Вы находитесь на странице: 1из 113

The Pennsylvania State University The Graduate School College of Earth and Mineral Sciences

EFFECT OF COAL RANK DURING OXY-FUEL COMBUSTION: ROLE OF CHARCO2 REACTION

A Thesis in Energy and Mineral Engineering by Sakthivale Roshan Dhaneswar

2011 Sakthivale Roshan Dhaneswar

Submitted in Partial Fulfillment of the Requirements for the Degree of

Master of Science

August 2011

The thesis of Sakthivale Roshan Dhaneswar was reviewed and approved* by the following:

Sarma V. Pisupati Associate Professor of Energy and Mineral Engineering Thesis Advisor

Sharon F. Miller Research Associate at EMS Energy Institute Graduate Faculty, Department of Energy and Mineral Engineering

Yaw D. Yeboah Professor of Energy and Mineral Engineering Head of the Department of Energy and Mineral Engineering

*Signatures are on file in the Graduate School

iii

ABSTRACT
The role of coal rank in oxy-fuel combustion is investigated using a suite of four coals: a low volatile bituminous coal, a high volatile bituminous coal, a subbituminous coal and a lignite. The study was conducted using the -170+200 mesh fraction coal in a drop-tube reactor (DTR) at 1,873 K to generate chars at three residence times under air, 21% O2/79% CO2 (oxy-fuel combustion) and 30%O2/70%CO2 (enhanced oxy-fuel combustion) atmospheres. It was observed that oxy-fuel combustion produced a higher carbon conversion compared to that in air for the low-rank subbituminous coal and lignite and also for the high volatile bituminous coal at longer residence times. This effect was even more pronounced at shorter residence times than at longer residence times for all ranks of coal. For the high-rank low volatile coal, air combustion produced a higher carbon conversion than oxy-fuel combustion. Also, as the fraction of O2 was increased in the mixture of O2/CO2, the carbon conversion was found to increase. The higher carbon

conversion in low-rank coals under oxy-fuel conditions was attributed to the char-CO2 gasification reactions occurring at high temperatures. The activation energy of the char-CO2 reaction for low-rank coals was more sensitive to temperature than high-rank coals, and the effect was more pronounced in enhancing the carbon conversion. Investigation of the catalytic activity (specifically in catalyzing the char-CO2 reaction) of ion-exchangeable cations in lower rank coal showed that there was a marked decrease in the isothermal reactivity at lower temperatures in a TGA. However, the same effect was not noticed in the drop tube reactor at higher temperatures. The lack of catalytic activity at higher temperatures (> 1,000 K) was attributed sintering of CaO at high temperatures The BET surface area of the chars was found to decrease and the average pore diameter was found to increase with decrease in coal rank showing the higher concentration of macropores in low-rank coals, leading to higher reactivity. Modified burning and reactivity profiles were generated using a thermogravimetric analyzer. The theoretically extrapolated

iv reaction rates corroborated well with the combustion test results with the lower rank coals showing higher char-O2 and char-CO2 reactivity. The effectiveness factors for the char-O2 and char-CO2 reactions indicated a higher degree of diffusion control for the char-O2 reaction. At similar temperature, CO2 molecules are able to access a higher percentage of the internal surface area compared O2 molecules due to relatively lower reactivity of char-CO2 reaction. The results also indicated that carbon conversion was a function of both overall reactivity and effectiveness factor. The influence of the char-CO2 reaction on particle temperature was theoretically modeled and the particles were at much higher temperatures compared to the gas at lower temperatures due to absence of the char-CO2 reaction. However, the particles approached the gas temperature at higher temperatures due to the endothermic char-CO2 reaction.

TABLE OF CONTENTS
LIST OF FIGURES..................................................................................................................viii LIST OF TABLES......................................................................................................................xi ACKNOWLEDGEMENTS.......................................................................................................xii Chapter 1 Introduction ............................................................................................................. ....1 Chapter 2 Literature Review .................................................................................................... ....4 2.1 Gas and Particle Temperatures..................................................................................4 2.2 Flame Characteristics................................................................................................5 2.3 Ignition Characteristics.............................................................................................6 2.4 Effect of Pressure......................................................................................................7 2.5 Char Morphology......................................................................................................7 2.6 Char Conversion and Reactivity...............................................................................8 2.7 Char Conversion rate Parameters Estimation.........................................................22 2.8 Process Feasibility and Economics.........................................................................25 2.9 Summary.................................................................................................................26 Chapter 3 Problem Statement .................................................................................................. .28 3.1Methodology............................................................................................................28 3.2Hypothesis...............................................................................................................29 Chapter 4 Experimental Details ............................................................................................... 31 4.1 Coal Sample Preparation........................................................................................31 4.2 Drop Tube Reactor.................................................................................................31 4.2.1 Combustion Test Conditions........................................................................33 4.2.2 Devolatilization Test Conditions..................................................................34

vi 4.3 Microproximate Analysis.......................................................................................35 4.4 Determination of Char Reactivity..........................................................................36 4.5 Modified burning Profiles......................................................................................37 4.6 Compositional Analysis..........................................................................................38 4.6.1 CHN Analysis...............................................................................................38 4.6.2 Sulfur Analysis..............................................................................................38 4.7 Surface Area Analysis.............................................................................................39 4.8 Major/Minor Oxide Analysis of Ash......................................................................39 4.9 Removal of ion-exchangeable cations....................................................................40 Chapter 5 Results and Discussion ............................................................................................ .42 5.1 Combustion Test Results - Parent coals..................................................................42 5.2 Devolatilization and Isothermal Reactivity Studies................................................48 5.3 Theoretical char-O2 and char-CO2 reaction rate study............................................58 5.4 Theoretical char particle temperature model...........................................................67 5.5 Combustion Test Results - AAW coals..... .............................................................73 5.6 Thermogravimetric modified burning profiles under O2/N2 and O2/CO2 atmospheres...................................................................................................................75 Chapter 6 Summary, Conclusions and Recommendations ...................................................... .78 6.1 Summary and Conclusions.....................................................................................78 6.2 Future Recommendations.......................................................................................80 Appendix A Calculation of extrapolated rate at combustion condition using intrinsic model ........................................................................................................................ .81 Appendix B Example of char conversion calculation..... ................................................ .84 Appendix C Sample calculation to extract intrinsic rate parameters from an isothermal char reactivity plot .................................................................................. .85 Appendix D Proximate and Ultimate Analysis of DTR chars..........................................86 Appendix E Plots of Rd, Rs, Rac and R as a function of temperature for the four

vii coals...........................................................................................................................90 Appendix F Sample char particle temperature calculation ............................................. .94 References..................................................................................................................................96

viii

LIST OF FIGURES
Figure 1-1. General flow process of oxy-fuel combustion..........................................................2 Figure 2-1. Plot of reaction rate vs. inverse temperature depicting three zones..........................9 Figure 2-2. Schematic diagram depicting a shrinking particle with an unreacted shrinking core, and concentration profile of gas reactant..........................................................................10 Figure 2-3. General shape of the effectiveness factor vs. temperature curve............................19 Figure 2-4. Comparison of char reaction with O2 (oxidation) and CO2 (gasification)..............21 Figure 2-5. Low temperature TGA kinetic data extrapolated to high temperatures (spheres: O2, rectangles: steam, triangles: CO2)........................................................................22 Figure 2-6. Reactivity profiles in air at 673 K for chars obtained under different pyrolysis conditions...............................................................................................................24 Figure 4-1. Schematic of Drop tube reactor...............................................................................32 Figure 5-1. Carbon conversion (%) vs. residence time (s) for Pocahontas coal........................46 Figure 5-2. Carbon conversion (%) vs. residence time (s) for Pittsburgh coal..........................46 Figure 5-3. Carbon conversion (%) vs. residence time (s) for Dietz coal..................................47 Figure 5-4. Carbon conversion (%) vs. residence time (s) for Beulah lignite...........................47 Figure 5-5. Char-O2 reactivity profiles for Pocahontas coal......................................................52 Figure 5-6. Char-O2 reactivity profiles for Pittsburgh coal........................................................53 Figure 5-7. Char-O2 reactivity profiles for Dietz coal...............................................................53 Figure 5-8. Char-O2 reactivity profiles for Beulah lignite.........................................................54 Figure 5-9 Char-CO2 reactivity profiles for Pocahontas coal....................................................54 Figure 5-10. Char-CO2 reactivity profiles for Pittsburgh coal...................................................55 Figure 5-11. Char-CO2 reactivity profiles for Dietz coal...........................................................55 Figure 5-12. Char-CO2 reactivity profiles for Beulah lignite....................................................56

ix Figure 5-13. Extrapolated oxidation rates for the four coals used in the study.........................60 Figure 5-14. Char-O2 reaction effectiveness factors for the four coals used in the study...........................................................................................................................................60 Figure 5-15. Extrapolated gasification rates for the four coals used in the study......................61 Figure 5-16. Char-CO2 reaction effectiveness factors for the four coals used in the study...........................................................................................................................................61 Figure 5-17. Variation of effectiveness factor and intrinsic reaction rate for char-O2 reaction of Pittsburgh coal.........................................................................................................62 Figure 5-18. Variation of effectiveness factor and intrinsic reaction rate for char-CO2 reaction of Pittsburgh coal.........................................................................................................62 Figure 5-19. Extrapolated oxidation Rac and Rd for the four coals used in the study................64 Figure 5-20. Extrapolated gasification Rac and Rd for the four coals used in the study.............64 Figure 5-21. Char-O2 effectiveness factors for two particle sizes of Dietz coal........................65 Figure 5-22. Char-O2 effectiveness factors for two particle sizes of Dietz coal........................65 Figure 5-23. Computed char particle temperature (K) vs. oxygen concentration(%) at a gas temperature of 1,373 K under oxy-fuel conditions..............................................................70 Figure 5-24. Computed char particle temperature (K) vs. oxygen concentration(%) at a gas temperature of 1,573 K under oxy-fuel conditions..............................................................70 Figure 5-25. Computed char particle temperature (K) vs. oxygen concentration(%) at a gas temperature of 1,673 K under oxy-fuel conditions..............................................................71 Figure 5-26. Computed char particle temperature (K) vs. oxygen concentration(%) at a gas temperature of 1,773 K under oxy-fuel conditions..............................................................71 Figure 5-27. Computed char particle temperature (K) vs. oxygen concentration(%) at a gas temperature of 1,873 K under oxy-fuel conditions..............................................................72 Figure 5-28. Simulated temperature (oC) profiles for an oxy-fuel PC boiler.............................72

Figure 5-29. Carbon conversion (%) vs. residence time (s) for Dietz and AAW Dietz coal.............................................................................................................................................73 Figure 5-30. Carbon conversion (%) vs. residence time (s) for Beulah and AAW Beulah lignite.............................................................................................................................75 Figure 5-31. Thermogravimetric burning profile generated under air for the four coals and a switchgrass sample used in the study...............................................................................76 Figure 5-32. Thermogravimetric modified burning profile generated under 21%O2/79%CO2 for the four coals and a switchgrass sample used in the study...................................................76 Figure C-1. Plot of ln (k) vs. 1/T for Beulah char-O2 reactivity................................................85 Figure E-1. Extrapolated oxidation Rs, Rac, Rd and R for Pocahontas coal...............................90 Figure E-2. Extrapolated gasification Rs, Rac, Rd and R for Pocahontas coal............................90 Figure E-3. Extrapolated oxidation Rs, Rac, Rd and R for Pittsburgh coal.................................91 Figure E-4. Extrapolated gasification Rs, Rac, Rd and R for Pittsburgh coal..............................91 Figure E-5. Extrapolated oxidation Rs, Rac, Rd and R for Dietz coal.........................................92 Figure E-6. Extrapolated gasification Rs, Rac, Rd and R for Dietz coal.....................................92 Figure E-7. Extrapolated oxidation Rs, Rac, Rd and R for Beulah lignite...................................93 Figure E-8. Extrapolated gasification Rs, Rac, Rd and R for Beulah lignite...............................93

xi

LIST OF TABLES
Table 1-1. Typical flue gas composition......................................................................................3 Table 2-1. Char-CO2 reaction activation energy values reported in literature...........................23 Table 4-1. Typical PC boiler residence times............................................................................34 Table 4-2. DTR combustion test parameters.............................................................................34 Table 5-1. Analysis of coal and switchgrass samples................................................................44 Table 5-2. Major and minor elemental oxide analysis of the ash reported as oxides (%) for the coals used in the study...............................................................................................................45 Table 5-3. Reproducibility of Combustion tests........................................................................48 Table 5-4. Analysis of char samples from DTR........................................................................49 Table 5-5. Surface area analysis of devolatilization char samples.............................................50 Table 5-6. Char-O2 reactivity for the four coals........................................................................57 Table 5-7. Char-CO2 reactivity for the four coals......................................................................57 Table 5-8. Oxidation and gasification rate parameters for the four coals..................................57 Table 5-9. Rate Parameters used in the theoretical particle temperature model........................68 Table 5-10. Characteristic temperatures under air and oxy-fuel atmospheres...........................77 Table D-1. Proximate and Ultimate Analysis of Pocahontas combustion test chars.................86 Table D-2. Proximate and Ultimate Analysis of Pittsburgh combustion test chars...................86 Table D-3. Proximate and Ultimate Analysis of Dietz combustion test chars..........................87 Table D-4. Proximate and Ultimate Analysis of Beulah combustion test chars........................88 Table D-5. Proximate and Ultimate Analysis of AAW Dietz and Beulah combustion test chars...........................................................................................................................................88

xii ACKNOWLEDGEMENTS

First, I wish to express my deep sense of gratitude to my advisor, Dr. Sarma V. Pisupati for his constant support, guidance and motivation. The reason I have been able to complete this work successfully is because of the essential skills of critical thinking and analysis and technical writing he has helped me imbibe over the course of my research. Secondly, I thank the Department of Energy and Mineral Engineering and Dr. Yaw D. Yeboah for providing me support by means of Teaching Assistantship throughout my Masters. I am grateful to my committee members, Dr. Yaw D. Yeboah and Dr. Sharon Miller for their inputs in refining my thesis. Next, I thank the Energy Institute for permitting me to use their facilities and equipment. I acknowledge the support and cooperation of Ron Wincek, Keith Miska, Ron Wasco, Tom Motel, Magda Salama and Henry Gong for helping me perform my analysis and experiments and troubleshooting problems. I also acknowledge the contribution of research group members, Vijay, Nari and Aime and Research Associate, Dr. Nandakumar Krishnamurthy towards my research my means of evoking interesting technical discussions. I am grateful to Dr. Prabhat Naredi, previous research group member whose thesis was also based on oxy-fuel combustion. His thesis helped me get started with my research, and in gaining a deeper understanding of the technology. I owe a special thanks to my group of friends particularly my roommates, Githin and Vivek Raja for providing me a life away from my lab and making me feel at home. And finally I thank my parents, Sakthivale and Vijayalakshmi and sister, Priyankari for their constant love and support and belief in me.

Chapter 1 Introduction
During the 1990's, the need for reducing greenhouse gases was recognized. Among the greenhouse gases, CO2 emitted into the atmosphere from burning fossil fuels, contributes about 82% to "global warming"[1]. Carbon dioxide capture and storage (CCS) is now included in most OECD countries' energy policies and R&D programs as one of the strategies to mitigate carbon dioxide emissions from large emitters[2]. The power generation industry accounts for nearly 40% of the CO2 emissions[3]. Alternatives to conventional air-blown pulverized coal (PC) combustion, such as coal gasification, integrated gasification combined cycle (IGCC) and pressurized combustion are being investigated to enhance energy efficiency, reduce greenhouse gas and pollutant emissions, and minimize the size and capital cost of future coal-based power plants. However, one of the most promising near-term alternatives to conventional PC combustion is oxy-fuel combustion (atmospheric pressure PC combustion in mixtures of oxygen and recirculated flue gas)[4] and is being considered as one of the leading carbon capture technologies for the power generation industry[2, 5-7]. Conventional pulverized fuel coal-fired boilers use air for combustion in which nitrogen in air (approximately 79% by volume) dilutes the CO2 concentration in the flue gas. The capture of CO2 from such dilute mixtures by stripping using chemicals such as amines is relatively expensive. During oxy-fuel combustion, a combination of oxygen (typically greater than 95% purity)[8] and recycled flue gas (RFG) is used for combustion of the fuel. By recycling the flue gas, a gas consisting mainly of CO2 (~90%[9]) and water vapor is generated, which can be sequestered (or stored as solid CO2 in gas hydrates[10]) without stripping of the CO2 from the gas

2 stream[11]. The water vapor is usually removed before compression and storage of CO2 so that it does not condense and form acids. Alternatively, CO2 can be used for methane recovery from unmineable coal seams, enhanced oil recovery (EOR) from depleted oil wells or even for production of value added chemicals. A general flow process of oxy-fuel combustion is shown in Figure 1-1. An oxy-fuel PC fired power plant consists of four main units: oxygen generation, oxy-fuel (O2/CO2) combustion, flue gas treatment and CO2 recovery/disposal. Wet (hot) gas recycling rather than dry gas (cold) recycling has been shown to improve economy[12]. The H2O concentration in the flue gas stream is generally between 10-40% depending on fuel moisture content and whether it is wet or dry recycling[9]. The recycled flue gas is used to control flame temperature and to make up the volume of missing N2 to ensure that there is enough gas to carry the heat through the boiler[11].

Hot RFG

Cold RFG

Figure 1-1. General flow process for oxy-fuel combustion[11] *ASU - Air separation unit To produce high purity oxygen, air is separated into oxygen using four methods of separation: cryogenic separation method, adsorption method, membrane separation method and chemical separation method. Of these four methods, cryogenic separation yields the best results[12] for large quantities. Burdyny and Struchtrup[13] have recently proposed a new method wherein air is drawn into an O2/N2 separation membrane using vacuum to produce

3 oxygen enriched air and then using cryogenic distillation to produce high purity oxygen, thereby reducing total energy requirements. The advantages of oxy-fuel combustion include: (1) no need to separate CO2 from the flue gases, (2) improved boiler efficiency, (3) low power consumption due to small amount of flue gas is involved (20% lesser than conventional air combustion[9]), and (4) no denitrification and desulphurization requirement[14]. Although the SO2 concentration is found to be 3-4 times higher for oxy-fuel combustion due to accumulated effect of recycled flue gas and reduced gas volumes, the mass emission rate (lb/MMBtu) is significantly lower than that in conventional air combustion. The NOx concentration also increases due to reduced gas volume but the mass emission rate is significantly lower than that in conventional air combustion. Furthermore, most of the NOx is reduced due to recycle in the flame region of the furnace[9, 15]. Table 1-1 shows the typical flue gas composition for a conventional PC system and an oxy-fuel system. Table 1-1. Typical flue gas composition[16] Parameter Conventional PC Combustion O2 (%) CO2 (%) H2O (%) NOx (ppm) 3.2 14.65 5.85 154 3.1 69 27.5 82 Oxy-fuel PC Combustion

Another major advantage of oxy-fuel combustion is that the same furnace that is used for conventional combustion may be used for oxy-fuel combustion with an appropriate exhaust gas recycle ratio[17, 18]. Thus, with rising global need to control CO2 emissions, oxy-fuel combustion offers a quick and economic means to retrofit existing plants. This research attempts to understand the effect of coal rank on carbon conversion.

Chapter 2 Literature Review


The major difference between oxy-fuel combustion and conventional air combustion is the replacement of N2 with recycled flue gas (primarily CO2). This leads to changes in flame propagation speed, flame stability and flame temperatures[19]. Also, CO2 has a higher specific heat[8] and emissivity[15] compared to N2 which means that CO2 can absorb more heat and radiation as compared to N2 and affect the heat transfer characteristics within the boiler.

2.1 Gas and Particle Temperatures The gas temperatures are lower in O2/CO2 combustion than in O2/N2 combustion with similar O2 concentrations though higher O2/CO2 ratios produce similar temperature profiles as combustion in air. It has been shown that a 30%O2/CO2 mixture can produce matching gas temperature profiles to those of combustion in air[8]. Particles burn at lower temperatures and require longer combustion times in O2/CO2 compared to combustion in O2/N2[20] and this is due to the lower gas temperatures in an oxy-fuel environment caused by the higher specific heat of CO2 (1.6 times that of N2)[21]. The particle temperatures are also reported to be lower compared to those in air combustion. Since CO2 is a reactive gas, Hampartsoumian et al.[22] attributed the reduction in particle temperatures to endothermic char-CO2 reaction taking place in oxy-fuel combustion. An increase in furnace temperature by 200 K, results in an increase of over 100 K in both gas and char temperatures[20]. Many studies [8, 15, 20, 23] also have reported that char particle temperatures are similar in air and 30% O2/CO2 mixture. As the oxygen mole fraction in

5 the combustion medium is increased, char particle temperatures increase and combustion times decrease in both atmospheres. Single particle studies have shown an increase in particle temperature from 1,850 K at an O2 mole fraction of 0.2 to 3,200 K at an O2 mole fraction of 1.0. The same study also has shown that peak char particle temperatures for lower rank coals like lignites (2700 K) were similar, if not slightly higher than those for bituminous coals (2650 K)[20]. While char-CO2 reaction consumes carbon and increases carbon conversion, the reaction being endothermic, also reduces the temperature and therefore the reaction rate. Hence, the net effect depends on the balance between the two. The effect of this when burning different ranks of coal is important.

2.2 Flame Characteristics The flame is less bright in oxy-fuel combustion due to CO2 molecule's ability to absorb IR radiation[15] and the volatile flame temperature is reduced by as much as 200 K under oxyfuel conditions as compared to air[20]. Hjartstam et al.[24] and Smart et al.[25] have shown that with reduced amount of recycled flue gas, more intense combustion and higher flame temperatures could be achieved. A higher concentration of O2 in CO2 (35%O2/CO2 mixture) has been shown to produce flame characteristics similar to those of combustion in air[15, 19] and this is because higher O2 concentrations lead to higher flame temperatures[4]. Andersson et al.[26] observed no significance difference in total radiation (sum of particle and gas radiations) between oxy-fuel and air firing. They concluded from their study that the particle radiation constitutes a significant fraction of the total radiation and that the total intensities become similar for air and oxy-fuel firing as long as the gas temperatures are similar due to similar particle radiation in each case. Smart et al.[27] conducted radiation and convective heat transfer studies in a pilot-scale test facility and have shown that radiative heat flux is inversely related to the recycle ratio though

6 convective heat flux increases with increasing recycle ratio and that it was possible to have a working range of recycle ratios where both convective and radiative heat fluxes were comparable to air.

2.3 Ignition Characteristics Ignition is delayed in oxy-fuel combustion as compared to air combustion of similar oxygen concentrations[28-31]. To study the ignition characteristics in a CO2 rich atmosphere, Kiga et al.[17] measured the flame propagation speed of a pulverized-coal cloud in O2/CO2, O2/N2 and O2/Ar atmospheres using a microgravity combustion chamber. They found the flame propagation speeds in an O2/CO2 atmosphere were markedly lower compared to that in O2/N2 and O2/Ar which they attributed to the difference in specific heat (CO2 has the highest specific heat among all the gases) since the difference in thermal conductivity of each gas mixture was not large. Shaddix and Molina[28] reported that a higher O2 concentration in oxy-fuel combustion can produce matching ignition times to air combustion. Particle devolatilization is delayed in oxyfuel combustion[28] though it has been shown that the volatile yield is higher[32, 33] as compared to air combustion. In thermogravimetric studies, Rathnam et al.[32] attributed the higher volatile yield in a CO2 environment to the onset of the char-CO2 gasification reaction at 1,030 K, which significantly increased the mass loss. Zhang et al.[34] said that coal particle ignition occurs in oxy-fuel environments as a volatile cloud due to the accumulation of unburnt volatiles rather than as individual particles in air combustion. Qiao et al.[31] claim that the endothermic char-CO2 reaction is responsible for the higher particle ignition temperature in air compared to oxy-fuel combustion.

7 2.4 Effect of Pressure Pressurized oxy-fuel combustion has been shown to increase the net efficiency during oxy-fuel combustion due to lesser fuel requirements, lesser pollutants and CO2 emissions. Higher pressure reduces the distance of the volatile flame from the particle and increases the temperature of the flame. As a result, heat transfer from the flame to particle and the devolatilization rate and extent are increased. The pressure also affects the rate of gas phase reactions (combustion of CO to CO2) in the boundary layer in the char combustion stage. The combustion in the boundary layer also increases heat transfer to the particle, increasing the char reactivity[35]. Hong et al.[36] and Deng et al.[37] compared cases operating at atmospheric pressure and pressurized conditions and found a 3% increase in power plant net efficiency in the latter case. They performed a thermodynamic analysis to identify the performance difference in an ambient and pressurized boiler process and found that the net efficiency improvement of 3% was due to an increase of 4.14% in boiler efficiency, a decrease of 2.73% in steam cycle efficiency, and an increase of 1.61% in efficiency from auxiliary load decrease. They further stated that a pressurized system helps to recover more heat from the flue gas and aids purification and compression of a concentrated CO2 stream, thus increasing the efficiency.

2.5 Char Morphology Borrego and Alvarez[19] characterized chars obtained under both air and oxy-fuel environments. They found more vesiculated particles in oxy-fuel chars compared to air combustion chars while network structures and voids distributed throughout the particle's surface are more common in air combustion chars indicating a lower capacity of the bubbles (large voids) to coalesce. Li et al.[38] found similar characteristics in the chars generated under an oxy-fuel environment and air with oxy-fuel chars having thicker surfaces and compact pores, contributing to a reduction in fragmentation. Also chars obtained under air were found to be round in shape

8 with a limited amount of secondary devolatilization within the voids and hence were considered to be more isotropic compared to oxy-fuel combustion chars. In terms of reflectance, vitrinite reflectance increases for chars compared to the parent coal reflectance value due to volatile release in the furnace and carbon enrichment. Reflectance values are higher for oxy-fuel chars compared to air combustion chars. However, in higher O2 concentrations in oxy-fuel environments, the values for both oxy-fuel chars and air combustion chars were found to be similar[19]. A similar trend was noticed for inertinite reflectance with oxy-fuel inertinites having higher values compared to air combustion inertinites. Reflectance values showed a drop with increasing O2 concentration. An increase in reflectance values indicates an ordering of char structure and carbonization which could have a negative impact on carbon conversion. In general, it was found that for low O2 concentrations, the chars had intact walls and granular appearance while more extensively burned chars had large coalesced voids[19]. Since this study looks into the effect of coal rank, the above results may not be valid for every rank of coal due to difference in conversion rates in air and oxy-fuel atmospheres.

2.6 Char Conversion and Reactivity Char reactions can occur in three zones as shown in Figure 2-1. At low temperatures, the chemical reaction rate is slow compared to diffusion and the chemical reaction is the rate determining step (zone I). In zone II, the overall reaction rate is controlled by both chemical reaction and pore diffusion and this occurs at moderate temperatures. In zone III, mass transfer limitations in the boundary layer of the particle control the overall reaction rate and this occurs at high temperatures.

Figure 2-1. Plot of reaction rate vs. inverse temperature depicting three zones[39]

The unreacted shrinking core model shown in Figure 2-2 is the most simple and best macroscopic approach to model gas-solid reactions[39]. Reactions take place on the surface of the particle and the reaction front recedes towards the center of the particle with time as shown in the figure. As the reaction takes place by consuming carbon, an ash layer forms. The reactant gas has to diffuse to the particle and product gases have to diffuse away from the particle.

10

Figure 2-2. Schematic diagram depicting a shrinking particle with an unreacted shrinking core, and concentration profile of gas reactant[40] (*CAc is the concentration of gaseous reactant at the center of the particle, CAs is the concentration at the surface and CAb is the bulk concentration)

11 This sequential operation involves three basic processes: gas film diffusion, ash layer diffusion and surface reaction of the central core. The overall rate coefficient depends on these three basic resistances, any of which can be rate controlling[39]. Char reacts heterogeneously with oxidizer gas (O2/CO2/H2O) and the reactions are given by[41], C + 1/2O2 C + CO2 C + 2H2O CO 2CO 2CO + H2 Ho = -9.2*106 J/kg Ho = +1.439*107 J/kg Ho = +1.096*107 J/kg 2.1 2.2 2.3

The char reaction rate varies with reactant media having the following order, Rair >> RH2O ~ RCO2 > RH2. The catalytic effect of inorganic impurities (discussed in detail later in this chapter) and surface area have also been shown to play an important role[42]. Radovic et al.[43] stated that active surface area and not total surface area played a role in determining the reactivity and that oxygen chemisorption capacity of chars at 375 K and 0.1 MPa was a good indicator of the concentration of carbon active sites. In coal chars, active sites would include sites bonded to heteroatoms (principally H), nascent sites (created during pyrolysis and gasification), dangling carbon atoms (single bonded), edge carbon atoms (double bonded) and trigonally bonded basal carbon atoms. Temperature has the highest influence on reactivity[44] with rate increasing with increase in temperature. The reactivity of chars decreases with increasing severity of pyrolysis conditions and residence times[23, 45, 46] due to a decrease in active sites (i.e. feeder pores[47]) either by thermal annealing (T ~ 973 - 1,373 K) (the char becomes structurally ordered), a reduction in catalytic activity caused by species like CaO sintering or due to the loss of hydrogen and oxygen[39, 43]. Gale et al.[46] attributed the decrease in reactivity to the flattening,

smoothing, or ordering of carbon-layered planes during the depletion of non-aromatic components in the char matrix, thereby increasing the relative concentration of aromatic compounds as mass is released. This is further proved by the fact that up to 1,473 K, heat

12 treatment of chars that are treated with pulses of oxygen have higher combustion reactivity than chars prepared under inert atmospheres. The mechanism has been attributed to chemisorption and formation of intercalation compounds, limiting the extent of graphene layers stacking and rearrangement upon heat treatment[48]. Walker et al.[49] stated that the reactivity of US coals may be attributed to three primary factors: catalysis of gasification, active site concentration, and the accessibility of reacting gas to the active sites. They also stated that the presence of active catalysts and the possession of a high concentration of active sites are necessary but not sufficient conditions for the high reactivity of coal chars. Overall reactivity decreases[20, 42, 50] and characteristic temperatures (initial temperature, peak temperature and burnout temperature) increase[51] with an increase in coal rank with the reactivity of lignites being of the order 20-40 times that of anthracites[52] due to a higher quantity of feeder pores and reactive sites on the surface[47]. The rate of the gasification reaction is enhanced by the presence of alkali and alkalineearth metal salts or oxides[52-56] and their importance increases for lower rank of coal (C < 80%)[53]. Ye[56] studied the gasification of Bowmans coal with CO2 and steam and found the reaction to be strongly catalyzed by Na, K and Ca. The effectiveness of catalysis of the CO 2 gasification reaction by alkaline-earth elements is in the following order: Be Mg < Ba Sr < Ca[54]. Kapteijn et al.[54] have indicated that the oxide (CaO) is the active species in the CO2 gasification reaction. Radovic et al.[52] and Walker et al.[49] attributed the relatively high gasification reactivity of lignites primarily to the catalytic activity of highly dispersed CaO on the char surface. They also found that the char reactivity decreases with increasing pyrolysis residence time, caused by CaO sintering and subsequently decreasing its dispersion. The high initial dispersion of inherent catalysts on the char surface is due to the presence of abundant exchangeable cations (mainly Ca2+) on the carboxylic groups[43]. The initial reactivity of Ca around 1,000 K has been found to be equal to that of potassium, a well known gasification

13 catalyst. However with increasing carbon burn-off, it shows severe deactivation due to sintering[54]. Chen[57] studied the sintering behavior by varying the percentage of CaO (between 5-10%) in CaO-MgO-Al2O3-SiO2 glass-ceramics and showed the samples to sinter at temperatures around 1173 K. Miura et al.[53] have said that the decrease in reactivity is caused by a decrease in number of active carbon sites and in catalytic activity. Prinsloo et al. [55] conducted CO2 gasification studies on bituminous coals and found the rate of catalyzed gasification reactions to first increase with carbon conversion and then decrease at higher carbon conversion which was attributed to the collapse of particle structure, pore plugging, cation loss due to migration of Na into the pores or Ca sintering or reaction of alkali metals with mineral matter. Other possibilities include formation of intercalated compounds or stable aluminosilicates and vaporization[53]. The catalytic activity of CaO can be explained by the following set of reactions[54]: CaO + CO2 CaO . O + C C[O] CO CaO . O + CO CaO + C[O] 2.4 2.5 2.6

Kapteijn et al. [54] proposed that the rate determining step in the CO2 gasification of carbon is the release of CO from the carbon structure. The increase in site density can be achieved by the formation of more CO groups by an oxygen transfer mechanism. A more detailed mechanism represented in the work of Miura et al.[53] shows the importance of the edge carbon atoms as compared to the basal carbon atoms in oxygen adsorption on carbonaceous materials. The activation energy of the reaction does not change due to catalytic effect, which suggests that the catalyst only increases the number of active sites on the char surface[54]. The two types of reactions basically contributing to char conversion in oxy-fuel combustion are char-O2 and char-CO2[41] reactions. In general, it has been shown that higher amount of O2 in CO2 than in N2 is required for similar burnout[19] despite contribution from

14 char-CO2 reaction and this has been attributed to lower particle temperatures[21]. Char burnout times are similar for air and 30%O2/CO2 mixture with burnout increasing, higher the O2 concentration[4, 41, 58]. Burnout efficiency is the best under 30% O2/CO2 mixture followed by air and lastly 21%O2/CO2 mixture. Li et al.[38] and Li et al.[59] conducted their studies in a droptube furnace and found that the lower char reactivity under oxy-fuel environments as compared to air was the reason for lower burnout. Brix et al.[60] studied char combustion at high temperatures (~1,673K) and low oxygen concentration (3.1-3.7%) and obtained lower conversions under an O2/CO2 environment as compared to O2/N2 environment, which they attributed to the lower molecular diffusion of O2 in CO2 as compared to N2. Borrego and Alvarez[19] attributed the lower conversion in oxy-fuel environments to CO2 forming cross-links on the surface of the char particle. Studies conducted by Varhegyi et al.[61, 62] showed that coal combustion in O2/CO2 atmosphere, the char-CO2 reaction has a much lower rate than the char-O2 reaction and that the net reaction rate was proportional to the partial pressure of O2 (PO2). They also stated that CO2 does not participate in the elementary reactions of oxidation and that any changes in ambient CO2 concentration do not affect the concentration of evolved CO2 at the molecular level. The CO2 only influences secondary reactions of the CO and CO2 formed (a detailed mechanism is shown later in the chapter). Maximum reaction rates are reached when increased concentrations of O2 are used. Increased O2 concentrations also lead to increased devolatilization rates due to closer proximity of the volatiles flame to the coal particle and the increased temperature of the volatiles flame[4]. Murphy and Shaddix[4] said that the best fit to their combustion results from an entrained flow reactor were shown by the n-th order Langmuir-Hinshelwood equation given by, 2.7

15 where k1 and k2 are constants dependent on temperature. The apparent reaction order, n was found to vary between 0.1 for near-diffusion-limit oxygen-depleted conditions to 0.5 for oxygenenriched conditions. The dependence of reaction rate on temperature and fractional burnoff is given by, dj/dt = Aj exp(-Ej/RT)g(PO2,PCO2)fj(j) 2.8

where Aj and Ej are pre-exponential and activation energy, respectively; g and f are empirical functions; and g(PO2,PCO2) is proportional to PO2 and does not depend on PCO2[61]. The studies by Varhegyi et al.[61, 62] were conducted at high pressure and moderate temperatures (~1,223 K) and they have reported no influence of the char-CO2 reaction as stated above. Liu et al.[44] modeled the char-CO2 gasification reaction at high temperature and pressure by extrapolating char reactivity data obtained at moderate temperatures. They showed that for low-rank coal char at 1,123 K, both the apparent and intrinsic reactivity increased with CO2 partial pressure. They also showed that the apparent reaction rate increased two orders of magnitude as the coal rank decreased and that the activation energy of the char-CO2 gasification reaction generally decreased as the coal rank decreased. The char-CO2 reaction is composed of the following reactions: Cf + CO2 C(O) CO + Cf CO C(CO) 2CO + C(O) CO2 + 2Cf C(O) + CO 2.9.1 2.9.2 2.9.3 2.9.4 2.9.5

CO2 + C(CO) CO + C(CO)

The intrinsic reaction rate in the form of the Langmuir-Hinshelwood expression is given by, 2.10

16 where k1, k2, k3 and k4 are temperature dependent rate constants which can be represented by an Arrhenius type equation, ki = Ai e-Ei/RT where Ai is the pre-exponential factor and Ei is the activation energy. Rathnam et al.[32] conducted reactivity studies in a thermogravimetric analyzer after generating char in a drop-tube reactor. They also showed an increase in weight loss at temperatures exceeding about 1,030 K due to the char-CO2 reaction. The authors also conducted combustion tests in the drop-tube reactor on four coals but the drawback of this study was that the authors did not provide answers as to why two of their selected coals showed higher conversion in an oxy-fuel environment as compared to air. The selected coals did not represent a wide range in rank thus not giving a clear rank effect. Saastamoinen et al.[35] conducted pressurized oxy-fuel combustion under different concentrations of O2 and CO2. They reported that when the gas oxygen concentration is low and the carbon dioxide concentration is high, both char oxidation and gasification may be important in the char mass reduction. This is more significant at high temperatures when the char oxidation rate is limited by transport of O2 to the particle surface (diffusion limitations). Under these conditions, CO oxidation (in the boundary layer of the particle) is faster, which further consumes the remaining small O2 content on the surface and the char oxidation rate increases the CO2 concentration on the surface of the char particle. They also claim that under certain conditions (high CO2 concentration, high gas temperature, high gas pressure, large particle size (diffusion control) and porous char particles (large internal area open to CO2)), the char gasification reaction may become important compared to the char oxidation reaction. However the char-O2 reaction is many orders of magnitude faster than the char-CO2 reaction, when the O2 concentration at the particle's surface is substantially reduced by CO oxidation at the boundary layer, the gasification reaction can compete favorably with the char-O2 reaction in the overall heterogeneous reaction 2.11

17 rate. Shaddix and Murphy (as referenced by Buhre et al. [11]) found that in oxygen enriched combustion, CO2 gasification of char becomes important at practical temperatures. The authors measured particle burning rates vs. temperature and saw little difference in a O2/CO2 or O2/N2 atmosphere. They also stated that a decrease in burning rate in an O2/CO2 environment was due to a decrease in O2 diffusion through the particle boundary layer though the authors did not provide any conclusive evidence that this was the reason. Zhang et al.[34] investigated the combustion of brown coal in O2/N2 and O2/CO2 mixtures. They found that up to 25% of the nascent char may undergo gasification to yield CO to improve the reactivity of the local fuel/O2 mixture. The subsequent homogeneous oxidation of CO released extra heat for the oxidation of both volatiles and char. The authors did not provide any concrete experimental or theoretical evidence as to how they concluded that 25% of the nascent char undergoes gasification. Li et al.[59] produced chars under a CO2 environment in a drop-tube reactor at 1673 K and gasified surfaces were analyzed using an SEM. They also reported that under oxy-fuel conditions when the oxygen partial pressure is low in the later stages of combustion, the gasification of unburned char would have a significant effect on the char conversion. The authors conducted combustion tests on three Indonesian low-rank coals and all three coals showed higher conversions in air as compared to 21%O2/79%CO2. They do not provide an explanation for this though they showed the occurrence of the gasification reaction through SEM images. Naredi[21] has shown theoretically that there is an exponential increase in the rate of the char-CO2 reaction beyond 1,800 K with very little influence below that temperature. Also he claimed that at higher temperatures, an increase in rate caused by the char-CO2 reaction possibly compensates for a decrease in O2 availability and lower particle temperatures during oxy-fuel combustion.

18 When the temperature exceeds 1,700 K under entrained flow conditions, the char reactions are limited by mass transfer into the porous structure of the char particles[44]. An effectiveness factor , which is the ratio of the actual rate per unit internal surface area to the rate attainable if no pore diffusion resistance existed should be used. The apparent reaction rate is then calculated by, Rapp = RinS 2.12

where S is the internal surface area. The effectiveness factor primarily depends on particle temperature and size and can be calculated by the well known Thiele modulus approach (sample calculations shown in Appendix A)[44]. The Thiele modulus depends on char properties, reaction conditions and particle shape. The basic form of the Thiele modulus, derived for solid-gas catalytic reactions, includes reactant concentration, rate constant, the effective diffusivity and a shape factor (F) which determines the geometry of the particles, given as[63],

2.13 The effective factor is then calculated as[21], 2.14 The effectiveness factor is 1 in zone I, and ranges from 1 to 0 in zone II and is equal to 0 in zone III[63]. It has been shown for the char-CO2 reaction, when the temperature is raised from 1,400K to 2,000 K, the calculated effectiveness factor decreases from unity to 0.13 indicating that the reaction transfers from chemical reaction limited to pore diffusion limited. Figure 2-3 shows the general trend of effectiveness factor as a function of temperature.

19

Figure 2-3. General shape of the effectiveness factor vs. temperature curve[63] An effective diffusivity coefficient, Deff is significant for the calculation of the effectiveness factor and is strongly dependent on the pore size within the particle. The total porosity of the char particle is written as the sum of the macro-, meso- and micro-porosity, which is written as, T = a + e + i where, T, a, e, and i are total, macro-, meso- and micro-porosity respectively[44]. Hodge[63] has shown that at high temperatures (zone II conditions), the activation energy of the char-CO2 reaction is half that measured under zone I conditions. Based on the results of the study, the author also stated that the transition from zone I to zone II depended on char morphology and reactivity but it could be safely assumed that under entrained flow conditions, it occurred at temperatures >1,473 K. Tree et al.[64] modeled both char-O2 and char-CO2 reactions in their study on oxy-fuel combustion. They obtained the oxidation and gasification rate parameters from various literature 2.15

20 sources[65-67]. They included the gasification reactions in their model although gasification reactions are much slower than oxidation reactions because of the high concentration of CO2 in oxy-fuel combustion. Figure 2-4 shows a comparison of oxidation and gasification reactions for various ranks of coal and in general it is claimed that rates increase with decreasing rank. Other studies have shown a similar effect of an increase in rate with CO2 partial pressure[21, 63, 68] and decrease in coal rank with the reaction order being around 0.7-0.8[63]. Goetz et al.[66] conducted char gasification studies on four coals: a lignite, a subbituminous coal, a high volatile A coal and a high volatile C coal. They found both combustion and gasification reactivity of the chars to follow the order: lignite > subbituminous > hvCb > hvAb and that the gasification reactivity of the least reactive char was lower by a factor of 10 compared to the most reactive char. They also found that the combustion and gasification rates were a strong function of pore structure, temperature and reactant gas concentration. Out of all the coals they studied, only the hvAb coal swelled though it showed the least gasification reactivity. Figure 2-5 shows results from a study by Roberts and Harris[69] where low temperature kinetics data from char-O2, charCO2 and char-H2O reactions have been extrapolated to high temperatures approaching mass transfer limitations. The exponential increase in rate of the char-CO2 and char-H2O as temperature increase can be clearly seen.

21

Figure 2-4. Comparison of char reaction rates with O2 (oxidation) and CO2 (gasification)[64]

A computational study conducted by Mann and Kent[70] for full scale boilers showed that with the incorporation of the char-CO2 and char-H2O reactions in their model, the accuracy of the burnout predictions improved and this was thought to be due to the significant role played by these species in oxygen deficient regions. Also, they found that the higher concentration of CO2 in the furnace in comparison to H2O led to greater mass loss from the char-CO2 reaction in comparison to the char-H2O reaction. An important conclusion of this study was that the high temperatures in the furnace were sufficient to make the transfer rates the important controlling factor but chemical kinetics remained important for the char-CO2 and char-H2O reactions. The drawback of the results of this study was that the conclusion was inferred based on the model results and there was no substantial evidence provided to prove that chemical kinetics was an important factor at high temperatures for the char-CO2 and char-H2O reactions.

22

Figure 2-5. Low temperature TGA kinetic data extrapolated to high temperatures ( : O2, : steam, : CO2)[69]

2.7 Char conversion Rate Parameters Estimation Le Manquais et al.[71] compared chars obtained at high heating rate (drop-tube reactor) and low heating rate (thermogravimetric analyzer (TGA)) for their reactivities. They stated that conflicting trends have been reported when TGA rate parameters were applied to pulverized coal combustion. They found that the drop-tube reactor chars showed an increased burnout propensity while moving from zone II to zone III. Char morphologies were different for both types of chars, with the TGA chars resembling the raw coal with an undeveloped pore network and the drop-tube chars being highly porous, swollen and with a high surface area. The drop-tube chars had an order

23 of magnitude higher reactivity as compared to the TGA chars as their higher porosity reduced mass transfer limitations and would more closely resemble pulverized coal combustion. There are a number of ways to extract rate parameters from experimental data leading to a wide scatter in pre-exponential factor and activation energy even for the same rank of coal[21]. Table 2-1 shows how the rate parameters can vary based on the method used. Table 2-1.Char-CO2 reaction activation energy values reported in literature
Ref. [72] Method X=0, X=0.5 Operating Temperature 1273-1673 K Coal Rank bituminous subbituminous lignite [73] Shrinking core 973-1173 K lignite subbituminous high-volatile bituminous lignite [74] Random pore model 1173-1333 K subbituminous high-volatile bituminous Activation Energy (kJ/mol) 62 82 98 146 151 155 79 147 180

Jenkins et al.[47] used the maximum value (shown in Figure 2-6) in the rate profile to compare char reactivity. Scaroni et al.[75] determined rate parameters using averaged rate values during different extents of burnoff given by, 2.16

where Ru = (dW/dt)/W is the reactivity determined on the basis of unburnt carbon. They also felt that determining rate parameters using the maximum value was not appropriate because of an extremely short rectilinear region. Fletcher et al.[46] and Dugwell et al. [45] determined rate

24 based on fixed value of burnoff. Naredi and Pisupati[23], noticed a variation in activation energy over the 10-30% burnoff range indicating that the char burning is not always under kinetic control. They attributed the initial part of the rate curve up to the maximum to intraparticle diffusion limitations and also due to lower partial pressure of the reaction gas and so they estimated the rate parameters from the slope of the region after the occurrence of the maximum. They also noticed that the rate parameters determined using this method closely matched those determined using the maximum and hence concluded that rate parameters could be extracted at any conversion level after the occurrence of the maximum. Figure 2-6 shows a typical rate profile from the study of Naredi and Pisupati[23] depicting the occurrence of the maximum. It was decided to utilize with the method of estimation of rate parameters adopted by Naredi and Pisupati based on their results showing that the initial part of the curve up to the maximum was subject to diffusion limitations.

maximum

Figure 2-6. Reactivity profiles in air at 673 K for chars obtained under different pyrolysis conditions[23]

25 2.8 Process Feasibility and Economics There have been many recent studies[7, 16, 36, 37, 76-80] conducted on evaluating oxyfuel combustion on the utility scale. Pak et al.[76] theoretically compared an oxy-fuel combined cycle system with CO2 liquefaction and utilizing low pressure steam to a conventional steam turbine power generation system utilizing low pressure steam. They found that the oxy-fuel system could generate 2.03 times greater electric power than the conventional system with a net CO2 reduction of 180 kilotons/year and an exergy efficiency of 54.2%. The system was evaluated to be economically feasible with it surpassing the conventional system if a CO2 credit of $30/ton was applied to captured CO2. Xiong et al.[78] also showed that the cost of electricity for an oxyfuel plant would be greater (1.5-1.7 times) than a conventional PC combustion plant though, a CO2 sale price of $17-22/ton would level the costs. However, the system has a 2.41% degradation of net power generation efficiency. Liszka and Ziebik[77] also theoretically compared an oxyfuel coal-fired power unit to a conventional unit and saw an efficiency drop of 10.89% due to CO2 compression and purification. The said that the high impact of CO2 compression system on the overall efficiency could be limited by sub-critical liquefaction and better heat regeneration systems. Zhou et al.[81] computed process requirements for a boiler retrofit with minimal impact on thermal and emission performance and found that an optimal wet flue gas recycle ratio depended on the type of coal and exit oxygen concentration and was typically around 0.7-0.75. They also said that dry flue gas recycle could be used to enhance flame temperatures in oxy-fuel combustion. Their computational fluid dynamics model showed oxy-fuel combustion and air combustion to be very similar in terms of thermal characteristics. Hjartstam et al[24] showed in their modeling study that with appropriate adjustment of the recycle rate, desired combustion stability and structure of coal-fired oxy-fuel flames may be achieved without any significant impact on the emission level.

26 2.9 Summary It has conventionally been thought that oxy-fuel combustion produced lower conversions[19, 38, 59, 60] compared to air combustion due to lower particle reactivity, particle temperatures and heating rate. The char-CO2 reaction has been the focus of many studies[21, 30, 32, 35, 44, 59, 61-63, 66, 70]. There is a lot of scatter in literature regarding the contribution of CO2 to char conversion in an oxy-fuel environment. Certain studies[61, 62] show no effect of CO2 on char conversion due to the gasification rate being much slower than the oxidation rate. There are recent studies[32, 34, 59] that have observed the char-CO2 reaction though have not provided substantial evidence to prove the same. Both the char-O2 and char-CO2 reactions are influenced by temperature and intrinsic reactivity of the coal and hence, there is expected to be an effect of coal rank and temperature on conversion. The oxidation and gasification reactivities increase with decrease in rank of coal and this has been shown theoretically and experimentally by Tree et al.[64] have and Goetz et al.[66], respectively. Hodge[63] has shown that the activation energy of the char-CO2 reaction under zone II conditions is half that under zone I conditions and Liu et al.[44] have stated that the activation energy of the char-CO2 reaction generally decreases with increase in coal rank. Roberts et al.[69] have shown the reactivity of the char-CO2 and char-H2O reaction to increase exponentially with an increase in temperature as compared to the char-O2 reaction. From the literature review, it has been shown that under certain conditions of high temperature, high CO2 partial pressure, large particle size and porous char particles, the gasification reactions can compete favorably with the oxidation reactions in the overall heterogeneous rate[35]. Mann and Kent[70] have shown the importance of the char-CO2 reaction in oxygen deficient regions of the boiler and how computational model predictions improve on incorporation of the char-CO2 reaction. Naredi[21] has shown numerically that higher conversions can be obtained by burning coals more reactive to CO2 and also that there is an

27 exponential increase in the rate of the char-CO2 reaction at temperatures beyond 1800 K. Based on the above discussion, some of the questions that this thesis tries to address are: Would there be any difference in using high furnace temperatures? Particles generally reach higher temperatures compared to the gas temperature and using higher furnace temperatures, leads to higher particle temperatures. It is believed that other studies[61, 62] did not report any influence of char-CO2 reaction because of the low furnace temperature (<1273 K) and heating rate used in their studies. What role does coal rank play during oxy-fuel combustion and how do carbon conversions differ under air and oxy-fuel atmospheres? Is there any role played by alkali and alkaline-earth metal salts and oxides in catalyzing the char-CO2 reaction, leading to enhanced carbon conversion under oxy-fuel atmospheres? The gasification catalytic activity has been studied previously by several authors including Walker et al.[49] and Radovic et al.[52] but the possibility of catalyst deactivation at high temperatures due to sintering has also been shown[54]. The aim of this work was to investigate catalytic activity under high temperature conditions and to clearly identify rank effect and catalytic effect on the char-CO2 reaction under such conditions. If there is a difference in carbon conversions between air and oxy-fuel combustion, what is the primary contributing factor to this? What is the reason for enhancement in contribution of char-CO2 reaction towards char conversion at high temperatures and how well does it compete with the oxidation reactions in the overall carbon conversion?

28

Chapter 3 Problem Statement

3.1 Methodology The aim of this work was to provide an experimental and theoretical analysis of the effect of coal rank and char-CO2 reaction under oxy-fuel conditions. To investigate the contribution of the char-CO2 reaction in increasing overall conversion under oxy-fuel conditions, combustion tests were conducted in a lab-scale drop-tube reactor with four coals (of different ranks) at three different residence times at 1873 K. The chars generated under air and oxy-fuel environments were analyzed for carbon conversion. Chars at conversion approximately equal to the ASTM volatile content of the parent coal were also generated in a reactive environment (21% O 2/79% CO2). This was done to ensure maximum removal of volatile matter, to obtain representative coal char samples. In order to quantify the result of the combustion tests and to obtain a more fundamental understanding of char-O2 and char-CO2 reactivity for the four coals, modified burning profiles were obtained on the parent coals using a bench-scale thermogravimetric analyzer (TGA). Isothermal reactivity profiles were also obtained on the char samples. The generated intrinsic char-O2 and char-CO2 rate parameters were extrapolated to high temperatures taking into consideration diffusion limitations using a theoretical model (Appendix A). The char samples were characterized by using proximate analysis, ultimate analysis and surface area analysis to identify differences in structure and composition. Since catalytic activity due to ionexchangeable cations is well recognized in low-rank coals, samples were washed with ammonium

29 acetate to remove the ion-exchangeable cations and to compare the effect of coal rank on a catalytic activity-free basis. Catalytic activity is generally high at lower temperatures in a TGA but in an actual boiler at high temperatures, its influence is lower. To investigate their importance at high temperatures on reactivity, combustion tests were performed on ammonium acetatewashed coal samples and the carbon conversions were compared with those obtained for the parent coals in a DTR. Also, the char particle temperature in oxy-fuel atmospheres was theoretically modeled. The effect of varying various parameters such as reactive gas partial pressure, gas temperature and coal rank on particle temperature was also studied. In summary, the main objectives of this study were: To include the char-CO2 reaction at high temperatures in computing the carbon conversion, specifically in an oxy-fuel environment. To study the role of coal rank and the char-CO2 reaction under simulated boiler conditions (high temperature, high heating rate, and similar residence times) in air and oxy-fuel environments. To determine the role of catalytic activity of alkali and alkaline-earth metal salts and oxides under TGA and at high temperatures. To gain a better understanding of the primary factor contributing to the difference in carbon conversion in air and oxy-fuel environments and to theoretically model the phenomenon.

3.2 Hypothesis At high temperatures (~1,873 K) in a drop-tube reactor, combustion in oxy-fuel environments produces higher carbon conversion than that in air for low-rank coals. This is due to the increasing contribution of the char-CO2 reaction at high temperatures along with the char-O2

30 reaction thereby, increasing the overall carbon conversion. The reactions take place under zone II conditions and due to high temperatures and high CO2 partial pressure in oxy-fuel combustion, the gasification reactions become important and compete favorably with the oxidation reactions. For low-rank coals, the rate of the char-CO2 gasification reaction is much higher than that for higher rank coals. The activation energy of the char-CO2 reaction is higher than the char-O2 reaction and hence, it is more sensitive to an increase in temperature with the rate increasing exponentially at high temperatures. Hence, oxy-fuel combustion may be more suitable for lowrank coals in decreasing the unburnt carbon because of their increased reactivity in an O 2/CO2 environment.

31

Chapter 4 Experimental Details


This chapter describes the various lab equipment used and the procedures followed.

4.1 Coal Sample Preparation A suite of four coals of different ranks was selected from the Penn State Coal Sample Bank: a low volatile, Pocahontas #3 Seam (DECS-19) bituminous coal; a high volatile, Pittsburgh Seam (DECS-34) bituminous coal; a subbituminous, Dietz Seam (DECS-38) coal and a Beulah Seam lignite (DECS-11). The coals were dried in an air drying oven at 333 K until the surface moisture weight loss was less than 0.1%/hr. The coals were then ground in a ball mill and sieved to obtain particles in the size range -170+200 mesh (74 - 88) that were used in the drop-tube reactor for pyrolysis and combustion studies. The reason for selecting a narrow size range was to limit the effect of particle size as reactivity varies greatly with particle size.

4.2 Drop Tube Reactor The drop-tube reactor (DTR) used in this study was a laminar flow furnace with controlled wall temperatures and particle residence times, that simulated flow conditions, high heating rate and temperatures of an actual PC boiler. The particle residence time is a function of gas flow rates, particle size and insertion length of the probe. Figure 4-1 shows the schematic of the DTR.

32 The reactor is a single zone, electrically heated furnace capable of being operated at a maximum temperature of 1,873 K. The reactor tube is made of high purity alumina refractory material positioned vertically. Six U-shaped Kanthal-super heating elements are attached on the wall. A rotary type feeder, Acrison GMC-60 supplies coal into an entrained flow of a primary gas through a water-cooled injection tube. The position of the tip of the injector is at the level of the bottom of a mulliteTM flow-straightener which supplies the preheated secondary gas to the

Figure 4-1. Schematic of Drop tube reactor[21]

33 furnace. Secondary gas from two inlets enters the furnace through the top of the preheater and exits the flow straightener along with the coal-laden primary gas. A char collection probe is inserted from the bottom of the furnace and positioned at a desired height. Char particles are collected isokinetically through the water cooled collection probe by using a vacuum pump. The particles are collected on a Whatman 541 ashless filter paper and the gas is then passed through a condenser to remove water vapor before being sent to a continuous emission monitoring (CEM) system. Calculations were performed to ensure laminar flow condition in the DTR and the Reynolds number was observed to vary between 127 and 1,774 for a mean gas velocity between 0.36 m/s to 5.08 m/s assuming air as the gas.

4.2.1 Combustion Test Conditions Three gas compositions (by volume) were used: 21%O2/79%N2 (air), 21%O2/79%CO2 (oxy-fuel combustion) and 30%O2/70%CO2 (enhanced oxy-fuel combustion). The combustion tests were conducted to measure carbon conversion of the coals at three residence times. These were the first set of tests to be conducted on the coals to investigate the rank effect. The residence time was varied by changing the location of the collection probe and, keeping the gas flowrates constant. The gas residence times investigated were: 0.8 s (0.254 m), 1.2 s (0.406 m) and 1.7 s (0.559 m). The residence time was calculated from the volumetric flow rate of the gas taking into account gas expansion at high temperature. The length is the distance of the probe from the top of the furnace. The typical residence times of PC boilers are shown in Table 4-1 and the results obtained in this study would be practically applicable to boilers in the size range 60-210 MW. The other parameters used in the combustion tests are given in Table 4-2.

34 Table 4-1. Typical PC boiler residence times*[82] Boiler size (MW) 60 MW 110 MW 210 MW 500 MW *mid-burner region to furnace outlet Table 4-2. DTR combustion test parameters Coal feedrate (g/min) Total gas flowrate (l/min) Primary gas flowrate (l/min) Secondary gas flowrate (l/min) Pre-heater temperature (K) Furnace temperature (K) 0.4 10 4 6 873 1,873 Residence time (s) ~0.8 ~1.2 ~2-2.5 ~3.5

4.2.2 Devolatilization Test Conditions To measure the reactivity of the coals used in the study in a thermogravimetric analyzer (TGA), chars were generated in the DTR under an oxy-fuel atmosphere (21%O2/79%CO2) to simulate conditions such as high heating rate and particle temperature of a real boiler. The conditions used were the same as those shown in Table 4-2 except for the residence time. A series of tests at short residence times were conducted for each coal sample to obtain a char sample that had a carbon conversion (%) approximately equal to the ASTM volatile matter of the coal.

35 4.3 Microproximate Analysis A Perkin-Elmer 7 thermogravimetric analyzer (TGA) was used to conduct microproximate analysis on the DTR feed coal and chars generated to determine the char conversion. This was done because of the limited amount of sample generated and sample requirements for further analysis. The procedure adopted is similar to that followed by Man et al.[83] in their study. About 4-5 mg of char was placed in the TGA pan for analysis. The furnace was held at room temperature in an inert atmosphere (N2 at 100 cc/min) for 1 hour. The sample was then heated to 380 K at the rate of 5 K/min and held at that temperature for 2 hours to drive off the moisture. It was then heated to 1,223 K at the rate of 50 K/min and was held at that temperature for 7 minutes to expel the volatile matter. The sample temperature was then decreased to 873 K and the gas was switched to air to initiate oxidation. It was then heated at the rate of 10 K/min to 1,023 K and held for 2 hours to determine the ash yield. The carbon conversion of the chars generated was calculated by the ash tracer technique using the formula[84]: 3.3.1 where W is the carbon conversion (%), A0 is the ash content of feed coal (%, dry basis) and A1 is the ash content of the collected char (%, dry basis). A sample calculation is shown in Appendix B. The underlying assumptions of the ash tracer technique are that coal ash is conserved within the char and that the ash fraction in the coal is not affected by temperature-time history of the particle. This requires that the ash species are not volatilized, or at least the extent of

volatilization which occurs in the experiment is the same as that which occurs under the conditions of the standard ash test so that carbon conversion may be calculated from an ash balance of the coal[84].

36 4.4 Determination of Char Reactivity Char-O2 and char-CO2 reactivity tests were conducted in the TGA. These tests were carried out to try to explain the difference in carbon conversions obtained during the DTR combustion tests in air and oxy-fuel atmospheres for the various coal samples. The procedure was adapted from the work of Naredi[21] who performed several preliminary tests (on the same TGA) to make sure that the operating parameters eliminated mass transfer limitations. About 4-5 mg of sample was placed in a platinum crucible for analysis. The furnace was held at room temperature in an inert atmosphere (N2 at 100 cc/min) for 30 min. The sample was then heated to 383 K at the rate of 10 K/min and held there for 1 hour to drive off moisture. It was then heated to 1,223 K at the rate of 50 K/min and held there for 7 minutes to drive out the residual volatiles. The sample was then cooled down to the required reaction temperature at a rate of 20 K/min and held there for 10 min to establish a thermal equilibrium. The gas was then switched from N2 to O2 or CO2 to initiate char oxidation or/gasification. The weight of the sample was recorded every 30 s until ~90% (dry and ash free (daf)) weight loss was achieved. After each experimental run,

conversion-time data was converted into rate of weight loss normalized with respect to initial mass (m0 (daf)) to construct reactivity plots with the rate calculated as, 3.4.1 For the char-O2 reactivity tests, reaction temperatures of 673K, 698K and 723K were used and for the char-CO2 reactivity tests, reaction temperatures of 1,123K, 1,148K and 1,173K were used. Plots of ln k vs. 1/T were generated and rate parameters were extracted after the occurrence of the maximum in the reactivity plots. The slope of the line gives (-E/R) and the intercept gives ln A. A and E are the pre-exponential factor and activation energy, respectively. A example calculation is shown in Appendix C.

37 4.5 Modified Burning Profiles Burning profiles of fuel samples indicate relative combustion rates and heat release profiles in a boiler when samples are heated non isothermally at a constant rate from room temperature to about 1,273 K. The procedure was developed by Wagoner and Duzy[85] to determine the relative rate of combustion and heat release profiles. The burning profiles were generated using the TGA based on the procedure outlined by Pisupati[51]. The term 'modified' is used as the existing procedure which is performed under air has been adopted for runs under oxyfuel environments. About 5 mg of sample was placed in the crucible for analysis. The furnace was then raised to enclose the sample which was held at room temperature in an inert atmosphere (N2 at 100 cc/min) for 30 min. The gas was then switched to air or oxy-fuel atmosphere (21%O2/CO2) and the sample was heated to 1,273 K at the rate of 10 K/min. While performing tests under oxy-fuel conditions, a special blended gas (21%O2/CO2) cylinder was used to supply the gas. Data was recorded every 60 s and then the burning profile was generated by plotting derivative weight % change (%/min) vs. temperature (K). From these profiles the initial temperature (IT), peak temperature (PT) and burnout temperature (BT) were obtained to compare various coals. The initial temperature (IT), is arbitrarily defined as the temperature at which the weight loss exceeds 0.1 %/min after the initial moisture peak. The peak temperature (PT) is defined as the temperature at which the weight loss is maximum. Burnout temperature is arbitrarily defined as the temperature at which the rate of weight loss decreases to 1.0 %/min towards the end of conversion. Dmax is the value of derivative weight % at the peak temperature (PT). The relative peak mass loss rates and the temperature range provide the relative rates of combustion of various fuels.

38 4.6 Compositional Analysis 4.6.1 CHN Analysis The carbon, hydrogen and nitrogen content of the feed coals and char samples were measured using an analyzer, LECO TruSpec CHN. First, blank calibration of the instrument was performed by running three blank samples. Then, a coal CHN calibration standard, AR 1706 was run 3-5 times to calibrate the instrument. A furnace temperature of 1,223 K and afterburner temperature of 1,123 K was maintained. Also, a system check and leak check were performed to ensure proper functioning of the instrument. About 0.1-0.2 g of sample was weighed into a tin foil which was compressed into a ball and pushed into the furnace pneumatically. After complete combustion, the flue gases were passed through various detectors to measure carbon, hydrogen and nitrogen contents and vented out. The system was purged and the baseline was established before each run. 4.6.2 Sulfur Analysis The total sulfur content of the feed coals and the char samples were measured using a sulfur analyzer, LECO SC 132. A few blank runs were performed using a random coal sample to saturate the pores of the sulfur adsorbent. Then, the calibration sample was run about 3-5 times until the value stabilized and a new calibration curve was generated. About 0.1-0.2 g of sample was placed in the sample boat for analysis. The system was purged and the baseline was established before each run. The sample boat was then pushed into the furnace to combust the sample in an atmosphere of enriched air at 1,673 K. The flue gases were then passed through a series of adsorbents and particle filters before passing through the sulfur detectors to measure the total sulfur content. The oxygen content was then determined by difference (100 - %C - %H - %N - %S). The higher heating value (HHV) for the coal samples was provided by the Penn State Coal Sample Bank.

39 4.7 Surface Area Analysis This analysis was performed on the DTR generated four coal char samples and one ammonium acetate-washed sample so as to use the pore area analysis as an empirical tool to explain the TGA char reactivity profiles. The surface area of the sample was determined using Micrometrics ASAP 2020 surface area analyzer and the analysis was performed by personnel at the Materials Research Laboratory (MRL), Penn State. About 0.4-0.5 g of sample was outgassed for a minimum of 12 hours at 383 K. The sample temperature was decreased to 77K and nitrogen gas was then introduced in controlled increments. After each dose, the pressure was allowed to equilibrate for 45 s. The volume of gas adsorbed at each pressure (at constant temperature) defines an adsorption isotherm, from which the quantity of gas required to form a monolayer over the external surface area of the sample and its pores is determined. With the area covered by each gas molecule known, the surface area can be calculated. The CO2 surface area (micropore surface area) was calculated similarly using the Dubinin-Astakhov method at 273 K. To determine the porosity, pore size and pore distribution, the process was extended to allow the gas to condense in the pores by increasing the pressure until saturation pressure was reached, at which time all pores are filled with liquid. The adsorptive gas pressure was then reduced incrementally to allow the condensed gases to evaporate from the system. Evaluation of the adsorption and desorption branches of these isotherms and the hysteresis between them reveals information about the pore size, pore volume, pore shape, and pore area.

4.8 Major/Minor Oxide Analysis of Ash The oxides (SiO2, Al2O3, TiO2, Fe2O3, P2O5, MgO, CaO, Na2O, K2O and SO3) present in ammonium acetate-washed (Section 4.9) coal's ash was determined using a Perkin-Elmer Optima 5300DV inductively coupled plasma - atomic emission spectroscope (ICP-AES) and the analysis was conducted by personnel at the Materials Research Institute (MRI), Penn State. This analysis

40 was done primarily to confirm the removal of ion-exchangeable cations. The analysis of the other unwashed feed coals was provided by the Penn State Coal Sample Bank. Before analysis, the sample was ashed in a muffle furnace at 1,023 K following which, a dissolution step (Lithium metaborate fusion technique) was carried out. About 0.1 g of sample was weighed out and mixed and shook in vials containing 1 g of lithium metaborate. The contents of the vials were moved to graphite crucibles which were placed in an oven, pre-heated to 1,173 K. 100 ml of 5% nitric acid was pipetted out into Teflon beakers and the contents of the graphite crucibles were transferred to the beakers. The solution in the beakers was stirred for atleast 15 minutes using a magnetic stirrer and stored in polyethylene bottles. Rock standards were used to calibrate the instrument. Then the sample solution (from the dissolution step) was nebulized into a fine aerosol and introduced into an argon plasma with a temperature ranging from 7,000-10,000 K. The high energies encountered excite the electrons in the atoms and ions in the aerosol to higher energy levels. When the electrons return to their ground state, they emit characteristic wavelengths of energy. Each element has a unique wavelength and these are measured on a charge-coupled device (CCD) chip. By measuring the intensities of these wavelengths and comparing them to those generated by standards, the concentrations of different atoms were determined using a calibration curve.

4.9 Removal of ion-exchangeable cations Low-rank coals tend to contain ion-exchangeable cations which can catalyze the oxidation and gasification reactions. When the effect of rank on combustion in an oxy-fuel environment was analyzed, there was a possibility of lower rank coals being more reactive due to the catalytic activity of these ion-exchangeable cations. To identify and eliminate the possibility of catalytic effect, samples were washed with ammonium acetate to remove organically-bound (mainly carboxyl groups) ion-exchangeable cations such as K, Ca and Na. The method was adapted from the work of Miller et al.[86]. Twenty five grams of sample was weighed out in a

41 container. A 1 M solution of ammonium acetate was prepared in a beaker and placed on a hot plate with a magnetic stirrer. The hot plate was maintained at ~330 K and stirred at 270 rpm. The coal was mixed into the solution and the beaker was covered with parafilm to prevent evaporation. After 24 hours, the sample solution was filtered using a Buchner funnel. The solid residue was placed on a Pyrex plate and dried in an air dying oven at 333 K. The dried coal sample was then used for combustion and devolatilization tests.

42

Chapter 5 Results and Discussion


Table 5-1 shows the proximate analysis, ultimate analysis, higher heating value (HHV) and petrographic composition of the DTR feed and ammonium acetate-washed (AAW) coals. AAW coals were considered as part of the study to account for the catalytic activity of alkali and alkaline-earth metal salts and oxides. Table 5-2 lists the major and minor element oxide concentrations for parent and AAW coals. The decrease in concentration of the ion-exchangeable elements for the AAW coals can be seen in Table 5-2. No inorganic mineral matter was removed during the ammonium acetate wash process and only organically-bound ion-exchangeable cations were removed. The switchgrass (Panicum virgatum) sample (-60 mesh) was only used for thermogravimetric burnout studies in Section 5.6.

5.1 Combustion Test Results - Parent coal The combustion test conditions are given in Section 4.2.1. Figures 5-1 to 5-4 show the plots of carbon conversion (%) vs. residence time (s) for the fours coals used in the study at three different residence times and different gas atmospheres. The analysis of the resultant chars is given in Appendix D (Tables D-1 to D-5). From Figure 5-1, it can be seen that for the Pocahontas coal, which is a high-rank low volatile bituminous coal, air combustion shows a higher carbon conversion (49% at a residence time of 0.8 s) than oxy-fuel combustion (21%O2/79%CO2) (30% at a residence time of 0.8 s). Pittsburgh coal, a high volatile bituminous coal, shows similar results (Figure 5-2) as those for Pocahontas coal although the difference in carbon conversion

43 between air and oxy-combustion is less (~1%). For subbituminous rank Dietz coal and Beulah lignite, as seen in Figures 5-3 and 5-4 respectively, combustion in 21%O2/79%CO2 produced a higher carbon conversion (98% for Dietz and 94% for Beulah at a residence time of 0.8 s) compared to that in air (85% for Dietz and 90% for Beulah at a residence time of 0.8 s). Another important result is that at higher oxygen concentrations in the mixture of O2/CO2 (enhanced oxycombustion), the carbon conversion increased and this can be seen in Figures 5-1, 5-2 and 5-4. The increase in carbon conversion in an enhanced oxygen atmosphere is primarily attributed to an increase in particle temperature due to the exothermic char-O2 reaction and this is discussed in further detail in Section 5.4. From the results, it appears that as the coal rank decreases, the contribution from the char-CO2 reaction increases, enhancing the net carbon conversion. It should be noted from Figures 5-1 to 5-4 that the difference in conversion between air and oxy-fuel combustion decreases as the residence time increases. At the highest residence time (1.7 s), there is very minimal difference (1%) especially for the Pittsburgh coal, Dietz coal and Beulah lignite. From this and Table 4-1, it can be said that oxy-fuel combustion will provide maximum benefit in terms of carbon conversion for smaller boilers (60 - 110 MW) firing low-rank coals apart from the benefit of providing a pure stream of CO2 at the exit of the boiler, ready for sequestration. This study has been able to show the effect of residence time on carbon conversion in both atmospheres and it is thought that this might be one reason why certain earlier studies did not notice any significant benefit of oxy-fuel combustion. The reproducibility of results for carbon conversion was within 3% and is shown in Table 5-3. Subsequent sections of this chapter aim to further explain the role of the char-CO2 reaction in increasing the carbon conversion of low-rank coals under oxy-fuel conditions.

44 Table 5-1. Analysis of coal and switchgrass samples Pocahont as Proximate Analysis (wt%, as determined) Moisture Volatile matter Fixed carbon Ash 0.21 19.98 73.47 4.02 0.64 35.15 56.73 7.48 4.65 38.35 50.03 6.97 7.33 39.16 45.70 7.81 6.39 39.95 49.74 3.92 6.33 42.39 45.05 6.23 2.7 82.2 11.0 4.1 Pittsburgh Dietz Beulah DietzAAW BeulahAAW Switchgrass

Ultimate Analysis (wt%, as determined) Carbon Hydrogen* Nitrogen Sulfur Oxygen (by difference)* Heating value (Btu/lb)** Total Vitrinite (%) Total Huminite (%) Total Inertinite (%) Total Liptinite (%) ASTM Rank 89.2 3.75 0.88 0.78 1.16 15,690 77.1 4.78 0.85 1.46 7.69 14,280 70.7 4.91 0.58 0.39 11.8 10,390 59.1 4.34 0.42 0.54 20.46 7,940 81.0 5.2 2.6 0.75 10.45 -66.4 4.51 3.07 0.75 25.27 -46.6 5.69 0.1 0.2 43.31 7670

86.9 --

79.1 --

-85.5

-73.8

---

---

---

9.9

12.6

11.7

19.5

--

--

--

0.5 lvb

3.9 hvAb

2.8 subB

6.7 ligA

---

---

---

*excluding H and O in moisture **mineral matter free, as received

45

Table 5-2. Major and minor elemental oxide analysis of the ash reported as oxides (%) for the coals used in the study

Oxide (%) SiO2 Al2O3 TiO2 Fe2O3 MgO CaO Na2O K 2O P2O5 SO3 Total Ash (%)

Pocahontas

Pittsburgh

Dietz

DietzAAW

Beulah

BeulahAAW

28.6 19.6 0.95 17.0 2.49 15.2 1.14 0.83 0.04 13.0 98.85 6.34

48.5 23.2 1.00 14.84 0.76 2.49 0.69 1.87 0.53 1.95 95.83 7.48

30.45 15.12 1.00 5.91 6.63 19.57 2.08 0.50 0.35 16.81 98.42 6.97

33.2 19.3 1.43 6.69 3.40 13.1 1.17 0.38 0.34 17.0 96.01 4.19

11.5 11.1 0.57 9.29 9.94 27.4 9.39 0.59 0.49 18.2 98.47 7.81

24.8 13.2 0.71 10.1 4.65 21.2 2.16 0.31 0.31 16.0 93.44 6.65

46
100 90

Carobn Conversion (%)

80 70 60 50 40 30 20 0.8 1.2 1.7 air 21%O2/CO2 30%O2/CO2

residence time (s)


Figure 5-1. Carbon conversion (%) vs. residence time (s) for Pocahontas coal

100

Carbon Conversion (%)

90 80 air 70 60 50 40 30 20 0.8 1.2 1.7 21%O2/CO2 30%O2/CO2 40%O2/CO2

residence time (s) Figure 5-2. Carbon conversion (%) vs. residence time (s) for Pittsburgh coal

47
100 90

Carbon Conversion ( %)

80 70 60 50 40 30 20 0.8 1.2 1.7 air 21%O2/CO2

residence time (s)


Figure 5-3. Carbon conversion (%) vs. residence time (s) for Dietz coal

100 90

Carbon Conversion (%)

80 70 60 50 40 30 20 0.8 1.2 1.7 air 21%O2/CO2 30%O2/CO2

residence time (s)


Figure 5-4. Carbon conversion (%) vs. residence time (s) for Beulah lignite

48 Table 5-3. Reproducibility of Combustion tests Coal Date Residence time (s) Atmosphere Carbon Conversion (%) 94.1 92.7 96.7 50.5 47.1 51.4 91.7 89.5 93.2 2 3 Error

Beulah

04/18/2010 04/19/2010 04/20/2010

0.8 0.8 0.8 1.2 1.2 1.2 1.2 1.2 1.2

Oxy Oxy Oxy Oxy Oxy Oxy Air Air Air

Pocahontas

03/26/2010 03/27/2010 03/28/2010

Pittsburgh

04/29/2010 04/30/2010 05/01/2010

5.2 Devolatilization and Isothermal Reactivity Studies Char samples were generated in the DTR at carbon conversions approximately equal to the ASTM volatile matter content of four coals and Beulah-AAW coal. The reason for this was to try to explain the combustion test results (Section 5.1) by means of using low temperature, benchscale TGA reactivity studies and then theoretically extrapolating these results to high temperature (Section 5.3). The test conditions have been discussed previously in Section 4.2.2. Another reason for this is the wide scatter in literature for reactivity parameters even for the same rank of coal. It was decided to obtain intrinsic rate parameters for the four coals used in the study since, as will be discussed in Section 5.3, char conversion is very sensitive to the intrinsic rate parameters and a high degree of accuracy is necessary in obtaining these parameters. The analysis of the resultant char samples is shown in Table 5-4. Isothermal and non-isothermal TGA along with surface area

49 analysis were performed on the char samples. Table 5-5 shows the surface area analysis of the char samples and it can be seen that the BET surface area increases with increase in coal rank. Another observation is that the average pore diameter

Table 5-4. Analysis of char samples from DTR

Pocahontas

Pittsburgh

Dietz

Beulah

BeulahAAW

Proximate Analysis (wt%) Moisture Volatiles Fixed carbon Ash Ultimate Analysis (wt%) Carbon Hydrogen Nitrogen Sulfur Oxygen (by difference) H/C ratio (*10-2) Carbon conversion (%)

0.09 17.7 74.9 7.31

0.16 25.12 62.6 12.12

2.0 30.3 57.1 10.6

3.5 34.8 50.2 11.5

2.4 34.7 53.2 9.7

85.9 3.2 0.9 0.9 1.79 3.7 14

75.1 3.1 1.4 1.9 6.38 4.1 41.1

73.2 3.5 0.6 0.4 11.7 4.8 34.9

66.7 3.2 0.4 1.1 17.1 4.8 32.1

63.3 3.81 1.31 0.85 30.73 6.0 35.4

50 Table 5-5. Surface area analysis of devolatilization char samples Pocahontas BET surface area(m2/g) Micropore surface area (m2/g) Average pore diameter (Ao) 17.6 196.4 Pittsburgh 14.7 113.4 Dietz 6.2 205.6 Beulah 1.8 192.6 Beulah-AAW 2.4 201.7

83.5

63.9

132.3

284.5

375.3

follows an opposite trend. Dietz and Beulah show higher values average pore diameters (132.3 Ao and 284.5 Ao, respectively) compared to the higher rank coals, Pocahontas and Pittsburgh (83.5 Ao and 63.9 Ao respectively). It has been seen that in general, the average pore size decreases with increase in coal rank. Macropores dominate lower rank coals and micropores dominate in higher rank (vitrinite-rich) coals. Also, macropores and mesopores pores act as feeder pores allowing the access of reactant to the interior surface, increasing the reactivity of lower rank coals. This study attempts to provide an empirical explanation of the char reactivity in terms of the surface area measurements but the shortcoming of this is that, char reactivity depends on several other factors such as rank, maceral composition and the conditions of char generation[87] and in general, simple char properties such as carbon content and pore surface area are not suitable indices[53]. From the results in Table 5-5, it can be seen that Pocahontas char has the largest percentage of micropores as it has the largest BET surface area and smallest pore diameter amongst the four coals. Similarly the trend follows with increasing percentage of macropores and decreasing percentage of micropores as the rank of coal decreases. A possible reason for the high BET surface area of high-rank chars, Pocahontas and Pittsburgh could be due to micro- and mesopore enlargement and creation of new mesopores from micropores during the

51 devolatilization process, leading to better access for N2 to these pores[88]. Since the chars in this study were generated at intermediate carbon conversions, it is thought that the BET surface areas were around the maximum attainable as beyond the initial peak, surface area begins to drop with carbon conversion. The reasoning was based on the surface area vs. conversion profiles generated for four Japanese coals by Adschiri and Furusawa[89] showing the surface areas to increase until a maximum (at conversions <40%) and then decrease there on. Figures 5-5 to 5-12 show the oxidation and gasification reactivity profiles obtained from isothermal TGA as shown in Section 4.4. Tables 5-6 and 5-7 show the reactivity values obtained after the occurrence of the maximum in the curves. It can be noticed from the reactivity plots that the curves get flatter as the rank decreases and this is related to the surface area of the chars. For the Pocahontas coal, which has the least percentage of macropores, the reactant gases consume these sites first and the steep decrease in reactivity is due to the increasing difficulty for the gases to access the large number of micropores. For the Beulah lignite, which has the largest percentage of macropores, the reactivity profile remains flat and shows a steep dip towards the end once all the macropores are consumed. Another reason for the flat profile of lower rank coals is due to their well connected pore structure leading to better access of reactant gas to the active sites located in the micropores leading to higher reactivity as opposed to higher rank coals that have negligible feeder pores[49]. The steep decrease in the later stages of conversion could also occur due to decrease in surface area caused by pore coalescence[87]. It can be seen from Figures 5-8 and 5-12 that the reactivity of the AAW Beulah char shows a considerable decrease especially for the char-CO2 reaction. This is attributed to the catalytic effect of highly dispersed (organically-bound) Ca2+ cations. The CaO gasification catalysis in lignites has already been studied by other authors[49, 52, 88] and similar results were obtained. Among the four coals used in the study, the Beulah lignite has the highest calcium content and is thus expected to have the highest gasification catalytic effect. Another aspect to be noted in Figure 5-8 is that the curves for

52 the AAW Beulah char are not flat as is the case with parent coal and this is because of absence of active sites, which are not being created due to the absence of CaO. CaO is known to increase the concentration of active sites[54] and this is the reason for the prolonged high reactivity (flat profile) of whole Beulah lignite. It can be noticed from Table 5-5, that AAW Beulah char has a higher BET surface area and larger percentage of macropores as compared to Beulah char, showing that surface area is not the only determining factor for char reactivity and that catalytic activity of alkali and alkaline-earth metal salts and oxides also plays an important role. Miura et al.[53] have stated that the gasification of low-rank coals proceeds mainly via the catalytic route while for higher rank coals, it proceeded via a non-catalytic reaction sequence related to the number of active sites. Wigmans et al.[90] have stated that the maximum gasification catalytic activity correlates well with the oxygen content of the carbon and as shown in Table 5-1, lowrank coals, Dietz and Beulah, have a higher oxygen concentration compared to the higher rank coals.

0.25 0.2 0.15 673 K 0.1 0.05 0 0 20 40 60 80 698 K 723 K

Reactivity(mg/g.s)

Conversion(%) Figure 5-5. Char-O2 reactivity profiles for Pocahontas coal

53
0.5 0.45 0.4

Reactivity(mg/g.s)

0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0 20 40 60 80 673 K 698 K 723 K

Conversion(%)

Figure 5-6. Char-O2 reactivity profiles for Pittsburgh coal

1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0 20 40 60 80 100 648 K 673 K 698 K

Reactivity(mg/g.s)

Conversion(%) Figure 5-7. Char-O2 reactivity profiles for Dietz coal

54
2 1.8 1.6 1.4 623 K 648 K 673 K 698 K 623 K_AAW 648 K_AAW 673 K_AAW

Reactivity(mg/g.s)

1.2 1 0.8 0.6 0.4 0.2 0 0 20 40 60

Conversion(%)

80

Figure 5-8. Char-O2 reactivity profiles for Beulah lignite

0.16 0.14

Reactivity(mg/g.s)

0.12 0.1 0.08 0.06 0.04 0.02 0 0 20 40 60 80 1123 K 1148 K 1173 K

Conversion(%)
Figure 5-9. Char-CO2 reactivity profiles for Pocahontas coal

55
0.12 0.1

Reactivity(mg/g.s)

0.08 0.06 0.04 0.02 0 0 20 1123 K 1148 K 1173 K

Conversion(%)

40

60

80

Figure 5-10. Char-CO2 reactivity profiles for Pittsburgh coal

0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 20 40 60 80 100 1123 K 1148 K 1173 K

Reactivity(mg/g.s)

Conversion(%) Figure 5-11. Char-CO2 reactivity profiles for Dietz coal

56
2.5

1123 K 1148 K 1173 K

Reactivity(mg/g.s)

1.5

1123 K_AAW 1148 K_AAW 1173 K_AAW

0.5

0 0 20

Conversion(%)

40

60

80

100

Figure 5-12. Char-CO2 reactivity profiles for Beulah lignite It can be noticed from Tables 5-6 and 5-7 that there is a clear effect of rank with both char-O2 and char-CO2 reactivities increasing with decrease in coal rank. For example, for the Pocahontas coal char-O2 reactivity increases from 0.08 to 0.23 mg/g.min as the temperature increases from 673 K to 723 K while for the Beulah lignite it increases from 1.02 to 6.03 mg/g.min within the same interval. Table 5-8 shows the oxidation and gasification rate parameters extracted from reactivity plots, which are used in Section 5.3. The activation energies for the char-CO2 reaction, as seen in Table 5-8, decrease with a decrease in rank of coal and this is discussed further in Section 5.3. Also, the activation energy for the char-O2 and char-CO2 reaction (especially char-CO2) is different for the AAW coal and parent Beulah lignite which is contrary to what was reported by Kapteijn et al.[54] that the catalyst did not change the activation energy of the reaction and only increased the number of active sites. This shows that any increase or decrease in active site concentration by an inherent property or catalytic effect leads to a change in the activation energy of the reaction.

57 Table 5-6. Char-O2 reactivity for the four coals T (K) Pocahontas 623 648 673 698 723 --0.08 0.14 0.23 Reactivity (mg/g.min) Pittsburgh --0.13 0.23 0.44 Dietz -0.43 0.65 1.50 2.78 Beulah 0.27 0.60 1.02 1.74 6.03 Beulah-AAW 0.22 0.41 0.90 ---

Table 5-7. Char-CO2 reactivity for the four coals T (K) Pocahontas 1123 1148 1173 0.04 0.07 0.12 Reactivity (mg/g.min) Pittsburgh 0.02 0.05 0.10 Dietz 0.50 0.67 0.79 Beulah 1.25 1.72 2.14 Beulah-AAW 0.65 1.04 1.53

Table 5-8. Oxidation and gasification rate parameters for the four coals Coal Oxidation A (kg/m2.s.Pa) Pocahontas Pittsburgh Dietz Beulah Beulah-AAW 0.03 0.3 1.9 4.8 12.7 E (kJ/mol) 87.4 97.1 92.5 89.3 97.3 Gasification A (kg/m2.s.Pa) 1,061.4 289,119 0.005 0.2 150 E (kJ/mol) 251.1 306.1 103.1 117.3 187.9

58 5.3 Theoretical char-O2 and char-CO2 reaction rate study Char-O2 and char-CO2 reaction rates were extrapolated to high temperatures using intrinsic rate parameters (Section 5.2, Table 5-8) and the details are given in Appendix A with all calculations being performed for a gas atmosphere of 21% O2/79% CO2. The model used in this study was adapted from the work of Naredi[21]. The motivation behind this study was to understand oxidation and gasification char reactivity of the four coals over a broad range of temperatures and since obtaining intrinsic rate parameters at high temperatures experimentally was not possible due to diffusion limitations. The study would also help understand the combustion results of Sections 5.1. The overall reaction rate was calculated as a sum of the contributions from surface reaction rate and gas diffusion rate. The effectiveness factor was calculated from the Thiele modulus method (assuming spherical particle shape) with the effective diffusivity calculated as a sum of the Knudsen and molecular diffusivities. Figure 5-13 shows the extrapolated oxidation rates of the four coals used in the study. The profiles show a rank effect with the lower rank coals showing higher rates compared to the higher rank coals. Figure 5-14 shows the plot of effectiveness factors as a function of temperature for the char-O2 reaction. As can be seen, there is a significant decrease in effectiveness factors with an increase in temperature. The oxidation effectiveness factors show a minor rank effect with the Pocahontas coal having the highest values and the Beulah lignite having the lowest, though the difference is not very significant. Figure 5-15 shows the plot of extrapolated gasification rates and shows a rank effect similar to Figure 5-13 though the higher rank coals show a significant increase in reaction rate with increase in temperature especially for the Pocahontas coal that shows an exponential increase in rate beyond 1,400 K. Figure 5-16 shows the gasification effectiveness factors and it can be seen that the values are much higher for the lower ranks coals compared to the higher rank coals at equivalent temperatures. Also, on comparing effectiveness factors for the char-O2 and char-CO2 reaction, it

59 was noticed that the values were higher for the char-CO2 reaction compared to the char-O2 reaction at corresponding temperatures. This indicates that the char-O2 reaction approaches diffusion control much faster than the char-CO2 reaction and that the CO2 molecule has a higher degree of access to the char internal surface area compared to the O2 molecule, especially in lower rank coals like Dietz and Beulah. This also means that at high temperatures, a larger portion of the char's internal surface area is consumed by the CO2 molecule as compared to the O2 molecule. In order to explain the decrease in effectiveness factors with increase in temperature and the effect of rank, a number of model parameters were compared. Figures 5-17 and 5-18 show the variation of effectiveness factor and intrinsic rate, Rs with respect to temperature. As can be seen that the intrinsic reaction rate is what determines the effectiveness factor, with the effectiveness factor decreasing proportionately with an increase in intrinsic reaction rate. This is because, the gas penetration length decreases as the intrinsic reaction rate increases, and thus the effectiveness factor (the fraction of total internal surface area used in the reaction) decreases[39]. In simple words, as the intrinsic reaction rate approaches very high values, the reactant gas molecule reacts only on the surface of the char and do not penetrate inside the char particle. The variation of gas diffusion rate, Rd and reaction rate per unit area of solid, Rac were also compared and are shown in Figures 5-19 and 5-20 for the group of coals with individual profiles shown in Appendix E (Figures E-1 to E-8). Rd does not have any rank dependence as it varies only with particle temperature and diameter. Rac is a function of , Rs, Ag and dp.

60
1.80E-01 Reaction rate (kg/m2.s.Pa) 1.60E-01 1.40E-01 1.20E-01 1.00E-01 8.01E-02 6.01E-02 4.01E-02 2.01E-02 1.00E-04 700 900 1100 1300 1500 1700 1900 Dietz Pocahontas Pittsburgh Beulah

Temperature (K)

Figure 5-13. Extrapolated oxidation rates for the four coals used in the study

1.2

Effectiveness factor

1 0.8 0.6 0.4 0.2 0 700 900 1100 1300 1500 1700 1900 Deitz Pocahontas Pittsburgh Beulah

Temperature (K) Figure 5-14. Char-O2 reaction effectiveness factors for the four coals used in the study

61
1.00E+00 700 log(Reaction rate(kg/m2.s.Pa)) 1.00E-01 1.00E-02 Dietz 1.00E-03 1.00E-04 1.00E-05 1.00E-06 Pocahontas Pittsburgh Beulah 900 1100 1300 1500 1700 1900

Temperature (K)

Figure 5-15. Extrapolated gasification rates for the four coals used in the study

1.2

Effectiveness factor

1 0.8 0.6 0.4 0.2 0 700 900 1100 1300 1500 1700 1900 Deitz Pocahontas Pittsburgh Beulah

Temperature (K)

Figure 5-16. Char-CO2 reaction effectiveness factors for the four coals used in the study

62
Intrinsic reaction rate (kg/m2.s.Pa) Intrinsic reaction rate (kg/m2.s.Pa) 1.2 1 Effectiveness factor 0.8 0.6 0.4 0.2 0 500 700 900 5.00E-04 4.50E-04 4.00E-04 3.50E-04 3.00E-04 2.50E-04 2.00E-04 1.50E-04 1.00E-04 5.00E-05 0.00E+00 1100 1300 1500 1700 1900

eta Rs

Temperature (K)

Figure 5-17. Variation of effectiveness factor and intrinsic reaction rate for char-O2 reaction of Pittsburgh coal

1.2 1 Effectiveness factor 0.8 0.6 0.4 0.2 0 700 900 1100 1300 1500 1700

4.00E-04 3.50E-04 3.00E-04 2.50E-04 2.00E-04 1.50E-04 1.00E-04 5.00E-05 0.00E+00 1900

eta Rs

Temperature (K)

Figure 5-18. Variation of effectiveness factor and intrinsic reaction rate for char-CO2 reaction of Pittsburgh coal

63

It can be seen that for the oxidation reaction there is a rank effect. The lower rank coals (i.e., Dietz and Beulah) have a higher Rac compared to the higher rank coals (i.e., Pocahontas and Pittsburgh), throughout the entire char conversion process. For the gasification reaction, there is a rank effect at lower temperatures, with the lower rank coals having a higher R ac compared to the higher rank coals. At higher temperatures, there is an exponential increase in Rac for the higher rank coals, especially the Pocahontas coal. This is related to the activation energy of the char-CO2 reaction for these coals. As seen in Table 5-8, the high-rank coals (i.e., Pocahontas and Pittsburgh) have significantly higher gasification activation energies compared to low-rank coals which results in higher reaction rates at high temperatures. As seen from Figures 5-14 and 5-16, the effectiveness factors drop significantly for high-rank coals at high temperatures suggesting that the overall char conversion depends both on the occurrence of reaction (kinetics) and gas diffusion. Alternately, this can also be explained by the shrinking core model where, reactivity increases with an increase in temperature. An increase in reactivity increases the gas diffusion and ash layer resistances. So, even though the reaction rate is high, the reactant gas is unable to access the internal surface area of the char particle (shown by the drop in effectiveness factor), therefore, char conversion depends on all these factors and not just the reaction rate. Comparing the above results with the combustion test results in Figures 5-1 to 5-4, the higher carbon conversion of lower rank coals (i.e., Dietz and Beulah) compared to high-rank coals (i.e., Pocahontas and Pittsburgh) under oxy-fuel atmospheres can be primarily attributed to both the char-CO2 reaction rate and higher effectiveness factors at high temperatures. Also, it can be inferred that most of the char conversion process occurs under zone I and zone II conditions for low-rank coals where both the reaction rate and effectiveness factor are high enough for char conversion to take place. In zone III, though the reaction rates are very high, the effectiveness factors are very low, leading to lower conversion.

64
0.001 500 0.0001 0.00001 log(Rd) 0.000001 0.0000001 1E-08 1E-09 1E-10 Temperature (K) log(Rac-Dietz) log(Rac-Pittsburgh) log(Rac-Pocahontas) log(Rac-Beulah) 1000 1500 2000

Figure 5-19. Extrapolated oxidation Rac and Rd for the four coals used in the study
1 0.01 0.0001 0.000001 1E-08 1E-10 1E-12 1E-14 1E-16 Temperature (K) Rd log(Rac-Dietz) log(Rac-Pittsburgh) log(Rac-Pocahontas) log(Rac-Beulah) 500 1000 1500 2000

Figure 5-20. Extrapolated gasification Rac and Rd for the four coals used in the study

It can be concluded that conversion is a function of both overall reactivity and effectiveness factor and this may be explained for each coal. For Pocahontas coal, oxy-combustion produces lower conversion compared to conventional air combustion though the gasification reactivity is high at high temperatures, because the char-CO2 effectiveness factors are low at high

65 temperatures. For Pittsburgh coal, the lower conversion in an oxy-fuel environment is a combination of both low gasification reactivity and very low effectiveness factors. For low-rank coals, Dietz and Beulah, the higher conversion in an oxy-fuel environment is due to both high gasification reactivity and high char-CO2 effectiveness factors. The effect of particle size was also studied by using two particle sizes, 81 m and 237.5 m and is shown in Figures 5-21 and 5-22. The profiles that are shown here are for Dietz coal and very similar profiles were obtained for the other coals. It was assumed that both particle sizes have the same rate parameters. As can be seen, the effectiveness factors for the larger particles are smaller and decrease more rapidly with an increase in temperature as compared to the smaller particles. Although the two particles have equivalent total specific surface areas (m2/kg), the increase in size makes it difficult for the reactant gas to access the internal surface area, causing a decrease in effectiveness factor. From this, it can be said that smaller particles could undergo higher char conversion as they approach zone III at much higher temperatures (much later) compared to larger particles provided both particles have similar reactivities.
1.2 1 Effectiveness factor 0.8 0.6 d=81 micron 0.4 0.2 0 500 1000 1500 2000 d = 237.5 micron

Temperature (K)

Figure 5-21. Char-O2 effectiveness factors for two particle sizes of Dietz coal

66
1.2 1 Effectiveness factor 0.8 0.6 d=81 micron 0.4 0.2 0 500 1000 1500 2000 d=237.5 micron

Temperature (K)

Figure 5-22. Char-CO2 effectiveness factors for two particle sizes of Dietz coal

As Rac and the Thiele modulus,

are directly proportional to both particle diameter, dp and

intrinsic rate, Rs, a sensitivity study indicated the model to be highly sensitive to a change in Rs as compared to dp therefore, in order to model char conversion accurately, it is very important to know the intrinsic rate parameters to a high degree of accuracy.

67 5.4 Theoretical char particle temperature model Section 5.3 showed the increasing importance of the char-CO2 reaction on carbon conversion at high temperatures due to high effectiveness factors compared to the char-O2 reaction. A theoretical model was formulated to quantify the effect of gas temperature and CO2 partial pressure on char particle temperature in an oxy-fuel atmosphere (O2/CO2). The objective of this study was to further demonstrate the importance of the char-CO2 reaction especially at high temperatures and high CO2 partial pressures and to substantiate the results of Section 5.3. The governing equation is the particle heat balance equation that was solved using ordinary differential equation solver, 'ode45' in MATLAB. The equation is given by[91], 5.4.1

where, Cp is the specific heat of the char particle was calculated from the correlation developed by Eisermann et al.[92], 5.4.2 The convective heat transfer coefficient, h was evaluated assuming spherical particle shape as[93],

5.4.3 where, Rep is the Reynolds number based on the particle diameter and velocity, and Pr is the Prandtl number. A particle diameter, dp of 45 and a density of 1,000 kg/m3, travelling at a velocity of 0.6 m/s was assumed based on the work of Wang et al.[41]. In equation 5.4.1, Ap (m2) is the surface area of the char particle; mp (kg) is the mass of the particle; Tp (K) is the particle temperature; Tg (K) is the gas temperature; Tw (K) is the temperature of the environment; is Stefan-Boltzmann constant (5.67*10-8 W/m2K4) and , the

particle emissivity was taken as 0.5[26].

68 The two char reactions considered along with combustion enthalpy of carbon are given by, 5.4.4 5.4.5 The char burning rate, qcmb (kg/m2.s) is given by the relation[91], 5.4.6 where, fmac is the maceral correction factor, given by, 5.4.7 Vit and In are the percentages of vitrinite and inertinite respectively in the parent coal. Typically northern hemisphere coals are rich in vitrinite, while southern hemisphere coals are rich in inertinite[91]. To account for the effect of maceral types on char reactivity, a maceral correction factor was used. In this study, a value of 82.8 and 13.2 for Vit and In respectively were chosen, typical of Pittsburgh#8 coal. In equation 5.4.6, Ri is the intrinsic reaction rate; Pg (atm) is the partial pressure of the reacting gas and is the effectiveness factor calculated from the Thiele modulus approach given in Appendix A (equations A.5-A.9). Reaction order, n for the char-O2 reaction was assumed as 0.5 and 0.8 for the char-CO2 reaction. The intrinsic rate parameters for a high volatile C coal used in this model are given in Table 5-9. A sample calculation is shown in Appendix F.

Table 5-9. Rate Parameters used in the theoretical particle temperature model[21] A (1/s.atm) Char-O2 Char-CO2 1.33E+05 3.5E+05 E (kJ/mol) 123 189

69 Figures 5-23 to 5-27 show plots of computed char particle temperature (K) vs. oxygen concentration (%) at gas temperatures of 1,373 K, 1,573 K, 1,673 K, 1,773 K and 1,873 K respectively. The results showed that as the gas temperature increases, the difference between the char particle and gas temperature (Tp-Tg) decreases, indicating an increasing influence of the char-CO2 gasification reaction in lowering the particle temperature. At 1,373 K, the difference between char particle and gas temperature was 570 K while at 1,573 K, it was 321 K. This effect was pronounced at very low oxygen concentrations and can clearly be seen at gas temperatures of 1,773 K and 1,873 K. At an oxygen concentration of 3% and gas temperature of 1,873 K, the difference between char particle and gas temperature was just 83 K. Another aspect to notice from Figures 5-23 to 5-27 is that the slope of the plot increases with increase in gas temperature which may be attributed to the influence of the char-CO2 reaction occurring at high temperatures. From the results, it may be said that at high gas temperatures when the oxygen concentration drops to very low levels, it is possible that the char-CO2 gasification reaction could lower the char particle temperature significantly, possibly even below that of the surrounding gas temperature under certain conditions. This model has demonstrated the importance of consideration of the char-CO2 reaction especially at high temperature and high CO2 partial pressure. In utility scale boilers, the char-CO2 reaction would play a major role especially in regions where the O2 concentration is low (after the primary combustion zone) and this would increase the overall efficiency of the process. Figure 5-28 shows a computationally simulated gas temperature profile for a typical oxy-fuel PC boiler. It can be seen from that the primary combustion zone and the region above it have very high gas temperatures (>1500 K) and a significant portion of the char is expected to be consumed by the char-CO2 reaction in these regions. If fuel feed is staged into the boiler and coal is injected along with CO2 as carrier gas in the region marked in red, the effect of char-CO2 reaction will increase further.

70
2100 2000

1373 K

Temperature (K)

1900 1800 1700 1600 1500 1400 1300 10 15 20 25 30 35 Tp Tg

% O2 Figure 5-23. Computed char particle temperature (K) vs. oxygen concentration(%) at a gas temperature of 1,373 K under oxy-fuel conditions

2100 2000

1573 K

Temperature (K)

1900 1800 1700 1600 1500 1400 1300 10 12 14 16 18 20 22 Tp Tg

% O2 Figure 5-24. Computed char particle temperature (K) vs. oxygen concentration(%) at a gas temperature of 1,573 K under oxy-fuel conditions

71
2100 2000

1673 K

Temperature (K)

1900 1800 1700 1600 1500 1400 1300 10 12 14 16 18 20 22 Tp Tg

% O2 Figure 5-25. Computed char particle temperature (K) vs. oxygen concentration(%) at a gas temperature of 1,673 K under oxy-fuel conditions

2200 2100 2000 1900 1800 1700 1600 1500 1400 1300 10 12 14

1773 K

Temperature (K)

Tp Tg

16

18

20

22

% O2 Figure 5-26. Computed char particle temperature (K) vs. oxygen concentration(%) at a gas temperature of 1,773 K under oxy-fuel conditions

72
2300 2200 2100 2000 1900 1800 1700 1600 1500 1400 1300 0 5 10 15 20 Tp Tg

Temperature (K)

% O2 Figure 5-27. Computed char particle temperature (K) vs. oxygen concentration(%) at a gas temperature of 1,873 K under oxy-fuel conditions

Primary combustion zone

Figure 5-28. Simulated temperature (oC) profiles for an oxy-fuel PC boiler[94]

73 5.5 Combustion Test Results - AAW coals From the reactivity profiles in Section 5.2, the catalytic effect of Ca was evident and it is possible that carbon conversion could be catalyzed by ion-exchangeable cations, especially in low-rank coals. These profiles are generated at low temperatures (673 K - 1173 K). However, at higher temperatures that are typical in utility scale boilers, the catalytic activity may not be significant due to sintering phenomenon. In order to confirm this phenomenon as to whether the higher carbon conversion during oxy-fuel combustion (parent coals) was due to catalytic effect or due to coal rank, combustion tests in an oxy-fuel atmosphere were conducted at a temperature of 1,873 K for the AAW Beulah lignite and Dietz coal in the DTR and the char conversions are compared to those obtained for whole coal. The reason for selecting the two low rank coals was because in general, the gasification of low-rank coal chars (especially lignites[43]) proceeds via the catalytic route while for high-rank chars, it proceeds via a non-catalytic reaction sequence[53]. Figures 5-29 and 5-30 show the comparison carbon conversion versus residence time plots for Dietz coal and Beulah lignite, respectively. It can be seen that contrary to what was expected, the AAW coals showed similar carbon conversion to the parent coals. This may be
100 90

Carbon Conversion ( %)

80 70 60 50 AAW-21%O2/CO2 40 30 20 0.8 1.7 21%O2/CO2

residence time (s) Figure 5-29. Carbon conversion (%) vs. residence time (s) for Dietz and AAW Dietz coal

74

attributed to CaO sintering at the high particle temperatures in the DTR leading to loss of catalytic activity. Kapteijn et al.[54] found the initial reactivity of Ca around 1,000 K to be equal to that of K, a well known gasification catalyst although with increasing carbon burn-off, it shows severe deactivation due to CaO sintering. Gopalakrishnan et al.[88] conducted tests in the droptube reactor at 1900 K on Zap lignite and Dietz coal and showed very similar results of loss of reactivity primarily due to CaO sintering. The isothermal reactivity profiles in Section 5.2 were conducted at low temperatures between 623 K - 1,173 K and that is the reason why the catalytic effect of CaO was noticed while in the DTR a furnace temperature of 1,873 K was used with the particle experiencing a very high heating rate (103 - 104 K/s) and hence no enhanced carbon conversion due to catalysis was noticed. Cope et al.[95] conducted char reactivity studies on untreated, acid-washed and Ca doped Zap lignites by generating chars using a flat-flame burner. They reported very similar findings that Ca catalysis was very significant at low temperatures but the decreasing rate with conversion was due to CaO sintering with time of exposure to high temperature environment. At this point, it has been established that the higher carbon conversions of low-rank coals under oxy-fuel environments as compared to air is purely an effect of rank and that there is no influence of catalytic activity on carbon conversion at high temperatures (>1173 K).

75
100 90

Carbon Conversion (%)

80 70 60 50 40 30 20 0.8 1.7 AAW21%O2/CO2 21%O2/CO2

residence time (s)


Figure 5-30. Carbon conversion (%) vs. residence time (s) for Beulah and AAW Beulah lignite

5.6 Thermogravimetric modified burning profiles in O2/N2 and O2/CO2 atmospheres The modified burning profiles (as described in Section 4.5) generated under air and 21%O2/79%CO2 are shown in Figures 5-31 and 5-32 respectively for the four coals and a switchgrass sample, and the characteristic temperatures are tabulated in Table 5-10. The aim of this study was to further substantiate that there is a rank effect as discussed in Section 5.2 and investigate the relative importance of the oxidation and gasification reaction at low temperature and low heating rate. The results obtained are very similar to those obtained by Morgan et al.[96] and show an influence of rank. The initial, peak and burnout temperatures are all found to increase with an increase in rank with switchgrass and Pocahontas having the lowest and highest characteristic temperatures respectively. This implies that lower rank coals (fuels) undergo faster conversion. Pisupati[51] has shown the burning profile to be a good tool for evaluation of the relative combustion characteristics of fuels at a laboratory scale. The author, after conducting combustion tests in a pilot-scale research boiler, found that the combustion efficiency was higher for coals with lower initial temperature. The purpose of using the switchgrass sample was to show

76
2

Derivative weight % (%/min)

0 -2 -4 -6 -8 -10 -12 -14 200 400 600 800 1000 1200 1400 Pocahontas Pittsburgh Deitz Beulah Switchgrass

Temperature (K) Figure 5-31. Thermogravimetric burning profile generated under air for the four coals and a switchgrass sample used in the study

Derivative Weight % (%/min)

0 -2 -4 -6 -8 -10 -12 -14 200 400 600 800 1000 1200 1400 Pocahontas Pittsburgh Deitz Beulah Switchgrass

Temperature (K)

Figure 5-32. Thermogravimetric modified burning profile generated under 21%O2/79%CO2 for the four coals and a switchgrass sample used in the study

77 that oxy-fuel combustion could have a beneficial effect for biomass fuels considering its effect on increasing carbon conversion for low-rank coals. From Figures 5-31 and 5-32, no effect of gas atmosphere was seen showing that at low temperatures (below 1,000 K), O2 is a substantially higher reactive gas as compared to CO2 (no influence of char-CO2 reaction). This is the reason why the combustion tests (Section 5.1) were conducted at high temperatures (typically experienced by particles in PC boilers) and the charCO2 reaction was observed due to enhancement in reaction rate.

Table 5-10. Characteristic temperatures under air and oxy-fuel atmospheres

Sample Air

IT (K) Oxy* Air

PT (K) Oxy*

BT (K) Air Oxy*

Dmax Air

Dmax Oxy*

Pocahontas

608

598

763

763

853

851

10

10

Pittsburgh

555.2

549.7

721

721

801

802

10

10

Dietz

485

483

660

658

730

722

13

13

Beulah

423

429

620(1) 731(2)

617(1) 749(2) 562(1) 679(2)

699

691

10(1) 1.6(2)

9(1) 2(2) 10(1) 2(2)

Switchgrass

391

392

563(1) 680(2)

639

645

9(1) 2(2)

*Oxy - 21%O2/79%CO2 (1) and (2) represent temperature values when two peaks are observed

78

Chapter 6 Summary, Conclusions and Recommendations


6.1 Summary and Conclusions The research has successfully identified the effect of coal rank during oxy-fuel combustion in increasing carbon conversion of low-rank coals and the role of the char-CO2 reaction in explaining the differences. The following are the main findings of this work: Combustion tests carried out in a drop-tube reactor showed that oxy-fuel combustion produced a higher carbon conversion than in air for low-rank subbituminous coal and lignite. For the high-rank bituminous coals, combustion in air produced higher carbon conversion than oxy-fuel combustion, thus showing an effect of rank. The reason for this was the char-CO2 reaction taking place at high temperatures. The influence is more pronounced for low-rank coals as they have a higher char-CO2 reactivity and a higher reaction sensitivity to an increase in temperature, leading to higher carbon conversions compared to high-rank coals. The study also showed the effect of residence time with oxy-fuel combustion being most beneficial in terms of carbon conversion for smaller boilers (60-110 MW) firing low-rank coals. These boilers tend to have a residence time of 0.8 seconds. Char-O2 and char-CO2 reactivities of DTR generated char samples showed an influence of rank with low-rank coals having higher reaction rates compared to high-rank coals. Surface area analysis showed a higher concentration of macropores in low-rank coals leading to higher reactivities. Isothermal TGA rate parameters were extrapolated to

79 higher temperatures taking into account diffusion limitations. The results showed that conversion is a function of both overall reactivity and effectiveness factor. Low-rank coals showed both high char-CO2 reactivity and effectiveness factors compared to highrank coals, leading to higher conversions. One exception was for the Pocahontas coal that showed high char-CO2 reactivity at high temperatures but due to the low effectiveness factors, the overall conversion was low. At equivalent temperatures, the effectiveness factors for the CO2 molecule are higher than the O2 molecule and at high temperatures, it is expected that a large portion of the char's internal surface area is consumed by CO2. Gasification catalytic activity of ion exchangeable cations was pronounced at low temperatures in the reactivity profiles in the TGA. However, the same catalytic activity was not seen with ammonium acetate-washed coals at higher temperatures and heating rates as used in DTR, producing similar carbon conversions as the parent coals. The reason for this is possibly due to CaO sintering at high temperatures (>1,000 K) leading to catalyst deactivation. Hence, the higher carbon conversion of low-rank coals in an oxyfuel environment compared to air was purely an effect of rank with no influence of catalytic activity of ion-exchangeable cations. Char particle temperatures in an oxy-fuel environment were successfully modeled and the enhanced effect of the char-CO2 reaction at high temperature and high CO2 partial pressure in lowering the particle temperature was seen. The difference in particle and gas temperature decreased with increase an gas temperature and CO2 partial pressure. It is possible that at high temperature and high CO2 partial pressure, the particle temperature could reduce below that of the gas temperature, thus showing the importance of the charCO2 reaction under such conditions. Modified burning profiles generated under O2/N2 and O2/CO2 atmospheres through nonisothermal TGA studies showed an effect of rank in lower rank fuels demonstrating faster

80 conversion than higher rank fuels which would lead to more efficient combustion in utility scale boilers. The profiles also indicated that O2 is a substantially higher reactive gas compared to CO2 at low temperatures (<1,000 K) with the char-CO2 reaction having no influence under such conditions. Overall this research has highlighted the importance of the char-CO2 reaction especially at high temperatures and high CO2 partial pressures and its consideration in any oxy-fuel combustion modeling effort. The experimental and theoretical results corroborate well to show the exponential increase in rate of the char-CO2 reaction at high temperatures especially for low-rank coals and how it can compete favorably with the oxidation reactions in enhancing the carbon conversion. 6.2 Future Recommendations Oxy-fuel combustion of biomass fuels could show beneficial effects considering that biomass fuels are highly reactive fuels (as seen in Section 5.6) and that oxy-fuel combustion has been shown to increase carbon conversion of low-rank fuels. Incorporate wet recycle with oxy-fuel combustion to study the effect of steam gasification and if there is any beneficial effect including other oxidation and CO2 gasification reactions. Detailed surface area vs. carbon conversion studies to look at the development of pore surface area (macropores, mesopores and micropores) for coals of various ranks under air and oxy-fuel atmospheres. Pilot-scale tests to confirm the rank effect observed at the lab-scale (drop-tube reactor) Computational fluid dynamics (CFD) models to validate experimental results and to study the sensitivity and importance of reaction rate parameters. This way, very efficient combustion predictive models may be developed.

81

Appendix A Calculation of extrapolated rate at combustion condition using intrinsic model Overall rate of char consumption, R per unit external area (kg/m2s) is calculated from, A.1 where Rac is the reaction rate per unit area of solid (kg/m2sPa) given by, A.2 where dp is the diameter of the particle (m) and Ag is the total surface area of solid (m2/kg) and Rd, the rate of gas diffusion (kg/m2sPa) is given by, A.3 In equation A.2, Rs is the intrinsic reactivity of solid per unit total surface area (kg/m2sPa), given by, A.4 where, A is the pre-exponential factor of intrinsic gas-solid reaction (kg/m2sPa) E is the intrinsic activation energy (J/mol), R is the universal gas constant (J/mol.K) and T is the temperature (K) In equation A.2, is the effectiveness factor, ratio of actual rate to the rate in absence of pore diffusion and is calculated by the Thiele modulus approach using the following correlation, A.5 where, is the Thiele modulus for a spherical particle is estimated using the following correlation,

A.6

82 where, Sb is the stoichiometry factor, p is the particle density (kg/m3), Pg is the partial pressure of oxidizer in the bulk gas stream (Pa), De is the effective diffusion coefficient through pore structure (m2/s) and g is the density of the oxidizer gas (kg/m3) In equation A.6, the effective diffusivity, De is estimated from the relation, A.7 where, Dkn, the Knudsen diffusion coefficient is estimated from, A.8 where, Tp is particle temperature (K), Mg is molecular weight of the diffusing gas and rp is mean pore radius of the particle (m) and, DA the molecular diffusion coefficient is estimated from, A.9 Sample Calculation: The following calculation is done for a subbituminous coal. Certain quantities have been assumed and have been obtained from the work of Naredi[21]. dp = 8.1*10-5 m p = 1300 kg/m3 Ag = 100000 m2/kg = 0.5 (assumed) = 2 (assumed) C = 5*10-12 A = 0.0143 kg/m2.s.Pa E = 8.36*104 J/mol rp = 100 A0 (assumed) Sb = 2.67 (assumed) Pg = 20,000 Pa Mg = 32 DA,0 = 1.53*10-5 m2/s Tg = Tp = 1873 K g = 0.273 kg/m3 From equation A.9, DA = 8.16 m2/s From equation A.8, Dkn = 7.42*10-6 m2/s From equation A.7, De = 9.28*10-7 m2/s From equation A.4, Rs = 6.68*10-5 kg/(m2.s.Pa) From equation A.6, = 1.73*103 From equation A.5, = 0.0017

83 From equation A.3, Rd = 1.76*10-5 kg/(m2.s.Pa) From equation A.2, Rac = 1.562*10-7 kg/(m2.s.Pa) From equation A.1, R = 3.1*10-3 kg/(m2.s)

84

Appendix B Example of char conversion calculation Suppose the proximate analysis (wt%) of the feed coal is:

Moisture Volatiles Fixed Carbon Ash Ash (dry basis)

0.205 19.98 73.47 6.341 6.354

Suppose ash (dry basis) content of char generated is 12.044 %, Using eqn. 3.1, Conversion, W (%) = [10000*(12.044-6.354)]/[12.044*(100-6.354)] = 50.45%

85

Appendix C Sample calculation to extract intrinsic rate parameters from an isothermal char reactivity plot

0.5

0 ln (k) 0.0014 0.00145 0.0015 0.00155 0.0016 0.00165 -0.5

-1 y = -10738x + 15.968 R = 0.9943

-1.5

1/T (1/K)

Figure C-1. Plot of ln (k) vs. 1/T for Beulah char-O2 reactivity

From the above plot, ln (k) = -10738 (1/T) + 15.96 So, (-E/R) = -10738 and ln A = 15.966 E = 89275.7 J/mol and A = 8589061.2 mg/g.s BET surface area of Beulah char = 1781.1 m2/kg A = 8589061.2*10-3/1781.1 = 4.8 kg/m2.s.Pa

86

Appendix D Proximate and Ultimate Analysis of DTR chars Table D-1. Proximate and Ultimate Analysis of Pocahontas combustion test chars Gas composition Proximate Analysis(%) Moisture Volatiles Fixed Carbon Ash Ultimate Analysis(%) Carbon Hydrogen Nitrogen Sulfur Oxygen Air0.8 s Oxy0.8s EnOxy0.8s Air1.2s Oxy1.2s EnOxy1.2s Air1.7s Oxy1.7s EnOxy1.7s

0.57 8.63 79.14 11.66

0.37 7.83 82.97 8.83

0.13 6.12 83.74 10.01

1.22 7.96 74.01 16.81

0 5.65 82.30 12.05

0.06 4.44 80.54 14.96

1.30 16.24 57.25 25.21

0.48 8.97 73.74 16.81

0 6.68 69.31 24.01

68.5 0.77 0 1.83 28.9

83.8 0.96 0 0.53 14.71

72.5 1.19 0 0.72 25.59

28.7 1.31 0 0.87 69.12

76.4 6.67 0 0.63 16.3

39.5 1.14 0 0.43 58.93

7.4 1.47 0 2.01 89.12

65.9 2.85 0 2.01 29.24

13.2 0.38 0 0.29 86.13

*EnOxy - 30%O2/70%CO2; SEnOxy - 40%O2/60%CO2

Table D-2. Proximate and Ultimate Analysis of Pittsburgh combustion test chars Gas composition Moisture Air0.8 s 0.04 Oxy0.8s 0.13 Air-1.2 0 Oxy1.2s 0 Air1.7s 0 Oxy1.7s 0 EnOxy- SEnOxy1.7s 1.7s 0.11 0

87 Volatiles Fixed Carbon Ash Ultimate Analysis(%) Carbon Hydrogen Nitrogen Sulfur Oxygen 13.18 52.04 34.74 13.98 64.20 21.69 5.51 45.03 49.46 13.27 56.15 30.58 1.38 25.19 73.43 1.63 14.14 84.23 2.18 13.54 84.17 3.39 4.80 91.81

56.8 0.95 0 0.92 41.33

68.0 2.49 0 1.57 27.94

52.3 0.73 0.04 0.74 46.19

53.4 1.29 0 0.79 44.52

28.4 0.47 0 0.43 70.7

29.3 1.41 0 0.37 68.92

38.8 1.54 0 0.31 59.35

31.6 8.32 0 0 60.08

Table D-3. Proximate and Ultimate Analysis of Dietz combustion test chars Gas composition Proximate Analysis(%) Moisture Volatiles Fixed Carbon Ash Ultimate Analysis(%) Carbon Hydrogen Nitrogen Sulfur Oxygen Air-0.8s Oxy-0.8s Air-1.2s Oxy-1.2s Air-1.7s Oxy-1.7s

1.69 7.65 56.85 33.81

0.38 9.27 13.37 76.98

0.24 6.57 25.83 67.36

0 1.02 0.06 98.92

0 2.23 0.99 96.78

0 1.60 0.13 98.27

59.9 0.60 0.16 1.02 38.32

22.0 0.92 0 0.93 76.15

20.9 0.7 0 0.57 77.83

0.84 0.56 0 0.21 98.39

1.84 0.63 0 0.47 97.06

1.23 0.66 0 0.34 97.77

88 Table D-4. Proximate and Ultimate Analysis of Beulah combustion test chars Air-0.8s Gas compositi on Proximate Analysis(%) Moisture 0.67 Volatiles Fixed Carbon Ash Ultimate Analysis(%) Carbon Hydrogen Nitrogen Sulfur Oxygen 6.43 44.57 48.33 Oxy-0.8s Air-1.2s Oxy-1.2s Air-1.7s Oxy-1.7s EnOxy1.7s

0.36 8.54 30.52 60.58

0 3.82 11.13 85.05

0 2.21 0.12 97.67

0.02 3.44 12.34 84.20

0 3.03 0.26 96.71

0 6.34 0 93.66

48.9 0.44 0 0.86 49.8

42 0.49 0 0.99 56.52

16.0 0.64 0 0.79 82.57

3.16 0.58 0 0.60 95.66

16.1 0.34 0 1.14 82.42

3.25 0.91 0 1.59 94.25

4.7 0.28 0 0.57 94.45

Table D-5. Proximate and Ultimate Analysis of AAW Dietz and Beulah combustion test chars Gas composition Proximate Analysis (%) Moisture Volatiles Fixed Carbon Ash AAW Dietz Oxy - 0.8s AAW Dietz Oxy - 1.7s AAW Beulah Oxy - 0.8s AAW Beulah Oxy-1.7s

0.11 6.92 5.31 87.66

0 1.29 0.25 98.46

0.65 10.09 26.28 62.98

0 3.77 4.04 92.19

89 Ultimate Analysis (%) Carbon Hydrogen Nitrogen Sulfur Oxygen

18.6 0 0 0.88 80.52

0.88 0 0 1.52 97.6

37.0 0.92 0 0.86 61.22

5.55 0.02 0 0.14 94.29

90

Appendix E Plots of Rd, Rs, Rac and R as a function of temperature for the four coals
1 0.1 500 0.01 0.001 0.0001 0.00001 0.000001 0.0000001 1E-08 1E-09 1E-10 Temperature (K) log(Rs) log(Rac) log(Rd) log(R) 700 900 1100 1300 1500 1700 1900

Figure E-1. Extrapolated oxidation Rs, Rac, Rd and R for Pocahontas coal

1 700 0.01 0.0001 0.000001 1E-08 1E-10 1E-12 1E-14 Temperature (K) log(Rs) log(Rac) log(Rd) log(R) 900 1100 1300 1500 1700 1900

Figure E-2. Extrapolated gasification Rs, Rac, Rd and R for Pocahontas coal

91

1.00E+00 1.00E-01 500 1.00E-02 1.00E-03 1.00E-04 1.00E-05 1.00E-06 1.00E-07 1.00E-08 1.00E-09 1.00E-10 Temperature (K) log(Rs) log(Rac) log(Rd) log(R) 700 900 1100 1300 1500 1700 1900

Figure E-3. Extrapolated oxidation Rs, Rac, Rd and R for Pittsburgh coal

1.00E+00 1.00E-02 1.00E-04 1.00E-06 1.00E-08 1.00E-10 1.00E-12 1.00E-14 1.00E-16 Temperature (K) log(Rs) log(Rac) log(Rd) log(R) 700 900 1100 1300 1500 1700 1900

Figure E-4. Extrapolated gasification Rs, Rac, Rd and R for Pittsburgh coal

92

1 0.1 500 0.01 0.001 0.0001 0.00001 0.000001 0.0000001 1E-08 1E-09 Temperature (K) log(Rs) log(Rac) log(Rd) log(R) 700 900 1100 1300 1500 1700 1900

Figure E-5. Extrapolated oxidation Rs, Rac, Rd and R for Dietz coal

1 0.1 500 0.01 0.001 0.0001 0.00001 0.000001 0.0000001 1E-08 1E-09 1E-10 1E-11 1E-12 1E-13

700

900

1100

1300

1500

1700

1900

log(Rs) log(Rac) log(Rd) log(R)

Temperature (K)

Figure E-6. Extrapolated gasification Rs, Rac, Rd and R for Dietz coal

93

1 0.1 500 0.01 0.001 0.0001 0.00001 0.000001 0.0000001 1E-08 1E-09 Temperature (K) log(Rs) log(Rac) log(Rd) log(R) 700 900 1100 1300 1500 1700 1900

Figure E-7. Extrapolated oxidation Rs, Rac, Rd and R for Beulah lignite

1 0.1 500 0.01 0.001 0.0001 0.00001 0.000001 0.0000001 1E-08 1E-09 1E-10 1E-11 1E-12 1E-13

700

900

1100

1300

1500

1700

1900

log(Rs) log(Rac) log(Rd) log(R)

Temperature (K)

Figure E-8. Extrapolated gasification Rs, Rac, Rd and R for Beulah lignite

94

Appendix F Sample char particle temperature calculation The right hand side of equation 5.2.1 consists of the reaction term, convective term and radiation term in the same order. The following char temperature calculation is done at a gas temperature of 1873 K and for a gas atmosphere of the composition 12%O2/88%CO2. Radiative term: = 5.67*10-8 W/m2K4 = 0.5 Tg = Tw = 1873 K so, (Tp4 - Tw4) = 0.5*5.67*10-8*( Tp4 - 18734) = (1.814*10-16 Tp4 - 2.2325*10-3) W Convective term: dp = 45.146 up = 0.6 m/s Vp = 4.8154*10-14 m3 Ap = 6.3998*10-9 m2 p = 1000 kg/m3 mp = 4.8154*10-11 kg convective term = h*Ap*(Tp - 1873) = h*6.3998*10-9*(Tp-1873) properties of the surrounding gas atmosphere, (@1873 K for 12%O2/88%CO2) = 6.13*10-5 kg/m.s Cp (@1873 K for 12%O2/88%CO2) = 1338.90 J/kg.K k (@1873 K for 12%O2/88%CO2) = 0.1376 W/m.K

95 (@1873 K for 12%O2/88%CO2) = 0.2768 kg/m3 Pr = Cp*/k = 1338.9*6.13*10-5/0.1376 = 0.5966 Red = up*dp */ = 0.6*45.146*10-6/0.0002216 = 0.12336 h = (2 + 0.6*Red*Pr1/3)*k/dp = (2 + 0.6*0.12336*(0.5966)1/3)*0.1376/45.146*10-6 = 6286.9 W/m2.K so, convective term = 6238.9*6.3998*10-9*(Tp - 1873) = (4.0235*10-5 Tp - 0.07536) W Reaction term: Vit = 82.8; In = 13.2 nO2 = 0.5; nCO2 = 0.8 fmac = 1.68*82.8 - 0.6*13.2 = 131.184 mp*fmac = 4.81554*10-11 * 131.184 = 6.3165*10-9 @ 1,873K, 12%O2 = 0.08 for char-O2 and 0.37 for char-CO2 Ri*(Pg)n* = (1.33*105*exp(-123*103/8.314*Tp)*9200*1000*0.08*(0.12)0.5) - (3.5*105*exp(189*103/8.314*Tp)*14390*1000*0.37*(0.88)0.8) so, reactive term = 6.3165*10-9*(3.3909*1010*exp(-14794.32/Tp) 1.6824*1012*exp(-

22732.74/Tp)) Left hand side term: Cp = -0.218 + 3.807*10-3*Tp - 1.758*10-6 * Tp * Tp assuming Tp = 1600 K, Cp = 1.37272*103 J/kg so, mp*Cp*dTp/dt = 4.8154*10-11*1.37272*103*dTp/dt = 6.6096*10-8dTp/dt

96

References
1. 2. 3. 4. 5. Information on Global Warming Potential. 15 June 2004, FCCC-TP-2004-3, United Nations Framework Convention Climate Change. International Oxy-Combustion Research Network for CO2 Capture: Report on Inaugural (1st) Workshop. 2006/4, July 2006, IEA Greenhouse Gas R&D Programme (IEA GHG). Inventory of U.S. Greenhouse Gas Emissions and Sinks. 2007, website, U.S. Environmental Protection Agency. Murphy, J.J. and C.R. Shaddix, Combustion kinetics of coal chars in oxygen-enriched environments. Combustion and Flame, 2006, 144(4): p. 710-729. Davison, J., Electricity systems with near-zero emissions of CO2 based on wind energy and coal gasification with CCS and hydrogen storage. International Journal of Greenhouse Gas Control, 2009, 3(6): p. 683-692. Plasynski, S.I., Litynski, J.T., Mcllvried, H.G., and Srivastava, R.D., Progress and New Developments in Carbon Capture and Storage. Critical Reviews in Plant Sciences, 2009, 28(3): p. 123-138. McDonald, D.K., Moorman, S., Darde, A., and de Limon, S., Oxy-Combustion: A Promising Technology for Coal-Fired Plants. Power, 2011, 155(1): p. 52-57. Liu, H., R. Zailani, and B.M. Gibbs, Comparisons of pulverized coal combustion in air and in mixtures of O2/CO2. Fuel, 2005, 84(7-8): p. 833-840. Liu, H. and Shao, Y. J., Predictions of the impurities in the CO2 stream of an oxy-coal combustion plant. Applied Energy, 2010, 87(10): p. 3162-3170. Giavarini, C., F. Maccioni, and M.L. Santarelli, CO2 sequestration from coal fired power plants. Fuel, 2010, 89(3): p. 623-628. Buhre, B.J.P., Elliott, L. K., Sheng, C. D., Gupta, R. P., and Wall, T. F., Oxy-fuel combustion technology for coal-fired power generation. Progress in Energy and Combustion Science, 2005, 31(4): p. 283-307. Okawa, M., Kimura, N., Kiga, T., Takano, S., Arai, K., and Kato, M., Trial design for a CO2 recovery power plant by burning pulverized coal in O2/CO2. Energy Conversion and Management, 1997, 38(1): p. S123-S127. Burdyny, T. and H. Struchtrup, Hybrid membrane/cryogenic separation of oxygen from air for use in the oxy-fuel process. Energy, 2010, 35(5): p. 1884-1897. Kimura, N., Omata, K., Kiga, T., Takano, S., and Shikisima, S., The Characteristics of Pulverized Coal Combustion in O2/CO2 Mixtures for CO2 Recovery. Energy Conversion and Management, 1995, 36(6-9): p. 805-808. Tan, Y.W., Croiset, E., Douglas, M. A., and Thambimuthu, K. V., Combustion characteristics of coal in a mixture of oxygen and recycled flue gas. Fuel, 2006, 85(4): p. 507-512. Zhou, W. and Moyeda, D., Process Evaluation of Oxy-fuel Combustion with Flue Gas Recycle in a Conventional Utility Boiler. Energy & Fuels, 2010, 24(3): p. 2162-2169. Kiga, T., Takano, S., Kimura, N., Omata, K., Okawa, M., Mori, T., and Kato, M., Characteristics of pulverized-coal combustion in the system of oxygen recycled flue gas combustion. Energy Conversion and Management, 1997, 38(1): p. S129-S134. Yamada, T., Kiga, T., Okawa, M., Omata, K., Kimura, N., Arai, K., Mori, T., and Kato, M., Characteristics of pulverized-coal combustion in CO2-recovery power plant applied O2/CO2 combustion. JSME International Journal Series B-Fluids and Thermal Engineering, 1998, 41(4): p. 1017-1022.

6.

7. 8. 9. 10. 11.

12.

13. 14.

15.

16. 17.

18.

97 19. Borrego, A.G. and D. Alvarez, Comparison of chars obtained under oxy-fuel and conventional pulverized coal combustion atmospheres. Energy & Fuels, 2007, 21(6): p. 3171-3179. Bejarano, P.A. and Y.A. Levendis, Single-coal-particle combustion in O2/N2 and O2/CO2 environments. Combustion and Flame, 2008, 153(1-2): p. 270-287. Naredi, P., Coal combustion in O2/CO2 medium for power generation: Numerical modeling of char burnout, NOx and CO emissions. 2009, PhD Thesis, The Pennsylvania State University. Hampartsoumian, E., Murdoch, P. L., Pourkashanian, M., Trangmar, D. T., and Williams, A., The Reactivity of Coal Chars Gasified in a Carbon-Dioxide Environment. Combustion Science and Technology, 1993, 92(1-3): p. 105-121. Naredi, P. and S.V. Pisupati, Interpretation of char reactivity profiles obtained using a thermogravimetric analyzer. Energy & Fuels, 2008, 22(1): p. 317-320. Hjartstam, S., Andersson, K., Johnsson, F., and Leckner, B., Combustion characteristics of lignite-fired oxy-fuel flames. Fuel, 2009, 88(11): p. 2216-2224. Smart, J., Lu, G., Yan, Y., and Riley, G., Characterisation of an oxy-coal flame through digital imaging. Combustion and Flame, 2010, 157(6): p. 1132-1139. Andersson, K., Johansson, R., Hjartstam, S., Johnsson, F., and Leckner, B., Radiation intensity of lignite-fired oxy-fuel flames. Experimental Thermal and Fluid Science, 2008, 33(1): p. 67-76. Smart, J.P., O'Nions, P., and Riley, G. S., Radiation and convective heat transfer, and burnout in oxy-coal combustion. Fuel, 2010, 89(9): p. 2468-2476. Shaddix, C.R. and A. Molina, Particle imaging of ignition and devolatilization of pulverized coal during oxy-fuel combustion. Proceedings of the Combustion Institute, 2009, 32(2): p. 2091-2098. Li, Q.Z., Zhao, C. S., Chen, X. P., Wu, W. F., and Li, Y. J., Comparison of pulverized coal combustion in air and in O2/CO2 mixtures by thermo-gravimetric analysis. Journal of Analytical and Applied Pyrolysis, 2009, 85(1-2): p. 521-528. Zhang, L.A., Binner, E., Qiao, Y., and Li, C. Z., In situ diagnostics of Victorian brown coal combustion in O2/N2 and O2/CO2 mixtures in drop-tube furnace. Fuel, 2010, 89(10): p. 2703-2712. Qiao, Y., Zhang, L. A., Binner, E., Xu, M. H., and Li, C. Z., An investigation of the causes of the difference in coal particle ignition temperature between combustion in air and in O2/CO2. Fuel, 2010, 89(11): p. 3381-3387. Rathnam, R.K., Elliott, L. K., Wall, T. F., Liu, Y. H., and Moghtaderi, B., Differences in reactivity of pulverised coal in air (O2/N2) and oxy-fuel (O2/CO2) conditions. Fuel Processing Technology, 2009, 90(6): p. 797-802. Duan, L.B., Zhao, C. S., Zhou, W., Qu, C. R., and Chen, X. P., Investigation on Coal Pyrolysis in CO2 Atmosphere. Energy & Fuels, 2009, 23(7): p. 3826-3830. Zhang, L.A., Binner, E., Qiao, Y., and Li, C. Z., In situ diagnostics of Victorian brown coal combustion in O2/N2 and O2/CO2 mixtures in drop-tube furnace. Fuel, 2010, 89(10): p. 2703-2712. Saastamoinen, J.J., Aho, M. J., Hamalainen, J. P., Hernberg, R., and Joutsenoja, T., Pressurized pulverized fuel combustion in different concentrations of oxygen and carbon dioxide. Energy & Fuels, 1996, 10(1): p. 121-133. Hong, J.S., Chaudhry, G., Brisson, J. G., Field, R., Gazzino, M., and Ghoniem, A. F., Analysis of oxy-fuel combustion power cycle utilizing a pressurized coal combustor. Energy, 2009, 34(9): p. 1332-1340.

20. 21.

22.

23. 24. 25. 26.

27. 28.

29.

30.

31.

32.

33. 34.

35.

36.

98 37. Deng, S.M. and R. Hynes, Thermodynamic Analysis and Comparison on Oxy-Fuel Power Generation Process. Journal of Engineering for Gas Turbines and Power-Transactions of the ASME, 2009, 131(5): p.707-714. Li, Q.Z., Zhao, C. S., Chen, X. P., Wu, W. F., and Lin, B. Q., Properties of char particles obtained under O2/N2 and O2/CO2 combustion environments. Chemical Engineering and Processing, 2010, 49(5): p. 449-459. Laurendeau, N.M., Heterogeneous Kinetics of Coal Char Gasification and Combustion. Progress in Energy and Combustion Science, 1978, 4(4): p. 221-270. Homma, S., Ogata, S., Koga, J., and Matsumoto, S., Gas-solid reaction model for a shrinking spherical particle with unreacted shrinking core. Chemical Engineering Science, 2005, 60(18): p. 4971-4980. Wang, C.S., Berry, G. F., Chang, K. C., and Wolsky, A. M., Combustion of Pulverized Coal Using Waste Carbon-Dioxide and Oxygen. Combustion and Flame, 1988, 72(3): p. 301-310. Mahajan, O.P., R. Yarzab, and P.L. Walker, Unification of Coal-Char Gasification Reaction-Mechanisms. Fuel, 1978, 57(10): p. 643-646. Radovic, L.R., P.L. Walker, and R.G. Jenkins, Importance of Carbon Active-Sites in the Gasification of Coal Chars. Fuel, 1983, 62(7): p. 849-856. Liu, G.S., Tate, A. G., Bryant, G. W., and Wall, T. F., Mathematical modeling of coal char reactivity with CO2 at high pressures and temperatures. Fuel, 2000, 79(10): p. 1145-1154. Cai, H.Y., Guell, A. J., Chatzakis, I. N., Lim, J. Y., Dugwell, D. R., and Kandiyoti, R., Combustion reactivity and morphological change in coal chars: Effect of pyrolysis temperature, heating rate and pressure. Fuel, 1996, 75(1): p. 15-24. Gale, T.K., C.H. Bartholomew, and T.H. Fletcher, Effects of pyrolysis heating rate on intrinsic reactivities of coal chars. Energy & Fuels, 1996, 10(3): p. 766-775. Jenkins, R.G., S.P. Nandi, and P.L. Walker, Reactivity of Heat-Treated Coals in Air at 500 Degrees C. Fuel, 1973, 52(4): p. 288-293. Senneca, O., P. Salatino, and S. Masi, Heat treatment-induced loss of combustion reactivity of a coal char: the effect of exposure to oxygen. Experimental Thermal and Fluid Science, 2004, 28(7): p. 735-741. Walker, P.L., Matsumoto, S., Hanzawa, T., Muira, T., and Ismail, I. M. K., Catalysis of Gasification of Coal-Derived Cokes and Chars. Fuel, 1983, 62(2): p. 140-149. Niu, S.L., Lu, C. M., Han, K. H., and Zhao, J. L., Thermogravimetric analysis of combustion characteristics and kinetic parameters of pulverized coals in oxy-fuel atmosphere. Journal of Thermal Analysis and Calorimetry, 2009, 98(1): p. 267-274. Pisupati, S.V., An examination of burning profiles as a tool to predict combustion behavior of coals. Abstracts of Papers of the American Chemical Society, 1996, 211: p. 13-18. Radovic, L.R., P.L. Walker, and R.G. Jenkins, Effect of Lignite Pyrolysis Conditions on Calcium-Oxide Dispersion and Subsequent Char Reactivity. Fuel, 1983, 62(2): p. 209212. Miura, K., K. Hashimoto, and P.L. Silveston, Factors Affecting the Reactivity of Coal Chars during Gasification, and Indexes Representing Reactivity. Fuel, 1989, 68(11): p. 1461-1475. Kapteijn, F., H. Porre, and J.A. Moulijn, CO2 Gasification of Activated Carbon Catalyzed by Earth Alkaline Elements. AICHE Journal, 1986, 32(4): p. 691-695. Prinsloo, F.F. and Schneider, M., Kinetics of catalyzed CO2 coal gasification reactions. Proceedings of 23rd Biennial Conference on carbon, 1997, p. 426-427.

38.

39. 40.

41.

42. 43. 44.

45.

46. 47. 48.

49. 50.

51.

52.

53.

54. 55.

99 56. 57. Ye, D.P., Gasification of South Australian Lignite, in Chemical Engineering. 1994, Thesis, University of Adelaide, Adelaide. Chen, G.H., Effect of replacement of MgO by CaO on sintering, crystallization and properties of MgO-Al2O3-SiO2 system glass-ceramics. Journal of Materials Science, 2007, 42(17): p. 7239-7244. Borrego, A.G., Osorio, E., Casal, M. D., and Vilela, A. C. F., Coal char combustion under a CO2-rich atmosphere: Implications for pulverized coal injection in a blast furnace. Fuel Processing Technology, 2008, 89(11): p. 1017-1024. Li, X.C., Rathnam, R. K., Yu, J. L., Wang, Q., Wall, T., and Meesri, C., Pyrolysis and Combustion Characteristics of an Indonesian Low-Rank Coal under O2/N2 and O2/CO2 Conditions. Energy & Fuels, 2010, 24(1): p. 160-164. Brix, J., P.A. Jensen, and A.D. Jensen, Coal devolatilization and char conversion under suspension fired conditions in O2/N2 and O2/CO2 atmospheres. Fuel, 2010, 89(11): p. 3373-3380. Varhegyi, G., Szabo, P., Jakab, E., Till, F., and Richard, J. R., Mathematical modeling of char reactivity in Ar-O2 and CO2-O2 mixtures. Energy & Fuels, 1996, 10(6): p. 12081214. Varhegyi, G. and F. Till, Comparison of temperature-programmed char combustion in CO2-O2 and Ar-O2 mixtures at elevated pressure. Energy & Fuels, 1999, 13(2): p. 539540. Hodge, E.M., The coal char-CO2 reaction at high temperature and pressure, in Chemical and Industrial Chemistry. 2009, Thesis, University of New South Wales. Tree, D.R., Mackrory, A.J., and Fletcher, T., A Mechanistic Investigation of Nitrogen Evolution and Corrosion with Oxy-Combustion. 2009, DE-FG26-05NT42530, DOE. Brown, B.W., Smoot, L. D., Smith, and P. J., Hedman, P. O., Measurement and Prediction of Entrained-Flow Gasification Processes. Aiche Journal, 1988, 34(3): p. 435446. Goetz, G.J., Nsakala, N.Y., Patel, R.L., and Lao, T.C. Combustion and Gasification Characteristics of Chars from Four Commercially Significant Coals of Different Rank. in Proceedings of the second annual EPRI Contractors' Conference on Coal Gasification. 1983, p. 15-1-15-27. Douglas Smoot, L., and Smith, P.J., Coal Combustion and Gasification. The Plenum Chemical Engineering Series, ed., D. Luss. 1985, Plenum Press. Cope, R.F., L.D. Smoot, and P.O. Hedman, Effects of Pressure and Coal Rank on Carbon Conversion in an Entrained-Coal Gasifier. Fuel, 1989, 68(6): p. 806-808. Roberts, D.G. and D.J. Harris, Char gasification with O2, CO2, and H2O: Effects of pressure on intrinsic reaction kinetics. Energy & Fuels, 2000, 14(2): p. 483-489. Mann, A.P. and J.H. Kent, A Computational Study of Heterogeneous Char Reactions in a Full-Scale Furnace. Combustion and Flame, 1994, 99(1): p. 147-156. Le Manquais, K., Snape, C., McRobbie, I., Barker, J., and Pellegrini, V., Comparison of the Combustion Reactivity of TGA and Drop Tube Furnace Chars from a Bituminous Coal. Energy & Fuels, 2009, 23(9): p. 4269-4277. Peng, F.F., I.C. Lee, and R.Y.K. Yang, Reactivities of in-Situ and Ex-Situ Coal Chars during Gasification in Steam at 1000-1400-Degrees-C. Fuel Processing Technology, 1995, 41(3): p. 233-251. Kwon, T.W., S.D. Kim, and D.P.C. Fung, Reaction-Kinetics of Char-CO2 Gasification. Fuel, 1988, 67(4): p. 530-535.

58.

59.

60.

61.

62.

63. 64. 65.

66.

67. 68. 69. 70. 71.

72.

73.

100 74. Ochoa, J., Cassanello, M. C., Bonelli, P. R., and Cukierman, A. L., CO2 gasification of Argentinean coal chars: a kinetic characterization. Fuel Processing Technology, 2001, 74(3): p. 161-176. Tsai, C.Y. and A.W. Scaroni, Pyrolysis and Combustion of Bituminous Coal Fractions in an Entrained-Flow Reactor. Energy & Fuels, 1987, 1(3): p. 263-269. Pak, P.S., Y.D. Lee, and K.Y. Ahn, Characteristics and economic evaluation of a power plant applying oxy-fuel combustion to increase power output and decrease CO2 emission. Energy, 2010, 35(8): p. 3230-3238. Liszka, M. and A. Ziebik, Coal-fired oxy-fuel power unit - Process and system analysis. Energy, 2010, 35(2): p. 943-951. Xiong, J., Zhao, H. B., Zheng, C. G., Liu, Z. H., Zeng, L. D., Liu, H., and Qiu, J. R., An economic feasibility study of O2/CO2 recycle combustion technology based on existing coal-fired power plants in China. Fuel, 2009, 88(6): p. 1135-1142. Normann, F., H. Thunman, and F. Johnsson, Process analysis of an oxygen lean oxy-fuel power plant with co-production of synthesis gas. Energy Conversion and Management, 2009, 50(2): p. 279-286. Lee, K., Kim, S., Choi, S., and Kim, T., Performance Evaluation of an Oxy-coal-fired Power Plant. Journal of Thermal Science and Technology, 2009, 4(3): p. 400-407. Zhou, W., Swanson, L., Moyeda, D., and Xu, G. A., Process Evaluation of Biomass Cofiring and Reburning in Utility Boilers. Energy & Fuels, 2010, 24(8): p. 4510-4517. Krishnamurthy, N., personal communication, 2011, The Pennsylvania State University. Man, C.K., Gibbins, J. R., Witkamp, J. G., and Zhang, J., Coal characterisation for NOx prediction in air-staged combustion of pulverised coals. Fuel, 2005, 84(17): p. 21902195. Ballantyne, T.R., P.J. Ashman, and P.J. Mullinger, A new method for determining the conversion of low-ash coals using synthetic ash as a tracer. Fuel, 2005, 84(14-15): p. 1980-1985. Wagoner, C.L. and A.F. Duzy, Burning Profiles for Solid Fuels. ASME Annual meeting, 1967, 67-WA/FU-4. Miller, S.F. and Miller, B.G., The occurrence of inorganic elements in various biofuels and its effect on ash chemistry and behavior and use in combustion products. Fuel Processing Technology, 2007, 88(11-12):p. 1155-1164. Alvarez, D. and A.G. Borrego, The evolution of char surface area along pulverized coal combustion. Energy & Fuels, 2007, 21(2): p. 1085-1091. Gopalakrishnan, R. and C.H. Bartholomew, Effects of CaO, high-temperature treatment, carbon structure, and coal rank on intrinsic char oxidation rates. Energy & Fuels, 1996, 10(3): p. 689-695. Adschiri, T. and T. Furusawa, Relation between CO2-Reactivity of Coal Char and Bet Surface-Area. Fuel, 1986. 65(7): p. 927-931. Wigmans, T., H. Haringa, and J.A. Moulijn, Nature, Activity and Stability of Active-Sites during Alkali-Metal Carbonate-Catalyzed Gasification Reactions of Coal Char. Fuel, 1983, 62(2): p. 185-189. Pallares, J., I. Arauzo, and A. Williams, Integration of CFD codes and advanced combustion models for quantitative burnout determination. Fuel, 2007, 86(15): p. 22832290. Eisermann, W., P. Johnson, and W.L. Conger, Estimating Thermodynamic Properties of Coal, Char, Tar and Ash. Fuel Processing Technology, 1980, 3(1): p. 39-53. Serbin, S.I. and Matveev, I.B., Modeling of the Coal Gasification Processes in a Hybrid Plasma Torch. IEEE Transactions on Plasma Science, 2007, 35(6): p. 1639-1647.

75. 76.

77. 78.

79.

80. 81. 82. 83.

84.

85. 86.

87. 88.

89. 90.

91.

92. 93.

101 94. 95. 96. Sturgeon, D., Oxyfuel Combustion Technology, in Official Opening of the OxyCoal Clean Combustion Test Facility. Technical Seminar, 2009, Doosan Babcock Energy. Cope, R.F., C.B. Arrington, and W.C. Hecker, Effect of Cao Surface-Area on Intrinsic Char Oxidation Rates for Beulah Zap Chars. Energy & Fuels, 1994, 8(5): p. 1095-1099. Morgan, P.A., S.D. Robertson, and J.F. Unsworth, Combustion Studies by Thermogravimetric Analysis .1. Coal Oxidation. Fuel, 1986, 65(11): p. 1546-1551.

Вам также может понравиться