Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Fuel 113 (2013) 780–786

Contents lists available at SciVerse ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Effects of firing coal and biomass under oxy-fuel conditions in a power


plant boiler using CFD modelling
S. Black, J. Szuhánszki, A. Pranzitelli ⇑, L. Ma, P.J. Stanger, D.B. Ingham, M. Pourkashanian
Energy Technology and Innovation Initiative (ETII), Faculty of Engineering, University of Leeds, Leeds, LS2 9JT, United Kingdom

h i g h l i g h t s

 A CFD model was developed to investigate oxy-fuel conditions for coal and biomass in a 500 MWe utility boiler.
 The simulation was used as a predictive tool for retrofitting a utility boiler.
 Heat transfer to the tube banks as well as temperature increases with oxygen concentration.
 The results suggested that heat transfer similar to air firing could be achieved by appropriate O2 enrichment.
 The large particle size of biomass limited burnout and heat transfer in air and under oxy-fuel conditions.

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

Article history: One of the most promising technologies for carbon capture and storage (CCS) is oxy-fuel combustion. This
Received 1 October 2012 study uses a commercial computational fluid dynamics (CFD) code to simulate the firing of coal and
Received in revised form 26 March 2013 biomass under air and oxy-fuel conditions in an existing full-scale 500 MWe coal-fired utility boiler.
Accepted 27 March 2013
Results are presented for conventional air–coal combustion that corresponds well against available
Available online 13 April 2013
experimental data and an in-house empirical model. Maintaining the same thermal input and exit oxygen
concentration, CFD was used as a predictive tool with standard physical submodels, to examine the
Keywords:
effects of firing under air–biomass, oxy-coal and oxy-biomass conditions. The oxy-fuel conditions were
Biomass
Coal
investigated at oxygen concentrations of 25% and 30% by volume for a wet flue gas recycle. The effects
CFD of firing biomass in both air and oxy-fuel conditions are predicted to have a lower total heat transfer
Carbon capture to the tube walls, with a lower furnace exit temperature in the boiler than the coal-fired cases. This
Oxy-fuel may be attributed to the effects of large biomass particles, which have a lower total surface area and
therefore causes a reduction in the radiative heat transfer to the tube walls as well as an increase in
carbon in ash (CIA) predictions. For oxy-coal firing, the study suggested that the optimum oxygen
concentration for heat transfer to be closely matched with air–coal, lies between 25% and 30%, but for
oxy-biomass firing a value greater than 30% may be needed. This study highlights the possible impact
of changing the fuel and combustion atmosphere on the heat transfer characteristics of an existing power
plant boiler, underlining that minor redesign may be necessary when converting to biomass firing under
air and oxy-fuel conditions.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction Biomass is also seen as a flexible resource and can increase


energy security when part of a balanced energy mix [4]. Biomass
Coal is one of the main sources of energy used in power gener- has the potential for near zero CO2 emissions, provided that it is
ation and accounts for approximately 40% of the total energy pro- sustainably produced. This means that the harvested biomass is
duced [1]. However, coal fired power stations are a major source of replaced with another generation of plants, which removes the
CO2 emissions that contribute to anthropogenic climate change. CO2 from the atmosphere. Moreover, biomass fuels include various
The utilisation of biomass, either co-firing or replacing the coal en- forest or agricultural residues, which otherwise would decay to
tirely, and carbon capture and storage technologies such as oxy- produce methane, which has a significantly higher greenhouse
fuel combustion in coal-fired power plants, can help reduce CO2 gas impact than CO2. Therefore, directly fired biomass residue
emissions in line with international and UK government targets can further reduce greenhouse gas emissions [5]. Biomass fuels
[2,3]. generally contain significantly less sulphur and nitrogen compared
to coal, reducing SOx and NOx emissions, and together with the
increased levels of UK government support under the revised
⇑ Corresponding author. Tel.: +44 (0)113 343 2481; fax: +44 (0)113 246 7310. Renewable Obligations Scheme [6], they provide additional
E-mail address: A.Pranzitelli@leeds.ac.uk (A. Pranzitelli). incentives for full conversion to biomass firing.

0016-2361/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.fuel.2013.03.075
S. Black et al. / Fuel 113 (2013) 780–786 781

Nomenclature

Roman letters ep particle emissivity (dimensionless)


A pre-exponential factor (1/s) r Stefan Boltzmann constant (W/m2-K4)
Ap particle surface area (m2) hR radiation temperature (K)
cp particle specific heat (kJ/kg-K)
dp particle diameter (m) Abbreviations
EA activation energy (J/kmol) ASU air separation unit
f fraction of heat absorb by the particle (dimensionless) CCS carbon capture and storage
h film heat transfer coefficient (W/m2-K) CFD computational fluid dynamics
Hr heat of reaction released (J/kg) CIA carbon in ash
k turbulent kinetic energy (m2/s2) RFG recycled flue gas
mp mass of particle (kg) FRH final reheater
SF shape factor (dimensionless) RANS Reynolds-averaged Navier–Stokes
t time (s) SSH secondary superheaters
T temperature (K) WSGGM weighted sum of gray gases model

Greek letters
e dissipation of turbulent kinetic energy (m2/s3)

Oxy-fuel combustion is an emerging CCS technology for new to allow the tube banks to be visible. The dimensions of the
and existing power stations where the fuel is burned in a mixture modelled section are approximately 50 m (height)  30 m
of oxygen and recycled flue gas (RFG) instead of air. This produces a (width)  9 m (depth).
flue gas consisting mainly of carbon dioxide and water vapour, The boiler is equipped with 48 Doosan Babcock Mark-III
which after subsequent purification can be compressed and trans- Low-NOx coal burners arranged in 4 rows of 12. A simplified
ported to a suitable geological storage site or utilised, for example geometry is used for each burner, as shown in Fig. 2, to reduce
for enhanced oil recovery [7]. Oxy-fuel technology can be retrofit- the overall computational cost. The coal is introduced in the
ted to existing power stations, recycling a portion of the exit gas primary registers with carrier air, and the secondary and tertiary
into the boiler to moderate the flame temperature. It has been sug- registers deliver swirled combustion air. The swirl is described
gested in the literature that by adjusting the amount of recycle and by a swirl angle to model the effect of the blades, and is alternated
oxygen concentration it is possible to achieve similar radiative and between the adjacent burners.
convective heat transfer profiles as for air firing [8]. Finally, com-
bining oxy-fuel technology with biomass firing can achieve an
overall negative emissions balance, with the potential to remove 2.1. Source of data for CFD input and comparison
800 million tonnes of CO2 from the atmosphere every year by
2050, using only sustainable biomass [9]. However, there is limited Experimental data and predictions from a RWE in-house power
experience in firing coal and biomass in an industrial boiler under station modelling tool (based on [17]) were used to validate the
oxy-fuel conditions, and the use of numerical methods such as CFD predictions for a selected set of conditions involving the com-
computational fluid dynamics (CFD) can help explore technical bustion of coal in air. The operating conditions of the air–coal case
challenges, such as heat transfer, inside a boiler. are shown in Table 1. The coal used was Pittsburgh 8 and its
CFD is an engineering design tool that can give insight into the
boiler, for example flow field, temperature and chemical species
distributions that would be difficult to obtain experimentally.
CFD for combustion problems has been successfully applied to
air–coal [10,11], oxy-coal [11,12], air–biomass [13] and air–coal–
biomass blend combustion [14], however most of this work is on
pilot scale, based on experimental rigs. To the authors’ knowledge,
there has been little modelling work on oxy-biomass combustion
and no published work on a full-scale utility boiler. CFD is there-
fore used in this study to evaluate the effects on heat transfer, from
the firing of biomass and coal under air and oxy-fuel conditions in
an existing 500 MWe sub-critical coal-fired boiler within the UK
power sector.

2. Utility boiler

The boiler considered in this study is a 500 MWe subcritical


coal-fired utility boiler located at Didcot Power Station in the UK.
The dimensions and details of the boiler are given in [15,16], but
a brief description will be given here. The boiler consists of the fur-
nace, primary superheater platens, the secondary superheater
(SSH) and the final reheater (FRH). The geometry after the FRH is
not considered and due to the symmetry of the boiler, half of the
geometry was modelled. In Fig. 1, the top section has been opened Fig. 1. The boiler geometry.
782 S. Black et al. / Fuel 113 (2013) 780–786

gated. These considered coal or biomass in air or under an oxy-fuel


condition with 25% or 30% volumetric inlet oxygen concentration.
The cases are named in accordance with the environment, followed
by the fuel type. The investigated conditions are outlined in Tables
1 and 3. Four variables are kept constant throughout each case: the
thermal input, the exit oxygen concentration, number of firing
Fig. 2. Simplified burner geometry. burners and air leakage. Coal was maintained to be the same as
in the air–coal case (Pittsburgh 8) and biomass was chosen to be
Table 1 dried pelletised wood with properties summarised in Table 2.
Operating conditions for all test cases. The mass flow of coal was kept constant for all of the oxy-coal
Fuel Air Oxy25 Oxy30 cases, whilst the flow rate of biomass was increased to maintain
the same thermal input. In reality there are constraints on the
Fuel feed rate (kg/s) Coal 46.7 46.7 46.7
Biomass 73.2 73.2 73.2 amount of biomass that can be carried by the primary air, but for
Gas feed rate (kg/s) Coal 529 500 402 the purposes of this study it was assumed possible. The pulverised
Biomass 534 461 375 coal particles were assumed to be spherical in shape and their size
Recycle ratio (%, wet) Coal 0 71 65 ranged between 1 lm and 350 lm, with the mean diameter of
Biomass 0 65 58
Exit O2 (vol%, dry) All fuels 5 5 5
75 lm. However the size of the fibrous wood, assumed to be cylin-
Thermal input (MWth) All fuels 1275 1275 1275 drical here, was significantly larger, with particle lengths ranging
Air leakage (kg/s) All fuels 16 16 16 from 90 lm up to 3 mm and the length to diameter ratio was as-
Operating burners All fuels 36 36 36 sumed to be 10.
For the oxy-fuel cases, a wet-recycle was assumed and since the
exit oxygen concentration was maintained at 5%, the 25% and 30%
cases corresponded to recycle ratios of 71% and 65% for coal, and,
Table 2
Coal and biomass properties. 65% and 58% for biomass. The oxygen is assumed to be supplied
by an air-separation unit (ASU) with a purity of 95% and 5% inert
Pittsburgh 8 Wood
gases. The oxygen concentration in the primary register was fixed
Ultimate analysis (wt%) (daf) at 21% by volume, as O2 enrichment of the primary stream would
Carbon 83.4 52.3
increase the risk of fire and dust explosions [7], while the second-
Hydrogen 5.5 6.4
Oxygen (by diff.) 6.9 41.1 ary and tertiary registers were enriched to achieve either a 25% or
Nitrogen 1.6 0.2 30% overall volumetric oxygen concentration for the oxy25 and
Sulphur 2.6 Trace oxy30 cases respectively. The composition of the RFG was calcu-
Proximate analysis (wt%) (ar) lated from an overall mass balance for each case.
Fixed carbon 50.3 18.9
Volatile matter 31.0 72.6 3. Computational modelling
Ash 10.3 5.7
Moisture 8.4 2.8 A commercial CFD package, ANSYS Fluent version 14.0 [18], was
GCV (MJ/kg) 28.54 18.90
used to model the geometry. A grid independent study was per-
formed on 3 different meshes consisting of 3.2, 4.1 and 4.8 million
cells, referred to as coarse, medium and fine, respectively, where
properties have been summarised in Table 2. The mass flow rates, predictions of heat transfer and furnace exit temperature were
temperatures and oxygen concentrations for the primary, second- used as criteria. The medium and fine meshes performed similarly,
ary and tertiary annulus are given in Table 3. Air leakage into the and therefore the medium mesh was adopted for this study. The
boiler was also assumed at 16 kg/s for the section of the boiler mesh is multiblock unstructured, finer in the burner regions, and
investigated, and was assumed to come through the hopper region consists of mostly hexahedral cells. A small region of polyhedral
corresponding to disruptions in the water seal. cells is present at the exit of the furnace section.

2.2. Investigated cases 3.1. Turbulence

Following the comparison of the air–coal case with the available A steady state Reynolds-averaged Navier–Stokes (RANS) model
data outlined in Section 2.1, five cases were numerically investi- was used where turbulence closure was given by the realizable k–e

Table 3
Operating conditions for the burners for all cases.

Air–coal Air–biomass Oxy25-coal Oxy25-biomass Oxy30-coal Oxy30-biomass


Mass flow rate (kg/s)
Primary 2.9 2.9 2.8 2.6 2.2 2.1
Secondary 2.2 2.2 2.1 1.9 1.7 1.6
Tertiary 9.5 9.7 9.0 8.3 7.3 6.8
Temperature (K)
Primary 363 363 363 363 363 363
Secondary 530 530 530 530 530 530
Tertiary 530 530 530 530 530 530
Oxygen concentration (mass %)
Primary 23.2 23.2 19.1 20.6 19.1 20.6
Secondary 23.2 23.2 23.8 25.5 29.7 31.7
Tertiary 23.2 23.2 23.8 25.5 29.7 31.7
S. Black et al. / Fuel 113 (2013) 780–786 783

Table 4 3.3. Homogeneous combustion


Mean values for the air–coal case at various positions in the boiler.

Exp. In-house model CFD The combustion of volatiles evolving from the devolatilising
Air–coal fuel particles was modelled using the Eddy Dissipation Model
Heat transfer (MW) [32], using a two-step global mechanism as follows:
Water walls – 456 457 Volatiles þ O2 ! CO þ H2 O
Platen 1 – 106 99
Platen 2 – 110 136
SSH – 110 101 CO þ 0:5O2 ! CO2
FRH – 79 52
Total – 861 846
This model assumes that combustion is controlled by the large-
eddy mixing time scale, k/e, and combustion proceeds whenever
Temperature (K)
Furnace exit 1591 1656 1670
turbulence is present (k/e > 0).
Platen 1 exit plane – 1135 1208
Platen 2 exit plane – 1282 1299 3.4. Particle combustion
SSH exit plane – 1173 1140
FRH exit plane/outlet – 1054 1094 The methodology used to describe the combustion of particles
has been described in previous papers for coal [10,11] and for bio-
mass [13,33]. The fluid flow was modelled with the conventional
Eulerian treatment, whilst the motion of the coal and biomass
particles was tracked within a Lagrangian frame of reference.
model. Since no detailed flow measurements were available, it was Exchange of momentum, heat and mass transfer between the
not possible to validate the turbulence model chosen. However, the phases were accounted for using the source/sink terms in the
realizable k–e model has been used successfully by other groups governing equations for the two phases.
modelling coal and oxy-coal combustion [14,19]. It also performed The particle combustion is divided into the following distinctive
better when the convergence of the solution was assessed, with steps: inert heating, drying, devolatilisation, char combustion, and
other more numerically stiff models (RNG k–e and RSM) resulting inert heating/cooling of the ash particles. The particle temperature
in significant fluctuations when local temperature and velocity is evaluated using the following equation:
were monitored within the flame zone.
dT p dmp
mp cp ¼ hAp ðT 1  T p Þ þ ep Ap rðh4R  T 4p Þ  f Hr ð1Þ
dt dt
3.2. Heat transfer where mp, cp, Tp, Ap, ep are the mass, specific heat, temperature, area
and emissivity of the particles, T1 is the temperature of the gas, hR,
Radiation is the dominant form of heat transfer in the furnace is the radiation temperature, Hr is the heat of reaction released by
and was modelled using the Discrete Ordinates (DO) model [20], the surface reaction and f is the fraction of heat absorbed by the
with three directions (3  3) to discretize each octant of angular particle. The film heat transfer coefficient, h, was evaluated with
space. The gas absorption coefficient was calculated with the do- the correlation of Ranz and Marshall [34], based on the particle
main based Weighted Sum of Gray Gases Model (WSGGM) Reynolds number and Prandtl number of the continuous phase
[21,22]. The WSGGM has been successfully applied in numerous assuming a spherical shaped particle.
CFD studies involving coal combustion in air [10,11] where due The rate of devolatilisation was calculated using the single
to the high level of non-absorbing gases, the mixture can be as- kinetic rate model [35], where the volatile release depends on
sumed to be gray. However, under oxy-fuel conditions, this the temperature history of the particle. The Arrhenius rate con-
assumption is no longer valid due to presence of high concentra- stants, pre-exponential factor, A, and activation energy, EA, used
tions of CO2 and H2O. Several modifications and new constants for Pittsburgh 8 coal were 3.8  1014 and 2.3  108 J/kmol respec-
have been proposed to improve the applicability of the original tively. These were calculated with the FG-DVC code, assuming
WSGG model to oxy-firing, which are available in the literature the heating rate of 105 K/s, representative of typical of pulverised
[23–27]. The full spectrum correlated k-distributions (FSCK) model coal flames [36]. For wood the values of A = 6.0  1013 and
tested by the authors [28] has proved to be able to provide more EA = 2.5  108 J/kmol were used based on [13].
accurate results in a typical oxy-fuel with RFG environment com- The intrinsic char combustion model [37] was used to model
pared to the WSGGM and its usage is therefore advised. The major the burnout of the char particles. This model assumes that the
limitation of implementing this model is related to the computa- oxygen order of the surface reaction is unity and that the surface
tional resources required, as discussed in [28]. For this work, due reaction rate includes the effects of both bulk diffusion and
to the scale of the boiler considered and the computational re- chemical reaction rates. The model constants for Pittsburgh 8
sources required by the FSCK model, the standard WSGGM was and wood were obtained from [36] and [5], respectively.
used, however with a cautionary note on the impact of the radia- Coal and biomass particles were assumed to be spherical and
tion on the predictions. cylindrical in shape, respectively, based on experimental observa-
Soot formation was modelled by the coal-derived Moss-Brookes tions [5]. When modelling the combustion of biomass, a shape fac-
soot model [29,30] and soot-radiation interaction was also in- tor, SF, was used to account for the effects of deviation of shape.
cluded [31]. This is defined as the ratio of surface area of an equivalent sphere
Heat transfer to the walls and tube walls were modelled as thin to the surface area of a cylindrical particle. When calculating heat
walls and an effective wall resistivity was used, which incorpo- transfer to and from the wood particle, the surface area in Eq. (1) is
rated the resistivity due to the deposition layer, metal tube wall modified according to the particle shape factor as:
and steam side film heat transfer coefficient. A previous study 2
1=4pdp
found that the value of 330 W/m2 K, which lies within the range Ap ¼
suggested by industry, was an appropriate representation of the SF
overall resistivity [15]. The outer temperature was set to the steam where Ap and dp are the area and spherical diameter of the equiva-
temperature inside the tubes, and the wall emissivity was assumed lent volume particle. The method of Haider and Levenspiel [38],
to be 0.8. described in more detail in [33], is used to determine the drag
784 S. Black et al. / Fuel 113 (2013) 780–786

Fig. 4. Prediction of heat transfer for air–coal and air–biomass.

thermal radiation is proportional to the fourth power of tempera-


ture. A cross section of the temperature distribution through the
third column of burners and the furnace exit temperature is shown
in Fig. 3a. The four separate rows of flames can be clearly seen,
where the release and combustion of volatiles takes place. The
burnout of the char particles is then completed in the region above
the volatile flames in the combined stream of hot gases rising
towards the superheaters. Comparison with available data in
Table 4 shows the heat transfer to the water walls and the temper-
atures at each section are matched closely, but the predicted distri-
bution of heat transfer differs more between the superheaters, FRH
and SSH sections. This can be possibly attributed to the differences
in modelling approaches between the CFD code and in-house
model. Without further experimental data, such as flow field or
chemical species measurements, it is difficult to fully validate the
CFD predictions. However, the total heat transfer to all the heat
exchanger sections considered is predicted with a 2% accuracy
between the CFD and in-house models. The CFD results are there-
fore deemed to give a good representation of the heat transfer in
the boiler and are used for comparison in subsequent sections.

4.2. Air–biomass

Fig. 4 shows the effect on heat transfer at various sections in the


boiler when biomass is fired instead of coal in air when thermal in-
put, exit oxygen concentration and air in-leakage are kept the same
as the air-coal case. Oxygen in the biomass was assumed to be
available for combustion and inlet velocities were similar. The re-
sults show a reduced heat transfer to all of the tube bank sections
compared to the air–coal case, resulting in an overall lower total
heat transfer by approximately 15%.
Comparing Fig. 3a and b, a different temperature distribution
and lower furnace exit temperature for air-biomass can be seen.
Fig. 3. Temperature distribution (a) air–coal, (b) air–biomass, (c) oxy25-coal, (d) High temperature regions near the burner exit corresponds to
oxy25-biomass, (e) oxy30-coal, and (f) oxy30-biomass. the intense volatiles flames that are shown in both cases. However
when comparing the burnout region above the flames the temper-
ature of the burnout zone is visibly reduced for the biomass case.
coefficient of the wood particles using the shape factor mentioned This may be attributed to differences in particle size, as coal parti-
above. cles are much finer and char combustion takes place mainly in the
furnace, however, the combustion of large biomass particles will be
4. Results and discussion slower and therefore the temperature in the burnout zone, and be-
fore the exit of the furnace, is reduced. Predictions of the carbon in
4.1. Air–coal ash (CIA) evaluated after the FRH section showed higher values for
the biomass than for coal, with values of 11% and 0.64%, respec-
The CFD predictions of heat transfer to the tube bank walls and tively. An increase in CIA is expected when firing biomass, due to
the gas temperatures leaving the tube bank sections are shown presence of larger particles as well as the ash content being gener-
against available data in Table 4, and are in good agreement with ally much lower for biomass, which results in the remaining car-
the in-house model. Heat transfer to the water walls is dominated bon taking up a larger percentage of the fly ash.
by radiation, and the rate of radiative heat transfer is closely The lower temperatures also have a direct impact on the radia-
related to the correct resolution of the temperature field, since tion and results in the lower heat transfer to the tube banks, as
S. Black et al. / Fuel 113 (2013) 780–786 785

within the furnace. The plots of temperature for the oxy25-coal


and oxy30-coal cases are shown in Fig. 3c and e respectively. As ex-
pected, the increase in O2 concentration at the inlet, which is
achieved by reducing the dilution of the recycle from 71% to 65%,
increases the overall temperature observed in the boiler. As the
temperature in the furnace section increases, the radiative heat
transfer to the water walls and platen superheaters also increase.
We can expect as a result of a lower inlet gas mass flow passing
through the furnace (Table 1), as the amount of recycle is reduced,
which would lead to less heat transfer to the convective sections
(not modelled here). This trend has also been observed in pilot
scale experiments [8].
For the purpose of retrofitting an existing coal boiler, it is
Fig. 5. Heat transfer (MW) for oxy25-coal and oxy30-coal compared against air–
important to match the radiative and convective heat transfer of
coal.
the boiler, and also to maintain the gas temperatures similar to
the design conditions [7]. The results imply that both temperature
and heat transfer profiles could be matched closely to air combus-
tion, by the appropriate selection of O2 enrichment, with the opti-
mum level lying between the modelled 25% and 30%, which agrees
with values in the literature [8].

4.4. Oxy-biomass

Results from the oxy-biomass cases are compared with air–bio-


mass study for the differences in heat transfer in Fig. 6. A higher
concentration of oxygen leads to an increased transfer of heat to
the tube walls due to higher temperatures. This is the same trend
observed for the oxy-coal cases, outlined in Section 4.3. Tempera-
Fig. 6. Heat transfer (MW) for oxy25-biomass and oxy30-biomass compared ture contour plots of oxy25-biomass and oxy30-biomass are
against air–biomass. shown in Fig. 3d and f, respectively. The increased oxygen concen-
tration appears to help the intensity of the flame, which is a result
of a faster rate of devolatilisation near the burner and lower inlet
discussed in Section 4.1 and observed in Fig. 4. As mentioned pre- mass flow rates inside the furnace, shown in Table 1. However,
viously, the differences in temperature distribution may be attrib- in comparison with the air–coal case, the total heat transfer is low-
uted to the larger biomass particles. These have an overall lower er in the oxy25-biomass and oxy30-biomass cases by 17% and 8%
total surface area compared to finely ground coal, which leads to respectively. Therefore, for oxy-biomass combustion increasing
a reduction in radiative emissions from the particles and subse- oxygen concentration levels above 30% could be used to achieve
quently heat transfer to the tube banks. As the amount of ash is similar heat transfer characteristics, compared to air-coal.
also significantly lower in the biomass, the radiative emissions
from fly ash particles passing the superheaters and subsequent
convective sections is also reduced. On the basis that particle emis- 5. Conclusion
sions play a significant role in radiative heat transfer [39], it can be
argued that the reduction in particle emissions are one of the In this study, a full-scale utility boiler has been simulated using
causes of lower heat transfer being observed in Fig. 4. Fig. 3b sug- CFD with standard physical submodels. The simulation was vali-
gests the particle size may also have an impact on flame stability dated under air–coal conditions against experimental data and
and shape, since the biomass flames appear less symmetric than an in-house power station model. By using a validated CFD simu-
in the case of coal-firing. lation as a predictive tool it was possible to examine the changes
Since the physical models in CFD neglect thermal gradients in heat transfer and temperature when firing coal or biomass under
within the particle, which may be important for larger particles air or oxy-fuel conditions with a wet RFG. For all of the investigated
[14], as well as simplifications to particle size distribution and cases, the thermal input, exit oxygen concentration, air leakage and
shape, the results are not quantitative. However, the results do number of operating burners were kept the same. The predictions
suggest that minor redesign may be needed for some elements of suggested the firing of biomass would result in a lower heat trans-
the boiler when switching to biomass firing, such as dedicated bio- fer to the tube walls in both air and oxy-fuel conditions compared
mass burners and better milling techniques to reduce the particle to coal firing, which could be attributed to the effects of large bio-
size and, increase the total surface area. mass particles such as a lower total surface area and an increase in
CIA. For oxy-biomass combustion it would appear increasing
oxygen concentration above 30%, could lead to similar heat trans-
4.3. Oxy-coal fer characteristics as air-coal firing. The oxy-coal cases found
higher temperatures and heat transfer with increased oxygen
The results for the change in heat transfer and temperature in concentration and could be similar to that of air if between 25%
the oxy25-coal and oxy30-coal cases are compared against the and 30% volumetric inlet oxygen was used.
air–coal case in Fig. 5. The heat transfer to the radiative sections, The results in the study suggest CFD can be used as a predictive
and overall total heat transfer is increased as the oxygen concen- tool to highlight the effects of different fuels and environments in
tration is changed from 25% to 30% by volume at the inlet. The in- an industrial boiler. However, to assess the results quantitatively a
crease in heat transfer can be associated to the rise in temperature number of updated physical models will be needed which include
786 S. Black et al. / Fuel 113 (2013) 780–786

radiation modelling, char gasification reactions as well as proper- [18] ANSYS FLUENT, version 14.0, 2012. <http://www.ansys.com>.
[19] Hjärtstam S, Johansson R, Andersson K, Johnsson F. Computational fluid
ties of biomass such as size, shape and thermal gradients.
dynamics modelling of oxy-fuel flames: the role of soot and gas radiation.
Energy Fuels 2012;26:2786–97.
Acknowledgements [20] Murthy JY, Mathur SR. A finite volume method for radiative heat transfer using
unstructured meshes. J Thermophys Heat Transfer 1998;12:313–21.
[21] Hottel HC, Sarofim AF. Radiative heat transfer. New York: McGraw-Hill; 1967.
Financial support from E-ON is gratefully acknowledged. We [22] Smith TF, Shen ZF, Friedman JN. Evaluation of coefficients for the weighted
would like to thank RWE npower for furnace geometry and mesh sum of gray gases model. ASME 1982;104:602–8.
as well as the data from the in-house model. [23] Johansson R, Andersson K, Leckner B, Thunman H. Models for gaseous radiative
heat transfer applied to oxy-fuel conditions in boilers. Int J Heat Mass Transfer
2010;53:220–30.
References [24] Johansson R, Leckner B, Andersson K, Johnsson F. Account for variations in the
H2O to CO2 molar ratio when modelling gaseous radiative heat transfer with
[1] IEA. Power generation from coal – ongoing developments and outlooks; 2011. the weighted-sum-of-grey-gases model. Combust Flame 2011;158:893–901.
[2] Kyoto Protocol. <http://unfccc.int/resource/docs/convkp/kpeng.pdf>, [accessed [25] Kangwanpongpan T, Silva RC, Krautz HJ. Prediction of oxy-coal combustion
29.09.12]. through an optimized weighted sum of gray gases model. Energy 2012;41:
[3] DECC. The electricity and gas (carbon emissions reduction) Order 2008, 244–51.
<http://www.legislation.gov.uk/uksi/2008/188/pdfs/uksi_20080188_en.pdf>, [26] Kangwanpongpan T, França FHR, Silva RC, Schneider PS, Krautz HJ. New
[accessed 29.09.12]. correlations for the weighted-sum-of-gray-gases model in oxy-fuel conditions
[4] DECC. 2012. UK Bioenergy Strategy. based on HITEMP 2010 database. Int J Heat Mass Transfer 2012. http://
[5] Lu H, Warren R, Peirce G, Ripa B, Baxter LL. Comprehensive study of biomass dx.doi.org/10.1016/j.ijheatmasstransfer.2012.07.032.
particle combustion. Energy Fuels 2008;22:2826–39. [27] Krishnamoorthy G. A new weighted-sum-of-gray-gases model for CO2–H2O
[6] DECC. The renewable obligation (RO) 2012. <http://www.decc.gov.uk/en/ gas mixtures. Int Commun Heat Mass Transfer 2010;37:1182–6.
content/cms/meeting_energy/renewable_ener/renew_obs/renew_obs.aspx>, [28] Porter R, Liu F, Pourkashanian M, Williams A, Smith D. Evaluation of solution
[accessed 29.09.12]. methods for radiative heat transfer in gaseous oxy-fuel combustion
[7] Toftegaard MB, Brix J, Jensen PA, Glarborg P, Jensen AD. Oxy-fuel combustion environments. J Quant Spectrosc Radiat Transfer 2010;111:2084–94.
of solid fuels. Prog Energy Combust Sci 2010;36:581–625. [29] Brookes SJ, Moss JB. Prediction of soot and thermal radiation in confined
[8] Smart JP, O’Nions PO, Riley GS. Radiation and convective heat transfer, and turbulent jet diffusion flames. Combust Flame 1999;116:486–503.
burnout in oxy-coal combustion. Fuel 2010;89:2468–76. [30] Brown AL. Modelling soot in pulverized coal flames [dissertation]. Brigham
[9] ZEP. EU Zero Emissions platform – biomass with CO2 capture and storage Young University; 1997.
(Bio-CCS) – The Way forward for Europe; 2012. [31] Sazhin SS. An approximation for the absorption coefficient of soot in radiating
[10] Backreedy RI, Jones JM, Ma L, Pourkashanian M, Williams A, Arenillas A, et al. gas. Fluent Europe Ltd.; 1994.
Prediction of unburned carbon and NOx in a tangentially fired power station [32] Magnussen BF, Hjertager BH. On mathematical models of turbulent
using single coals and blends. Fuel 2005;84:2196–203. combustion with special emphasis on soot formation and combustion. Symp
[11] Gharebaghi M, Irons RMA, Ma L, Pourkashanian M, Pranzitelli A. Large eddy Int Combust Proc 1976;16:719–29.
simulation of oxy-coal combustion in an industrial combustion test facility. Int [33] Ma L, Gharebaghi M, Porter R, Pourkashanian M, Jones JM, Williams A.
J Greenhouse Gas Control 2011;5:S100–10. Modelling methods for co-fired pulverised fuel furnaces. Fuel 2009;88:
[12] Toporov D, Bocian P, Heil P, Kellerman A, Stadler H, Tschunko S, et al. Detailed 2448–54.
investigation of a pulverized fuel swirl flame in CO2/O2 atmosphere. Combust [34] Ranz WE, Marshall WR. Vaporation from drops, Part I. Chem Eng Prog
Flame 2008;155:605–18. 1952;48(3):141–6.
[13] Ma L, Jones JM, Pourkashanian M, Williams A. Modelling the combustion of [35] Badzioch S, Hawksley PGW. Kinetics of thermal decomposition of pulverized
pulverized biomass in an industrial combustion test furnace. Fuel 2007;86: coal particles. Ind Eng Chem Process Des Develop 1970;9(4):521–30.
1959–65. [36] Backreedy RI, Fletcher LM, Ma L, Pourkashanian M, Williams A. Modelling
[14] Gubba SR, Ingham DB, Larson KJ, Ma L, Pourkashanian M, Tan HZ, et al. pulverised coal using a detailed coal combustion model. Combust Sci Technol
Numerical modelling of the co-firing of pulverised coal and straw in a 2006;178:763–87.
300 MWe tangentially fired boiler. Fuel Process Technol 2012;104:181–8. [37] Smith IW. The combustion rates of coal chars: a review. In: 19th Symp. (Int’l)
[15] Edge PJ, Heggs PJ, Pourkashanian M, Stephenson PL, Williams A. A reduced on combustion 1982. p. 1045–65.
order full plant model for oxyfuel combustion. Fuel 2012;101:234–43. [38] Haider A, Levenspiel O. Drag coefficient and terminal velocity of spherical and
[16] Edge PJ, Heggs PJ, Pourkashanian M, Stephenson PL. Integrated fluid dynamics nonspherical particles. Powder Technol 1989;58:63–70.
process modelling of a coal-fired power plant with carbon capture. Appl [39] Marakis JG, Papapavlou C, Kakaras E. A parametric study of radiative heat
Therm Eng 2012. http://dx.doi.org/10.1016/j.applthermaleng.2012.08.031. transfer in pulverised coal furnaces. Int J Heat Mass Transfer 2000;43:
[17] Mobsby JA. Thermal modelling of fossil fired boilers. In: 1st UK Heat transfer 2961–71.
conference. I. Mech E. and I. Chem E., Leeds; 1984.

You might also like