Jurnal 3 Udara

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Overview and Some Issues Related to Co-Firing

Biomass and Coal


Jianjun Dai,* Shahab Sokhansanj, John R. Grace, Xiaotao Bi, C. Jim Lim
and Staffan Melin
Department of Chemical and Biological Engineering, University of British Columbia, 2360 East Mall, Vancouver,
BC, Canada V6T 1Z3

Low heating values, variable chemical compositions, peculiar physical properties, high investment cost and insecurity of biomass feedstocks
supply limit the applications of biomass for energy and other processes. Co-firing biomass and coal has potential for the development of
biomass-to-energy capacity with significant economic, environmental, and social benefits. However, co-firing is not straightforward, and some
questions need to be addressed due to the differences in chemical compositions and physical properties of biomass and coal. This paper highlights
key issues related to co-firing, including reactor types, feeding, hydrodynamics, ash sintering, fouling, and corrosion, based on previous studies,
as well as calculations and analysis. Direct co-firing is the most common option for biomass and coal co-firing currently, mostly due to relatively
low investment needed to turn existing coal power plants into co-firing plants. For direct co-firing, the physical characteristics and chemical
compositions of the fuel entering the combustors or gasifiers are critical to an optimum operation. Any biomass mixed with coal needs to have
acceptable physical properties. More research is needed on co-firing biomass and coal, including work on: preparation, handling, storage, and
feeding of biomass feedstocks (e.g. drying, torrefaction, pelletization); co-firing mechanisms; hydrodynamic analysis of co-firing combustors
and gasifiers; boiler/gasifier capacity, slagging, fouling, corrosion, efficiency, reliability, fuel flexibility; lower emissions and gas cleaning; catalyst
poisoning; investment and operating costs.

Les applications de la biomasse pour les procédés énergétiques et d’autres procédés sont limités par les faibles pouvoirs calorifiques, des composi-
tions chimiques variables, des propriétés physiques particulières, des coûts d’investissement élevés et l’insécurité quant à l’approvisionnement en
biomasse comme matière première. La co-alimentation biomasse-charbon est potentiellement intéressante pour la transformation de la biomasse
en énergie, avec des avantages économiques, environnementaux et sociaux significatifs. Toutefois, la co-alimentation n’est pas simple et certaines
questions se posent du fait des différences dans les compositions chimiques et les propriétés physiques de la biomasse et du charbon. On souligne
dans cet article les principaux aspects entrant dans la co-alimentation, notamment le type de réacteurs, l’alimentation, l’hydrodynamique, le
frittage des cendres, l’encrassement et la corrosion, d’après des études antérieures ainsi que des calculs et analyses. La co-alimentation directe est
l’option la plus commune pour la co-alimentation biomasse-charbon, principalement en raison de l’investissement relativement faible requis pour
transformer les centrales électriques alimentées au charbon actuelles en centrales hybrides. Pour la co-alimentation directe, les caractéristiques
physiques et les compositions chimiques du fioul entrant dans les chambres de combustion et les gazéifieurs sont critiques pour un fonctionnement
optimal. Toute biomasse mélangée à du charbon doit posséder des propriétés physiques acceptables. De nouvelles recherches sont nécessaires sur
la co-alimentation biomasse-charbon, notamment sur: la préparation, la manutention, le stockage et l’alimentation des charges de biomasse (p.
ex., la torréfaction, la pelletization); les mécanismes de co-alimentation; l’analyse hydrodynamique des chambres de combustion et des gazéifieurs;
la capacité des chaudières/gazéifieurs, l’entartrage, l’encrassement, la corrosion, l’efficacité, la fiabilité, la flexibilité du fioul; les émissions faibles
et le lavage du gaz; l’empoisonnement du catalyseur; et enfin, les coûts d’investissement et de fonctionnement.

Keywords: co-firing, co-feeding, biomass, coal, fluidized bed

INTRODUCTION every aspect of the development of biomass-to-energy capacity,


especially in the presence of economic incentives to replace coal.

P
artial substitution of coal by biomass feedstocks or other
materials (e.g. waste) in coal-fired power plants requires co-
firing (Fernando, 2005). Co-firing also occurs when biomass ∗ Author to whom correspondence may be addressed.
feedstocks (or other materials) are partially replaced by coal E-mail address: djianjun@chml.ubc.ca
Can. J. Chem. Eng. 86:367–386, 2008
(Leckner, 2006). In most instances, co-firing of biomass in exist-
ing coal-fired boilers provides an attractive approach to nearly © 2008 Canadian Society for Chemical Engineering
DOI 10.1002/cjce.20052

| VOLUME 86, JUNE 2008 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 367 |
There are many successful co-firing systems with different
Table 1. Properties of different biomass fuels compared with coala
reactors (fixed bed, fluidized bed and entrained flow). Several
Fuel LHV Volatile Ash Ultimate analysis different types of biomass can be co-fired with coal, including
(daf) matter content % w/w (daf) wood, residues from forestry and related industries, agricul-
MJ/kg % w/w % w/w tural residues, and biomass in refined form, such as pellets.
(daf) (dry) C H O N S Energy crops are also potential candidates for co-firing (IEA, 2005;
Straw 18.2 81.3 6.6 49 6 44 0.8 0.2 Maciejewska et al., 2006). Clean wood waste, especially when
Wood 18.7 83 1.8 50.5 6.1 43 0.3 0.1 pelletized, is an excellent fuel with low ash and alkali concentra-
Bark 16.2 76 7 50.5 5.8 43.2 0.4 0.1 tions, and several commercial-scale co-firing demonstration tests
have been completed without deposition problems with up to
Rape oil 35.8 100 0 77 12 10.9 0.1 0
10% biomass on an energy basis (Tillman, 2000; Savolainen,
Peat 19 74.2 2.7 52.6 5.8 40.6 0.9 0.1 2003; Baxter, 2005). Danish tests with up to 20% on an energy
Bituminous 31.8 34.7 8.3 82.4 5.1 10.3 1.4 0.8 basis indicate that straw can be co-fired with coal without severe
coal deposition or corrosion problems (Frandsez, 2005). There is also
a
Chmielniak and Sciazko (2003). one large coal-burning CHP plant in Sweden which successfully
converted to 100% wood pellets (Melin, 2007).
In recent years, some groups (Sjostrom et al., 1994; Kurkela
Low heating values (LHV), varying chemical compositions et al., 1994; Madsen and Christensen, 1994; Reinoso et al., 1994;
(Table 1), peculiar physical properties (e.g. wide range of par- Brage et al., 1995; Chen et al., 1995; de Jong et al., 1998; Collot
ticle size, high moisture content (MC), irregular shapes, low et al., 1999; Sjostrum et al., 1999; Brown et al., 2000; Pan et al.,
bulk densities), as well as high investment costs and insecu- 2000; Xie et al., 2001; Chmielniak and Sciazko, 2003; McLen-
rity of feedstock supply, are major concerns when stand-alone don et al., 2004) have reported co-gasification of biomass and
biomass plants are built. Large biomass units (>300 MWe) may coal. Co-gasification brings environmental benefits (e.g. reduced
be economically impractical based on present economic crite- CO2 emissions, decreased sulphur and nitrogen oxide emissions),
ria (Fernando, 2005). Coal can mitigate the effects of variations while also reducing problems that occur in biomass operations
in biomass feedstock quality and buffer the system when there associated with the production of tar. Air-steam gasification facil-
is insufficient biomass feedstock (Rickets, 2002; Viewls, 2005), itates high conversion of solid feedstocks such as biomass and
whereas biomass brings environmental and social benefits to coal coal into gas (Hanaoka et al., 2005). However, there have been
plants. When co-firing occurs in large units with high thermal few reports on co-gasification of woody biomass and coal with air
efficiency, specific operation costs are likely to be lower than in and steam from the viewpoint of the supply of syngas for synthe-
small-scale systems (Rickets, 2002; Viewls, 2005), and the costs of sis of liquid fuels (de Jong and Hein, 1999; Pinto et al., 2003).
adapting existing coal power plants should be lower than building There have also been few reports on co-gasification where the
new dedicated biomass systems (Fernando, 2005). Recent reviews relative proportions of biomass and coal have varied over wide
of co-firing identified over 100 successful field demonstrations in ranges (Kumabe et al., 2007). Research is needed on pre-treating
16 countries, utilizing many types of biomass in combination with biomass to facilitate co-feeding and dosing to the gasifier. The
various types of coals and boilers (Baxter, 2005) (Table 2). There ash, slagging, fouling and corrosion behaviour of typical biomass
have also been extensive studies concerning the co-firing of coal minerals have to be assessed (Boerrigter et al., 2006).
and biomass for energy generation (Leckner and Karlsson, 1993; This paper highlights the technical difficulties related to
Armesto et al., 1997, 2003; Desroches-Ducarne et al., 1998; Hein co-firing of biomass and coal based on previous experience. Cal-
and Bemtgen, 1998; Dayton et al., 1999; Amand et al., 2001; Sami culation and analysis are also provided to deepen understanding
et al., 2001; Laursen and Grace, 2002; Ross et al., 2002; Skodras of co-firing issues and problems, especially for direct co-firing.
et al., 2002; Gayan et al., 2004; Hupa, 2005; Huang et al., 2006;
Zulfiqar et al., 2006; Nevalainen et al., 2007).
Three basic types of technological configurations can be COAL AND/OR BIOMASS
identified for biomass co-firing in power plants (Zuwala and COMBUSTION/GASIFICATION
Sciazko, 2005): direct co-firing (Figures 1 to 4), parallel co-firing
and indirect co-firing (Figure 5). Currently the most common TECHNOLOGY AND EQUIPMENT
option is direct co-firing, where biomass and coal are utilized
together in the same boiler, mainly due to relatively low capital
Coal/Biomass Combustion
cost required to convert an existing coal-fired power plant into a Information on biomass co-combustion with coal can be found in
co-firing operation. For direct co-firing of biomass, two methods several articles that have summarized the state-of-the-art in this
have been developed: (a) blending the biomass and coal in the field (Sami et al., 2002; Demirbas, 2005). In general, three types
fuel handling system, with the blended fuel then being fed; and of combustion systems can be identified (Table 3):
(b) separate fuel handling and separate burners for the biomass,
thereby avoiding impact on the conventional coal delivery (1) Packed bed combustion systems use grate-fired furnaces and
system (Brem, 2005). Parallel co-firing units (where biomass and underfeed stokers. Different types of grate furnaces (up to
coal are fed into separate boilers, jointly producing steam for 20 MWth ) are available: fixed, moving, travelling, rotating,
power generation) are also popular, especially in the pulp and and vibrating (van Loo and Koppejan, 2004). Underfeed stok-
paper industry. The indirect option is expensive and involves ers are used in small- and medium-scale systems up to a
a separate biomass gasifier. Hence this option is rarely adopted nominal boiler capacity of 6 MWth (van Loo and Koppejan,
(Maciejewska et al., 2006). However, some researchers regard 2004). Packed bed boilers are generally not good candidates
indirect co-firing as the most effective method of introducing for direct co-firing compared to fluidized beds, although they
large quantities of biomass with coal (BDC, 2007). can be a component of more advanced modes, for example for

| 368 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 86, JUNE 2008 |
Table 2. Examples of co-firing projects

System description Fuel type and heating value Blend parameters and Technical difficulties Reference
HHV (kJ/kg) feeding methods
Grate-fired boiler Coal/wood chips blends. 10–20% (energy basis) Difficult blend mixing; Sampson
(electricity or steam Wood: birch, aspen, spruce. wood; blend feeding, Stoker capacity problems et al. (1991)
generation) HHVcoal = 22 605; spreader stokers with
HHVwood = 17 742 travelling grates; fly ash
re-injection system;
35–41% moisture for
blend fuel
Spreader stoker-fired Coal and refuse-derived fuel Blend feeding and separate Brouwer
boiler (140–300 kW (softwood waste) feeding; for the pulverized et al. (1995)
(fuel)) and a pulverized coal facility, blend feeding,
coal boiler (38 kW and separate feeding with
(fuel)) biomass as a reburn fuel
after the recirculation zone
Multi-circulating Coal/straw/wood chips 18–49% biomass (mass Coal and wood injected at Hansen et al.
fluidized bed blends; heating values not basis); blend feeding and bottom, straw injected (1995)
combustor (MCFBC) available separate feeding with secondary air; no
(power generation, 20 steady output of gaseous
MWe) alkali metals
Pressurized FBC (1.6 Blend of straw with coal Blend feeding Co-firing reduced the CO, Andries et al.
MWth ) NOx , and SO2 (1997)
concentrations in the
freeboard
Fluidized bed Coal and wood, straw, and Blend feeding; emissions of Wood is the most Van Doorn
combustor municipal sewage sludge SO2 , CO, and NOx favourable co-firing fuel in et al. (1996)
decreased with increasing terms of ease of
wood/coal ratio combustion and reduced
emissions of NOx and SO2 .
For co-firing straw, HCl
concentration increased
with larger straw/coal
ratios; co-firing sewage
sludge with coal caused
agglomeration
CFBC combustion Federal coal, straw, and Jacquet et al.
systems (250 MWe sewage sludge (1994),
CFBC at Gardanne) Rajaram
(1999),
McIlveen-
Wright et al.
(2007)
CFB boiler (295 MWth ) Milled peat, wood fuels Blend feeding and separate Melting of ash did not Rauhalahti
(Rauhalahti Municipal (sawdust, bark, cutter chips, feeding occur. However, deposits (2005)
CHP Plant, Finland) forest chips) (20%, heat formed on superheaters
basis) and coal. MC (wt%, when the combustion of
wb): 45% (peat); 45–50% fresh forest chips started
(wood fuels); LHV (daf,
MJ/kg): 10 (peat); 7–9
(wood fuels)
BFB boiler (42 MWe) Wood waste (35%), coal Blend feeding; the fuel mix Bed temperatures are Tacoma
(Steam plant #2, (50%) and RDF (15%, heat is fed to the FBCs overbed, maintained at ∼840◦ C to (2005)
Tacoma, Washington) basis) while limestone is added minimize ash
directly to the beds for agglomeration and
SO2 absorption maximize sulphur capture
CFB boiler (132 MWth ) Coal (59%, energy basis), Direct co-firing, not Spring
recycled pulping chemicals available for blend or (2005)
(35%), bark and wood separate feeding
waste (5%), and oil (1%)
(Continued)

| VOLUME 86, JUNE 2008 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 369 |
Table 2. Examples of co-firing projects

System description Fuel type and heating value Blend parameters and Technical difficulties Reference
HHV (kJ/kg) feeding methods
CFBC boiler (35 MWe) Coal (60%, heat basis) and a Direct co-firing, not Slough
(Slough Heat and densified RDF material available for blend or (2005)
Power Ltd.) (40%). HHV (daf, MJ/kg): 18 separate feeding
(RDF)
CFBC boiler (55 MWth ) Wood chips, bark, sawdust Direct co-firing, not Stora (2005)
and rejects from cardboard available for blend or
production. MC(wt%, wb): separate feeding
10% (coal) and 46%
(biofuels). LHV (daf, MJ/kg):
24.5 (coal) and 8.2 (biofuels)
CFBC (88 MWth ) Coal and straw (50% energy Separate feeding. The Several short-duration Grena (2005)
(Elsam, Denmark) basis) shredded straw was fed failures occurred due to
pneumatically through air excessive cutter wear and
locks to the boiler injection compacted bales with wet
loop seals intrusions. Dry
comminuted and/or
pelletized biomass fuels
were used with separate
storage and supply systems
CFBC boiler (110 Coal, RDF, wood waste, Blend and separate Lenzing
MWth ) (Austria Energy sewage sludge, and a range feeding; fuels of lower bulk (2005)
and LLB Lurgi, Lenzing, of specific industrial wastes density that are RDF, wood
Austria) waste, and specific
industrial wastes are
injected pneumatically into
the combustion chamber.
Coal and sewage sludge
were fed into the return
leg from the seal pot
Front wall-fired, Bituminous coal and Direct co-firing, blend Mixing and handling Colbert
pulverized coal boilers sawdust (5% mass basis); feeding problems; moisture (2005)
(182 MWe) about 95% of the screened content variation (the
material was <1/8 and blends were mixed by the
67% was <1/16 in size natural path through the
transfer points and the
bunker and pulverizers)
Wall - fired pulverized Coal and sawdust (<12% Separate injection Boiler efficiency loss was Seward
coal boiler (32 MWe) heat basis) ∼0.5%; slight impact on (2005)
(Seward Station) unburnt carbon; CO
emissions always were <20
ppmv, indicating no
problem with combustion
completeness. Favourable
impact on SO2 , NO, and
CO2 emissions
Wall-fired boiler Coal/switchgrass blends. 15% co-firing (mass basis); No slagging, normal unit Aerts et al.
Burners (50 MW (fuel)) HHVcoal = 25 500; blend feeding; 12 wt% operation, NOx decreased (1997)
HHVswitchgrassl = 15 997 (wb) moisture in biomass by 20%; some traces of
partially burned
switchgrass in ash
Wall-fired boiler Coal/straw/cereal blends. 0–100% biomass firing Three different burner Siegel et al.
Burners (500 kW (fuel)) Heating values not available (heat basis); fuel with configurations studied (1996)
higher nitrogen content (burner efficiency and
should be injected in fuel optimization)
rich zone to reduce NOx ;
Optimum co-firing ratio
60%; blend and separate
feeding
(Continued)

| 370 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 86, JUNE 2008 |
Table 2. Examples of co-firing projects

System description Fuel type and heating value Blend parameters and Technical difficulties Reference
HHV (kJ/kg) feeding methods
Wall-fired dual fuel Coal/sawdust blends. Coal and sawdust fed 81–90% burnout; NOx Abbas et al.
burner (500 kW (fuel)) HHVcoal = 32 260; separately. Coal: 74% <90 reduced, optimum (1994)
HHVsawdust = 18 140 ␮m, sawdust: 75% <1.4 co-firing ratio: 30% (mass
mm basis) for maximum
burnout and minimum
NOx ; fuel injection mode
depends on reactivity and
N2 in biomass
Tangentially fired and 5% wood derived biofuel Blend feeding Due to pulverizer Sami et al.
wall-fired PC boilers (heat basis) was co-fired performance and fuel (2001)
(Kingston and Colbert particle size, 5% (heat
power plants) basis) co-firing was found
to be the limiting case
Laboratory-scale Pulverized coal, rice husk, Fuel mixture supplied by a VM and MC very Chao et al.
pulverized fuel and bamboo; variable speed screw important in affecting (2008)
combustion testing Coal size: 75–106; conveyor from storage bin; combustion time, particle,
facility Biomass: 100–300 ␮m, pulverized coal and and PAH emissions. Particle
MC: 8–16 wt% (wb); biomass were transported size did not significantly
HHVcoal = 27 463. by primary air. Biomass affect combustion
HHVrice husk = 16 054. blending ratio 0%, 20%, performance
HHVbamboo = 17 296 30%, 40%, 50%, and
100% on mass basis
Cyclone coal boiler Coal and crossties (<25 Blend feeding, wood Blending wood with coal Thomas
(175 MWe) wt%). Crossties size: <1 mm blended with coal by a with further reduction of (2005)
(manufactured by coal yard scraper and wood particle size worked
Babcock & Wilcox) dozer to measure and mix well; SO2 emissions
the materials in the coal decreased by 7%;
yard prior to pushing the particulate emissions
blend to the reclaim decreased by 12%; NOx
hoppers. Blend was then emissions increased 8%.
loaded on conveyor belt No significant handling
prior to crusher house problems during testing
Cyclone coal boiler Bituminous coal and paper Blend of coal and paper Biggest problem in Gannon
(165 MWe) (Gannon pellets (5 wt%) pellets was bunkered on a co-firing was pluggage in (2005)
Generating Station) 24-h basis during the 21 d. the conveyor feeder; there
No control over the blend was no impact on
once it was introduced to unburned carbon in the
the bunker flyash; problem
experienced for low and
high feed rates was
variability in the steaming
rate of the boiler due to
the variations in heat input
from the varying blend
going to the boiler
Cyclone-fired Coal/b-dRDF blends. 12% co-firing (mass basis), NOx reduction of 2–3%, Ohlsson
Combustor (440 MWe HHVcoal = 14 388; blend feeding; 19 wt% SO2 reduction 17%, (1994)
Power generation) HHVRDFl = 12 955 (wb) moisture in biomass particulate concentration
(heat basis) increased by
about 50%
Down-fired concentric Coal/manure blends 100 g/min blend feed rate; Crushing manure to same Frazzitta
swirl burner (35.4 kW HHVcoal = 26 535; 20% manure (mass basis) size as coal difficult; SOx et al. (1999)
(fuel)) HHVmanure = 8650 and NOx decreased with
blend combustion, easy
ignition with blend
Supercritical PF Coal and straw (20% on Separate burners Flue gas desulphurization Breihofer
coal-fired power station energy basis); coal and (FGD) et al. (1991),
(600 MWe Amer 9 sewage sludge (20% on Gramelt
power station) energy basis) (1994),
McIlveen-
Wright et al.
(2007)

| VOLUME 86, JUNE 2008 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 371 |
Koppejan, 2004). Particle segregation in bubbling fluidized
beds is primarily caused by differences in particle density
of the two species, so this can be a problem when coal and
biomass are present. The combustion temperature has to be
kept below about 900◦ C to prevent ash sintering, which could
cause de-fluidization (van Loo and Koppejan, 2004). Hence,
fluidized bed combustion is better suited for woody biomass
(ash melting point >1000◦ C) than for herbaceous biomaterial
(e.g. straw, ash melting point <700◦ C) (Maciejewska et al.,
2006). The potential significant drawback, especially in CFB,
is incomplete combustion of fuels in the bed. This occurs
because the unburned carbon, often appearing as soot, is
Figure 1. Simplified process layout of direct co-firing unit (Zuwala and
very light and may not be captured by cyclones (Maciejew-
Sciazko, 2005). ska et al., 2006). Among the three basic types of combustion
systems, fluidized bed combustion seems to have the highest
fuel flexibility with respect to MC, heating value and ash con-
tent, enabling the use of fuel mixes and increasing the range of
fuels which can be fed or co-fed in existing power plants (van
den Broek et al., 1996; Bhattacharya, 1998; Werther et al.,
2000).
(3) Pulverized fuel or dust combustion systems (Figure 7) are fed
pneumatically. Fuels include coal, sawdust, and fine shavings
(van Loo and Koppejan, 2004). Fuel quality in dust combus-
tion needs to be maintained, with a maximum fuel particle
size of 10–20 mm and MC of no more than 20 wt% (wb) (van
Loo and Koppejan, 2004).

Coal/Biomass Gasification
Gasification is an important process related to indirect co-firing
(Maciejewska et al., 2006). Gasifiers are used in various appli-
cations (Hotchkiss et al., 2002). There are two major types of
gasification: direct and indirect gasification.
Fixed bed, fluidized bed and entrained flow gasifiers consti-
Figure 2. Possible bottlenecks in direct co-firing: BM, biomass; SCR, tute the three basic types of gasifiers. Fixed bed gasifiers can
selective catalytic reduction; ESP, electrostatic precipitator; FGD, flue gas be subdivided into updraft and downdraft gasifiers. Both require
desulphurization (Kiel, 2005).
mechanically stable fuel particles of limited size (e.g. ≤10–30
mm) to facilitate passage of gas through the bed. Therefore the
parallel or in-direct co-firing (Maciejewska et al., 2006). Some biomass can be preferred as pellets or briquettes. The scaling-up
studies of packed bed direct co-firing of biomass and coal have of fixed bed gasifiers, tar generation and NOx emissions are major
also been conducted (Sampson et al., 1991; Brouwer et al., concerns affecting the relative merits of different configurations
1995). (Maciejewska et al., 2006).
(2) Fluidized bed combustion (FBC) (Figure 6) has been reported Two types of fluidized bed gasifiers can again be identified (Fig-
to be most efficient and suitable for converting agricul- ure 6): BFB and CFB. Both have favourable fuel flexibility, being
tural and wood residues into energy, as well as for co-firing able to treat fuels of different origins (Belgiorno et al., 2003). Ash
(Elanchezian and Antonio, 1993; van den Broek et al., 1996; sintering and bed agglomeration are of concern when biomass is
Philippek and Werther, 1997; Bhattacharya, 1998; Werther the fuel (Heinrich and Weirich, 2002). The efficient performance
et al., 2000; Tsai et al., 2002; Backreedy et al., 2005). There of fluidized bed gasifiers requires relatively small fuel particles
are two major types of FBC systems: bubbling fluidized beds to ensure good contact with bed material, as in fluidized bed
(BFB) and circulating fluidized beds (CFB). Hupa (2005) combustion.
reviewed interactions of various fuels in various FBCs and Entrained flow gasifiers (EFG) convert mixtures of fuel and oxy-
found that factors such as flue gas emissions, fouling, and gen into a syngas at high temperatures (significantly >1200◦ C,
bed-sintering seldom depend in a simple linear manner on even as high as 2000◦ C) in very short periods of time (a few sec-
the composition of the fuel mixture; instead, non-linear rela- onds) and at high pressures (∼50 bar). The typical oxidizing agent
tionships tend to be the norm (Wan et al., 2007). Hot, inert, is oxygen, in order to reduce the nitrogen content in the gas and
and granular material (usually silica sand and limestone resulting NOx emissions. The production of pure oxygen and the
or dolomite) provide thermal inertia and can stabilize the high pressure result in high costs (Heinrich and Weirich, 2002).
combustion process (also capturing sulphur), constituting In order to achieve reliable feeding and high conversion of the
90–98% of the total in-bed mixture by mass, the balance being feedstock, particles should be smaller than 1 mm, or liquid (e.g.
fuel particles (van Loo and Koppejan, 2004). pyrolysis oil) feedstock can be used (Maciejewska et al., 2006).
Given their good mixing, fluidized beds can accept vari- Producing sufficiently fine powders from biomass for entrained
ous fuels (e.g. wood and straw), but require close control of flow gasifiers is also expensive. Slurry feeding can reduce the
fuel particle size (BFB < 80 mm, CFB < 40 mm) (van Loo and overall cost of solid fuel feeding at high-pressures, but it cannot

| 372 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 86, JUNE 2008 |
Figure 3. Organization of fuel supply during co-firing in PC and FBC boilers (Leckner, 2006).

Figure 4. Schematic diagram of CIEMAT’s 300 kW CFB reactor (Gayan et al., 2004).

| VOLUME 86, JUNE 2008 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 373 |
ties with ash melting (Maciejewska et al., 2006). Emery Energy
hinted at using an entrained flow concept for biomass, but details
are not publicly available (Maciejewska et al., 2006). Supercritical
water gasification operates at temperatures of 500–700◦ C and pres-
sures of 200–400 bar. The process is suitable for wet feedstocks,
but, as it is a relatively new concept, experience is restricted to
laboratory and pilot scale experiments (Tuurna, 2003). Commer-
cialization of supercritical water gasification also faces challenges
in both economic and technical aspects, depending on specific
conditions.

MECHANISM AND PHENOMENA RELATED


TO CO-FIRING
Combustion of biomass differs from coal combustion due to differ-
ences in chemical compositions and physical properties (Figures 8
and 9). Chao et al. (2008) conducted thermogravimetric analysis
(TGA) and differential thermal analysis (DTA) of coal, rice husk,
and bamboo at a heating rate of 5◦ C/min in air. High volatile
matter (VM) contents and low activation energies of rice husk
Figure 5. Biomass-coal co-gasification in Ruien 5 (Kurkela, 2002).
and bamboo made the pyrolysis and subsequent volatile oxidation
start earlier than for coal. The higher VM content of biomass led to
be applied to biomass because of the low energy density and the two distinct stages of weight losses, with gas phase oxidation at
high moisture capacity of biomass (Maciejewska et al., 2006). the beginning and char oxidation in the second stage, whereas
There are several commercial pneumatic feeding designs avail- the latter dominated the entire process for coal. The TG/DTA
able for coal (GE/ChevTex, Conoco E-gas, Shell), but these do results indicated that a substantial fraction of the energy from
not work with more than 10–15% biomass on an energy basis in the biomass combustion came from VM reaction, whereas almost
a coal blend (Maciejewska et al., 2006). Entrained-flow gasifica- all of the energy for coal came from char oxidation (Table 4).
tion is a mature/commercial technology for petroleum residues The time scales of the VM gas-phase reactions are much less than
and coal. However, experience is (very) limited with biomass. that of the char oxidation reaction of the residual carbon in the
Not every biomass type is appropriate in any case due to difficul- solid phase. The latter depends on the diffusion of oxygen to the

Table 3. Comparison of different combustion technologiesa

Reactor Advantages Disadvantages


type
Packed bed Low investment costs for plants <20 MWth and low operating Mixtures of wood fuels can be used, but mixtures of fuels with
combustion costs (van Loo and Koppejan, 2004); can use almost any type of different combustion behaviour and ash melting points (e.g.
(grate wood (Veijonen et al., 2003); appropriate for biomass fuels with blends of wood with straw or grass) are not possible (van Loo
furnaces) high moisture content (10–60 wt% wb) (Tuurna, 2003; van Loo and Koppejan, 2004); increase of temperature may cause ash
and Koppejan, 2004). Suitable for fuels with high ash content melting and corrosion (Tuurna, 2003)
and varying particle sizes (with a limitation regarding the
amount of fine particles) (van Loo and Koppejan, 2004)
Fluidized Large fuel flexibility in calorific value, moisture content, and ash Despite the flexibility with regard to fuel specifications, it is not
bed content, enabling fuel diversification and increasing the scope of always possible to use the existing feeding system for biomass
combustion fuels in existing power plants; combustion temperature in bed is by premixing the fuels (the cheapest option). In cases where the
low, resulting in low NOx emissions (EC, 2000; van Loo and feeding characteristics of the co-fired fuels vary too much from
Koppejan, 2004); provides an option to directly inject limestone the primary fuel, a separate feeder needs to be installed;
to remove sulphur cost-effectively (instead of FGD equipment), slagging and fouling on boiler walls and tubes when burning
maximized combustion efficiency even with low-grade fuels; fuels with high alkali content; Bed agglomeration when burning
environmental performance of FBC installations is good, with fuels of high alkaline and/or aluminum content; Cl-corrosion on
low emissions of CO (<50 mg/Nm3 ), NOx (<70 mg/MJ, after heat transfer surfaces (e.g. superheater tubes) (EC, 2000); high
the boiler, eventually reduced to less than 10 mg/MJ when using investment costs, interesting only for plants >20 MWth for BFB
SCR) and high boiler efficiencies (about 90%) (EC, 2000); and >39 MWth for CFB, low flexibility in particle size, high dust
Fluidized bed technology can be converted from coal to load in the flue gas, loss of bed material with the ash (van Loo
biomass/coal co-combustion with relatively little investment and Koppejan, 2004); incomplete combustion of fuels and high
(Veijonen et al., 2003) unburned carbon content in the ash, especially in CFB
(Maciejewska et al., 2006)
Pulverized Increased efficiency due to low excess oxygen, high NOx Particle size of biomass is limited to <10–20 mm (van Loo and
fuel or dust reduction possible when appropriate burners used (van Loo and Koppejan, 2004). Low moisture content required (typically <15
combustion Koppejan, 2004) wt%, wb) for pneumatic feeding and decreased efficiency for
high-moisture fuels
a
Maciejewska et al. (2006).

| 374 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 86, JUNE 2008 |
Figure 6. Fluidized bed combustion systems: (a) BFB and CFB furnaces (van Loo and Koppejan, 2004); (b) BFB and CFB gasifiers (Belgiorno et al., 2003).

surface of the carbon residues, followed by surface chemistry. faster when they are non-spherical. Kinetics of the major steps in
This difference in time scale is crucial in explaining combustion biomass combustion are not fully understood (Backreedy et al.,
performance and pollutant emissions (Chao et al., 2008) (Fig- 2005). Pyrolysis, ignition, and combustion of coal and biomass
ure 9). Hence, the mechanisms of biomass and coal combustion particles (Sami et al., 2001; Backreedy et al., 2005) are compared
differ somewhat, although biomass follows the same sequence of in Table 4.
pyrolysis, devolatilization and combustion as for combustion of Combustion modelling for coal/biomass blends is complex due
low-rank coal (Sami et al., 2001). Biomass can burn more inten- to gas and two particulate phases, as well as chemical reactions.
sively and may give rise to higher local peak temperatures due Two chemically different fuels are involved, with biomass much
to its higher reactivity than coal. The combustion rate of biomass more reactive and having higher VM and MC than coal. Most
char is slightly higher because of a more disordered carbon struc- reactor models contain sub-models for fluid-mechanics, particle
ture (Backreedy et al., 2005), while biomass char burning rates are dispersion, fuel devolatilization, gaseous combustion, heteroge-
comparable to burning rates of high-VM bituminous coal chars neous char reaction and pollutant formation. Combustion models
(Sami et al., 2001). In fact, the reactions of the major components based on coal need to be modified to account for the effects of
of wood, hemicellulose, cellulose, and lignin, are interconnected biomass co-firing on the overall combustion behaviour (Gayan
at high temperatures, with the wood reacting at one composite et al., 2004). The suitability of the sub-models for biomass com-
rate. Wood containing a high proportion of lignin, for example in bustion is a key factor in selecting an appropriate code, including
knots, reacts more slowly (Backreedy et al., 2005). Lignin does CFD models (Backreedy et al., 2005). Similarity between the coal
not burn completely, especially if >0.5 mm in diameter. and biomass sub-models can be assumed, despite differences in
Biomass particles are large and physically complex, influencing mechanisms and kinetics (Backreedy et al., 2005). Sami et al.
heat and mass transfer. Particle shape and size affect char burnout (2001) revised the modelling of co-firing based on models for pul-
because biomass does not melt, and irregular shapes are main- verized or swirls burners. Saastamoinen et al. (2005) presented
tained during combustion. Larger particles of a given mass burn a burning regimes model covering the combustion of coal, wood

| VOLUME 86, JUNE 2008 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 375 |
Figure 7. Pulverized fuel and dust combustion systems: (a) pulverized fuel technology and muffle dust furnace in combination with a water-steam tube
boiler (van Loo and Koppejan, 2004); (b) example of entrained flow gasifier (Heinrich and Weirich, 2002).

chips and their mixtures. The model assumes that a burning fuel 2004; Goh, 2005; Lu et al., 2007). This is attributed to the larger
particle initially loses mass due to drying and devolatilization, particle sizes and higher MC of the biomass. Premixing biomass
causing its average density to decrease, while its diameter remains and coal can enhance the combustion of the two fuels, whereas
approximately constant. In practice, wood particles may shrink poorly mixed biomass and coal tend to burn independently at dif-
in size, whereas some coals swell (Nevalainen et al., 2007). CFB ferent rates (Lu et al., 2007). Test results have suggested that, due
models for burning coal (Adanez et al., 1995) and biomass (de to the varying physical and chemical properties of the biomass
Diego et al., 2002; Adanez et al., 2003) and blended biomass and fuels, their additions have a significant impact on the character-
coal (Gayan et al., 2004) have been developed in the past decade istics of the flame, particularly the flame front and brightness.
with a focus on predicting combustion efficiency, fouling, and However, flame stability has been found to be little affected by
emission of pollutants for different fuels and their mixtures in the amount of biomass added in all cases studied, provided that
commercial-scale fluidized bed combustors. the addition is less than 20% by mass (Lu et al., 2007).
When a small amount of biomass is added to a coal flame, the McIlveen-Wright et al. (2007) analyzed 25 biomass processes
reaction environment is primarily determined by the combustion in CFBC systems based on actual power plants. It was shown that
of the coal rather than by the biomass kinetics. Biomass additions CFBC power plants of different sizes could operate effectively and
have led to a slight delay in the ignition of the blended fuels, efficiently with a range of biomass types and loadings in co-firing
although biomass has a lower ignition temperature (Demirbas, applications, with lower net CO2 emissions (compared to cases

| 376 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 86, JUNE 2008 |
Figure 8. Comparison of coal and biomass elemental composition
(van Loo, 2002).

where coal is the sole fuel), and improved compliance with NOx
and SOx emission limits.

Figure 9. Schematic of co-firing phenomena (Biagini and Tognotti,


HYDRODYNAMIC CHARACTERISTICS OF CFBC 2002).
CFB hydrodynamic characteristics have been analyzed and mod-
elled (e.g. Johnsson et al., 1992; Johnsson and Leckner, 1995;
Pallares and Johnsson, 2000; Gayan et al., 2004). The riser was regimes for different particles and to optimize their physical prop-
divided into three zones: a bottom zone, characterized by a dense erties (e.g. size, MC, density) for blended fuels. Mixtures of
bed, similar to a bubbling bed; a splash zone with predominantly biomass and coal can be well fluidized only when the biomass
homogeneous particle clustering flow; and a transport zone with constitutes less than 50% by volume. Minimum fluidization veloc-
a core-annulus structure. In the splash and transport zones, the ity of particle blends can be calculated by the Ergun equation with
vertical distribution of solids was characterized by an exponential reasonable accuracy (Bi, 2005). A sample calculation is given in
decay model, with the solid concentration assumed to be the sum Table 5 for different fuels, including coal and various biomass
of contributions from a cluster phase and a dispersed phase. The feedstocks. It is commonly assumed that the density and shape of
hydrodynamic model can predict mean voidage, annulus and core biomass particles do not change and that no fragmentation occurs
voidages, core radius, upward solids flow in the core, downward (Nevalainen et al., 2007). Non-spherical shapes may affect the
solids flow in the annulus and external circulation solid flux, all drag coefficients and require the use of shape factors (Clift et al.,
as functions of height (Gayan et al., 2004). For more information 1978; Backreedy et al., 2005). Biomass particles differ in density,
about fluidized bed hydrodynamics and reactor modelling, see size and shape from coal particles, and this can cause different
Grace et al. (2003). trajectories and reaction locations in the furnace.
Hydrodynamic analysis is also important for both blended feed- Ganser (1993) introduced two shape factors K1 and K2 applica-
ing and separate feeding in order to estimate the hydrodynamic ble in the Stokes and Newton’s regimes for the estimation of the

Table 4. Comparison of coal and biomass combustiona

Items Biomass Coal


Particle size Relatively large and wide range Relatively small and narrow range
Particle size distribution Wide Narrow
Particle shape Irregular Relatively regular
Moisture content High Low
Reactivity Higher Lower
Volatile matter content Higher Lower
Devolatilization Lower temperature Higher temperature
Pyrolysisb Earlier Later
Specific heating value of volatiles (MJ/kg) Lower Higher
Fractional heat contribution by volatiles ∼50–70% (Chao et al., 2008) ∼30%
Char More oxygen Less oxygen
Combustion rate of char Slightly higher Slightly lower
Ash More alkaline, chlorine Less alkaline
a
Sami et al. (2002) and Demirbas (2005).
b
Temperature at which pyrolysis occurs depends on fuel type and heating rate.

| VOLUME 86, JUNE 2008 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 377 |
Table 5. Sample calculations for coal and some biomass feedstocks
Fuel types American Federal Coala Wood pelletsa,b Ground wood pellets-1a,b Wheat strawa,c Switchgrassa
Proximate analysis
Water (wt%, ar) 6.3 5.43 5.43 10.6 7.17
Ash (wt%, ar) 6.22 2.55 2.55 4.07 4.62
Volatiles (wt%, ar) 87.48 79.16 (daf) 79.16 (daf) 85.33 88.21
Fixed carbon (wt% ar) 47.91(daf) 47.91(daf)
Ultimate analysis (wt%, daf)
Carbon 84 51 51 48.84 43.58
Hydrogen 5.7 6 6 7.08 5.83
Oxygen 6.06 42.9 42.9 42.36 50.19
Nitrogen 1.5 0.05 0.05 1.28 0.4
Sulphur 2.6 0.05 0.05 0.16 0
Chlorine 0.14 0 0 0.28 0
HHV (MJ/kg daf) 35.64 18.71 18.71 19.9 21.61
LHV (MJ/kg daf) 18.2
Physical properties
Average size (mm)d 0.5 9.8 4 0.91 0.91
Shapee Spheroid Cylinder (6.5 × 15) Cuboid (3.2 × 3.2 × 3.2) Disk (1 × 0.5) Disk (1 × 0.5)
Sphericity 1 0.82 0.81 0.83 0.83
Particle density (kg/m3 ) 1500 1200 1200 500 500
Bulk density (kg/m3 ) 810 630 485 160 160
Voidage 0.46 0.48 0.60 0.68 0.68
Calculationsf
(1) Sole coal or biomass
Power (HHV, MW) 60 60 60 60 60
Fuel flow rate (kg/s, daf) 1.68 3.21 3.21 3.02 2.78
Fuel flow rate (kg/s, wb) 1.92 3.48 3.48 3.53 3.15
Oxygen molar flow rate (mol/s) 130 150 150 135 111
Air mass flow rate (kg/s) 17.8 20.59 20.59 18.54 15.23
Air volumetric flow rate (m3 /s) 43.88 50.75 50.75 45.69 37.54
Bed radius (or side dimension) (m) 0.9 0.9 0.9 0.9 0.9
Superficial gas velocity in reactor (m/s) 17.24 19.94 19.94 19.59 16.09
Min. fluidization velocity-1, Umf -1 (m/s)g 0.12 3.4 2.76 0.37 0.37
Min. fluidization velocity-2, Umf -2 (m/s)h 0.09 2.98 2.32 0.27 0.27
Drag coefficient, Cd i 2.93 1.13 1.08 3.62 3.62
Terminal velocity, Ut , (m/s)j 2.80 17.86 11.71 1.96 1.96
(2) 50 wt% biomass in coal/biomass blend
Average blend bulk density (kg/m3 )k 709 607 267 267
Average blend particle density (kg/m3 )l 1333 1333 750 750
Average voidage of blend fuelm 0.47 0.54 0.64 0.64
Average sphericityn 0.9 0.89 0.87 0.87
Average particle size (mm)o 5.67 2.44 0.81 0.81
Average minimum fluidization velocity (m/s)p 2.41 1.52 0.29 0.29
a
Huang et al. (2006).
b
Lu et al. (2007).
c
McIlveen-Wright et al. (2007).
d
Equivalent volume diameter.
e
Assumed particle shape.
f
Based on 830◦ C, 1
atm and 10% excess air, bed materials (e.g. silica sand and dolomite) are not considered.
g
Based on Remf = C12 + C2 Ar − C1 with consideration of sphericity and voidage of particles.
h
Based on modified Ergun equation.
i
Based on modified drag coefficient for irregular single particle without wall effects.
j
Based on ut = 4gdv ( p − f )/(3 f CD ) for single particle without wall effects.
b,ave = (mcoal + mbiomass )/(mcoal / b,coal + mbiomass / b,biomass ).
k

p,ave = (mcoal + mbiomass )/(mcoal / p,coal + mbiomass / p,biomass ).


l
m
␧ave = (1 − b,ave / p,ave ).
n
Calculated and averaged according to particle volume fraction of each type of particle in the blend.
o
Calculated and averaged according to particle volume fraction of each type of particle in the blend.
p
Based on modified Ergun equation.

| 378 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 86, JUNE 2008 |
Figure 10. Schematic of co-firing aspects (Biagini and Tognotti, 2002).

drag coefficient, Cd : Figure 11. Synergetic co-firing effects (Leckner, 2006).

Cd 24 0.4305 where p is the particle density, and f is the fluid density.


= + The Ergun equation (Ergun, 1952):
K2 Re K1 K2 [1 + 0.1118(Re K1 K2 )0.6567 ] 1 + 3305/Re K1 K2
(1) P ␮f Umf (1− ␧mf )2 f Umf (1− ␧mf )
2
=A +B =( p − f) )(1− ␧mf )g
Hb 2 dv ␧mf
2 3
dv ␧3mf
where Re is the particle Reynolds number based on the volume
 −1 (3)
equivalent diameter, dv ; K1 = (dn /3dv ) + (2/3) −0.5 ; K2 =
)0.5743
101.8148(−log ; dn is the projected-area-equivalent diameter;
can be used to estimate the minimum fluidization velocity (Nemec
and is the sphericity.
and Levec, 2005; Keyser et al., 2006). Here P is the pressure drop,
The terminal velocity, ut , can then be estimated from:
A is the Blake–Kozeny–Carman constant, B is the Burke–Plummer
 constant, Hb is the bed height, Umf is the minimum fluidization
4gdv ( p − f) velocity, ␮f is the dynamic viscosity of fluid, and ␧mf is the void
ut = (2)
3 f Cd fraction at minimum fluidization. For spherical particles, A = 150

Table 6. Physical and chemical characteristics of biomass feedstocks and their effects on co-firinga

Properties Effects

Physical properties Moisture content


Storage durability
Dry-matter losses
Low LHV
Self ignition
Bulk density Fuel logistics (storage, transport, handling) costs; storage and feeding problems (e.g. bridging and
stoppage)
Ash content Dust, particulate emissions, ash utilization problems, disposal costs
Particle size, size determines fuel feeding system, Determines combustion technology, drying properties, dust
distribution, and shape formation, operational safety during fuel conveying
Chemical composition Carbon (C) HHV (position)
Hydrogen (H) HHV (positive)
Oxygen (O) HHV (negative)
Chlorine (Cl) Corrosion
Nitrogen (N) NOx , N2 O, HCN emissions
Sulphur (S) SOx emission, corrosion
Fluorine (F) HF emissions, corrosions
Potassium (K) Corrosion (heat exchangers, superheaters), lowering of ash melting temperatures, aerosol
formation, ash utilization
Sodium (Na) Corrosion (heat exchangers, superheaters), lowering ash melting temperature, aerosol formation
Magnesium (Mg) Increase of ash melting temperature, ash utilization
Calcium (Ca) Increase of ash melting temperature, ash utilization
Phosphorus (P) Increase of ash melting point, ash utilization
Heavy metals Emissions of pollutants, ash utilization and disposal issues, aerosol formation
a
EBA (2000), van Loo and Koppejan (2004) and Maciejewska et al. (2006).

| VOLUME 86, JUNE 2008 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 379 |
and B = 1.75. For cylindrical particles (including disks), Nemec reduce corrosion in boilers burning biomass, even at tempera-
and Levec (2005) recommend: tures as high as 550◦ C. Ammonium sulphate may be injected into
the flue gas after combustion to convert gaseous potassium chlo-
150 1.75 ride into potassium sulphate (a much less corrosive compound),
A= , B=
3/2 4/3 resulting in reductions in corrosion and deposition rates by 50%.
The corrosion issues in co-firing or biomass systems can also be
Since cylinders are reasonably similar in shape to cuboids, these addressed by pretreatment of biomass by leaching with water,
relations are also give reasonable predictions for cuboidal parti- thereby reducing the content of alkalis, sulphur, and chlorine in
cles (e.g. ground wood pellets in Table 5). Calculated minimum the feedstock (Jenkins et al., 1996; Jensen et al., 2001; Davids-
fluidization velocities calculations for different types of particles, son et al., 2002). More information on biomass pre-treatment
including 50:50 blends of biomass and coal, are shown in Table 5. through washing (both biomass washing and char washing) and
Biomass particles are not easy to fluidize due to their large size and its benefits for co-firing systems was provided by Maciejewska
irregular shapes. Wood pellets can be better suited to fluidization et al. (2006). Biomass-related deposit formation and corrosion
after grinding. are linked (EUBION, 2003); erosion and corrosion also interact
(Stack and Jana, 2004). Therefore it is difficult to address these
issues separately.
TECHNICAL CONSTRAINTS RELATED TO Baxter (1993) concluded that the ash deposition rate in
CO-FIRING COAL AND BIOMASS biomass combustion peaks at early times and then decreases
Constraints related to co-firing can include fuel preparation, han- monotonically. The tenacity and strength of biomass combus-
dling, storage, milling and feeding problems (e.g. high MC, tion deposits tend to be higher than for deposits from coal
low bulk density, hydrophilic, non-friable character, biodegrad- combustion, with smooth deposit surfaces and low poros-
ability), different combustion behaviour, possible decreases in ity. This means that the deposits from biomass combustion
overall efficiency (e.g. relatively low calorific value, high MC), tend to be difficult to remove, requiring additional cleaning
deposit formation (slagging and fouling), agglomeration, cor- effort.
rosion and/or erosion (e.g. low ash melting point, chemical
composition with potentially high alkaline metals and chlorine
content) and ash utilization (e.g. high alkaline metals and chlo-
Pollutant Emissions
rine content). Most of these issues are related to fuel properties Blending coal and biomass can lead to reductions in pollutant
(Figures 8 to 11, and Tables 1 and 6). With proper combi- emissions (Leckner and Karlsson, 1993; Nordin, 1995; Armesto
nations of these elements, a number of power plants practice et al., 1997, 2003; Gulyurtu et al., 1997; Desroches-Ducarne et al.,
co-firing without major problems (Figure 11) (Tillman, 2000; Aho 1998; Hein and Bemtgen, 1998; Dayton et al., 1999; Werther
and Ferrer, 2005; Aho et al., 2005; Baxter, 2005; Ferrer et al., et al., 2000; Amand et al., 2001; Laursen and Grace, 2002; Ross
2005). et al., 2002), with the levels of pollutants decreasing as the pro-
portion of biomass increases. Dayton et al. (1999) investigated the
Ash, Slagging, Fouling, and Corrosion Problems interactions between co-fed biomass and coal during combustion.
In indirect co-firing, as well as parallel co-firing, the ash pro- The results revealed the synergetic effects of co-firing for HCl, KCl,
duced in the process is kept separate. In direct co-firing, coal and and NaCl. The amounts of NOx and SO2 detected suggested that
biomass ash are mixed together. Mixed ash is not easy to utilize any decrease resulted from dilution of N and S in the fuel blend,
in the same applications as coal ash (e.g. in the construction although alkaline ash from biomass may capture some SO2 gen-
industry). The degree of difficulty depends on the quality and erated during combustion. The fuel nitrogen content of biomass
percentage of biomass in the fuel blend, type of combustion is mainly converted to ammonia during combustion, contributing
and/or gasification, co-firing configuration, and coal properties. to reduced NOx for co-firing (Gayan et al., 2004). Hydrocarbons
Therefore, when analyzing the environmental impacts of co- from biomass can also react with NOx , producing molecular N2 .
firing, the options for ash utilization must be assessed, especially Hence, biomass has the potential to be an effective additional
for high biomass/coal ratios. fuel when coal is the primary fuel. In addition, NH3 found in
The major mechanisms and rates of ash deposition are related the biomass (e.g. animal wastes) or formed during combustion of
to the inorganic material (e.g. chlorine, sulphur, aluminium, biomass may contribute to the catalytic reduction of NOx (Sami
and alkaline) in the fuel and to the combustion conditions et al., 2001).
(EUBION, 2003). Deposits may be caused by light sintering, or Circulating fluidized bed (CFB) technology has been used to
complete fusion due to the lower ash melting-point of biomass burn coal and biomass because of its ability to handle low-quality,
ash. The degree of fouling and slagging varies throughout the high-sulphur fuels. Leckner and Karlsson (1993) measured exper-
boiler, depending on local gas and tube temperatures, tube ori- imental emissions of NO, N2 O, SO2 , and CO from combustion of
entation, gas velocity and fuel composition (EC, 2000; Jensen mixtures of bituminous coal and wood in a CFB. They concluded
et al., 2001; EUBION, 2003; Benetto et al., 2004). Deposits tend that emissions from the combustion of mixtures are related to
to cause deterioration in the heat transfer to tubes, reducing the mass fractions of the fuels and to their properties. Nordin
combustion efficiency (EUBION, 2003). Although slagging and (1995) optimized sulphur retention during co-combustion of coal
fouling may occur quickly, corrosion may progress slowly over and biomass fuels in a fluidized bed using statistical experi-
a long period, with or without associated slagging or fouling mental designs for operating variables. When Van Doorn et al.
(EUBION, 2003). Chlorine-rich deposits (NaCl and KCl) induce (1996) and Sami et al. (2001) fired blended coal, wood, straw,
hot corrosion of heat transfer surfaces, but high-risk chlorine and municipal sewage sludge into a fluidized bed combustor, they
compounds can react with sulphur and aluminum silicate com- found wood to be the most favourable co-firing fuel in terms of
pounds, releasing HCl, which is less harmful (EUBION, 2003). ease of combustion and reduced emissions of NOx and SO2 . No
CORIDS (2005) reported that suitable materials and additives can particle agglomeration was observed. Emissions of SO2 , CO and

| 380 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 86, JUNE 2008 |
NOx decreased with increasing wood-to-coal ratio. Similar effects Handling and Feeding of Biomass
were observed when co-firing straw, but the HCl concentration Different chemical compositions and peculiar physical properties
increased with larger straw-to-coal ratios due to its relatively high (e.g. low bulk density and high MC) can significantly influence
chlorine content. Huang et al. (2006) reported that co-firing wood the design and operation of handling and feeding systems. A
chips resulted in lower NOx emissions, whereas co-firing straw or separate feeding system is frequently provided for the biomass
sewage sludge slightly increased NOx in pressurized fluidized bed component of the fuel. Drying, size reduction, storage, transporta-
combustors (PFBC). NOx may decrease or increase when co-firing tion, feeding, and handling of biomass fuels present problems in
coal and straw (or other biomass feedstocks) depending on the achieving stable conditions. Large particle size, high MC, irregular
blending ratio, fuel properties, and combustion conditions. shapes and low bulk density tend to promote feed rate irregular-
Hein and Bemtgen (1998) studied the co-combustion of dif- ities. Co-feeding of blended fuels, for example coal and biomass,
ferent biomass materials with coal in a range of pilot plants presents more problems than separate feeding.
and large-scale power stations. They found that CFBs could be Although direct co-firing affects combustion behaviour and ash
designed to handle the size of wood chips and that biomass addi- handling, if the proportion of biomass in the coal is small (e.g.
tion suppressed SO2 emissions significantly for all FB facilities. <10% on an energy basis), the effect of biomass addition has
Higher excess air for co-combustion of biomass and coal rela- been found to be insignificant (Table 2) for packed bed, fluidized
tive to pure coal combustion in a CFB was recommended by bed, and entrained flow reactors, with significant economic and
Werther et al. (2000) and Amand et al. (2001). Armesto et al. environmental benefits (Sami et al., 2001). Feeding of blend fuels
(2003) combusted a blend of coal and olive-oil-industry residue is not straightforward. For example, pre-mixing of some biomass
in a bubbling fluidized bed pilot plant to study the effect of oper- feedstocks (e.g. straw) and coal was not feasible due to segre-
ating conditions on the emissions and combustion efficiencies. gation of the two materials. In addition, slightly higher MC of
They found that the share of residue in the mixture (10–25% on the biomass can cause feeding of blend fuels to be unstable or
a mass basis) did not affect the combustion efficiency, although to fluctuate (Dai, 2007). Separate feeding mitigates the feeding
there was a significant influence on SO2 emissions due to the cal- and ash problems for co-firing of biomass and coal at the expense
cium and potassium content of the biomass. Generally the flue of higher investment costs. In many co-firing plants, biofuels are
gas passes to an electrostatic precipitator or bag filter to have pre-mixed with coal (or other materials) before feeding into the
particulate matter removed. Sulphur can be removed using flue boiler (Granada et al., 2006). If the limestone is fed with the coal
gas desulphurization, whereas oxides of nitrogen can be con- for capture of sulphur, the limestone may also be pre-mixed with
trolled by modifications to the burners. Clean-up systems for the coal and biomass.
NOx such as selective catalytic reduction (SCR) and selective Various measures can be applied to avoid or reduce prob-
non-catalytic reduction (SNCR) can also be adopted. Each of lems in biomass or blend feeding. Densified biomass (pellets
these technologies can be used in co-firing systems with little or and briquettes) is one option. Pellets are appropriate for coal-
no modification. fired plants (Bergman et al., 2005; Maciejewska et al., 2006)
Large portions of Cl-rich biomass (meat and bone meal (MBM) because modification of biomass properties addresses the source
and refuse-derived fuel (RDF)) have been co-fired with selected of the problems, rather than their consequences. The high costs
coals without operational problems (Aho et al., 2008). As men- of pelletization can be justified by better operability of the
tioned above, the sulphur and aluminosilicates present in coal can fuel (handling, transportation, storage, and feeding), resulting
capture alkalis from alkali chlorides and release HCl, preventing in improved boiler and combustion performance. The impor-
Cl from condensing on superheaters as alkali chlorides. HCl does tance of pre-treatment is likely to increase with the tendency
not bring chlorine to the deposits. The key reactions are sulpha- to utilize low-quality biomass. Another interesting option, espe-
tion and alkali aluminum silicate formation (Aho et al., 2008). cially for herbaceous biomass (currently rarely considered for
Increased kaolinite (Al2 Si2 O5 (OH)4 ) and decreased alkali contents co-firing) might be a pre-treatment process combining torrefac-
in the coals improved alkali capture, allowing larger contents of tion and pelletization to allow co-utilization of high ratios of
Cl-rich biomass in co-firing without Cl deposition (Ferrer et al., low-quality biomass with coal in existing coal systems without
2005). Ca/S ratios >3 can provide effective SO2 capture. High major modifications. This pre-treatment option has not yet been
S/Cl ratios (>4) can facilitate sulphation (Salmenoja et al., 1996; commercialized, so that environmental impacts, large-scale per-
Fernandez, 1998). High Al/Cl ratios can lead to effective alkali formance and economics are currently unknown (Maciejewska
aluminum silicate formation if a significant portion of the alu- et al., 2006).
minum is present as active kaolinite (Fernandez and Curt, 2004; Bulk material handling and feeding are widely described in the
Aho and Ferrer, 2005; Ferrer and Aho, 2005). (Na + K)/Cl > 1 literature. Various options such as hoppers or lock hoppers, screw
indicates an excess of alkalis for formation of alkali chlorides feeders (Dai, 2007), conveyor belts (Abbas et al., 1994), and pneu-
(Aho et al., 2008). Chlorine concentration in some fuels, such as matic feeding systems (Tmej and Haselbacher, 2000; Sami et al.,
straw, can be reduced by fuel pre-treatment with water. This can 2001; Dai, 2007), have proved to be suitable for different kinds
also have a beneficial effect on ash fusion temperatures (Jenkins of biomass. The feeding system should be designed to handle the
et al., 1998). specific fuel flow properties. The most common feeding system
Dioxins that might be expected appear to be destroyed within for pellet stoves is a screw auger driven by a slow-moving high-
the furnace when temperatures are >1000◦ C, but may be torque motor fed from a hopper (Granada et al., 2006). Screw
formed in the cooling region downstream by de novo syn- feeders may cause fuel flow fluctuations and segregation of pellet
thesis. The fate of certain trace elements in biomass and and forest residues when fed by the same screw. Because of seg-
wastes has not been fully established. Most heavy metals regation during storage and different feeding behaviour of pellet
appear to be trapped within the ash. Although further tests are and forest residue, different chambers are needed in a hopper to
needed in this area, co-firing appears to be advantageous in obtain steady flow and to control mixing (Granada et al., 2006).
many respects. Pelletized biomass (dried during processing to a low MC) can be

| VOLUME 86, JUNE 2008 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 381 |
successfully processed by coal mills, but this route is expensive dv volume-equivalent diameter (m)
(Segrest et al., 1997). More research is required for blend feeding. Hb bed height (m)
 −1
K1 = (dn /3dv ) + (2/3) −0.5
0.5743
CONCLUDING REMARKS K2 = 101.8148(−log )
Re particle Reynolds number
(1) Both biomass and coal can benefit from co-firing. Co-firing
Umf minimum fluidization velocity (m/s)
in coal plants can strongly increase biomass use and reduce
ut terminal settling velocity (m/s)
the emissions of greenhouse gases and other pollutants at low
capital and operational cost (compared to dedicated biomass
plants). Greek Symbols
(2) Direct co-firing is the most popular current option for biomass P pressure drop (Pa)
and coal co-firing, with modest investment cost to turn exist- ␧mf void fraction at minimum fluidization
ing coal power plants into co-firing plants. Direct co-firing ␮f dynamic viscosity of fluid (Pa s)
of biomass and coal takes advantage of the high efficiencies density (kg/m3 )
obtainable in large coal-fired power plants and improves com- f fluid density (kg/m3 )
bustion due to the higher volatile content of the biomass. The p particle density (kg/m3 )
cost of parallel co-firing is significantly higher than the direct sphericity
option, but may assist in optimizing the combustion process
and in utilizing difficult fuels with high alkali and chlorine
contents. Indirect co-firing can keep the biomass ashes sepa- Abbreviations
rate from the coal ashes, while allowing very high co-firing ar as received
ratios. However, indirect co-firing requires relatively high unit b-dRDF binder-enhanced densified refuse-derived fuel
investment costs. BFBC bubbling fluidized bed combustion
(3) Although more research is needed, there is already a wealth CFBC circulating fluidized bed combustion
of practical experience for different conditions. For direct co- daf dry ash-free
firing, the physical characteristics and chemical composition db dry basis
of the fuel entering the combustors or gasifiers are critical to DTA differential thermal analysis
their operation. Any biomass mixed with coal needs to have EFG entrained flow gasification
acceptable physical properties. For low co-firing ratios (<10% FBC fluidized bed combustion
thermal), there appears to be no irresolvable issues. Higher HC hydrocarbons
capital costs of advanced co-firing configurations may be jus- HHV higher heating value
tifiable due to better operability and flexibility of the system. LHV lower heating value
For higher co-firing ratios, additional research is needed. The MBM meat and bone meal
trend in co-firing is to increase the ratio of biomass/coal, and MC moisture content
to utilize a wider range of biomass fuels. PFBC pressurized fluidized bed combustion
(4) Combining torrefaction and pelletization, with leaching SCR selective catalytic reduction
biomass, and combining biomass pyrolysis with char washing SNCR selective non-catalytic reduction
are interesting options for pre-treatment processes, especially TGA thermogravimetric analysis
for herbaceous biomass (which currently is not often consid- VM volatile matter
ered for co-firing). wb wet basis
(5) Chemical composition and particle physical properties affect wt weight
reactor performance (e.g. fouling, agglomeration, and quality
of fluidization). Trouble-free feeding is crucial for the success
of co-firing. Subscripts
(6) More research is needed on co-firing biomass and coal includ- ave average
ing work on: preparation, handling, storage, and feeding of b bulk density
biomass feedstocks (e.g. drying, torrefaction, pelletization); mf minimum fluidization velocity
co-firing mechanisms; hydrodynamic analysis of co-firing p particle
combustors and gasifiers; boiler/gasifier capacity, slagging, v volume
fouling, corrosion, efficiency, reliability, fuel flexibility; lower
emissions and gas cleaning; catalyst poisoning; investment
and operating costs. REFERENCES
(7) In all the co-utilization technologies considered, there are Abbas, T., P. Costen, N. H. Kandamby, F. C. Lockwood and J. J.
technical problems and limitations that have not yet been Ou, “The Influence of Burner Injection Mode on Pulverized
fully resolved. However, none of the perceived technical Coal and Biosolid Co-Fired Flames,” Combust. Flame 99,
issues appears to be unsolvable. 617–625 (1994).
Adanez, J., L. F. de Diego, P. Gayan, L. Armesto and A.
Cabanillas, “A Model for Prediction of Carbon Combustion
NOMENCLATURE Efficiency in Circulating Fluidized Bed Combustors,” Fuel 74,
A Blake–Kozeny–Carman constant 1049–1056 (1995).
B Burke–Plummer constant Adanez, J., P. Gayan, L. F. de Diego, F. Garcia-Labiano and A.
Cd drag coefficient Abad, “Combustion of Wood Chips in CFBC: Modeling and
dn projected area diameter (m) Validation,” Ind. Eng. Chem. Res. 42, 987–999 (2003).

| 382 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 86, JUNE 2008 |
Aerts, D. J., K. M. Bryden, J. M. Hoerning and K. W. Ragland, Bi, J., “Co-Gasification of Biomass and Coal in a Fluidized bed,”
“Co-Firing Switchgrass in a 50 MW Pulverized Coal Boiler,” 2nd Biomass-Asia Workshop Bangkok, December 13–15,
Proceedings of 59th Annual American Power Conference, 2005.
Chicago, IL, 59, 1997, pp. 1180–1185. Biagini, E. and L. Tognotti, “Fundamental Aspects of
Aho, M. and E. Ferrer, “Importance of Coal Ash Composition in Biomass/Coal Co-Firing,” CCS 23rd Meeting, Italy, 2002,
Protecting the Boiler against Chlorine Deposition During http://iea.ccs.fossil.energy.gov/docs/Events/tognotti.pdf.
Combustion of Chlorine Rich Biomass,” Fuel 84, 201–212 Boerrigter, H., H. van der Drift and J. Kiel, “Biomass and Coal
(2005). Co-Gasification for Biosyngas Production,” Accelerating the
Aho, M. A., A. Gil, R. Taipale, P. Vainikka and H. Vesala, “A Biomass-to-Liquids (BTL) Implementation, 2006.
Pilot-Scale Fireside Deposit Study of Co-Firing Cynara with Brage, C., Q. Z. Yu and K. Sjostrom, “Characterization of Tars
Two Coals in a Fluidized Bed,” Fuel 87, 58–69 from Coal-Biomass Gasification,” Proceedings of Third
(2008). International Symposium on Coal Combustion Science and
Aho, M., P. Vainikka, R. Taipale, H. Vesala and K. Veijonen, “A Technology, Beijing, 1995, pp. 45–52.
Solution to Alkali Chlorine Deposition using a Protecting Fuel Breihofer, D., A. Mielenz and O. Rentz, “Emission Control of
and Simultaneous Measurement with Impactor, FTIR and SO2 , NOx and VOC at Stationary Sources in the Federal
Deposit Probes,” Proceedings of 14th European Biomass Republic of Germany,” Karlsruhe, 1991.
Conference, 2005, pp. 1323–1327. Brem, G., “Biomass Co-Firing: Technology, Barriers and
Amand, L., H. Miettinen-Westberg, M. Karlsson, B. Leckner, K. Experiences in EU,” G GCEP Advanced Coal Workshop, 2005.
Luecke, S. Budinger, E. U. Hartge and J. Werther, Brouwer, J., W. D. Owens, S. Harding, M. P. Heap and D. W.
“Co-Combustion of Dried Sewage Sludge and Coal/Wood in Pershing, “Co-Firing Waste Biofuels and Coal for Emissions
CFB: A Search for Factors Influencing Emissions,” Reduction,” Proceedings of 2nd Biomass Conference of the
Proceedings of 16th International Conference on FBC, New Americas, Portland, August, 1995, pp. 390–399.
York: ASME, 2001. Brown, C. R., Q. Liu and G. Gorton, “Catalytic Effects Observed
Andries, J., M. Verloop and K. Hein, “Co-Combustion of Coal during the Co-Gasification of Coal and Switchgrass,” Biomass
and Biomass in a Pressurized Bubbling Fluidized Bed,” Bioenergy 18, 499–506 (2000).
Proceedings of 14th International Conference on Fluidized Chao, C. Y. H., P. C. W. Kwong, J. H. Wang, C. W. Cheung and
Bed Combustion, Vancouver, BC, Canada, 1, 1997, G. Kendall, “Co-Firing Coal with Rice Husk and Bamboo and
pp. 313–320. the Impact on Particulate Matters and Associated Polycyclic
Armesto, L., A. Cabanillas, A. Bahillo, J. J. Segovia, R. Escalada, Aromatic Hydrocarbon Emissions,” Bioresour. Technol. 99,
J. M. Martinez and J. E. Carrasco, “Coal and Biomass 83–93 (2008).
Co-Combustion on Fluidized Bed: Comparison of Circulating Chen, G., K. Sjostrom, E. Bjornborm, C. Brage, C. Rosen and Q.
and Bubbling Fluidized Bed Technologies,” Proceedings of Z. Yu, “Coal/Wood Co-Gasification in a Pressurized Fluidized
14th International Conference on FBC, New York, ASME, Bed,” Proceedings of 3rd International Symposium on Coal
1997, pp. 301–311. Combustion Science and Technology, Beijing, China,
Armesto, L., A. Bahillo, A. Cabanillas, K. Veijonen, J. Otero, A. September 1995, pp. 383–390.
Plumed and L. Salvador, “Co-Combustion of Coal and Olive Chmielniak, T. and M. Sciazko, “Co-Gasification of Biomass and
Oil Industry Residues in Fluidized Bed,” Fuel 82, 993–1000 Coal for Methanol Synthesis,” Appl. Energy 74, 393–403
(2003). (2003).
Backreedy, R. I., L. M. Fletcher, J. M. Jones, L. Ma, M. Clift, R., J. R. Grace and M. E. Weber, “Bubbles, Drops and
Pourkashanian and A. Williams, “Co-Firing Pulverized Coal Particles,” Academic Press, New York (1978).
and Biomass: A Modeling Approach,” Proceedings of Colbert Fossil Plant #1, http://www.ieabcc.nl/database/cofiring.
Combustion Institute, 30, 2005, pp. 2955–2964. html, Tuscumbia, AL, U.S.A., 2005.
Baxter, L., “Ash Deposition during Biosolid and Coal Collot, A., Y. Zhuo, D. Dugwell and R. Kandiyoti, “Co-Pyrolysis
Combustion: A Mechanistic Approach,” Biomass Bioenergy 4, and Co-Gasification of Coal and Biomass in Bench-Scale
85–102 (1993). Fixed-Bed and Fluidized Bed Reactors,” Fuel 78, 667–679
Baxter, L., “Biomass-Coal Co-Combustion: Opportunity for (1999).
Affordable Renewable Energy,” Fuel 84, 1295–1302 CORDIS, “Solving Corrosion Issues in Combustion of Biomass,”
(2005). http://icadc.cordis.lu/fepcgi/srchidadb?CALLER=OFFR TM
BDC (Biomass Development Company), “Reducing the Use of EN& ACTION=D& RCN=2243, 2005.
Coal,” http://www.sovereign-publications.com/biomassdev. Dai, J., “Biomass Granular Feeding for Gasification and
htm, 2007. Combustion,” Ph.D. Thesis, The University of British
Belgiorno, V., G. Feo de, R. C. Della and R. M. A. Napoli, Columbia, Vancouver, BC, 2007.
“Energy from Gasification of Solid Wastes,” Waste Manage. Davidsson, K. O., J. G. Koresgren, J. B. C. Pettersson and U.
23, 1–15 (2003). Jaglid, “The Effects of Fuel Washing Techniques on Alkali
Benetto, E., P. Rousseaux and J. Blondin, “Life Cycle Assessment Release from Biomass,” Fuel 81, 137–142 (2002).
of Coal By-Products based Electric Power Production Dayton, D. C., D. Belle-Oudry and A. Nordin, “Effect of Coal
Scenarios,” Fuel 83, 957–970 (2004). Minerals on Chlorine and Alkali Metals Released during
Bergman, P. C. A., A. R. Boersma, R. W. R. Zwart and J. H. A. Biomass/Coal Co-Firing,” Energy and Fuel 13, 1203–1211
Kiel, “Torrefaction for Biomass Co-Firing in Existing (1999).
Coal-Fired Power Stations,” BIOCOAL, ECN, ECN-C-05-013, de Diego, L. F., F. Garcia-Labiano, A. Abad, P. Gayan and J.
2005. Adanez, “Modeling of the Devolatilization on Nonspherical
Bhattacharya, S. C., “State of the Art of Biomass Combustion,” Wet Pine Wood Particles in Fluidized Beds,” Ind. Eng. Chem.
Energy Sources 20, 113–135 (1998). Res. 41, 3642–3650 (2002).

| VOLUME 86, JUNE 2008 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 383 |
de Jong, W. and K. R. G. Hein, “Coal/Biomass Co-Gasification in Goh, B., “Biomass Combustion with Coal,” Institute of Physics,
a Pressured Fluidized Bed Reactor,” Renew Energy 16, Combustion Group Young Researchers Meeting,
1110–1113 (1999). Loughborough, September 21, 2005.
de Jong, W., J. Andries and K. R. G. Hein, “Coal-Biomass Grace J. R., H. T. Bi and M. Golriz, “Circulating Fluidized Beds,”
Gasification in a Pressurized Fluidized Bed Gasifier,” ASME Chapter 19, in “Handbook of Fluidization and Fluid-Particle
International GT and Aerospace Congress, Stockholm, SE, Systems,” W. C. Yang, Ed., Marcel Dekker, New York
June, 2–5, 1998, pp. 1–7. (2003).
Demirbas, A., “Combustion Characteristics of Different Biomass Gramelt, S., “FGD System for 600 MWe Coal Fired Power
Fuels,” Prog. Energy Combust. Sci. 30, 219–230 (2004). Plant-Process Description, PFD, Mass and Energy Balances,
Demirbas, A., “Potential Applications of Renewable Energy Equipment Specifications,” Deutsche Babcock Anlagen
Sources, Biomass Combustion Problems in Boiler Systems GMBH, Private Communication, 1994.
and Combustion Related Environmental Issues,” Prog. Energy Granada, E., G. Lareo, J. L. Miguez, J. Moran, J. Porteiro and L.
Combust. Sci. 31, 171–192 (2005). Ortiz, “Feasibility Study of Forest Residue Use as Fuel
Desroches-Ducarne, E., E. Marty, G. Martin and L. Delfosse, through Co-Firing with Pellet,” Biomass Bioenergy 30,
“Co-Combustion of Coal and Municipal Solid Waste in a 238–246 (2006).
Circulating Fluidized Bed,” Fuel 77, 1311–1315 (1998). Grena Kraftvarmeværk, http://www.ieabcc.nl/database/cofiring.
EBA (European Biomass Association), “Wood Pellets in Europe. html, Denmark, 2005.
State of the Art, Technologies, Activities, Markets,” Thermie B Gulyurtu, I., E. Frade, H. Lopes, F. Figueiredo and I. Cabrita,
DIS/2043/98-AT, Industrial Network on Wood Pellets “Combustion of Various Types of Residues in a Circulating
http://www.energyagency.at/(en)/publ/pdf/pellets net en. Fluidized Bed Combustor,” Proceedings of 14th International
pdf, 2000. Conference on FBC, New York, ASME, 1997,
EC (European Commission), “Addressing the Constraints for pp. 423–431.
Successful Replication of Demonstration Technologies for Hanaoka, T., S. Inoue, S. Uno, T. Ogi and T. Minowa, “Effect of
Co-Combustion of Biomass/Waste,” booklet DIS 1743/98-NL, Woody Biomass Components on Air-Steam Gasification,”
2000. Biomass Bioenergy 28, 69–76 (2005).
Elanchezian, C. and F. Antonio, “Successful Firing of Paper Mill Hansen, L. A., H. P. Michelson and K. Dam-Johansen, “Alkali
Sludges in Ahlstrom Pyroflow CFB Boilers,” Proceedings of Metals in a Coal and Biosolid Fired CFBC—Measurements
the 12th International Conference of Fluidized Bed and Thermodynamic Modelling,” Proceedings of the 13th
Combustion, New York, ASME, 1, 1993, pp. 231–238. International Conference on Fluidized Bed Combustion,
Ergun, S., “Fluid Flow through Packed Columns,” Chem. Eng. Orlando, FL, 1995, pp. 39–48.
Prog. 48, 89–94 (1952). Hein, K. R. G. and J. M. Bemtgen, “EU Clean Coal Technology
EUBION (European Bioenergy Networks), ALTENER, “Biomass Co-Combustion of Coal and Biomass,” Fuel Process. Technol.
Co-Firing—An Efficient Way to Reduce Greenhouse Gas 54, 159–169 (1998).
Emissions,” http://europa.eu.int/comm/energy/res/sectors/ Heinrich, H. and F. Weirich, “Pressurized Entrained Flow
doc/bioenergy/cofiring eu bionet.pdf, 2003. Gasifiers for Biomass,” Forschungszentrum Karlsruhe, IT3’02
Fernandez, J. C., in “Energy Plant Species,” N. El Bassam, Ed., Conference, New Orleans, Louisiana, 2002.
James & James (1998). Hotchkiss, R., W. Livingston and M. Hall, “Waste/Biomass
Fernandez, J. and M. D. Curt, “Low-Cost Biodiesel from Cynara Co-Gasification with Coal,” Report No. COAL R216 DTI/Pub
Oil,” Proceedings of the 2nd World Conference on Biomass URN 02/867, 2002.
for Energy, Industry and Climate Protection, 2004, Huang, Y., D. McIlveen-Wright, S. Rezvania, Y. D. Wang, N.
pp. 109–112. Hewitt and B. C. Williams, “Biomass Co-Firing in a
Fernando, R., “Fuels for Biomass Cofiring,” IEA Clean Coal Pressurized Fluidized Bed Combustion (PFBC) Combined
Centre, 2005. Cycle Power Plant: A Techno-Environmental Assessment
Ferrer, E., M. Aho, J. Silvennoinen and R. V. Nurminen, Based on Computational Simulations,” Fuel Process. Technol.
“Fluidized Bed Combustion of Refuse-Derived Fuel in 87, 927–934 (2006).
Presence of Protective Coal Ash,” Fuel Process. Technol. 87, Hupa, M., “Interaction of Fuels in Co-Firing in FBC,” Fuel 84,
33–44 (2005). 312–319 (2005).
Frandsez, F. J., “Utilizing Biomass and Waste for Power IEA, Co-Firing Database, http://www.ieabcc.nl/database/
Production—A Decade of Contributing to the Understanding, cofiring.html, 2005.
Interpretation and Analysis of Deposits and Corrosion Jacquet, L., J. Jaud, G. Ratti and J. P. Klinger, “Scaling Up of CFB
Products,” Fuel 84, 1277–1294 (2005). Boilers the 250 MWe GARDANNE CFB project,” Proc. Am.
Frazzitta, S., K. Annamalai and J. Sweeten, “Performance of a Power Conf. 56, 930–936 (1994).
Burner with Coal and Coal: Manure Blends,” J. Propulsion Jenkins, B. M., R. R. Bakker and J. B. Wei, “On the Properties of
Power 15, 181–186 (1999). Washed Straw,” Biomass Bioenergy 10, 177–200 (1996).
Gannon, F. J., Generating Station #3, http://www.ieabcc.nl/ Jenkins, B. M., L. L. Baxter, T. R. Miles Jr. and T. R. Miles,
database/cofiring.html, Florida, U.S.A., 2005. “Combustion Properties of Biomass,” Fuel Process. Technol.
Ganser, G. H., “A Rational Approach to Drag Prediction of 54, 17–46 (1998).
Spherical and Non-Spherical Particles,” Powder Technol. 77, Jensen, P. A., B. Sander and K. Dam-Johansen, “Pretreatment of
143–152 (1993). Straw for Power Production by Pyrolysis and Char Wash,”
Gayan, P., J. Adanez, L. F. de Diego, F. Garca-Labiano, A. Biomass Bioenergy 20, 431–446 (2001).
Cabanillas, A. Bahillo, M. Aho and K. Veijonen, “Circulating Johnsson, F. and B. Leckner, “Vertical Distribution of Solids in a
Fluidized Bed Co-Combustion of Coal and Biomass,” Fuel 83, CFB Furnace,” Proceedings of the 13th International
277–286 (2004). Conference on FBC, New York: ASME, 1995, p. 671.

| 384 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 86, JUNE 2008 |
Johnsson F., A. Svensson, and B. Leckner, “Fluidization Regimes Nordin, A., “Optimization of Sulfur Retention in Ash When
in Circulating Fluidized Bed Boiler,” in “Fluidization VII,” O. Co-Combusting High Sulfur Fuels and Biomass Fuels in a
Potter and D. Nicklin, Eds., Engineering Foundation Small Pilot Scale Fluidized Bed,” Fuel 74, 615–622 (1995).
Conference, New York (1992), p. 471. Ohlsson, O., “Results of Combustion and Emissions Testing
Keyser, M. J., M. Conradie, M. Coertzen and J. C. Van Dyk, When Co-Firing Blends of Binder-Enhanced Densified
“Effect of Coal Particle Size Distribution on Packed Bed Refuse-Derived Fuel (b-dRDF) Pellets and Coal in a 440 MWe
Pressure Drop and Gas Flow Distribution,” Fuel 85, Cyclone Fired Combustor,” Vol. 1, Test Methodology and
1439–1445 (2006). Results. Report No. DE94000283. Argonne National
Kiel, J., “Co-Utilization of Coal, Biomass and Other Fuels,” Laboratory, IL, 1994, p. 60.
Presented at JRC-Integration & Enlargement Workshop on the Pallares, I. D. and F. Johnsson, “Project Report JOR3CT980306,”
Perspectives for Cleaner Fossil Fuel Energy Conversion Department of Energy Conversion, Chalmers University of
Technologies in an Enlarging EU, Petten, the Netherlands, Technology, Sweden, 2000.
ECN, 2005. Pan, Y. G., E. Velo, X. Roca, J. J. Manya and L. Puigjaner,
Kumabe, K., T. Hanaoka, S. Fujimoto, T. Minowa and K. “Fluidized-Bed Co-Gasification of Residual Biomass/Poor Coal
Sakanishi, “Co-Gasification of Woody Biomass and Coal with Blends for Fuel Gas Production,” Fuel 79, 1317–1326 (2000).
Air and Steam,” Fuel 86, 684–689 (2007). Philippek, C. and J. Werther, “Co-Combustion of Wet Sewage
Kurkela, E., “Gasification-Based Co-Firing of Biomass and Sludge in a Coal-Fired Circulating Fluidized-Bed Combustor,”
Recovered Fuels in Coal-Fired Boilers,” Presentation to J. Inst. Energy 70, 141–150 (1997).
Amsterdam Biomass Conference, June 2002. Pinto, F., C. Franco, R. N. Andre, C. Tavares, M. Dias and I.
Kurkela, E., J. Laatikainen and P. Stahlburg, “Co-Gasification of Gulyurtlu, “Effect of Experimental Conditions on
Biomass and Coal,” APAS Clean Coal Technology Programme, Co-Gasification of Coal, Biomass and Plastics Wastes with
1992–1994, 3, C9, (1994). Air/Steam Mixtures in a Fluidized Bed System,” Fuel 82,
Laursen, K. and J. R. Grace, “Some Implications of 1967–1976 (2003).
Co-Combustion of Biomass and Coal in a Fluidized Bed Rajaram, S., “Next Generation CFBC,” Chem. Eng. Sci. 54,
Boiler,” Fuel Process. Technol. 76, 77–89 (2002). 5565–5571 (1999).
Leckner, B., “Possibilities and Limitations of Co-Firing of Rauhalahti Municipal CHP Plant, http://www.ieabcc.nl/
Biomass,” 1st Project Conference AGS, Stockholm, October database/cofiring.html, Finland, 2005.
2006. Reinoso, C., A. Cuevas, K. Janssen, M. Morris, K. Lassing, T.
Leckner, B. and M. Karlsson, “Emissions from Circulating Nilsson, H. P. Grimm, L. Puigjaner, G. P. Ying, E. Velo, M.
Fluidized Bed Combustion of Mixtures of Wood and Coal,” Zaplana, J. T. McMullan, B. C. Williams, E. P. Sloan and D.
Proceedings of 12th International Conference on FBC, New McIlveen-Wright, “Fluidized Bed Combustion and
York, ASME, 1993, pp. 109–115. Gasification of Low-Grade Coals and Biomass in Different
Lenzing, http://www.ieabcc.nl/database/cofiring.html, Austria, Mixtures in Pilot Plants Aiming to High Efficiency and Low
2005. Emission Processes,” APAS Clean Coal Technology
Lu, G., Y. Yan, S. Cornwell, M. Whitehouse and G. Riley, “Impact Programme, 1992–1994, 3, C5 (1994).
of Co-Firing Coal and Biomass on Flame Characteristics and Rickets, B., “Technology Status Review of Waste/Biomass
Stability,” Fuel 87, 1133–1140 Co-Gasification with Coal,” IChem Fifth European
(2008). Gasification Conference, Netherlands, 2002.
Maciejewska, A., H. Veringa, J. Sanders and S. D. Peteves, Ross, A. B., J. M. Jones, S. Chaiklangmuang, M. Pourkashanina,
“Co-Firing of Biomass with Coal: Constraints and Role of A. Williams, K. Kubica, J. T. Andersson, M. Kerst, P.
Biomass Pre-Treatment,” http://ie.jrc.ec.europa.eu/ Danihelka and K. D. Bartle, “Measurement and Prediction of
publications/scientific publications/2006/EUR22461EN.pdf, the Emission of Pollutant from the Combustion of Coal and
EUR 22461 EN (2006). Biomass in a Fixed Bed Furnace,” Fuel 81, 571–582 (2002).
Madsen, M. and E. Christensen, “Combined Gasification of Coal Saastamoinen, J., H. Hasa, J. Pitsinki, A. Tourunen and J.
and Straw Coal,” APAS Clean Coal Technology Programme, Hamalainen, “A Simplified Method to Predict Heat Release
1992–1994, 3, C2, (1994). Profiles in a Circulating Fluidized Bed Reactor,” Circulating
McIlveen-Wright, D. R., Y. Huang, S. Rezvani and Y. Wang, “A Fluidized Bed Technology VIII, Beijing, 2005, pp. 313–320.
Technical and Environmental Analysis of Co-Combustion of Salmenoja, K., M. Makela, M. Hupa and R. Backman,
Coal and Biomass in Fluidized Bed Technologies,” Fuel 86, “Superheater Corrosion in Environments Containing
2032–2042 (2007). Potassium and Chlorine,” J. Inst. Energy 69, 155–162
McLendon, T. R., A. P. Lui, R. L. Pineault, S. K. Beer and S. W. (1996).
Richardson, “High-Pressure Co-Gasification of Coal and Sami, M., K. Annamalai and M. Wooldridge, “Co-Firing of Coal
Biomass in a Fluidized Bed,” Biomass Bioenergy 26, 377–388 and Biomass Fuel Blends,” Prog. Energy Combust. Sci. 27,
(2004). 171–214 (2002).
Melin, S., Personal Communication about Co-Firing, (2007). Sampson, G. R., A. P. Richmond, G. A. Brewster and A. F.
Nemec, D. and J. Levec, “Flow through Packed Bed Reactors. 1. Gasbarro, “Co-Firing of Wood Chips with Coal in Interior
Single-Phase Flow,” Chem. Eng. Sci. 60, 6947–6957 Alaska,” For. Prod. J. 41, 53–56 (1991).
(2005). Savolainen, K., “Co-Firing of Biomass in Coal-Fired Utility
Nevalainen, H., M. Jegoroff, J. Saastamoinen, A. Tourunen, T. Boilers,” Appl. Energy 74, 369–381 (2003).
Jantti, A. Kettunen, F. Johnsson and F. Niklasson, “Firing of Segrest, S. A., D. L. Rockwood, J. A. Stricker and A. E. S. Green,
Coal and Biomass and Their Mixtures in 50 kW and 12 MW “Biomass Co-Firing with Coal at Lakeland Utilities,” Final
Circulating Fluidized Beds-Phenomenon Study and Report to The United States Department of Energy,
Comparison of Scales,” Fuel 86, 2043–2051 (2007). 1997–1998.

| VOLUME 86, JUNE 2008 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 385 |
Seward Generating Station #12, http://www.ieabcc.nl/database/ Veijonen, K., P. Vainikka, T. Järvinen and E. Alakangas,
cofiring.html, Pennsylvania, 2005. “Biomass Co-Firing—An Efficient Way to Reduce Greenhouse
Siegel, V., B. Schweitzer, H. Spliethoff and K. R. G. Hein, Gas Emissions,” http://ec.europa.eu/energy/res/sectors/doc/
“Preparation and Co-Combustion of Cereals with Hard Coal in bioenergy/ cofiring eu bionet.pdf, 2003.
a 500 kW Pulverized-Fuel Test Unit,” Biomass for Energy and Viewls, “Biofuel and Bio-Energy Implementation Scenarios,
the Environment, Proceedings of the 9th European Bioenergy Modeling Studies,” Final Report of VIEWLS WP5, 2005.
Conference, Copenhagen, Denmark, 2, 1996, pp. 1027–1032. Wan, H. P., Y. H. Chang, W. C. Chien, H. T. Lee and C. C. Huang,
Sjostrom, K., E. Bjornbom, G. Chen, C. Brage, C. Rosen and Q. “Emissions during Co-Firing of RDF-5 with Bituminous Coal,
Yu, “Synergetic Effects in Co-Gasification of Coal and Paper Sludge and Waste Tires in a Commercial Circulating
Biomass,” APAS Clean Coal Technol 1992–1994, 3, C3 (1994). Fluidized Bed Co-Generation Boiler,” Fuel 87, 761–767
Sjostrum, K., G. Chen, Q. Yu, C. Brage and C. Rosen, “Promoted (2008).
Reactivity of Char in Co-Gasification of Biomass and Coal: Werther, J., E. U. Hartge, K. Luecke, M. Fehr, L. E. Amand and B.
Synergies in the Thermo-Chemical Process,” Fuel 78, Leckner, “New Air-Staging Techniques for Co-Combustion in
1189–1194 (1999). Fluidized Bed Combustors,” VGB-Conference Research for
Skodras, G., P. Grammelis, P. Samaras, P. Vourliotis, E. Kakaras Power Plant Technology, 2000, pp. 1–2.
and G. P. Sakellaropoulos, “Emissions Monitoring during Xie, Z., J. Feng, W. Zhao, K. C. Xie, K. C. Pratt and C. Z. Li,
Coal Waste Wood Co-Combustion in an Industrial Steam “Formation of NOx and SOx Precursors during the Pyrolysis of
Boiler,” Fuel 81, 547–554 (2002). Coal and Biomass. Part IV. Pyrolysis of a Set of Australian and
Slough Trading Estate, http://www.ieabcc.nl/database/cofiring. Chinese Coals,” Fuel 80, 2131–2138 (2001).
html, UK, 2005. Zulfiqar, M., B. Moghtaderi and T. F. Wall, “Flow Properties of
Spring Grove Paper Mill (York), http://www.ieabcc.nl/database/ Biomass and Coal Blends,” Fuel Process. Technol. 87,
cofiring.html, PA, U.S.A., 2005. 281–288 (2006).
Stack M. M. and B. D. Jana, “Modeling Particulate Zuwala, J. and M. Sciazko, “Co-Firing Based Energy
Erosion—Corrosion in Aqueous Slurries: Some Views on the Systems-Modeling and Case Studies,” Paper presented at the
Construction of Erosion—Corrosion Maps for a Range of Pure 14th European Biomass Conference & Exhibition Biomass for
Metals,” Wear 256, 986–1004 (2004). Energy, Industry and Climate Protection, Paris, 2005.
Stora Enso Fors Ltd. (Fors), http://www.ieabcc.nl/database/
cofiring.html, Sweden, 2005.
Tacoma Steam plant #2, Tacoma http://www.ieabcc.nl/database/ Manuscript received December 24, 2007; revised manuscript
cofiring.html, Finland, Washington, U.S.A., 2005. received February 15, 2008; accepted for publication February 21,
Thomas Hill Energy Center #2, http://www.ieabcc.nl/database/ 2008.
cofiring.html, Missouri, MO, U.S.A., 2005.
Tillman, D. A., “Biomass Cofiring: The Technology, the
Experience, the Combustion Consequences,” Biomass
Bioenergy 19, 365–384 (2000).
Tmej, C. and H. Haselbacher, “Development of Wood Powder
Feeding into Gas Turbine Combustion Chambers,” First World
Conference on Biomass for Energy and Industry, Sevilla,
2000.
Tsai, M. Y., K. T. Wu, C. C. Huang and H. T. Lee, “Co-Firing of
Paper Mill Sludge and Coal in an Industrial Circulating
Fluidized Bed Boiler,” Waste Manage. 22, 439–442 (2002).
Tuurna, S., VTT Processes, State of the Art Report-Lifetime
Analysis of Boiler Tubes, Research report No. TUO74-021828,
(2003).
van den Broek, R., A. Faaij and A. van Wijk, “Biomass
Combustion for Power Generation,” Biomass Bioenergy 11,
271–281 (1996).
Van Doorn, J., P. Bruyn and P. Vermeij, “Combined Combustion
of Biomass, Municiple Sewage Sludge and Coal in an
Atmospheric Fluidized Bed Installation,” Biomass for Energy
and the Environment, Proceedings of the 9th European
Bioenergy Conference, Copenhagen, Denmark, 2, 1996,
pp. 1007–1012.
van Loo, S., “Biomass Co-Firing with Coal: An Overview of
Possibilities and Constraints,” Accelerating the Deployment of
Renewable Energy in the Baltic’s Riga, Latvia, 2002,
www.ieabioenergy-task32.com.
van Loo S. and J. Koppejan, Eds., “Handbook of Biomass
Combustion and Co-Firing,” Prepared by Task 32 of the
Implementing Agreement on Bioenergy under the Auspices of
the International Energy Agency, Twente University Press,
Enschede, Netherlands (2004).

| 386 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 86, JUNE 2008 |

You might also like