OVERDUE FINES: 25¢ per day per item RETURNING LIBRARY MATERIALS: N Place in book return to rent charge from cirgulation rec: A MATHEMATICAL MODEL FOR THE PYROLYSIS OF OIL SHALE USING DISTRIBUTION OF ACTIVATION ENERGY KINETICS BY John Arthur Marlatt A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Chemical Engineering 1980 ABSTRACT A MATHEMATICAL MODEL FOR THE PYROLYSIS OF OIL SHALE USING DISTRIBUTION OF ACTIVATION ENERGY KINETICS BY John Arthur Marlatt Mathematical simulations of commercial oil shale retort Operations require an understanding of the rate processes which occur in individual oil shale particles. Previously, the kinetics of oil shale devolatilization have been modeled using simple kinetic schemes but have failed to predict the production of volatile products over a wide range of pyrolysis conditions. A prototype model was developed in this study which incorporated the kinetics of devolatilization, a mechanism for intraparticle mass tranSport of volatile products, and the kinetics for intra- particle product degradation. The complex kinetics associated with oil shale devolatilization were interpreted using the distribution of activation energies theory develOped pre- viously for coal pyrolysis. A Gaussian distribution function was used to represent the activation energies for the reactions producing volatile products. The parameters in the model were estimated from isothermal weight loss data John Arthur Marlatt for western oil shale. A favorable comparison was made between the model prediction and data from a nonisothermal eXperiment. Dedicated to Janet Lee Marlatt ii ACKNOWLEDGMENTS The author gratefully acknowledges the guidance and editorial assistance of Dr. Charles Petty. Special appreciation is also expressed to my wife, Jan, for her encouragement and help in the preparation of this thesis. iii TABLE OF CONTENTS Page LIST OF TABLES vii LIST OF FIGURES. . . . . . . . . . . viii LIST OF NOMENCLATURE . ix Chapter 1. INTRODUCTION . . . . l 1.1 Physical Description of Oil Shale 2 1.2 Description of Oil Shale Pyrolysis. 4 1.3 Kinetic Models for Pyrolysis. . . 7 1.4 Objectives of This Reseach . 15 2. A MATHEMATICAL MODEL FOR THE PYROLYSIS OF OIL SHALE. . . . . . . . . . 17 2.1 Devolatilization Kinetics. 18 2.2 Coking Kinetics . 21 2.3 Material and Energy Balances. . 22 2.3.1 Material Balance for the Reactive Volatile Species in the Gas Phase of the Particle. . . . . 22 2.3.2 Material Balance for the Non- reactive Volatile Species in the Gas Phase of the Particle . 26 2.3.3 A Material Balance for the Volatile Species Bound to the Solid Phase . . 27 2.3.4 Material Balance for the Weight of Volatiles Collected . . . 28 2.3.5 Energy Balance . . . . . . 28 3. SPECIAL CASES TO BE STUDIED . . . . . . 31 3.1 Chemical Kinetics Controlling Regime-- Isothermal . . . . . . . . . 31 3.1.1 Reactive Species . . 31 3.1.2 Nonreactive Species . . . 32 3 1.3 Weight of Volatiles Collected . 33 iv Chapter 3.2 Chemical Kinetics Controlling Regime-- Nonisothermal . . . 3.3 Diffusional Limitations for Pyrolysis-- Isothermal . . . . . . . . . 3.3 l Reactive Species . . . . 3.3.2 Nonreactive Species . . . 3 3 3 Weight of Volatiles Collected 3.4 Convection Limitations for Pyrolysis-- Isothermal . . . . . . . . . 3.4.1 Nonreactive Species . . 3.4.2 Reactive Species . . . . 3.4.3 Weight of Volatiles Collected 4. THE EFFECT OF KINETIC PARAMETERS ON "OIL" YIELD . . . . . . . . . . 4.1 The Effect of 0 on "Oil" Yield . . 4.2 The Effect of E0 on "Oil" Yield. . 4.3 The Effect of k0 on "Oil" Yield. 4.4 The Effect of the Heating Rate on Oil Yield' . . . . . . . . . 5. PARAMETER ESTIMATES USING DATA FOR WESTERN OIL SHALE. . . . . . . . . . . 5.1 Estimates of Parameters for the Distri- bution of Activation Energy Model . 5.2 Estimates of Parameters for a Lumped First Order Model. 5.3 Prediction of Nonisothermal Pyrolysis Using Kinetic Parameters Estimated from Isothermal Data. . 5.4 Prediction of Isothermal Pyrolysis for Large Particles . . . . . 6. MODEL PREDICTIONS FOR EASTERN OIL SHALE 6.1 Chemical Kinetics Controlling Regime . 6.2 Diffusional Limitations for Pyrolysis. 6.3 Convection Limitations for Pyrolysis 7. CONCLUSIONS AND RECOMMENDATIONS FOR FURTHER RESEARCH . . . . . . . . . . . . Page 34 36 36 37 38 4O 40 41 43 48 49 51 S4 56 59 59 65 66 7O 75 77 79 79 APPENDICES Appendix A. Procedure for Estimating the Devolatilization Parameters . . . . . . . . . . . . B. Computer Program for Estimating the Devolatilization Parameters . . . . . C. Procedure for Estimating Y and C*(l - e). D. Procedure for Estimating the Kinetic Para- meters in the Lumped First Order Model. . E. Computer Program for Calculating the Volume of Oil Collected for Nonisothermal Conditions for the Chemical Kinetics Controlling Regime. F. Computer Program for Calculating the Weight of Volatiles Collected as a Function of Time for Isothermal Convection Limitations . . . . LIST OF REFERENCES . . . . . . . . . . . . vi Page 87 91 97 101 103 107 111 LIST OF TABLES Table Page 1. Expected Values for the Kinetic Parameters in the Distribution Model . . . . . . . . 4S 2. Physical and Transport PrOperties for Western Oil Shale . . . . . . . 46 3. Physical and Transport Properties for Eastern Oil Shale . . . . . . . . . . . 47 vii Figure 1. 10. 11. 12. 13. 14. 15. LIST OF FIGURES A schematic representation of an oil shale particle . The The The The The effect effect effect effect effect Results of of of of of of g on the "oil" yield at 648°K 0 on the "oil" yield at 775°K E0 on the "oil" yield at 775°K. k0 on the "oil" yield at 775°K. the heating rate on oil yield parameter estimation at 648°K for the distribution and first order models . . . Page 23 50 52 53 55 57 61 Results of parameter estimation at 673°K to obtain Distribution model with diffusional limtations at 703°K Distribution model with convection limitations at 703°K y and C*(1 - e). . . . . . . . . . 'The distribution and first order models for a heating rate of 0.033°K/sec. . . . . The effect of the heating rate . . . . Distribution and second order models for the chemical kinetics controlling regime. . . Distribution and second order models for diffusional limitations . . . . . Distribution and second order model for convection limitations . . . . . . . viii 63 67 69 71 73 78 80 82 iR iBR iBR iRco iNR iBNR iBNR CiNRm LIST OF NOMENCLATURE (kg sec} collection of constants given on page 44 [ 3 heating rate [°K/Sec] collection of contants given on page 44 [“—§&Q-] (m sec) mass concentration of a reactive volatile Species i in the gas phase of the particle [kg/m3 of gas phase in the particle] mass concentration of a reactive volatile Specie? i bound in the solid phase of the particle [kg/m of solid phase in the particle] initial mass concentration of a reactive volatile species i bound in the solid phase of the particle [kg/m3 of solid phase in the particle] mass concentration of a reactive volatile Species i in the sweep gas far from the particle [kg/m3 of gas phase in the particle] mass concentration of a nonreactive volatile Species i in the gas phase of the particle [kg/m3 of gas phase in the particle] mass concentration of a nonreactive volatile Species i in the solid phase of the particle [kg/m3 of solid phase in the particle] initial mass concentration of a nonreactive volatile Species i bound in the solid phase of the particle [kg/m3 of solid phase in the particle] mass concentration of a nonreactive volatile s ecies i in the sweep gas far from the particle [kng of gaS phase in the particle] ix NR Roo NR°° C* total mass concentration of all reactive volatile Species in the gas phase of the particle [kg/m3 of gas phase in the particle} total mass concentration of all nonreactive volatile species in the gas phase of the particle [kg/m of gas phase in the particle] total mass concentration of all reactive volatile Species in the sweep gas far from the particle [kg/m3 of gas phase in the particle] total mass concentration of all nonreactive volatile species in the sweep gas far from the particle [kg/m3 of gas phase in the particle] total volatile content of the particle [kg/m3 of solid phase in the particle] heat capacity of the solid phase of the particle [J/(kg°K)] diameter of the particle [m] effective diffusion coefficient [mZ/sec] activation energy associated with the devolatiliza- tion reaction liberating species i [cal/g-mole] activation energy [cal/g-mole] mean activation energy associated with the Gaussian distribution function [cal/g-mole] distribution function for the activation energies ratio defined in Eq. (4) convective heat transfer coefficient between the sweep gas and the particle surface [J/(m2°K sec)] heat of reaction for each devolatilization reaction liberating Species i [J/kg] Arrhenius rate constant associated with the reaction liberating Species i [sec‘l] Arrhenius frequency factor associated with the rate constant ki [sec‘ ] Arrhenius frequency factor for any reaction i [sec' ] k" iR iNR Arrhenius rate constant for the decomposition of the reactive volatile Species mass transfer coefficient [sec-l] [m/sec] thermal conductivity of gas [J/(m sec °K)] thermal conductivity of the solid phase of the particle [J/(m sec °K)] typical molecular weight for a reactive Species [q/g-mole] typical molecular weight for a nonreactive species [g/g-mole] effective mass flux of a reactive volatile Species i in the particle [kg/(mzsec)] effective mass flux of a nonreactive volatile species i in the particle total effective mass flux Species i in the particle total effective mass flux volatile Species i in the total effective mass flux Species at the surface of total effective mass flux [kg/(mzsec)] of all reactive volatile [kg/(mzseC)] of all nonreactive particle [kg/(mzsec)] of all reactive volatile the particle [kg/(mzsec)] of all nonreactive volatile Species at the surface of the particle [kg/(m sec)] number of particles partial pressure of reactive volatiles in the gas phase of the particle [psia] partial pressure of nonreactive volatiles in the gas phase of the particle [psia] total pressure of volatiles in the gas phase of the particle [psia] heat flux at the surface of the particle [J/(mzsec)] intrinsic rate of reaction for a devolatilization reaction i [k g/ (m3sec)] radius of the particle [m] xi w(t) W(t) O. (l - Y) ideal gas constant intrinsic rate of devolatilization [kg/(sec m3 of solid phase)] intrinsic rate of coking [kg/(sec m3 of gas phase)] surface area of a Spherical particle [m2] time [sec] uniform temperature of the particle [°K] initial temperature of the particle [298°K] temperature of the sweep gas [°K] velocity of volatile Species in the gas phase of the particle as they are convected out of the particle [m/sec] weight of reactive volatiles collected at time t [kg] weight of reactive and nonreactive volatiles collected at time t [kg] Greek Letters thermal diffusivity of solid [mZ/Sec] fraction of C* that represents the total reactive volatiles content fraction of C* that represents the total nonreactive volatiles content void Space of the particle [m3 of gas phase in the particle 1 m3 of particle Darcy permeability [m2] characteristic viscosity of the nonreactive volatile Species in the gas phase of the particle [kg/m sec] 3.1416.... density of the solid phase of the particle xii Da Nu Pr Re So Sh standard deviation of the Gaussian distribution of activation en integrating fact ergies or [cal/g-mole] Dimensionless Groups R k" Damkéhler number Nusselt number Prandtl number Cp' Reynolds number where Um is the superficial velocity of the sweep gas Schmidt number p Sherwood number Thiele Modulus V u *0 kg Pu u AB k mi! xiii D 2R 1 . INTRODUCTION In view of dwindling petroleum reserves and the ubiquitous nature of Oil shale deposits, oil from Oil shale is being considered once again as a source of chemical feedstock. Although more than one trillion barrels of oil exist in the eastern oil shale formations in Michigan (Katz and Goddard, 1964) and another 700 billion barrels of oil are present in the western formations of Colorado, Wyoming, and Utah (Lewis and Rothman, 1975), it is unclear whether significant fractions of this oil can be recovered. Combustion and hot gas retorting are two methods currently being investigated as a means of recovering oil from oil shale. The proposed methods include both above ground and in Situ processing (Jones, 1976; Lewis and Rothman, 1976). Mathematical Simulations of these large scale com— mercial Operations are presently being develOped by Braun and Chin (1977) and Crowl and Piccirelli (1979) to study alternative Operating strategies. These models necessarily require an understanding of the rate processes which occur in individual Oil shale particles under pyrolysis conditions. Devolatilization of shales has been modeled using simple kinetic schemes but, unfortunately, this strategy fails to predict the production of volatile products under a wide range of pyrolysis conditions. A satisfactory model has not been developed which incorporates the kinetics of devolatilization, a mechanism for intraparticle mass trans- port of volatile products, and the kinetics for intraparticle product degradation. Therefore, a prototype model will be developed in this study which describes all these rate processes during the pyrolysis of Oil Shale particles. The complex kinetics characteristic of oil shale devolatiliza— tion will be interpreted using the distribution of activation energies theory developed previously for coal pyrolysis (see, eSp., Anthony and Howard, 1976). The kinetic model will be combined with material balances to describe the pyrolysis of large particles. For these larger particles, diffusion and bulk flow will affect the production of volatile products. 1.1. Physical Description of Oil Shale Oil Shale is a low porosity material which yields gaseous volatile products upon heating. Volatiles, other than water, which condense at temperatures above -78°C are classified as "oil." The noncondensable products are typically CO, CO CH4, and C2H6' The nominal void 2! H2! fraction of eastern oil shale particles is 0.05 (Crowl and Piccirelli, 1979); the western oil shales typically have higher porosities on the order of 10% (Tisot and Murphy, 1965). The solid phase of the particle is composed of inorganic and organic material. The solid organic phase contains kerogen, a complex polymeric material. The structure of kerogen is highly napthenic with aromatic, nitrOgen, and sulphur heterocylic ring systems distributed throughout the kerogen molecule (Jones and Dickert, 1965). When an oil shale particle is heated, the volatile products are produced by devolatilization reactions which occur in the solid organic phase at temperatures between 475°K and 775°K. The inorganic solid phase of western oil shales contains dolomite, CaMg(CO3)2; calcite, CaCO3; and magnesium carbonate, MgCO3. At temperatures of 840°K or higher, Jukkola, et a1. (1953) determined that these inorganic carbonates decomposed by the following endothermic reactions: + CaMg(CO3)2 + MgO + CO2 + CaCO3 CaCO3 + Ca0 + CO2 Mg’CO3 + MgO + C02. Allred's (1967) eXperiments also Show that the inorganic carbonates do not decompose if the temperature is below 880°K. These endothermic reactions act as heat sinks in the retorting process and their presence could potentially limit the efficiency of a particular recovery strategy. However, for the temperature range of the pyrolysis of kerogen (i.e., 473°K to 773°K) the decomposition reactions of the carbonates will not occur, so the heat sinks associated with these reactions should not limit the extent of pyrolysis. 1.2. Description of Oil Shale Pyrolysis AS an oil shale particle is heated the bonds of the kerogen molecule begin to break. Parts of the molecule are released as volatile Species into the gas phase of the particle. These bond breaking processes are the devolatili- zation reactions which occur in the solid organic phase uniformly throughout the particle. Reactions producing certain classes of volatile species may be identified with a characteristic activation energy associated with the strength of the chemical bond being broken by thermal energy. Moreover, the volatiles released into the gas phase may be classified as either reactive or nonreactive, depending on their tendency to coke. The reactive volatiles, associated with "oil" forma- tion, are tarry, higher molecular weight Species possibly existing as free radicals (Russell, et a1., 1979). These reactive volatiles are subject to gas phase coking and cracking reactions and may deposit as inert char inside the particle instead of being released as "oil." The nonreactive volatiles are typically lower molecular weight species such as C0, C02, CH4, C2H6’ hydrogen, and water. Although these Species are not subject to gas phase decomposition for T < 775°K, they do react at higher temperatures. The activa- tion energies associated with the gas phase decomposition of the nonreactive species are of the order of 80-104 kcal/g- mole (Benson, 1968; Murphy, et al., 1958; Palmer, 1963); whereas, the activation energies for the gas phase decomposi- tion of the reactive volatiles are of the order of 30-60 kcal/g-mole (Benson, 1968; Murphy, et al., 1958; Palmer, 1963). In typical cracking and coking reactions, additional nonreactive volatiles are produced when reactive volatiles decompose (Hirt and Palmer, 1963; Palmer and Cross, 1966; Hudson and Heicklen, 1968; Heicklen, et al., 1969); however, in this study it is assumed that the volatile species pro- duced by the decomposition reactions are negligible. Upon release from the solid, the volatiles are transported to the particle surface by convection and diffusion, but the reactive volatiles are subject to decom- position reactions. Therefore, the yield is limited by internal as well as external resistances to mass transfer. These limitations cause the buildup of reactive volatiles in the gas phase of the particle which increases the rate Of gas phase decomposition. For large particles the rate of volatiles production may be affected by mass transport processes. Anthony (1974) points out that very little eXperimental work has been con- ducted to find this limiting particle size for coal pyrolysis. It is noteworthy that no influence on the rate of coal devolatilization has been observed for particle diameters up to 200 microns. Anthony (1974) states that the onset of mass transport limitations arises somewhere in the particle size range of 200-2000 microns. It will be assumed that the limiting particle Size for oil shale devolatilization will also fall in this range. Shih and Sohn (1978) have demonstrated that for large particles internal temperature gradients within the particle affect the retOrting process; however, if the particle sizes are sufficiently small, the particle can be considered spatially isothermal. This assumption can be justified by estimating the time needed for a solid Sphere of radius R and thermal diffusivity a(EkS/pSCp) to approach a uniform temperature when subjected to a step change at its boundary. According to Bird, et a1. (1960), this occurs when oat/R2 a 0.4. Granoff and Nuttall (1977) and Campbell, et a1. (1977) have measured the following physical properties for Colorado oil shale: cp = 1.13 x 103 J/kg °c p8 = 2.25 x 103 kg/m3 kS = 1.25 J/m sec °C. Thus, the thermal diffusivity, a, corresponding to these prOperties is ~ 5 x 10-7 m2/sec. For particle radii ranging from hundreds of microns (i.e., 10-4 -3 m) to tenths of a centimeter (i.e., 10 m), the real times corresponding to the dimensionless time 0.4 are approximately 10—2 sec and 1 sec, respectively. The time scale for devolatilization in most experiments is on the order of 105 sec; therefore, regardless of whether the environment of the particle is isothermal or nonisothermal, the particle will behave as though it is Spatially isothermal. This will be true for particle diameters up to approximately one centimeter. As the solid organic phase of the Oil shale pyrolyzes, the void Space within the particle will increase as volatiles leave the solid. For non—coking lignite coal, the particles maintain their porous structure throught out pyrolysis and leave behind a very porous structure of ash and char (see, Russell, et al., 1979). However, coking bituminous coal particles swell and soften during pyrolysis and volatiles escape as bubbles. Although the physical structure of oil shale is comparable to that of coking, bituminous coals, the prototype model developed hereinafter assumes that the shale particle maintains its geometric configuration and integrity throughout pyrolysis. 1.3. Kinetic Models for Pyrolysis Hubbard and Robinson (1950) conducted isothermal experiments on samples of oil shale in the form of cylinders one centimeter in diameter and six centimeters in length. They determined the amount of kerogen reacted by measuring the amounts of oil, gas, and bitumen produced as a function of time and interpreted their data with a simple mechanism involving two consecutive first order reactions, viz., k1. k2. kerOgen + bitumen + Oil + gas. Their experiments showed that a carbonaceous residue was formed as a pyrolytic product. However, no measurements were made on the amounts of residue formed, and the residue does not appear in the reaction mechanism. Allred (1967) was able to calculate the amount of residue directly from the data of Hubbard and Robinson (1950). He also conducted nonisothermal experiments in which the weight of volatiles collected were measured as a function of time and found that the production of oil and gas occurred over temperatures ranging from 477°K to 744°K. No further appreciable weight loss was observed until 880°K where inorganic carbonates began to decompose giving off additional volatile Species. Allred tried to explain ..is data and that of Hubbard and Robinson using the kinetic mechanism prOposed by Hubbard and Robinson. However, his measurements of the overall rate constant showed three distinct regions in which k vs. l/T was linear. This observation led him to the following mechanism k kerogen +1 gas + bitumen + carbon residue k bitumen +2 oill + gas At temperatures below 744°K, the decomposition of bitumen was observed to be rate limiting with an Arrhenius activation energy of 40 kcal/g-mole. Above 744°K the vaporization of oily liquids became rate limiting with an activation energy of 15 kcal/g-mole. Jones and Dickert (1965) state that the activation energies of 45 kcal/g-mole reported for kerOgen pyrolysis is to be expected for reactions involving carbon- carbon bond breaking. Braun and Rothman (1975) observed that the experi- ments of Hubbard and Robinson were not true isothermal experiments because of the large particle size. They con- sidered a kinetic scheme similar to that of Allred's but introduced a heating up period before retorting started. No devolatilization occurred until a certain thermal induction time was reached. Braun and Rothman Observed that the duration of this initial stage decreases linearly with increasing temperature. The reaction scheme employed by Braun and Rothman is k kerogen +1 fl bitumen + f3 gas + carbon k bitumen +2 f2 oil + f4 gas + carbon where the stoichiometric coefficients f f f f and ll 2' 3’ 4, the initial concentration of kerogen are functions of the pyrolysis temperature. The rate constants kl and k2 are _ -lO,650 cal/ -mole -1 k1 — 14.4 exp ( R T ) , sec G k = 2.025 x 1010 -421443 cal/g-mole) e1. exp ( R T , sec G 10 Although these results are in general agreement with Allred's eXperimentS in which the decomposition of bitumen is the rate limiting step at temperatures below 760°K, above 760°K kerogen decomposition is rate determining. Apparently, the decomposition of kerogen involves the breaking of relatively weak chemical bonds, while the decomposition of bitumen involves breaking much stronger bonds. The data of Hubbard and Robinson (1950) and Cummins and Robinson (1972) Show that the weight of volatiles col- lected as a function of time asymptotes to a value character- istic of a particular temperature. If the temperature is increased this asymptote also increases. The upper limit of this asymptote occurs around 773°K which Allred measured as the upper limit for oil and gas production. Braun and Rothman show that the limiting value of the fraction of kerogen converted to oil is 0.62 and occurs at temperatures over 748°K. Because the molecular structure of oil shale is not well defined, Johnson, et a1. (1975) concluded that it was unrealistic to describe the mechanism for each chemical Species undergoing pyrolysis. By lumping groups of similar products, they were able to arrive at a more complicated kinetic scheme, viz., kl k2 k5 k7 kerogen + rubberoid + bitumen + semicoke + coke i * k4 .¥k3 y k6 k8 k9 gas Oil oil heavy oil + oil. 3‘ klO gas 11 With this assumption they were able to interpret a wide range of isothermal and nonisothermal experiments conducted with various particle size shales. Their model contains no tem- perature dependent coefficients but contains twenty adjustable kinetic parameters. Such agreement between their model and experimental data is expected given the number of adjustable parameters. Furthermore, such a complex kinetic scheme may not be convenient for describing large scale retorts. Campbell, et al. (1978) assumed that kinetic experi- ments could be interpreted as an average of many processes occurring simultaneously. They postulated that devolatili- zation of shale could be modeled as a single first order reaction with an effective activation energy of 52 kcal/g- mole. As with the two previous simple kinetic models, the initial concentration of reactive kerogen must be interpreted as a function of temperature to describe experiments at different temperatures. In another study Campbell, et al. (1977) Observed that the heating rate in nonisothermal experiments also affected the final yield of volatiles. They postulated that the heating rate affected some intra- particle mechanism degrading the oil once it had been liberated from the solid phase. The bond breakages that occur during pyrolysis include carbon-carbon, carbon-oxygen, and carbon-hydrogen bonds. The bond strengths of these three groups depend on neighboring functional groups but range from 80 kcal/g-mole to 104 kcal/g-mole. Anthony and Howard (1976) point out a 12 study by Juntgen and Van Heek (1970) where it was demonstrated theoretically that a set of parallel first order reactions can be represented by a single first order expression having a lower activation energy. This phenomena can explain the lower activation energies, 40-60 kcal/g—mole, Observed in kinetic experiments. This might also explain the low value of 19 kcal/g-mole obtained by Cummins and Robinson (1972) in their experimental study. Many of the qualitative features of oil shale kinetics have also been observed by researchers investigating coal devolatilization. Anthony and Howard (1976) give an extensive review of this work. Pitt (1962) adapted Vand's (1943) treatment of a large number of simultaneous parallel rate processes with differing activation energies to interpret coal devolatili- zation experiments. In a refinement of Pitt's theory, Anthony and Howard (1975, 1976a, 1976b) described the devolatilization of the solid organic phase of coal as an infinite set of simultaneous, parallel, irreversible, first order reactions. The intrinsic rate of reaction for one of the reactions can be described by ri = -kiCiB' These reactions are all first order in the mass concentration of unreacted material and have rate constants of the Arrhenius form 1“ J-.J v-n _"‘ ki = floi exp(§gf). The Arrhenius frequency factors, koi’ for each reaction i can be estimated from transition state theory to have a value of 1013 sec—l, provided the entropy of activation is close to zero (Anthony and Howard, 1976; Benson, 1968). The activation energies for all the reactions represent a continuous distribution characterized by some distribution function. Anthony and Howard assumed a Gaussian distribution characterized by the parameters E0, the mean activation energy, and f, the variance of the distribution. If the frequency factor is the same for all the reactions, then only three, intrinsic, adjustable, kinetic parameters are introduced: E 0, and k0. If a o’ more complex kinetic scheme is used, as in the Johnson, et al., model, two more adjustable parameters are required for each new reaction proposed. By using the distribution of activation energies approach and assuming a Gaussian distribution, only one additional parameter, c, is intro- duced, although the number of reactions in the kinetic scheme is infinite. The distribution of activation energies model retains the complexity of many simultaneous first order reactions yet keeps the number of adjustable para- meters to a minimum. The theory is able to predict the weight loss at low temperatures since the reactions which produce volatiles at 14 these temperatures have characteristically low activation energies. As the temperature is increased, the additional yield of volatiles is again predicted. Volatiles released at higher temperatures are produced by reactions which have much higher activation energies than the mean, so these reactions occur extremely Slow at low temperatures. Hill, et a1. (1967) observed that the density of the oil produced in pyrolysis eXperiments increased with tempera- ture due to increasing amounts of higher molecular weight Species produced at higher temperatures. This experimental evidence provides additional motivation to develop a theory of Oil shale pyrolysis similar to the one already developed for coal. Basically, the idea is that higher molecular weight volatiles are bound more tightly to the solid and are liberated by reactions having high activation energies; whereas, the lower molecular weight volatiles are produced by reactions which have low activation energies. Because of the complex nature of kerogen, a continuous distribution of reactions characterized by different activation energies occurs. The value for the initial concentration of volatiles in the distribution of activation energies theory represents the ultimate volatile content of the oil shale. This is in contrast to the initial concentrations in the Simpler models which represent apparent volatile concentrations for a particular experiment. Anthony and Howard (1976a) 15 successfully used this approach to predict the weight of volatiles collected as a function of time for non-coking lignite coals. 1.4. Objectives of This Research The kinetic model to be developed in this study uses the distribution of activation energies theory develOped by Anthony and Howard and others. By comparing the kinetic model to experimental data for powdered Oil shale samples where no mass transfer limitations exist, the parameters of the model can be determined for a particular shale. The kinetic model can then be used to predict weight loss as a function of time in cases where intraparticle and inter- particle mass transfer limitations are important. Simple kinetic approaches are not able to predict experimental data for a wide range of eXperimental conditions without the use of temperature dependent initial concentra- tions and stoichiometric coefficients. Anthony (1974) was able to overcome these problems in his coal devolatilization study by using the distribution of activation energies theory. Because the pyrolysis behavior of oil shale is similar to that of coal, it is natural to assume that this approach will also apply to oil shale. Anthony (1974) did not extend his theory to situa- tions where mass transfer limitations are important. Although Russell, et a1. (1979) developed a model for coal pyrolysis which included intraparticle mass transfer as well 16 as gas phase decomposition, they did not employ the distri- bution of activation energy kinetics as developed herein- after. In the approach develOped here, the heating rate enters the problem eXplicitly through the boundary condition for the temperature. This is in contrast to the approach of Campbell, et a1. (1977) where the heating rate appears implicitly in the kinetic rate parameters. It is anticipated that the appearance of the heating rate in the boundary condition will allow for more accurate model predictions over a wider range of experimental conditions. 2. A MATHEMATICAL MODEL FOR THE PYROLYSIS OF OIL SHALE The porous oil shale particle will be treated as an effective homogeneous mixture of organic and inorganic compounds. Each devolatilization reaction which liberates a volatile Species i occurs homogeneously throughout the particle. Because the porosity of the particle is very small and the rate of devolatilization is slow compared to the intraparticle mass transport, the pseudo-steady state assumption will be applied to the gas phase material balances. The mechanism for intraparticle mass transfer includes both diffusion and convection. The diffusive flux will be described by Fick's law of diffusion characterized by an effective diffusion coefficient, and the convective flux will be described by Darcy's law. A resistance to mass transfer will also occur between the gas surrounding the particle and the surface of the particle. This resistance occurs across an imaginary film surrounding the particle and is characterized by a convective mass transfer coeffi- cient. For very small particles, intraparticle mass tranSport occurs extremely fast, so the concentration of 17 18 Species i will be the same everywhere in the particle. Therefore, the only resistance to mass transfer will occur across the imaginary film surrounding the particle. A finite resistance to heat transfer will also occur across this imaginary film. This resistance to characterized by a convective heat transfer coefficient, which will be estimated from correlations found in Bennett and Myers (1962) for a sphere suspended in a gas stream. An estimate of the mass transfer coefficient will be made using the Colburn analogy (i.e., jm = jh). The temperature of the particle is approximately uniform (see, Section 1.2) for particle diameters less than one centimeter. However, the endothermic heats of reaction for the devolatilization reactions will act to lower the temperature of the particle relative to the surroundings. Hence, the temperature of the particle will be determined by the amount of heat transferred between the particle and its surroundings and the heat removed by the pyrolysis reactions. 2.1. Devolatilization Kinetics The intrinsic rate of reaction for Species 1 under- going devolatilization can be described by r. = ki C. (1) where 19 -E. _ 1 ki — kO exp(§—T). (2) G C. is the mass concentration of species i bound to the solid 1B (i.e., mass of Species i per unit volume of solid phase). Eqs. (1) and (2) can be combined to give .3. _ ___1 o ri — kO eXp(RGT)g(Ei,t) CiB (3) where c. g(Ei,t) s 7:31 . (4) C13 The ratio in Eq. (4) will be determined later by a material balance over the solid phase of the particle. The constant CIB represents the initial mass con- centration of species i which decomposes by a reaction having an activation energy between Bi and Bi + AEi. The probability that a reaction, which produces Species i, has a certain activation energy is represented by f(Ei)AEi. Thus, 0 _ CiB — C* f(Ei) AEi (5) where C* is the total initial mass concentration of volatile species that decompose. Because ZCIB = C*, the density 1 function must satisfy I f(Ei) AEi = 1. (6) l 20 The total intrinsic rate of reaction is obtained by summing the contributions from all species i. Eqs. (3) and (5) imply that -Ei ‘2‘ - * .____. a r. - k C E exp(R T)g(Ei,t)f(Ei)AEi. (7) J. 1 G If the number of different activation energies in the reaction set is very large, then the sum appearing in Eq. (7) can be calculated as an integral with the result that z: r = R = k c* Jbexp’i)g(E t)f(E)dE (8) . i ' o a ‘R T ' l G where the limits of the integral in Eq. (8) define the interval over which f(E) is defined. Eq. (8) is the total intrinsic rate of devolatiliza- tion for all volatile species. To describe the intrinsic rate of devolatilization for the reactive volatile Species, Eq. (8) is multiplied by y, the fraction of C* which yields reactive volatiles. Similarly, the intrinsic rate of reaction for the nonreactive species is given by Eq. (8) multiplied by (l-Y). The density function f(E) for the activation energies is unknown. In this study it is assumed that EO>>0 and that o<> MB’ R R and the total pressure is independent of CR’ Therefore, RTC p~—§——§3 (21) _ MNR This simplification uncouples the material balances for C R and CNR' The boundary conditions for Eq. (15) are dC R - dE_ r=o _ 0 (22a) anr=Rp = kg (CR‘r=Rp ’ CROO) ' (22b) CR is the mass concentration of reactive species far from the particle and k9 is a convective mass transfer coeffi- cient. For situations where a sweep gas flows past the R;*0. The first boundary condition specifies symmetry for particle, C the concentration profile at the center of the particle. The second boundary condition requires continuity of mass flux at the surface of the particle. The mass transfer coefficient will be taken to be a constant independent of concentration and temperature. kg will be calculated by using the Colburn analogy (see, p. 646 in Bird, et al., 1960) 26 j = Nu Re'lpil/3 = j = Sh Re'15c‘l/3. H In An estimate of Nu can be made using the following correla- tion (see, p. 386 in Bennett and Myers, 1962) Nu = 2 + 0.6 (Pr)l/3(Re)l/2. This correlation was developed for a sphere suspended in a gas stream. The sweep gas will be assumed to be nitrOgen. As an approximation, the diffusivity of methane in nitrogen is used to calculate Sh and So. The transport prOperties of the nonreactive volatile Species appearing in Eqs. (19) and (21) will also be assumed to be those of methane. For powdered shale particles, Nu== 2 will be employed. 2.3.2. Material Balance for the Nonreactive Volatile Species in the Gas Phase of the Particle A mass balance for a nonreactive volatile species i in the interstices of the particle is 8(C a) iNR - . at + e(v EiNR) - (1 - {)(l e)ri (23) where EiNR is the intraparticle mass flux of nonreactive volatiles. Once again, if Eq. (23) is summed over all nonreactive species, then V'BNR = il—g—El (1 - v) R . (24) provided pseudo-steady state applies. In spherical coordi- nates, Eq. (24) is 27 1., r 2 _ - _ I (r nNR) - ———:—- (l x) R. (25) (LID; H The intraparticle mass flux of nonreactive Species includes both diffusion and convection dC __ NR r1NR De dr + CNRVr' (26) v is given by Eqs. (19) and (21) as before. r The boundary conditions for Eq. (25) are dC NR _ dr Ir=o - 0 (27a) nNer=Rp _ kg (CNer=Rp - CNRw)' (27b) The mass transfer coefficient for the nonreactive species is estimated in the same way as for the reactive Species. CNRmis approximately zero, if a sweep gas flows past the particle. 2.3.3. A Material Balance for the Volatile Species Bound to the Solid Phase A material balance for species i bound to the solid phase of the particle is d(CiBVp(l - s)) -E = - __£ - r dt koexp(RGT)CiBVp(l -). (28) If the volume of the particle and the void fraction are constant, then Eq. (28) is 28 dCiB 'Ei dt = ‘ko eXp(RGT) Cis‘ (29) Thus, c. -E. —£§ = eXp[-k ft exp(-—$)dt] (30) o o R T ciB 0 G which defines g(Ei,t) appearing in Eq. (4). 2.3.4. Material Balance for the Weight of Volatiles Collected The weight of volatiles collected as a function of time from an arbitrary sample of oil shale particles can be described by performing an overall material balance around all the particles. This yields dW _ 5E ‘ Sp€(nRs + nNRS)’ (31) The surface area of a particle is S (i.e., 4nRé), p is the number of particles, n and nNRS are the mass fluxes of RS reactive and nonreactive Species at the surface of the particle, and W is the total weight of volatiles collected. 2.3.5. Energy Balance The energy balance for the particle is d(vppstT) dt = qS - R]AHIVp(l - e) (32) where Vp is the volume of the particle, Cp is the heat capacity of the solid phase, ps is the density of the solid 29 phase, T is the uniform temperature of the particle, and [AH{ is the endothermic heat of reaction. The heat flux at the surface of the spherical particle is q and equals q = h(Tg - T) (33) where h is a convective heat transfer coefficient and Tg is the temperature in the gas far from the particle. The heat transfer coefficient in Eq. (33) is assumed to be independent of concentration and temperature. For pseudo-steady state, Eqs. (32) and (33) imply that h(T ~ T)S = RIAHIV (1 - e), (34) g p which yields RIAHXR (1 - s) T = T - P 9 3h (35) where Rp is the radius of the particle. Eq. (35) shows that the temperature of the particle will differ from that of the gas depending upon Rp and h. The magnitude of this difference can be investigated by calculating the second term on the right hand side of Eq. (35) for large and small particles. The heat of pyrolysis for Colorado Oil shale is 335 kj/kg (see, Campbell, et al., 1977), and the porosity is 10%. With Nu = 2 and kga§=0.03 J/(m sec °K), h=65 J/(m2 sec °K) for 0 = 8 x 10‘4m; and, h = 4 J/(m sec °K) for 0 = 0.0127 m. 30 These particle diameters were used in experiments to R be investigated later, so the values of giH| are 0.7 3 3 (m sec °K)/kg and 17 (m sec cI)/kg. For isothermal experiments, R is a maximum at t = 0 and decreases as the reaction progresses. In nonisothermal experiments, R increases to a maximum and then decreases to zero. It will be assumed as a first approximation that the R|AH|R (l-E) P 3h to Tg' If this is the case, then the temperatures of the particle and the sweep gas will be approximately the same. values of in Eq. (35) are small compared This assumption will be verified a posteriori when the values of E0, 0, kc, and C* are determined and values for R are calculated. For nonisothermal experiments the heating rate of the particle enters the problem through T9. The heating rate does not affect the intrinsic kinetic rate parameters as postulated by Campbell, et a1. (1977) but enters the rate expression through the eXplicit temperature dependence of k.. 1 3. SPECIAL CASES TO BE STUDIED 3.1. Chemical Kinetics Controlling Regime--Isothermal The equations develOped in the previous section can be solved for the situation where intraparticle mass tranSport occurs much faster than chemical kinetics. These cases are typical of small particles, and the concentration of volatile Species in the gas phase of the particle will be the same everywhere in the particle. 3.1.1. Reactive Species Integrating Eq. (15) and applying the boundary con- ditions (22a) and (22b) yields 2 (l ' F) YRRP: 12.939: Rp kg(CR-CRm) = s 3 - 3 . (36) Solving Eq. (36) for CR gives R _p (1 - e) c = 3 Y __-E__- R + kgCRm (37) R k + R ° 9 32 k" Substituting Eq. (37) into Eq. (22b) gives the flux of reactive species out of the particle 31 32 Rp -—-———Y(l - E) R — R k"CR°° = p nRs 3 + Da (38) where Da is the Dakahler number Da = —E—— . (39) The Dakahler number is the ratio of the rate of gas phase decomposition to the rate of interphase mass transfer. 3.1.2. Nonreactive Species Integrating Eq. (25) and applying the boundary conditions (27a) and (27b), results in 2 (l-€)(l-‘{)RR3 - = p Rp kg(CNR CNRm) 3s (40) Thus, (1 - a) (1 - Y) NR ' 38 k RRp + CNRm (41) Substituting Eq. (41) into Eq. (27b) gives the flux of nonreactive Species out of the particle 1 [11 - a) (l - Y)RRP; NRS =| 3e J' n (42) Note that the flux does not depend on the external resistance to mass transfer. This is a consequence of the pseudo-steady state assumption and the absence of coking reactions. As the external resistance to mass transfer becomes very small (i.e., kg+m, Da+0), CR-rCRoo and CNRICNRw' As the external resistance increases (i.e., kaeo, Da+m), CR+0 since R goes to zero at long times. This indicates that all reactive volatiles will eventually go to coke if not transported out of the particle. If k"=0, then Eqs. (37) and (38) reduce to Eqs. (41) and (42), respectively. 3.1.3. Weight of Volatiles Collected Substituting Eqs. (38) and (42) into Eq. (31) gives P dw i ‘1’ Vppe k "CR“: 3? ‘ Vp‘”l ‘ ”’)(i‘¥‘5a) + (l ‘ I) R + Da ° (43’ L 3— (l + 3“) Eq. (43) describes the rate of volatiles collected as a function of time and is applicable in both isothermal and nonisothermal situations. The only difference will be the form of g(E,t) as given by Eqs. (4) and (30). Substituting Eqs. (30), (9), and (4) into Eq. (8) gives for the isothermal case k c* +w O -E ~15: R = f exp(——-) exp[-k t exp(—-)) x /2Ro -m RGT O RGT '(E - EO)2 exp[ 2 ] dB. (44) 20 When CR3? 0, Eqs. (44) and (43) yield 34 gig L._ 5 Da + (1 - y) x v/Z’T 7 '(l + T) f exp(-——) exp[-k t exp(-——)] exp[ ]dE.(45) _w RGT o RGT 202 Integration of Eq. (45) over time gives W(t) = v p(1 - e)c*[———Y——— + (1 - «M x p Da (1 + 3—) i 1 +m -E -(E - Ho)2 1 - I__ f eXp[-k t exp(-——)] exp[ ]dE (46) o R T 2 /2w O -w G 20 Eq. (46) describes the total weight of volatiles collected as a function of time for isothermal situations where chemical kinetics controls the production of volatiles. The weight of "oil" collected is given by Eq. (46) without the term (1 - y) and will be represented by the variable w(t). 3.2. Chemical Kinetics Controlling Regime-~Nonisothermal Many experiments are designed so the temperature far from the particle increases linearly with time, i.e., T = TO + bt. The parameter b is a constant heating rate. Substituting Eqs. (30), (9), and (4) into Eq. (8) gives 35 k0c* +m E t _ -(E-EO)2 R = f expf—-¥:<———] exp(-k f exp[-—fi———-dt) exp[————————jdE. R m VERO -m G(Toot) O RG(igbti 20(2 (47) IfR———:§—— = x, then C(To+bt) -E kOC* +co _E -k E RG(TO+bt) ex (x) -(E-EO)2 R = f “MFR—67:7] exp[ R b x 412—31") exp[“z—ldfi- /2F0 -m G o G -E x 20 RGTO (48) Recoqnizing that TO + bt = T -.:32 R T _,_ 2 koc* +m -E -koE G exp(x) (L E ) R = f exp(§f¥)eXp[ R b .f 2 dx] exp[ 2 ] dE. /§;J -m G G -E x 20 RGTO (49) Integrating over x gives kOC* +w _E -(E-EO)2 R = _ £00 exp(E—f) exp[ 2 ] X /2fl0 G 20 - exp(:§*) -koE -E -E RGT EXP R—'b_‘ [E 1(R—Tf)‘ ELI-(W) "E + G G G o (-—) R T - G -E eXP‘R—GT" -E ] dB. (50) ( ) RGTO _ where Ei(') is the exponential integral function and T0 is the initial temperature of the particle. Substitution of Eq. (50) into Eq. (43) with CRm=O yields 36 dw __ L. Y _ V 3 5E; - Vpp(l -) [(l+2§)+ (1 g)] b (5].) 3 where R is defined by Eq. (50). Eq. (51) can be integrated numerically to obtain the weight of volatiles collected as a function of temperature for nonisothermal situations where chemical kinetics controls the rate of volatiles production. The weight of "oil" collected (Ew(t)) is Eq. (51) without the term (1 - v). 3.3. Diffusional Limitations for Pyrolysis--Isothermal For very short times or for very low pyrolysis tem- peratures, the concentration of reactive and nonreactive volatiles in the interstices of the particle will be rela- tively small. If the mass concentration of nonreactive volatiles is small, the total pressure given by Eq. (21) will be small. If this pressure is low enough so that Cer and CNRV in Eqs. (l8) and (26) are small relative to dC r chR De-EE and De d , then the dominant mechanism for intra- r r particle mass transport of volatiles will be diffusion. 3.3.1. Reactive Species For intraparticle diffusional limitations, Eqs. (15) and (18) can be combined to give dCR e dr (1 - a) (I'D )=-——€——YR-R. (52) ~13. (LID: H H the general solution of Eq. (52) is 37 C ‘? C ‘7 _ - v c =--]-'-Coshx/}5-r+-—25inh /-;—-r+-(—];———EZ--‘-—R. (53) r e r R e Applying the boundary conditions Eqs. (22a) and (22b) results in :{k R 2(1 — C)R _ _ S;E - _ R acc ’3:;r C = )0“; P QR“ Sinh / DE + “/(l _ E) R R D [Sinh¢-¢cosh¢]-k R sinh¢ r *———;:g e g p ‘ ~J (54) where .111 $ = v De Rp. (55) Substitution of Eq. (54) into Eq. (22b) gives the flux of reactive volatile species out of the particle. - q (ILLEZ_EL R - k" CRm)R — 53 (56) RS $2 Da - [l - ¢ coth @] - J For ¢+0 (i.e., De+w), Eq. (56) reduces to Eq. (38) for the case of no intraparticle mass transfer limitations. By applying L'Hospital's rule, it is simple to show that lim o2 ¢+0 (1 - ¢coth¢) =-3 3.3.2. Nonreactive Species For intraparticle diffusional limitations, Eqs. (25) and (26) can be combined to give 38 (r21) NR) = (l " ”fl " S) R. (57) 8‘)“ (D D.) H :l 2 r The solution of Eq. (54) subject to the boundary conditions, Eqs. (27a) and (27b),is F 2 2 (l - ‘rHRp -r ) (l - Y)R NR ‘ 60 3k e g _ J (58) Substitution of Eq. (58) into Eq. (27b) gives the flux of nonreactive volatile species - . - a R (l ()(l ) P nNRs = 35 R' (59) Eq. (58) shows that the concentration profile is affected by diffusional limitations; however, Eq. (59) is identical to Eq. (42) owing to the pseudo-steady state assumption. As k"+0 (i.e., ¢+0 and Da+0), Eq. (56) and Eq. (59) are the same except for the parameter y. Thus, n=n +n RS + (l - €)RpR/3€. NRS 3.3.3. Weight of Volatiles Collected Eqs. (56) and (59) can be substituted into Eq. (31) with the result that . 3V pek"C m 3% ‘ VP? (1 - E) 3‘2 + (l - Y) R + -p 2R Da-(-——$———-) Da—(——¥g———.) _ l-¢coth¢ l-¢coth¢ (60) 39 Eq. (60) describes the rate of volatiles collected as a function of time for both isothermal and nonisothermal situations where diffusional mass transfer limitations exist. For isothermal cases the total intrinsic rate of devolatilization is given by Eq. (44). Substitution of Eq. (44) into Eq. (60) and assuming CRufO yields ‘ 1 V l - €)k C* V w - pp( _ O 3! + (l __ Y) X “it (‘2'?) v Dal—(J1..-) E l-ocotho L - r+w -E -E '(E-EO)2 exp(———) exp[-k t exp(~—~)] exp[--—-~—]dE. RGT o RGT 202 (61) Integrating Eq. (61) over time gives _ - * 3Y . W(t) - Vpp(l - c)C 2 + (1 - f) X _ Da-(l-¢coth$) l +m -E -(E"Eo)2 l - I__ f exp[-kot exp(E—T)] exp( 2 )dE /2n 0 -m G 20 L (62) Eq. (62) describes the total weight of volatiles collected as a function of time for isothermal situations where diffusional limitations affect the production of volatiles. Note that the only difference in this result and Eq. (46), which describes the situation of no intraparticle mass transfer limitations, is the constant multiplicative factor containing o. The weight of "oil" collected is Eq. (62) v. 92 Ct \ufi 40 without the term (1 - Y) and will be represented once again as w(t). 3.4. Convection Limitations for Pyrolysis--Isothermal For high pyrolysis temperatures, the concentration of reactive and nonreactive volatiles in the interstices of the particle will be large. Therefore, the total pressure within the shale particle relative to the surroundings may be high enough to cause the convective terms in Eqs. (18) and (26) to dominate the diffusive terms. Thus, intra- particle mass transport of volatiles will be controlled by convective processes. 3.4.1. Nonreactive Species For intraparticle convection limitations, Eqs. (25) and (26) can be combined to give (1 - a) (T) A 1 ‘ “R. (63) Substituting Eqs. (19) and (21) into Eq. (63) and inte— grating the result yields (' .- 2 u (l - Y)(l - €)R A NR 1 3K€RGT - 1 where Al and A2 are arbitrary constants of integration. (64) Applying the boundary conditions to Eq. (64) gives 38k + cNRoo + g . L. -J 2 _ { (l y)(l €)RPR NR 41 uMN (l - Y)(l - €)(R 2 - r2)R) % R p (65) 3K€RGT Substituting Eq. (65) into Eq. (27b) gives the flux of nonreactive species out of the particle (l - Y)(l - €)R R WNRS = 36 p ° (66) Eq. (66) is the same as Eqs. (42) and (59), because coking is ignored and pseudo-steady state is employed. Substitution of Eq. (66) into Eqs. (19) and (21) gives the velocity profile within the shale particle (1 - y)(l - €)r R (1 - ()(1 - s)RER 2 Vr = 36 38k + CNR + L g 2 2 )‘% MkRU.-€)U.-y)mp-r)Ri .- (67) 3euommwm Annmav Hanuusz cam mwocmuo AUmmnmEV\o mN.H mx .wmmza cflaom mo >ua>wuozccoo Hmsumne mmfluquOHm uuommcmue Annmav Hamuusz com wwocmuo ms\@x mmm *0 .ucmucoo waflumao> mumEHuH: Amhmav cmenuom can csmum No.0 >.Hao ou Umuum>coo cmmouox mo cofluomum Anwmav .Hm um .Hawnmemo mx\hx mmm :4 .:0Huommu mo pom: Ammmav Haaouwooflm can Hzouu we macoa x vmm.a Amamcm cumummmv y .>uHHHnmmEHmm Amomav snaps: can acmfle oH.o o .suwmouom Ashmao .Hm um .Hamnmsmo ms\mx moH x mm.m ma .mmanm oflHOm mo suflmcma Annmav Hamuusz can mmocmuu Ava mxv\hmoa x MH.H mo .ummn owwflowmm mofluummoum Havamwnm umumfimumm oozoumuwm 05Hm> HmcflEoz I! II .-'..l|. ‘ll I‘l‘ .iv "li‘ a- .I.yulll. . I .‘I‘I .0 II .51- .-| I’I. .III .mamnm HHO cuvummz “Om mmfluuomoum uuommcmue can HmOflmxnmtl.N mange 47 m oom\E mo. 1 mo. x .ucmwOfiwwmoo ummmcmuu mmmz 0mm NE\n ow : v S .ucmflOflwmwoo umwmcmuu umwm Aonmav pamflwumuumm 0mm\mE OHIOH x H mQ .ucmonwmmoo :OHmDMMflc m>fluommwm Amhmav Haaoufiooam cam Hzouo Axoum-e\h mmo.o mx .ommcm pflHOm mo >uw>wuoswcoo Hmeumse mofluuomoum uuommcmue : : mE\@x sum *0 .ucoucoo mafiumao> mumeuH: 83: 33.383 new 386 3. s .30 3 nouuogoo ammouox «0 coflomum :8: .E S .2838 3 \E mmm :< .8383 mo 83: = : me mH-oH x va.H y .s»a~wnmmsumm : : mo.o w .wufimouom : : mE\mx MOM x vv.m mo .mmmzm Ufiaom mo muflmcmo Ammmav Haamuflooflm ocm Hzouu Axolaxv\n moH x mflo.H mo .ummn Dawwommm mmfiuuomoum amOflm>nm mocvuowom msam> Hmcfleoz umuwfimumm .-..- It'll; (.Y(- i) U. I: Q I‘l'- 'l'lli -.‘l| A- II -.!I.|I| luuiiy) I I I Q. I -llbllo! (1"-‘ll' .'l' ~10 I.-t'.l It :l’u|tvl, 'Jl:|.". ‘A'It’ill‘l: .moamzm Hfio cumummm MOM mofluuomoum uuommcmnfi Ucm Hmofim>£mnu.m magma 4. THE EFFECT OF KINETIC PARAMETERS ON "OIL" YIELD For very small particles, the weight of "oil" pro— duced will be limited by chemical kinetics. The effect of the kinetic parameters on the weight of "oil" collected for small particles can be calculated using Eq. (46) with the term containing (1 - Y) deleted. In what follows, the fraction of the total possible oil yield will be calculated for different values of k0, Eo’ and o. The base case for these calculations will be k0 = 6.95 x 1013sec-l, E0 = 55, 333 cal/g-mole, and o = 2000 cal/g-mole. For this set of calculations the Damkohler number will be set equal to zero. The effect of the heating rate on the weight of oil collected for small particles can be calculated using Eqs. (50) and (51). The integration of E over the interval (-m, +m) was approximated by integrating over the finite interval [BO-20, Eo+201. Even with this truncation, 95.45% of the reactions are still included in the reaction set. The error introduced by the truncation will be small since the probability that a reaction has an activation energy in the truncated region is very small. The integral was calculated 48 49 using Simpson's rule (see, p. 79 of Carnahan, et al., 1969). It was determined that thirty applications of Simpson's rule were necessary to obtain an error of less than 10-6 in the evaluation of the integral. 4.1. The Effect of 0 on "Oil" Yield Figure 2 is a plot of the dimensionless oil weight collected versus time at 648°K for three values of O. The sigma values of 2 and 10 kcal/g-mole represent narrow and broad distributions. At 648°K, the curve representing 0 = 10 kcal/g-mole displays the greatest "oil" yield during the time represented. This behavior is a consequence of the higher probability that a reaction has an activation energy much lower and much higher than E0. For these reactions, the rate constants will be relatively large despite the low temperature, and a significant production of volatiles will occur. For large and small values of O, the ultimate "oil" yield is the same. However, 0 influences the time required for pyrolysis significantly. Note that for 0 = 2 kcal/g-mole the probability that a reaction has an activation energy less than or greater than E0 is much smaller than the case with 0 = 10 kcal/g-mole. Thus, with ZRGT60 sec. This occurs because ZRGT< HMO :0 mumu mcwummn mnu mo uommwm one .m mudmflm AwDHung. umapcmmmZMH oom owe 8a com 3m r . c . _ . (u . nu . \\\ .\. \ \ \ omm\xommo.one \\ \\mw \\ a / \\\\ Z omm\xomeao.oun \.\ \\.\ x \\ ommEOmmooeun \ it \\ \\\\ \\ omm\xoaeoo.un \ . \ .AH a. \ 1 \ \\\ omm\xomooo.oun \\ \\\ xx 1. knlllu \ \\\\x In: \ \\ (\AH.\ "n unnrxAv \. .L 8H3 031331103 1I0 30 3H010A 58 qualitatively and nearly quantitatively follows observed experimental behavior even with Da = 0 lends support for this approach. It is anticipated that the model will also give satisfactory predictions for cases where the heating rate is nonlinear. S. PARAMETER ESTIMATES USING DATA FOR WESTERN OIL SHALE Campbell, et a1. (1978) conducted several isothermal and nonisothermal experiments on powdered samples of Colorado oil shale. The diameter of the particles used in the experi- ments was 800 microns, and the oil content of the shale was 22 gal/ton as measured by the Fisher Assay. In one isother- mal experiment, the weight of oil collected as a function of time at 648°K was measured. From this experiment it was , o, and C*(1 - 5) possible to determine the parameters k0, EO for the distribution of activation energies model. For this experiment Da/3=3 x 10-8, which implies that very little degradation of the liberated "oil" is expected to occur. This is a consequence of the small particle size and the low temperature of pyrolysis. 5.1. Estimates of Parameters for the Distribution of Activation Energy Model Values for E0, 0, k and (l - c)C* were obtained 0! by curvefitting w(t) given by Eq. (46) to the experimental data using a direct search Optimization scheme (see, Hooke and Jeeves, 1961). Details of the procedure can be found 59 60 in Appendix A. The results of the Optimization gave the following values for the four parameters: 13 -1 k0 = 6.95 x 10 sec E0 = 55,333 cal/g-mole = 1740 cal/g-mole yC*(l - 5)==223 kg/m3 The results of the estimating procedure can be seen in Figure 7. Because the first nine data points were used in the Optimization scheme, the model accurately predicts the data for the first 40 ks. However, the long time response of the model deviates from the data. In a second isothermal experiment, Campbell, et al. (1978) measured the total weight loss as a function of time at 673°K. For this experiment, Da/3 = 8 x 10-8, indicating once again that "oil" degradation will be negligible. The values of y and (1 - s)C* were obtained by fitting Eq. (46) for the total weight collected to the experimental data E using the values for k and 0 obtained previously. 0’ 0’ Details of this procedure can be found in Appendix C. The values for y and (l - s)C* are: y'=O.65 (1 - €)C* = 344 kg/(m3 of particle) The total initial mass concentration of volatiles as indicated by (l - s)C* represents 15 wt% of the initial weight of the particle. This corresponds favorably with the 15 wt% of organic material as measured by Jukkola, et a1. (1953) for 28 gal/ton Colorado oil shale. It also compares 61 ms\mx mmm u 10 . Hoeo> .meoe-m\emo oven u o .H-ommmnoexmm.e n ox .maosum\nmo mmm.mm n om .mHmOOE HOOHO umunm UGO cenusnfluumnw mcu How xomvo um cofiumeumm Hmumamuma mo muasmmm .5 musmflm Amx. mzup om" om— cv— ON— co“ am am O¢ ow o _ . _ r _ r! _ p _ . _ . p . _ . . . mu 0 0 0 w z 0 HOOOE MOOMO umuflm/ [mu .7 I'I'IIIIIIII Amend. .Hm um .Hnmnasmo we mean x \ \\ \\\ HOOOE Genusnfluumflo 1 031331103 1M 110 SS31NDISN3N10 62 well with the initial concentration of kerogen (i.e., 377.3 kg/m3) for Michigan oil shale (see, Crowl and Piccirelli, 1979). The experiments of Granoff and Nuttall (1977) also showed that the total weight loss at 793°K for 22 gal/ton Colorado oil shale was 15 wt% of the initial weight of the shale. Braun and Rothman (1975) determined that the fraction of kerOgen converted to oil was 0.62. The value for y is consistent with this value. The value for the mean activa— tion energy E0 is typical of what has been observed pre- viously for oil shale pyrolysis (see, Section 1.3). The variance of the distribution is much smaller than 9,380 cal/g-mole obtained by Anthony (1974) for lignite coal. This smaller value of 0 apparently indicates that the bonds which make up the kerOgen molecule may be more uniform than those of a coal molecule. The results of the last estimating procedure are shown in Figure 8. The total weight of volatiles collected is made dimensionless with the initial weight of the parti- cles. It is apparent from Figure 8 that the distribution model does not predict a sharp plateau at long times. The absence of a distinct plateau could be eXplained by the shape of the distribution function. For instance, the Gaussian distribution of activation energies requires that the activation energies for all the reactions producing volatile Species occur in the interval [EC-23, EO+20]. The experiments of Allred (1967) have shown that the reactions producing gaseous nonreactive Species nominally occur at 63 6 ms\ x eem u 16 . H..O .me.o u > .meos-m\3mo seen u e .Huommmaoaxmm.e n ox .mnos-m\nmo mmm.mm u on .Aw I HV«U can > :flmuno ou Momhm um conumfinumm umuOEmqu mo muasmmm .m musmnm Amzu mzuh cu m w c N o . _ . _ . _ . _ . _ . F0 I Y Amend. .Hm um .Hamnusmo no memo .. 031331103 1M SS31N013N3H10 64 lower temperatures than those reactions producing reactive Species (i.e., "oil"). Given the low value of O, the lower limit of the distribution function [EC-20] might not be expected to include the activation energies for those reactions producing nonreactive species. Thus, the distri- bution function might better be described by a bimodal dis- tribution. One peak of this bimodal distribution would be characterized by 01 and E01 and would be centered arOund the mean activation energy of the reactions producing non— reactive species. The second peak of this bimodal distri- bution would be characterized by E02 and 02 and would be centered around the mean activation energy of the reactions producing reactive Species. A bimodal distribution would improve the model prediction in Figure 8 if E01 and E02 differed significantly and if ol< .meos-m\nmo eeea u o .H-ommmnoaxmm.e u ox .maosnm\amo mmm.mm u om .omm\xommo.o mo oumu mcnumm: a ROM mHmOOE umcuo umnnw can GOHDDQHMHmHO one .m musmflm .mDHOOMO. manpmmumzm~ owe ow? emu own Dom — p r0 \IIII‘) IP b 0 00m x . u xx .6 H- neon m e x \\\ I OHOEIO\HOO www.mm n om \\ \\ . HOOOE HOOHO umuHh/V\ .7 \\ I I III\ 1 \ 9 HOOOE coflusnnuumflJV\ Ammmav .Hm um .HHOOQEOU mo mama \ -9 (€N3J 031331103 110 30 3Nfl10A 68 E0 and the heating rate. The effect of varying a will be negligible since such a small value of o was found by the Optimization scheme. However, in general, the effect of o in the distribution model is not negligible as discussed in Section 4.1. The effect of E0 on the "oil" yield is shown in Figure 4. By reducing Eo by 5% and 10%, the response is altered significantly. Even an error of 1% in the deter- mination of EO would affect the response. Such an error is possible since the Optimization procedure identified several local minima very close to each other. The least of all the local minima was chosen as the global minimum. If the optimization process had continued, possibly one of the other local minima could have been identified as the global minimum. This procedure would have required excessive computer time, and the global minimum that was arrived at was considered satisfactory. Figure 10 shows the experimental data along with the predictions given by the distribution model for four different heating rates. The data can be represented quite well if the heating rate of the experiment were 0.0165°K/sec. However, this would have required that the heating rate be lower than reported by a factor of two. For this nonisothermal experiment the maximum value of R occurs at 723°K and equals 0.16 kg/m3sec. The corre- RPIAHI (1 - €)R 3h Sponding value of in Eq. (35) is 0.10°K. 69 ms\mx man ”10 u aceo» .H-ommmneflxmm.e u ox .mnosnm\nmo mmm.mm u on .muou manummn may mo pomwmm one .oa muamwm AmsHunuui washcmmmEMH om. om. omm oem nu .OHOEum\Hmo owha u 6 com _ b . \ \ \. \ \ \ I Z UQW\MOMMO.O H D \ //V\ \ \ \\ x. \ Tb. 663310385 u n \ x \ \ . _\ \\ omm\mommeo.o u n \ 9 \ x \A I \ omm\xoneoe.e u a .\ \\\ T \ E: A... em .3338 no 33 8 \ 4. I \ . \ .I Ol“! (9N3) 031331103 110 30 3N010A 70 Therefore, for the duration of the experiment the tempera- tures of the particles and the sweep gas will be approxi- mately the same. 5.4. Prediction of Isothermal Pyrolysis for Large Particles Granoff and Nuttall (1977) conducted an isothermal eXperiment at 703°K with a 12.7 mm diameter sphere and measured the total weight loss as a function of time. The oil shale that was used in the experiment contained the same oil content as that used by Campbell, et a1. (1978). Therefore, the parameters that were determined from the data of Campbell, et al. (1978) will be used in this set of calculations. For this large particle size, the mechanisms for intraparticle mass transfer are expected to be diffusion for the short time data and convection for the long time data. However, Da=8 x 10-5, which indicates that degradation of the liberated "oil" will be negligible. Eq. (62) is used to calculate the total weight of volatiles collected in the presence of diffusion limitations. Figure 11 is a comparison of the data with the distribution model for the total weight loss using three values of ¢. Up to 300 seconds, the cases where ¢=20 and ¢=40 represent the data quite well. The values of the effective diffusion coefficients corresponding to ¢=20 and -11 -ll ¢=4O are 4 x 10 mz/sec and 10 mZ/sec, respectively. The case where ¢=0 represents no intraparticle mass transfer 71 .OHOE(m\HmO ownau O .Hno om.~ no.u v0.0.fim.~ on r _ _ _ hhmHv Hamuusz Cam wmocmuw m0 mama ms\mx 44m u 13 - H016 O 0 mm onxmm.e u x .mnosum\nmo mmm.mm u m ma .somos um mcoHumueEflH Hmeonsuuec can; Hence cenusneuumno .HH ousmfim “0:0 mzuh .~ mg.“ N0.0 00.0 0v.0 m~.0 00.0 b _ b _ _ p 0 000' osc~o szo'o 0313311103 1M SSB1NOISNBNIO SLO'O 001'0 SZI'O 72 limitations. Apparently, at short times there are diffu- sional limitations present as hypothesized in Section 3.3. To calculate the total weight of volatiles collected versus time for convection limitations, Eq. (81) was inte- grated. A fourth order Runge—Kutta technique (see, p 363 of Carnahan, et al., 1969) was used with a step size of ten seconds. The program that was used in making this calcula- tion is presented in Appendix F. Figure 12 is a comparison of the data with the convection case and the case where 6:0. Both cases lag behind the data after 360 seconds but begin to approach the data at longer times. This lag in the weight loss at short times for both the case where ¢=O and the convection case gives further support for a bimodal distribution. As was pointed out in Section 4.2, if reactions with lower activation energies are included in the reaction set, the rate of volatiles production increases. An increase in the production of volatiles would increase the initial time response of the model. Depending on the magnitude of this increase, the convection case may be expected to repre- sent the data more accurately. It is noteworthy that pre- vious attempts to interpret this data with other models (cf., Granoff and Nuttall, 1977; Shih and Sohn, 1978) were also relatively unsuccessful and gave similar predictions as those given by the distribution model. For this isothermal experiment the maximum value of R occurs at t=0 and equals 0.35 kg/mBSec. The corre- RPIAHI(1 - €)R 3h Sponding value of in Eq. (35) is 5.50°K. 73 ms\mx eem n 10 n 3016 O .mHos-m\Hmo came u e . omm oaxmm.e u x .maosum\emo mmm.mm n H: ma Amv_. m:10h om.~ po.~ 4e.1 18.1 om.1 m1.1 mm.o mo.o m¢.o m~.o so. _ _ _ _ _ L. _ _ b _ \ x. \ _\ .\\ _\\.\ \\. \\ \\x .\ COHUUWN/COO \ \ \ \ \¢\ \ \ \ \ N \ \ \ \ \ \ AbhmHv Hamuusz UGO mchOHO mo mama 0 SLO'O 030-0 920-0 ooo~o 031331103 1M 8831N013N3010 001'0 SZI'O o m .Momon um mcoflumuflafla 20nuom>coo can: HOOOE :oHuanHumno .NH whomnm 74 O This term is approximately equal to one after 23 minutes into the experiment. Because of the large particle size, the temperature of the particle will be cooler relative to the surroundings. This small difference is not eXpected to affect the diffusion, convection, and kinetic responses significantly. 6. MODEL PREDICTIONS FOR EASTERN OIL SHALE The kinetics of devolatilization of eastern oil shales has been modeled as a single, second-order, irre- versible reaction (Crowl and Piccerelli, 1979). The rate expression for this reaction is R = 2.4868 x 1013 exp ( cp ~30,337 K) D2 kg (84) T K m3sec where 9K is the concentration of unreacted kerOgen. The initial concentration of kerogen was measured to be 377.3 kg/m3 or 15.26 wt % of the initial weight of the shale. Crowl and Piccirelli (1979) used this second order expression in an intraparticle mass transport model to predict the pyrolytic behavior of a semi-infinite slab of oil shale. In order to make comparisons between the intrapar- ticle mass transfer models using second order kinetics versus using the distribution of activation energy kinetics, it was necessary to obtain values of E0, 0, and k0 typical of eastern oil shale. The determination of these para— meters requires weight loss versus time data for the devola- tilization of eastern oil shale. Since no experimental data are presently available in the literature for eastern oil 75 76 shales, hypothetical isothermal oil loss data at 775°K for powdered oil shale were generated using the second-order kinetic model. Eq. (46) for the weight of oil collected was curvefitted to the initial time data as before with yC*(l - a) = 377.3 kg/m3. The results of this parameter determination are k0 = 5.88 x 1013 m3/kg sec E0 = 53,333 cal/g-mole o = 1563 cal/z—mole With these parameters Eqs. (46), (62), and (81), can be used to calculate the weight of oil collected as a function of time at different temperatures. These equations are made dimensionless with the maximum possible weight of oil which can be produced (i.e., VppyC*(l - 6)). To obtain predictions for the second order model, Eqs. (46), (62), and (81) will differ only in the form of R. For the second order model i‘ “O 2 -E V ’K R = k exp(-——) (85) cp o R T -E o I G 51 + k0 exp(§—§)pk ; G 1 where 00K = 377.3 kg/m3 and k0 and E are the kinetic para- meters appearing in Eq. (84). The isothermal, dimensionless oil weight collected versus time was calculated for each model for the chemical kinetics controlling regime, diffusional limitations, and convection limitations using a hypothetical 5 cm in diameter Sphere of eastern oil shale. The calculations were made 77 at temperatures of 648°K, 703°K, and 775°K. Internal tem- perature gradients were ignored since the main purpose of this study was to compare the differences between second order and distribution of activation energy kinetics. The values for Da at 648°K, 703°K, and 775°K corresponding to a particle size of 5 cm are 4 x 10-4, 3 x 10-3, and 2 x 10-2, respectively. 6.1. Chemical Kinetics Controlling Regime For the chemical kinetics controlling regime the dimensionless weight of oil collected under isothermal conditions according to the second order model is - l l w(t) = -———-——— [l - ]. (86) (l + Da) -E 0 §—- (1 + k0 eXp(§;T)DK t Figure 13 shows the reSponse of the distribution model and the second order model. The close agreement between the models is expected at 775°K since this was the temperature at which the parameter determination was made. It is note- worthy that a single first order expression (i.e., o = 0 kcal/g-mole) will not represent "data" from the second order model. However, the distribution model represents the "data" quite well because of the flexibility introduced by the parameter 0. The model predictions at the other temperatures are also very close. 78 .mEnmmH manaaouncoo mOAHOOAx HmoHEmzo mzu H00 mHmOOE umwuo Ocoomm 0cm coflusnnuumno .ma mudmflm Amazouum. 020p owe can own owe om. own can ouw ems ow" so lllll dmnda.anuQ.mwmwmww. \ \\ I). Momvw um AOOOE coausnnuumflo .\ \ \ xomon um H0008 HOOHO Ocoomm/JY .\ Memos um HOOOE :oflusnnuumna xomvw um H0008 HOOHO ocoomm /////JV \ omnn um HOOOE :ofiasnnuumno \ Do 0 2'0 031331103 1M 110 SS31N013N3H10 79 6.2. Diffusional Limitations for Pyrolysis With diffusional limitations, the dimensionless weight of oil collected for the second order model under isothermal conditions is 0(t) = 31 2 [1 - l _E O 1 (87) ¢ (1 + k exp(-—-)p t) Da‘ (m) 0 RGT k Figure 14 shows the response of the distribution model and the second order model for ¢=40. Close agreement is again seen at 775°K. The weight loss occurs over a much longer time scale but all the curves will eventually attain the final value of 3Y2 . Once again, close agreement _A_) l-Ocoth6 is seen at all temperatures. Da-( 6.3. Convection Limitations for Pyrolysis For convection limitations, the isothermal, dimen- sionless weight of oil collected is % - l d 3:) = 235 f eXp [[2 (l - r 2)} X p O . 3 K _ i B g (AR[l - r ] - (AR) 3 _2 _ _2 8 g g r dr (88) _lAR[l - r ] + (AR) where A and B are defined by Eqs. (82) and (83). Eqs. (81) and (88) were integrated using a fourth order Runge-Kutta technique. It was determined that a step size of 10 seconds 80 .oeue .mconumnesna HmconSMMHO MOM mHmOOE HOOHO OOOOOm cam :oHuOQHHumHQ .va muzvflm \ 70'0 20’0 00 90'0 .wx0 mzuh 00.0 h0.~ v0.— «0.0 mm." m_._ «0.0 00.0 0v.0 m~.0 00.0 _ _ _ u . . p u .1... nu XMmmwleImwOMEIwmvmw OGOOOmN. \ M6060 um H0008 OOAOOOHHumHO \ \. \ .\ MoMOh #6 H0008 MOOHO OOOOOm \.\. \/ \\\\\\ /...mon \\\\\\ um HOOOE coflusnnuumwo \ \ Momhn um HOOOE uwcuo Ocoommx/JY 1... somhn 06 Hence coHusaeuumeQ 0‘ 80'0 01'0 031331103 1M 110 SS31NOISN3NIO 81 was necessary to avoid truncation error. Figure 15 shows the response of the distribution model and the second order model. Once again close agreement is seen at 775°K. The weight loss occurs over the same time scale as the diffusion case. The final value for the dimensionless weight collected are 0.83 for the second order model and 0.82 for the distri- bution model showing that a yield loss has occurred. This yield loss was caused by the buildup of reactive species in the gas phase of the particle because of intraparticle con- vection limitations. This buildup increased the rate of coking, and consequently, a yield loss occurred, character- ized by the asymptote at 0.83. It appears that the curves for the other two temperatures will eventually approach this asymptote. Close agreement at all three temperatures is seen again. 82 .mcoflumuAEnH :0nuom>coo How HOOOE HOOHO Ocoomm 0cm COHuOQHHumHQ .mH musmfim Amxv mzmh om.~ he.~ em." 18.1 om.1 m1.1 «6.9 mo.o oe.e m~.o oo.o _ b _ 110 . _ . b u nu momvm um HOOOE HOOHO ocoomm V.1.11.1.|.l.l.1.11 111111111 \ || “| 00 031331103 1M 110 SS31NOISN3N10 \ somee pm demos codesnfluumfla \ mu 7.. _\ \_ .\ \ 0 \ .Ilo \ .7 Momon um HOOOE HOUHO Ocoomm 1111 M1 \|\\\\ 0 1111.1 9 ..1 1.1. zones 66 36668 863683060636 Roman pm 36608 menusnnuumna// 1111/ 11111111111111 momhn um H0005 MOOHO Ocoomm 0'1 7. CONCLUSIONS AND RECOMMENDATIONS FOR FURTHER RESEARCH The low value of o that was determined for western oil shale indicated that the bonds of the kerogen molecule are very uniform. The predictions given by the second order and distribution models for eastern oil shale were very similar. This implies that perhaps these simpler kinetic expressions are adequate in describing the kinetics of devolatilization. However, the failure of the first order model and the success of the distribution model in non- isothermal situations indicates that the distribution model is required. The need for the more complicated distribution model to describe the kinetics of devolatilization remains unknown until additional data from isothermal experiments conducted with powered oil shale particles over a wide range of temperatures are available. The determination of the parameters k , E , O, y, o and C* requires a set of data from an isothermal experiment measuring both the weight of reactive and nonreactive volatiles collected versus time for powdered oil shale 83 84 particles. Ideally, the temperature of this experiment should be 775°K, the data from such an experiment would not be expected to show an apparent weight loss plateau. In the isothermal experiment which was used to determine the para- meters in this study, only the weight of oil collected as a function of time at 648°K was measured. The data from this experiment show that the weight of oil collected as a function of time approaches an asymptote characteristic of 648°K. Steps were taken in the parameter determination scheme so that this apparent weight loss plateau would not be imposed on the model. However, it is anticipated that the parameters could be determined more accurately if data from isothermal experiments at the highest temperature of pyrolysis were used. The experimental observation that nonreactive and reactive volatiles are produced at different temperatures led to the suggestion that a bimodal distribution might be more apprOpriate than a Gaussian in characterizing the activation energies of the devolatilization reactions. This suggestion was given further support since model predictions for the total weight loss lagged behind the data at short times (see, Section 5.4). This lag suggested that reactions with lower activation energies should be included in the reaction set in order to Speed up the short time response (see, Section 4.2). If a bimodal distribution is indeed the case, the parameters E01 and G which characterize one 1 peak of the bimodel distribution could be determined by the 85 weight of nonreactive volatiles collected versus time in an isothermal experiment. The parameters E02 and 02 which characterize the second mode could be determined by the weight of reactive Species (i.e., "oil") collected versus time. If the distribution is actually bimodal, then the parameters E0, 0, y, and C* which were estimated in this thesis characterized only the peak representing the activa— tion energies of the reactions producing reactive Species. The value of C*(1 - s), the ultimate volatile con- tent of the oil shale, was determined to be 15.26 wt % of the initial weight of the particle and correSponded to values from other experimental investigations. ED was determined to be 55,333 cal/g-mole and is typical of activation energies represented in other studies. The fraction of kerogen con- verted to oil, y, was determined to be 0.65. This value was nearly identical with the value of 0.62 obtained by Braun and Rothman (1975). The model predictions for nonisothermal experiments where the temperature of the sample increased linearly with time were consistent with experimental data. These model predictions were made using no further adjustable parameters and the heating rate entered the calculation explicitly through the boundary condition for the temperature. Although favorable predictions were obtained using linear heating rates, it is anticipated that the distribution model will also predict situations where nonlinear heating rates are used. 86 For experiments with 12.7 mm particles it was not clear if the production of volatiles was affected by intra- particle convection limitation (see, Figure 12). However, Figure 11 suggests that the production of volatiles at short times may be affected by diffusion with an effective diffu- ll m2/sec. The calculations comparing sion coefficient of 10- the convection response for second order and distribution of activation energy kinetics for a 5 cm in diameter sphere (see, Section 6.3) show that the production of volatiles is limited by intraparticle convection. The limitations cause a buildup of reactive volatiles in the particle and conse- quently an increase in the rate of gas phase decomposition. Further eXperiments need to be conducted to define the limiting particle size where the rate of production of volatile species is no longer controlled by chemical kinetics but by intraparticle mass transfer. APPENDICES APPENDIX A PROCEDURE FOR ESTIMATING THE DEVOLATILIZATION PARAMETERS APPENDIX A PROCEDURE FOR ESTIMATING THE DEVOLATILIZATION PARAMETERS The parameters in the distribution of activation energies kinetic model can be determined using the weight of oil collected versus time isothermal data at 648°K. The weight of oil collected given by Eq. (46) can be made dimensionless by dividing by the initial weight of the particle. Thus, ‘ * — oo - V_rl£1 = (C ‘1 E) [l - 7:17— f+ 8Xp[-kot exp (FET’I ppcs ps V20 0 -m G -(E-EO)2 x exp[—TIdE]. (21.1) 20 The estimating procedure seeks to find a minimum in the sum of the squares of the differences between the data and the model. For this strategy, the objective function is n - " l +00 -E J(A k ,E ,O) = Z [y.-A [l- f exp[-k t. exp(‘——)] X o o 0 i=1 1 o /20 O _m o 1 RGT 87 88 2 -(E-EO)2 eXp[-—-—,3--—]dE] (A.2) 20“ : 7C*(l - a) . o where AO _ O and yi are the data p01nts at 648 K. s In order to begin the Optimization it was necessary , o, and to choose starting values for the parameters k0, EO A0. The starting points were determined using the Michigan State University IMSL subroutine ZSRCH. ZSRCH generates a specified number of points in an n-dimensional space for use as starting points in nonlinear Optimization routines. For each set of starting values, a direct search Optimization scheme was used to find the best values for the parameters that minimized the objective function Eq. (A.2) (see, Hooke and Jeeves, 1961; Beveridge and Schechter, 1970; and Walsh, 1975). The Objective function is highly nonlinear and has many local minima. Because many sets of starting values will converge to different local minima, several sets of starting values were tried. It was found that twenty starting sets of parameters were sufficient to identify the various local minima. The Hooke and Jeeves method and ZSRCH require that constraints be placed on the parameters. The following constraints were imposed: l3 l4 -1 1.0 x 10 :koil.0 x 10 sec 40,000:Eo:60,000 cal/g-mole 0:0510,000 cal/g-mole 05A030.15 , 89 The global minimum of the objective function was used to define the parameter estimates. For the data of Campbell, et a1. (1978), the devolatilization parameters were found to be k0 = 6.95 x 1013 sec“1 E0 = 55,333 cal/g-mole o = 1740 cal/g—mole A0 = 0.09906 , The step sizes and the number of times the step sizes were halved were chosen to economize the computing time without sacrificing accuracy. Several combinations of step sizes and number of halvings were tried. It was determined that the starting step sizes should be Ako = 0.1 x 1013 sec"1 AEO = 2000 cal/g-mole AG = 1000 cal/g-mole AAO = 0.01. It was also determined that halving five times gave suffi- cient results. Further halving did not decrease the least squares error appreciably and only led to excessive computing costs. Campbell, et a1. (1978) show twelve data points in their experiment measuring the weight of oil collected as a function of time. The data show the weight of oil collected varying nearly linearly with time and then leveling off to 90 a plateau characteristic of 648°K. If all data points were used in the estimating routine, it was felt that the value for yC*(l - s) characteristic of 648°K would be imposed upon the model and not the true yC*(l - 6). Therefore, only the first nine data points were used since none of these were in the plateau region. Hence, in Eq. (A.2), n=9. In terms of experimental accuracy, these first nine data points are probably the most accurate. Ideally, an isothermal experiment should be conducted at 775°K where the total weight loss would be realized. The data from this type of experiment would not impose an apparent yC*(1 - s) on the model provided no degradation of the liberated oil takes place. Appendix B contains the program as well as sample outputs of the curve- fitting strategy. I" APPENDIX B COMPUTER PROGRAM FOR ESTIMATING THE DEVOLATILIZATION PARAMETERS APPENDIX B COMPUTER PROGRAM FOR ESTIMATING THE DEVOLATILIZATION PARAMETERS A*91¢Xc‘CSTAR*99X0*FUNCTION "(0).PTQ)QTINE(120)0LEAST 1H5Xfl14dh5anqdh5XJJQJHSWflJQ )9 9L GM A564 FLAG 9.31 *95XQEIQ.8.5X9E140895XoElQ-BgfiXoEl“ INSXJJQJMSXJJQJM5XJJQJMflXHU“ B 9 X :T(*0*9*1NTERMEDIATE P01NT=*.5X.E1§.8.5X9E14.8g5X.E14.8.5X9£14 9 £3LD Rod *916 A~< #vm vAC um. mvx (3h Amt-0H h... 2AA I Q¢¢Q hvvx =mtc Coo. vamx Ucnc Neon 00x0 (.0. ZAZI ¢ G Ofififi. ZavaUUva («uh9hohoh K w&OK.KOE. o 1o o 0 «a N «a c: 0 c: 0 “‘0'." 0 11111 INITIALIZATION 0F ARGUEHENTS FOR ZSRCH 91 INITIALIZATION OF TIME DATA UUU 131.9)lll94.g2388..5373oo7164.913134..16716.g207§1.g } CFC Zq thocb I—LJ-C" 0". Da-2’.O)C:.r U o 4 d) ‘ P 3 oooowxmuu OZH¢ h4NfiGb 301x 000m chddzhOuo C 2 H CH «#2 Im mwo WU h mpgcur WKFGAZD—Zu U umuonuompncm UNU 528% kW 0 gunk uJu N KUHZO m>zr0 QZQ~QJu~~hm> H h Jazz “2 me onozwcn 0000030 u: 3 ~0>mh war—0m p¢u>002hm010 <00 thdCz h N25” ¢ b HJ> mm WOdszuJO t 030:0 4 Hz HJZJ HTH>JL hm< oth<3£ at #0440 cm CNUIZC>~Q¢C~ x0 0~r~<:z Lucmoc~uchpm Ircwhmu~oz zmmphai mutt: mu>c2 >¢cp02 zxuphqm < mux4cmiz~ m— mun 0b mxumxu 0_0oz pac~4n4;20 302 m~ au>uufi 02¢ Lxccr Acu—o3~.zomod~ox.2o:o 0. p L man (Am (in: z_?_ #20 0:1: —.z< 24% 1:.\ «\N Qua O. C by- 0&1- o4 t A O- d? a )5 CC 0 O H ~1’ Q P— .-1 i. K4 uJ CL' -- 0 a D x IL LL \1 O l .AA 0 In 00:14.; A 1 xx 0 o o A Cx’ u Ecuador. " .4 O 'v-QLLv—n] v Q. 1 ‘ILu '9‘ 0 3' o A 00‘ O; + ‘5 FOX .XQ A 3 L Swamp u w W Q U .-4 .J) g V ' Ad vztuxt.‘ x - (N; O O. 0 yo fi '— nv O l: wa€g O 9. Par- L1H OH!— 0 m:mago AiLZGQG§dHU mocfiowud cm ca .auxamcao mzduh can 33m ZDMCON"...DW L_:‘~h~:pu Auimquonfiiq ..ouaaqio ~;_ud...o¢dC/.ohua1.ow vhmmaozrwu 23m 4.~zmqu¢«u m 00.97“..— .N c: oun>3m .wzauh Km>w rsm .qzaqa.m;oqt.mzaaa. H:aq..rqog.»o4n.rchahoumuoo Aa.. a<-..:a h v' .v" 3.1 p4 .‘ ’ 0:1. u '1'..IL... <2 I! I \ I 1“ A o J HQ 5', 05-.J *.l< ¢ P'J‘ «1". .‘LJ'! k 3 " l." xa~.0w qvv O .."(JJ-hd: 0 codqnm 1\'5’5uc ”5014? QC Fur. H U¢LLLLXP + “F0 "40-4 108 J) 51"? .-l .V 5/(VOIn-CIZMA~? .- h— 0 ”2.6 JUC‘v.-fi-h- <1 d")\.’""H/‘ 09... I‘- gu\\\U00“i U -‘q.vr‘[n (V)“/C TIMFQHT D‘ P‘ L OthO L "C (444cmd‘r. \xw «rad . . 2‘3“} I I I4\ U? QQ'OEQAC‘ 0 I L v~UdV~nvxwhwflafififish~d 13 mm" II II II II H H H UH .7 oi ‘IHNF‘Q’U‘mf-a‘i-f. #F-HH 100313113172c4:7+h& HNHmuqomauUH 4~~zrco H H H<acC..:«¢C.c:aqa.~:aqn.tr_p. ».o~o> C.o~m.pw . J?w rxzbuu W\1+w}~huutmh H r» cv A .CUo4.u~ O’SuJfiohw HCHAJVY n.I,_ .u.¢cc.C.:.L>ucrouum 2.,uCoUL» Hc¢..niln bmxnu aazoaru¢::C+NC C+ CLC..r:uch om»:~w 53.0.3CCOO VHIDW CC L...:.~.—...;CU ACQJCoLioq.o u:hm~0+tDU OCCH )DMDQC 5OW<.UCQ"~‘ morfloHHS m CD o.H$DwDOC Accac so «a Cthch+azonz :.m TcUSQQQHm CC Jaw“! a CD ouHSDm Amfcnon.o<.LT. I I..L.:F.I AIRLFHfl ..(n....f .IV...n:.LI\ :6... ”6.0.1....“ T Tm.r.I...p TCTCZ..§HD?R CLCILL.TIU r...n.,_.1.r.fl.II\T.H I..rL Cylr .1..- .7 ..er CCVILLTD HFILLTICI.» ,I/L.r.IPI.IL:uI:. LIPICIST UDV ..rawFflonr)CleD ..1 LI ST OF REFERENCES LI ST OF REFERENCES Allred, V. D. 1967. "Shale Oil Developments: Kinetics of Oil Shale Pyrolysis." Proceedings of the Fourth Oil Shale Symposium, Quarterly of the Colorado School of Mines, 61, 657. Anthony, D. B. 1974. "Rapid Devolatilization and Hydro- gasification of Pulverized Coal." ScD. Thesis, Department of Chemical Engineering, Mass. Instit. Techol., Cambridge. Anthony, D. B. and Howard, J. B. 1976. "Coal Devolatiliza- tion and Hydrogasification." AICHE Jl., 22, 625. Anthony, D. B., Howard, J. B., Hottel, H. C. and Meissner, H. P. 1976. "Rapid Devolatilization and Hydrogasi- fication of Bituminous Coal." Fuel, fig, 121. Anthony, D. B., Howard, J. B., Hottel, H. C., and Meissner, H. P. 1975. "Rapid Devolatilization of Pulverized Coal." Fifteenth Symposium (International) on Com- bustion, 1303, The Combustion Institute, Pittsburg, PA. Bennett, C. O. and Myers, J. C. 1962. Momentum, Heat, and Mass Transfer. New York: McGraw-Hill. Benson, S. 1968. Thermochemical Kinetics. New York: Wiley. Beveridge, G. S. and Schechter, R. S. 1970. Optimization: Theory and Practice. New York: McGraw-Hill. Bird, R. B., Stewart, W. B., and Lightfoot, E. N. 1960. Transport Phenomena. New York: Wiley. Braun, R. L. and Chin, R. C. Y. 1977. Progress Report on Computer Model for In Situ Oil Shale Retorting. Report UCRL-52292, Lawrence Livermore Laboratory, Livermore, Cal. 111 112 Campbell, J. H., Koskinas, G. H., Coburn, T. T., and Stout, N. D. 1977. Oil Shale Retorting: Part 1, The Effects of Particle Size and Heating Rate on Oil Evolution and Intraparticle Oil Degradation. Report UCRL-52256, Lawrence Livermore Laboratory, Livermore, Cal. Campbell, J. H., Koskinas, G. H., and Stout, N. D. 1978. "Kinetics of Oil Generation from Colorado Oil Shale." Fuel, 51, 372. Carnahan, B., Luther, H. A., and Wilkes, J. O. 1969. Applied Numerical Methods. New York: Wiley. Crowl, D. A. and Piccirelli, R. A. 1979. Transport Pro- cesses in Pyrolysis of Antrim Oil Shale. College of Engineering Energy Center, Wayne State University, Detroit, Mi. Cummins, J. J. and Robinson, W. E. 1972. "Thermal Degrada- tion of Green River Kerogen at 150° to 350°C: Rate of Production Formation." Bureau of Mines Report of Investigations, 7620. Granoff, B. and Nuttall, H. E. 1977. "Pyrolysis Kinetics for Oil Shale Particles." Fuel, 56, 234. Heicklen, J., Hudson, J. L., and Armi, L. 1969. "Theory of Carbon Formation in Vapor Phase Pyrolysis--II. Variable Concentration of Active Species." Carbon, 7, 365. Hill, G. R., Johnson, D. J., Miller, L., and Dougan, J. L. 1967. "Direct Production of Low Pour Point High Gravity Shale Oil." Ind. Eng. Chem. Prod. Res. and Dev., 6, 52. Hirt, T. J. and Palmer, H. B. 1963. "Kinetics of Deposi- tion of Pyrolytic Carbon Films from Methane and Carbon Suboxide." Carbon, 1, 65. Hooke, R. and Jeeves, T. A. 1961. "'Direct Search' Solu- tion of Numerical and Statistical Problems." Association for Computing Machinery Journal, g, 212. Hubbard, A. B. and Robinson, W. E. 1950. "A Thermal Decom- position Study of Oil Shale." Bureau of Mines Report of Investigations, 4744. Hudson, J. L. and Heichlen, J. 1968. "Theory of Carbon Formation in Vapor-Phase Pyrolysis--I. Constant Concentration of Active Species." Carbon, g, 113 Johnson, W. F., Walton, D. K., Keller, H. H., and Couch, E. J. 1975. "In Situ Retorting of Oil Shale Rubble: A Model of Heat Transfer and Product Formation in Oil Shale Particles." Proceedings of the 8th Oil Symposium, Quarterly of the Colorado School of Mines, 71, 237. Jones, J. B. "Paraho Oil Shale Retort." Proceedings of the Ninth Oil Shale Symposium, Quarterly of the Colorado School of Mines, 1;, 120. Jones, D. G. and Dickers, J. J. 1965. "Composition and Reactions of Oil Shale of the Green River Forma- tion." AICHE Chem. Engin. Symp. Ser., 61, 33. Jukkola, E. B., Denilauler, A. J., Jensen, H. B., Barnet, W. I., and Murphy, W. I. 1953. "Thermal Decomposi- tion Rates of Carbonates in Oil Shale." Ind. and Eng. Chem., 45, 2711. Juntgen, H. and Van Heek, K. H. 1970. "Reaktionablaufe unter Nichtisothermen Bedingungen." Fortschritte der Chemischen Forschung, I}, 601. Katz, D. L. and Goddard, J. D. 1964. A Study of the Dow Oil Recovery Process. Report 06061-1-F, The Dow Chemical Company, Midland, Mi. Lewis, A. E. and Rothman, A. J. 1976. Rubble In Situ Extraction (RISE): A PrOposed Program for Recovery of Oil from Oil Shale. Report UCRL-51768, Lawrence Livermore Laboratory, Livermore, Cal. Murphy, D. B., Palmer, H. B., and Kinney, C. R. 1958. "A Kinetic Study of the Deposition of Pyrolic Carbon Films." Industrial Carbon and Graphite, 77, Society of Chemical Industry, London. Palmer, H. B. 1963. "An Analysis of Carbon Deposition Kinetics In Isothermal Flow Systems." Carbon, 1, 55. Palmer, H. B. and Cross, W. D. 1966. "Carbon Films from Carbon Suboxide Decomposition. Inhibition by Carbon Monoxide and the Heat of Formation of C20." Carbon, 3, 475. Pitt, G. J. 1962. "The Kinetics of the Evolution of Volatile Products from Coal." Fuel, 41, 267. Russell, W. B., Saville, D. A., and Greene, M. I. 1979. "A Model for Short Residence Time HydrOpyrolysis of Single Coal Particles." AICHE Jl. 25, 65. 114 Satterfield, C. N. 1970. Mass Transfer in HeterOgeneous Catalysis. M.I.T. Press, Cambridge, Mass. Shih, Shin-Min and Sohn, S. Y. 1978. "A Mathematical Model for the Retorting of a Large Block of Oil Shale: Effect of the Internal Temperature Gradient." Fuel,‘§l. Tisot, P. R. and Murphy, W. R. 1965. "Physical Structure of Green River Oil Shale." AICHE Symp- Ser., El, 25. Vand, V. 1943. "A Theory of the Irreversible Electrial Resistance Charges of Metalic Films Evaporated in Vacuum." Proceedings of the Physical Society (London), A55, 222. Walsh, G. R. 1975. Methods of Optimization. New York: Wiley.