) 627–636. [36] M. Lupion, B. Navarrete, P. Otero, V.J. Cortés, Experimental programme in
[13] B. Arias, C. Pevida, F. Rubiera, J.J. Pis, Effect of biomass blending on coal ignition CIUDEN's CO2 capture technology development plan for power generation,
and burnout during oxy-fuel combustion, Fuel 87 (2008) 2753–2759. Chemical Engineering Research and Design doi:10.1016/j.cherd.2010.10.017.
[14] S. Badzioch, P.G.W. Hawksley, Kinetics of thermal decomposition of pulverized [37] R.K. Rathman, L.K. Elliot, T.F. Wall, Y. Liu, B. Moghtader, Differences in reactivity of
coal particles, Industrial and Engineering Chemistry Process Design and pulverised coal in air (O2/N2) and oxy-fuel conditions (O2/CO2), Fuel Processing
Development 9 (1970) 521–530. Technology 90 (2009) 797–802.
[15] ANSYS FLUENT, 2009. Version 12, Ansys Inc. USA. [38] R.I. Backreedy, L.M. Fletcher, J.M. Jones, L. Ma, M. Pourkashanian, A. Williams, Co-
[16] S.A. Orszag, V. Yakhot, W.S. Flannery, F. Boysan, D. Choudhury, J. Maruzewski, B. firing coals and biomass: a modelling approach, Proceedings of the Combustion
Patel, Renormalization group modeling and turbulence simulations, In Interna- Institute 30 (2005) 2955–2964.
tional Conference on Near-Wall Turbulent Flows, 19938, Tempe, Arizona. [39] P.A. Berejano, Y.A. Levendis, Single-coal particle combustion in O2/N2 and O2/CO2
[17] W.P. Jones, B.E. Launder, The prediction of laminarization with a two-equation enviroments, Combustion and Flame 153 (2008) 270–287.
model of turbulence, International Journal of Heat and Mass Transfer 15 (1972) [40] L. Zhang, E. Binner, Y. Qiao, C.-Z. Li, In situ diagnostics of Victorian brown coal
301–314. combustion in O2/N2 and O2/CO2 mixtures in drop tube furnace, Fuel 89 (2010)
[18] B.E. Launder, B.B. Spalding, The numerical computation of turbulent flows, 2703–2712.
Computer Methods in Applied Mechanics and Engineering 3 (1973) 269–289. [41] H. Liu, R. Zailani, B.M. Gibbs, Comparison of pulverized coal combustion in air and
[19] R. Viskanta, M.P. Menguc, Radiation heat transfer in combustion systems, Progress mixtures of O2/CO2, Fuel 84 (2005) 833–840.
in Energy Combustion Science 13 (1987) 97–160. [42] Y. Tan, E. Croiset, M.A. Douglas, K.V. Thambimuthu, Combustion characteristics of
[20] R.I. Backreedy, J.M. Jones, L. Ma, M. Pourkashanian, A. Williams, A. Arenillas, B. coal in a mixture of oxygen and recycled flue gas, Fuel 85 (2006) 507–512.
Arias, J.J. Pis, F. Rubiera, Prediction of unburned carbon and NOx in a tangentially [43] G. Várhegyi, P. Szabó, E. Jakab, F. Till, Mathematical modelling of char reactivity in
fired power station using single coals and blends, Fuel 84 (2005) 2196–2203. Ar–O2 and CO2–O2 mixtures, Energy and Fuels 10 (1996) 1208–1214.
[21] L. Ma, M. Gharebaghi, R. Porter, M. Pourkashanian, J.M. Jones, A. Williams, [44] Y.-G. Kim, J.-D. Kim, B.-H. Lee, J.-H. Song, Y.-J. Chang, C.-H. Jeon, Experimental
Modelling methods for co-fired pulverised fuel furnace, Fuel 88 (2009) investigation into combustion characteristics of two sub-bituminous coals in O2/N2
2448–2454. and O2/CO2 environments, Energy and Fuels 24 (2010) 6034–6040.

You might also like