You are on page 1of 12

Energy xxx (2016) 1e12

Contents lists available at ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

A modeling approach to co-firing biomass/coal blends in pulverized


coal utility boilers: Synergistic effects and emissions profiles
rez-Jeldres a, *, Pablo Cornejo b, Mauricio Flores c, Alfredo Gordon a,
 n Pe
Rube
Ximena García a
a
Department of Chemical Engineering, University of Concepcion, P.O. Box 160-C, Concepcion, Chile
b
Department of Mechanical Engineering, University of Concepcio n, P.O. Box 160-C, Concepcio
n, Chile
c n, P.O. Box 4051, Coronel, Chile
Technological Development Unit, University of Concepcio

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

Article history: Pulverized coal power plants in Chile are evaluating to reduce CO2 emissions by co-firing coal with
Received 7 November 2015 biomass, which is CO2-neutral. A computational fluid dynamics model was used in this study to predict
Received in revised form the performance of a 150 MW commercial boiler co-firing pulverized coal with pine sawdust. Synergistic
26 October 2016
effects were identified by burnout, thermal and hydrodynamic profiles. Co-firing was simulated with 5%
Accepted 21 November 2016
Available online xxx
of biomass substitution, and feeding in the first level of burners. The model was validated using data
from the power plant. The results show an expected decrease in SO2 emissions and a negligible reduction
in heat transferred to the water tubes (0.6%). Biomass presence increased the burning rate of fuel par-
Keywords:
Co-firing
ticles, as shown by higher CO2 emissions and a lower CO concentration, per unit of thermal power. The
Sawdust model reveals synergistic effects, proved by an increase in temperature, due to an early combustion of
Synergistic biomass particles, increase in the coal combustion rate, and a better temperature distribution in the
Pulverized coal boiler. These synergistic effects were compared with results obtained at bench scale reported in the
Tangentially boiler literature. Thus, it was concluded that a relatively small replacement of coal by biomass could signifi-
Biomass cantly improve the fuel combustion process and the boiler performance.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction hectares of native and forest plantations and an average yield of


20e40 m3/ha/y [3]. Furthermore, at about 4 million of tonnes/year
The generation of electricity is the essential engine and the base of woody residues (14,000 GWh/y) are produced from forestry
for sustainable economic development of a country. Thus, low-cost activities. These residues can replace an important fraction (viz.
power generation along with an independence from fossil fuels in 25%) of the internal coal demand for electricity production in coal-
its energy matrix would allow South American countries to reduce fired systems leading to a more sustainable energy matrix [3].
their fixed costs and offer the opportunity to enter into new mar- Therefore, there is an interesting opportunity to transform the
kets. In recent years, Chile has made a great effort to include con- installed generation units into systems capable to process biomass/
ventional and unconventional renewable energies in their energy coal blends, which will favor the reduction of fossil-derived emis-
matrix. However, the transition from conventional energy to sions as well as the increment in the share of local resources used
renewable energy sources and clean technologies to generate en- for energy production.
ergy independence from fossil fuels is slow, requiring regulations to To achieve this goal, co-firing of coal with biomass has become a
introduce these technologies into the energy matrix [1]. In this feasible alternative to existing pulverized coal power plants.
context, Chile has recently approved a new legal regulation that However, extensive and costly field tests are frequently needed to
demands the utilization of 10% of unconventional renewable en- analyze the performance in existing utility boilers, and the co-firing
ergy in the thermoelectric sector [2]. of coal with biomass may interfere with the normal operation of
As special feature, Chile accounts for more than 15 million the power plant [4]. Numerical simulations may be used to study
the co-firing process, predicting the behavior of the combustion
process within the boiler, reducing the number of experimental
tests by designing appropriate field tests, and establishing the
* Corresponding author.

http://dx.doi.org/10.1016/j.energy.2016.11.116
0360-5442/© 2016 Elsevier Ltd. All rights reserved.

rez-Jeldres R, et al., A modeling approach to co-firing biomass/coal blends in pulverized coal utility boilers:
Please cite this article in press as: Pe
Synergistic effects and emissions profiles, Energy (2016), http://dx.doi.org/10.1016/j.energy.2016.11.116
2 R. Perez-Jeldres et al. / Energy xxx (2016) 1e12

Nomenclature Abbreviations
BFG Blast Furnace Gases
A Pre-exponential constant [s1] CEMS Continuous Emission Monitoring System
AP Particle area [m2] CFD Computational Fluid Dynamics
CP Specific heat [kJ kg1 K1] COG Coke Oven Gas
Do Oxygen diffusion coefficient [m2 s1] CPD Chemical Percolation Devolatilization model
E Activation Energy [kJ kmol1] DO Discrete Ordinate radiation model
dP Particle diameter [m] DPM Discrete Phase Model
de Mean particle diameter [m] DTRM Discrete Transfer Radiation Method
h Convective coefficient [W m2 K1] EDM Eddy Dissipation Model
I Radiation intensity [W sr1] EDC Eddy Dissipation Concept model
mP Mass particle [kg] EPA Environmental Protection Agency
mp,0 Initial mass particle [kg] FG Functional Group devolatilization model
mp,daf Initial mass dry ash free particle [kg] LHV Lower Heating Value [kJ kg1]
R Surface kinetic rate [m2 s1] PDF Probability Density Function for transport species
TP Particle temperature [K] PKE Palm kernel extract
T∞ Local temperature of the gas phase [K] RANS Reynolds Averaged Navier-Stokes (equations)
Yi Mass fraction of the i-species [kg kg1] RNG Renormalization Group model
RSM Reynolds Stress Model
Greek symbols RTE Radiative Transfer Equation
εP Absorption coefficient SST Shear Stress Transport model
ƟR Radiation temperature [K] SIMPLE Semi-implicit Method for Pressure-Linked Equations
g Dispersion factor TGA Thermogravimetric Analysis
s Stefan-Boltzmann constant (5.67  108 [W m2 K4]) WSGGM Weighted Sum of Gray Gases Model

optimum operating conditions. Among the different combustion elevations [7]. In the case of small-diameter biomass particles, CFD
technologies, integration of a biomass firing system into a pulver- simulation results show that if biomass is fed into the lowest level
ized coal boiler could be far more challenging than integration of a of burners, higher temperatures and heat fluxes to the walls are
biomass firing system into a grate-fired boiler or a fluidized bed obtained, than when biomass is fed at higher levels [14]. 4) The
combustor because the burner aerodynamics and fuel properties operational parameters are unaffected by co-firing biomass at a low
have a much greater impact on combustion and the formation of thermal loading [7]. However, the temperature at the furnace exit is
pollutant emissions [5]. slightly increased at a low thermal loading of biomass, so there is no
Computational fluid dynamics (CFD) has been applied success- expectation of intensification of slagging/fouling in the convective
fully to study pulverized coal combustion and co-firing coal with zone [7]. However, higher ratios biomass/coal than 20 wt% in the
biomass in different aspects, e.g., to study the effects produced in co-firing process could produce a decrease in the furnace exit
the performance of a low-NOX multi-fuel burner by substituting the temperature, thus enhancing the possibility of slagging/fouling [9].
feed from coal to straw [6]. A significant number of CFD models 5) A decrease in SO2 and NOX emissions should be obtained because
have been developed for co-firing of coal with biomass but only a of the lower contents of sulfur and nitrogen in the biomass fuel. A
limited number of the CFD models have been applied on an in- reduced availability of nitrogen in the blend of course results in a
dustrial scale or have been validated by experimental tests in large fuel NOX decrease [7]. 6) When biomass is incorporated into a co-
power plants (See Table 1),. Some of these models have been used fired burner, or when an existing low NOX coal-fired burner is
to study the influence of operational factors related to the biomass retrofitted with biomass, the air supply should be finely tuned. The
feeding conditions: mean particle size, substitution level of coal by air supply should be sufficient to ensure that the excess air will not
biomass and feeding location in the furnace. reduce the flame temperature but will provide sufficient oxygen in
The main conclusions of the studies summarized in Table 1 are zones where volatiles are released to promote the homogeneity
as follows: 1) A limit exists in the maximum biomass substitution and char combustion [15]. Normally, in pulverized coal combustion,
level and particle size that can maintain a reasonable boiler effi- an adequate air excess is normally between 15 and 20% [16].
ciency [4]. Some authors show a maximum of 20% of mass substi- These CFD models have been developed to study the macro-
tution due to technical limitations [13], e.g. the milling capacity and scopic response of coal plants under co-firing conditions. However,
power consumption. Pallare s et al. [4] propose to maintain the mill limited information is available about the possible synergies be-
power consumption at optimum values for a cost-efficient com- tween coal and biomass during the combustion process. In this
bustion process. The mean diameter of the biomass particles must field, thermogravimetric studies under an inert atmosphere and a
be in the range of 0.5e1 mm. 2) The mean particles size is the most low heating rate of 5e50 K/min are typically used to study syner-
influential factor in biomass combustion efficiency to obtain an gistic effects during the co-firing process. The results are quite
adequate burnout. Low mean particles diameter (0.5e1 mm) in- controversial. Some studies indicate that no interactions occur in
crease the combustion rate and the temperature in the near burner the pyrolysis and combustion of coal-biomass blends [17e20],
region, requiring low residence time and obtaining higher com- whereas several other studies present an increase in the gas/vola-
bustion efficiencies than particles with mean diameters higher than tile yield and composition, so a higher reactivity of the coal-
1 mm [4]. 3) Particles of biomass with larger diameters run the risk biomass blends depends on the mixing ratio and temperature
of reaching the hopper without being completely burned. Thus, it is [21e24]. Even if there is no interaction between the coal and the
desirable to feed these particles to the burners located at higher biomass during the devolatilization process, the heat released from

rez-Jeldres R, et al., A modeling approach to co-firing biomass/coal blends in pulverized coal utility boilers:
Please cite this article in press as: Pe
Synergistic effects and emissions profiles, Energy (2016), http://dx.doi.org/10.1016/j.energy.2016.11.116
R. Perez-Jeldres et al. / Energy xxx (2016) 1e12 3

Table 1
Resume of the pulverized combustion boilers models under co-firing condition.

Technology Fuels Turb. Devol. model Rad. Gas reaction Char comb. Ref
model model.

Pulverized Combustion Boiler Coal - Cyanara k-ε Single rate P-1 PDF Single film model [4]
350 MWe cardunculus
Front Wall burners
Replace:4% in energy basis
Pulverized combustion - Test Coal - Straw SST k-u Single rate DO Finite rate/eddy dissipation Single rate [6]
Facility
Low NOX Multiburner
Replace: 0e100% in mass basis
Pulverized Combustion Boiler Lignite - Cardoon k-ε Single rate P-1 Finite rate/eddy dissipation UDF - User Defined Function [7]
300 MWe
Tangentially burners
Replace: 5e10% in energy basis
Pulverized Furnace - Test Coal - k-ε RNG FG. DO Eddy dissipation Smith's Intrinsyc Model [8]
Facility Milled wood
0.5 MW PKE
Replace: 15e20% in energy Olive
basis. Miscanthus
Pulverized Furnace - Test Coal e k-ε RNG Two competing P-1 PDF Kinetics/diffusion limited rate [9]
Facility Wheat rates model
5.8 MW Straw
Replace: 10e20% in energy basis
Pulverized Combustion Coal k-ε CPD P-1 Eddy dissipation Kinetic/diffusion Model [10]
Tangentially Fired Utility Boiler BFG
200 MW COF
Replace: 20% in energy basis
Pulverized Combustion Coal - k-ε Single rate DTRM Three steps defined by the Three steps defined by the [11]
Tangentially Fired Utility Boiler Lignite authors authors
550 MW
Replace: 20e60% in energy basis
Pulverized Combustion Coal k-ε Two competing DO Finite rate Kinetics/diffusion limited rate [12]
Wall Fired Utility Boiler Wood chip rates model
700 MW Wood pellet
Replace: 0e100% in energy basis Sawdust

Note: BFG ¼ Blast Furnace Gases. CPD ¼ Chemical Percolation Devolatilization model. COG ¼ Coke Oven Gas. DO ¼ Discrete Ordinate radiation model. DTRM ¼ Discrete
transfer radiation method. FG ¼ Functional Group devolatilization model. PKE ¼ Palm kernel extract. RNG ¼ Renormalization Group. SST ¼ Shear Stress Transport.

homogeneous-heterogeneous combustion of biomass particles 2.1. Plant description


could generate a synergistic effect. This synergistic effect could be
manifested as a variation in burnout rates, exhaust gas concentra- The Tocopilla power plant consists of two pulverized coal utility
tions and temperature profiles. boilers (Unit 14 and Unit 15), with a total generation capacity of
Therefore, this study aims to answer three important questions: 300 MW. Unit 14 (U14) was designed by Mitsubishi Heavy In-
i) On an industrial scale, if a small fraction of coal is replaced by dustries to operate at 18.8 MPa absolute and 813 [K], resulting in a
biomass, will synergistic effects be detected during co-firing? ii) steam evaporation capacity of 108 [kg s1]. A schematic of the
Because of the low LHV of biomass, will there be a substantial installed boilers is presented in Fig. 1, indicating main furnace zones
decrease in heat flux that could affect evaporator tubes and steam (focus of the numerical model) and the distribution of fuel feeding
production? iii) By replacing 5% of the coal by biomass, would it be ports. The dimensions of the main cross section of the furnace are
possible to observe an increase in temperature and combustion 9.0  9.2 [m]; the height of the boiler is 32.4 [m], and the com-
efficiency? bustion chamber volume is 2130 [m3].
The boiler has four levels of burners, with four burners each,
which are tangentially distributed. During 100% of coal usage
(normal operation condition) in the power plant, the boiler burns
2. Description of fuel and facilities two types of fuels namely Adaro and Hatillo, which compositional
analysis and properties are summarized in Table 2 along with those
The above-mentioned research questions are addressed by a of sawdust. The modeled boiler (Unit 14) burns Adaro coal at in-
CFD model, which is implemented to simulate the process of co- termediate burner levels and Hatillo coal in the upper and lower
firing coal with sawdust in a commercial 150 MW utility boiler burner levels, mainly to avoid an excessive flame temperature and
located in Tocopilla, Chile. A commercial code (ANSYS FLUENT 14) thus to prevent thermal NOX formation. A 5% (mass-based) of
was used to solve the transport equations for momentum, energy replacement at the first level burners (level A) was considered for
and mass. The model was validated using the results of a field test modeling purposes. The conservative replacement used is due to
(large-scale) operating with 100% pulverized coal. The co-firing the necessity of the power plant to test the boiler and the mills,
study was carried out for a mass-base replacement of 95/5 wt% operating at a small quantity of biomass and not to modify the
coal/sawdust mixture in the first level of burners. Sawdust is hydrodynamics within the boiler. The results of this work will allow
considered to be derived from Pinus radiata, since this specie ac- to the power plant to obtain the environmental authorization to
counts for more than 60% of forest plantations in the country. develop his co-firing test.
Furthermore, Arteaga-Pe rez et al. [25] have recently reported on After milling (each level has its own rod mill), the pulverized
the environmental benefits of using it as fuel in co-firing stations.

rez-Jeldres R, et al., A modeling approach to co-firing biomass/coal blends in pulverized coal utility boilers:
Please cite this article in press as: Pe
Synergistic effects and emissions profiles, Energy (2016), http://dx.doi.org/10.1016/j.energy.2016.11.116
4 R. Perez-Jeldres et al. / Energy xxx (2016) 1e12

Fig. 2. Unstructured mesh of a cross section of level-A burners.

The modeled zones were represented using a grid structure


based on a non-structured hexahedral mesh, as illustrated in Fig. 2,
where 96% of the cells have a quality equal to or lower than 0.6 (0
means a very good quality and 1 a very poor quality). To ensure the
Fig. 1. Burner levels and secondary air distribution. independence of the results, several tests were performed to
evaluate the influence of the grid size on the results. The mesh was
refined, and the computations were repeated until the variation of
fuel particles are pneumatically transported by the primary air to
the results was lower than 5%, resulting in an optimal cell number
the corresponding burner. The secondary air flow is fed by a wind-
of approximately 322,000 cells.
box located between the burners, as shown in Fig. 1. The operating
conditions are shown in Table 2.
3.1. Gas-solid reaction model
3. Mathematical model
An Eulerian-Lagrangian approach was used to model the gas-
The boiler geometry comprises a burners and a radiative/ solid two-phase flow. The two-way coupling model was chosen
convective zone (see Fig. 1). To simplify the model, the superheater because the solid phase affects the transfer of mass, momentum
was not included because it has a negligible influence on the overall and energy in the gas phase. The latter was modeled using the
combustion process, considering the low gas temperature at its Reynolds Averaged Navier-Stokes (RANS) equations, where the
inlet and the low char and charcoal reactivity far from the reactive Reynolds stress tensor is described by the Boussinesq hypothesis.
zone [4]. The RANS equations were coupled with a k-ε model to incorporate

Table 2
Fuel analysis, properties and feed conditions.

Adaro Coal Hatillo Coal Sawdust

Proximate analysis (wt% as received)

Fixed carbon (by difference) 33.6 44.9 11.5


Volatile matter 35.7 35.3 62.8
Moisture 27.7 14.2 25.4
Ash 2.93 5.68 0.30

Ultimate analysis (wt% dry ash free basis)

Carbon 72.7 79.0 48.9


Hydrogen 4.80 5.40 6.00
Oxygen 21.4 13.1 45.0
Nitrogen 0.92 1.70 0.10
Sulfur 0.18 0.80 0.01

Fuel properties/Burner feed level B/C A/D A

Lower heating value (as received) [kJ kg1] 20,524 25,623 11,190
Rosin-Rammler dispersion factor (1) 0.678/1.419 1.092/1.008 1.092
Rosin-Rammler mean diameter [mm] (1) 0.090/0.150 0.110/0.090 0.110

Feed conditions/Burner level B/C A/D A

Primary air flow [kg s1] 7.50/6.67 6.67/6.94 For co-firing conditions see Table 4.
Secondary air flow [kg s1] 26.8/26.8 26.8/26.8
Primary air temperature [K] 884/884 884/884
Secondary air temperature [K] 1114/1114 1114/1114
Combustible mass flow [kg s1] 4.00/3.97 4.3/4.03

Note: The symbol “/” separates the value corresponding to one level of burners from another level.

rez-Jeldres R, et al., A modeling approach to co-firing biomass/coal blends in pulverized coal utility boilers:
Please cite this article in press as: Pe
Synergistic effects and emissions profiles, Energy (2016), http://dx.doi.org/10.1016/j.energy.2016.11.116
R. Perez-Jeldres et al. / Energy xxx (2016) 1e12 5

the turbulent viscosity. Different models for turbulence were 3.4. Kinetics of coal and sawdust combustion
compared with a non-linear model, the Reynolds Stress Model
(RSM), proving that the k-ε model is capable of representing the At the discrete phase, the mass and heat transfer rates over the
hydrodynamics of the system at a lesser computational cost than particles are calculated considering four stages: heating, drying,
the RSM but with a maximum difference of 10%. The SIMPLE al- devolatilization (char formation) and char oxidation. In the first
gorithm was used to predict the flow, temperature and gas con- stage, the particle is heated until its temperature reaches the water
centrations within the boiler in addition to solving time-averaged vaporization temperature (373 [K]). At this point, the temperature
expressions for mass, momentum, and enthalpy of flue-gas species. of the particle remains constant, and the moisture content is
released during the endothermic drying. When the particle reaches
3.2. Discrete phase model the devolatilization temperature and the moisture mass fraction is
zero, the devolatilization process begins and persists until all vol-
The discrete phase model (DPM) in a Lagrangian reference atile materials are released.
system was used to describe the disperse solid phase. This model
idealizes the discrete phase as dispersed particles in the continuous Zt
phase. The model calculates the trajectories, heat, momentum and Zt  ðk1 þ k2 Þdt
mv ðtÞ
mass transfer to and from the particles. The trajectories were ob- ¼ ða1 $k1 þ a2 $k2 Þ$e 0 dt (4)
tained by the integration of the force balance in the particle. The mp;0  mp;daf
0
inertial force is thus equal to the forces acting on the particle
(weight and drag force). The drag coefficient for spherical particles E1 E2
was calculated using the Morsi and Alexander correlation [26]. k1 ¼ A1 $eRT ; k2 ¼ A2 $eRT (5)
The discrete phases were modeled in separate ways for coal and
The coal devolatilization rate was simulated using the well-
biomass particles. The Rosin-Rammler distribution was used to
known two competing rates model (Eq. (4)). In the case of coal,
characterize the fuel particle sizes (Eq. (1)), as presented in Table 2.
the values of activation energy and pre-exponential constants re-
The mean diameter (dp) and dispersion factor (g) were calculated
ported by Kobayashi et al. [28] were used, while for biomass these
from sieve fractions provided by the power plant. Coal and biomass
values were taken from Pallare s et al. [4] (Table 3). This model was
particles were assumed to have a spherical shape, which will in-
selected because it can change the devolatilization rate as a func-
fluence the fluid-dynamic behavior and the particle heating rate.
tion of the temperature. The model assumes that the light volatile
  g  matter (higher H/C ratio) is released at low temperatures, and the
dp
Yd ¼ exp  (1) heavier products (lower H/C ratio) are released at higher temper-
de
atures, having a higher release rate than the light products, e.g.,
pulverized coal combustion [29]. Mass stoichiometric factors for
low temperatures (a1) are assume to be equal to the VM content,
3.3. Heat transfer mechanisms but at higher temperatures, the mass stoichiometric factor (a2) is
equal to 1.0, as recommended in Ref. [30]. The biomass devolatili-
Radiation was quantified using the discrete ordinates model zation was modeled using the single rate model expressed in an
(DO). This model solves the radiative transfer equation (RTE) for a Arrhenius form. Despite the LHV of biomass, the higher content of
three-dimensional finite number of angles, each associated with a VM in biomass is expected to promote the ignition of the coal
director vector fixed in the global Cartesian system. In the DO particles, being observed in the burnouts, temperatures and con-
model, the RTE equation becomes a transport equation for the ra- centration profiles.
diation intensity in spatial coordinates. In this study, transport When the coal particle is dried and its volatile matter is released,
equations in five directions were solved. The solution method is the coal particle burns until the char mass fraction is reduced to
identical to the one used to solve the flux equations. zero. Under the normal operating conditions of pulverized-char
The heat transfer between the homogeneous and discrete combustion, the char reaction will be partly controlled by the
phases was calculated from the heat balance applied to the particle diffusion of oxygen to the surface of the char (physical control),
(Eq. (2)), relating the particle temperature TP, the convective coef- itself a function of the particle diameter, and partly by intrinsic
ficient (h) and the absorption/emission of radiation (εP ) at the chemical reaction kinetics or by a combination of pore diffusion and
surface of the particle (AP). reaction kinetics (chemical control) [16,31].
The diffusion of oxygen to the external surface of the particles
dTP   becomes the major rate-controlling influence for temperatures
mp Cp ¼ hAP ðT∞  TP Þ þ 3 P AP s q4R  Tp4 (2) above 1823 [K]. For temperatures below 773 [K], chemical reactions
dt
control the surface reaction rate [33]. For temperatures above 1005
The radiation temperature (ƟR) was calculated from the incident
[K] the diffusional effects become important, controlling the
radiation by integrating the radiation intensity in all directions over
the solid angle.
Table 3
UZ¼4p Combustion kinetics for coal and biomass [4,28,32].
1 ! !
q4R ¼ $ Ið r ; s Þ dU (3)
4s Adaro Hatillo Sawdust
U¼0 Devolatilization parameters
E - Activation Energy [kJ kmol1] E1 ¼ 1.1  105 1.7  104
The absorption coefficient (εp) was calculated using the
E2 ¼ 1.7  105
Weighted Sum of Gray Gases Model (WSGGM) as a function of the A - Pre-exponential constant A1 ¼ 2.0  105 7.7  1011
concentration and partial pressures of H2O and CO2 and a charac- A2 ¼ 1.3  105
teristic length. The wall temperature was assumed to reach the Heterogeneous combustion parameters
steam saturation temperature, and the emissivity of the wall was Echar - Activation Energy [kJ kmol1] 8.3  104 9.25  104 1.2  105
C2 - Pre-exponential constant 2.0  103 2.35  103 9.3  103
considered to be equal to 0.75, obtained from the literature [27].

rez-Jeldres R, et al., A modeling approach to co-firing biomass/coal blends in pulverized coal utility boilers:
Please cite this article in press as: Pe
Synergistic effects and emissions profiles, Energy (2016), http://dx.doi.org/10.1016/j.energy.2016.11.116
6 R. Perez-Jeldres et al. / Energy xxx (2016) 1e12

reaction velocity [34]. To include both effects, the heterogeneous monitoring system (CEMS) under stationary conditions. For all of
char combustion was simulated using the kinetic/diffusion surface the operational parameters, the gas concentrations represent the
reaction model (Eq. (6)). mean value at the 95% level of confidence.

dmP rRg T∞ YO2 1 5. Results and discussion


¼ AP 1
(6)
dt MWO2 R þ D10
5.1. Model validation
The overall rate is determined by oxygen diffusion to the surface
(D0) and the rate of surface reaction (R), which depend on tem- The model was validated by comparing the simulated concen-
perature and composition of the gaseous environment, as well as trations of CO2, SO2 and O2 with experimental data obtained for the
the size, porosity and temperature of the particle [35]. The kinetic utility boiler (U14) operating with 100% coal and at the operation
rate (R) incorporates the effects of chemical reaction on internal conditions indicated in Table 2. In addition, the simulated and
char surfaces (intrinsic reaction) and pore diffusion. The oxygen experimental gas concentrations were compared with those
diffusion rate and the charcoal/char reaction rate are calculated calculated from the EPA AP 42 [40] standards using the operational
using the field correlations [36]: conditions (see Fig. 3). The comparison shows that, for the studied
 0:75 technology, the concentrations measured in U14 are similar to both
Tp þ T∞ the model-predicted values and those determined by the EPA
D0 ¼ C1 (7)
dP standards.
By comparing the simulation results for co-firing with EPA AP 42
 
E calculations using the same emission factors for coal combustion
R ¼ C2 exp  char (8) and co-firing, both results show a decrease in the concentrations of
RTP
CO2 and SO2 and an increase in the O2 percentage. The increase of
Because the pre-exponential factor and the activation energy 7% in O2 presents the possibility of reducing secondary air to
depend on fuel characteristics [35], the kinetic parameters were optimize the combustion and to reduce the power consumption of
calculated from Hurt and Mitchell's correlation [32] using the fuel the induced draft fan, thus promoting an increase in the plant
elemental analysis presented in Table 3. The diffusion rate is ex- efficiency.
pected to be higher in biomass-derived char particles because the To assure the independence of C-CO2 calculations from the
biomass has a higher volatile matter content than coal and burns selected model, a sensitivity analysis using the eddy dissipation
rapidly leaving a highly porous char, thus accelerating the char concept model (EDCM) and the PDF-transport species (PDF) was
combustion as well [13]. performed. The similarity in values obtained in the concentrations
For pulverized coal combustion, it has been shown that volatiles of CO (5.5  105 for EDM, 7.34  105 for EDCM and 8.36  108
are burned extremely quickly in areas close to the burner region [mg Nm3] for PDF) suggest that low concentrations of CO seem not
[37] and that the overall rate of reaction is controlled by turbulent to have a relationship with the EDM but could be associated with
mixing in the reaction zone (Damko €hller number ≪1). Under these the particle size distribution. This distribution was obtained from
assumptions, the eddy-dissipation model (EDM) is used to describe four samples in the discharge line from two of the four mills. It is
the entire VM combustion process. Applications on the industrial possible that the mills that were not sampled operate under
scale in pulverized coal combustion modeling show that EDM offers different particle size distributions than the sampled mills. How-
a practical and reasonably accurate prediction to describe the ho- ever, in practical terms, this difference would not affect the quali-
mogeneous combustion process [10,38,39]. tative analysis of the combustion. Considering the accuracy of the
To verify the convergence of the results, mass balance monitors model, it may be extended to the analysis of plants with similar
were created for carbon, sulfur and mass flow. The convergence combustion technology operating under steady-state conditions.
criteria of these monitors was 1  103, while the relative
convergence criteria for energy, continuity and momentum equa- 5.2. Contour analysis
tions were 1  106, 1  104 and 1  104, respectively. The tur-
bulence kinetic energy (k) and dissipation rate (ε) convergence The burnout of char particles is a reference index to assess the
criteria was 1  103. combustion performance of a fuel mixture under specific reaction
conditions. For level A of burners, the burnout zones for the co-
4. Case study firing case reveals a decrease as compared to a normal operation
(Fig. 4). This could be due to a higher content of VM in biomass in
International experience has demonstrated that a maximum of reference to that of coal (See Table 2). Abbas et al. [37] showed the
20 wt% substitution of coal by biomass in pulverized coal utility early ignition of sawdust volatiles produces a faster devolatilization
boilers can be possible without changing the design parameters rate of the coal particles. The heat liberated by the combustion of
[13]. In this research, we studied the performance of Unit 14 of sawdust VM will affect the surrounding coal particles through the
Tocopilla, operating with a 20% replacement of the coal mass flow acceleration of both heat transfer and the kinetics of the hetero-
rate by feeding biomass into the burner level A. This replacement geneous and homogeneous processes. Indeed, the previous state-
corresponds to 5% of the total feedstock rate. The simulated con- ment could be corroborated by analyzing the shape and area of the
ditions used in this study are presented in Table 4. These were burnouts zones for co-firing (Fig. 4), which suggest that a higher
proposed by the power plant and are quite conservative in order to number of particles have finished or are developing their com-
guarantee to his clients a continuous and steady supply of elec- bustion process as compared to 100% of coal combustion.
tricity during the co-firing. Because the replacement is mass-based, If the morphology of the burnout contour is analyzed in terms of
and thus the efficiency and the power level would be affected. To distances and areas, one of the remarkable characteristics observed
establish a comparative basis for emissions, between normal is that the average area bounding the burnouts near the burner and
operation (100% coal) and co-firing, the concentrations of CO and the average distance from the centroid area to the burner, both
SO2 are normalized by the power level. The flue gas concentrations remained constant (see Table 5). The analysis also revealed a
are measured in the chimney boiler using a continuous emissions reduction of 26% in the burnouts and 17% in the area occupied by

rez-Jeldres R, et al., A modeling approach to co-firing biomass/coal blends in pulverized coal utility boilers:
Please cite this article in press as: Pe
Synergistic effects and emissions profiles, Energy (2016), http://dx.doi.org/10.1016/j.energy.2016.11.116
R. Perez-Jeldres et al. / Energy xxx (2016) 1e12 7

Table 4
Operating conditions for the case study.

Operational variable/Burner level A B C D

Coal/Biomass flow rate [kg s1] 3.22/0.81 4.03/0.00 3.97/0.00 4.03/0.00


Primary/Secondary air flow [kg s1] 6.67/26.8 7.50/26.8 6.67/26.8 6.94/26.8
Primary/Secondary air temperature [K] 611/841 611/841 611/841 611/841
Boundary conditions.
Wall temperature [K] 373
Wall internal emissivity 0.75
Gauge pressure outlet [Pa] 3056

Note: The symbol “/” separates the value corresponding to one level of burners from another level.

Fig. 3. Flue gas concentrations from the model, experimental data and EPA AP 42 calculation.

Fig. 4. Contours of burnouts at the level of burners fed coal-sawdust blend (level A).

the burnouts into the flame vortex (zone enclosed by a dashed particles. The previous, suggests that combustion was brought
white line) associated with a major quantity of completely burned forward due to the catalytic effect of volatiles released from

rez-Jeldres R, et al., A modeling approach to co-firing biomass/coal blends in pulverized coal utility boilers:
Please cite this article in press as: Pe
Synergistic effects and emissions profiles, Energy (2016), http://dx.doi.org/10.1016/j.energy.2016.11.116
8 R. Perez-Jeldres et al. / Energy xxx (2016) 1e12

biomass, as reported by Gil et al. [41]. 5.3. Flame vortex analysis


The contour of gas-phase temperature for burner in level A
(Fig. 5) suggests that high-temperature zone is localized close to the Because the combusting particles are concentrated in the flame
burner, reaching a peak of 1820 [K] for pure coal combustion and vortex, the concentration of the flue gas and the temperature gra-
1710 [K] for co-firing. A slight decrease (7 [K]) in the mean tem- dients are analyzed within this vortex (Fig. 10).
perature allows inferring a better distribution of the flame tem- The calculations show a maximum reduction of 42% in the CO
perature. This trend will also benefit the ignition and burning of the concentration per unit of thermal power for co-firing but maintain
fuel particles as they are fed through the ignition ports to keep the the position of the maximum CO peak. The CO concentrations
heat flux close to the boiler water tubes. This temperature decrease reached lower values in level A of the burners, being consistent
was previously observed by Kastanaki and Vamvuka [42], during a with the unburned profiles presented in Fig. 4, except for those
lab-scale study, where a comparative study on the reactivity and zones close to the burner where, due to the devolatilization process,
kinetics of the combustion of coal-biomass blends using TGA, was CO concentrations were higher.
provided. Their results showed that for 0/20 wt% coal-wood blends, A lower presence of CO is commonly associated with higher
the burnout (final) temperature of the blends was lowered by 4e10 combustion efficiency and an increase in the concentration of CO2.
[K]. In this study, the increase in emissions of CO2 shown in Fig. 10 is due
The velocity field within the boiler is not excessively affected by to the lower thermal power fed into the co-firing of coal with
the inclusion of biomass, which can be corroborated by inspecting sawdust than for pure coal combustion. Co-firing assays at pilot
both the confinement of the flame and the vortex profile (Fig. 6). plant scale [37], in which the partial replacement of coal is based on
This visualization is consistent with the mean velocity values of flue energy, coincide in the increase of the CO2, which is a result of
gas at burner level A, of 11.7 and 11.3 [m s1] for coal combustion earlier devolatilization of the sawdust. Furthermore, it allows
and co-firing cases, respectively, indicating a 3% reduction of the accelerating the combustion of coal particles as well. However,
mean flue gas velocity. Similar results were obtained by Backreedy even if the synergistic effect described is not considered, sawdust
et al. in a 1 [MW] pulverized coal combustion test facility [43]. The inclusion enough for the same power fed than in the case of pure
study showed that if a smaller amount of the coal was replaced by coal will increase the emission of CO2, due to higher biomass flows
pinewood (3 wt%), the temperature and the velocity fields of the into the boiler to compensate the lower thermal power of that
base coal flame are not significantly changed because the replace- biomass, thus increasing the amount of carbon present in the
ment was made on a mass basis. At higher percentages re- combustor.
placements (20, 40 and 60%), Bhuiyan et al. [11] showed a decrease At the exit of the furnace, Fig. 10 shows a 5% decrease in SO2
in the velocity, temperature and CO2 profiles within the boiler, so as emissions, by unit of thermal power, which is attributed to the
an increased in the unburned carbon in the ash. lower Sulfur content of the coal-sawdust blend (0.57 [kg MWh1])
The simulations yield a mean turbulence intensity value of 15.8% than for pure coal (0.6 [kg MWh1]). In addition, a maximum of 10%
for coal and 20.1% for co-firing conditions (see Fig. 7). These values reduction of SO2 emissions is observed in the hopper zone. This
are close to those reported by Baxter [44] for pulverized coal reduction not only is due to lesser content of Sulfur in the blend. In
combustion (10% typical and 30% maximum). High levels of tur- fact, lower gas temperatures retard Sulfur oxidation and the higher
bulence are within expectations: the main feature of tangential reactivity of sawdust than of coal allows a rapid decrease of particle
distribution of burners (which is the case of Unit 14) is the gener- size and increasing entrainment which diminishes the possibility of
ation of high levels of turbulence, thus optimizing the fuel-air particles reaching the hopper [45].
mixture and improving the efficiency of the combustion. A comparison of the temperature profiles shows similar tem-
Figs. 8 and 9 reveal a slight decrease of 0.6% in the radiation and perature values from the burner area to the exit of the furnace. The
the total heat flux in the co-firing case. However, the mean maximum difference between the flame temperatures obtained
convective heat flux reached 32.8 [MW] for both cases. The mean from burning blends of biomass-coal and pure coal was 180 [K], and
convective heat flux was calculated by subtracting the radiation it occurred in the region of the hopper. For heights below 8 [m],
from the total heat flux. Therefore, the reduction in the total flux to lower temperatures were observed in the co-firing of coal-sawdust
the walls is due only to the decrease in the radiation flux, which is blends that in the combustion of pure coal. This trend is due to the
related to the decrease in the temperature of the flame, as a lower LHV of sawdust than of coal, as well as the sawdust is fed into
consequence of the inclusion of biomass (see Fig. 5). The total heat the first level of the burners (level A in Fig. 1), located at a height of
flux reveals a decrease in the heat flux on the hopper zone and a 9 [m] (see Fig. 10). In addition, the high reactivity and volatile
constant radiation heat flux (See radiation heat flux contour). content of sawdust, as compared to coal, allows a larger and faster
Therefore, the reduction in the total heat flux could be associated release of volatile matter, which burns at heights above the level of
with a decrease in the velocity located in the hopper. burners [7]. Thus, a displacement of the temperature profile and a
mild temperature-increase at the exit of the hopper was obtained.
What is remarkable about this result is that a small amount of coal
Table 5 replacement by sawdust allows keeping the temperature profile in
Results of the morphology analysis of the burnout contour e Burner level A. the regions of flow to the evaporator walls of the boiler.
Burner ID Combustion of pure coal Co-firing coal-sawdust
blends 5.4. Synergistic effect analysis
dburner [m] Area [m2] dburner [m] Area [m2]
The contour analysis, like the axial flame vortex variability, al-
E 1.07 1.4 1.14 1.4 lows the identification of synergistic effects during the co-firing
F 1.33 1.6 1.14 1.4
process. The decrease in the burnouts and the increase of the
G 1.18 1.2 1.12 1.1
H 1.00 1.6 1.34 1.9 temperature close to the water tubes lead to consideration of an
increase in the amount of fuel particles that has ended the com-
Mean 1.14 1.5 1.18 1.5
bustion process because the mean heat flux to the walls is slightly
Flame vortex n/c 19.9 n/c 16.5 decreased (0.5%).
Note: n/c e not calculated. The synergistic effect during co-firing of coal/biomass blends

rez-Jeldres R, et al., A modeling approach to co-firing biomass/coal blends in pulverized coal utility boilers:
Please cite this article in press as: Pe
Synergistic effects and emissions profiles, Energy (2016), http://dx.doi.org/10.1016/j.energy.2016.11.116
R. Perez-Jeldres et al. / Energy xxx (2016) 1e12 9

Fig. 5. Contours of mean temperature at the level of burners fed coal-sawdust blend level A).

Fig. 6. Contours of flue gas velocity at the level of burners fed coal-sawdust blend (level A).

Fig. 7. Contours of turbulence intensity at the level of burners fed coal-sawdust blend (level A).

rez-Jeldres R, et al., A modeling approach to co-firing biomass/coal blends in pulverized coal utility boilers:
Please cite this article in press as: Pe
Synergistic effects and emissions profiles, Energy (2016), http://dx.doi.org/10.1016/j.energy.2016.11.116
10 R. Perez-Jeldres et al. / Energy xxx (2016) 1e12

and therefore an increase in the local temperature, as shown in


Fig. 5. The increase in the gas temperature will allow the coal
particles to reach the devolatilization temperature earlier.
Analyzing the effect of temperature on the heterogeneous
combustion and using the kinetic/diffusion limited surface reaction
model, an augmentation of the gas temperature would result in an
increase in the temperature of the char/charcoal particles and a
decrease in the diffusional and surface kinetics resistances (see Eqs.
(6)e(8)). In a macroscopic pore model like the one used in this
study, the diffusional resistance is inversely proportional to the
average temperature between the particle and the environment
[46]. Thus, an increase in the temperature will induce the diffusion
of oxidant to the surface of the particle, favoring the reaction in
terms of oxidant availability. Meanwhile, the kinetic resistance is
inversely proportional to the surface reaction rate. For coal, this
reaction has been demonstrated to be of first order with an
Arrhenius-type dependence on the temperature. Furthermore, an
increase in temperature will lead to an exponential increase in the
reaction rate and a higher degree of conversion of the combusting
particles will be obtained [46], thus explaining the diminution of
the particle presence in the burnout contour and the CO concen-
tration and the increment in the CO2 concentrations (Figs. 4 and
10). According to the literature, similar effects are observed for
the combustion of char-charcoal blends. The literature reveals an
increase in the reactivity of the char particles is due to the increase
Fig. 8. Contours of the total heat flux to the water tubes walls of the boiler. in the pre-exponential factor, while the activation energy remains
practically invariable [42].

6. Conclusions

A CFD model was validated using gas concentration values ob-


tained from a commercial power plant and calculations using EPA
factors. For coal combustion, the predictions of CO2 and O2 showed
relative errors close to 6.0% and 1.1%, respectively, while the relative
error for SO2 was below 9%. Biomass was fed into Level A at 5%, and
the SO2 and O2 concentrations per unit of thermal power decreased
by 5.0% and 1.1%, respectively. The proportional SO2 decrease re-
flected the negligible sulfur content of biomass. The concentration
of CO was maintained near zero, and CO2 (9.4%) was similar to the
reference conditions. The mean temperature of the flue gas
increased by approximately 2%, while the heat transferred to the
tubes decreased slightly (0.6%), as expected. The burnout
morphology analysis revealed a displacement in the direction of
the burnout centroid and 17% of the area occupied by the burnout
into the flame vortex, which is associated with a major quantity of
completely burned particles.
Following up the synergistic effect demonstrated on the labo-
ratory scale, here we are demonstrating analyzing the combustion
process step by step on an industrial scale. This effect could be
associated with the heat release, physically manifested as a tem-
perature increase, resulting in catalyzed combustion of the coal
particles in terms of diffusional and kinetic resistance diminution
and expressed as a decrease in the burnout particles, as an increase
Fig. 9. Contours of radiation heat flux to the water-tube walls of the boiler.
in the particle conversion. We observed an increase in the CO2
concentration and the temperature of the flue gases and a decrease
in the CO2, CO and SO2 concentrations without greatly affecting the
vorticity, the residence time of the fuel particles, the radiative and
could be associated with an early release of heat during the com- convective heat fluxes, the steam production or the temperature of
bustion of the biomass particles. The higher biomass reactivity the flue gases. We observed a slight increase in the CO2 concen-
compared with coal, expressed in terms of activation energy and tration and the temperature of the flue gases and a decrease in the
pre-exponential factor (see Table 3), allows the biomass to begin CO and SO2 concentrations without greatly affecting the vorticity,
the devolatilization and homogeneous combustion process earlier the residence time of the fuel particles, the radiative and convective
than coal, producing a heat release in the zones close to the burners heat fluxes, or the steam production.

rez-Jeldres R, et al., A modeling approach to co-firing biomass/coal blends in pulverized coal utility boilers:
Please cite this article in press as: Pe
Synergistic effects and emissions profiles, Energy (2016), http://dx.doi.org/10.1016/j.energy.2016.11.116
R. Perez-Jeldres et al. / Energy xxx (2016) 1e12 11

Fig. 10. Profiles of pollutant emissions and mean temperature per unit of thermal power. Orange arrows shows the location of the level A of burners. (For interpretation of the
references to colour in this figure legend, the reader is referred to the web version of this article.)

Acknowledgments combustion in a 200 MW tangentially fired utility boiler. Energy Fuels


2012;26:313e23.
[11] Bhuiyan A a, Naser J. CFD modelling of co-firing of biomass with coal under
The authors acknowledge the staff of Engie Chile Company for oxy-fuel combustion in a large scale power plant. Fuel 2015;159:150e68.
their support in providing key information for the study, FONDEF [12] Tamura M, Watanabe S, Kotake N, Hasegawa M. Grinding and combustion
Project D09I1173, CONICYT - Becas Chile, CONICYT PAI 7813110021, characteristics of woody biomass for co-firing with coal in pulverised coal
boilers. Fuel 2014;134:544e53.
CONICYT/FONDAP Project-15130015 for its financial support, the [13] Sami M, Annamalai K, Wooldridge M. Co-firing of coal and biomass fuel
Technological Development Unit (UDT) and the Mechanical Engi- blends. Prog Energy Combust Sci 2001;27:171e214.
neering Department of University of Concepcion for providing the [14] Agraniotis M, Nikolopoulos N, Nikolopoulos A, Grammelis P, Kakaras E. Nu-
merical investigation of Solid Recovered Fuels' co-firing with brown coal in
computing resources. large scale boilers - evaluation of different co-combustion modes. Fuel
2010;89:3693e709.
[15] Yin C, Rosendahl L, Kaer S, Condra T. Use of numerical modeling in design for
References co-firing biomass in wall-fired burners. Chem Eng Sci 2004;59:3281e92.
[16] Smoot LD, Pratt DT. Pulverized-coal combustion and gasification: theory and
[1] Bo€hringer C, Hoffmann T, Rutherford TF, Arbor A. Alternative strategies for applications for continous flow processes. Springle U. S; 1979.
promoting renewable energy in eu electricity markets. Appl Econ Quat 2007: [17] Biagini E, Lippi F, Petarca L, Tognotti L. Devolatilization rate of biomasses and
9e26. coal-biomass blends: an experimental investigation. Fuel 2002;81:1041e50.
[2] Chilean Mandatory #20257. “Modificaciones a la ley general de servicios [18] Moghtaderi B, Meesri C, Wall TF. Pyrolytic characteristics of blended coal and
ctricos respectos de la generacio
ele n de energía ele ctrica con fuentes de woody biomass. Fuel 2004;83:745e50.
energía renovables no convencionales”. [n.d]. [19] Vuthaluru HB. Investigations into the pyrolytic behaviour of coal/biomass
[3] Berg A, Bidart C, Espinoza D, Flores M, Moraga A, Müller N, et al. Reco- blends using thermogravimetric analysis. Bioresour Technol 2004;92:187e95.
mendaciones para la elaboracio n de una Estrategia Nacional de Bioenergía. [20] Lu KM, Lee WJ, Chen WH, Lin TC. Thermogravimetric analysis and kinetics of
Minist Energía 2013. co-pyrolysis of raw/torrefied wood and coal blends. Appl Energy 2013;105:
s J, Gil A, Corte
[4] Pallare s C, Herce C. Numerical study of co-firing coal and Cynara 57e65.
cardunculus in a 350 MWe utility boiler. Fuel Process Technol 2009;90: [21] Haykiri-Acma H, Yaman S. Interaction between biomass and different rank
1207e13. coals during co-pyrolysis. Renew Energy 2010;35:288e92.
[5] Yin C, Kaer S, Rosendahl L, Hvid S. Co-firing straw with coal in a swirl- [22] Wu Z, Wang S, Zhao J, Chen L, Meng H. Product distribution during Co-
stabilized dual-feed burner: modelling and experimental validation. Bio- pyrolysis of bituminous coal and lignocellulosic biomass major components
resour Technol 2010;101:4169e78. in a drop-tube furnace. Energy & Fuels 2015;29:4168e80.
[6] Mandø M, Rosendahl L, Yin C, Sørensen H. Pulverized straw combustion in a [23] Wang M, Tian J, Roberts DG, Chang L, Xie K. Interactions between corncob and
low-NOx multifuel burner: modeling the transition from coal to straw. Fuel lignite during temperature-programmed co-pyrolysis. Fuel 2015;142:102e8.
2010;89:3051e62. [24] Xu C, Hu S, Xiang J, Zhang L, Sun L, Shuai C, et al. Interaction and kinetic
[7] Karampinis E, Nikolopoulos N, Nikolopoulos A, Grammelis P, Kakaras E. Nu- analysis for coal and biomass co-gasification by TG-FTIR. Bioresour Technol
merical investigation Greek lignite/cardoon co-firing in a tangentially fired 2014;154:313e21.
furnace. Appl Energy 2012;97:514e24. [25] Arteaga-pe rez LE, Vega M, Rodríguez LC, Flores M, Zaror CA, Casas Y. Energy
[8] Ma L, Gharebaghi M, Porter R, Pourkashanian M, Jones J, Williams A. Modelling for Sustainable Development Life-Cycle Assessment of coal e biomass based
methods for co-fired pulverised fuel furnaces. Fuel 2009;88:2448e54. electricity in Chile : focus on using raw vs torre fi ed wood. Energy Sustain Dev
[9] Ghenai C, Janajreh I. CFD analysis of the effects of co-firing biomass with coal. 2015;29:81e90.
Energy Convers Manag 2010;51:1694e701. [26] Morsi S. a., Alexander a. J. An investigation of particle trajectories in two-
[10] Fang Q, Musa AAB, Wei Y, Luo Z, Zhou H. Numerical simulation of multifuel phase flow systems. J Fluid Mech 2006;55:193.

rez-Jeldres R, et al., A modeling approach to co-firing biomass/coal blends in pulverized coal utility boilers:
Please cite this article in press as: Pe
Synergistic effects and emissions profiles, Energy (2016), http://dx.doi.org/10.1016/j.energy.2016.11.116
12 R. Perez-Jeldres et al. / Energy xxx (2016) 1e12

[27] Mueller C, Selenius M, Theis M, Skrifvars B-J, Backman R, Hupa M, et al. [38] Xu G, Zhou W, Swanson LW, Moyeda DK, Nguyen Q. Evaluation of applying
Deposition behaviour of molten alkali-rich fly ashesddevelopment of a sub- low calorific fuel as reburn fuel in an opposed wall fired boiler. J Therm Sci
model for CFD applications. Proc Combust Inst 2005;30:2991e8. Eng Appl 2009;1:31007.
[28] Kobayashi H, Howard J, Sarofim A. Symp. Combust. Coal devolatilization at [39] Lundmark D, Mueller C, Skrifvars B-J, Hupa M. Computational fluid dynamic
high temperatures, vol. 16. Elsevier; 1977. p. 411e25. modeling of combustion and ash deposition in a biomass-cofired bubbling
[29] Du X, Gopalakrishnan C, Annamalai K. Ignition and combustion of coal particle fluidized bed boiler. Clean Air Int J Energy a Clean Environ 2007;8:155e69.
streams. Fuel 1995;74:487e94. [40] US Enviromental Protecion Agency-EPA. Compilation of air pollutant emission
[30] Kobayashi H. Coal devolatilization of pulverized coal at high temperature. factor - AP42. In: Stationary point and area sources. fifth ed., vol. I; 1995.
Massachusetts Institute of Technology; 1976. [41] Gil MV, Casal D, Pevida C, Pis JJ, Rubiera F. Thermal behaviour and kinetics of
[31] Baum MM, Street PJ. Predicting the combustion behaviour of coal particles. coal/biomass blends during co-combustion. Bioresour Technol 2010;101:
Combust Sci Technol 1971;3:231e43. 5601e8.
[32] Hurt RH, Mitchell RE. Unified high-temperature char combustion kinetics for a [42] Kastanaki E, Vamvuka D. A comparative reactivity and kinetic study on the
suite of coals of various rank. Symp Combust 1992;24:1243e50. combustion of coal-biomass char blends. Fuel 2006;85:1186e93.
[33] Durie A. The science of victorian brown Coal : structure, properties and con- [43] Backreedy R, Fletcher L, Jones J, Ma L, Pourkashanian M, Williams A. Co-firing
sequences for utilization. Butterworth-Heinemann; 1991. pulverised coal and biomass: a modeling approach. Proc Combust Inst
[34] Batchelder HR, Busche RM, Armstrong WP. Kinetics of coal gasification pro- 2005;30:2955e64.
posed mechanism of gasification. Ind Eng Chem 1953;45:1856e78. [44] Baxter L. Ash deposit formation and deposit properties. In: A comprehensive
[35] Wang Y, Yan L. CFD studies on biomass thermochemical conversion. Int J Mol summary of research conducted at Sandia's combustion research facility;
Sci 2008;9:1108e30. 2000.
[36] Field MA. Rate of combustion of size-graded fractions of char from a low-rank [45] Wei X, Guo X, Li S, Han X, Schnell U, Scheffknecht G, et al. Detailed modeling
coal between 1 200K and 2000K. Combust Flame 1969;13:237e52. of NO x and SOX formation in Co-combustion of coal and biomass with
[37] Abbas T, Costen P, Kandamby NH, Lockwood FC, Ou JJ. The influence of burner reduced kinetics. Energy Fuels 2012;26:3117e24.
injection mode on pulverized coal and biomass co- fired flames. Combust [46] Smith I. The Combustion rates of coal chars: a Review. Combustion 1982:
Flame 1994;99:617e25. 1045e65.

rez-Jeldres R, et al., A modeling approach to co-firing biomass/coal blends in pulverized coal utility boilers:
Please cite this article in press as: Pe
Synergistic effects and emissions profiles, Energy (2016), http://dx.doi.org/10.1016/j.energy.2016.11.116

You might also like