You are on page 1of 44

PERGAMON

Progress in Energy and Combustion Science 27 (2001) 171214 www.elsevier.com/locate/pecs

Co-ring of coal and biomass fuel blends


M. Sami, K. Annamalai*, M. Wooldridge 1
Department of Mechanical Engineering, Texas A & M University, College Station, TX 77843-3123, USA Received 4 August 1999; accepted 6 June 2000

Abstract This paper reviews literature on co-ring of coal with biomass fuels. Here, the term biomass includes organic matter produced as a result of photosynthesis as well as municipal, industrial and animal waste material. Brief summaries of the basic concepts involved in the combustion of coal and biomass fuels are presented. Different classes of co-ring methods are identied. Experimental results for a large variety of fuel blends and conditions are presented. Numerical studies are also discussed. Biomass and coal blend combustion is a promising combustion technology; however, signicant development work is required before large-scale implementation can be realized. Issues related to successful implementation of coal biomass blend combustion are identied. 2001 Published by Elsevier Science Ltd.
Keywords: Co-ring; Coal; Biomass; Emissions; Renewable energy

Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Fundamental combustion issues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. Material and combustion characteristics of coal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Pyrolysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3. Volatiles oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4. Char reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5. Homogeneous reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.6. Ignition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.7. Char combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.8. Pollutant emissions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. Biomass fuel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1. Pyrolysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Ignition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3. Char combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4. Fouling issues in biomass combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5. Summary comparison of coal and biomass combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4. Co-ring of blends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. Blend combustion efciency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Classes of co-ring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3. Coal and agricultural residues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4. Coal and RDF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172 178 178 179 179 179 180 180 181 182 183 186 187 187 189 191 191 192 192 193 195

* Corresponding author. Tel.: 1-979-845-2562; fax: 1-979-862-2418. E-mail address: kannamalai@mengr.tamu.edu (K. Annamalai). 1 Current address: Mechanical Engineering and Applied Mechanics Department, University of Michigan, Ann Arbor, MI 48109-2125, USA. 0360-1285/01/$ - see front matter PII: S0360-128 5(00)00020-4 2001 Published by Elsevier Science Ltd.

172

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

4.5. Coal and animal waste . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6. Combustion modeling for coal biomass blends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.7. Fouling issues in co-ring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5. Issues and opportunities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

200 203 208 209 211 212 212

1. Introduction The combustion of fossil fuels provides almost 85% of the energy requirement in the United States [1]. In 1997, the US electrical power utilities consumed 87% of the nearly 1.1 billion tons of coal produced [1]. Due to the large coal reserves, the United States relies heavily on coal for electricity generation. Coal was used to provide 51% of the total power generation in the United States in 1995 and is projected to provide 49% in 2020 [1]. Hence, coal will continue to be the dominant fuel for use in electricity production in the United States for the foreseeable future. As such, the air pollution emissions accompanying the coal combustion are signicant. Among these pollutants are oxides of sulfur (SOx) and nitrogen (NOx), which lead to acid rain and ozone depletion. In addition, greenhouse gas emissions (CO2, CH4, etc.) have become a global concern. Due to concerns over public health and the environment, federal regulations regarding the emission of air pollutants have become particularly demanding. The EPA New Source Performance Standards (NSPS) passed in 1990 require a 50% reduction of emissions that lead to acid rain (i.e. NOx and SO2). The previous standard was 520 g SO2/GJthermal (1.2 lb/MBtu), where GJthermal is the thermal energy from fuel combustion. The current standard is 260 g SO2/GJthermal (0.6 lb/MBtu). In Sweden and in some portions of the United States, the standards are even more severe, limiting emissions to 50 g SO2/GJthermal (0.115 lb/MBtu). As a direct result of these regulations, almost 50% reduction in SO2 from utility boilers has been achieved during the time period from 1980 to 1995 [1]. The current standard for NOx is 260 g/GJthermal (0.6 lb/MBtu). However, little progress has been made in reducing NOx or CO2 emissions [1]. A number of techniques and methods have been proposed for reducing gaseous emissions of NOx, SO2 and CO2 from fossil fuel combustion and for reducing costs associated with these mitigation techniques. Some of the control methods are expensive and therefore increase production costs. Among the less-expensive alternatives, co-ring has gained popularity with the electric utilities producers. Co-ring, in this context, is dened as the ring of a renewable fuel (i.e. biomass) along with the primary fuel (coal, natural gas, furnace oil, etc.). Recent studies in Europe and the United States [25] have established that burning biomass with fossil fuels has a positive

impact both on the environment and the economics of power generation. The emissions of SO2 and NOx were reduced in most co-ring tests (depending upon the biomass fuel used). The CO2 net production was also inherently lower, because biomass is considered CO2neutral. In addition, total fuel costs can be reduced in some cases if the biomass processing costs (transportation, grinding, etc.) are lower, on energy basis, than the primary fuel processing costs on an energy basis. In general, any organic fuel can be considered a biomass fuel. For the context of this discussion, biomass is used to describe waste products and dedicated energy crops. Waste products include wood waste material (e.g. saw dust, wood chips, etc), crop residues (e.g. corn husks, wheat chaff, etc.), and municipal, animal and industrial wastes (e.g. sewage sludge, manure, etc.). Dedicated energy crops, including short-rotation woody crops like hard wood trees and herbaceous crops like switchgrass, are agricultural crops that are solely grown for use as biomass fuels. These crops have very fast growth rates and can therefore be used as a regular supply of fuel. Biomass fuels are considered environmentally friendly for several reasons. First, there is no net increase in CO2 as a result of burning a biomass fuel (i.e. fossil generated CO2). Biomass consumes the same amount of CO2 from the atmosphere during growth as is released during combustion. Therefore, blending coal with biomass fuels can reduce fossil-based CO2 emissions [2,4]. Co-ring of biomass residues, rather than crops grown for energy, brings additional greenhouse gas mitigation by avoiding CH4 release from the otherwise landlled biomass. It is believed that CH4 is 21 times more potent than CO2 in terms of global warming impact. Most biomass fuels have very little or no sulfur, and therefore net SO2 emissions can also be reduced by co-ring coal and biomass. This attribute is particularly desirable when co-ring with high sulfur coals. The alkaline ash from biomass also captures some of the SO2 produced during combustion [2,6]. Typically, woody biomass contains very little nitrogen on mass basis as compared to coal. In addition, most of the fuel nitrogen in biomass is converted to NH radicals (mainly ammonia, NH3) during combustion [2,6]. The ammonia reduces NO to molecular nitrogen (essentially providing an in situ thermal DeNOx source). Hence, biomass co-ring can also result in lower NOx levels.

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

173

Nomenclature d d0 dp h hm hc m _ mc _ mO2 mv t tpyr B BO2 Bv Cp D DIV,c Dw E Ru Sh Tad Tg Tp Tp,I Y O2 ;w YO2 ; particle diameter at time t (m) initial particle diameter (m) particle diameter (m) convective heat transfer coefcient (W/m 2 K) mass transfer coefcient (kg/m 2 s) Heat of combustion (kJ/kg) mass ow rate (kg/s) char consumption rate (kg/s) oxygen consumption rate (kg/s) volatile matter in biomass (kg) time (s) pyrolysis time (s) transfer number pre-exponential factor for oxygen consumption (m/s) pre-exponential factor (m/s) specic heat of air (kJ/kg K) diffusion coefcient (m 2/s) fourth Damkohler number for carbon oxidation (BO2dpYO2 ; Ruhc)/(ShDwy O2ECp) diffusion coefcient at particle wall (m 2/s) activation energy (kJ/kmol) universal gas constant (8.314 kJ/kmol K) Sherwood number hmdp/r D adiabatic ame temperature (K) gas temperature (K) particle temperature (K) particle ignition temperature (K) mass fraction of oxygen at particle wall mass fraction of oxygen in ambient

h ch,b r r ch r s u ,HI

burn-out fraction of coal or biomass char mixture density (kg/m 3) char density (kg/m 3) ambient density (kg/m 3) StefanBoltzmann constant (W/K 4) dimensionless temperature Tp,I/(E/Ru)

Greek symbols a combustion rate constant (m 2/s) e emissivity y O2 stoichiometric coefcient for reactions (I) or (II) h mixture fraction h blend blend combustion efciency h coal coal combustion efciency

Abbreviations APAS activite de promotion, daccompagnement et de Suivi APF annular primary fuel Btu British thermal unit CFB circulating uidized bed CFD computational uid dynamics CPF central primary fuel DAF dry ash free EPA environmental protection agency FBC uidized bed combustion FC xed carbon FCM nished composted manure GI gas ignition HC hydrocarbons HI heterogeneous ignition HV heating value HVc heating value of char HVcoal heating value of coal HVv heating value of volatile matter HHV higher heating value MCFBC multi-circulating uidized bed combustion MSW municipal solid waste NSPS new source performance standards PCGC2 pulverized coal gasication and combustion code two-dimensional PC pulverized coal PCM partially composted manure RDF refuse-derived fuel SIT spontaneous ignition temperature TGA thermogravimetric analysis TEM transmission electron microscope VM volatile matter XN nitrogen containing volatile matter

Blending can also result in the utilization of less-expensive fuels with a possible reduction in fuel costs (depending upon the biomass fuel processing costs). Lastly, soil and water pollution can be mitigated depending upon the type of biomass fuel blended with coal. Stored biomass wastes anaerobically (i.e. in the presence of bacteria and moisture) release CH4, NH3, H2S, amides, volatile organic acids, mercaptans, esters and other chemicals. By combusting the biomass, ambient emissions of these pollutants are reduced. For example, if cattle manure is used as an inexpensive alternative biomass

fuel then many of the advantages described above are realized, in addition to avoiding contamination of soil, water and air due to otherwise stockpiled manure. Some typical biomass fuels found in coal co-ring studies are: cattle manure [7], sawdust and sewage sludge [8,9], switchgrass [10], wood chips [11,12], straw [9,1316] and refuse-derived fuels [8,17]. Table 1 provides a brief overview of some selected studies on co-ring coal and biomass fuels. The material properties of a number of biomass fuels used in co-ring are given in Tables 26.

Table 1 Selected coal:biomass co-ring studies System description Fuel type and heating value HHV (kJ/kg) Coal/wood chips blends. Coal: 22,605; wood: 17,742 Coal/manure blends. Coal: 26,535; manure: 8,650 Coal/sawdust blends. Coal: 32,260; pine sawdust: 18,140 Coal/straw/wood chips blends. Heating values not available Coal/manure blends. Coal: 26,535; manure: 8650 System capacity, blend parameters 3541% moisture, 1020% wood Results Issues Application Reference Remarks

174

Grate-red boiler burner

Only 1020% coring feasible, 0.040.09 grains/SCF opacity SOx and NOx decrease with blend combustion, easy ignition with blend 8190% burnout, NOx reduced, optimum coring ratio 30% for maximum burnout and minimum NOx

Down-red concentric swirl burner Wall-red dual fuel burner

35.4 kW (fuel) 100 g/min blend feed rate; 20% manure (mass basis) 500 kW (fuel); two-fuel injection modes, secondary swirl of 1.0 20 MW 1849% biomass

Blend mixing difcult. Stoker capacity problems Crushing manure to same size as coal difcult Fuel injection mode depends on reactivity and N2 in biomass

Electricity or steam generation

Sampson et al. [11]

Wood: birch, aspen, spruce. Break-even price $26 (45% moisture)

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Power generation

Frazzitta et al. [7]

Power generation

Abbas et al. [8]

Multicirculating uidized bed combustor (MCFBC) 15.2 cm (dia) 1.57 m (height) swirl combustor

No steady output of gaseous alkali metals

Power generation

Hansen et al. [77]

90:10 blend, three-mixture fraction approach

Wall-red boiler burners

Wall-red boiler burners

Coal/switchgrass blends. Coal: 25,500; switchgrass: 15,997 Coal/ straw/cereal blends. Heating values not available

50 MW (fuel) 15% co-ring mass basis, 12% moisture in biomass 500 kW 0100% biomass ring heat basis

Higher burnout with blend, differences in the near burner region of temperature and species concentration No slagging, normal unit operation, NOx decreased by 20%, lower opacity Fuel with higher nitrogen content should be injected in fuel rich zone to reduce NOx. Optimum co-ring ratio 60%

Power generation

Dhanaplan et al. [79]

Coal and sawdust fed separately. Coal: 74% 90 mm, sawdust: 75% 1.4 mm Coal and wood injected at bottom, straw injected with secondary air Comparison of two- and threemixture fraction approaches

Some traces of partially burned switchgrass in ash

Power generation

Aerts et al. [10]

Power generation

Siegel et al. [15]

Three different burner congurations studied

Table 1 (continued) System description Fuel type and heating value HHV (kJ/kg) Coal/railroad ties. Heating values not available System capacity, blend parameters 0.51 Mbtu/h 20% co-ring mass basis Results Issues Application Reference Remarks

Spreader-stoker

Down-red PC furnace

Coal/hard wood/soft wood blends. Heating values not available

38 kW 15% co-ring mass basis, 12% moisture in biomass

Cyclone-red combustor

Coal/b-dRDF blends. Coal: 14,388; RDF: 12,955

440 MW 12% co-ring mass basis, 19% moisture in biomass

Down-red combustor

Coal/ straw/sewage sludge blends. Coal: 30,140; straw: 17,090; sewage: 10,510

0.5 MW 20% coring of straw (mass basis), 11% moisture in straw

NOx reduced by 25% due to low nitrogen content of biomass. CO reduced considerably at low excess air ratios Co-ring: unstaged combustion, NOx decreased by about 17% at 50% coring. Staged combustion, no reduction until co-ring ratio 50%, 60% NOx reduction achieved in reburn. optimal stoichiometry for reburn 0.85 NOx reduction 23% SO2 reduction 17%, particulate concentration (heat basis) increased by about 50% as the RDF contained more ash than coal Max. particle size for complete burnout: 6 mm for straw. Burner conguration important in NOx reduction

Power generation

Brouwer et al. [76]

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Power generation

Brouwer et al. [76]

Both co-ring and reburning tests were conducted

Power generation

Ohlsson [17]

Lower sulfur content of RDF and heterogeneous reaction of SO2 with the binder lowered SO2 signicantly

Power generation

Spliethoff and Hein [6]

175

176

Table 2 Chemical analysis and properties of selected biomass fuels Fuel type Agricultural residue Moisture Ash Volatiles Fixed carbon C H O N S HHV (kJ/kg) HHV (kJ/kg) g A:F h AFT (K) i
a b c d e f g h i

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Corn stover a

Cotton gin b

Coconut shell c

Rice husk d

Olive husk e

Corn cobs a

Mustard stalk f 35 3.25 54.6 7.15 10,730

Barley straw f 10.3 20.9 39.92 5.27 43.81 1.25 17,310 17,288 4.48 2302

Wheat straw f 8.9 19.8 43.2 5.0 39.4 0.61 0.11 17,510 17,499 6.8 1981

35 3.25 54.6 7.15 42.5 5.04 42.6 0.75 0.18 10,730 10,718 3.24 1895

11.5 14.5 42 5.4 35 1.4 0.5 15,500 15,459 4.58 2273

25 (% wet) 0.8 (% dry) 79 (% dry) 20.2 20,000

9.96 20.61 54.68 15.02 34.94 5.46 38.86 0.11 13,524 13,515 3.7 2315

33 1.6 47.8 5.1 45.4 0.1 17,993 5.28 2342

15 1.4 76.6 7 48.4 5.6 44.3 .3 15,549 15,549 4.72 2179

Paul and Buchele [24]. LePori [23]. Mendis [67]. Hartiniati et al. [68]. Maschio et al. [83]. Ebeling and Jenkins [29]. HHV based on Boie equation. Air to fuel ratio (DAF mass basis). Adiabatic ame temperature calculated from the ultimate analysis.

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214 Table 3 Chemical analysis and properties of selected wood products Fuel type Moisture Ash Volatiles Fixed carbon C H O N S HHV (kJ/kg) HHV (kJ/kg) g A:F h AFT (K) i
a b c d e f g h i

177

Peat a 28 2.2 53.3 5.9 36.7 1.7 0.2 15,300 15,286 4.69 2115

Fuelwood b 10 2 70 18 16,100

Sawdust c 7.3 2.6 76.2 13.9 46.9 5.2 37.8 0.1 0.04 18,140 18,136 5.55 2279

Hardwood d 50.2 6.2 43.5 0.1 21,000 21,000 6.02 2438

Softwood d 52.7 6.3 40.8 0.2 22,000 22,000 6.46 2422

Switchgrass e 11.99 4.61 42.02 4.97 35.44 0.77 0.18 15,991 15,974 5.01 2168

Redwood f 0.36 19.92 50.64 5.98 42.88 0.05 0.03 20,720 20,717 6.02 2419

Tan oak f 1.67 17.4 47.81 5.93 44.12 0.12 0.01 18,930 18,929 5.64 2339

Black locust f 0.8 18.26 50.73 5.71 41.93 0.57 0.01 19,710 19,709 6.0 2339

Kurkela et al. [69]. Kandpal et al. [84]. Abbas et al. [8]. Ragland et al. [85]. Aerts et al. [10]. Ebeling and Jenkins [29]. HHV based on Boie equation. Air to fuel ratio (DAF mass basis). Adiabatic ame temperature calculated from the ultimate analysis.

The primary objective of this review is to summarize the state of knowledge on suspension burning of pulverized coal and biomass fuel blends. Some relevant material on uidized bed combustion is also included. However, an in-depth discussion of uidized bed combustion of fuel blends is beyond the scope of this review. Co-ring with furnace oil or natural gas and pulverizer perTable 4 Chemical analysis and properties of selected municipal residue Fuel type Moisture Ash Volatiles Fixed carbon C H O N S HHV (kJ/kg) HHV (kJ/kg) c A:F d AFT (K) e
a b c d e

formance are not included in this work. Previous work on the use of biomass as an energy source has primarily addressed direct combustion, pyrolysis, anaerobic digestion or fermentation for synthesis of ethyl alcohol [1826]. Until recently, there have been few studies concerning the co-ring of coal/biomass blends for energy generation.

MSW

Tires

Table 5 Chemical analysis and properties of selected energy crops Fuel type Poplar a 1.33 16.35 48.45 5.85 43.69 0.47 0.01 19,380 19,379 5.7 2374 Eucalyptus (Grandis) a 0.52 16.93 48.33 5.89 45.13 0.15 0.01 19,350 19,349 5.6 2388

1638 1120 6778 612 15,95017,533

0.5 6.1 65.2 28.7 81.5 7.1 3.4 0.5 1.4 36,800 36,671 11.66 2492

Moisture Ash Volatiles Fixed carbon C H O N S HHV (kJ/kg) HHV (kJ/kg) b A:F c AFT (K) d
a b c d

Saxena and Jotshi [86]. Christian and Unsworth [70]. HHV based on Boie equation. Air to fuel ratio (DAF mass basis). Adiabatic ame temperature calculated from the ultimate analysis.

Ebeling and Jenkins [29]. HHV based on Boie equation. Air to fuel ratio (DAF mass basis). Adiabatic ame temperature calculated from the ultimate analysis.

178

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Table 6 Chemical analysis and properties of selected food processing residue Fuel type Moisture Ash Volatiles Fixed carbon C H O N S HHV (kJ/kg) HHV (kJ/kg) b A:F c AFT (K) d
a b c d

Sugarcane baggasse a 11.27 14.95 44.8 5.35 39.55 0.38 0.01 17,330 17,329 5.2 2338

Almond shells a 4.81 21.74 44.98 5.97 42.27 1.16 0.02 19,380 19,378 5.4 2445

Olive pits a 3.16 18.19 48.81 6.23 43.48 0.36 0.02 21,390 21,388 5.75 2561

Walnut shells a 0.56 21.16 49.98 5.71 43.35 0.21 0.01 20,180 20,179 5.84 2424

Peach pits a 1.03 19.85 53.0 5.9 39.14 0.32 0.05 20,820 20,815 6.46 2326

Ebeling and Jenkins [29]. HHV based on Boie equation. Air to fuel ratio (DAF mass basis). Adiabatic ame temperature calculated from the ultimate analysis.

The discussion is organized in the following format. Fundamental concepts relevant to coal combustion are summarized in Section 2. In Section 3, a description of biomass fuels and their combustion behavior is presented. Studies related to direct combustion of coal/biomass blends are discussed in Section 4. Issues regarding successful implementation of co-ring are discussed in Section 5 and conclusions are presented in Section 6. 2. Fundamental combustion issues To facilitate a discussion on coal/biomass blends, a brief review of the characteristics important to coal combustion is presented. It will be seen later that biomass fuels behave similarly to low-rank coals. 2.1. Material and combustion characteristics of coal Coal is a complex polymer consisting primarily of carbon, hydrogen, oxygen, nitrogen and sulfur. It is a compact, aged form of biomass containing combustibles, moisture, intrinsic mineral matter (originating from dissolved salts in water) and extrinsic ash (due to mixing with soil). Coal is formed by the following sequence: Vegetation 3 peat 3 lignite low rank coal 3 anthracite high rank coal Plant materials have high cellulose (CH2O) content and high molecular weights (on the order of 500,000 kg/kmol). After plants die and are exposed to high pressure and heat over a long period of time in dense swampy conditions, anaerobic micro-organisms assist in converting plant debris

into peat-like deposits. Overtime, peat bed becomes covered in sediment, increasing the pressure. The temperature in the peat bed also increases and chemical decomposition occurs, lowering the oxygen and hydrogen content. Hence, coal is formed rst as lignite, then as sub-bituminous and nally as anthracite. The C/O and C/H ratio increase throughout the formation process. Anthracite is almost all carbon, with a corresponding increase in the heating value. It takes approximately 200300 million years to form coal. The chemical properties of coal depend upon the relative proportions of the chemical constituents present in the parent plant debris, the nature and extent of the changes, which the constituents have undergone since deposition, and the nature and quantity of the inorganic matter present. Coal rank indicates the relative proportions of volatile matter (VM) and xed carbon (FC) present in the coal. Upon heating coal in an inert atmosphere, combustible gases are evolved from the coal due to thermal decomposition of the solid. This process is called pyrolysis. The remaining skeletal matter in the solid is called char, and is predominantly FC. Coal rank increases with decreasing VM. Typically, a medium rank coal consists of 40% VM and 60% FC while a high-rank coal has about 10% VM. The older the coal, the higher the rank. The highest ranked coal is graphite in structure. The higher or gross heating value (HHV) for a particular coal sample can be either measured or estimated using the ultimate analysis of the fuel and the following relation [27]: HHV kJ=kg 35;160C 6280N 116;225H 10;465S 11;090O 1

where C, H, O, N and S are the elemental mass fractions in the coal. The HHV can be determined on an as-received

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

179

basis or dry-ash-free basis (DAF). Both VM and FC contribute to the heat released from coal. If the heat of pyrolysis is neglected, the heat of combustion of coal can be represented as a combination of the contribution from the VM (HVv) and the contribution from the FC (HVc), where VM and FC content is on a per unit mass basis: HVcoal VMHVv FCHVc 2

2.3. Volatiles oxidation Once released, the volatiles undergo oxidation within the gas lm surrounding the particle (see Fig. 1(a)). During the volatiles combustion period, the gas phase temperature is much higher than the particle temperature [38,39]. Shadow photography by McLean et al. [36] and holography by Seeker et al. [38] reveal that volatiles may burn in jets or as a ame envelope. An enveloping ame acts like a shroud, preventing oxygen from reaching the particle surface and therefore preventing heterogeneous oxidation of char. Some approaches used to analyze the oxidation of coal volatiles are: 1. local equilibrium assumption [40,41] which assumes instantaneous chemical reaction to the equilibrium composition as the volatiles are released; 2. methane oxidation kinetics [42]; 3. kinetics controlled by CO kinetics [43]; 4. cellulose kinetics [44]; 5. global oxidation kinetics [45]; 6. oxidation to CO and H2 followed by CO oxidation kinetics [46,47]. Typically, one-step kinetics mechanisms over predict the rate of release of chemical energy. Depending upon the scheme, the rate of heat release changes, affecting the volatile combustion time scale. Because the total time for volatile combustion is on the order of a few ms, the particular reaction scheme does not affect the overall coal combustion time scale, which is on the order of seconds. However, the choice of approach does become important in obtaining accurate estimates of NOx and SOx emissions. 2.4. Char reactions The skeletal char remaining after pyrolysis is essentially FC. The carbon undergoes heterogeneous reactions with gaseous species. Heterogeneous reactions are generally governed by the following processes [48]: (i) diffusion of gas phase oxidizing reactant species to the particle surface; (ii) adsorption of gas phase species E 32;000 kJ=kmol; (iii) chemical reaction of the adsorbed species; (iv) desorption of the solid oxides E 170;000 kJ=kmol; (v) diffusion of gas phase products through the boundary layer to the free stream. If boundary layer diffusion, (i) or (v), controls the overall reaction rate, the system can be readily analyzed using mass transfer and uid mechanics relations. Oxygen transfer to carbon/char can occur via O2, CO2 and/or H2O. The heterogeneous combustion of carbon/char occurs primarily via one or more of the following reactions: C 1=2O2 3 CO I

If FC 1 VM as in the case of DAF coals, the heating values of the volatiles, HVv, can be correlated to VM [31]. Eq. (2) can be used to determine HVv if HVcoal and VM are known. The heating value of char (HVc) can be assumed as 32,500 kJ/kg. 2.2. Pyrolysis Fig. 1(a) and (b) is a schematic representation of the various physical mechanisms important in the pyrolysis and combustion of coal. The following sections briey describe the signicant characteristics of these mechanisms. When a coal particle is placed in a furnace, it is heated and pyrolysis ensues (thermal decomposition of coal). Pyrolysis is somewhat similar to vaporization; however, it is a relatively slow chemical process compared to vaporization. The temperature at which pyrolysis occurs depends on the fuel type and the heating rate. Typically, bituminous coal pyrolyzes at about 700 K (1% mass loss) for heating rates 100 C/s [32]. This temperature is at or below the spontaneous ignition temperature (SIT) for hydrocarbons (HC) [33]. At high heating rates ( 10,000 C/s), pyrolysis is presumed to start around 1500 K. The products of pyrolysis are volatile gases and the composition of these gases also depends on the fuel. Pyrolysis products range from lighter volatiles (CH4, C2H4, C2H6, CO, CO2, H2, H2O, etc.) to heavier tars. Typically the composition of the volatiles from lignite (low-rank) coals at 1300 K is 3% CH4 and 38% CO and CO2 [34]. When experiments were conducted with a few micrograms of 60 mm coal particles sandwiched between an electrical wire mesh, pyrolysis started around 350400 C (1% mass loss) for very low heating rates of 10 4 C/s. At moderate heating rates of 100 C/s, pyrolysis occurred at 1200 C [35]. Volatiles were found to issue as jets from 80 to 100 mm diameter particles, while for Dp 40 mm; there was no evidence of volatiles jets [36]. Recent experiments on 1 mm particles using digital imaging system conrmed the jetting behavior [37]. Pyrolysis time scales are on the order of 10100 ms for a particle diameter of 1 mm. Apart from volatiles, nitrogen is also evolved from the fuel during pyrolysis in the form of NH3, HCN and other N2containing species which are generally represented as XN. Nitrogen evolution normally occurs during the later part of pyrolysis. Nitrogen evolved from fuel undergoes oxidation to NOx and is called fuel NOx to distinguish it from thermal NOx produced by oxidation of atmospheric nitrogen.

180

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

O2 Pyrolysis products (volatile matter) XN CO2, H2O coal coal

NOx

(a)

(i) Pyrolysis

(ii) Volatile oxidation

(iii) Char

O2 CO2 CO coal

CO CO2 H2 CO H2O

(b)

Fig. 1. (a) Schematic of coal combustion mechanisms [87]. Pyrolysis produces volatile matter which is oxidized, leaving a char skeletal matrix as a solid product; and (b) schematic depicting heterogeneous coal reactions for carbon particle [87].

C C C

O2 3 CO2 CO2 3 2CO H2 O 3 CO H2

II III IV

Assuming rst-order reaction for scheme (I), the oxygen consumption rate is given as, 2 3 E 2 _ mO2 pdp BO2 exp r Y 3 Ru Tp O2 ;w Similar expressions can be written using the other reactions. The dominant oxygen transfer mechanism at high temperatures is via reaction (I) with E=Ru 26; 200 K and BO2 2:3 107 m=s: Reaction (II) has an activation energy of E=Ru 20; 000 K and BO2 1:6 105 m=s: Reaction (III), the Boudouard reduction reaction, proceeds with an E/Ru of about 40,000 K. In general, char reaction with steam has been found to be 50 times faster than char reaction with CO2 for temperatures up to 2073 K (Montana Rosebud char, 75100 mm, 1 atm) [49]. While, at lower temperatures (1150 K), Matsui et al. [50] found that reactions (III) and (IV) proceed at approximately the same rate and that the reaction (I) proceeds 10 4 10 5 times faster than reaction (III) between 1100 and 1200 K. Reaction (II) is signicant at low temperatures (e.g. ignition conditions) while reaction (I) is dominant under typical combustion conditions. For char particle combustion in pure dry air, the CO2 and H2O mass fractions in the ambient are negligible.

However, in boiler burners, the fuel particles are closely spaced resulting in reduced oxygen availability to each particle. Combustion under such conditions is called cloud combustion. Under these conditions, the reduction reactions (III) and (IV) may become signicant, especially at high temperatures. 2.5. Homogeneous reactions Once CO and H2 are released via heterogeneous reactions (III) and (IV), they can undergo oxidation in the gas phase. The detailed CO reaction scheme is summarized by Mitchell et al. [51], Hottel et al. [43] and others [52,53]. 2.6. Ignition A brief discussion is presented on the ignition of an isolated coal particle and the relation of the ignition temperature to the reaction parameters for coal. The reader is referred to the review on ignition of coal by Essenhigh et al. [33] for a detailed discussion. Oil droplets which fully vaporize or a plastic which fully pyrolyzes (e.g. polymethyl methacrylate) ignite homogeneously (homogeneous ignition or gas ignition, GI). Carbon particles, which are non-gasifying upon heating in an inert environment, ignite heterogeneously (heterogeneous ignition, HI). Coal, like many polymers, is partially gasifying and is called a charring solid. Volatile matter, ammability of the volatiles and transport from the particle determine

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

181

whether ignition of an isolated coal particle occurs either heterogeneously or homogeneously. If the volatiles evolve early in the combustion sequence, they may remove oxygen from the internal pore surface area, thus preventing heterogeneous ignition. If coal is ignited homogeneously, the volatiles burn in the gas phase (similar to droplet ames) preventing oxygen from reaching the particle surface. If combustion is not complete or if volatile liberation occurs as jets or intermittently from the pores [36,54], then oxygen can still reach the particle surface. In such a situation, heterogeneous combustion of FC and combustion of in situ VM can proceed in parallel with gas phase combustion. The thermal explosion theory is commonly used to dene the onset of heterogeneous ignition. According to thermal explosion theory, ignition occurs if the rate of heat release due to chemical reaction is higher than the rate of heat loss due to convection, radiation, etc., and if the process results in runaway conditions. A correlation for the heterogeneous char ignition temperature can be determined using thermal explosion theory [55]: Tp;I ln 4 E=Ru 5 BO2 dp YO2 ; hc E 2 ShDw nO2 Ru Tp;I Cp 4

Fig. 2 shows the variation in heterogeneous ignition temperature with particle diameter. For small particles ( 30 mm or less), no realistic values for Tp,I could be found, which indicates that very small particles cannot be ignited heterogeneously. The rate of heat loss is too high compared to the rate of heat generation for smaller particles due to the high surface area to volume ratio. Fig. 3 shows a comparison of the experimentally measured non-dimensional minimum gas temperature (u ,HI) required for heterogeneous ignition [56] with those predicted by the steady state and transient models of Du and Annamalai [55]. The inverse of u ,HI is plotted as a function of the fourth Damkohler number for carbon oxidation. The Damkohler number is proportional to particle diameter, and u ,HI is proportional to particle temperature. It is clear that both theoretical and experimental results predict a decrease in ignition temperature as the particle size is increased. In other words, larger particles need lower temperatures for heterogeneous ignition to occur. Recent literature suggests that there is also an upper limit to particle size for heterogeneous ignition [57]. More details on the transient ignition phase can be found in Ref. [55], while studies on laser ignited particles are presented in Ref. [58]. 2.7. Char combustion

For temperatures below Tp,I, heterogeneous ignition of the coal particle will not occur. Note that Eq. (4) is an implicit relation for Tp,I. Eq. (4) was used to estimate the heterogeneous ignition temperatures for three different types of coal.

Once ignited, the combustion of high volatile content coal proceeds in two stages: combustion of the VM and combustion of the FC. Combustion of the VM is similar to the

Fig. 2. Variation of heterogeneous ignition temperature with particle diameter for various coal and biomass fuels; Sh 2; Dw 1 cm2 =s; YO2 ; 0:23:

182

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Fig. 3. Comparisons of experimental results for heterogeneous ignition with steady state and transient models, u;HI Ru Tp =E (X: BO2 35:6 m=s; E 46:3 kJ=mol; W: BO2 2388:1 m=s; E 103:4 kJ=mol; B: BO2 390:2 m=s; E 73:7 kJ=mol; A: BO2 66:8 m=s, E 56:0 kJ=mol; O: BO2 1903:4 m=s; E 82:8 kJ=mol [55].

combustion of vapors from an oil droplet. However, there are signicant differences. Consider a single oil droplet of radius a placed in a hot furnace. Due to energy supplied from the furnace to the droplet, the droplet is heated. Vapors released by the droplet move radially into the ambient where they mix with oxygen and establish a diffusion ame at a radius rf q a away from the droplet surface. The vaporization of the fuel drop is a non-chemically reacting process. The molecular composition of the vapor is the same as the drop molecular composition. On the other hand, coal particles partially pyrolyze releasing volatiles (a chemical decomposition process of the coal particles to gaseous hydrocarbon species) and leaving char particles (predominantly carbon). The gasication rate is controlled by the kinetics of pyrolysis. The volatiles diffuse into the surrounding atmosphere where they mix with oxygen and a ame is formed if the volatiles exist in sufcient proportion to the ambient air. After the volatiles are exhausted, only char (carbon and small amounts of hydrogen) remains. Now, oxygen diffuses to the char surface and reacts to form carbon monoxide and carbon dioxide (char oxidation). Unlike the combustion of volatiles, which diffuse towards the oxygen rich atmosphere (resulting in a large reaction area), the oxygen in heterogeneous combustion must be transported to the particle surface. Thus, the char combustion rate under diffusion limited conditions is very slow. In droplet combustion, the ame is located far away from drop surface and thus the reaction surface area is 4prf2 and the mass transfer coefcient for oxygen is on the order of rD=rf (the diffusion rate of O2 per unit area is slower). The oxygen consumption rate is on the order of 4prf rD and the burning rate is fast rf 30a for a quiescent ame). For a char/carbon particle

of radius a, the mass transfer coefcient is of the order of rD=a while the surface area is 4pa2 : The consumption rate of O2 is therefore on the order of 4parD: Under diffusion limited combustion of solid char (i.e. carbon) particles, the diameter can be shown as follows:
2 d 2 d0

at

where a 4ShrD=rch ln1 B; B YO2; =y O2 and y O2 may correspond to reactions (I) or (II). Similarly, steady state char particle temperature under diffusion limited combustion is given by: _ m c hc {hTp where, _ mc prDShdp ln1 B Tg

esT p4

2 Tj4 }pdp

_ For mc 0 and hc 0; Tp Tg and, for Tp Tg ; no combustion occurs. The diffusion-controlled burning rate is about 30 times slower for a carbon particle than an oil drop, mainly due to the reduced reaction surface area. The typical total combustion time for a 100 mm solid coal particle is on the order of 1s in boilers and is dominated by the time required for the heterogeneous combustion of the residual char particle. The pyrolysis time tpyr 2 106 s=m2 dp [32]) is on the order of 110% of the total burning time. 2.8. Pollutant emissions The nitrogen in coal, which mainly exists as XN (e.g. HCN, H3N, etc) readily oxidizes to form NOx (NO, NO2)

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

183

4500

4000 HHV/m3 of stoich. air

3500

3000

2500

2000
lls st ra w ck lo cu st co rn co bs co tto n gi eu ca n ly pt us ha rd w oo d m an ur e ol ive pi pe ts ac h pi ts rle y bl a pe at po pl ar re dw oo d ric e hu sk so ft w su oo ga sa d rc w an du e st ba gg sw ass e itc hg ra ss ta n oa w he k at wa stra w ln ut sh el ls on al m d ba sh e

Fuel
Fig. 4. DAF higher heating value per unit standard cubic meter of stoichiometric air for several biomass fuels.

at low temperatures (fuel NOx). The oxidation of atmospheric N2 to NOx (thermal NOx) occurs at higher temperatures ( 1800 K). The majority of NOx emitted from coalred plants originates from fuel nitrogen. Some NOx control technologies include:

1. Staged combustion, where oxidation near the burner occurs in a fuel-rich regime. Oxygen concentrations are lower in the high-temperature regions of the burner, therefore decreasing thermal NOx emissions. 2. Injection of NH3 at later stages of burning to reduce NOx to N2 (thermal DeNOx process). 3. Reburn combustion, where HC (or coal which releases HC during pyrolysis) are injected downstream of the primary combustion zone to reduce NOx to N2. Pollutant emissions are a growing concern as mentioned earlier and emission regulations are driving continuous development of new combustion technologies.

3. Biomass fuel Biomass fuels follow the same sequence of pyrolysis, devolatilization and combustion as seen in low-rank coal combustion mechanisms. However, there are some signicant differences between coal and biomass combustion. Coal densities typically range from 1100 kg/m 3 for low-rank coals to 2330 kg/m 3 for high-density pyrolytic graphite [59]. Biomass densities ranges from 100 kg/m 3 for straw to

500 kg/m 3 for forest wood [60]. Biomass usually consist of 7080% VM whereas coal consists of 1050% VM. The heating values of biomass fuels are appreciably lower than that of coals. Eq. (1) was recently used [28] to determine the HHV for biomass fuels and resulted in good agreement with the experimental results of Ebeling and Jenkins [29] for 62 types of biomass. A maximum error of 12% for rice straw and 4% for wheat dust was found [29]. It is known that the HHV per unit mass of stoichiometric oxygen mO2 ;s is approximately constant for most fuels (HHV/ mO2 ;s 12,50020,000 kJ/kg,O2,s) [30]. Converting to a per unit volume of stoichiometric air basis (Vair;g ) (using the density of air at 1 atm, 298 K) results in HHV/ Vair,s 34005400 kJ. Figs. 4 and 5 show the DAF HHV per cubic meter of stoichiometric air for several biomass and coal fuels, respectively. It is apparent from the gures that the DAF HHV/Vair,s is almost constant for all types of fuel (within ^2%). Fig. 6 is a plot of the adiabatic ame temperature as a function of the HHV/mO2 ;s (on an asreceived basis) for selected biomass fuels. Because the HHV/Vair,s are approximately the same value regardless of fuel type, the adiabatic ame temperature should remain constant when plotted on a DAF HHV/Vair,s basis, as shown in Fig. 7. The scatter observed in Fig. 6, in terms of the heating value, is less indicating that the ame temperature is approximately constant for all the fuels provided DAF fuels are used. Hence, the adiabatic ame temperature (Tad) can decrease if the ash and/or moisture content increases, or if the amount of stoichiometric air required for complete combustion increases. A parametric study of the effects of moisture and ash contents on the adiabatic ame temperature for sawdust is shown in Fig. 8.

184

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

3400 3200 HHV/m3 of stoich. air 3000 2800 2600 2400 2200 2000 meta anthracite low vol. bitum. Coal med. vol. bitum. coal Fuel
Fig. 5. DAF higher heating value per standard cubic meter of stoichiometric air for coal fuels (coal analysis from Ref. [31]).

high vol. bitum. Coal

sub bitum. Coal

An empirical curve t R2 0:99 for Tad is given as follows: a 1 bx gx cx hx2


2

coefcients are a 2336:7; b 24:5; c 0:0149; d 16:616;

Tad

dy iy

ey jy2

fy ky3

e g

0:3057; f 0:0037; 0:00727; h 7:70 10 6 ; i


6

0:0071;

where, x is the moisture content (% mass basis, 0 x 40 and y is the ash content (% mass basis, 0 y 40 and the
3000

0:00012 and k 1:44 10

2500

Adiabatic Flame Temper ature (K)

2000 cotton gin 1500 rice husk corn cobs saw dust 1000 switchgrass poplar eucalyptus 500 sugarcane baggase raw manure 0 0 5000 10000 15000 20000 25000

HHV (kJ/kg of stoichiometric oxygen)


Fig. 6. Adiabatic ame temperature as a function of higher heating value (HHV/kg of stoichiometric oxygen, as received basis) for several biomass fuels. The values are bracketed within 15,00020,000 kJ/kg, except for RM and rice husk.

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

185

Fig. 7. Adiabatic ame temperature as a function of DAF higher heating value per cubic meter of stoichiometric air for several biomass fuels and coals. The values are bracketed within 32004300 kJ/m 3.

If the temperature decreases below 1600 K, ame stability problems can occur in boilers and furnaces due to the reduced rates of chemical reaction. It is clear from Fig. 8 that at 40% ash content, the maximum moisture level to maintain the temperature at or above 1600 K is 34%. Similarly, at 30% ash, moisture should not be more than 40% in order to avoid ame instability.
2600

In order to illustrate further the differences in coal and biomass properties, Table 7 shows the proximate and ultimate analyses of Wyoming coal and feedlot manure, a biomass fuel. Feedlot manure (raw, PCM or FCM) contains approximately 82% VM on DAF basis as compared to Wyoming coal which contains only 36% VM. With aging, the VM in manure decreases as a result of the gradual

0% ash

Adiabatic Flame Temper ature (K)

2400 2200 2000 1800


critical temperature for flame stability

5% 10% 15% 20% 25% 30% 40%

1600 1400 1200 1000 0 10 20 30 40 50

Moisture (%)
Fig. 8. Adiabatic ame temperature for varying moisture and ash content for sawdust.

186

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Table 7 Chemical analysis and properties of coal and raw feedlot manure Parameter Moisture Ash Volatile matter Fixed carbon Heating value (as received), kJ/kg C H O N S Empirical formula a (DAF basis) Molecular weight b A:FStoichiometric b A:FStoichiometric b (DAF basis) Adiabatic ame temperature (K) c
a b c

Coal 10.8 5.68 30.72 52.80 26,535 54.9 4.33 23.32 0.76 0.34 CH0.94O0.32N0.012S0.0023 18.3 6.9 8.2 2505

Raw manure 36.61 25.25 31.57 6.57 7865 19.24 2.22 14.68 1.47 0.53 CH1.37O0.57N0.065S0.01 23.8 2.5 6.2 1705

Partially composted manure 30.02 28.01 34.11 7.86 8305 20.65 2.30 16.43 1.86 0.73 CH1.32O.597N.077S.0132 24.4 2.6 5.9 1790

Fully composted manure 35.35 30.73 27.86 6.06 6610 16.62 1.72 12.92 1.82 0.84 CH1.23O.585N.0937S.019 24.5 1.9 5.8 1685

Determined via ultimate analysis. Determined via empirical formula for fuel. Including ash and moisture in fuel. Stoichiometric condition; heating values conrmed with Boie equation.

release of hydrocarbon gases or dehydrogenation. The DAF heating value of feedlot manure, on average, is 15,000 kJ/kg (5000 Btu/lb) versus coal, which has a DAF heating value of 35,000 kJ/kg (14,000 Btu/lb). Even though the biomass has high VM contents, the heating values of biomass volatile are less than that of coal as shown in Table 8 (cf. Eq. (2)). Feedlot manure also has higher moisture ( 32%) and ash ( 28%) contents on mass basis than coal, which typically has 1025% moisture and 5% ash. Ash can be classied as extrinsic and intrinsic. Intrinsic ash is contained in the matrix of the biomass, whereas extrinsic ash comes from the biomass collection process. The ash contents of woody biomasses are much lower than for coal. However, the ash content of non-woody biomass fuels can vary widely. For example, rice hulls have 18% ash while almond shells have 5% ash. Most of the ash in high-ash biomasses are extrinsic in nature, consisting primarily of SiO2. Higher moisture and ash contents of biomass may cause ame instability during combustion if used in higher proportions
Table 8 Comparison of estimated heating contributions from volatiles Fuel Coal Sawdust Manure Rice husk Fuel wood Tires
a

in the blend. The potential impact on the burner temperature prescribes the maximum allowable percentage of biomass in the blend. 3.1. Pyrolysis Pyrolysis of biomass is thermal decomposition of the fuel. As with coal, pyrolysis is a relatively slow chemical reaction occurring at low temperatures. The reaction mechanisms are complex but can be dened in ve stages for wood [60]. Other biomass fuels with considerable woody matter would exhibit similar behavior. 1. Moisture and some volatile loss. 2. Breakdown of hemicellulose; emission of CO and CO2. 3. Exothermic reaction causing the wood temperature to rise from 250 to 359 C; emission of methane and ethane. 4. External energy is now required to continue the process. 5. Complete dissociation occurs.

HV of volatiles (kJ/kg VM) 31,375 17,994 18,256 15,945 14,773 42,360

VM a (%) 36.8 84.5 82.8 78.8 79.5 69.8

Heat from VM (%) 36.3 75.5 73.3 64.5 64.2 75.0

Heat from char (%) 63.6 24.5 26.7 35.5 35.8 25.0

Mass basis.

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214 Table 9 Pyrolysis kinetic data for selected biomass fuels and coals Fuel Coal Biomass Lignite a Bituminous a Hazel nut Rice husk Rice husk Forest wood Forest wood Hardwood Cellulose Lignin Hemicellulose Heat rate 2 C/s 100 C/min 100 C/min 2 C/s 2 C/s 0.83 C/s Temperature range ( C) 225350 350600 225325 700900 3001123 280350 390500 320400 A (1/s) 6.60 10 4 4.30 10 14 4.69 10 13 1.30 10 9 1.31 10 1 7.68 10 7 6.32 10 2 2.15 10 3 4.69 10 5 2.10 10 5 8.70 10 2 E (kJ/g mol) 105 229 89.8128.6 97.1 11.2 124.8 92.3 59.4 82.7 70.7 33.8

187

References [88] [89] [90] [91] [92] [93] [94]

Single coal devolatilization reaction assumed.

The main products of biomass pyrolysis depend on the temperature, heating rate, particle size and catalyst used. Typical gas composition of woody biomass pyrolysis includes CO, CO2, CH4 and H2 as major products along with other organic compounds. Usually, fast pyrolysis yields more gases than solids. The VM loss can be determined using an overall mass loss rate equation: dmv Bv mv exp dt  E RT  8

3.2. Ignition The ignition process of biomass is similar to that for coal except there is more VM available for reaction in a biomass fuel. It is, therefore, more likely that homogeneous ignition will occur for biomass fuels. The heterogeneous ignition temperature of biomass chars can be predicted using Eq. (5). Fig. 2 shows the variation of ignition temperature with particle diameter for biomass chars (wood and sewage sludge). BO2 was calculated using the char kinetic data given in Winter et al. [62]. For Ficus wood, the data given by Dasappa et al. [63] was used. Note that the biomass fuels exhibit a broader range of ignition temperature for a given particle size than coal does. 3.3. Char combustion The physical and chemical transformations of biomass during combustion have been studied by many researchers. Wornat et al. [64] investigated the combustion of switchgrass and pine char particles in a laminar ow reactor. Table 10 shows the fraction of each organic element as a function of char conversion. Large amounts of oxygen and hydrogen are lost from the char early during conversion, indicating the release of VM. The absence of a visible ame suggests that these volatiles are primarily CO, CO2 and H2O and not hydrocarbon gases. Unlike high-rank coals, oxygen levels remain high (almost an order of magnitude higher) in biomass char as also seen in low-rank coals. In addition, nitrogen is preferentially retained compared to carbon in the chars of switchgrass and pine. Table 11 gives the fraction of each inorganic element as a function of char conversion. During devolatilization, all metals are retained within the biomass char. After devolatilization, potassium and sodium vaporize. Although sodium has higher boiling point than potassium, vaporization of sodium is more pronounced. This may be due to the

Table 9 presents a comparison of kinetics data obtained for some coal and biomass fuels. Activation energies and pre-exponential constants vary considerably depending upon the pyrolysis conditions and fuel type. In biomass pyrolysis, distinct temperature ranges with different kinetics are found. In addition, the activation energies of coals are considerably higher than those for biomass fuels. Sweeten et al. [25] performed thermogravimetric analysis (TGA) of feedlot manure. The results for a heating rate of 80 C/min are shown in Fig. 9. Due to the high VM contents of manure and lower activation energy, the pyrolysis temperature was much lower than in the case of coal. Drying (outgassing of water vapor) occurred between 50 and 100 C. Pyrolysis started (i.e 1% mass loss occurs) at 185200 C and the minimum ignition temperature was approximately 528 C. Pan et al. [61] carried out TGA experiments to study the pyrolytic behavior of pine char and blends of pine with low-grade coals. Fig. 10 shows the results for pine chips. Most of the devolatilization ( 82%) was achieved within 225350 s and in a temperature range of 360560 C. For comparison purposes, Fig. 11 shows typical TGA results for a bituminous coal. When compared with coal, it is clear that biomass pyrolysis starts at lower temperatures, and the percent weight loss is higher in biomass due to the higher VM content.

188

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Fig. 9. TGA of feedlot manure at a heating rate of 10 C/min [25].

fact that potassium, being more electropositive than sodium, is capable of forming intercalation compounds with carbons. The intercalation can prevent vaporization. Reaction of sodium with SiO2 to form sodium silicates is also likely because silica is abundant in biomass. This reaction is facilitated at particle temperatures above 1900 K (melting point of SiO2) and biomass particles can exceed this temperature. Compared to the alkali metals (K, Na), the di- and trivalent metals (Mg, Ca and Al) are retained at higher levels. Conversion to silicates, coalescence, and sintering may account for the fairly high percentage of these metals in biomass char after devolatilization [64]. Fig. 12 shows the heterogeneous nature of biomass char. Particle size, shape and texture vary widely and upon combustion the aspect ratio decreases and structures become more lace-like. Fig. 13 shows images of pine and switchgrass chars before and after combustion. Before combustion, there is no crystalline order (Fig. 13(a)). Short-range order

develops during devolatilization but after devolatilization, very little additional ordering (graphitization) takes place, even at the highest levels of conversion. Graphitization depends on the ability of carbon crystallites to align and coalesce. Mobility is enhanced by hydrogen whereas oxygen hinders mobility by developing highly cross-linked rigid carbon structures. The biomass chars contain high levels of oxygen and low levels of hydrogen compared to coal. Hence, graphitic structures do not develop in biomass chars as they do in bituminous coal chars, which contain lower oxygen levels. The structural disorder may also lead to higher reactivities of biomass in the late stages of combustion since more edge carbon (which is more reactive) is available [64]. Experiments were conducted by Wornat et al. [65] to examine biomass char reactivity. Fig. 14 shows the particle temperature as a function of particle diameter for pine char, lignite coal and bituminous coal. Compared with the coal

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

189

switchgrass char as a function of residence time. As the residence time increases from 47 to 95 ms, the mean particle temperature decreases and particle distribution narrows. The changes may be due to the preferential depletion of carbon and the physiochemical transformations during combustion in biomass chars. As combustion proceeds, carbon is consumed from the biomass/ char particles leaving non-combustible ash-rich particles which causes the mean temperature to decrease. Combustion is also accompanied by preferential loss of catalytic elements such as K and Ca. These transformations would have a large impact on the reactivity of biomass chars at later times. Nevertheless, biomass chars are quite reactive in the early stages of char conversion and burn almost under diffusion control. Biomass char burning rates are comparable to burning rates of high-volatile matter bituminous coal chars. 3.4. Fouling issues in biomass combustion
Fig. 10. TGA of pine chips at a heating rate of 100 C/min [61].

particles, the pine particles burn over a wider range of temperatures. This is an indication of the heterogeneity of the biomass particles. Also shown are the temperatures under chemical diffusion limited combustion (with chemical heat contribution) and inert condition (without chemical heat contribution). Note that the biomass char temperatures span the limits between diffusion-limited combustion and inert combustion. Fig. 15 shows the particle temperatures and sizes for

Fouling of combustor surfaces is a major issue that has played an important role in the design and operation of combustion equipment. Slagging and fouling reduces heat transfer and causes corrosion and erosion problems, which reduce the lifetime of the equipment. The main contributions to fouling come from the inorganic material in the fuel. The behavior of these inorganics is less well understood than that of organic materials. Because biomass fuels contain a larger variety of inorganic materials compared to coal, issues of fouling, corrosion and pollutant emissions need to be explored. This is particularly true for some agricultural residues and new tree growth where the
Table 11 Inorganic element retention in the biomass chars [64] Char conversion (% DAF) Normalized fractional retention in the char

Table 10 Organic element retention in the biomass chars [64] Char conversion (% DAF) Mass a Normalized fractional retention in the char C Southern pine 0 52.8 b 73.0 b 86.4 b 94.6 Switchgrass 0 48.0 76.3 90.7 93.9
a

Na H O N Southern pine 0 52.8 b 73.0 b 86.4 b 94.6 Switchgrass 0 48.0 76.3 90.7 93.9 1 1 0.455 0.449 0.313 1 0.853 0.591 0.429 0.058

Mg

Al a

Ca

1 0.472 0.270 0.136 0.054 1 0.520 0.237 0.093 0.061

1 0.562 0.314 0.154 0.053 1 0.626 0.225 0.104 0.065

1 0.082 0.054 0.020 0.010 1 0.146 0.044 0.019 0.018

1 0.157 0.126 0.088 0.049 1 0.237 0.118 0.066 0.060

1 0.694 0.532 0.394 0.172 1 0.582 0.349 0.222 0.126

1 0.980 0.931 0.917 0.827 1 1 0.911 0.738 0.697

1 0.762 0.692 0.729 0.878 1 1 0.978 0.947 0.751

1 0.982 0.887 0.599 0.468 1 0.946 0.721 0.493 0.441

1 0.997 0.962 0.973 0.789 1 1 0.920 0.792 0.678

Normalized char mass on a DAF basis. b These samples produced at 6% O2; all other samples produced at 12% O2.

a Al values more uncertain for pine char, due to low absolute Al levels. b These samples produced at 6% O2; all other samples produced at 12% O2.

190

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Fig. 11. TGA of bituminous coal at several temperature levels [95] Particle size: 460 mm. A, B, C and D: temperature levels (250, 400, 450 and 500 C).

ash can have relatively high alkaline metal contents, particularly sodium and potassium [4]. Sodium and potassium lower the melting point of ash and, hence can increase ash deposition and fouling of boiler tubes. Baxter [66] addressed ash deposition and corrosion problems during coal and biomass combustion. He developed a mechanistic model to describe ash deposition in solid fuel combustors and postulated characteristics of ash deposits in biomass combustion. The major mechanisms of ash deposition were related to the types of inorganic material in the fuel blend and the combustion conditions. Ash deposition properties such as tenacity, emissivity, thermal conductivity

and morphology were discussed in relation to fuel characteristics and operating conditions. The theoretical predictions were validated by experiments carried out with a variety of coal types. Baxter concluded that the ash deposition rate in biomass combustion would peak at early times and then decrease monotonically. As compared to deposits from coal combustion, the tenacity and the strength of the biomass combustion deposits will be higher, with smooth deposit surfaces and little deposit porosity. This means that the deposits from biomass combustion may be hard to remove and may require additional cleaning effort.

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

191

Fig. 12. Scanning electron micrographs (magnication 100) of switchgrass chars: (a) uncombusted char; and (b) sample removed from reactor after burning in 12% O2 (mole basis) for 95 ms (90.7% conversion DAF) [64].

3.5. Summary comparison of coal and biomass combustion A comparison of pyrolysis, ignition and combustion of coal and biomass particles reveals the following: 1. Pyrolysis starts earlier for biomass fuels compared to coal fuels. 2. The VM content of biomass is higher compared to that of coal. 3. The specic heating value of volatiles in kJ per kg is lower for biomass fuels compared to those from coal fuel. 4. The fractional heat contribution by volatiles in biomass is of the order of 70% compared to 36% for coal. 5. Biomass char has more oxygen compared to coal. 6. Pyrolysis of biomass chars mostly releases CO, CO2 and H2O. 7. Biomass fuels have ash that is more alkaline in nature, which may aggravate the fouling problems. 4. Co-ring of blends In the combustion applications, biomass has been red directly either alone (as a sole source fuel) or along with a primary fuel (co-ring). Various technologies that utilize animal-based biomass as an energy source are summarized in Sweeten et al. [25] and Annamalai et al. [26]. These include on-site gasication [6770], uidized bed combustion [23,25,26] and circulating uidized bed combustion [71,72]. Some of the biomass technologies

have met with limited technical success. The limitations were primarily due to relying on biomass as the sole source of fuel, despite the highly variable properties of biomass. The high moisture and ash contents in biomass fuels can cause ignition and combustion problems (Fig. 8). The melting point of the dissolved ash can also be low which causes fouling and slagging problems. Because of the lower heating values of biomass accompanied by ame stability problems, the limited need for new electrical capacity in most of the US and the relatively low capital investment required for implementation, co-ring currently holds more appeal than any of the sole source technologies including more advanced conversion options such as integrated gasication combined cycles [7,73]. It is anticipated that blending biomass with higher-quality coal will reduce ame stability problems, as well as minimize corrosion effects. The co-ring approach will also have high potential for commercialization. The synergetic effects of blending coal and biomass may also lead to reductions in other pollutant emissions. For example, HC are known to react with NOx and produce molecular N2. By injecting coal beyond the combustion zone as a reburn fuel, the HC released from volatiles can be used to reduce NOx. The higher the VM content, the larger the reduction in NOx [74]. While coal contains 4050% VM, biomass contain up to 80% VM on a DAF basis. Hence, biomass has the potential to be a very effective reburn fuel when coal is used as the primary fuel. Another possible advantage of biomass blend combustion stems from the potential catalytic reduction of NOx by NH3 found in the

192

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

biomass. For example, NH3 is naturally present in animal waste. 4.1. Blend combustion efciency Since biomass fuels have higher VM compared to coal, the combustion efciency is typically limited by the extent of char combustion. If it is assumed that similar fractions of coal and biomass char, h ch,b, are completely burnt (and since volatiles are completely burnt), then overall combustion efciency of the blend can be given as: VMb Yb hblend hcoal VMc 1 Yb

Although biomass fuels have lower heating value, rapid release of volatiles may locally increase temperature near the burner compared to coal-only case. Such an increase in temperature may increase production of thermal NOx. Hence, changes in NOx emissions from ring coal only may not be simply proportional to the decrease in total nitrogen input in the blend [76]. 4.2. Classes of co-ring In order to present a discussion of coal biomass blend VMc 1 VMb 9

hch;b 1 VMb Yb 1 VMc hch;b 1 VMc

where, Yb is the mass fraction of the biomass in the blended fuel. Assuming VMb 0:8; VMc 0:4 and hch;b 0:67; we get the following linear relationship.

combustion, three classes of co-ring are dened. The class of co-ring depends upon the feeding method used for the coal and biomass fuels. (I) Separate feed lines and separate burners for coal and biomass fuels (see Fig. 16) [10]. (II) Separate feed lines and a common burner: (a) two inlets coal in the primary air and biomass in the swirling secondary air (or vice versa) (see Fig. 20) [8]. (b) three inlets two for primary air (central and annular), one for swirling secondary air [15]. Coal and biomass ow arrangements are shown in Fig. 18(a). (III) Common feed lines and a common burner with premixed coal biomass blends [7,11,77]. Class I co-ring has the advantage of better control over fuel ow rates. Thermal output similar to coal-only ring requires higher biomass feed rates. Thus separate feeders facilitate controlling the biomass feed rate independent of the coal feed rate. Using a single feed line for a blend fuel has the risk of agglomeration occurring in the supply line, which may lead to a disconnection or blockage in the fuel supply. On the other hand, separate feed lines and separate burners increase capital and maintenance costs. Firing low heating value biomass independently of coal also has a risk of poor combustion efciency. Class II co-ring is relatively inexpensive in the sense that a single swirl burner can be used to re the blend. The coal and biomass are fed separately, as in Class I coring. At the burner entrance, the two fuel streams are unmixed. In the quarl region of the burner, the two streams mix due to the swirling action of the secondary air. When good mixing is obtained, higher combustion efciencies and lower emissions result. However, if a swirler is used, feeding one of the pulverized fuels in the swirling secondary air is likely to cause damage to swirler blades and this issue needs to be addressed. The swirling fuel stream can be avoided using Class IIb ring (see Fig. 18) or introducing secondary air at an angle to achieve the swirl (see Fig. 20).

hblend =hcoal 1

0:16458Yb

10

Our recent detailed turbulent combustion modeling of blends conrms such a relationship [75].

Fig. 13. High-resolution transmission electron micrographs of carbon rich portions of biomass chars: (a) uncombusted pine char; (b) pine char sample at 53% conversion DAF; (c) uncombusted switchgrass; and (d) switchgrass char at 48% conversion DAF [65].

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

193

Fig. 14. Optical measurements of temperatures and sizes of single particles burning at 72 ms in 12% O2 (mole basis) for: (a) southern pine char; (b) Beluah lignite coal; and (c) Pittsburgh #8 bituminous coal. Particle diameters were determined using the area-averaging technique. Tg denotes mean gas temperature. D and I correspond to diffusion and inert ignition limits for spherical particles of equivalent diameter [65].

Damage to the swirl blades can also be avoided by pre-mixing the two fuels before they enter the burner (Class III ring). Class III ring is the least expensive method since no separate feed lines and burners are required and the existing fuel lines can be used. The Class III coring method provides good mixing, high combustion efciencies and low emissions. However, the risk of burner problems associated with feeding difculties and maintaining similar thermal input are obvious concerns for the boiler operators. In order to have the same heat throughput rates when co-ring, the following relationship is obtained. mblend mc HHVc =HHVblend 11

Typically, a 90:10 coal:sawdust blend will have a heating value about 3% less compared to coal. Using Eq. (10), the blend feed rate should be increased by about 3% to have the same heat throughput. The co-ring can be through suspension or stoker boilers. In addition, uidized bed combustion of coal/biomass blends falls under this category of co-ring. In the following sections, specic results of various coring studies are presented. Due to the broad range of biomass fuels studied, the discussion has been grouped by the type of biomass fuel. 4.3. Coal and agricultural residues Sampson et al. [11] reported results for co-ring wood chips mixed with coal (Class III) at a stoker-red steam plant. The boilers were mechanically red by spreader stokers with travelling grates and a y ash reinjection system. Three different types of wood chip were used

(higher heating values ranging from 19,350 to 19,690 kJ/ kg (83208420 Btu/lb)). The higher heating value of the coal was 25,080 kJ/kg (10,600 Btu/lb). However, the investigators did not provide the size distribution of coal and wood chips used. They red 1022% by dry mass basis of wood chips in the fuel mixtures and found a negligible effect on particulate emissions. Table 12 shows that the fuel blend particulate emissions ranged from 0.06 to 0.1 grains per SCF (0.00390.0065 gm/SCF). No data on NOx and SOx emissions were provided. Due to bunker and stoker capacity problems in the feeder system, they could not re 30% biomass in the fuel blend. The capacity problems and the particulate emissions can be reduced if the moisture content of the wood chips is reduced. An economic study, conducted for the 55,555 kg/h (125,000 lb/h) of steam power plant, concluded that energy derived from wood would be competitive with that from coal if more than 30 kton of wood chips were produced per year and hauling distances were less than 60 mile. Aerts et al. [10] carried out experiments on co-ring switchgrass with coal (Class I ring) in a 50 MW, radiant, wall-red, pulverized coal boiler with a capacity of 180 ton of steam at 85 bar and 510 C (see Fig. 16). Table 13 lists the ultimate analysis of the coal and switchgrass used. Note that switchgrass contains 60% less nitrogen than coal on mass basis. However, on heat basis, nitrogen content in switchgrass is only about 13% less than coal. Five co-red and ve coal-only combustion tests at switchgrass feed rates up to 3365 kg/h (10% heat input at 40 MW) were conducted for 4 h durations. The cumulative particle size distributions for coal and switchgrass are shown in Fig. 17. The switchgrass

194

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Table 12 Wood chip characteristics, steam load and particulate emissions [11] Date Fuel Chip MC (% wet basis) Wood as percent of total Weight (%) 6/2/87 6/2/87 6/2/87 6/3/87 6/3/87 6/3/87 6/4/87 6/4/87 6/4/87 6/5/87 6/5/87 6/5/87 6/8/87 6/8/87 6/8/87
a b

Average steam load (1000 lb/h)

Particulate emissions a

Btus 7.3 7.3 7.3 8.9 8.9 8.9 15.2 15.2 15.2 8.3 8.3 8.3 125 123 130 124 123 124 125 125 123 106 120 123 120 121 125 0.0730 0.0845 0.0466 0.0593 0.0665 0.0637 0.0781 0.0678 0.0604 0.0630 0.0985 0.0743 0.0742 0.0748 0.0850

Coal Coal Coal Aspenspruce b Aspenspruce Aspenspruce Aspenspruce Aspenspruce Aspenspruce Sprucebirch c Sprucebirch Sprucebirch Sprucebirch Sprucebirch Sprucebirch

34.7 34.7 34.7 35.3 35.3 35.3 41.6 41.6 41.6 40.4 40.4 40.4

9.9 9.9 9.9 12.1 12.1 12.1 21.9 21.9 21.9 12.1 12.1 12.1

Rate is in grains per dry standard cubic foot, corrected to 12% CO2. Aspenspruce was 75% aspen and 25% white spruce by volume and was chipped from green timber. c Sprucebirch was 75% white spruce and 25% paper birch by volume and was chipped from re-killed timber that had been standing dead for 3 years.

particles were signicantly larger than the coal particles, yet there were no adverse effects in co-ring with switchgrass. Unit operation was normal and slagging was not observed during co-ring. Table 14 provides SO2 emissions for coal and coal:switchgrass blend combustion. The SO2 levels, in general, did not change appreciably except at low load when the SO2 jumped to almost 1200 ppm. No explanation was provided. The NOx emissions decreased by about 20%, which most likely was due to low nitrogen content of the switchgrass (see Table 15). Some partially burned switchgrass nodes were observed in the bottom ash. The amount of unburned carbon in the cyclone and precipitator ash was

about 5% with co-ring, which was similar to coal-only operation. Such observations seem to suggest that the large particle size and lower heating value of the biomass fuel did not adversely affect combustor performance, probably due to the higher VM content of the biomass fuel. The VM burns rapidly and the higher VM content of the biomass can also result in a highly porous char, thus accelerating the char combustion as well. Fahlstedt et al. [12] carried out a series of tests on coring wood chips, olive pits and palm nut shells with coal at the ABB carbon, 1 MW uidized bed facility. Table 16 shows the results of the tests in terms of combustion efciency, heat output and system temperatures. It is interesting to note that the blend combustion had a slightly higher

Table 13 Analysis of coal and switchgrass [10] Coal (Kindill) As-received Dry HHV (kJ/kg) 25,498 29,826 Switchgrass As-received Dry 15,991 42.02 4.97 35.44 0.77 0.18 0.03 4.61 11.99 18,171 47.74 5.64 40.26 0.87 0.21 0.04 5.24 0.00 Coal Coal Coal Coal Co-re Co-re Co-re Co-re Co-re Table 14 Sulfur dioxide emissions for coal switchgrass blends (corrected to 3% excess O2) [10] Load (MW) 40.2 40.7 48.8 49.2 39.9 40.1 43.8 46.4 47.3 SO2 (ppm) 853 781 912 810 1198 910 949 896 930 SO2 (g/MJ) 1.04 0.94 1.07 0.96 1.31 0.98 1.00 0.98 1.04

Ultimate analysis (percent by weight) Carbon 62.97 73.66 Hydrogen 3.73 4.36 Oxygen 7.26 8.49 Nitrogen 1.36 1.59 Sulfur 1.34 1.57 Chlorine 0.02 0.02 Ash 8.80 10.30 Moisture 14.51 0.00

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214 Table 15 Emissions of nitrogen oxide for coal and switchgrass blends (corrected to 3% excess O2) [10] Load (MW) Coal Coal Coal Coal Co-re Co-re Co-re Co-re Co-re 40.2 40.7 48.8 49.2 39.9 40.1 43.8 46.4 47.3 NO (ppm) 385 386 450 422 395 313 355 350 377 NO (g/MJ) 0.22 0.22 0.24 0.23 0.20 0.16 0.18 0.18 0.20

195

content (coal in this example) should be injected into the fuel rich zone in order to reduce NOx emissions. This also explains the low NOx emissions at lower biomass loading and higher NOx at higher loading because the % primary air increases with reduced coal loading. Andries et al. [16] co-red straw with coal in a 1.6 MWthermal pressurized FBC test rig (Class III ring). Because of the high VM of the biomass, the temperature downstream of the free board (the surface of the uidized bed) was higher than the coal-only case by about 30 K. The location of the high-temperature region corresponds to the location of volatile combustion. Co-ring reduced the CO, NOx and SO2 concentrations in the free board. 4.4. Coal and RDF

efciency than coal-only combustion. The reason, as mentioned above, is likely due to the higher VM content of the biomass fuels. Increasing the wood chips co-ring ratio from 20 to 40% by mass resulted in a decrease in NOx of about 25%. The NOx levels were also low with other biomass fuels when compared with the coal-only case, probably again due to the lower nitrogen content in woody biomass fuel. No fouling was observed in general but in the case of palm nut shells there was some indication of an oxide layer on the bed surfaces which requires further investigation. Similar results were presented by Siegel et al. [15] who co-red straw and cereal with hard coal in a 500 kW pulverized-fuel test unit (Classes III and IIb co-ring, see Fig. 18(a)). Fig. 18(b) shows the effects of the different fuel injection schemes on NOx emissions. At higher biomass thermal loading ( 60%), it is obvious that injecting coal in the annular pipe results in a decrease in NOx, whereas using a central coal jet causes an increase in the emissions as also reported by Abbas et al. [8]. However, at lower biomass thermal loading ( 40%), injection of coal in the central jet has lower emissions. The burnout was almost 99% for a co-ring ratio of up to 60% (by heating value) for a variety of cereals and straw tested. However, above this co-ring ratio, a substantial drop in the combustion efciency occurs. The authors concluded that the fuel with higher nitrogen
Table 16 Results of coal biomass uidized bed pilot plant testing [12] Wood chips 20% Combustion efciency (%) Bed density (kg/m 3) In-bed heat transfer, (kW) NOx (mg/MJ) Bed temperature ( C) Freeboard temperature ( C) Cyclone temperature, ( C)
a b

A dual fuel burner designed for the co-ring of wastederived solid fuel with pulverized coal (Class IIa ring) was evaluated by Abbas et al. [8]. Figs. 19 and 20 are schematics of the burner facility and the burner, respectively. The inuence of fuel injection mode and co-ring ratio on combustion aerodynamics, ame stability and NOx emissions were studied. The biomass fuel was predominantly sawdust, although some sewage sludge data were also reported. The authors found that combustion efciency increased and NOx emissions decreased when saw dust particles were injected in the primary air with a swirling annular stream of coal particles surrounding the biomass particles (swirling primary fuel, SPF mode). They reasoned that the central sawdust stream pushed the coal volatiles into an oxygenlean zone, thereby reducing the initial amount of NO formed. The opposite results were obtained when the coal was injected through the center with an annular swirling sawdust stream (central primary fuel, CPF mode). The experimental results suggested an optimum co-ring ratio that resulted in maximum particle burnout and minimum NOx emissions. The optimum was achieved when sawdust provided 30% of the total heat input and the injection mode was SPF. They also compared results on 15% co-ring of sewage sludge in both the SPF and CPF modes and found the SPF mode most suitable for co-ring. The results on

Wood chips 40% 97.2 1167 418 57 840 779 731

Olive pips 24% 97.9 1244 439 90 843 794 743

Palm nut shells 24% 98.1 1181 420 101 840 794 752

Polish coal 100% 97.1


a a

97.6 1186 443 77 839 785 739

141 b 840 784 739

Too short a test for stable value. Too high excess air.

196

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Fig. 15. Optical measurements of temperatures and sizes of single particles of switchgrass char burning in 12% O2 (mole basis), initial particle size 75106 mm, particle residence times: (a) 47 ms; (b) 72 ms; and (c) 95 ms. Particle diameters were determined using the coded aperture technique. Tg denotes mean gas temperature [65].

sawdust and sewage sludge emphasize the need to consider the reactivity and the nitrogen content of all the fuels before selecting the fuel injection mode. Van Doorn et al. [9] red coal and wood, straw and municipal sewage sludge in a uidized bed combustor (Class III ring). They found wood to be the most favorable co-ring fuel in terms of ease of combustion and reduced

emissions of NOx and SO2. No agglomeration of fuel particles was observed. The emissions of SO2, CO and NOx decreased with increasing wood to coal ratio. In the case of co-ring straw, similar effects were observed. However, the HCl concentration increased with larger straw to coal ratios due to the relatively higher chlorine content of straw. Co-ring sewage sludge with coal caused agglomeration of

Fig. 16. Alternate fuel handling facility for biomass fuel at Blount St. Generating Station [10].

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

197

Fig. 17. Particle size distribution for coal and switchgrass [10].

fuel particles and high emissions. Similar tests on co-ring straw with coal were conducted at two pulverized coal red power plants by Christensen and Jespersen [14]. They found a reduction of 20% in NOx with 22% (by mass) co-ring ratio. The decrease is again attributed to the lower nitrogen content of straw. The authors also estimated that the corrosion rate with co-ring would be twice as high as that for coal alone. Ohlsson [17] carried out co-ring tests in a 440 MW cyclone red combustor and measured emission levels of

SO2 and NOx, among other pollutants (Class III co-ring). The RDF used was binder-enhanced, densied municipal solid waste (MSW). They red a blend of 12% RDF and 88% coal (mass basis) over a 10 h test period. The reduction in NOx emissions was 23% and was attributed to the low nitrogen content of the RDF. The sulfur dioxide emissions were 17% lower than the coal-only case. The lower sulfur content of the RDF and more importantly the heterogeneous reaction of SO2 with the binder (calcium hydroxide) in RDF helped in lowering the SO2 emissions. The particulate concentration was higher with blend ring as the RDF contained more ash than coal. On heat basis, the particulate concentration was about 1.5 times more than the coal-only case. Armesto et al. [71] investigated co-ring of coal and pine chips in a circulating uidized bed (CFB) and a uidized bed combustor (FBC). Operation of the combustors was normal and they did not encounter difculties. The CFB process had higher combustion efciencies than the FBC process and consequently lower CO emissions. The NOx and SOx emissions were also reduced in the two systems. In both cases, the author estimated an increase in combustion efciency with an increase in co-ring ratio. A sharp decrease in SOx emissions was reported by Rasmussen and Clausen [72], who co-red straw in an

Fig. 18. (a) Multi-fuel burner conguration for co-ring cereal and straw with coal [15]; and (b) NOx emissions for different burner congurations [15].

198

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Fig. 19. Experimental facility for co-ring studies of coal/waste derived fuels [8].

Fig. 20. Schematic of dual-fuel burner which can be operated in two modes. APF mode is obtained when fuel 1 is sawdust and fuel 2 is coal; CPF mode is obtained when fuel 1 is coal and fuel 2 is sawdust [8].

80 MWthermal CFB. SOx levels decreased substantially with increasing straw input because straw has lower sulfur content compared to coal. Due to decreasing temperature at higher co-ring ratios, the NOx emissions remained almost constant. Particulate emissions were below the detection limit. Brouwer et al. [76] carried out studies on emission reductions while ring coal blended with refuse-derived fuel (Class III co-ring). The RDF used was hard and softwood waste from manufacturing and chipped railroad ties. Two facilities were used: a spreader stoker-red boiler and a pulverized coal boiler. The stoker system was a pilot-scale facility with a ring rate capacity of 500,000 1,000,000 Btu/h. The pulverized coal facility was a 38 kW research facility. Two methods of ring biomass with coal in the pulverized coal facility were used. In the rst method, the biomass was pre-mixed with coal and injected through the main line. In the second method, the biomass was injected after the recirculation zone as a reburn fuel. In the stoker tests, railroad ties were red at 20% mass basis in the blend. As shown in Fig. 21, the NOx emissions were lowered by 25% with co-ring under clean conditions (excess air 50% and CO emissions 20 ppm). The emissions reduction was ascribed to the lower nitrogen content of the railroad ties (0.22% nitrogen). At low excess air levels, the CO emissions were considerably lower for blend combustion suggesting increased burnout with

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

199

0.8 0.7
NOx (lbs / MBtu)

0.6 0.5 0.4 0.3 0.2 0.1 0 0 20 40 60 80


Excess Air (%)
NO (coal) NO (coal+railroad ties) CO (coal) CO (coal+railroad ties)

500 450 400 350 300 250 200 150 100 50 0 100

Fig. 21. Effects of co-ring 20% (mass basis) railroad ties with coal on NOx and CO emissions [76].

blended fuels. Note that this reduced CO level corresponds to increased NOx. This is also evident by looking at Fig. 21. At lower excess air levels ( 20%), there is almost no difference in NOx between coal and the blend even though the blend has lower nitrogen content. The NOx emissions were further reduced with the injection of natural gas as a reburn fuel. With 28% excess air and 15% gas injection, the NOx levels decreased from about 0.45 lb/MBtu to almost 0.25 lb/ MBtu. In the pulverized coal furnace, both co-ring and reburning of wood waste was tested. In co-ring, wood waste (100% 24 mesh) was pre-mixed with coal (70% 200 mesh) and red at a constant heating rate of 38 kW. External staging was used to study NOx emissions. The results are shown in Fig. 22. With unstaged combustion, the NOx decreased as the co-ring ratio was increased, again due to the lower nitrogen content of the biomass fuel. However, the decrease is not as much as expected from the equivalent

Fig. 22. Nitrogen oxide emissions as a function of percentage of wood in the blend (on heat basis) [76].

reductions in fuel nitrogen. For example, 600 ppm of NOx was measured when coal, which contained 1.6% nitrogen (mass basis), was red alone. With 50% co-ring (by heating value) of wood waste (0.2% nitrogen by mass), the measured NOx was 500 ppm. Had there been a proportional decrease in NOx with the reduction in fuel nitrogen (0.9% nitrogen), the NO levels would have been 340 ppm. On the other hand, with staged combustion (reduction in secondary air at the burner and subsequent injection of air and fuel downstream of the burner), signicant reduction in NOx did not occur until the co-ring ratio was greater than 50%. This indicates that co-ring wood waste with coal may not lead to reductions in NOx emissions in a low-NOx conguration unless large co-ring ratios are used. In reburning, wood waste was not co-red but separately injected after the recirculation zone to reduce NOx. Lower reburn stoichiometric ratios led to increased wood reburning and NOx reduction. The wood waste reduced NOx just as effectively as coal or natural gas at a temperature of 1721 K. All fuels performed poorly when the NOx levels in the reburn zone were low (500200 ppm). For 200 ppm of NOx, wood and natural gas performed much better than coal as a reburn fuel, primarily due to the large amounts of VM provided by wood and natural gas. It can be concluded that co-ring in stoker system resulted in NOx reduction. Reburning with natural gas further decreases the emission level. In the pulverized coal biomass blend ring, NOx reductions were observed but they were not in proportion to the amount of wood combusted. Substantial reduction in NOx was achieved ( 60%) while injecting the biomass as a reburn fuel. At high temperatures, biomass works even better than coal as a reburn fuel to reduce NOx emissions. Reburn stoichiometry is the most important parameter that determines the effectiveness of reburning with the waste biomass fuels with optimal stoichiometric ratios around 0.85. Foster Wheeler Environmental and Reaction Engineering

CO (ppm)

200

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

International co-red wood-derived fuel with coal at existing facilities within the Tennessee Valley Authority (TVA) power generation system [5]. Four types of plant were evaluated: (1) a cyclone coal-red power plant; (2) a wall-red pulverized coal power plant; (3) a tangentially red pulverized coal power plant; and (4) a gas or oil-red power plant. Results of co-ring tests at the tangentially red PC unit were as follows: 1. The ame temperature decreases by about 40 K during co-ring. This suggests that there will be a small reduction in thermal NOx during co-ring. 2. 1015% wood co-ring (heating value basis) results in a decrease in boiler efciency of less than 1.5% relative to coal-only ring. 3. There is a signicant reduction in SO2 and NOx emissions primarily due to reducing the total amount of sulfur and nitrogen in the fuel blend. The Electric Power Research Institute (EPRI) initiated a program in 1992 to commercialize co-ring in utility stations. Hughes and Tillman [3] provided a detailed account of the co-ring tests carried out in full-scale power plants, pilot plants and laboratory scale facilities. At the Allen Fossil Plant of TVA, biomass/coal blends of up to 20% wood have been burned in cyclone boilers. Tires have also been burned in the fuel blend. The co-ring also conrmed a reduction in CO2, SO2 and NOx. The study also established that there is a trade off between boiler efciency and fuel costs. At the Kingston and Colbert power plants, 5% woodderived biofuel was co-red (Class III co-ring) in tangentially red and wall-red PC boilers, respectively. Due to pulverizer performance and fuel particle size, 5% co-ring was found to be the limiting case. Other co-ring tests performed in utility boilers and pilot plants are summarized in Tables 17 and 18. The plant capacity was maintained in all the co-ring tests. The results of these tests can be summarized as follows. 1. Co-ring can be performed at moderate and high percentages in cyclone boilers. 2. Co-ring can be performed at low percentages (by mass, 05%) in PC boilers. The co-ring ratio depends on the pulverizer performance which in turn depends on the type of biomass fuel used. Most biomass fuels have brous structure (e.g. wood, switchgrass) which are difcult to ground to the same sizes as coal used in coal pulverizers. Grinding costs determine the extent to which biomass can be economically pulverized. 3. Co-ring (510%, mass basis) in PC boilers may require separate fuel feed lines depending upon the capacity of existing pulverizers, type and condition of biomass fuels and the type of pulverizers used. 4. The potential for successful application of co-ring is site-specic. It depends upon the characteristics of

power plant being considered, the availability and price of biofuel within 50100 mile of the plant and the economic value of environmental benets. In 1996, DOEs Federal Energy Technology Center (FETC), became a co-funder of the overall biomass co-ring research program conducted by EPRI. After several successful parametric tests done at utility boilers [5], the DOEs Ofce of Energy Efciency and Renewable Energy, Biomass Power Program joined as a co-funder in 1998 to participate in several long-term demonstration tests. Coring at the NIPSCO Bailly Station, the GPU Seward Station and the Allen Station will demonstrate the longterm impacts and benets of biomass co-ring. The European Commission launched a 2 year project (APAS) in 1993 on co-ring biomass in laboratory, pilot and full-scale units. Twenty-ve partners from eight European countries participated in the project. Hein and Bemtgen [2] summarized the activities undertaken during this project. A description of participants and facilities is given as Table 19. Two general types of biomass fuel were considered: woody (wood, straw, paper, Miscanthus) and sewage sludge. Table 19(a) describes the woody biomass studies and Table 19(b) describes the sewage sludge studies. The combustion facilities were either PC or FBC. The results derived from these studies were: 1. Both modes of PC and FB are well suited for co-ring provided a fuel dependent feed and preparation system is installed. 2. No major negative effects on fuel conversion were found. 3. In order to avoid corrosion and slagging of the heat transfer surfaces, biomass fuel rich in chlorine and alkali metals should not be used in co-ring. 4. With respect to the emissions of hazardous gaseous compounds, no increases in concentrations in the ue gases were observed. In many cases, substantial emission reductions were found, which were a function of biomass composition and the fuel injection mode. The effect of emission reduction due to injection mode was particularly apparent for NOx emissions. 4.5. Coal and animal waste Frazzitta et al. [7] evaluated the performance of a smallscale boiler burner facility (see Fig. 23) while using coal and pre-mixed coal and manure blends (Class III ring). The coring ratio was 20% manure by mass. Three types of feedlot manure were used: raw (RM), partially composted (PC) and fully composted (FC). RM is dened as fresh manure just collected from the ground. When RM is stockpiled and turned in the outside air for about 30 days, it is considered PC. If the drying and turning period is greater than 120 days, the manure is considered FC. The authors measured the temperature distribution in the burner, emissions and composition of ue gases and

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214 Table 17 Co-ring tests performed at full-scale utility boilers [3] Descriptive title and approximate date Organizations Performers Commercial operations coring sander dust with coal in a cyclone boiler at NSP (1987 present) Co-ring forest debris from Hurricane Hugo in a pulverized coal boiler at Santee Cooper (1990) Commercial coal/biomass uidized-bed combustion at Tacoma (1991present) Waste wood co-ring tests at Plant Hammond, pulverized coal boiler of Georgia Power (1992) Tire-derived fuel co-ring test in a wall-red grate-equipped PC boiler (1992) Conring of low percentage of wood at the Colbert wall-red PC boiler of TVA (1992) Co-ring of low percentage of wood at the Kingston tangential-red PC Boiler of TVA (1993, 1994) Plastic/ber waste co-red in a PC boiler (1993) High-percentage wood co-ring in a pulverized coal boiler at Savannah Electric (1993) Wood co-ring up to 20% by mass in a Cyclone boiler at TVA (August 1994) Mid-percentage (10% by heat) co-ring in a pulverized coal boiler at NYSEG (1994) Wood and tire triring with coal up to 15% by mass at Cyclone boiler at TVA (August 1995) Wood co-ring up to 20% by mass in a Cyclone boiler at TVA (December 1995) Wood preparation (sawdust, right-of-way and poplar) and co-ring in a PC boiler (1995) Switchgrass co-ring in a wallred, grate-equipped PC boiler in Madison, WI (1996) Preliminary test of plastics, mill residues co-red in a PC boiler (1996) Northern States Power Funders Northern States Power Power ( 1990). EPRI conference 1992 and 1993 Report reference

201

Santee Cooper Electric Coop. (South Carolina)

Santee Cooper Electric Coop. (South Carolina)

Power (1993), SERBEP (1992)

Tacoma Public Utilities

Tacoma Public Utilities

Power ( 1993)

Georgia Power and Southern Company Services (SCS)

SCS and Georgia Power

EPRI Conference 1993 in Washington, DC (TR-103146, 12/93) EPRI TR-103851, 12/94

City of Ames, Iowa State University TVA

South Carolina E&G, Penelec: Centerior TVA

EPRI Meeting 12/93

TVA, Foster Wheeler

TVA

EPRI Meetings 12/93 and 11/94

South Carolina E&G Savannah Electric and SCS

South Carolina E&G Savannah Electric and SCS

South Carolina E&G 1994 Power (1995)

TVA, Foster Wheeler

TVA/EPRI

Foster Wheeler 12/94, EPRI Gray Cover 7/96 NYSERDA Report No. 96-01 (January 1996) EPRI Gray Cover 5/96

NYSEG

NYSERDA, NYSEG, ESEERCO TVA/EPRI

TVA, Foster Wheeler

TVA, Foster Wheeler

TVA/EPR

Foster Wheeler Report and Paper 5/96 EPRI Gray Cover 7/96

GPU/Penelec, Foster Wheeler

State of PA, DOE/PETC, EPRI, GPU/Genco EPRI, DOEs Great Lakes Reg. Biomass Prog., MG&E, others Duke Power

Madison Gas and Electric, University of Wisconsin Duke Power

EPRI Meeting 6/96

EPRI Meeting 6/96

202

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Table 18 Co-ring tests performed at pilot plants and laboratory scale facilities [3] Descriptive title and approximate date Organizations Performers RDF co-ring in utility boilers and performance calculations via RDFCOAL ( 1985) RDF co-ring update (1996) Strategic analysis of biomass/ waste fuels for utilities, and BIOPOWER Calculations (1993) Wood fuel sources, transportation and cost/supply for co-ring at TVA plants (19921995) Case studies of wood co-ring concepts and costs for TVA power plants (1993) Conceptual designs and costs for co-ring wood in systems based on natural-gas red gas turbines (1993) Wood reburn for NO, control in coal-red boilers: lab tests and rough economics (1994) Wood reburn concept design/ cost for NO, control in TVA cyclone-red boiler (1995) Wood/coal blends: storage and cold-ow tests simulating bin feed for cyclone boilers (1994) Fuel characteristics of mill residues for TVA co-ring (1994) Biomass fuel resources for Indiana (1994) Designs, costs and COFIREI spreadsheet (1995) Biomass resources and power plants for potential co-ring at Union Electric (1995) Wood resources and power plants for potential co-ring projects at Pennsylvania Electric (1995) Willow energy crop and wood co-ring feasibility in New York State (1995) Grass crops and wood crops for co-ring in a coal-red power plant in Iowa (1995) Bench-scale test of switchgrass co-ring (1995) Lab drop tube test and engineering study of waste plastic co-ring for a PC boiler (1995) MRI, Iowa State University, EPRI Iowa State University Appel, SFA Pacic, EPRI Funders EPRI, DOE/Argonne EPRI CS-5754.6/88 Report reference

EPRI EPRI, NYSERDA, DOE/ SERBEP

EPRI 9/96 EPRI TR-102773, 12/93; -774, 3/95

University of Tennessee

TVA, EPRI

University of Tennessee 8/93, EPRI 8/96

Ebasco (now Foster Wheeler)

EPRI, TVA, DOE/SERBEP

EPRI Gray Cover 4/94

Ebasco (now Foster Wheeler)

EPRI, TVA, DOE/SERBEP

EPRI Gray Cover 7/96

REI, Foster Wheeler, University of Utah Foster Wheeler, REI

DOE/SERBEP, EPRI, NSF

REI Paper 1995 EPRI

8/96

TVA/EPRI

EPRI Gray Cover 5/96

REI, Foster Wheeler

TVA/EPRI

EPRI Gray Cover 7/96

Foster Wheeler, TVA

TVA/EPRI

EPRI Gray Cover 7/96

NEOS Foster Wheeler Foster Wheeler

DOE/GLRBEP TVA.EPRI.DOE Union Electric, EPRI

GLRBEP Report EPRI EPRI 8/96 9/96

Foster Wheeler

GPU/Penelec, EPRI

EPRI Gray Cover 10/95

Antares, NYSEG, NMPC, SUNY Iowa State University, IES, others DOE/PETC ABB-CE, Duke Power

DOE/NREL, EPRI, NYSEG, NMPC, others DOE/NREL, EPRI, IES, others

EPRI TR.105250 11/95. NREL also NREL (also an EPRI summary in TR-105854) PETC soon ( 8/96) EPRI Gray Cover 7/96

DOE/PETC Duke Power, EPRI, American Plastics Council

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214 Table 18 (continued) Descriptive title and approximate date Organizations Performers Biomass co-ring prospects for Northern Indiana Public Service (1996) Biomass co-ring and CO2 options for NIPSCO Wood fuel preparation at NYSEG: equipment selection and equipment tests (1996) Foster Wheeler Funders NIPSCO, EPRI EPRI soon ( 8/96) Report reference

203

Moll Associates NYSEG

NIPSCO, EPRI EPRI, NYSERDSA, ESEERCO, NYSEG

EPRI soon ( 8/96) EPRI 11/96

determined the combustion efciencies. Fuel blends were injected in a pre-heated burner with a hot secondary air stream of about 200 C. As the ring started with a constant air feed rate, the burnt fraction gradually increased with increasing gas temperatures. When the burnt fraction was about 95%, the temperature distribution with and without the manure was approximately the same suggesting that manure did not adversely affect the ame stability. The burnt fraction was determined using the exhaust gas analysis and was found to be 97% for both coal and coal/manure blends. The NOx levels decreased from 555 ppm for FC manure, to 500 ppm for RM and to 480 ppm for PC manure at 95% burnt fraction. The decrease in NOx may be due to the varying amounts of nitrogen and moisture in the three types of manure (see Table 7). On a DAF basis, FC manure has more nitrogen than PC manure. Fig. 24(a) shows the amount of nitrogen in NO normalized with nitrogen in fuel for coal and coal/manure fuel blends. The ratio is highest for coal and lowest for RM and gradually increases from raw to PC to FC manure blends. The same excess air was used in all three experiments and O2 concentration in exhaust was 3%. Fig. 24(b) shows that the NOx evolution starts even at low burnt fractions that occur during the warm-up period of the reactor. Most importantly, lower SOx emissions were measured for blended fuel. This may be due to SO2 capture by the alkaline ash of the feedlot manure, however, these results on SO2 need to be rechecked. 4.6. Combustion modeling for coal biomass blends Coal/biomass blend combustion modeling is a complex problem that involves gas and particle phases along with the effects of turbulence on the chemical reactions. For turbulence closure, k e model has been used in many combustion codes. It is relatively easy and simple to implement compared with other turbulence models. Computational time is relatively small and the results obtained are reasonably accurate. The standard k e model is modied to include the effects of particles on gas phase turbulence. However, for ows with strong swirl, the k e model does not yield good results. This is due to the assumption of

isotropic turbulence made in the model. Strong swirl ows are known to be highly non-isotropic in nature and therefore need a more elaborate modeling than k e model. Most coal combustion codes (e.g. PCGC-2) are based on mixture fraction-equilibrium chemistry approach. Fast chemistry is assumed and thus mixing of fuel and oxidizer determines the combustion process. The mixture fractions are dened as,

hi

mi i mk
k0

12

where k 0 refers to secondary air and the index i refers to primary air i 1; coal-off gas i 2 and biomass-off gas i 3: The predictions are obtained by numerical solution of the time-averaged conservation equations for the gas and particle phases. The basic idea is to de-couple the gas and particle equations using the particle source terms as tear variables. Tear variables are the terms in the gas phase equations that contain the sources of mass, momentum and energy from the coal particles. These terms make it possible to solve the gas and the particle phase equations separately. The gas phase is solved in the Eulerian domain while the particles are treated in a Lagrangian frame. Most mathematical models consist of sub-models for turbulent uid mechanics, gaseous combustion, particle dispersion, coal devolatilization, heterogeneous char reaction, pollutant formation and radiation. Existing coal combustion models are modied to include the effects of biomass co-ring on the overall combustion behavior. The problem in blend combustion is that two chemically different fuels (coal and biomass) are involved. Therefore, the off-gases from each solid fuel must be tracked separately to capture the interaction of coal and biomass combustion. There are few modeling studies on blend combustion in the literature probably due to the fact that co-ring is a developing technology still in the testing phase. Abbas et al. [8] developed a mathematical model for

204

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Table 19 Description of participants and facilities involved in the European Union biomass utilization program [2] Fluidized bed Atm. bubbling Press. bubbling Atmospheric circulating Low-rank coal High-rank coal Pulverized fuel Low-rank coal High-rank coal

(a) Woody biomass KTH 10 kW straw, wood, misc. 1100 C pyrolysis, reactivity tests, N-release ECN 0.3 MW INETI 0.3 MW pellet, biomass straw, wood 700900 C 700950 C 030% 050% CIEMAT 1 Tu Delft 1 MW RWE 1 MW straw, CIEMAT 1 MW MW wood, straw straw, Miscanthus, wood, straw 700900 C Miscanthus, 750950 C 7001000 C 030% 750950 C 030% 030% 030%

IVD 0.35 MW Miscanthus, straw 9001200 C 050% RWE 1 MW straw, Miscanthus, 9001100 C 030%

IVD 0.35 MW Miscanthus, straw 9001200 C 050% KEMA 1 MW wood 10001300 C 030% IFRF 2.5 MW w.paper, straw 10001300 C 050% ELSAM 131 MW straw 1300 C 0.25%

ELSAM 80 MW straw 700900 C 060% (b) Sewage sludge Pulverized fuel High-rank coal

VEAG 100 MW Miscanthus, wood 9001100 C 010%

IVD 0.35 MW sewage sludge 9001200 C 050% Imp. Coll. 0.5 MW straw. Wood 700950 C 050% IFRF 2.5 MW sewage sludge 10001300 C 050% NEI 88 MW sewage sludge 12001600 C 030% Saarbergwerke 150 MW sewage sludge 1500 C 015%

Fluidized bed High-rank coal DMT 0.3 MW pellet, sewage sludge 700950 C 030% ECN 0.3 MW pellet, sewage sludge 700950 C 030% Stadtw.SB 2 MW sewage sludge 700900 C 030%

Other systems

Berzelius sewage sludge lead production

Fechner 11 MW sewage sludge 700900 C 020%

blend combustion based on the k e turbulence model [78], a turbulence decay model for volatile combustion and a diffusive radiation model. A single step model was used for the pulverized fuel devolatilization, and a kinetic and diffusion

model was used for the char reactions. In order to account for the variability in the properties of the volatiles from each type of fuel, they used two mixture fractions to track coaland straw-off gases separately. The coal-off gas and straw-off

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

205

Fig. 23. Schematic of experimental facility used for coal/manure blend studies [7].

gas fractions were calculated on a staggered Cartesian mesh using a hybrid differencing scheme. The gas phase equations were based on Eulerian reference frame while the particle phase equations were modeled using Lagrangian reference frame. The authors compared the results with their experiments on co-ring coal and sawdust and found good agreement between the experiments and the model results. They predicted an earlier devolatilization of coal particles due to early ignition of sawdust volatiles in the near burner region. Similar predictions were obtained by Dhanaplan et al. [79]. The increased devolatilization rate led to lower NO formation (40% reduction) when the sawdust was introduced through the middle of the annular coal jet. A full-scale coal and straw-red utility boiler was modeled by Kaer et al. [13]. The authors used a commercial CFD code (CFX4.2) with an extended particle formulation model. The steady state three-dimensional NavierStokes equations and Lagrangian particle tracking equations were solved in a cartesian mesh. The calculations were based on physical data from a full-scale co-ring facility. The calculations incorporated a k e turbulence model, a two-step gasphase combustion formulation including chemical kinetics and a kinetic-diffusion model for the coal and straw char particles. The results showed a marked difference in the

combustion behavior (temperature, species concentration, etc.) due in part to the large volumetric concentration of straw near the burner mouth. Devolatilization and burnout of the larger straw particles occurred further away from the burner mouth, which changed the combustion behavior. In addition, the trajectories of the chopped straw were quite different from those of the coal particles. Dhanaplan et al. [79] obtained similar results when they numerically studied coal-only and coal/manure blend combustion in a swirl burner using the combustion code, PCGC2 [80]. In the near burner region, there were signicant differences in temperature and species concentrations between coal-only and coal/manure blend combustion. This is due to the different chemical compositions of coal and manure. For blend combustion, PCGC-2 was modied to incorporate a three-mixture fraction approach. The original PCGC2 tracks two-mixture fractions (primary and coal-off gases) only. In the modied code, a third mixture fraction was added. The third fraction, manure-off gas, was used because the properties of manure and coal are very different and consolidating fuel-off gases into one mixture fraction would lead to erroneous results. The authors compared the modeling results for 90:10 coal/manure blend (mass basis) as predicted by the three-mixture fraction model with those

206

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Fig. 24. (a) Ratio of nitrogen in NO to nitrogen in the fuel for four different fuels [7]; and (b) emission of NOx from coal and coal/manure blend studies [7].

from the original two-mixture fraction model. Similar sizes of coal and manure particles were used (70% 60 mm). They found that in the near burner regions there were significant differences between the two models in predicting the temperature distribution and species concentrations. Fig. 25 shows the results for the centerline temperature proles for the two- and three-mixture fraction approaches. The threemixture fraction model yields higher temperatures than those predicted by the two-mixture fraction model for distances greater than 0.2 m. The maximum temperature difference is about 350 K, which can be signicant, particularly when determining thermal NOx emissions. However, the ignition distance remained almost unaltered due to the small amounts of manure in the blend. Fig. 26 shows species concentration results (CO, CO2 and O2) at the centerline of the burner. The NOx concentrations were not evaluated. The

three-mixture fraction model predicted lower levels of oxygen and higher levels of CO and CO2, indicating slightly higher burnout. The results can be attributed to the carbon atom fraction, which was more accurately predicted in the three-mixture fraction approach. Separate treatment of the coal-off and manure-off gases avoids averaging coal- and manure-off gas properties and thus leads to improved results. Sami et al. [75] also modeled coal and manure blend combustion in a swirl burner using the PCGC2 code modied for three-mixture fractions. A 90:10 coal/manure blend (by mass) was studied. The burner dimensions are described elsewhere [7,79]. Manure particle sizes were twice that of coal particles (coal: 70% 60 mm, manure: 70% 110 mm). The effects of the primary and secondary airow rates on burnout and temperature proles were studied,

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

207

Fig. 25. Predicted centerline temperature proles for 90:10 coal/manure fuel blends using the two- and three-mixture fraction approach [79].

while maintaining the equivalence ratio at a constant value. Since, the stoichiometric air to fuel ratio for manure (5.2 kg air/kg manure) is much lower than that for coal (9.7 kg air/ kg coal), the equivalence ratio (w ) decreases when coal is replaced by a small percentage of manure. There are two ways to maintain the equivalence ratio at the base case (coal-only) value: (i) increasing the fuel loading in the primary air (equivalent to decreasing the primary air); or (ii) decreasing the secondary air. Fig. 27 shows a comparison of the burnout for three different cases when w was kept constant at 0.89. It is interesting to note that a higher burnout is obtained for the case where the secondary air ow was decreased (Case 3) as compared to the case where the fuel ow rate was increased (Case 4). However, the associated temperature prole for Case 3 resulted in lower values, as seen in Fig. 28. In fact, the Case 3 temperature prole was almost identical to that predicted for the coal-only case. This is due to the fact that the net heat input in Case 3 is less than that for coal-only case. The reduction in secondary air compensates the reduction in heat input such that the resulting temperatures are similar. There was also a shift in the location of the peak temperature for Case 4. For axial distances greater than 0.14 m, the centerline temperature increased with an increase in fuel ow rate. This was due to the fact that the _ heat capacity mf C of the fuel increased with an increase in

fuel ow rate. It therefore took more time to heat the fuel to the pyrolysis temperatures. Hence, ignition was delayed and the peak temperature shifted downstream of the burner. The relative changes may be interpreted as a change in the ame location. Unlike coal-only combustion, blend combustion produced more combustibles in the near-burner region for a given amount of primary air. More combustibles mean more energy in a given control volume. Thus, in Case 4, temperature decreased slowly when compared with the coalonly case for a given secondary airow rate. The conclusions of Sami et al. [75] are summarized below: 1. Blend combustion resulted in improved combustion efciencies compared to coal-only combustion. 2. Increasing fuel loading resulted in higher temperatures compared to the coal-only case. A downstream shift in the location of the peak temperature was also observed. Higher temperatures may also increase thermal NOx levels. 3. Decreasing the secondary air resulted in almost the same temperature proles (hence same thermal NOx level) as those of the coal-only case, however the burnout was improved signicantly. 4. In order to maintain the same equivalence ratio, it is better to reduce the secondary air ow than to increase

208

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Fig. 26. Predicted centerline species prole for 90:10 coal/manure fuel blends using the two- and three-mixture fraction approach [79].

the fuel ow rate. However, it should be noted that the heat throughput will also decrease slightly. Although some combustion models exist for fuel blend combustion, they are limited in their scope and complexity. Most of these models incorporate simplied assumptions in describing various aspects of combustion like kinetics, turbulence and particle dispersion, and thus the predictions obtained are qualitative in nature. For better predictions, accurate experimental data determining the chemical reac0.8 0.75 0.7 0.65 0.6 0.55 0.5 0.45 0.4 0.05

tions kinetics and a set of minimum assumptions must be employed. Fewer model assumptions can increase computational time many folds, however, with the rapid advancement in computing technology, it will be possible in the near future to run detailed combustion codes in a reasonable amount of time. 4.7. Fouling issues in co-ring Hansen et al. [77] investigated ash deposition in a multicirculating uidized bed combustor (MCFBC) red with fuel blends of coal, wood and straw. The ow rates were 650 kg coal, 1640 kg straw, 2330 kg wood chips and 25.8 metric ton of air per hour. Fig. 29 is a sketch of the combustor. Coal wood blend is injected in the bottom of the MCFBC above the primary air. Secondary air mixed with straw is introduced above the dense bed. The average bed temperature was 980 K. Concentrations of alkali metals, sodium and potassium, were measured at six different locations inside the burner as shown in Figs. 30 and 31. More potassium than sodium was found at all locations. More alkali vapors were measured just above the dense bed than above the riser section. The authors also developed a semiempirical model based on thermodynamic equilibrium calculations to predict stable forms of alkali metals sodium and potassium. A code (MINGTSYS) was used, which is

burnout

coal Blend_case4 Blend_case3

0.1

0.15

0.2

Axial distance (m)


Fig. 27. Predicted burnout along burner centerline for coal and coal/ manure blends [75]. A 90:10 blend was used (Case 3 reduction in secondary air, Case 4 increase in fuel ow rate).

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214


2200 Co al 2100 90:10 B lend_case4 90:10 B lend_case3

209

Tem p era ture (K)

2000 1900 1800 1700 1600 1500 0.05 0.1 0.15 0.2

0.25

0.3

Axial distance (m)


Fig. 28. Predicted centerline temperature proles along burner centerline for coal and coal/manure blends [75]. A 90:10 blend was used (Case 3 reduction in secondary air, Case 4 increase in fuel ow rate).

based on the concept of Gibbs minimum free energy. In the calculation, 120 different compounds of C, H, O, N, Cl, Na, K or Si were examined. They found good agreement, except in the riser section, where predicted potassium concentrations were higher than measured values. Clearly more studies are needed to understand the fouling mechanism in coal biomass blend combustion. The successful implementation of co-ring technology depends upon nding new ways to mitigate fouling and corrosion associated with the combustion of biomass fuels. This apparently is the most important issue that needs to be addressed.

sion burning in pulverized coal boilers and implementation costs. 1. Fuel availability can be dealt with if ample sources are close to the power plant. For example, in Texas, large quantities of feedlot manure are available. In Kansas, wheat and corn residues are readily available. Moreover, dedicated energy crops have been grown in some parts of the United States and Europe for the sole purpose of energy generation. 2. The price of biomass depends on the collection, transportation, drying and grinding processes. These costs need be lower than the primary fuel cost, on energy basis, in order to obtain a cost advantage. 3. Deposition and high-temperature corrosion are significant concerns to boiler operators. High concentrations of potassium and chlorine in biomass can cause serious problems, such as slagging, fouling and corrosion. However, in pulverized fuel boilers, this phenomenon is not observed to the extent that it is observed in uidized beds and stoker red boilers [81] where ash agglomeration is a serious problem. It should be noted, however, that these phenomena require long test times to evaluate the impact on the combustor performance. Short duration tests (i.e. less than a few hours) may not provide meaningful data on slagging, fouling and corrosion rates. 4. Flame stand-off distance can be altered when blends are red in coal burners. In order to have the same heat throughput, more biomass fuel is red due to the lower heating value of biomass fuels. This causes the ame to shift downstream in the combustion chamber. This shift in ame location may cause ame instabilities and increase in NOx levels. 5. The maximum size of biomass particle for suspension ring varies depending upon the type of fuel. Spliethoff and Hein [6] recommend maximum particle sizes

5. Issues and opportunities Some of the main issues in the co-ring of coal and biomass fuel blends are availability, fuel price, hightemperature corrosion, slagging and fouling effects, ame location, size of the biomass particles for suspen-

Fig. 29. Schematic of MCFBC. R1R3 and P1P3 indicate measurement locations [77].

210

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

Fig. 30. Measured concentrations of Na and K at locations R1R3 in the riser section. Measurement locations are shown in Fig. 29. Measurements carried out simultaneously are connected by lines [77].

of 4 mm for straw and 6 mm for wood. This issue needs to be addressed further in order to determine the relationship between particle size, type of biomass and the economics of pulverizing the fuel. 6. The implementation costs depend on the class of coring desired, the biomass paticle size and the coring ratio. If the co-ring percentage is very small, then pre-blended coal and biomass fuels can be red in existing facilities with very few modications. As the size of biomass particles increases, non-premixed

blends should be used. The biomass can then be cored using separate feed lines to avoid clogging. Pulverizer performance is also a factor if more than 5% biomass is red [5]. In this case, a dedicated biomass feeding mechanism may be required increasing implementation cost. Although there are still many important issues that are yet unanswered regarding coal biomass blend combustion, there are numerous attractive features to

Fig. 31. Measured concentrations of Na and K at locations P1P3 in the pre-separator. Measurement locations are shown in Fig. 29. Measurements carried out simultaneously are connected by lines [77].

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214

211

blend combustion. These opportunities include additional incentives (such as tax cuts and allowances), reduction in emissions, disposal of waste, jobs in agricultural sector (with the cultivation of dedicated energy crops) and low implementation costs. 1. The EPA awards allowances and tax credits to utilities who comply with agreed emissions levels. The allowances and credits offset costs incurred by the utilities in retrotting and/or buying new equipment. Some states in the US have laws that favor the use of biomass in utilities. In addition to these programs, some states have implemented renewable energy credits requiring specic generation or capacity levels or other green power initiatives [82]. 2. A strong motivation for biomass fuel usage comes from the fact that gaseous emissions are reduced when biomass is red with coal. 3. Waste disposal is a perennial problem with some biomass sources, e.g. manure. Co-ring these biomass fuels also reduces waste accumulation and attendant soil, water and air pollution. 4. When dedicated energy crops are cultivated to ensure continuous supply of biomass fuels to utilities, more jobs will be created in the agriculture sector. More people will be employed and economy may improve in the rural areas. 5. Finally, provided the blend ratio is small and the size of the biomass particles is suitable, co-ring implementation cost would be very low. If biomass fuel costs are also low, a net prot can be obtained compared to coal-only ring. The future of coal and biomass blend combustion in utility boilers looks very bright. Based on the positive results of recent co-ring studies, coupled with more strict environmental regulations and associated penalties, utilities are seriously considering co-ring locally available biomass fuels with coal in their boilers. TVA, for example, has a research and development program for co-ring wood residue with coal in utility boilers. DOE and EPRI are actively engaged in co-ring different biomass fuels in coal red boilers. In addition, many sections of the United States have made using a percentage of biomass mandatory for electric utilities. Some of these states include Connecticut, Massachusetts, Nevada, Maine, California, Colorado, Iowa, Minnesota, New York and Wisconsin.

can be reduced because biomass is a CO2 neutral fuel. The co-ring programs carried out in the United States (mainly under the auspices of DOE and EPRI) and Europe (under the European Union) have demonstrated that co-ring biomass with coal in large utility boilers can be benecial to the utilities as well as to the environment. Co-ring may also reduce fuel costs, minimize waste and reduce soil and water pollution depending upon the chemical composition of the biomass used. Co-ring technology, however, faces some technological problems. First, the issue of combustor fouling and corrosion due to the alkaline nature of the biomass ash needs attention. Ash deposits reduce the heat transfer and may also result in severe corrosion at high temperatures. Compared to deposits generated during coal combustion, deposits from biomass materials are denser and more difcult to remove. Second, the maximum particle size of a given biomass that can be fed to and burned in a given PC boiler through a given feeding mechanism requires additional studies. However, this issue is a combination of economics, and combustion characteristics and more work needs to be done in this area. Third, practical pulverizer performance needs to be examined. Biomass fuels may require separate pulverizers to achieve high blend ratios and good combustion performance. Since biomass fuels have lower heating value compared to coal, blend ow rate has to be increased in order to have a heat throughput same as in coal-only case. This increased fuel ow rate may cause the ame to move away from the burner mouth, thereby creating ame stability problem. Lifted ames are also known to cause higher NOx levels. Fundamental combustion studies must be performed, particularly for pre-mixed coal and biomass fuel blends, in order to determine combustion behavior characteristics in controlled laboratory settings. Interaction between biomass and coal particles during combustion is an area in particular need of study. The results of this basic research will aid in the design and optimization of practical coal and biomass blend facilities. Despite all the issues and concerns, coalbiomass blend combustion appears to be a promising combustion technology for electric utilities. Co-ring has moved from engineering studies to parametric tests to long-term demonstrations. Future long-term demonstrations will address many of the issues mentioned above and will help in making the co-ring technology easily available to the industry at an optimal cost.

Acknowledgements 6. Conclusions Coal and biomass fuels are quite different in composition. Co-ring biomass fuels with coal has the capability to reduce both NOx and SOx levels from existing pulverizedcoal red power plants. In addition, overall CO2 emissions This work was supported by the Advanced Technology Research Program of the State of Texas and the Western Biomass Regional Program (DOE) through the Texas Engineering Experiment Station at Texas A & M University.

212

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214 Combustion, Vancouver, Canada, 1114 May, vol. 1, 1997. p. 31320. Ohlsson O. Results of combustion and emissions testing when co-ring blends of binder-enhanced densied refuse-derived fuel (b-dRDF) pellets and coal in a 440 MWe cyclone red combustor, vol. 1, test methodology and results. Subcontract report No. DE94000283. Argonne National Laboratory, Argonne, IL, 1994. p. 60. Walawender WP, Fan LT, Engler CR, Erickson LE. Feedlot manure and other agricultural wastes as future material and energy resources: II. Process descriptions. Contribution 30, Department of Chemical Engineering, Kansas Agricultural Experiment Station, Manhattan, KS, 1973. 30 p. Huffman WJ. Alternate manure recycling systems for energy recovery. Methane production from livestock manure. In: Proceedings of the Great Plains Extension Seminar and Tour, Liberal, KS, 15 February, 1978. p. 319. Kreis RD. Recovery of by-products from animal wastesa literature review, EPA-600/2-79-142, Robert S. Kerr Environmental Research Laboratory, US Environmental Protection Agency, Ada, OK, 1979. 50 p. Beck SR. Economic feasibility of cattle manure as a chemical feedstock, livestock waste: a renewable source. In: Proceedings of the 4th International Symposium, Amarillo, TX. American Society of Agricultural Engineers, St. Joseph, MI, 1517 April, 1980. p. 3146. Raman KP, Walawander WP, Fan LT. Gasication of feedlot manure in a uidized bed: effect of temperature. Ind Engng Chem Proc Des Dev 1980;10:6239. LePori WA. Fluidized bed combustion and gasication of biosolid. Agric Energy 1980;2:3304. Paul WC, Buchele WF. Development of a concentric vortex agri-residue furnace. Agric Energy 1980;2:34956. Sweeten JM, Korenberg J, LePori WA, Annamalai K, Parnell CB. Combustion of feedlot manure for energy production. Energy Agric 1986;5:5572. Annamalai K, Ibrahim YM, Sweeten JM. Experimental studies on combustion of cattle manure in a uidized bed combustor. Trans ASME J Energy Resour Technol 1987;109:4957. Boie W. Wiss Z Tech Hochsch Dresden 1952/53;2:687. Annamalai K, Sweeten J, Ramalingam S. Estimation of gross heating values of biomass fuels. Trans ASAE 1987;30:12058. Ebling JM, Jenkins BM. Physical and chemical properties of biomass. Trans ASAE 1985;23(3):898902. Annamalai K, Sibulkin M. Ignition and ame spread tests of cellular plastics. J Fire Flammab 1978;9:44558. Babcock, Wilcox. Steam: its generation and use. Babcock and Wilcox Co., 39th ed., 1978. Essenhigh RH. The inuence of coal rank on the burning times of single captive particles. J Engng Power 1963;85:18390. Essenhigh RH, Misra MK, Shaw DW. Ignition of coal particles: a review. Combust Flame 1989;77:330. Smoot L, Pratt DT. Pulverized coal combustion and gasication. New York: Plenum Press, 1979. Juntgen H, Van Heek KH. Gas release from coal as a function of the rate of heating. Fuel 1968;47:10317. McLean WJ, Hardesty DR, Poul JH. Direct observations of devolatlizing pulverized coal particles in a combustion environment. In: 18th Symposium (International) on Combustion, 1981. p. 123948.

References
[1] Department of Energy, Energy Information Administration, Annual energy outlook 1998, DOE/EIA-0383(98), Washington, DC, December, 1997. [2] Hein KRG, Bemtgen JM. EU clean coal technology, cocombustion of coal and biomass. Fuel Process Technol 1998;54:15969. [3] Hughes EE, Tillman DA. Biomass coring: status and prospects 1996. Fuel Process Technol 1998;54:12742. [4] Easterly JL, Burnham M. Overview of biomass and waste fuel resources for power production. Biomass Bioenergy 1996;10(23):7992. [5] Gold BA, Tillman DA. Wood coring evaluation at TVA power plants. Biomass Bioenergy 1996;10(23):718. [6] Spliethoff H, Hein KRG. Effect of co-combustion of biomass on emissions in pulverized fuel furnaces. Fuel Process Technol 1998;54:189205. [7] Frazzitta S, Annamalai K, Sweeten J. Performance of a burner with coal and coal: manure blends. J Propulsion Power 1999;15(2):1816. [8] Abbas T, Costen P, Kandamby NH, Lockwood FC, Ou JJ. The inuence of burner injection mode on pulverized coal and biosolid co-red ames. Combust Flame 1994;99:61725. [9] Van Doorn J, Bruyn P, Vermeij P. Combined combustion of biomass, municiple sewage sludge and coal in an atmospheric uidised bed installation. In: Biomass for energy and the environment, Proceedings of the 9th European Bioenergy Conference, Copenhagen, Denmark, vol. 2, 2427 June, 1996. p. 100712. [10] Aerts DJ, Bryden KM, Hoerning JM, Ragland KW. Co-ring switchgrass in a 50 MW pulverized coal boiler. In: Proceedings of the 1997 59th Annual American Power Conference, Chicago, IL, vol. 59(2), 1997. p. 11805. [11] Sampson GR, Richmond AP, Brewster GA, Gasbarro AF. Coring of wood chips with coal in interior Alaska. For Prod J 1991;41(5):536. [12] Fahlstedt I, Lindman E, Lindberg T, Anderson J. Co-ring of biomass and coal in a pressurized uidised bed combined cycle. Results of pilot plant studies. In: Proceedings of the 14th International Conference on Fluidized Bed Combustion in Vancouver, Canada, vol. 1, 1997. p. 2959. [13] Kaer SK, Rosendahl L, Overgaard P. Numerical analysis of co-ring coal and straw. In: Proceedings of the 4th European CFD Conference, Athens, Greece, 711 September, 1998. p. 11949. [14] Christensen J, Jespersen P. Straw-ring tests at Amager and Kyndby power stations. Biomass for energy and the environment. In: Proceedings of the 9th European Bioenergy Conference, Copenhagen, Denmark, 2427 June, vol. 2, 1996. p. 10138. [15] Siegel V, Schweitzer B, Spliethoff H, Hein KRG. Preparation and co-combustion of cereals with hard coal in a 500 kW pulverized-fuel test unit. Biomass for energy and the environment. In: Proceedings of the 9th European Bioenergy Conference, Copenhagen, Denmark, 2427 June, vol. 2, 1996. p. 102732. [16] Andries J, Verloop M, Hein K. Co-combustion of coal and biomass in a pressurized bubbling uidized bed. In: Proceedings of the 14th International Conference on Fluidized Bed [17]

[18]

[19]

[20]

[21]

[22]

[23] [24] [25]

[26]

[27] [28] [29] [30] [31] [32] [33] [34] [35] [36]

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214 [37] Kharbat E. Digital image processing applications in the ignition and combustion of coal. MS thesis, Mechanical Engineering Department, Texas A & M University, 1992. [38] Seeker WR, Samuelsen GS, Heap MP, Trolinger JD. The thermal decomposition of pulverized coal particles. In: 18th Symposium (International) on Combustion, 1981. p. 121326. [39] Nettleton MA, Stirling R. The combustion of clouds of coal particles in shock-heated mixtures of oxygen and nitrogen. Proc R Soc Lond, Ser A 1971;322:20721. [40] Stickler DB, Ubhayaker SK, Becker FE. Combustion of pulverized coal in high temperature preheated air, 17th Aerospace Sciences Meeting, New Orleans, LA, 1979. [41] Lau CW, Niksa S. The combustion of individual particles of various coal types. Combust Flame 1992;90:4570. [42] Annamalai K, Durbetaki P. A theory on transition of ignition phase of coal particles. Combust Flame 1977;29:193208. [43] Hottel HC, Williams GC, Nerheim NM, Schneider GR. Kinetic studies in stirred reactors: combustion of CO and propane. In: 10th Symposium (International) on Combustion, 1965. p. 11121 (see also p. 141320). [44] Biswas BK, Essenhigh RH. The problem of smoke formation and its control, air pollution and its control, AIChE Symp Series No. 126, vol. 68, 1972. p. 20715. [45] Shaw DW, Zhu X, Misra MK, Essenhigh RH. Determination of global kinetics of coal volatiles combustion. In: 23rd Symposium (International) on Combustion, 1990. p. 115562. [46] Edelman RB, Fortune O, Weilerstein G. Some observations in reacting ows described by coupled mixing and kinetics. In: Cornelius W, Agnew WG, editors. Proceedings of the GM Symposium on Emissions from Continuous Combustion Systems, New York: Plenum Press, 1972. p. 5587. [47] Smoot LD, Horton MD, Williams GA. Propagation of laminar pulverized coal-air ames. In: 16th Symposium (International) on Combustion, 1977. p. 37587. [48] Smoot L, Smith PJ. Coal combustion and gasication. New York: Plenum Press, 1985. [49] Howard JB, Sarom AF. Gasication of coal char with carbon dioxide and steam at 12001800 C, Energy Laboratory Report, Department of Chemical Engineering, Massachusetts Institute of Technology, 1978. [50] Matsui K, Tsuji H, Makino A. Estimation of the relative rates of CO2 and C-H2O reactions. Carbon 1983;21(3):3201. [51] Mitchell RE, Kee RJ, Glarborg P, Coltrin ME. The effect of CO conversion in the boundary layers surrounding pulverized coal-char particles. In: 23rd Symposium (International) on Combustion, 1990. p. 116976. [52] Kuo K. Principles of combustion. New York: Wiley, 1986. [53] Glassman I. Combustion. 2nd ed. New York: Academic Press, 1987. [54] Maloney DJ, Monazam ER, Woodruff SD, Lawson LO. Measurements and analysis of temperature histories and size changes for single carbon and coal particles during the early stages of heating and devolatalization. Combust Flame 1991;84:21020. [55] Du X, Annamalai K. The transient ignition of isolated coal particles. Combust Flame 1994;97:33954. [56] Karcz H, Kordylewski W, Rybak W. Evaluation of kinetic parameters of coal ignition. Fuel 1980;59:799802. [57] Tran P, Annamalai K. Heat and mass transfer analysis of the

213

[58]

[59]

[60] [61] [62]

[63]

[64]

[65]

[66]

[67]

[68]

[69] [70]

[71]

[72]

[73]

[74]

ignition and extinction of solid fuel particles. J Heat Transfer, Trans ASME 1999;121:88693. Ubhayakar SK. Burning characteristics of a spherical particle reacting with ambient oxidizing gas at its surface. Combust Flame 1976;26:2434. Essenhigh RH. In: Elliott MA, editor. Fundamentals of coal combustion, Chemistry of coal utilization, vol. II. New York: Wiley/Interscience, 1979 [chap. 19]. Tillman DA, Rossi AJ, Kitto WD. Wood combustion: principles, processes and economics. New York: Academic Press, 1981. Pan YG, Velo E, Puigjaner L. Pyrolysis of blends of biomass with poor coals. Fuel 1996;75(4):4128. Winter F, Prah ME, Hofbauer H. Temperatures in a fuel particle burning in a uidized bed: the effect of drying, devolatilization and char rcombustion. Combust Flame 1997;108:30214. Dasappa S, Sridhar HV, Paul PJ, Mukunda HS, Shrinivasa U. On the combustion of wood-char spheres in O2/N2 mixtures experiments and analysis. In: 25th Symposium (International) on Combustion. The Combustion Institute, 1994. p. 56976. Wornat MJ, Hurt RH, Yang NYC, Headley TJ. Structural and compositional transformations of biomass chars during combustion. Combust Flame 1995;100:13143. Wornat MJ, Hurt RH, Davis KA, Yang NYC. Single-particle combustion of two biomass chars. In: 26th Symposium on Combustion. The Combustion Institute, 1996. p. 307583. Baxter LL. Ash deposition during biosolid and coal combustion. A mechanistic approach. Biomass Bioenergy 1993;4(2):85 102. Mendis MS. Biomass gasication: past experiences and future prospects in developing countries. Pyrolysis and Gasication. In: Proceedings of the International Conference in Luxembourg, 2325 May, 1989. p. 11128. Hartiniati SA, Youvial M. Performance of a pilot scale uidized bed gasier fueled by rice husks, pyrolysis and gasication. In: Proceedings of the International Conference in Luxembourg, 2325 May, 1989. p. 25763. Kurkela E, Stahlberg P, Simell P, Leppalahti J. Updraft gasication of peat and biosolid. Biosolid 1989;19:3746. Christian R, Unsworth J. Pilot plant demonstration of used tires vacuum pyrolysis. Pyrolysis and gasication. In: Proceedings of the International Conference in Luxembourg, 2325 May, 1989. p. 1809. Armesto L, Cabanillas A, Bahillo A, Segovia JJ, Escalada R, Martinez JM, Carrasco JE. Coal and biomass co-combustion on uidized bed: comparison of circulating and bubbling uidized bed technologies. In: Proceedings of the 14th International Conference on Fluidized Bed Combustion in Vancouver, Canada, vol. 1, 1997. p. 3019. Rasmussen I, Clausen JC. ELSAM strategy of ring biosolid in CFB power plants. In: Proceedings of the 13th International Conference on Fluidized Bed Combustion, Orlando, FL, 710 May, vol. 1, 1995. p. 55763. Turnbull JH. Biosolid power as a strategic business investment. In: Proceedings of the 1996 International Joint Power Generation Conference, part 1, vol. 1, 1996. p. 35564. Liu H, Gibbs BM. Hampartsoumian. The signicance of reburning coal rank on the reduction of NO in drop tube furnace. In: 8th International Symposium on Transport phenomena in combustion, San Francisco, CA, 1995.

214

M. Sami et al. / Progress in Energy and Combustion Science 27 (2001) 171214 [83] Maschio G, Lucchesi A, Stoppato G. Production of syngas from biosolid. Bioresour Technol 1994;48:11926. [84] Kandpal JB, Maheshwari RC, Kandpal TC. Particulate pollution from biosolid combustion in three cookstoves. Int J Energy Res 1995;19:43341. [85] Ragland KW, Aerts DJ, Baker AJ. Properties of wood for comobustion analysis. Bioresour Technol 1991;37:1618. [86] Saxena SC, Jotshi CK. Fluidized-bed incineration of waste materials. Prog Energy Combust Sci 1994;20:281324. [87] Annamalai K, Ryan W. Interactive processes in gasication and combustion-II. Isolated carbon, coal and porous char particles. Prog Energy Combust Sci 1993;19:383446. [88] Kobayashi H. 16th Symposium (International) on Combustion. The Combustion Institute, Pittsburgh, 1977. p. 41125. [89] Solomon PR, Serio MA, Carangelo RM, Markham JR. Very rapid coal pyrolysis. Fuel 1986;65:18294. [90] Demirba A. Kinetics for non-isothermal ash pyrolysis of hazel shell. Bioresour Techonol 1998;66(3):24752. [91] Sharma A, Rao TR. Kinetics of pyrolysis of rice husk. Bioresour Technol 1998;66(3):5359. [92] Reina J, Velo E, Puigjaner L. Thermogravimetric study of the pyrolysis of waste wood. Thermochim Acta 1998;320(12):1617. [93] Rocca PAD, Cerrella EG, Bonelli PR, Cukierman AL. Pyrolysis of hardwoods residues: on kinetics and chars characterization. Biomass Bioenergy 1999;16(1):7988. [94] Rao TR, Sharma A. Pyrolysis rates of biomass materials. Energy 1998;23(11):9738. [95] Vargas JM, Perimutter DD. Interpretation of coal pyrolysis kinetics. Ind Engng Chem Proc Des Dev 1986;25:4954.

[75] Sami M, Annamalai K, Dhanapalan S, Wooldridge M. Numerical simulation of blend combustion of coal and feedlot waste in a swirl burner. In: ASME Conference, Nashville, TN, November, 1999. [76] Brouwer J, Owens WD, Harding S, Heap MP, Pershing DW. Coring waste biofuels and coal for emissions reduction. In: Proceedings of the 2nd Biomass Conference of the Americas, Portland, OR, August, 1995. p. 3909. [77] Hansen LA, Michelsen HP, Dam-Johansen K. Alkali metals in a coal and biosolid red CFBC-measurements and thermodynamic modeling. In: Proceedings of the 13th International Conference on Fluidized Bed Combustion, Orlando, FL, 7 10 May, vol. 1, 1995. p. 3948. [78] Jones WP, Launder BE. The prediction of laminarization with a two-equation turbulence model. Int J Heat Mass Transfer 1972;15:30114. [79] Dhanaplan S, Annamalai K, Daripa P. Turbulent combustion modeling coal: biosolid blends in a swirl burner. Energy Week, vol. IV, ETCE, ASME, January, 1997. p. 41523. [80] PCGC2, revised users manual. Advanced Combustion Engineering Research Center, Brigham Young University and University of Utah, 1988. [81] Hansen PFB, Andersen KH, Wieck-Hansen K, Overgaard P, Rasmussen I, Frandsen FJ, Hansen LA, Dam-Johansen K. Co-ring straw and coal in a 150 MWe utility boiler: in situ measurements. Fuel Process Technol 1998;54:207 25. [82] Department of Energy. Energy Information Administration, annual energy outlook 1998, DOE/EIA-0383(99), Washington, DC, December, 1997.

You might also like