TrHesrs 1,122,231! acumen $1318 Uh.“ “52!? This is to certify that the dissertation entitled APPLICATION OF A CONCENTRIC VORTEX-CELL BIOMASS FURNACE TO GRAIN DRYING presented by Eliud Ng'ang'a Mwaura has been accepted towards fulfillment of the requirements for .Ph.D. degree in Agricultural Engineering [7/3%//M Major professor 1729/8 74 MSU is an Affirmative Action/Equal Opportunity Institution 0. 12771 lllllllllllljlljllllflljfljLllllllfllljlllll MSU LIBRARIES RETURNING MATERIALS: Place-in book drop to remove this checkout from your record. FINES will be charged if book is returned after the date stamped below. APPLICATION OF A CONCENTRIC VORTEX-CELL BIOMASS FURNACE TO GRAIN DRYING By Eliud Ng'ang'a Mwaura A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Agricultural Engineering 1984 ABSTRACT APPLICATION OF A CONCENTRIC VORTEX-CELL BIOMASS FURNACE T0 GRAIN DRYING By Eliud Ng'ang'a Mwaura Artificial crop drying is relatively energy intensive and uses about 60% of the energy required to produce corn, exclusive of irriga- tion. Biomass fuels, such as wood chips and corncobs, are potential alternatives to fossil fuels for crop drying. A concentric vortex-cell furnace (CVCF) with a maximum heat output of 690 kw was built and incorporated in an in-bin counterflow (IBCF) corn dryer with an hourly capacity of 3 m3 at an initial grain temperature of 10°C, a drying temperature of 71°C, and corn dried from 25% to 15.5% moisture content (wet basis). The system was tested using wood chips and corncobs. Experimental tests included: (1) fuel con- sumption and drying capacity. and (2) corn contamination with polycyclic aromatic hydrocarbons (PAHs) and heavy metals. The results were com- pared with the performance of a propane (LP) fueled system. The results demonstrate that the CVCF biomass furnace is a technically viable alternative to conventional fossil fuel burners. The drying capacity of the CVCF/IBCF system is approximately 90% of the LP fueled system. The reduction in the drying capacity is a result Eliud Ng'ang'a Mwaura of the reduction in the flow rate of the drying air mass due to the heating of the air before the fan entrance. The CVCF/IBCF system operates at about 70% efficiency in con- verting biomass fuels into energy for grain drying. The average specific energy consumption (kJ/kg of water removed) of the system is about 6,100 kJ/kg compared with 4.600 kJ/kg for the LP fueled system. There is no objectionable discoloration or objectionable odor in the corn dried with the CVCF/IBCF system. The grain contains no dangerous concentration levels of PAH or heavy metals. The operating costs (exclusive of labor and fixed costs) of the CVCF/IBCF system are 30 to 40% of the energy costs of the LP fueled system. The total break-even costs of the CVCF/IBCF system are about 20% higher than the total costs of the LP system (at 1983 prices) due to the high capital investment required to convert the LP system into a biomass system. The CVCF furnace does not have immediate economic feasibility in the USA at the 1983 fossil fuel prices and interest rates. .:-I . . _.=1:." " To Kinnthi and his colleagues 'i'f'iitm), a“. . f-. _ ACKNOWLEDGMENTS The author expresses deep appreciation to Professor F. v. Bakker-Arkema for his guidance and encouragement during the course of this research. He has provided moral and technical support since 1979 as a teacher, mentor, and friend. Sincere thanks are expressed to Dr. G. R. Van Ee (Agricultural Engineering), Dr. J. F. Steffe (Agricultural Engineering), Dr. E. N. Braselton (Animal Health Diagnostic Laboratory), Dr. E. Grulke (Chemi- cal Engineering), and to Mr. S. J. Kalchik (Kalchik Farms, Bellaire, MI) for serving as guidance committee members. Their advice, along with that of Dr. A. Dhanak (Mechanical Engineering) and Dr. A. Atriya (Mechanical Engineering), is greatly appreciated. Special gratitude is sincerely expressed to the Netherlands Government and the University of Nairobi, Kenya, for financial support and scholarship, and to Michigan Department of Commerce (Energy Administration), and Shivvers Corporation, Corydon, Iowa, for providing funds for this project. The author is indebted to Mr. and Mrs. Steven Kalchik and Mr. and Mrs. Larry Kalchik, Bellaire, MI, on whose farm this research was conducted. Their help and hospitality is sincerely acknowledged. Two fonmer students who assisted in data collection during the early days of this project deserve special mention: Juan Rodriguez and Carlos Fontana. Their assistance is appreciated. iii Sincere thanks are extended to the entire community of Agricultural Engineering Department and International Studies Center for making the author's stay at Michigan State University a happy one. Special thanks goes to my family, Eva, Njoki, Hangari, and Mumbi for their companionship, love, and encouragement. The author can be reached at the following address: Department of Agricultural Engineering University of Nairobi (Kabete Campus) Post Office Box 30197 NAIROBI2 KENYA iv TABLE OF CONTENTS Page LIST OF TABLES . viii LIST OF FIGURES xi LIST OF APPENDICES . . . . . . . . . . . . . . xiii LIST OF SYMBOLS . xiv Chapter 1. INTRODUCTION . . . . . . . . . . . . . . 1 1.1 Michigan Grain Production 3 1.2 Grain Preservation . . . . . . 4 1.3 Biomass Fuels for Grain Drying . . . . . 6 1.4 Grain Production and Storage in Kenya . 8 1.5 Objectives . . . . . . . . . 11 2. LITERATURE REVIEW . . . . . . . . . . . . 13 2.1 2.2 2.3 Biomass Furnaces for On- Farm Grain Drying. . . 13 2.1.1 Major Products of Combustion . . . . 14 2.1.2 Design Criteria for Biomass Furnaces . . 15 2.1.3 Direct- Combustion Biomass Furnaces . . . 16 2.1.4 Gasification-Combustion Furnaces . . . 24 2.1.5 Direct-combustion Versus Gasification- Combustion Systems . . . . 27 2.1.6 Emissions from Biomass Furnaces and Grain Contamination . . . . . . 30 Biomass Combustion Theory . . . 31 2.2.1 The Properties of Biomass Relevant to Combustion . . . . . 32 2.2.2 The Process of Biomass Combustion . . . 45 Grain Drying . . . . . 65 2.3.1 Importance of Grain Drying . . . 66 2.3.2 Corn Quality as Affected by Drying Methods . . . . . . . 67 2..33 Drying Systems . . . . . . . . . 74 Chapter Page 2.4 Grain Drying Theory and Simulation . . . . . 90 2.4 Thin- -Layer Drying Models . . . . 91 2.4.2 Equilibrium Moisture Content Models . . 95 2. 4. 3 Deep- bed Drying Models . . . . 96 3. FURNACE DESIGN, CONSTRUCTION, AND OPERATION . . . . 99 3.1 Design and Construction . . . . . . 99 3.2 Combustion Principle of the Concentric Vortex- Cell Furnace . . . . . . . 103 3.3 Furnace Operation . . . . 104 3.4 Concentric Vortex- Cell Furnace Design Theory. . 104 3.4.1 The Design Heat Output. . . . 105 3.4.2 Grate Design . . . . . . . . . . 105 3.5 Fuel Feed System . . . . . . . . 107 3.5.1 Furnace Air Supply System . . . . . 112 4. THEORY . . . . . . . . . . . . . . . . 123 4.1 Furnace Perfonmance Simulation . . . . . . 123 4.1.1 Preliminary Estimates . . . . 124 4.1.2 Heat Transfer from Combustion Chamber . . 128 4.2 Heat Transfer Between the Furnace and the Hood . 136 4.3 Grain Drying Simulation . . . . . 140 5. EXPERIMENTAL INVESTIGATIONS . . . . . . . . . 143 5.1 Furnace/Dryer Performance . . . . . . . . 143 5.2 Chemical Analysis . . . 144 5.2.1 Determination of Corn Contamination with H PA . 5.2.2 Detennination of Corn Contamination with Heavy Metals . . . . . . . . . . 148 6. RESULTS AND DISCUSSION . . . . . . . . . . . 149 6.1 Introduction . . . . . . . . . . . . 149 6. 2 Fuel Consumption . . . . 149 .2.1 Factors Affecting the Fuel Feed Rate . . 158 6. 2.2 Effect of Fuel Moisture on Fuel Consump- ti on . . 162 6.2.3 Effect of Excess Air on Furnace Perform- ance . . . . . . . . 166 6.3 In-Bin Counterflow Corn Drying . . . . . . 176 6.3.1 Biomass Fuel Feed Rate . . . 179 6.3.2 Experimental Specific Energy Consumption . 179 6.3.3 Drying Capacity . . . 181 6.3.4 Standardized Dryer Performance . . . . 181 vi Chapter Page 6.4 Co orn Drying Simulation . . . . . . . . . 184 6.4.1 Cycle Times . . . . . . . . . . 184 6.4.2 Drying Capacity . . . . . . . 190 6.4.3 Specific Energy Consumption . . . . . 190 6.4.4 Airflow Rate . . 194 6.4.5 Comparison of Experimental and Simulated Grain Moisture Contents . . 194 6.4.6 Comparison of Experimental and Simulated ‘ Drying Data . . . . . 196 6.5 Predicted Drying Performance Parameters . . . 198 6.5.1 Effect of Refill on Dryer Performance . . 200 6.5.2 Effect of Dryeration on System a Performance . . . 202 .5.3 Effect of Drying Air Temperature on System Performance . . . 203 6.5.4 Effect of Initial Moisture on System Performance . . 204 6.5.5 Effect of Ambient Temperature on System Perfonnance . . . 205 6.6 Economics of the CVCF/IBCF System . . . . . 205 6.6.1 Operating Costs . . . . . . . 206 6. 6. 2 Capital Budgeting Analysis . . . 209 6.7 Corn Contamination in Drying by Direct Biomass Heating . . . . 213 6.8 General Observations . . . . . . . . . 217 6.8.1 Fuel Feed System . . . . . . . . . 217 6. 8. 2 Grate Performance . . . . . . . . 218 6. 8. 3 The Combustion Chamber . . . . . . . 219 6. 8. 4 Furnace- -to-Dryer Duct . . . . . . . 219 6.8.5 Safety . . . . . . . . . . 220 6. 8. 6 Miscellaneous . . . . . . . . . . 220 7. SUMMARY . . . . . . . . . . . . . . . . 221 8. CONCLUSIONS . . . . . . . . . . . . . . 224 9. SUGGESTIONS FOR FURTHER RESEARCH . . . . . . . 226 APPENDICES . . . . . . . . . . . . . . . . . 228 REFERENCES . . . . . . . . . . . Table 1.1 1.2 2.1 2.2 2.3 2.4 2.5 5.1 6.1 6.2 6.3 6.4 LIST OF TABLES World production (1981) of main cereal grains, 1,000 metric tons . World production of soybeans and pulses (1981), 1,000 metric tons . Proximate analysis of selected biomass fuels (weight, percent, dry basis) Ultimate analysis (percent dry basis) and higher heat- ing values (MJ/kg) of selected fuels . The effect of moisture content on heat recovery and combustion efficiency of wood . . Thermal conductivity of selected wood species Numerical grades and sample grade requirements for S corn . - Excitation/emission wavelengths, retention times, and detection limits of different PAHs on the fluores— cence detector . . . . The experimental and simulated fuel consumption rates of the CVCF biomass furnace at different ambient conditions and different drying air temperatures Experimental and simulated airflow, drying energy, total energy losses and furnace/system efficiencies for tests 1- 7 given in Table 6.1 . Effect of ambient air conditions on fuel consumption, furnace and system efficiency, and total heat loss from CVCF/IBCF drying system; dr rying air temp = 80° C, drying airflow = 12. 3 m3/m -m1n . . . Furnace/IBCF system perfonmance using woodchips at different moisture contents (percent wet basis) for different drying air temperatures . . . viii Page 34 36 39 42 68 147 150 152 159 163 01 .10 O! .11 The theoretical effect of excess combustion air (percent) on the CVCF furnace perfonmance Experimentally measured higher heating value (MJ/kg) of wood chips and corncobs at different moisture contents (percent wet basis) . . . . . . Actual drying rates, energy (fuel) consumption rates and operating costs (1983 prices) for seven drying tests at different ambient conditions at the Kalchik Farms, Ballaire, MI . . . Standardized energy (fuel) consumption and operating costs (1983 prices) for different fuels and fuel feed rates for the CVCF- IBCF drying system . . A comparison of simulated and experimental CVCF- IBCF performance parameters for Test No. . . . Simulated operating conditions during Test 1 Comparison of CVCF-IBCF experimental and simulated results for seven tests under different drying conditions . . . . . . . . Simulated energy consumption, drying rate, and operat- ing costs for different dryer operation modes at average ambient conditions . Estimates and assumptions for a 12-year budgeting analysis (1983 prices) of the CVCF/IBCF system drying in Mode 1 . . Economic analysis results of different biomass fuels compared to propane (1983 prices) in drying 380 tons of corn annually from 26% to 15. 5% w. b. . . Concentration of polycyclic aromatic hydrocarbons on corn dried with biomass energy and LP- -gas Experimental results: Test No. 1 Experimental results: Test No. 2 Experimental results: Test No. 3 4 Experimental results: Test No. Page 168 177 178 182 185 186 197 201 210 212 215 232 233 234 235 Page Experimental results: Test No. 5 . . . . . . . 236 i Experimental results: Test No.6 . . . . . . . 237 7 ’_.;Experimental results: Test No. 7 . . . . . . . 238 '1': Figure 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 2.10 2.11 3.1 3.2 3.3 4.1 4.2 4.3 4.4 LIST OF FIGURES The Cell Furnace . Vortex Furnace European Sloping Grate Principle Schematic Diagram of Husk Fired Cyclone Furnace Up-draft Gasification and Close—coupled Combustion Down-draft Gasification and Close-coupled Combustion . Batch Column Dryer Schematic of a Low-temperature In—bin Drying System Schematic of a Crossflow Grain Dryer . . . . . . Schematic of a Concurrent Flow Dryer Schematic of the Internal View of an In-bin Counter— flow Drying "Shivvers System“ . . Schematic of the Concentric Vortex-cell Furnace Schematic Diagram of Magnetic and Bin Switches for Fuel Level Control in the Fuel Hopper . . The Schematic Diagram of the Eductor Schematic Flow Diagram of the MSU Concentric Vortex- cell Furnace . . . . . Simplified Furnace Diagram Radiation-Convection Heat Transfer . . . . . . . Complete Radiation Network for the System 1,2, 3, and 4 in Figure 4. 2 . . . . . xi Page 17 19 21 23 26 28 77 79 83 85 87 100 110 117 125 131 133 135 Figure 4.5 6.1 6.2 6.3 6.4 6.5 6.6 Radiation Network for the Surfaces 3, 5, and 8 in Figure 4. 2 . . Fuel Feed Rate as a Function of Fuel Moisture The Effect of Excess Air on the Combustion Chamber and in the Exit Flue Gas Temperature . The Effect of Excess Air on the Drying Air Tempera- ture, Furnace Efficiency, and the System Efficiencey . Cycle Times vs Number of Cycles for Test No. 1 . Simulated Specific Energy Consumption as a Function of Number of Cycles . . . . . Simulated Airflow Rates at the Beginning of Each Cycle xii Page 137 164 169 170 188 193 195 LIST OF APPENDICES Appendix A. Sample of CVCF Computer Input/Output Data B Experimental Results C. CVCF Computer Program 0 In-Bin Counterflow Drying Computer Program xiii Page 229 231 239 258 LIST OF SYMBOLS Area, m2 Ash content, percent Inert heat absorbing compound Concentration kg/m3 Convective thermal resistance 9C/N Specific heat kJ/kg °C Diameter m Diffusivity constant mZ/hr Diameter m Activation energy kJ Radiated thermal energy kJ/h View factor for radiation from surface m to surface n Friction factor Fractional void volume Mass of wet grain kg Grashoff number Gravitational constant 9.8 kgm m/kgfs2 Enthalpy kJ/kg Absolute humidity kg/kg Higher heating value kJ/kg Heat of combustion kJ Radiocity w/m2 xiv Ra RH rh 59 St SECO N><£ CO2 + 0 (2.33) 0H + H + B + H20 + B (2.34) CO+0+B+COZ+B (2.35) The equilibrium and rate constants for these individual reactions have been published (Jensen and Jones, 1978). However, no overall expres- sions are available due to the many variables and complexities associated with the oxidation of wood volatiles (Tillman, 1980). 2.2.2.4 Char Combustion Bradbury and Shafizadeh (1980) described the empirical formula for cellulose char as C6 7H3 30. The combustion of char is similar to the combustion of carbon described by Kanury (1975): 55 1. Diffusion of oxygen to the fuel surface; 2. Absorption of the oxygen on the surface; 3. Reaction of oxygen within the solid to form absorbed products; 4. Desorption of the absorbed products from the surface; 5. Diffusion of desorbed products away from the surface. The slowest (H’ the above steps determines the burning rate. In the case of carbon and char, combustion steps (2) and (4) are extremely fast. According to Glassman (1977), char oxidation is considered to be mass transport limited. This is due to the large particles and relatively high temperatures associated with char combustion. Thus, step (3) is much faster than steps (1) and (5). The burning rate is, therefore, controlled by the diffusion rate of oxygen to the char particles. Several simultaneous char combustion mechanisms have been suggested. Mulcahy and Young (1957) proposed the following mechan- isms based on hydroxy radicals: 20H + C + C0 + H20 (2.36) 0H + C + 00 + H (2.37) The classic Boudouard reaction was proposed by Glassman (1977): C + C02 + 2C0 (2.38) 56 Bradbury and Shafizadeh (1980) proposed: C* + 02 + C(O)m + C0 + CO (2.39) 2 C* + 02 + C(0)S + CD + 002 (2.40) The asterick designates an active reaction site, the subscript m is a mobile species and the subscript s is a stable species. All the solid carbon, including the remains on the grate and in the fly ash, must be oxidized via the above reactions to form C02. The reactions must have sufficient residence time, high temperature, and turbulance (the tree Ts) to complete the reactions in order to elim- inate particulate emissions from the furnace. 2.2.2.5 Concentric Vortex-Cell Furnace Combustion Theory The design and the operating principles of the concentric vortex-cell biomass furnace are discussed in section 2.1.3.2 The CVCF is similar to the spreader-stoker furnace if primary combustion air is introduced into the fuel bed through the grate supporting the fuel. The difference between a spreader-stoker and CVCF is that in the former furnace, most of the combustion air is introduced into the furnace through the grate. In the CVCF, secondary combustion air is supplied tangentially through a set of tuyeres located above the fuel bed. In some CVCF designs. the hearth is solid and no undergrate air is supplied (Claar et al., 1980; Wahby et al., 1981). Adams (1979, 1980) has developed a theoretical model for a spreader-stoker wood-waste-fired boiler that predicts the flight times 57 and mass reduction of combustible particles entrained in the furnace exhaust gases. The model also predicts the influence of fuel and operating parameters on particulate emissions. The spreader-stoker furnace employs the underfeed and overfeed principles described by Nicholls and Eilers (1934). Thus, the analyzed furnace is similar to the CVCF with undergrate air. Claar et al. (1981) suggested that this model can be modified to analytically model the CVCF combustion mechanism. However, the Adams (1979. 1980) models have not been validated. Several authors (Claar et al., 1980; Tuttle, 1977; Sukup, 1982) stressed the need to maintain high combustion temperatures and turbulence, and to provide sufficient time for complete combustion. The CVCF design provides suitable conditions for fuel combustion (Claar et al., 1980). However, there has not been a study corre- lating the combustion rate and the amount of particulate emissions from the CVCF to the temperature. turbulence, and time aspects of the CVCF. Thus, concentric vortex-cell furnaces have been designed and built by trial and error, rather than by optimization of the three Ts for a given furnace size and capacity (Claar et al., 1980 and 1981; and Wahby et al., 1981). 2.2.2.6 Heat Transfer in Furnaces Heat losses from a furnace determine the efficiency of fuel conversion into useful energy. The net energy loss from a furnace at steady-state conditions may be determined by the application of the first law of thermodynamics. The conservation of energy equation for 58 a steady state flow, open system with one inlet and one outlet is (Holman, 1980): where: Q Ah Qout in ri pi AHR out = 2 2 Win + H1 - H2 + (U2 - U1)/29c (2.41) n n = 1.Elllmith. - AHR - iEIMPlAhPl (2.42) hT - ho net heat loss (kJ/h) net work done on the system (kJ/h) enthalpy of products entering (1), or leaving (2) the system enthalpy at temperature T(K) (kJ/kg) enthalpy at 298 K (standard reference temperature) mass flow rate (kg/h) reactant i (i = air, fuel dry matter, H20 in fuel, etc.) product i of combustion (i = C02, H20, etc.) heat of combustion at a reference temperature of 298 K (kJ/kg) gravitational constant (9.81 kg-m/kg $2). velocity (m/s) 59 The heat loss determination using equations (2.41) and (2.42) requires that the temperature and the composition of the flue gases are known. 'Hueheat losses in the different components of the furnace can- not be determined from equation (2.41). Heat transfer in furnaces occurs in a combination of the three modes of heat transfer, namely: radiation, convection, and conduction. Heat transfer by convection, radiation, and conduction can be analyzed using standard techniques available in numerous heat transfer books such as Holman (1981) and Siegel and Howell (1981). However, heat transfer by radiation and convection in flames within the combustion chamber requires special treatment due to the thermal properties of flames. Also, the combined radiation-convection heat transfer should be studied since it occurs in a furnace. 2.2.2.6.] Convective Heat Transfer from Flames Calculation of heat transfer in flames involves the determina- tion of the convective heat transfer coefficient for the products of combustion. The thermal properties of flue gases and particulates entrained therein are quite different from those of air. Moreover, the use of the arithmetic mean of the surface and gas temperatures as the film temperature is inadequate due to the large variation in the physical and thermal properties with temperature. Also. chemical recombination reactions occur when the hot gases are cooled; consequently there may be some heat release within the boundary layer (Gray et al., 1976). 60 The thermal and transport properties of flue gas can be cal- culated if the composition of the gas is known. The properties are determined as follows: 1. specific heat (Holman, 1980): cm = m1C1+ mzc2 + N303 + . . . + mncn (2.43) 2. viscosity (Brokaw, 1958): : E “i (2.44) where pi an uj arc the viscosities of pure components i and j (N S/mz). (2.45) where: x, = the mole fraction of component i. xj = the mole fraction of component j. ”1 = the molecular weight of component i. Mj the molecular weight of component 5. 61 3. thermal conductivity (Mason and Saxena, 1958): n xi ki kmix : 131 E (2.46) 1;... j=1 x3 1) Jri ¢ij has the same value as that given by equation (2.45) (4) density: pi = P/(RiT) (2.47) 1/0 = m1/01 + m2/02 + . . . mn/on (2.48) where P is the total pressure (bars); R1 the gas constant of component i (J/kg K); p is the density of a mixture (kg/m3); pi is the density of component i; mi temperature of the mixture (K). is the mass fraction of component i; and T is the 5. Prandtl number: Pr = Cu/k (2.49) Due to the large variation of temperature and, therefore, the thermophysical properties, weighted averages of the above properties are used. The following equation was suggested by Gray et al. (1976): 62 T e "ave g éudT/(Te-Tw) (2.50) W where Te and Tw are the flame and the furnace wall temperature (K). Weighted averages for thermal conductivities, specific heats, densities, and Prandtl numbers are defined in a similar way as the dynamic viscosity (equation 2.50). Adams (1979) devel0ped empirical equations for determining the thermophysical properties of combustion gases, assuming that the combustion products behave as air at the same temperature: om = Mg/RuTe (2.51) M9 = 30.5 - 12.5 xW (2.52) “m = (182 + .245)~'r10"7 (2.53) where M9 is the molecular weight of flue gas; Ru is the universal gas constant (8.2*10'3 atm per kg-mole per K; Xw is the mole fraction of water (steam) in the flue gas; and Te is the flue gas temperature (K). The validity of the above equations is uncertain since the products of combustion do not behave as air. Empirical heat transfer correlations have been developed by several researchers and are generally accepted for estimating convec- tive heat transfer in tubes. plates and other geometrical shapes (Holman, 1981). The equations are of the form: 63 Nu c PrmRen (2.54) Nu th/k (2.55) The heat transfer is then given by: Q = Ahc(Te'Tw) (2.56) The following is the defining equation for the Reynolds number: Re = UpL/u (2.57) The magnitude of the Reynolds number determines whether the flow is laminar (Re < 2,300) or turbulent (Re > 10,000). Transition from laminar to turbulent flow takes place in the range of Reynolds numbers between 2,300 and 10,000 (Holman, 1981). 2.2.2.6.2 Radiative Heat Transfer from Flames The flame and the combustion products absorb and emit thermal radiation. Both gases and particulate materials present in the flame contribute to the absorbing/emitting potential of the flame. The gases present in flames from biomass fuels include 002, H20, CO, H2, 02. H, 0, 0H, N2, N and NO. The atomic and the homo- nuclear diatomic molecules (H,0,N and H2, 02) are the only non- absorbing constituents (Gray et al., 1976). The rest absorb and emit radiation over a specific wavelength range or band within the thermal spectrum. Carbon dioxide and water vapour are the most important absorbing/emitting gases in combustion products. They are the only 64 two gases considered in engineering calculations; for more precise calculations other gases should be included (Hottel and Sarofim, 1967). The range of wavelengths over which absorption occurs in carbon dioxide and water vapor gases have been given by (Gray et al., 1976) as 2.4-3, 4-4.8, and 12.5 to 16.5 pm for C02. and 2.2-3.3, 4.8-8.5, and 12-25 pm for H20. The absorptivities and emissivities of combustion gases vary considerably with wavelength. For engineering calculations, it is assumed that the gases are grey; thus, average values of absorptivities and emissivities for the whole of the wave- length range under consideration are used. The absorptivity and emissivity of CO2 and water vapour in a furnace depend upon the flame temperature, the mean beam length of the gas to the furnace walls, the partial pressure of the gases, the total pressure, the composition and the amount of the other gases in the combustion products, and the wall temperature. Adams (1979) suggested the following equation for calculating the emissivity of CO2 and H20 in wood-fired boilers: w = 493[0.6¢A (x +x )1'45/1 9 w co e (2.58) 8c where e is the combined emissivity of carbon dioxide and water CW vapor; A is the grate area (m2); Xw and xco are the mole fractions of 9 the water vapor and carbon dioxide, respectively; and Te is the flame temperature (K). 65 Soot, fly ash, and particulates in the flame emit and absorb radiant thermal energy. The emissivity and absorptivity of particu- lates depend upon their size, shape, concentration, composition, and the temperature. Also, the wavelength of the radiant energy affects the emissivity/absorptivity of the flue gases. Adams (1979) suggested the following empirical equation for the emissivity of soot in a wood fired boiler: as = 1-exp(-0.12/Ag) (2.59) Adams (1979) also suggested an equation for calculating the flame emissivity from the values of gas and particulate emissivities: e = as + Esw - (es)(esw) (2.60) Having determined the emissivity of the flame, the heat radiated to the wall is given by: 0r = Awea(T: - 1:) (2 61) 2.3 Grain Drying In this study a concentric vortex-cell biomass furnace (CVCF) is used to provide heat energy for an in-bin counterflow grain dryer (IBCF). The dryer was experimentally tested by Silva (1980) using liquid propane gas. The IBCF drying temperature is between 50 and 100°C depending upon the type of grain dried and the intended use of the grain (Silva, 1980; Kalchik et al., 1979). When feed corn is 66 dried with air at 65 to 100°C, the IBCF dryer is a high-temperature, high-capacity dryer. High-temperature drying may affect grain quality in different ways depending upon the initial grain quality and moisture content, and the drying/cooling rate. Grain and seeds are highly perishable if poorly handled. If well harvested, and stored at a low moisture content and low tempera- ture, grain will retain the original germinability and other desirable qualities for a long period of time. In the following sections, grain drying is reviewed. 2.3.1 Importance of Grain Drying The importance of grain drying has been discussed by Brooker et al. (1974). Drying facilitates early harvest, thus, reducing field losses from storm, insect damage, and natural shattering. Field con- ditions are often better for harvesting early in the season. Early harvest permits early and timely seedbed preparation for the next crop; this is particularly important in some tropical areas where two or more crops are raised in one year. Grain drying permits farmers to better plan the harvest season, and to make better use of labor and machinery since harvesting is not dependent on fluctuations of the grain moisture content in the field. Finally, the early harvest enables farmers to take advantage of higher prices early in the harvest season. The most important advantage of grain drying is that it permits long-time storage without deterioration in quality. By removing excess moisture from the grain, the possibility of natural heating of the 67 grain due to molding and respiration is reduced. Thus, grain via- bility is maintained during storage. Dried grain (at moisture content below 13%) is less prone to insect, mites, and fungal damage than wet grain. 2.3.2 Corn Quality as Affected by Drying Methods The desirable corn properties are dependent upon the intended use of the corn. In the US corn is mainly used as animal feed, with smaller usage as human food, seed and industrial starch manufacturing. Corn quality is dependent upon several factors: (1) the variety characteristics. (2) the environmental conditions during growth, (3) the time and the harvesting procedure, (4) the drying method, and (5) the storage practice (Brooker et al., 1974). During the drying process, corn quality is affected by the grain temperature and the drying rate. Corn in the US is officially graded for quality under the Grain Standards Act. The grades and grade requirements are listed in Table 2.5. There are other properties which are of importance to specific corn users that are excluded from the standard, such as nutritional value, millability, viability, and susceptibility to breakage. 2.3.2.1 Effect of Drying on Nutritional Feed Value The most important quality factor of corn for animal feed is the nutritional value. The effect of drying temperature on the nutri- tional value of corn for animal feed has received considerable research attention. Hathaway et al. (1952) found that corn dried at temperature 68 Table 2.5.--Numerical grades and sample grade requirements for US corn. Maximum Limits Minimum Damaged kernels, % Grade testweight Moisture BCFM LB/bu % % heat damaged Total 1 56 14.0 2.0 0.1 3.0 2 54 15.5 3.0 0.2 5.0 3 52 17.5 4.0 0.5 7.0 4 49 20.0 5.0 1.0 10.0 5 46 23.0 7.0 3.0 15.0 Sample grade shall be corn which does not meet the requirements for any of the grades numbers 1 to 5; or which contains stones; or which is musty, sour, or burned; or which has any commercially objectionable foreign odor; or which is otherwise of distinctly low quality. Source: Brooker et al. (1974). above 60°C significantly decreased its energy content and palata- bility. Sullivan et al. (1975) reported that heat has a definite effect on the nutritional value of corn; also, that the decrease in commercial quality due to drying at high temperatures may not necessarily result in a decreased value of corn as animal feed. Jensen et al. (1960) reported that drying temperatures of 60°C, 82.2°C, and 104°C, have no deleterious effect on the nutritive value of corn for swine as measured by growth rate and feed use. Jensen (1978) showed that roasting corn at 15% and 23% moisture at 127°C and 150°C reduced the availability of lysine; he found that niacin is unaffected by roasting temperature, but pyridoxine (vitamin 86) 69 availability is significantly reduced in 14% moisture corn when it is heated at 160°C. From the above review it appears that corn which reaches temperatures above 60°C undergoes some minor nutritional changes. Nutritionists do not agree on the effects of drying temperature on the feed value of corn (Brooker et al., 1974). It is generally recognized that physical and chemical properties such as consistency, energy content. palatability, hardness, color, moisture, vitamins. protein and amino acid profile are affected by high drying temperature (Williamson, 1975). 2.3.2.2 Effect of Drying on Corn Milling Quality Farmers and elevator operators who dry corn often consider only its feed value. Corn millers are concerned about the increasing volume of artificially dried corn being marketed (Freeman, 1978; Rutledge, 1978). High starch yield (millability), maximum yield of selected fractions and prime products mix, and low fat content are the most important desirable characteristics of corn for milling. Brekke et al. (1973) compared the corn milling response of in-bin natural air drying with artificial drying in a small experimental fluidized dryer at air temperatures from 32 to 143°C (maximum corn temperatures were 32 to 104°C, respectively). The yield of total grits recovered by sieving, aspiration, and flotation decreased with increasing drying air temperature; also, the fat content of the grits increased with increasing air temperatures. Prime products recovered by rolling and 70 grading followed similar patterns; however, sometimes yields and fat contents were less satisfactory. The results also showed that the cold paste viscosity of selected products increased if corn dried at elevated temperatures. The corn dried with natural air had the best dry-milling quality; drying at 60°C yielded corn of acceptable drying- milling quality except for a high percentage of stress cracks. Freeman (1978) discussed the quality factors affecting the value of corn for wet milling. Drying at high temperatures causes "case hardening" of proteins. Case-hardened protein affects the millability by impairing separation and purification of starch. The result is starch with a high protein content and reduced viscosity. The drying temperature, drying rate, and the initial corn moisture content determine the degree of case hardening. High drying tempera- tures may also destroy some amino acids, especially lysine. Watson and Mirata (1962) concluded that since kernel viability is evidently more easily altered by drying conditions than the other properties examined, corn dried to preserve viability should invari- ably be suited for starch manufacture. The drying temperature should not exceed 71°C. 2.3.2.3 Drying Corn for Seed Generally the techniques used to dry seeds do not differ greatly from those used to dry grain for other purposes. However, extra dryer control and management must be practiced in order to ensure a high degree of germination (Copeland, 1976). The drying air temperature, the drying rate, and the initial moisture content are 71 the most critical factors affecting the germinating quality. Copeland stated that the product temperature limit varies with the type of seed, but should not exceed 38°C; the highest safe temperature also depends on the initial moisture content. Ulileman and Ullstrup (cited in Hukill, 1954) showed that seed corn can be dried safely at 49°C, if the moisture content is less than 25%; for moisture above 25%, 38°C, is the upper limit. An excessive drying rate may cause stress cracks. Over-dried seeds are susceptible to mechanical damage, which is detrimental to seed quality. 2.3.2.4 The Effect of Drying on Commercial Grade The effects of artificial corn drying on its composition, nutritional value, viability as seed, and industrial processing have been discussed in the previous sections of this review. These factors are not included in the determination of commercial grade. As shown in Table 2.5, the only factors considered in the grain standard code for corn are: (1) the testweight, (2) the moisture content, (3) the broken corn and heat damaged corn, and (4) the presence of foreign materials. Artificial drying has a direct effect on the testweight, the moisture content, and on the percentage of heat damaged corn, and has an indirect effect on broken corn. These factors will be reviewed in the following sections. 72 2.3.2.5 Testweight The corn testweight is its bulk density and is influenced by grain shape, grain surface texture, moisture content, type and amount of impurities, size and uniformity, temperature, and other factors that affect the packing characteristics. According to Freeman (1978), testweight may indirectly indicate the wet milling quality of corn. High-temperature drying may reduce the extent of kernal shrinkage resulting from moisture removal and, hence, result in low testweight. Hill (1975) studied the effect of drying temperature on test- weight of shelled corn. According to his observations, the testweight increases during drying. The increase is due to shrinkage with a loss of moisture and decrease of the coefficient of friction on the surface, thus permitting closer packing of the kernels. The testweight increase is less at higher drying temperatures, possibly due to case hardening. The testweight reaches a maximum and declines with further drying. The maximum testweight is reached at 14 to 16% moisture (Brooker et al., 1974). The amount of testweight increase with drying depends upon: (1) the degree of kernal damage, (2) the initial moisture content, (3) the temperature reached by the grain during the drying process, (4) the final moisture content, and (5) the grain variety. Early harvested, high moisture grain is not exposed to much weathering and shows a higher testweight after drying than the same grain harvested later at a lower moisture content (Brooker et al., 1974). 73 A higher testweight corn offers some saving in storage since less storage volume is required to store the same amount of dry matter. 2.3.2.6 Stress Cracks and Broken Corn Although drying per se does not directly affect the number of broken kernals, it is well known that grain is physically and physio— logically damaged when dried at excessively high temperatures. The degree of damage depends upon the maximum temperature reached by the grain and the length of the period during which the high temperature is sustained. The drying and cooling processes directly affect the degree of stress cracking, and thus determine the susceptibility of corn to breakage during subsequent handling. Thompson and Foster (1963) defined stress cracks as the fissures in the endosperm, or starch inside the kernel, in which the seed coat is not ruptured. Ross and White (1972) studied the effect of overdrying in stress cracking in white corn. Their results show a general decrease in stress cracking as the white corn dried to lower moisture content, and as drying started at lower moisture contents. These phenomena are difficult to explain. There may be some physical and chemical changes such as gelatinization during over-drying which make the grain kernel more resistant to cracking during the cooling period. Generally. stress cracking decreases with decreasing drying air temperature. Slow cooling of both the white and yellow corn after drying results in a dramatic reduction in the number of checked kernels, particularly after subjecting the corn at drying air temperatures above 71°C. 74 Gustafson et al. (1978) concluded that there is no significant increase in breakage susceptibility when corn is dried to 18% moisture (or above) in a high-temperature dryer; the product of the heating time and the change of moisture content (under 18%) appears to be the best predictor of the change in breakage. Freeman (1973) indicated that broken kernels too large to be removed by screening for wet milling may release starch granules during steeping. Free starch in the steeping water causes fouling of evapora- tor surfaces during steep water concentration. 2.3.2.7 Corn Color Artificial drying affects other grain characteristics including color. Ross and White (1972) concluded that darkening and yellowing of white corn was apparent when it was dried with air at 88°C and 104°C, and when drying started at an initial moisture content above 25%. Discoloration was only slight in samples dried at 71°C from any initial moisture content, and for samples dried from 25% initial moisture con- tent (mc) to 14% final me with drying air at 104°C. 2.3.3 Drying Systems There are three basic methods of grain drying: (1) high- temperature drying, (2) low-temperature drying, and (3) combination drying. In the US, high-temperature drying has been the primary tech- nique for more than 25 years. This method is fast, but has a high fossil-fuel requirement, and can result in a low grain quality. Low- temperature grain drying (energy may be obtained from electricity, 75 liquid propane, solar energy, or any other heat source) is an energy efficient process and often results in high-quality grain, if properly managed. Mold spoilage risk is the main problem encountered in warm and humid areas. Natural drying is a low-temperature drying method and takes place when grain is either left standing (or stacked) in the field, or harvested and dried in a crib. The latter method is often practiced in the Third World and is the most risky since it exposes grain to the weather, insects, rodents, birds, and other destructive elements. Combination drying processes for drying shelled corn started in the US in the late 1970's (Brooker et al., 1978). In these processes high-speed batch or continuous flow drying is combined with low heat or natural air in-bin drying. The high-speed, high-temperature dryers dry the corn to a moisture range of 18-23%. The corn is then trans- ferred to storage where it is slowly dried to a safe storage moisture content. Combination drying offers a number of advantages, including: (1) increased output, (2) increased fuel efficiency, (3) improved product quality (compared to corn dried by high-speed, high-temperature processes). Brooker et al. (1978) subdivided the on-the-farm high- and low-temperature drying methods into the following categories: (1) high- speed, high-temperature batch and continuous dryers; (2) continuous in-bin drying systems; (3) batch-in-bin drying systems with and without stirring; (4) low-heat and no-heat in-bin drying systems with and with- out stirring; and (5) combination systems, in which high-speed batch or continuous flow systems are combined with low-heat in-bin drying sys- tems. 76 2.3.3.1 Column Batch Dryers Column batch dryers are stationary bed dryers, in which the air moves across a stationary grain column (see Figure 2.7). The dryer is often portable so that it can be moved from location to location when not filled with grain. According to Brooker et al. (1974), column batch dryers have the following characteristics: (1) column thickness is usually from 30 to 46 cm; (2) column batch dryers operate at high air- flow rates (over 40 m3/min/m3); (3) drying air temperatures vary from 82 to 116°C; (4) due to the high flow rate coupled with a narrow column, the moisture gradient across the column is less than with batch in-bin systems; and (5) drying is completed in a 1 to 3 hour period, depending on the initial grain moisture content and the need for cool- ing. Column dryers are popular because of theirsimple construction and operation, and because the initial cost is generally lower than of continuous flow types (Sutherland, 1975). They are suitable for mod- erate grain volumes (250-650 tons annually) with high initial moisture contents. Because the dryer has no storage function, it requires well planned and coordinated handling and storage systems. The fuel consumption and, therefore, the Operating costs depend upon the moisture removal range. The fuel efficiency decreases with decreasing moisture removal range. It requires (at 30% initial mc) about 5,800 kJ/kg of water removed when the final moisture is 25%, and 6,970 to 8,140 kJ/kg of water when the final moisture content is 15%. 77 Jg" GRAIN HOPPER LEVELING AUGER GRAIN COLUMN PERFORATED SCREENS \ HEAT AND COOL PLENUM \ / DISCHARGE AUGER Figure 2.7. Batch Column Dryer (from Brooker et al., 1974). 78 The thicker the grain column is in a column dryer, the lower is the specific fuel consumption. However, thicker grain columns result in an increase in the moisture content gradient across the column. Kirk (1959) investigated column thicknesses of 10.2, 20.3, 30.5, and 40.6 cm and concluded that: (1) the 20.3, 30.5, and 40.6 cm columns are similar in drying air requirements; (2) the operating costs do not significantly increase with an increase of static pressure of up to 0.5 kPa for a grain column thickness of 20.3 to 40.6 cm; (3) in the static pressure range of 0.06 to 0.5 kPa, the drying capacity increases linearly with static pressure; and (4) there is no signifi- cant difference in drying capacity with drying column thickness between 20.3 and 40.6 cm. The moisture and the temperature gradients across the drying column in a column batch dryer along with the dryer operating costs can be reduced either by decreasing the drying airflow rate at a con- stant air temperature, or by decreasing the drying air temperature at a constant airflow rate (Morey et al., 1976). 2.3.3.2 In-Bin Dryers In-bin drying systems dry and cool the grain in a bin designed either as a batch in-bin dryer or as a drying-storage bin. In the latter case the grain is left in the same bin for storage (see Figure 2.8). In-bin drying may be categorized in different ways. Brooker et al. (1974) categorized in-bin drying as (1) full-bin drying, (2) layer-drying, and (3) batch in-bin drying. 79 vqr-Q-rfll ‘ ‘- .Eoumam mewaLa can-:H mcaumaoaem» -384 a mo upuueosum .o: 1.: n M. u a r. .. SPF... L m N masawd 80 In full-bin drying a single batch of grain (up to 5 m) is dried at relatively low airflow. To avoid grain deterioration during drying, the initial moisture content of the corn must not be above 21-23% (Brooker et al., 1974). In the layer-drying process an initial shallow layer of grain is placed in the bin and drying is started. As the drying zone moves upwards, other layers of grain are added periodically, so that a layer of wet grain remains ahead of the drying zone. Eventually the bin is filled; drying continues until the entire grain mass is dried. This process allows the initial layers of grain to be relatively wet; they receive more drying air (and therefore dry faster) than the subse- quent drier grain placed at the top of the bin. Batch-in-bin drying is a process in which batches of grain are dried in a bin and transferred to a storage bin. This process utilizes large quantities of heated air which is forced through relatively shallow grain beds. Therefore, the drying rate is considerably higher than in the other in-bin drying systems. In-bin dryers may be operated with high or low temperatures (heated air) or with natural air (unheated air). High-temperature in-bin drying is undesirable in full-bin and layer drying due to excessive overdrying at the bottom of the bin. However, stirring devices may be used to solve this problem. Natural air and low-termperature drying are similar processes (Bakker-Arkema et al., 1978). The difference is that no heat is added in case of natural air drying. Low-temperature drying requires raising the drying air 3 to 6°C above the ambient temperature by either 81 electric heat, solar energy, or other heat sources (Zink et al., 1978). Liquid propane and electricity are the most common heat sources for low-temperature drying; both require low capital investment. Liquid propane gas is usually not used since it requires interval timers to limit the heating rate. The airflow rate required for a drying system depends on the drying system. the harvest date, harvest moisture content, and the location. Adding heat, even in small amounts, increases the drying capacity in a low-temperature drying system (by decreasing the drying air relative humidity). The temperature increase also results in faster mold development. To reduce mold growth, the average tempera- ture in the bin should be below 10°C. With addition of low heat, the airflow can be limited to 2.4m3/min-m3 for 24-26% mc corn, 1.6m3/min-m3 for 22-24% mc, and 0.8 m3/min-m3 for 20-22% mc corn (Brooker et al., 1978). Although low-heat and natural air drying are slow, the quality of the finished grain is frequently high due to the low application of heat (Kalchik et al., 1979). The main disadvantage of in-bin drying is that the grain at the bottom may be overdried and at the top underdried. Drying is stopped after the average moisture content in the bin reaches a desired value. Since unloading does not allow perfect mixing and blending to obtain a uniform moisture content, the corn may deteriorate during storage. 82 Several improvements have been incorporated in in-bin drying to reduce the vertical moisture gradient. This includes grain recir- culating, in-bin counterflow drying, grain stirring, and drying with alternating heated and unheated air (Browning et al., 1971). According to Brooker et al. (1978), the specific fuel con- sumption of deep bin-in-storage drying is low (3,500 kJ/kg, or lower in some cases). Therefore, these systems are usually used in combina- tion with other drying systems. 2.3.3.3 High-Speed Continuous Crossflow Dryers Crossflow dryers have a wet grain holding bin at the top (see Figure 2.9). Grain flows by gravity from the top to the bottom through drying and cooling columns which are 20-45 cm wide. A column thickness of 28-36 cm is most common. Two fans. one for heating and one for cooling, are usually used. The drying rate and the final moisture content are mainly dependent upon the grain velocity, the air tempera- ture, the airflow rate, and the initial moisture content. Moisture is usually controlled by regulating the grain flow rate by a metering auger at the bottom of the dryer, while maintaining a constant air temperature and airflow rate. The auger speed responds to a tempera- ture sensor located in the grain column near the lower edge of the drying section. The drying characteristics of crossflow dryers are similar to those of column batch dryers. The grain on the plenum side is over- dried while that on the exhaust side is underdried. 83 FILLING AUGER FAN -— AND —— ;: HEATER HEATED AIR PLENUM Gnamcowun COOLING AIR PLENUM GRAIN METER UNLOADING AUGER Figure 2.9. Schematic of a Crossflow (Continuous Flow) Grain Dryer (From Brooker et al., 1974). 84 In some crossflow designs. the cooling air is reused. This allows recovery of heat from the cooled grain and also reduces the moisture content gradient across the grain column (Bakker-Arkema et al., 1980). Some designs incorporate a metering device that causes the grain on the plenum side to move faster than the grain on the outside (Rodriguez, 1982). Other designs use a "turnflow device" to interchange the grain on the plenum side with the grain on the exhaust side (Fontana, 1983). These improvements are intended to reduce the moisture content gradient across the grain columns. 2.3.2.4 Concurrentflow Dryers Concurrent flow drying is a relatively new grain drying tech- nique (Dalpasquale, 1981; Bakker-Arkema et al., 1983). In a concurrent- flow dryer, high temperature (BO-300°C) air flows in the same direc- tion as the grain (see Figure 2.10). The hot air only encounters the cool and wet grain. Intense heat and mass transfer takes place at the grain/air inlet. causing rapid evaporative cooling of the air and heating of the grain. The latter is accompanied by rapid drying of the grain. The grain temperature remains considerably below the inlet air temperature. As the grain and the air flow through the drying bed, an equilibrium is reached (the equilibrium temperature is between the inlet air temperature and the initial grain temperature). The absolute and relative humidity of the air progressively increase as the grain dries.' The major advantage of concurrentflow drying is that all the grain kernels are subjected to the same thermal process. Thus, there Figure 2.10. mmmonwzb Wet grain Drying floor Exhaust air Tempering section Air recirculation Counterflow cooler Metering rolls 85 H: I: J: L: M: Schematic of a Concurrent Flow Dryer. Ambient air Hot air--First stage Hot air--Second stage Ambient air--Cooler Dry grain 86 is no moisture difference among the kernels. This results in better grain quality. Counterflow cooling has usually been combined with concurrent- flow drying (Bakker-Arkema et al., 1982). Cold air first encounters the coldest grain, thus limiting thermal stresses. Several authors have reported the advantages of concurrentflow drying including: (1) high capacity (in multi-stage concurrentflow dryers), (2) improved grain quality, (3) greater flexibility for adaptation to different craps, and (4) high termal effiency (Westelaken and Bakker-Arkema, 1978; MDhlbauer et al. , 1978; Bakker-Arkema et al. , 1982, 1983; Dalpasquale. 1982; Fontana. 1983). 2.3.3.5 In-Bin Counterflow Dryers In-bin counterflow drying is an intermittent process in which the grain in a bin flows downwards, and the drying air flows in the opposite direction (see Figure 2.11). The drying process is the same as for the batch-in-bin dryer. Dried grain is removed from the bottom of the bin by the means of a tapered sweep auger which removes a thin- layer of grain from the bottom of the grain bed and delivers it to a vertical auger; the auger, in turn, discharges the hot dry grain into a cooling bin. The sweep auger is activated by a temperature sensing element placed about 46 cm above the false floor. As drying progresses, the temperature in the region of the sensor increases. When a pre- selected temperature is reached, the sweep auger is activated; it makes one complete turn around the bin removing a 7-9 cm thick layer of dry warm grain. As the auger completes the turn, damp grain 87 =.soumam mco>>Fsm= mcwxgo zopecmpczou cwnlca cm mo zmw> pecamucH use we upumemzum .HH.~ mcamra ; . . . .Onlrwfin.rw. ”4 88 moves into the sensor's region and the temperature at that point drops. The auger is turned off and waits for the next cycle. Thus, the mode of operation is intermittent. In a countinuous-flow, counterflow dryer, the sweep auger speed is set so that the grain receives suffi- cient residence time in the dryer such that the grain is dried to a desirable moisture content by the time it is discharged. The grain discharge rate is used to control the final moisture content. In the counterflow drying process, the warm saturated air leaving the drying zone passes the cool and relatively wet incoming grain. Some energy is used to warm the cool grain, and condensation may occur on the cool grain especially if the bed is deep and the initial grain temperature is low (Brooker et al., 1974). Silva (1980), Bakker-Arkema et al. (1980), and Kalchik et al. (1981) investigated the performance of the in-bin continuous flow counterflow dryer. The specific energy consumption decreases with increasing drying temperature (4,390 to 5,110 kJ/kg of water for drying air temperatures of 93 and 50°C, respectiveIYI; corn quality is not seriously affected by drying at high temperatures (82-93°C). 2.3.3.6 Combination Drying In combination drying, wet grain is dried down to 18-24% mc (wb) in a high-temperature batch or continuous dryer. The grain is transferred to a bin in which drying is completed by a low-temperature in-bin drying process (Shove, 1978). Dryeration is a form of combina- tion drying (Foster, 1964). In the dryeration process, grain is dried in three phases. The first phase consists of drying the grain down 89 to 2-3 percentage points above the desired final moisture in a high- temperature dryer. The partially dried grain is transferred to a tempering bin (second phase) and held hot for 6 to 10 hours. In the third phase the grain is finally cooled by aeration at low airflow (.01-.03 m3/min-m3) (Bakker-Arkema et al., 1978; McKenzie et al., 1972). ’ Gustafson et al. (1976) and Shove and White (1977) showed that the susceptibility to breakage in the 15-18% mc range is substantially reduced by eliminating immediate cooling after high-temperature drying. According to Kalchik et al. (1981), combination drying results in increased fuel efficiency and increased drying capacity. The average specific energy consumption for high-temperature drying is 5000 kJ/kg compared with 3800 kJ/kg for combination drying when corn is dried from 26 to 15.5% mc. 2.3.3.7 Drying Systems Suitable for Biomass Heating The drying systems discussed previously have been designed for use with conventional fossil fuels as energy sources for heating the drying air. A batch dryer has distinct loading, heating (drying), cooling, and unloading phases during a drying cycle. Thus, the fuel burner is turned on during the heating phase and is turned off during the cooling, unloading, and the loading phases. In contrast, continu- ous flow dryers require a continuous supply of heated air for the drying sections. The concentric vortex-cell furnace and other solid fuel burners require a continuous supply of fuel and combustion air to maintain 90 steady-state conditions. Sudden interruption of either fuel or air supply results in undesirable transient conditions characterized by the emission of products of incomplete combustion. Thus, a biomass furnace cannot be used for heating drying-air for a batch dryer without incorporation of a heated air diversion system to divert the heated air to some other use during the loading, cooling, and unloading phases. Unlike the batch dryers, continuous flow dryers are easily convertible to the use of biomass heated air. 2.4 Grain Drying Theory and Simulation Much research has been conducted to study the processes by which water is removed from biological materials. The drying process consists of simultaneous heat and mass (moisture) transfer. Heat is required to provide the energy for evaporating moisture from the drying grain. The moisture is removed from the surface of the grain by an external drying medium, usually air. Bakker-Arkema et al. (1974) listed six possible modes of mois— ture removal from cereal grains: (1) surface forces (capillary flow); (2) liquid movement due to moisture concentration differences (liquid diffusion); (3) liquid movement due to diffusion of moisture on the pore surfaces (surface diffusion); (4) vapor movement due to moisture concentration differences (vapor diffusion); (5) vapor movement due to temperature differences (thermal diffusion); and (6) water and vapor movement due to total pressure differences (hydrodynamic flow). The exact manner in which water leaves the grain is dependent upon drying 91 air temperature, air velocity, moisture concentration, and product type and condition. Two drying rate periods have been identified: (1) the constant- rate period during the initial drying of extremely moist grain, and (2) the falling-rate period. Cereal grains dry during the falling- rate period (Bakshi and Singh, 1979). The drying rate decreases con- tinuously as the moisture content decreases during the course of drying. The thermal and the physical properties of the grain which are moisture content dependent also change as drying progresses. Thus, the prediction of the drying rate during the falling-rate period is more complicated than during the constant-rate period. The basic drying theory consists of a thin-layer drying equa- tion (which is equivalent to drying a single grain kernel), coupled to a series of deep-bed drying equations. 2.4.1 Thin-Layer Drying Models A "Thin-layer" of grain refers to a layer of grain that is approximately one kernel deep. A thin-layer drying equation predicts the drying rate of single kernels of grain as a function of the drying conditions. Empirical, theoretical, and semi-theoretical thin-layer equa- tions have been developed for different crops. Empirical equations are developed by analyzing statistically the data obtained by drying thin layers of grain. Theoretical equations are based upon the diffusion theory (Crank, 1979). Thin- layer drying equations for several cereal grains have been reviewed and 92 published (Bakker-Arkema et al., 1983; Steffe, 1979; Dalpasquale, I981; Bakshi and Singh, 1979). First the thin-layer equations for corn will be reviewed. Thompson et al. (1968) developed an empirical thin-layer drying equation for corn in the range of 60 to 150°C. t = Atn (MR) + B (1n(MR)]2 (2.62) where: A = 0.004880 - 1.86178 B = 427.36 exp (-0.0330) The most commonly used thin-layer equations are modified versions of the Thompson equation (Pfost et al., 1976). Flood et al. (1969) developed an empirical drying equation for corn in the temperature range of 2 to 22°C: MR = exp (-kt0'663) (2-54l where: MR = (M-Me)/(Mo-Me) k = exp (~xty) x = [6.014 + 1.45*10'4(rh)2]'5 -(1.86 + 32)[0.33.41:10'3 + 3111043040215 y = 0.125 - 2.1971110‘3 - [1.88 + 32][2.3*10'5(rh) + 5.8*1o’5] Rugumayo (1979) developed an equation from mathematical diffu- sion theory similar to the model presented by Chu and Hustrulid (1968) for low temperature thin-layer drying of corn. The Troeger and Hukill 93 (1970). the Muh (1974), and the Misra (1978) equations have been compared by Rugumayo (1979). In the development of the above equations, it is assumed that moisture is uniformly distributed within the individual kernels of the grain. Therefore, the equations cannot be used to predict the moisture gradient within the kernels. The knowledge of the moisture content gra- dient within single kernels is needed in order to simulate the tempering process in grain-drying (Rodriguez, 1982; Fontana. 1983; Sabbah, 1971). The Crank (1979) spherical diffusion equation can be used to calculate the moisture gradient in the grain kernels during the dry- ing process: 0C 33C 20C '3? ‘ ”('51: * 'r—af) (2'64) 07': 30! 3'14 23” a " ”Par-1 * a?) ”‘55) The following are the boundary conditions: M(v.6) ‘ "111111111 M(o,t) = Me c(r,o) = cinitial c(o,t) = C8 where C is moisture concentration (kg/m3) and M is the moisture content (decimal). The analytical solution of equation (2.64) for the moisture 94 content distribution and the average moisture content of various regularly shaped bodies can be found 'h1 Crank (1979). For a sphere the average moisture content is (Sabbah, 1971): _ 1 "2112 2 MR - '57 exp - [—17— X ] (2.66) :3401 "MB n 1 where: X =%-(Dt)2 : The diffusion coefficient (0) is a function of the moisture content, the grain temperature, and the type of grain. Sabbah (1971) using a spherical diffusion equation found for corn: 0m = 5.196*10‘3M[exp(-2673.33/8)] (2.67) A diffusion constant equation should be used with the equation for which it was developed. The Chu and Hustrulid (1968) equation is a mass concentration-based diffusivity and is used to solve the equation: M - M 2 - H* _ fi =.%gxp[ n/(R Dc)l (2,53) . e where: M* = 1'0655Mo - 0.0108 R radius of a sphere, m 1.51exp {(0.0451a - 5.49)Me - 2513/Ta} Dc The grain shape is not assumed spherical; the equivalent radius is defined as: 95 R - 3V/s kernel volume, m3 where: V kernel surface area, m2 V) II The temperature and the moisture content Ta and Me’ respectively, are the drying air temperature and the corresponding equilibrium moisture content. 2.4.2 Equilibrium Moisture Content Models The thin-layer equations discussed above require the grain equilibrium moisture content at the drying air conditions (temperature and relative humidity). The following are some of the equations used for predicting the equilibrium moisture contents of corn: Thompson et al. (1968): Me = [£n(a - rh/{0.382(1.8T + 82}]'5 (2.69) Kalchick et al. (1979): H 0.027“ °°°°2(R”) /1n1 052 (2.70) Me 0.3968xp(.5*RH)/£hT 5299.99 (2.71) where Me is the equilibrium moisture content (decimal); RH and rh are the percent and the decimal air relative humidity, respectively; and T is the air temperature (°C). Other equilibrium moisture content equations can be found in Bakker-Arkema et al. (1983). The equilibrium 96 moisture content models should be used with the thin-layer equations for which they were validated. 2.4.3 Deep-bed Drying Models For the research presented in this dissertation an in-bin semicontinuous flow dryer was used to dry corn. The in-bin drying process can be modeled using a deep-bed simulation model. Bakker- Arkema et al. (1974) developed a series of deep bed drying simulation models. The models are subject to the following assumptions: 1. there is no shrinkage of the grain bed during the drying process; 2. the temperature and the moisture content gradients within an individual particle are negligible; 3. particle to particle thermal conductivity is negligible; 4. the airflow is plug type (no wall effects); 5. there is no heat loss through the dryer (or bin) walls and the walls have negligible heat capacity; 6. the heat capacity of the moist air and of the grain is constant during short-time periods; 7. accurate thin-layer, moisture equilibrium isotherms and latent heat of vaporization equations are known. The models (Bakker-Arkema et al., 1974; Rugumayo, 1979) con- sider the basic laws of mass and heat transfer in the analysis of the drying process. Thin-layer equations developed by Rugumayo (1979), Thompson et al. (1968), and Misra (1978), are used to predict the drying rate of corn. Drying processes of other crops can be modeled if 97 the appropriate auxiliary equations are included. The models solve for the grain and air temperature, the air absolute humidity and relative humidity, and the grain moisture content as a function of time and position in the drying or cooling bed. The accuracy of the results depends upon the accuracy of the drying rate equations, and theequations used to predict the thermal and the physical pr0perties of a given product. 2.4.3.1 Stationary Deep-bed Model Brooker et al. (1974) analyzed stationary bed drying by making energy and mass balances on a differential volume located at an arbitrary position in the stationary bed. The following set of four differential equations was obtained: ST -h a ( _ = 1-9) (2.72) 82 Ga ca + Ga CVH h + c (T-e) 32': h+a *M (T-G) ' f C +v 1M5 .90 (2’73) ”9 Cp 09 cw 0P P pp cw ‘ a 32 3H .. - 32.311 .52 _ 93 at (2.74) %¥-= an appropriate thin-layer drying equation (2.75) The boundary and initial conditions are: a. T(0.t) = Tinlet b' e(x,o) = einitial C. d. H(o.t) = Hinlet M(x,o) = Minitial. 98 CHAPTER 3 FURNACE DESIGN, CONSTRUCTION, AND OPERATION 3.1 Design and Construction The concentric vortex-cell biomass furnace (CVCF) is a modifi- cation of a furnace developed at Iowa State University and described by Claar et al. (1981). The furnace consists of two concentric steel cylinders (see Figure 3.1). The outer cylinder is 1.22 m in diameter and 3.66 m high and is formed from 11 gage (GA) mild steel. The inner cylinder is 96.5 cm in diameter and 3.35 m high and made of 3 GA mild steel. The top 2.75 m of the inner cylinder is lined with 11.4 cm thick firebricks. The firebricks are supported by a round steel table with a hole in the middle. The outside and the inner diameters are 96 and 66 cm, respectively. The height of the table is 56 cm. The cylinders are covered with lids through which a chimney, 61 cm high and 36 cm in diameter, made of 3 GA mild steel is inserted. The inner lid is insulated (on the inside) with 2.5 cm thick blanket insulation with a thermal conductivity of 3.444 W/m-°C at 800°C. A 16 GA eductor is horizontally connected at the top of the chimney. The eductor consists of a 147 cm long diffuser with a throat diameter of 36 cm and a diffuser tail end diameter of 66 cm, and an 8-cm nozzle (see Figure 3.4). The increase in the air velocity at the nozzle tip and the expansion of the flue gases in the diffuser result in a 99 100 .momcaad ppmulxoaco> o4cucmocou use we o4pasmgom .4.m ocsmwd mmmrwdflflxnxvsqx ...1.......... . 1.3.45.1. ...P? u. u... 44...... 4...)... .. .. .. .. .... a. a}. e ee ewe. .. ‘v . ... ”I I ‘0. Ag ”459098 Be. agnaafl eeeee AD 39m “Pm zwmilfila. II I. I mg Egan . I ‘ u a D O n D u .unhiuucn 101 0.25-0.75 kPa pressure reduction at the top of the chimney inside the combustion chamber. The pressure reduction depends upon the air velocity at the nozzle tip. A 76 cm diameter sloping (12 degree) grate with a cross 2 is located 61 cm above the bottom of the sectional area of 0.43 m inner cylinder to support the fuel pile. An opening in the cylinder below the grate provides access to remove the undergrate ashes. The opening is controlled by means of two doors, the outside is hinged on the outer cylinder and the inside is bolted to the inner cylinder. A 15 cm diameter pipe is located in a hole cut on the inside door. A disc slides inside the pipe to regulate the undergrate airflow rate. The disc is adjusted manually by means of an axial bolt welded on the disc and rotating through a nut welded on the outside door (the disc-pipe mechanism acts as a gatevalve for regulating the undergrate airflow). Both doors can be opened to remove the ash below the grate. A 25 cm2 square opening_is cut 90 cm above the furnace base in" both cylinders to provide access to the combustion chamber for lighting and observation of the fire. The opening is covered with firebricks on the inside and a mild steel door on the outside. The inner cylinder (lined with firebricks) has two rows of tuyeres. three tuyeres per row, positioned at an angle of 18 degrees with respect to the cylinder wall. The tuyeres are fabricated from 5 cm nominal diameter, schedule 40, seamed stainless-steel pipe. Each tuyere has a cross-sectional flow area of 22 cm2. 102 The fuel is placed onto the sloping grate by an automatic auger feeder system. The system consists of an electric motor-driven feed wagon, a feed hopper and a stoker-auger driven by a 0.37 kw (1/2 Hp) motor. The rotational speed of the stoker-auger is controlled by a speed reduction system consisting of 4 to 1 variable speed belt drive, a 200 to 1 double reduction gearbox, and finally a 4 to 1 chain-drive reducer. The minimum rpm is 0.5; the maximum 2. The combustion and the eductor air streams are supplied by a 0.74 kw (1Hp), 36 cm axial fan. The fan forces air into a 36x43x43 cm plenum from which the combustion and the eductor air streams are supplied. The combustion air is forced into the furnace plenum (between the two cylinders) through a 20 cm diameter duct. The air is preheated to 93-205°C in the furnace plenum by convective and radiant heat transfer. The undergrate primary combustion air is introduced through the adjustable gatevalve previously described. Air for secondary combustion is introduced via the furnace plenum into the secondary combustion zone through the tuyeres. The combustion airflow is regu- lated by a butterfly valve. The air for the venturi eductor nozzle is supplied by a 10 cm diameter duct connected to the fan plenum. The 10 cm duct is reduced to 8 tun diameter at the nozzle exit. The eductor airflow depends upon the combustion airflow rate. If the butterfly valve (regulating the combustion airflow) is fully open, less air is available for the eductor, and vice versa. 103 A 1.8 m diameter 24 GA sheet-metal shroud surrounds the top 1.3 m section of the furnace; it extends 1.2 m above the furnace to improve heat recovery and to facilitate coupling of the furnace to the in-bin counterflow drying bin. 3.2 Combustion Principle of the Concentric Vortex-Cell Furnace In the concentric-vortex furnace, combustion air is preheated between the two steel cylinders. The primary air enters the fuel pile at the bottom of the grate where gasification and pyrolysis of the fuel takes place. The producer gases travel upwards into the secondary combustion zone of the furnace. Secondary air from the furnace plenum enters into the upper secondary combustion zone through the tuyeres. The tuyeres cause a vortex and mixing of the flue gases. Unburned material or fly ash is centrifugally separated from the flame spiral. The particulates fall onto the top of the burning pile and are reignited. There are no tuyeres located directly at the top of the burning pile. The high velocity air from such tuyeres would agitate the pile and cause excessive unburned particles to accelerate upwards and escape through the chimney. The vacuum developed at the tap of the chimney by the eductor ensures that smoke and other volatile gases do not escape through the feed auger and other leaks. Thus, the vacuum prevents backfiring through the feed system and accelerates the flue gases from the combus- tion chamber into the chimney. The exhaust flue gases are blended with the ambient air in a 91.4 x 91.4 cm duct that connects the furnace shroud to the dryer fan 104 intake. The temperature of the blended air is between loo-150°C depending on the desired drying temperature. The drying system consists of an in-bin counterflow unit built into a 5.50 m diameter, 3.70 m high bin. It is equipped with a 9 kw (12.5 Hp) centrifugal fan. The drying system is described in section 2.3.3.5. 3.3 Furnace Operation The furnace is started by a propane torch and requires about 30 minutes before the desired operating temperature is reached. During the start-up period, a door in the heat duct between the furnace and the dryer fan intake is opened to allow the products of incomplete combus- tion to be discharged into the atmosphere. The dryer fan is turned on when the exit temperature of the furnace flue gas is about 540°C. The flue gases are blended directly with ambient air to obtain the desired drying air temperature. The fuel feed rate is regulated to maintain the proper drying temperature. The combustion air is kept at 120-170% of the stoichiometric air to avoid smoke and soot in the exhaust gas and to keep the exit temperature of the flue gas below 700°C. Metal parts in contact with the flue gas experience accelerated damage when the exhaust gas temperature exceeds 1095°C. 3.4 Concentric Vortex-Cell Furnace Design Theory Concentric vortex-cell furnaces have been built by trial and error (Claar et al., 1981). The theoretical basis for the design of CVCF and other biomass furnaces is lacking in the literature. In the 105 following section the theoretical considerations are discussed that will contribute to the design of CVCF. 3.4.1 The Design Heat Output The heat output of a furnace depends upon the environmental conditions, the furnace design, the grate area, the combustion chamber volume, and the efficiency of mixing the combustion air and the combustible materials in the furnace, and the inherent fuel character- istics (as defined by the particle size, the moisture content, and the fuel type). Claar et al. (1981) reported a heat output of 5.3 GJ/h per m2 of grate area, of 1.9 GH/h per m3 of combustion chamber volume. The heat output of the MSU (Michigan State University) CVCF is 2.5 GJ/h (grate area = 0.43 m2, combustion chamber volume - 1.3 m3); with 20% moisture content of wood, the required maximum fuel feed rate is approximately 160 kg/h (assuming a net heating value of 15.63 MJ/kg). 3.4.2 Grate Design The 1981 model of the MSU CVCF had a solid hearth similar to the hearth by Claar et al. (1980b). It was impossible to continuously separate the ash from the burning fuel on the solid hearth. The ash accumulated on the hearth, smothering the fire. Also, the increasing amount of the ash raised the bed-depth of the fuel pile, resulting in large amounts of fly ash in the flue gases. Complete combustion of char was impossible due to the inability of air to penetrate the mixture of unburned combustible materials and ash. This resulted in inefficiency due to unburned fuel and in a high resistance to the fresh fuel flow 106 onto the hearth. To overcome the above limitations, the grate was redesigned in order to ensure complete char combustion and continuous separation of the ash from the combustible materials. Junge (1979) suggested the following criteria for designing grate systems in spreader -stoker boilers: 1. the system must provide uniform distribution of the undergrate air; 2. the system must permit the removal of ash with minimum labor; 3. the system must be able to withstand high tempera- tures; 4. the system must be able to expand and contract with thermal cycling without buckling, warping, or opening large gaps; 5. the system must be designed for minimum maintenance. To satisfy the above requirements, a slightly sloping grate 3 ash receptacle. (with 12-degree slope) was placed above a 0.25 m The grate was made of 16 GA, 4-mesh screen. Firebricks were randomly spaced on the screen to protect it from high temperatures. The screen (with 6.35 mm holes) was selected to facilitate combustion of wood chips and corncobs as well as large grains such as corn and soybeans. During combustion the undergrate air cools the grate and provides oxygen required for complete char combustion. Combustible materials are held on the grate while the ashes fall through the grate. Combustion of char particles is completed under the grate. 107 The 12-degree slope facilitates uniform fuel distribution on the grate. Also, a combustion front from the lower end of the grate (and advancing towards the feed) is established. Thus, three regimes are established in the furnace: (1) char combustion on and under the grate, (2) new fuel drying on the grate, and (3) fuel pyrolysing and burning between the char and the new fuel. The undergrate airflow is regulated to maintain char combustion (to provide pyrolysis heat) without fluidizing the fuel bed. Fluidization of the fuel bed results in excessive fly ash development and possibly undesirable combustion of fuel in the feed hopper. 3.5 Fuel Feed System The designing of a fuel feed system consists of determining the feed auger size, maximum rpm and power requirement, and the feed auger speed control system. The screw conveyer (auger) designed consists of a standard 25 cm diameter auger rotating in a semicircular trough. The auger is driven by an electric motor thrdugh the speed reduction system described in section 3.1. The theoretical capacity of the auger was calculated using the following equation (Henderson and Perry, 1976): 108 m = 1.86 x 10‘5 (D2 - d2)Ppr (3.1) where: m = capacity, kg/h D = screw diameter, cm d = shaft diameter, cm p = pitch, cm 9b = material bulk density, kg/m3 R = auger speed, rpm Equation (3.1) assumes that the auger is 100% full of the con- veyed material. The volume of the auger occupied by the biomass material depends upon the particle size, flowability, abrasiveness, and packing characteristics of the fuel. The Link-Belt Company classi- fied bulk solids (including wood chips) for auger conveyor design (Perry and Chilton, 1973). Wood chips are described as irregular in size with a sluggish flowability (angle of repose more than 45 degrees), nonabrasive, very light and fluffy, and interlocked to resist digging. The average bulk density of wood chips is 160-481 kg/m3. Given these characteristics, wood can occupy only 45% of the available screw conveyor volume. Thus, a 25 cm (diameter) screw has a maximum capacity of 0.227 m3/h per rpm using equation (3.1) with a factor of 0.45. This translates to 91 kg/h per rpm if an average bulk density of 400 kg/m3 is assumed. The auger used on the MSU CVCF is designed for a minimum and maximum speed of 0.54 and 2.0 rpm, respectively. Thus, its minimum and maximum capacities are 49 and 181 kg/h, respectively. The maximum rate is slightly higer than the 160 kg/h required to give the maximum heat output of the furnace (section 3.3.1). 109 The power requirement for screw conveyors is the sum of the power necessary to drive the screw empty and that necessary to move the material. The first component is a function of conveyor length, speed of rotation, and friction in the conveyor bearings. The second term is a function of the total weight of the conveyed material per unit time, the conveyed length, and the depth to which the trough is loaded. In addition to the above components, extra power is required to overcome transient overloads due to extra friction on the system. The extra power is necessary when conveying irregularly sized materials such as wood chips and corncobs. The following equation was used to estimate the power required to convey the materials (Henderson and Perry, 1976): Pk = M L x 1.535 x 10‘4 (3.2) where PK is the conveying power, kw; M is the material flow rate kg/h; and L is the length, m. According to equation (3.2), the power required to convey 181 kg/h for 1 m is less than 0.03 kw. The power required to drive the screw empty and to overcome transient resistances is considerable. A 0.37 kw (1/2 HP) motor was selected. The fuel feed control system consists of two diaphragm switches (located at the top and bottom of the fuel hopper)and a magnetic relay switch (see Figure 3.2). The switches control the motor that drives the feed wagon discharge system. When the feed h0pper is empty, the 110 .gmaaoz pun; one :4 4ocacou Po>m4 4oz; Lo» mmsuuwzm cwm was umamcmmz mo Emgmm4o uwmemnum .~.m c.3m.. JIM ‘li 1|.“ ..4 111 contacts in the switches A and B (see Figure 3.2) are closed, and, therefore, the relay switch K is energized and closes contacts K1, K2, and K3 to energize the feed wagon motor M. Fuel is discharged into the fuel hopper. Upon reaching level 1 the fuel pushes against the diaphragm switch A and opens the contact. The current-flow through the relay is maintained through switch 8 at the top of the hopper and contact K1. Upon reaching level 2, the fuel pushes against switch B and opens its contact. Current-flow to the relay is interrupted and contacts K1, K2, and K3 are opened. The feed wagon motor stops. As the stoker-auger continues to move the fuel into the furnace, the fuel level falls below level 2, and contact 8 closes. Since contacts A and K remain Open, the relay is not energized. When the fuel level falls below level 1, the pressure on switch A is released and contact A is closed. The relay is energized and the cycle is repeated. Thus, the intermittent action of the feed wagon motor maintains the fuel between levels 1 and 2 in the fuel hopper. A vibrator (Model 20N manufactured by Eriez Magnetics) is attached to the side of the fuel hopper to prevent bridging of the fuel at the bottom of the hopper. The stoker-auger speed (and therefore, the fuel feed rate) is manually set to maintain the desired drying air temperature by adjust- ing the variable speed belt drive. If the fuel conditions do not change drastically during the combustion process, the fluctuation in the drying air temperature is limited to i 5°C. 112 3.5.1 Furnace Air Supply System The furnace air supply system in the MSU CVCF has three func- tions: (1) to supply oxygen for primary combustion (undergrate air) and secondary combustion (overfire air); (2) to provide high velocity air to the eductor for maintaining a vacuum in the combustion chamber for preventing backfiring (premature ignition of fuel) in the fuel hopper, especially during start-up: (3) to provide the vortex motion for centrifugal separation of fly ash and unburned particles from the flue gas. The following equation can be used to estimate the stoichio- metric combustion air (Perry and Chilton, 1973): Ma = 1.38(C/12 + H/4 - 02/32 + S/32) (3.3) where Ma is the combustion air (kg/kg fuel dry matter); and C, H, 02 and S are the percentages of carbon, hydrogen, oxygen, and sulfur, respectively. The total amount of combustion air (for both the primary and the secondary combustion) is determined from equation (3.3) using an average ultimate analysis of wood, and the maximum fuel feed rate (160 kg/h) previously determined. The average ultimate analysis of wood (Table 2.2) is 50.5% carbon, 6.2% hydrogen, 46.1% oxygen, 0.88% ash, and 0.0% sulfur. Using these values in equation (3.3), the stoichiometric air required is 6.1 kg/kg of fuel dry matter. This is equivalent to 4.9 kg of air per kg of wood at 20% moisture. Thus, the stoichiometric air required in the MSU CVCF is 4.9x160 = 784 kg/h at 113 the maximum fuel (wood) feed rate. Excess air is required for complete combustion due to the poor mixing of air and combustible materials in the furnace. Also, wet fuels require more air than dry fuels. For design purposes, 50% excess air was assumed. This gives the maximum combustion airflow rate of 1180 kg per hour. The minimum undergrate air may be determined if the ultimate analysis of char is known. (Table 2.2 contains the available ultimate analysis of char.) The stoichiometric air required for complete com- bustion of char is 10.5 kg/kg (using the average ultimate char analy- sis) or 2.1 kg/kg fuel dry matter since char constitutes about 20% of the products of pyrolysis (Admas, 1980). Thus, the undergrate airflow is approximately 4.2 kg/kg fuel dry matter with 50% excess air. The eductor airflow requirement may be computed by applying the Bernoulli equation for adiabatic, nonviscous flow to the eductor nozzle. The assumption of adiabatic flow is justified by the fact that the heat transfer per unit mass is small compared to the kinetic energy changes. The velocity in the nozzle is too large to allow enough time for the air to come to thermal equilibrium with the walls of the short eductor nozzle. The Bernoulli equation for compressible flow is (assuming no elevation differences): U2 k(P,/o ) U2 k(P2/o ) ;+__1_=—£+———2— (34) 2 ’ k-l 2 k-l ' Where U is the gas velocity (m/s); P is the static pressure (kPa); p is the density (kg/m3); and k is the ratio of specific heat capacities 114 (k = Cp/Cv = 1.4 for air). The subscripts 1 and 2 denote the cross section at the nozzle entrance and exits. respectively. The equation of mass continuity can be used to eliminate one of the unknown velocities, U1 and U2: Alulp1 = Azuzp2 (3.5) Algebraic rearrangement of the above equations along with the isentropic relationship Pl/pE == Pz/p: results in the following equation: 1/2 u2 = [2kRT1C1/C2] (3.6) . _ (k-1)/k where. C1 - (Pa/P1) - 2 Z/k c2 - [(42/41) “2”.) - 1101-1) Since k = 1.4 for air, equation (3.6) can be rewritten as: _ 1/2 u2 - 37.88[T1C1/C2] (3.7) Equation (3.7) is suitable for adiabatic isentropic flow through converging/diverging nozzles. A similar equation for straight ducts can be developed from equation (3.4): u2 = [2RT1C1k/(1-k)]1/2 (3.8) For air, k = 1.4 and R = 287, and u2 = 44.82[T1(-C1)]1/2 (3.9) 115 Equations (3.6) and (3.9) were used to design the axial fan for both the eductor nozzle and the combustion airflows. For the eductor to maintain a sufficient vacuum in the combustion chamber, a pressure of 0.25 kPa below the atmospheric pressure is required. An axial fan operating at approximately 0.5 kPa above atmospheric pres- sure was selected. The eductor nozzle is made of a 10 cm schedule No. 40 steel pipe connected to a 7.62 cm tapered reducer. The inside cross-sectional areas of the pipes are 0.0082 m2 and 0.00477 m2, respectively. At an operating temperature of 50°C, the nozzle exit velocity and the mass flow rate are 38 m/s and 715 kg/h. A 14-inch diameter, 1 HP axial flow fan was available with a capacity of 2,100 kg air per hour at 10°C. With 715 kg/h flowing through the eductor, approximately 1385 kg/h is available for combustion (some of the eductor air is used to complete the combustion of unburned volatiles that may be contained in the flue gases). Equation (3.9) was used to predict the airflow rate in the tuyeres. The tuyeres are constructed of schedule No. 40, 2 inch. nominal diameter pipes with a flow area of 0.0022 m2. The pressure in the furnace plenum is between 0.125 and 0.25 kPa above the combustion chamber pressure (slightly below the atmospheric pressure) depending upon the undergrate airflow rate, the drying airflow and the combustion chamber temperature. The maximum furnace plenum temperature is about 100°C. Thus, the airflow velocity and the total airflow in the tuyeres are approximately 20 m/s and 879 kg/h (the undergrate airflow rate under these conditions is approximately 500 kg/h). 116 The above analysis is used for estimating the fan size for supplying the eductor and combustion air when the furnace is operated without the drying system. During drying the drying fan imposes an additional vacuum on the furnace resulting in additional airflow supplied to the furnace. 3.5.1.1 Eductor Design The eductor consists of the eductor nozzle, a suction chamber, and a diffuser (see Figure 3.3). The eductor is a jet pump in which a high velocity fluid (the motive fluid) is forced through the nozzle. The momentum and the kinetic energy of the motive fluid are used to entrain and pump a second fluid stream, referred to as the induced fluid. At least three processes are involved (Davieset al., 1967): 1. acceleration of the induced fluid by the impact of the particles of the‘motive fluid from the nozzle; 2. entrainment of the induced fluid by viscous friction at the periphery of the nozzle jet; 3. overexpansion of the jet to a pressure below that of the induced fluid with consequent flow of the induced fluid towards the axis of the jet. The design of an eductor requires consideration of the temperature and pressure of the motive and the induced fluids as well as the eductor size and geometry. The diffuser is the most critical part of an inductor. The diffuser consists of the diffuser entrance, the diffuser throat, and the diffuser outlet or tail pipe (see Figure 3.3). According to Kroll 117 mapa 4 ..m h n.3,. umuauc. 40V u.=Pc w>..o= 4m. WHQFLUmnzm 4.43xv a... zap. ...: 4xv wczumgmgsmh u h 4m¢xv mgammmgm “moggp me=444o u m mucmcuzm gmmavw4o .gopusum mg» 4o Eugam4o u4umeacum on» II II (V) 3 1414er (4.6) b reac prod 127 and for control volume CV2: Q + M10 (T7-T1) a .- 2 MiAhi + AHR + Z M10111 (4.7) reac T prod T 2 s where Ahi = hi - hi T. T. J J 298 C = specific heat (kJ/kg°C). The mass flow rates and the ambient temperature are known in equations (4.2) to (4.7). The equations are solved by an iterative procedure that starts with an estimated temperature of the preheated combustion air T3. Consequently, the adiabatic flame temperature can be computed from Equation (4.3). The corresponding furnace tempera- ture Tl+ is obtained from equation (4.2). The eductor exit temperature is computed by applying mass and energy balances after mixing the flue gases and the eductor motive air at point 5. Thus, T: = ”“2555 I 1:384) <4-81 The temperature of T6 obtained from equation (4.8) is compared with the solution for T6 in equations (4.4) and (4.5). The iteration is repeated until a solution is obtained. This procedure results in approximate values for the combustion air preheated temperature T3. the furnace combustion chamber temperature T“, and the eductor exit temperature. The total heat loss cannot be predicted since equations 4.6 and 4.7 cannot be solved. 128 The final combustion chamber temperature, the preheated com- bustion air temperature, preheated diluting air temperature, the furnace flue gas temperature before and after dilution, the drying air temperature and the total heat losses can be computed from a heat transfer analysis as will be shown in the next section. 4.1.2 Heat Transfer from Combustion Chamber The heat losses from the combustion chamber occur in the form of radiative and convective heat transfer. 4.1.2.1 Radiative Heat Transfer The flame in the combustion chamber looses heat to the chamber walls by radiation and convection. In the model the radiative heat transfer is predicted using the following equation: 1, 11 QR = AwEFw-fflf ' Tw) (4'9) The flame emisSiVity is computed using the Adams (1980) equations for soot, carbon dioxide, and water vapor (equations 2.58-2.60). Since the flame is entirely enclosed within the furnace chamber, the view factor Ff—w is equal to one. 4.1.2.2 Convective Heat Losses The convective heat transfer from the flame to the combustion chamber walls is calculated by: Qe = Awhc(Te - Tw) (4-10) 129 The convective heat transfer coefficient hc is estimated from the empirical correlation for rough pipes (Holman, 1981): 2 StbPrf ’3 = f/8 (4.11) Stb = hC/(pCUm) (4.12) Equations (4.11) and (4.12) are rearranged to solve for the convective heat transfer: 11C = prUm/B (4 13) In equation (4.13) f is the friction factor. The friction factor depends upon the roughness of the surface and the Reynolds number, and varies between 0.008 and 0.1. In the MSU CVCF furnace the combustion chamber is lined with bricks and, therefore, it can be regarded as a rough pipe. The maximum value of 0.1 was selected for the friction factor of the chamber. 4.1.2.3 Total Heat Transfer from Combustion Chamber The net energy out of the furnace is the sum of the enthalpy change of the reactants (fuel and air) from the initial temperature to the flame temperature, and the net heat transfer from and to the cham- ber walls. The next heat transfer is given by: Qw = Aw[(ewa_fo(T; - 1;) + hc(Tf - Tw)] (4.14) 130 4.1.3 Radiative Heat Transfer Between Cylinders The furnace jacket consists of the annulus region (plenum) between the inner and the outer cylinders (marked D and E in Figure 4.2). Heat is transferred from the surfaces of the inner cylinder (6) to the outer surfaces of the cylinder (7) by radiation and convection. The surfaces (6) and (7) are considered concentric cylinders of infinite length. The radiation shape factors are (Howell, 1982): F5_‘7 = 1 (4.15) F7-5 = rG/r7 (4.16) F = 1 - F (4.17) 7-7 7.5 u u - 0 (T6- T7) Q6-7 - 1-66 1 1'87 (4°18) + + A666 A6F6_7 A76, The radiation thermal resistance between surfaces 6 and 7 is: 1 - as + 1 + 1 - e7 MFG”7 A787 (4.19) 1s - A566 Thus, equation (4.18) can be written as: o (T“ - T“) Q6 7 = “R 7 (4.20) 16 131 Figure 4.2. Simplified Furnace Diagram. The letters refer to spaces and the numbers to surfaces. 132 Figure 4.3 represents the radiation, the conduction, and the convection heat transfer network for the plenum area between the inner and the outer cylinders, and the furnace combustion chamber. Heat balances on the network of Figure 4.3 result in the following set of simultaneous equations: on surface 7: T'Eb (%—+%—)-Qc -QC -—=o (1.21) On surface 6: 1.4. _ 11..- a... R R 18 16 - Q = o (4.22) C6 0n the furnace wall W: Ebf Ebw Tf Tw Tw TE R C R 19 f » 18 The radiative heat transfer between the surfaces 1, 2, 3, and 4 is complex since all the surfaces radiate and absorb heat to and from each other. Surfaces 1, 2, 3, and 4 are formed by two coaxial cylinders (of inner and outer cylinders 2 and 4, respectively) and the two annular ends, 1 and 3. The radiation shape factors between vari- ous surfaces are calculated using the governing equations for the relevant shapes (C-86, C—87, and C-89) as presented by Howell (1982). \ 133 .448; mumcgs4 mg“ on msmpw as» 2014 mocmum4mmg «>4uum>cou "mu new "Ezcmpa ms“ :4 14m umpmmgmga an» op .446>wuomnmmc .cmuc4pxu gmuao mzu use gmucwpao smcc4 mg» 5014 mommo4 paw; m>4uum>cou "44uo 8 ..uo upcmwnsm one op 44m: Lmuzo 5°14 mucmpmwmmg m>4¢um>coo . 0 "ppm; Luggage :o4umsnsou mg» on damp. mg» 5014 mucmpmpmmg m>4u84umg .mHm 444m: mu4mc4 an» no oc4cw4 3041a ms» sonoggp mucmum4mmg co4uoancoo.:m 444v gmuc444u gmuao mg» on Amy 44m: gmvcmpau gmcc4 ms» 5°14 mucmpm4mwg m>4um4u~gmflm mucm4asm mg» op 44V44m2 . mumcgam Lmuao 2°14 museum4mm1 «>4um4umg ..m .34 asaumgmaeou um ¢=m4n5m o» 4» manumgmasmp um «omega; mgu sag» xgozwmz .gmmmcmgh paw: :o4uum>=ouuco4um4umm 9..“ 1.. o H“ flFUO m4“ m4“ 4.4m .m.¢ «13m.. 134 The radiative heat transfer among the surfaces 1, 2, 3, and 4 can be simplified by analyzing the radiation network shown in Figure 4.4. The radiation resistance R1 to R10 are defined in Figure 4.3. The net- work in Figure 4.4 is analyzed by the application of Kirchhoff's cur- rent law which states that the sum of the energy flows into a node is zero. The Ebi terms in Figure 4.3 represent the product ng, the radiocity from a node i (the total radiation which leaves a surface per unit time and per unit area). By applying the Kirchhoff law to the network in Figure 4.4, a set of four simultaneous equations is obtained: ELL-fiZ-fJ“ +J3_J (L+_1_+_L_+_1__)=0 (424) R R R R 1 R1 R2 R1, R10 ' 1 2 10 "1 %Z+J.+i1.i1-, 1 1 1.1)... (425) R3 R2 R9 R 2 R2 R R R9 . J J J ii+—1—+—2-+-3-J —1—+—1—+i-+-l-)=0 (4.26) R7 R10 R8 R6 11 R6 R7 R8 R10 E J J J _L 4 4 4- Jqlqlutho Mfl) R + R. + R9 + R5 J3 (R. R5 R5 R9) A fifth equation is obtained by writing an energy balance for surface 4 that includes the convective heat loss from the surface: ____£.1. .. - - R7 QCu-o QCu-i (4 28) 135 R 1 J1 R2 J2 R3 Ebl Eb2 R“ Re 3 E b3 Rs 33 Rs J1 R7 b“ R = 1'5 R a 1 R = 1’6 R = 1 1 A c A F A e A F 1 1 1 1'2 2 2 1 1'3 -e R5 8 1 3 R 8 1 R 3 1 R 8 1 A e 5 A F 7 A 9 A r 3 3 3 3‘4 3 3“» 2 2-1 R =- 1 R I 1 9 A F 1° A F 2 2-3 1 1-1 Figure 4.4. Complete Radiation Network for the System 1, 2, 3, and 4 in Figure 4.2. 136 4.2 Heat Transfer Between the Furnace and the Hood The surface marked 9 in Figure 4.2 is a reflective surface. Its emissivity and absorptivity are approximately zero. Thus, the thermal radiation resistance is high. Therefore, the four-surface system (3, 4, 8, and 9) reduces to a three-surface system consisting of surfaces 3, 5, and 8. The radiation network for the reduced sys- tem and radiation resistances R11 to R1 are presented in Figure 4.5. S The Kirchhoff analysis of the network in Figure 4.5 results in three simultaneous equations: __S+__?_+_i.-J L+;+-—1—- =0 4.29 R5 R13 R11 5 (R5 R11 R13) ( ) E J J bs 5 7 1 1 1 __+.__+ J -——-+—+— =0 4.30 12 R11 le 6 (R11 R12 R14) ( ) E J J be 5 5 1 1 1 _. ....-.) __+—_+— =0 4.31 R1s -§13 R11 7 (R13 R11 R15) ( ) The energy balance on surface 3 (including radiation and convective heat gains and losses on both sides of the surface) gives: Ja-Ebs-EbJL-J _Q R5 R5 C3‘0 - Q = 0 (4.32) Equation (4.33) is obtained from an energy balance on the top plate 5 (including convection from the top of the plate to the ambient and to the space between plate 5 and 3, and the radiation from the plate to the sky): 137 an o nm on w o 41mg“ w14 1 an mlmm mlmm m4 :4 m4 2 H u m H H m ma «4 m 4. m m an m m h. m 4... m m .N.¢ ogszR :4 m use .m .m mmuw$c=m as» no» xgozumz :o4um4umm .m.3 3.3m.1 44 138 J - E E - E 5 b5 b5 b3 _ R12 Ra ' ”WATS-Tm) - 1151-0545141— 0 (4.33) Equations (4.21) to (4.33) consist of a set of thirteen simul- taneous equations with 13 unknowns. A computer program was written to solve the set of the equations after estimating initial values of T2, T7, and the temperature of the preheated combustion air Td as functions of the combustion chamber temperature. The estimates are based on sta- tistical analysis of the experimental temperature data. The following are the equations used for the initial guesses of the temperatures: T, = 0.691f - 24.10 (4.34) T, = 0.241f - 139.80 (4.35) Td = 9.181f - 83.00 (4.36) Equations 4.1 to 4.8 are used for the initial guesses of the preheated air temperature, the furnace adiabatic temperature, and the actual flame temperature. The convective resistances and heat transfer rates are calcu- lated using standard procedures for the relevant flow regimes, shapes, and temperatures as presented in the heat transfer literature (Holman, 1981; Gebhart, 1971). The convective heat transfer coefficient on the outer surfaces is a function of the ambient temperature. The influence of wind and precipitation are not included in this analysis. Ixcomputer program was developed and used to analyze the furnace performance in terms of the heat losses and heat output at different 139 fuel feed rates, and at different fuel and ambient conditions. The inputs to the model include: a. b. C. The fuel type. The fuel moisture content (% wet basis). The fuel feed rate (kg/h). The ambient temperature (°C) and the relative humidity (percent). The drying bin plenum pressure (kPa) (the pressure is employed to calculate the drying airflow rate). The program solves for the maximum flame temperature and for the wail temperatures contained in equations (4.21) to (4.33). The output of the program includes (see Appendix A): 1. The a. b. C. temperatures in and around the furnace: The preheated combustion air temperature. The furnace combustion chamber (flame) temperature. The exit flue gas temperature. The diluted air temperature (immediately after mixing flue gases with ambient air). The drying air temperature and absolute humidity. mass flow rates in the system (kg/h): The fuel feed rate. The combustion air. The drying air. The ash accumulation (and fuel accumulation if the combustion air is less than stoichiometric). 140 3. The total thermal energy input (MJ/h). 4. The total heat loss from the system (Md/h). 5. The furnace and the system efficiencies (percent). 4.3 Grain Drying Simulation The in-bin intermittent counterflow drying process is simulated using the MSU fixed bed simulation model (Bakker-Arkema et al., 1974), equations (2.72) to (2.75). The equations are solved by finite dif- ference techniques. The Troeger and Hukill (1970) thin-layer equa— tion is used for drying air temperatures less than 71°C, and the Thompson et al. (1968) equation (2.78) for grain temperature above 71°C. The DeBoer equations (Bakker-Arkema et al., 1974) are used to esti- mate the equilibrium moisture content required in the thin-layer equa- tions. The intermittent in-bin counterflow dryer is similar to a fixed bed in-bin dryer. The grain is held in the bin until the bottom layer (7.5-10 cm deep) has dried to a predetermined moisture content. The dry grain is then removed and the process continues until the bin is empty (if grain is not added to the bin). This process is simulated by an in-bin fixed bed drying model which reduces the bed depth instan- taneously by the equivalent depth of the grain removed and increases the airflow rate accordingly. The grain conditions at various levels above the false floor are maintained and shifted downwards as drying progresses. During intermittent counterflow drying, the drying airflow increases with a decrease in grain depth and decreases each time some 141 grain is added to the bin. To simulate these changes, the character- istic fan equation, which predicts the airflow delivered by the centri- fugal dryer fan at a given static pressure, and the resistance of the grain to airflow should be known. The characteristic fan equation (4.37) was supplied by the manufacturer (Parkes, 1982): v = a - 8P° (4.37) where Y = volumetric airflow rate (m3/min) P = pressure (kPa) a = constant (368.81 m3/m1n) B = 80.87 0 < P < 1.9 8 = 56.23 1.9 5_p ( 2.4 C = 1.17 0 < P < 1.9 c = 1.95 1.9 E P < 2.4 The pressure drop due to the grain (corn) airflow resistance is predicted by equation (4.38) (Hukill and Shedd, 1955): _ 7 x 10'3 dTy P ’ 1n(1 - 0.512y) (4.38) where y is the airflow rate (m3/min - m2), and D is the corn depth (m). An additional constant 1 was introduced in the original Hukill and Shedd equation to account for fines, corn kernel size, parking of the corn in the bin, and foreign materials. The value of the constant T is found by solving equation (4.38) using experimental values of pres- sure (P) and grain depth (D). The value of T for the corn at the Kalchik Farms was found to vary between 1.3 and 1.75. 142 Equation (4.38) and (4.37) may be combined to a single equation (4.39): 3 . l/c ' A — - 7310 D72 _._ 4 .39 (Jr—2) m(1—o.512y) O ( ) Thus, for a given grain depth there exists y > 0 which satisfies equation (4.39). The value of y is found by the secant iterative method for finding the roots of “an equation (Conte and de Boor, 1980). The pressure drop is computed by equation (4.38), using the y value that satisfied equation (4.39). The furnace thermal analysis model and the in-bin counterflow grain drying models are included in Appendices C and D, respectively. CHAPTER 5 EXPERIMENTAL INVESTIGATIONS 5.1 Furnace/Dryer Performance The MSU concentric vortex-cell furnace (CNCF) was tested in conjunction with an in-bin counterflow dryer. The fuels used were wood chips at 14, 19, and 45 percent moisture content, and corncobs at 10 and 12%. Shelled corn was dried from different initial moisture contents (22-27%) to approximately 17-18% and dumped hot in a cooling bin in which it was held for 6 to 12 hours before cooling at a low airflow rate (0.5-1.0 m3/min-m2). Moisture contents after cooling varied from 14 to 16%. The airflow in the drying bin varied from 11 to 13 min3/min- m2; the drying air temperature from 65 to 94°C. The airflow in the cooling bin varied from 0:5 to 1 m3/min-m2; the cooling air temperature from 0 to 13°C. During each test the following temperatures were measured and recorded: 1. The temperature in the furnace plenum and above the grate (after primary combustion). 2. The temperature between the two sets of tuyeres. 3. The temperature above the top set of tuyeres. 4. The temperature at the chimney exit. 143 144 5. The temperature at the dryer fan-intake after diluting the flue gases with ambient air. 6. The temperature in the dryer plenum. In addition, wet and dry bulb temperatures of the exhaust air after drying were measured along with ambient wet and dry bulb temperatures. The furnace temperatures were recorded using Type K (chromel- alumel) thermocouples. All other temperatures were measured using Type T (copper-constantan) thermocouples. The limits of error are i 0.75% and i 0.83°C for Types K and T, respectively. The temperatures were recorded on a Kaye Digistrip II recorder at one minute intervals during start-up, and every five minutes there- after. The furnace airflows were measured using Alnor-Velometer series 6000-P airflow meter (error limit 1 0.5 m/min). The blended drying airflows were calculated from the characteristic fan curve after measuring the plenum pressure. The fuel feed rate was determined by recording the time required to burn a known weight of fuel. The corn and the fuel mois- ture contents were determined using the standard oven method (ASAE 1982). 5.2 Chemical Analysis The corn dried with the CVCF/IBCF system was analyzed for polycyclic aromatic hydrocarbon (PAH) and heavy metal contamination. The corn samples for the contamination analyses were taken at the dryer inlet and during each cycle as the grain was removed from the 145 bottom of the IBCF dryer and discharged into the cooling bin. The cycle time was approximately one hour. The initial corn depth varied from 1.3 to 2 m. The residence time of the corn in the dryer varied from two to ten hours. The corn that was transferred to the cooling bin in the first cycle had the shortest residence time, the corn transferred in the last cycle had the longest. A propane test was conducted in a batch column dryer for com- parison. 5.2.1 Determination of Corn Contamination with PAH The contamination of corn with PAHs was analyzed using a high performance liquid chromatography (HPLC). The chemicals and the reagents used for the analysis are described below. Hexane and methylene chloride of analytical reagent grade were used along with acetonitrile distilled in glass of U.V. grade. Water was distilled, deionized, purified of organics, and filtered through a 0.22u pore size membrane. Polynuclear aromatic hydrocarbon reference standards were obtained from Supelco, Inc. (Bellefonte, PA). Silica gel (SilicAR cc-7 special) was employed for the column chromatography. 5.2.1.1 Extraction and Purification of Samples Samples were extracted with methylene chloride in a soxhlet extractor and purified by silica gel chromatography using procedures described by Jacko et al. (1982). Shelled corn (20 g) was extracted with 125 m2 methylene chloride for six hours, with the temperature adjusted to give a cycle time of six minutes in the soxhlet. The 146 methylene chloride was evaporated on a rotary evaporator, and the sample redissolved in Smt of hexane. A 1.2 x 28 cm column containing 4 g silica gel was topped with 2 cm of anhydrous Na2504 and equi- librated with 25 m2 haxane. The corn sample was rinsed into the column with 5 m2 of hexane and the column eluted with 25 m2 of hexane which was discarded. Subsequently, the column was extracted with two 25 m2 fractions of CHZCl2 and hexane (40:60) which were combined and evaporated to near dryness on a rotary evaporator. The sample was transferred to a test tube with CHZClZ and evaporated to dryness in a stream of N2. The samples were redissolved in 2 m2 of acetonitrile and filtered through a 0.4u pore size membrane prior to analysis by high performance liquid chromatography (HPLC). 5.2.1.2 High Performance Liquid Chromatography The HPLC analysis was carried out as suggested by Ogan et al. (1979). The HPLC equipment consisted of a Waters Associates (Waltham, MA) Model 680 automated gradient controller, a Haters H6000 and M45 solvent delivery system with U6K injector, a Perkin-Elmer LS-4 fluorescence detector and a Hewlett-Packard 3390-A integrator recorder. The column was a Beckman Ultrasphere 3u-ODS (reverse phase), 4.6 mm 1.0., and 7.5 cm long. The solvents employed were pure acetonitrile (Solvent A) and acetonitrile/water mixture at a 20:80 ratio (Solvent B). The chromatography was performed at a flow rate of 1.5 mzlmin. The samples were eluted and the column re-equilibrated under the following program SEQUEHCE: 147 0-20 min: 40% to 100% Solvent A 20-30 min: 100% Solvent A 30-35 min: 100% to 20% Solvent A 35-40 min: 20% to 40% Solvent A 40-55 min: 40% Solent A, isocratic Polycyclic aromatic hydrocarbons were detected at excitation and emission wavelengths optimal for each compound as determined by preliminary scans of the reference materials on the LS-4 detector. The entrance and the exit slit widths were 10 nm each. Table 5.1 contains the wavelengths for the fluorescence detector used to deter- mine the PAHs, along with the expected retention time and the detec- tion limit of each compound. Actual elution times varied slightly from day to day, but were determined by establishing a reference mix before each unknown. Table 5.1.--Excitation/emission wavelengths, retention times, and detection limits of different PAHs on the fluorescence detector. Eggggggggn/ Retention Detection Compound wavelength time limit (nm) (min) (ppb) Naphthalene (Nph) 261/328 4.4 1.25 Fluorine (F1) 257/305 8.6 0.12 Anthracene (An) 251/400 10.8 0.06 Fluoranthene (Ft) 235/388 13.6 0.06 Benz [a] anthracene (BaA) 281/386 17.6 0.06 Benzo [a] pyrene (BaP) 305/410 20.0 0.06 148 5.1.2 Determination of Corn Contamination with Heavy Metals The concentration of heavy metal on the biomass and the propane dried corn was determined using inductively coupled argon plasma atomic emission spectroscopy (ICP-AES). Intra-analyzed, concentrated nitric acid was employed along with yttrium of analytical grade. Water was the same as described in section 5.2.1 One-gram corn samples were wet-ashed with 5 m2 of concentrated nitric acid in 30 mt containers and kept for 12 hours in a low tem- perature (75-80°C) oven. The samples were transferred to 25 m2 volumetric bottles spiked with 10 ppm yttrium which acted as an internal standard. The samples were diluted to a volume of 25 mt with distilled water. The plasma spectrometer (ICPaAES) employed for heavy metal analysis was a Model 955 Atom Comp manufactured by Jarrel-Ash Division (Naltham, MA). The results were printed directly in ppm. The follow- ing were the conditions during each analysis: Coolant (Argon) flowrate 18.0 t/min Sample carrier flowrate 0.5 £/min Sample flowrate 1.0 t/min Nebulizer pressure 1.8 bars Forward power 1 . 1 kw Refelected power < 5 N The detection limits for the heavy metals were the following: arsenic 3.0 ppm, cadmium 0.3 ppm. chromium 0.6 ppm, mercury 6.0 ppm, selenium 15.0 ppm, lead 3.0 ppm, and thallium 15.0 ppm. CHAPTER 6 RESULTS AND DISCUSSION 6.1 Introduction The primary objectives of this study were to develop and incorporate a concentric vortex-cell biomass furnace (CVCF) in an existing on-farm grain drying system, and to test and analyze the furnace/dryer system in order to determine its economic feasibility and safety aspects. In this chapter the experimental and the simulated results will be compared. The simulation models are used to predict the fuel requirement and the corn drying rate using different fuels under different ambient conditions. An economic analysis is made for average conditions. A corn contamination analysis is presented to assess the health hazards associated with biomass fuels. 6.2 Fuel Consumption Table 6.1 contains the experimental and the simulated fuel consumption rates for different ambient and drying conditions. The table also contains the minimum theoretical fuel consumption rate required in the case of zero heat losses from the furnace/drying system. 149 150 mnouccou u us mawcu.uooz u 3* .mmop paw; ogo~ an co_pasam=ou Push pmurumgomgp u E:E_:_z~ .m xwucmaa< momH mHH mmH mma o.m¢ u 3 enmfl mummfi mm mm m.o 5 mm mm co” H.oH - u Hmmfi mummfi mm om m.m~ m mm ad“ ONH m.mH . 3 mfimfi mmNmH mm mm m.o . m mm efia ooH ~.N~ . u omefi mmmnfi mm on m.m e um moH mag “.ma . o emefi mommfi mm mm o.~ m mm mm oHH m.¢~ - 3 mmaa omemfi mm oofi m.m~ N mm om mm m.¢~ . 3 MHNH Hmnwfi an cw ¢.NH H Nazspcwz umuupasmm Papcmsvcmaxu My »MHmuu zodflmu< ammp hm game gxmx amazu “wmwwwwmmwm>m «copuvucou umuh mum; vow» Foam Pmam Pmpcmswcmnxu p=m_asm mmucm>< ; .mmcaumcmaswu awn mcwagu pcmcmCCVu can mcopuvvcou acmpasm ucmcmwmwv um mumccaw mmmEowa uu>o use we moan; copua22mcoo Foam umumpzswm can Faucospgmaxo mgh--.~.u mpam» 151 The fuel consumption rate depends on the heat required to raise the drying air from ambient to the drying temperature, on the heat losses from the system, on the fuel heating value, and on the fuel moisture content. The fuel consumption varies slightly with the combustion air/fuel ratio (and therefore with the excess air). An in-bin counterflow drying system (IBCF) was operated at a 2 m bed depth, an airflow rate of 300 kg/min, a drying temperature of 77°C, and at 12.4°C ambient temperature. This requires a minimum energy input of 1213 MJ/h (without considering heat losses). The energy input is equivalent to approximately 78 kg/h of wood chips or 74 kg/h of corncobs with 14.6 and 12.1 percent moisture (wet basis), respectively, if the average heat values of these fuels are used (20.97 and 18.65 MJ/kg, respectively). The minimum theoretical fuel feed rate for other tests is similarly computed and is presented in Table 6.1. The actual fuel feed rate is the sum of the theoretical minimum feed rate and the feed rate required to offset the energy losses from the system. The experimental and the simulated fuel feed rates are included in Table 6.1. Table 6.2 contains the experimental and the simulated heat losses from the furnace/dryer system for the tests and ambient condi- tions presented in Table 6.1. The experimental heat losses are based upon the experimental furnace temperatures (of furnace surfaces and of the diluted air leav- ing the furnace), the drying air temperature in the drying bin plenum, 152 .m xFucmaa< mumH um um new on mumfl ean Nmmofi onon 5 H5 mm mnv oem mNNH How“ nmemfi mummfi m mm mm mmv eon mneH mHmH cums” ommmfi m mm mm wma cum mmmfi omea ammofi ammua a mu mm «me emu mmmH ewe“ Hmmug mama“ m mm mm mme mmm Hmma woe" aoomfi enema N an em mum wme «Hag mHNH monfi ”mum” H smumsm auacaza umuapaswm Pancasvcaaxm umpapzewm Paucas_cmaxm umuapas_m Papcmswzmaxm .oz xucowuwmwwcwmpmpaewm mmo— amumuo psych zucmumwucPAgo zoFCCVu muflmuu mmmcm>< amok .~.o mpamp cw cm>vm m-“ mummp go$ mm_ucmqummm smumxm \mumccsm new mommop xmgmcm Page» .Amsmcm mcwzgu .3opmgpm umumpaswm ucm paucwsvcmaxmua.m.m mpnmh 153 the ambient temperature, and the drying air mass flow rate (kg/h). The average experimental heat losses are calculated as follows (using test No. 1 results for illustration): Average experimental temperatures (Test 1): Top of furnace shroud 39°C Furnace side wall 91°C Diluted air at furnace exit 99°C Drying air (bin plenum) 77°C Ambient 12°C Sky temperature -7°C The sky temperature is estimated by an equation presented by Duffie and Beckman (1980): _ 1.5 TS - 0.0552Ta (6.1) where TS and Ta are the sky and the ambient temperatures (K), respec- tively. 154 1. Furnace radiation loss: 1 (a) From top of shroud (radiation): 4 S) er(T: - T (6.2) Qrt 5.66 x 10"8 x 2.6(3124 - 2664) 656 w (2,362 kJ/h) 1 (b) Wall radiation: 4 T4) 0.5A [F e (Tw - a rw ' w w-g w-g O I 4 4 + Fw_sew_s(Tw - 75)] (6.3) 3.3[(3124 - 2854) + (3124 - 2664)] 1,386 N (4,991 kJ/h) Total radiation loss - Qrt + Qrw = 7,353 kJ/h 2. Furnace convective heat loss: From top of shroud (convection): Tfilm = (39 + 12.4)/2 = 25.7 0 (299K) - 3 2 Gr - gB(Tw — Ta)L /v (6.4) = 1.37 x 1010 Ra - PrGr = 0.708 x 1.37 x 1010 = 9.7 x 109 nt = (0.58k/0)Ra'2 (Eq. 7.25 Holman, 1981) (6.5) 0.026 x 0.58 x (9.7 x 109)'2/1.65 4.6 w/m2 - °c 155 QCt = 4.6 x 2.6 x (39-12.4) = 318 u (1,145 kJ/h) Total furnace loss - Qrt + Qrw + QCt = 8,498 kJ/h The convective heat loss from the furnace wall is recovered through the shroud. 3. Duct/bin plenum heat loss: The heat loss from the duct connecting the furnace to the bin plenum is computed by the product of the temperature difference between the diluted air temperature at the furnace exit and in the dryer plenum, the air mass flow rate (kg/h), and the specific heat of air. Thus: oduct MaCaAT (6.6) 18,781 x (99 - 77) 413,182 kJ/h where Ma is the drying air mass flow rate (kg/h); Ca is the specific heat of air (kJ/kg°C); and.8Tis the difference between the drying air temperature at the furnace (after the flue gases are diluted with the ambient air) and in the drying bin plenum (°C). The total system heat loss is equal to 421,680 kJ/h, the sum of the total loss from the furnace (8,498 kJ/h) and the combined loss from the duct and the bin plenum (413,182 kJ/h). The total energy required is 1,635 MJ/h, being the sum of the energy required to raise the air from the ambient to the drying tempera- ture (1,213 MJ/h) and the heat losses (421.68 MJ/h). The furnace/dryer system efficiency is therefore 74.2% (100 x 1213/1635) at these 156 conditions. Using an average heating value of 20,965 kJ/kg of wood, the required fuel feed rate is 78 kg/h of wood chips dry matter, which is equivalent to 92 kg/h of wet wood chips at 14.6% moisture (wet basis). The experimental and simulated fuel feed rates (at 14.6% moisture) are 85 and 80 kg/h, respectively (see Table 6.1). In Table 6.2 the simulated and experimental heat loss values are compared. The simulated heat loss is computed using the procedure outlined in Chapter 4. The experimental and the simulated heat losses are similar. The experimental loss is greater than the simulated heat loss due to the following reasons: 1. The experimental flow rates of the air mass are higher than the simulated flow rates. The experimental volumetric airflow rates are determined from the characteristic fan curve using the experimental pressure data taken in the drying bin plenum. The volumetric flow rate is converted to mass flow rate by dividing the former by the specific volume of air as a function of the dry bulb temperature. However, the temperature in the bin plenum may not have been a true representa- tive of the average temperature in the bin (Silva (1980). Empirical equations were developed to compute the total airflow through the furnace and into the dryer bin plenum as a function of the bin plenum pressure and the fuel feed rate. The equations are based on linear regression analysis of the experimental pressure and air flow data. The relation between Pf, the furnace fan pressure, and Pb’ the pres- sure in the bin plenum in kPa is: Pf = 0.3076Pb + 0.3058 (6 7) 157 The furnace volumetric air flow (m3/min) was determined from the characteristic fan curve using the pressure Pf obtained from equa— tion (6.7). The mass flow rate (kg/h) is determined by dividing the volumetric flow rate by the specific volume at ambient temperature. The equation for determinine the velocity of air blown around the furnace by the dryer fan is: = 146.28 - 13.08P (6.8) Ff b where Ff is the air flow velocity around the furnace (m/min). The mass flow rate of the air around the furnace is determined by multi- plying the velocity by the cross-sectional area of the annular space between the shroud and the furnace, and dividing the product (m3/min) by the specific volume of air at ambient temperature. The total gas flow out of the furnace is the sum of the mass flow rate of the air around the furnace and through the eductor, and the flue gas from the combustion process, The mass flow rate of the flue gas is the sum of the combustion air mass and the fuel rate, less the ash fraction in the fuel. The predicted drying mass air flow is equal to the total gas flow out of the furnace calculated by the above procedure. The total air mass flow computed by the above method is slightly less than the experimental values (as shown in Table 6.2). This is due to leakage of air into the dryer fan through the duct section between the furnace and the dryer fan. The empirical equations do not account for the air leakage of the system. 158 2. The experimental temperature difference between the diluted air after the furnace and in the bin plenum may be in error due to the location of the thermocouples. If the experimental mass airflow rate and heat losses for Test No. 1 are used, the expected fuel feed rate is 92 kg/h. The experi- mental fuel feed rate is 85 kg/h, compared to the simulated fuel feed rate of 80 kg/h. ‘ Thus, for Test No. 1, the experimental method of determining the heat loss overestimates the actual heat loss by about 8% and the simulation method underestimates the heat loss by approxi- mately 6%. 6.2.1 Factors Affecting the Fuel Feed Rate The fuel feed rate depends upon the following factors: 1. Heat loss from the furnace/dryer system as affected by the ambient conditions (temperature, wind veloc- ity, and precipitation); 2. the fuel moisture content; 3. the required drying air temperature and air mass flow rate. 6.2.1.1 Effect of Ambient Temperature on Fuel Feed Rate The simulated fuel feed rate, heat loss and the furnace/system efficiency at varying ambient temperature and at a constant drying air temperature of 80°C are presented in Table 6-3. The furnace efficiency is defined as the ratio of the total useful energy out of the furnace to the total thermal energy input into the furnace. It is calculated by the equations: 159 Table 6.3.--Effect of ambient air conditions on fuel consumption, furnace and system efficiency, and total heat loss from CVCF/IBCF drying s stem; drying air tem = 80°C, drying airflow = 12.3 /m2-min. Ambient conditions Fuel Heat Furnace System Temperature Absolute consumption loss efficiency efficiency °C humidity* kg/h MJ/h % kg/kg 0 0.0029 115 482 95.9 75.0 2 0.0033 110 462 95.8 74.9 4 0.0038 106 447 95.6 74.9 6 0.0043 102 432 95.4 74.7 8 0.0050 98 417 95.3 74.6 10 0.0057 94 403 95.1 74.5 12 0.0065 90 388 94.9 74.3 14 0.0075 87 377 94.7 74.2 . *Ambient air absolute humidity at 75% relative humidity. 160 01.n = anu - MW) (6.9) n, = (1 - QL/Ain)100 (6.10) where Qin is the thermal energy input to the furnace (kJ/h); QL is the heat loss from the furnace (kJ/h); H is the fuel higher heating value (kJ/kg dry matter); MN is the fuel moisture content (decimal); and nf is the efficiency (percent). The system efficiency is defined as the ratio of the net energy available for drying divided by the total thermal energy input to the system: ns = (1 - 05/01.")100 (6.11) where DS is the total heat loss from the system (kJ/h). The total heat loss includes the heat loss from the furnace, from the duct work link- ing the furnace to the bin, and from the wall of the drying plenum of the drying bin. The heat loss from the furnace/dryer system increases linearly with a decrease of ambient temperature. A 1°C decrease in ambient temperature results in a 7.5 MJ/h increase in the heat loss. The fuel feed rate required t‘omaintain the same drying air temperature increases by 2 kg/h (equivalent to 33.5 MJ/h) with the decrease of 1°C in ambient air temperature. Therefore, the increase in energy input (in the form of fuel) is higher than the increase in the heat loss. The increase in the fuel consumption is mainly due to the energy required in heating the drying air from a lower ambient 161 temperature. Additional energy is utilized in preheating the combus- tion air and the fuel. The energy required to counterbalance the heat loss is approximately 22% of the total fuel rate consumption increase. According to the simulated results in Table 6.3, the heat loss varies from 377 MJ/h to 482 MJ/h when the ambient temperature decreases from 14 to 0°C. According to Silva (1980), the heat loss from the IBCF (when propane gas was employed) was 351 MJ/h at an ambient and a drying temperature of 10 and 71°C, respectively: the system efficiency at these conditions is 71%. The simulated heat loss from the entire CVCF/ IBCF system is 403 MJ/h at an ambient and a drying temperature of 10 and 80°C, respectively, with a system efficiency of 75% (Table 6.3). The heat loss in a biomass fueled system is slightly higher than in a propane fueled system due to the extra duct linking the furnace to the fan. There is no data available to compare the heat lossfrom a propane fueled system with the biomass fueled system for the full range of ambient temperatures (0 to 14°C) included in Table 6.2. According to Silva (1980), the heat loss from the propane fueled IBCF is equivalent to approximately 5.6 litres/h of liquid propane (at the ambient and drying temperatures of 10 and 71°C, respectively). The heat losses in Table 6.3 are equivalent to 22-25 kg kg of wood per hour and 30-32 kg of corncobs per hour at an ambient temperature of 14 and 0°C, respectively, when the fuel moisture con- tent is 20%. In monetary terms the heat losses are equivalent to 162 approximately $1.00/h and $1.26/h when using biomass and liquid pro- pane fuels, respectively. The furnace and the system efficiencies are nearly constant for the 0 - 14°C temperature range presented in Table 6.3. The furnace efficiency is 95.9% at 0°C and 95.7% at 14°C. The system efficiency varies from 75% at 0°C to 74.2% at 14°C. These small changes are due to the fact that the drying temperature is constant, and therefore, the heat loss increase is compensated by an increase in fuel feed rate. The furnace efficiency is approximately 95% for the conditions in Table 6.3. Thus, the furnace heat loss is only 5% of the total heat loss. 6.2.2 Effect of Fuel Moisture on Fuel Consumption The fuel moisture greatly influences the average performance of the CVCF and the required fuel consumption rate at a specific heat output. The moisture in the fuel reduces the net heating value of the fuel and the flame temperature, and therefore results in a high fuel consumption. Incomplete conbustion is: an additional undesirable effect of low flame temperature. Table 6.4 lists the simulated heat loss, flame temperature, fuel consumption rate, and furnace/system efficiencies, when employing wood chips at different fuel moisture contents and different drying air temperatures. The simulations are based on an average airflow rate of 13 m3/min-m2, and an ambient temperature of 4°C. The results are also plotted in Figure 6.1. 163 om mm cam Neofi -- NH“ om m.~m nm em mam mmfifl .. Nfifi oe «.mm on no man mmfifi .. NH“ on m.mn mm mm use mama .. Nag om m.mm as on mum Hmmfi -- NHH oH m.mm mm mm on oooH Hem NmN om m.mm us mm «mm eefifi mow mm” ca o.mm mm mm mmm momfi oufi HmH om m.mm mm mm mam «mNH oefi omH om ¢.em an em mum Ham“ eNH Nag oH m.mm mm mm mfie mmoH afim omH om «.mn um um “we mmHH mo“ oma ow ¢.m~ on no mme mafia mm“ emfi om m.m~ mm mm see new“ may mo~ om w.m~ an mm mmc mama mofi mm oH m.mn ow mm “mm Hmofi Nu“ NmH om m.mm Nu mm oem HmHH mmH oNH cc H.mm on ma mmm flag” NHH cog on n.9m mm mm Hem mmmfi mm mm om o.mm an em mom mum“ mm as ea o.mo smumaw muacsau mnouccou mm? o woo: ;\w: 96 a go a mmop mcapmcmasmp “\mx «coucou ocawmcmnsou new: ounces» cop» Ezmcou mcaumwoz cw so aucmvqumm Foam .mmcaumcmnswp awn mcngu acmcmwmvu so» nmwmmn um: ucmugmav mucmucou asapmroe acmgmmmvuzpmmawzu:uooz mcwm: mucmsgoCLma Seaman mommxoumcsauuue.m mpam» 164 .Amwmmm pm: pcmucmav mcaamwoz Fan; Co cowuucaa a mu Agxmxv mama vow; Foam .H.m mesa?“ .\.\ .\ .\ \.\ \.\. \III\\J ..\ 6666666 . .\ \\6\\ .\ \.. .\ A 0050 .\ \|I\\| .\\\\ .\\1. \ \ . \ .\ \ \ ome\x\\\ \\ \\ O “\“ \\\ ..\ z O.x\ns\ \ ,3— an— 1 ms— 1 Igw_ (Ii/530 HINGEJ‘BU 165 The three curves in Figure 6.1 fit an exponential curve: Fr = aebM (6.12) where Fr is the fuel feed rate (kg/h) and M is the fuel moisture (percent). The regression coefficient b is 0.018 for the three curves. The coefficient a depends upon the higher heating value of the fuel and the required total heat output (as a function of the drying air- flow rate and temperature as well as the ambient temperature). The values of the coefficient a for the conditions in Table 6.4 for wood chips are 60.18, 74.69, and 91.14; the values for corncobs are 67.45, 83.50, and 102.45 for average drying temperatures of 67, 80, and 94°C, respectively. The coefficients a for corncobs are higher than for wood chips (at the same drying air temperature) because corncobs have a lower heating value than wood chips. The coefficient of determination for all the curves is 0.989. It can be concluded that the fuel feed rate required for a given drying condition increases exponentially with the fuel moisture content. The increase in fuel feed rate is attributed to the reduced dry matter of the fuel and the heat required to evaporate moisture from the fuel. The increase in the fuel dry matter feed rate at a constant drying air temperature is relatively small compared to the increase in the wet fuel feed rate. At a constant drying air tempera- ture of 67°C, the wood chips dry matter consumption increases from 67 kg/h when the moisure content is 10% to 76 kg/h when it is 50%. Table 6.4 also contains information on the furnace performance characteristics at varying fuel moisture contents at a constant fuel 166 (wood chips) feed rate of 112 kg/h. The results indicate that for every 10 percent increase in the fuel moisture content, the furnace temperature decreases by an average of 62°C, and the drying air tem- perature by 10.6°C. This is due to the decrease in the fuel heating value since higher moisture fuel contains less dry matter and, there- fore, less net energy than the same quantity of the lower moisture fuel. At a constant feed rate (and constant ambient temperature), the heat losses decrease inversely with an increase in the fuel moisture content. A 10% increase in fuel moisture results in 77 MJ/h decrease in the heat losses. This can be attributed to the lower furnace tem- perature. The furnace efficiency remains constant, but the system effi- ciency at constant fuel consumption increases with an increase in the moisture content due to a reduction of the duct heat losses. The duct heat losses decrease since the drying (and therefore;the-duct) temperature decreases. The increase in the system efficiency is 6% when the fuel moisture content increases from 10 to 50%. 6.2.3 Effect of Excess Air on Furnace Performance Excess air is the quantity of the combustion air in excess of the stoichiometric air required for complete combustion. Some excess air is necessary in biomass furnaces to ensure complete fuel combustion (Babcock and Nilcox, 1978; Buchele et al., 1981). If combustion air is limited to the stoichiometric quantity, incomplete combustion may result due to a lack of thorough mixing of the fuel and the air. High 167 moisture fuels require high levels of excess air for drying the fuels prior to combustion (Tillman et al., 1981). Too much excess air slows down the combustion reaction by excessive cooling of the furnace combustion chamber. More importantly, the excess air increases the pressure in and the mass flow rate from the combustion chamber. This results in backfiring (an explosion of prematurely ignited fuel) in the feed system and in an increase of the furnace gas velocities. Unburned entrained materials are carried out of the furnace and into the drying bin under such conditions. In the MSU concentric vortex-cell furnace, the combustion and the eductor air are supplied by the same fan. Increasing the excess air reduces the eductor airflow. Consequently, the eductor develops less vacuum in the chamber causing backfiring and smoke-escape through the feed hopper. Table 6.5 contains the simulated temperatures in the combustion chamber and of the drying air along with the furnace and the system efficiencies at different excess combustion air percentages. The data are based on an ambient temperature of 4°C and an air flow rate of 12.5 m3/m1n-m2 (17,600 kg/h). The data in Table 6.5 are plotted in Figures 6.2 and 6.3. Figure 6.2 is a plot of the furnace combustion chamber and exhaust flue gas temperatures versus excess air (percent). The furnace and the system efficiencies as well as the drying air temperature are plotted against percent excess air in Figure 6.3. The furnace com- bustion chamber temperature decreases exponentially from 1,518 to 168 Table 6.5.--The theoretical effect of excess combustion air (percent) on the CVCF furnace performance. Furnace and drying Furnace and system Excess Heat air temperatures efficiency air loss °C % % MJ/h Combustion Exit Drying Furnace System 0 591 1518 579 82.6 92.9 72.3 10 510 1457 590 82.9 93.6 72.9 20 499 1400 600 83.2 94.3 73.4 40 481 1300 615 83.5 95.4 74.4 ‘50 472 1248 621 83.5 95.8 74.9 60 464 1203 626 83.4 96.2 75.3 80 446 1117 631 82.9 97.0 76.2 100 427 1038 632 82.0 97.7 77.3 140 379 896 622 79.0 99.2 79.8 180 345 763 594 74.5 99.4 81.6 NOTE: Ambient temperature 4°C, relative humidity 75%, drying airflow 12.5 m3/min-m2 (292 kg/min) 169 .mczumcmaea» mac «spa uwxm on» cw use amasmgu :o_um:nsoo mg» no cw< mmmuxm Co uumaam use .~.m mczmwu 0‘. Iciololololoiolo‘o‘o‘uu‘. g a: m5 3...". #36 ll.) g Babb—gill] 170 .xucmwuwmwm Embmaw may new .aocmwuawmm ounces; .AQOV ocuumcmnsmp cw< mcwzgo mg» no gw< ammuxu Co pummau asp m.m «gnaw; Czwumwmv m2 wmwuxm In In a" 8 I 8 n 2. Ill. \10\|I|\J a an m «a I z a m 1 8 / gain—u Eh... .....Ill. ruzuHuHuum wuom cow “mauve; Anna—V mumoo mcpausoao can muacc :o_uns:m:ou A—uapv aucuco .mouae mcvasu pasau<--.~.o u_a~» 179 ambient relative humidity varied from 70% to 100% during the two test periods. In Test No. 1 (Table 6.7) corn was dried from an initial moisture content of 26.2% to an intermediate moisture content of 18.8% in the high-temperature IBCF system and dumped hot(about 60°C) in the dryeration bin. The corn was held for approximately 6 hours after which it was cooled to the 16% final moisture content using ambient air at a low airflow rate (approximately 0.5 m3/min-m2 per m3 of corn). The corn was similarly dried in the Tests Nos. 2, 3, and 5. During the Test Nos. 4, 6, and 7 the corn was dried directly to a final moisture contents of 14.2%, 13.1%, and 14.6%, respectively, in the high-temperature system. The electric power consumption was 10.7 kWh/h during the heating-drying phase. The difference between the 10.7 kWh/h and the rates in Table 6.7 are due to the electrical energy required to operate the cooling fan in the dyeration bin. 6.3.1 Biomass Fuel Feed Rate The average fuel feed rates are included in Table 6.7. The fuel feed rate depends upon the total thermal energy required for the drying process as discussed in section 6.2 (see also Table 6.1 for the same information). 6.3.2 Experimental Specific Energy Consumption The specific energy consumption (SECO, kg/kg of water removed) is presented in Table 6.7 under two categories, namely the net SECD 180 (for drying only) and the gross $600 (which includes the heat loss from the system). The net SECD is calculated by multiplying the tem- perature difference between the ambient and the drying air (°C) by the specific heat of air (kJ/kg) and the mass flow rate of the drying air (kg/h), adding the product to the electricity used (kJ/h), and dividing the sum by the water removed from the grain (kg/h). The gross SECO is calculated by multiplying the fuel dry matter consump- tion by the higher heating value and adding the electrical energy to the product; the sum is divided by the amount of water removed. As can be seen in Table 6.7, the net and gross SECO depend upon the combined effects of the following factors: 1. The amount of water removed per ton of wet corn. 2. The moisture content percentage points removed after drying in the high-temperature phase. 3. The ambient temperature. 4. The drying temperature. 5. The airflow rate of the drying air. The effects of the above factors on the drying capacity and the SECO are discussed in detail in section 6.5. It is obvious that when drying is completed in the high-temperature phase (i.e., corn dried to 15.5% or less), the process results in high SECO values. Thus, Tests 4, 5, and 7 resulted in high SECO values compared to Test No. 1. The ambient and the drying air temperatures, as well as the drying airflow rates, have a direct effect on the SECD values. This effect is not obvious from Table 6.7 since some of the important para- meters varied simultaneously. 181 6.3.3 Drying Capacity The average drying capacity at an average drying temperature of 84°C for the seven tests shown in Table 6.7 is 2.2 wet tons per hour in drying from 25% to 15%. The drying capacities for Tests 6 and 7 are much lower than for the other tests (1.5 and 1.8 compared to 2.7 and 2.5 wet tons per hour for Tests 1 and 2, respectively). This is due to the fact that the corn was dried to a very low mois- ture content in the high-temperature phase (14.6% and 14.7% compared to 18.8% and 17.0% in Tests 1 and 2). If Tests 6 and 7 are excluded, the average capacity for the other tests is 2.4 wet tons per hour. This compares with a capacity of the IBCF fueled with propane of 2.6 wet tons per hour in drying from 26% to 17% at a drying air temperature of 71°C (Silva, 1980). Thus, there is a small loss in drying capacity when the IBCF is operated with biomass fuel. The slight decrease in capacity is due to the air flow reduction in heating the air before the fan entrance. 6.3.4 Standardized Dryer Performance The standardized dryer performance data are presented in Table 6.8. The data include the value of the energy consumption (fuel hg/ha, electricity kWh/ha), capacity (wet ton/h), SECO (kJ/kg of water removed), and the drying costs ($/wet ton). The standardized per- formance is based on drying corn from 26% to 18% moisture in the high-temperature phase and from 18% to 15.5% in the dryeration/cooling bin. The energy consumption per hectare is based upon a corn yield of 6.28 wet tons per hectare. 182 .n.3 oz NNH op oN soc» umwcu Hocum\:n ooHv msmpum: son mcop.upsuo2 NN.m mo uHmHa ceou co vmmumm .zzx\su xnwuwcauopm .Humzv ux\eom.m um mnooccou .Hpmzv mx\ene.e pm umaHm>mang woo: a .Emumzm coHuosngu cwnucH cm :H am.mH on umHoou ucm ewumxm mumH :H .N.@ «Ham» cw venom up emu mosaumgmnemu acmHasu mg» com magma ummw Hmzm new Hmnoucsou n u.mgH;u_uooz u 3V mHmam acmsmCCHu so» Ammuwsa mmmHv mumoo mcHumsmao ucm coHHa53mcou HHmzmv Hmsmcm umupucmccmumuu.w.o mHnmp mm.o N¢.m comm mm.N om Nee mNH Nw zuuo.me N mH.o mm.H Hmom ¢N.N mN mNN ooH mm uiuH.oH m MN.o me.N mmmm mm.N om mmN ONH Nm 2--m.mH m mH.o mm.H NNNm mm.N oN NeN ooH mm u--H.NH e HN.o ON.N mHHN NH.N mm Hmm mHH NN uu-H.NH m NN.o oN.N omNo oo.N mN moN oHH mm uuum.oH N HN.o 0N.N comm NH.N mm meN mm NN 3--o.¢H H i as am. is 5%... .: a. ......9... new”... 3...”. .nmpmou mcwago uHmHumnm Humumgmu . . acowqu=mcou Hmzm ch ocHHgo Hazy 9.5mumzm 183 The following is the standardization procedure: 1. The corn drying constant k for each test is derived from the drying equation: M - M MR = fi = exp(-kt) (6.17) Rearranging equation (6.17): k = -tn(MR)/t (6.18) 2. The theoretical time required to dry corn from 26 to 18% moisture is calculated as follows: tS = £n[(0.IB-Me)/(0.26-Me)]/k (6.19) 3. The standardized capacity (wet ton/h) is: CPs = CPe/tS (6.20) where CPS and CPe are the standardized and the experimental capacity (wet tons per hour). The other standardized data are based on the standardized capacity (CPS) and time (ts). 6.3.4.1 Standardized Drying Capacity The standardized capacity is 2.2, 2.5, and 2.7 wet tons per hour for a drying air temperature of 78, 82, 89, and 93°C, respectively. The average standardized capacity is 2.5 wet tons per hour at an inlet air temperature of 84°C. 184 6.3.4.2 Standardized Specific Energy Consumption The average standardized gross SECO for the CVCF/IBCF system is 6,099 kJ/kg of water, compared to 4,548 kJ/kg if propane is used (Silva, 1980). The difference is due to the extra heat loss through the additional duct work linking the furnace to the dryer (see section 6.2). 6.4 Corn Drying Simulation The in-bin counterflow grain drying model discussed in section 4.2 was used to simulate the drying of corn in the CVCF-IBCF system. The simulated and experimental results are compared in Tables 6.9 and 6.10. 6.4.1 Cycle Times The in-bin counterflow dryer (IBCF) is a semi-continuous counterflow dryer. Grain is dried in a stationary mode as in an in-bin batch drying system. When a thin layer (7.5-10 cm deep) of corn at the bottom of the bed is dried to a predetermined moisture content, it is transferred to a dryeration bin (see section 4.2). The transfer process takes approximately 25 minutes during which the drying proceeds. The time between two consecutive starts of the transfer process is defined as the cycle time. It is the time required to dry the 7.5 - 10 cm of grain that is transferred to the dryeration bin. The experimental and simulated cycle times, moisture content at transfer, capacity (wet tons per hour), airflow rate and specific energy consumption rate are tabulated in Table 6.9 for Test No. 1. (See the tables in Appendix B for an explanation of other tests.) 185 New xchHsas m>HHmHms «cownsm mmmcm>o NN.oN weapmwoe :Hmsm HmHuH:H .ua¢.NH assumcmaamp ucmmnsm mmmcm>m EN spawn :Hmsm HoHuH:H .ueaa msaumcmasmp mcngu mmmcm>m mucmucou «caumwoe mmmco>< “meoHuHucou mcHumcmao ; .msHp Hench” compasamcou xacmcm uHmHomam umz "ooumH came m.o m.o 1- -- o.oH o.aH 0.0 oo.NH aeHHoou - - N.mH - 1- mx.mH mm.NH Na.m Noo.oH Hmmo H.mH - o.m N.N m.mH m.mH Ha.o mm.o HH mmmo a.eH - o.m x.N N.NH - Ha.o mo.H oH moan e.¢H N.eH o.m a.N m.mH - HN.o om.o a Nome H.eH m.NH m.N N.N o.mH m.mH Na.o mm.o m mmvm N.mH m.mH a.N N.N m.wH m.mH Na.o ma.o N ammo m.mH 1- m.N a.N N.NH 11 ma.o mm.o o moon N.mH m.mH a.N N.N m.NH 1- ea.o aa.o m coma m.NH - m.N o.N m.mH o.mH mm.o aa.o a mamm a.NH m.NH m.N m.N m.mH a.mH oa.o No.o m wam ¢.NH m.NH ¢.N H.N m.mH - mm.o mm.o N onN o.NH o.NH N.H m.H H.NH a.mH me.H om.H H umpmpaswm arm axm EHm axw EHm axm EHm axe .oz 0N: mxxax . . A apexu Hoomm Ns\eso .onmsH< :xcou .xuwuaamu a 3 u .mcaumwoz ; .msHp «Homo co» mcmpmsmcma mucmscoacma aumH-ao>u Hmpcmspsoaxm new umpmstHm .H .oz pump Co campgmasou <-.m.m oHaaH 186 Table 6.10.--Simulated operating conditions during Test 1. Bottom layer Exhaust air moisture content Average Remaining Cycle % moisture bed No. Temperature RH in the bin depth °C % Before After: % m cycle cycle 1 26 100 26.2 18.7 26.2 1.85 2 38 100 19.9 18.5 26.8 1.69 3 43 100 20.7 18.8 27.0 1.53 4 44 100 20.9 18.6 26.5 1.37 5 45 100 20.6 18.8 26.0 1.21 6 47 92 20.5 18.8 25.4 1.05 7 48 65 20.3 18.6 24.7 0.09 8 54 42 20.2 18.6 23.8 0.74 9 59 27 20.1 18.5 23.0 0.58 10 64 18 20.0 18.8 22.5 0.42 11 70 10 20.0 18.8 21.5 0.26 187 The cycle times at different cycles are plotted in Figure 6.4. The first cycle requires the most time. The first cycle time is the time duration from the moment the drying fan is started until the start of the transfer of the first dried layer of corn to the dryeration bin. It requires more time to dry the first layer because of the following reasons: 1. The grain is initially cold, and it requires time (and energy) to heat the grain (and the drying struc- ture) before evaporation of moisture from the grain is initiated. 2. All the grain in the bin is nearly at the same high initial moisture content and low temperature. More time is required to dry the initial bottom layer to a specified moisture content than is the case for sub- sequent layers. 3. The airflow increases as the bed depth decreases. The length of the time interval between the start of the dry- ing process and the first transfer depends upon the initial grain temperature and moisture content, the grain depth, the drying tem- perature, and the final (transfer) moisture content. The cycle time of the second transfer is shorter than of the others. During the drying of the first cycle, the grain above the layer transferred is dried to a moisture close to the desired final moisture content. Thus, it takes a short time to dry the grain. The simulated average moisture contents of the transferred corn before and after drying, the average moisture of the grain left in the 188 .H .oz amok co» moHan ac amassz m> mmEHP mHuxo .e.m mesmHa mmHu>o “5 ”.3552 x a .newgb. m a \_\_V a. “mam Q. I R I A O 0 Ki .3 (HOOH) 3W I1 311MB 1 I J R ..i I 10 189 bin, and the exhaust drying air temperature and relative humidity are presented in Table 6.10. It can be seen that the initial mois- ture content for cycle 1 is 26.2%; for cycle 2, 19.9%; and for cycle 3 to 9 the initial moisture is about 20%. As the drying continues, the following changes occur: 1. The average grain temperature is increased. 2. The average moisture content of the grain remaining in the drying bin first has increased due to rewetting of the grain at the top, and then is decreased as the drying front moves upward (see Table 6.10). 3. The airflow rate is increased since the resistance to the airflow is decreased as the grain depth decreases. 4. The initial moisture content for each cycle (being the moisture content at the bottom layer after removal of dry grain in the previous cycle) is decreased gradually. 5. The time between consecutive cycles decreases and therefore, the drying rate (capacity) is increased due to the combined effect of factors 1-4 above. The experimental results in Tables 6.9 and 6.10 show some deviations from the expected trends, such as the experimental cycle time for cycle No. 10, which increased to 1.05 hrs compared to the previous and the following cycle times of 0.80 and 0.93 hrs, respectively. Also, the experimental cycle time increased from 0.77 hrs for the fifth cycle to 0.95 hrs in the 7th cycle. These irregular- ities are due to the falling drying air temperature, the difference in 190 the original moisture content of grain in different layers, or a combination of the two. 6.4.2 Drying Capacity The drying capacity (wet ton/h) is defined as Capacity = Gilt (6.21) ° 1 'l IIMZ where N is the total number of cycles dried after time t (hours) and Gi is the weight of wet grain dried in cycle i (ton). The average capacity can also be defined as the product of the number of cycles and the average amount of grain dried per cycle, divided by the total drying time. Both the simulated and the experimental capacities increase with time (and therefore with the number of cycles) as the cycle times decrease. The reasons for the reduced cycle time are discussed in section 6.4.1 There is close agreement between the simulated and the experi- mental capacities (3.0 and 2.7 wet ton/h, respectively). 6.4.3 Specific Energy Consumption The simulated specific energy consumption (SECO) presented in Table 6.9 is the net energy required to remove one kilogram of water from the wet grain. The net SECO does not include heat losses from the system and is computed as follows: 191 SECO( = N) (GaiCATAti + Ei)/Wi (6.22) IIMZ i 1 where Gai is the airflow during the drying of cycle i (kg/h); C is the specific heat of air (kJ/kg°C); AT is the temperature difference between the ambient and the drying air; Ati is the cycle time (h); E1 is the electrical energy used in drying cycle i (kJ); and W1 is the amount of water removed in cycle i (kg). The total amount of water removed after a given cycle is calculated as follows: ”i ‘ (moi ' mfi)GDi + (mBo ‘ mBi)GBi (5°23) where moi is the initial moisture content of the grain at the transfer layer dried in cycle i (decimal, dry basis); mfi is the final moisture content of the grain transferred in cycle i; mBO is the initial average moisture content in the bin above the layer removed in cycle i; mBi is the average moisture content of the grain remaining in the bin after cycle i has been transferred; GDi is the dry matter of the corn transferred in cycle i; and GBi is the dry matter of the corn remaining in the bin after cycle i. The specific energy consumption at the end of each transfer as defined by equations (6.22) and (6.23) has not been determined experimentally. It requires the grain moisture content of the bottom layer before and after the layer has been dried as well as the average moisture of the grain remaining in the bin. These values cannot be obtained in a commercial operation. 192 The variation of the simulated specific energy consumption as drying progresses is illustrated in Table 6.9. The SECO is low during the first cycle, reaches a maximum during the second cycle, and decreases until the 8th cycle is reached. After drying is started, 1.43 hours elapses (1.8 hours expe- rimental) before the first layer of dried grain (26-18.7%) is dis- charged from the dryer. During this time the grain immediately above the bottom layer is dried to 19.9%. Thus, the initial moisture for the second cycle is only 19.9%, compared to 26.2% for the first cycle. Therefore, considerably more water is removed in the first cycle than in the second (224 kg in the first cycle compared to 86 kg for the second cycle). The shorter drying time for cycle No. 2 results in a higher specific energy consumption because there is little water removed in drying the corn from 19.9 to 18.5%. Due to the shorter drying time for the second cycle, the grain above the layer transferred is 20.7% (compared to 19.9% for the second cycle). The third cycle requires 0.76 hours to dry. The fourth cycle time takes longer than the third cycle for the same reasons. However, after the fourth cycle, the cycle time progressively decreases because the grain is sufficiently heated namaintain a steady drying rate. After the third cycle, the amount of water removed per cycle remains nearly constant (143 kg/cycle). However, the cycle time progressively decreases since the average temperature of the corn is also increasing as the drying continues. The decrease in the energy comsumption is shown in Figure 6.5. 193 = ‘ 2 2 8 NUMBER (I: CYCLES 2%Z88 2 2 W: a\\\\\\\\\\\\\\\\\\\\\\jj 1 k \~ J l J l l (02H 641m NOIIdWIlSNOO AOUBNEI 013103-15 ‘ of Cycles. Figure 6.5. Simulated Specific Energy Consumption as a Function of Number 194 As drying progresses, the grain depth also decreases. The decrease in depth results in an increase in airflow rate. After the 6th cycle the exhaust air leaving the drying bin is no longer saturated (see Table 6.10). The exhaust relative humidity decreases and the temperature increases. This is an indication of the start of ineffie cient use of energy. The specific energy consumption reaches a minimum after the 8th cycle and thereafter increases. The increase is a result of the decreasing amount of water removed per cycle. The depth at which the minimum specific energy consumption is reached depends upon the drying air temperature, the initial grain depth and moisture, and the degree of drying. 6.4.4 Airflow Rate The simulated and the experimental airflow rates at the begin- ning of each cycle are tabulated in Table 6.9 and plotted in Figure 6.6. The experimental airflow rates were determined by measuring the static pressure (in the bin plenum) and the bed depth. The characteristic fan curve was also used to determine the airflow rate using the static pressure readings. The simulated airflows were predicted employing the procedure discussed in section 4.2. There is a close agreement between the experimental and simulated airflow rates. 6.4.5 Comparison of Experimental and Simulated Grain Moisture Contents Experimental and simulated moisture contents of each cycle are presented in Table 6.9. 195 I\\\\\\\\\\\\\\\\\§ = m\\\\\\\\\\\\§ = mxxw - t\\\\\\\\\\\8 «- m\\\\\\\ 7 NUMBER (I: CYCLES 7 / ? 18 "'3 V at (z. mpuflugm) ND'IJHIV ng of Each Cycle. Figure 6.6. Simulated Airflow Rates at the Beginni 196 The desired final moisture content is one of the inputs to the grain drying model. Whenever the average grain moisture content within the bottom layer (7.5-10 cm above the false floor) reaches the desired moisture, the layer is removed from the drying bin instantane- ously and dumped in the cooling bin. Therefore, the simulated mois- ture content is always equal to or slightly less than the desired value. The difference between the average experimental and simulated moisture contents is small (e.g., 18.8% experimental compared with 18.7% simulated for Test No. 1, Table 6.9). 6.4.6 Comparison of Experimental and Simulated Drying Data Table 6.11 contains the experimental and simulated corn drying results for seven tests under different drying and ambient conditions. The average drying capacities and net specific energy consumption results are compared. 6.4.6.1 Capacity The experimental capacity was determined by dividing the total amount of corn dried by the total drying time. The simulated capacity was determined by the procedure outlined in section 6.4.2. The data indicate close agreement between the experimental and simulated results. The difference between the experimental and the simulated capacity is due to the following factors: 1. The sweep auger does not remove equal amounts of grain each time when grain is transferred to the 197 vamHaswm "sHmN Hmucmswcmaxm "ame emNm amNm emNo NwNo w.H m.H o.eH e.NN Nm o.o a NmNe meme acme mmmm o.N m.H H.mH\N.eH m.mN mm N.NH m mmwm NNem emmo mmmm N.N o.N m.mH\N.oH e.oN Nm m.o- m mmHm Name mmHm Name e.N e.N N.eH N.NN mm m.m e mmem Name mmam mmHm N.N N.N e.eH\m.mH m.NN ma o.N m HNme Hmem eNmN moan m.N m.N o.eH\o.aH m.mN mm m.mH N oHoe HNmm Hmmm Noam o.m N.N o.mH\m.wH N.mN Na e.NH H NEHm Haxm NeHm Haxm aeHHOOe eczema; Neem Haxe Hee_e Feceeee a. a. .62 manuacmasmp waspmcwaewp Hawk Loam: mo 9:3 $2; {=3 33. N 23:8 953.85 9.35 p535 cowuasamcou xmcmcm uHewomam xuwuaamu mmmcm>< Hmucmchmaxm .mcoHuHucou mcwxgu acmLmCCHc cmuc: mummp cw>mm ace mszmmc umpaHaswm use Hmucmswcmaxm aumH1au>u eo comwceasou-.HH.m mHnmh 198 dryeration/cooling bin due to uneven grain flow characteristics such as density, broken kernels, foreign materials (BCFM), etc. 2. The simulation model uses average ambient and drying temperatures, and an average initial grain moisture content; these parameters vary somewhat during a test. 6.4.6.2 Simulated Versus Experimental Specific Energy Consumption The experimental and simulated net specific energy consumption values for the heating and cooling phases are also presented in Table 6.10. The experimental SECO data are based upon the average airflow rate, the average moisture content, and the average drying capacity. The simulated SECO values are based on the integration of the SECO values for every cycle as discussed in section 6.4.3. There are differences between the experimental and the spe- cific energy consumption values due to the methods used to compute them. The integration method (used to calculate the simulated SE00) is more accurate but requires a knowledge of the average moisture content of the grain in the bin, as well as the moisture of the grain at the bottom of the bin before and after each transfer. 6.5 Predicted Drying Performance Parameters The furnace and the drying models are used to predict the furnace/dryer performance under varying ambient conditions and dryer operating modes. The IBCF dryer is operated in four basic modes: Mode 1: Mode 2. Mode 3. Mode 4. 199 The bin is filled to a depth of 1.8 m. Drying is started and continues until the bin is empty. A total of 30 wet tons are dried in 8 to 12 hours depending on the drying temperature, the initial moisture content, and on whether drying is completed in the high-temperature phase or in the cooling bin. The bin is initially filled to a depth of 1.8 m. Drying is started and the bin is refilled with 8 tons of wet grain every three hours. Drying continues until grain loading is stopped after 18 hours. A total of seven batches are added during the drying period. Approximately 57 wet tons are dried (21 cycles). As in Mode 2 except that 16 wet tons per batch are added. Drying continues in this mode until the bin is filled up to 3.7 m after 11 hours. Four batches (a total of 64 wet tons) are added to the bin. A total of 12 cycles (32 wet tons) are dried. As in Mode 2, but 4 wet tons are added per batch. The drying rate is greater than the refill rate. Drying con- tinues until the bin is empty after 16 hours. A total of five batches (20 wet tons) are added to the initial 30 wet ton. Fifty wet tons (17 cycles) are dried during the 16-hour drying period. In addition to the four modes discussed above, the IBCF dryer is used to partially dry grain to an intermediate moisture content and 200 finally dry the grain in the dryeration bin (Method 0). Alternatively, the grain can be dried directly to the final moisture in the IBCF dryer (Method ND). Table 6.12 presents the predicted IBCF system performance parameters (simulated) in different drying modes and at different drying temperatures. The following parameters are studied: capacity, specific energy consumption, fuel feed rate, and drying costs. 6.5.1 Effect of Refill on Dryer Performance The IBCF performance is better when operating in Modes 2 and 3, than in Modes 1 and 4. The capacity is 3.2 and 3.1 wet ton/h in Modes 2 and 3, respectively, compared with 3.0 and 2.9 wet ton/h in Modes 1 and 4. The net specific energy consumption is 4656 and 4792 kJ/kg of water removed in Modes 2 and 3 compared with 5010 and 5126 kJ/kg in Modes 1 and 4. The operating (energy) costs are also lower in Modes 2 and 3 than in Modes 1 and 4 ($1.81, $1.86, $1.90, and $1.91, in Modes 2, 3, 1 and 4, respectively). The performance is superior in Modes 2 and 3 than in Modes 1 or 4 for the following reasons: 1. Replenishing the bin with wet grain ensures complete saturation of the exhaust air. Thus, the air water- carrying capacity is fully utilized, resulting in better energy utilization. 2. The wet grain is preheated by the exhaust air and therefore, dries faster upon reaching the bin bottom. 2(11 .uo>oeoc smug: mo mxxex .coHuaeanccu xacucu AccHNoguxuu on. aumH :— am.mH on vapgu :Lou uoz .um.mH on :oHuucoagu he umon—om smart aumH :— u: an o» vupgv :590 no e.c_ueaa «a: "Comm m "canvas uc—huow .m.m :o—auum c. vucpmou mouOtH mw.m mh.H am.~ Nh nNH mam ONum NQ.N 9 ON ON a." v om.m mo.~ Dm.H NN MNu mod Nosc so.” a oN om m.~ m cm.m no." “a." Nb MN" God can? o~.n a mN on N." N on.m no.H um.~ Nu MN~ mom mmmv 0H." 9 ON on N.N Nm.m mo.N QN1N um aha MNH Nsmm mh.N oz 0N em m.H Nm.m Nh.~ ma.H um and MNH Ova MN.m 0 ON cm m.H Hm.m N0.H OQ.H Nu MNH moH vuvs m~.m 02 NN om m1H o¢.¢ on.H Hm.H Nu MNH mad HNNo am.” a NN cm a." o¢.o ma.~ om.N NN MNH mcH numm vo.N az 0N om m.— no.m NN.~ om.H Nu MN" mOH oHom Oo.n a 0N om m.~ ¢Q.© OH.N mm.N no QOH om mnoo ON.N oz ON so Q.H mo.m «N.H em.~ no mcH mm oHom mo.N 9 ON mm m." H ...... mm ...”...x ea... aw ...... s... .. .. mmmmwm WM”; NWMuwwu u: ocauecoasuu guauo Hone: “unweemnflmea wuawowewceflwua aeexga .a.e_e_ ae.xga ...e.=_ .Hza nmu .uaev neoHHHucou acoHaEu «unease an autos coHuacoao soxgu «cocowaHu com mumou uc—uacmac use mung mcquv .coHuassmcou xmcoco vane-:sHm-1.NH.m oHauh 202 3. There is no increase in the airflow rate since the depth either remains constant (Mode 2) or increases with time (Mode 3); the airflow rate increases in Modes 1 and 4 with a decrease in grain depth. 6.5.2 Effect of Dryeration on System Performance The data in Table 6.12 indicate that the dryeration process results in a higher drying capacity, a lower specific energy consump- tion, and a lower drying energy cost compared to drying completely in the IBCF dryer. When the drying air temperature is 80°C, the simulated drying capacity is 3 and 2.6 wet ton/h with and without dryeration, respectively; the respective SECO values are 5,010 and 5,573 kJ/kg of water, and the energy cost per wet ton $1.90 and $2.16 (wood chip fuel). The improvement of the system performance with dryeration is attributable to the following factors: 1. Less moisture is removed by the IBCF dryer resulting in less drying time of each wet ton of grain in the counterflow dryer. 2. The last 2.5% moisture remaining in the corn after the high-temperature drying phase is the portion of the moisture most difficult to remove. This moisture is now removed slowly during the aeration process. 3. The aeration process uses a low airflow rate (0.5 - 1 m3/min-m2) and an insignificant amount of electrical energy to cool the grain and remove the last 2.5% of moisture. 203 6.5.3 Effect of Drying Air Temperature on System Performance The drying air temperature has a direct effect on the grain drying capacity, the specific energy consumption, and the drying costs. The effect of the drying air temperature is shown in Table 6.12. The drying capacity increases substantially with an increase in the drying air temperature. The drying capacities are 2.7, 3.0, and 3.2 wet ton/h when corn is dried from 26 to 18% at the drying air temperatures of 67, 80, and 94°C, respectively. The higher tempera- ture air increases the drying rate of the corn and has more water- carrying capacity since the saturation absolute humidity is higher. The effect of drying air temperature on the specific energy consumption depends upon the moisture percentage points removed in the high-temperature phase. If corn is dried from 26 to 18% the SECD is 5,010 kJ/kg at 67 and 80°C drying air temperatures, and 5,140 kJ/h at 94°C. However, if corn is dried from 26 to 15.5% in the high temperature phase, the SECO is 6,035, 5,573, and 5,972 at air tem- peratures of 67, 80, and 94°C, respectively. Thus, if drying is com- pleted in the high-temperature phase (i.e., corn dried down to 16% moisture content or less), the SECO decreases with increasing drying air temperature up to a point after which the SECD increases with the temperature increase. The inflection point depends upon the ambient temperature, the amount of water removed, and the rate of increase in drying capacity with the temperature increase. In the example given above, the drying capacity is 2.2 wet ton/h at 67°C and 2.6 wet ton/h at 80°C. Thus, the capacity increases 204 by 20%; the energy required'Ulheat the drying air increases by 19% resulting in an 8% reduction in the specific energy consumption. When the drying air temperature is raised to 94°C, the capacity increases by 26% while the energy increases by 40%. Thus, raising the drying temperature from 67 to 94°C results in a 1% reduction in the SECO when corn is dried from 26 to 15.5% in the IBCF system. When corn is dried from 26 to 18% moisture in the high- temperature phase and finally dried to 15.5% in the cooling phase, the drying costs are $1.94, $1.90, and $1.96 per wet ton at tempera- tures of 67, 80, and 94°C, respectively, compared to $2.33, $2.16, and $2.28 per wet ton if the corn is dried from 26 to 15.5% in the high-temperature phase. Thus, the drying costs vary inversely with the drying capacity. 6.5.4 Effect of Initial Moisture on System Performance Table 6.12 contains the simulated IBCF dryer performance para- meters when drying corn from 26 to 22% initial moisture contents. The results indicate that when corn is dried from 22% moisture, the drying capacity is higher and therefore the drying costs per wet ton lower. The drying capacity (with dryeration) is 3.8 and 3.0 wet ton per hour, the drying costs are $1.51 and $1.90 per wet ton for the 22% and 26% initial moisture content corn, respectively. The costs per kg of water removed is higher for the 22% initial moisture corn. The reason for the higher drying capacity is that less water is removed per ton of wet corn if the initial moisture is 22% than if it is 26%. 205 Corn with an initial moisture of 22% or less can also be dried in low-temperature or natural air drying systems (Bakker- Arkema et al., 1981). 6.5.5 Effect of Ambient Temperature on System Performance The ambient temperature has a direct effect on the specific energy consumption and the drying costs. At lower ambient temperature, the heat loss from the system is greater as well as the energy required to raise the drying air to a given drying temperature, than at a higher ambient temperature. The specific energy consumption is 5334 kJ/kg at 0°C ambient temperature, compared to 5010 and 4514 kJ/kg at 4 and 10°C, respectively. The drying costs (wood chip fuel) are $2.01, $1.90, and $1.77 for drying at 0, 4 and 10°C ambient temperature, respectively. The drying capacity does not change with ambient tem- perature if the drying temperature remains constant. 6.6 Economics of the CVCF/IBCF System The CVCF/IBCF drying system is a technically feasible system. It would be economically viable if it results in reduced drying costs compared to a propane fueled IBCF system. A substantial investment is required to convert the propane fueled system to a biomass system. The recovery of the additional capital costs must be through savings in fuel costs. In this section the fuel costs for operating the CVCF/IBCF system using wood chip and corncob fuels at different ambient and drying conditions are presented. The operating costs include the 206 biomass ‘fuel costs, as well as the cost of the electrical energy required to run the furnace and the dryer fans and the IBCF system control. The biomass costs are compared with the cost of operating a propane fueled IBCF system. A 12-year budgeting analysis is included to show the effect of fuel costs, labor, repairs and maintenance, taxes, depreciation, and inflation on the annual break-even costs of operating the CVCF/IBCF system. The budgeting analysis includes the life cycle costing of the system. The results are compared with the annual break-even costs for a propane fueled system. 6.6.1 Operating Costs The experimental drying costs are presented in Table 6.7. The costs represent operating costs; they do not include labor and invest- ment costs. The fuel costs are based upon the folloWing 1983 prices: Wet wood chips: F.0.B. cost $ 6.00 per ton Drying and storage 0.00 Freight 0.16 per ton/km Dry wood chips: F.0.B. cost 15.00 per ton Drying 3.75 per ton Storage 2.00 per ton Freight 0.08 per ton/km Dry corncobs: F.0.B. cost 25.00 per ton Drying and storage 0.00 Freight 0.36 per ton/km 207 Electricity: 0.07 per kWh Propane: 0.22 per liter The cost of biomass fuel depends upon the distance over which the fuel is hauled. The F.0.B. cost of biomass fuels is made up of handling expenses incurred in harvesting and/or gathering and storage. The storage and drying costs are zero for dry corncobs and wet wood chips. In the case of corncobs, the fuel is gathered during or after the harvest season and stored in storage bins which are not used for grain storage during the off-season. The corncobs dry naturally in storage. The corncob fertilizer value is approximately $25 per ton. In the case of wet wood chips, the fuel is used directly without storage or drying. The cost of dry wood chips includes the drying and the storage expenses. The fuel is obtained from a saw mill at approximately 50% moisture and dried to 30% moisture (or less). A special bin is required for storage and drying by aeration. The transportation costs depend upon the quantity per load and the distance. The transportation costs per km (kilometer) for the corncobs is high because of their low density compared with wood chips. The bulk density of dry corncobs is approximately 200 kg/m3 compared to an approximate value of 320 kg/m3 for dry wood chips. The total costs of the wood chips and corncobs used in this research are assumed to be $44.70/ton and $35.00/ton, respectively. High moisture content fuels are more expensive than dry fuels because the net heating value is much less for the wet fuels. The energy costs 208 (Table 6.7) is $5.82 per wet ton for Test No. 7 compared to $2.37 per wet ton for Test No. 6. The moisture content of the fuel (wood chips) for Test No. 7 is 45% compared to 10% for the corncobs used in Test No. 6. The higher the fuel moisture content, the higher is the fuel feed rate required to maintain the same drying air temperature. The average drying cost for the corn in Table 6.7 is $2.74/wet ton. This compares with $5.75/wet ton for a similarly operated propane fueled IBCF system (Silva, 1980). Thus, the operating (fuel) costs for the CVCF/IBCF system are lower than for an equivalent system fueled by propane (note that the investment and labor costs are excluded). The standardized energy drying costs are included in Table 6.8. The cost for drying with 45 and 19.3% moisture wood chips is $3.48 and $2.45 per wet ton, respectively, compared to $2.25 per wet ton if 14.6% moisture is used; the figure is $3.72 per wet ton if liquid propane is the fuel. Table 6.8 shows that dry corncobs are the least expensive fuel. The cost of drying is $1.58 and $1.68 per wet ton for 10% and 12% moisture corncobs, respectively, due to the low transportation costs (as discussed in section 6.6.1). The average standardized energy cost is $2.27 per wet ton for the conditions in Table 6.8. The costs for similar conditions when drying with propane fueled system is $3.72/wet ton (Silva, 1980). Thus, the operating (energy) costs are less for biomass fuel than for liquid propane. 209 6.6.2 Capital Budgeting Analysis The operating costs presented in Tables 6.7 and 6.8 do not include labor, maintenance, investment, interest on borrowed money, depreciation, and taxes. To analyze these costs, a 380 wet-ton annual corn drying system was analyzed using a capital investment model developed by Harsh (1983). The cost estimates are based upon a simulated energy consumption, drying rate, and operating costs for the CVCF/IBCF system operating in Mode 1, Method 0 (Table 6.12), at a drying temperature of 80°C. The relevant input data for the budgeting analysis are summarized in Table 6.13. The biomass furnace can be considered an addition to an IBCF system. The MSU furnace was manufactured at a cost of $11,070; this includes costs for materials ($8,570), labor ($2,000), and miscel- laneous expenses ($500). The retail price for the unit is estimated to be $16,000. The total cost of the CVCF/IBCF system is $43,000 (Kalchik, 1984) compared to $27,000 for the pr0pane fueled system. It is assumed that the additional $16,000 is borrowed at 12% annual interest rate and is payable over a five-year period. The accelerated cost recovery system (ACRS) outlined in the 1983 Famers Tax Guide (IRS, 1983) was applied to calculate deprecia- tion. The biomass furnace can be classified as a 5—year property with a recovery period of 12 years when using the ACRS code. A 12- year, rather than a 10-year, budgeting analysis is presented. A 10—year recovery is not permissible for a 5-year property under the ACRS method. Also, the expected salvage value of the property is not 210 Table 6.13.--Estimates and assumptions for a 12-year budgeting analysis (1983) prices) of the CVCF/IBCF drying in Mode 1 (see Table 6.12). Fuel Parameter . . Wood chips Corncobs °5tlm°t°° 20% M.C. 20% M.C. Pr°°°°° Fuel cost $/ton 1.90 1.72 5.32 $/ton-point 0.18 0.16 0.51 Labor cost $/ton 1.11 1.16 0.40 $/ton-point 0.11 0.11 0.04 Dryingrate ton/h 3.00 3.00 3.0 ton-point/h 31.50 31.50 31.5 Initial Investment: $27,000 pr0pane, $43,000 biomass Percent Borrowed: 0.% propane, 35% biomass Repayment Period: 5 years Interest Rate: 12% Tax Bracket: 20 Type of Property: 5 year property Depreciation Type: Straight line method Recovery Period: 12 years Number of Units Dried: 4,000 ton-point per year (381 ton dried from 26-15.5% w.b.) Losses due to Corrosion, etc.: $120 per year Repairs & Maintenance: $2,000 propane, $4,500 biomass Percent Return: 15% General Inflation Rate: 6% Fuel Inflation Rate: 10% propane, 2% biomass Biomass Fuel Moisture: 20% w.b. Drying Temperature: 80°C Ambient Temperature: 4°C Ambient Relative Humidity: 75% 211 subtracted from its basic price. The straight line depreciation method is used to compute the annual depreciation cost as a deductible item in tax computations under the ACRS system. The results of the 12-year budgeting analysis are presented in Table 6.14. The annual operating costs are expressed in $/wet ton- point. The unit represents the cost of drying one ton (wet) when one percentage point is removed (e.g., one ton wet corn dried from 26 to 25% MC, wet basis). This unit is preferred over the more common unit of $/wet ton. The latter unit cannot be readily used to estimate the drying costs when grain is dried from and to different moisture contents. The total and per wet ton investment expenditures are included in the table. The total annualized break-even drying costs for the biomass fuels is $3.00 per wet ton-point compared to $2.23 for the propane system if the annual rate of inflation of the fuel-cost is 2% for all the fuels. If the pr0pane cost inflation is 10% annually, the total drying costs for the propane fueled system if $2.26 per wet ton-point. The latter is considered a more realistic figure. Thus, the propane fueled system is the least expensive at the current propane price of $0.225 per litre. The pr0pane cost would have to increase to $0.378 per litre for the total costs of the prOpane system to be equal to that of the biomass fueled system. This constitutes an increase of 68% in the present propane cost. The variable annual break-even costs of the biomass system ($0.95 per wet ton-point) are about the same as for the propane fueled 212 umou wcaaoca :H mmmmcucp wuaHumEEH awe me wcmaoca :oHamecH mung Haze NN coHHmHecH mama Haze aoH HHV acmaoca N acmaosa e .ucwoa-cop pmzm .Aooo.HHw + ooo.mev mumccse mgu $0 Hmou uwmmn «an new: acmeumm>=H man we ooo.mmmN .Aooo.eaav eeaeezc mmeeeee 6:8 ca eeeea Panama neeeeeeme use use Hooo.ava :Hn aomH mcm>>wgm any we maHm> unoccau map Ho Sam mg» mH Hamspmm>:H ooo.memH oo.m oe.H He.H Hm.ox coo.xN Ana ”N.N me. He._ am.ox ooo.x~ ANV ae.~ mo.a He.“ em.ox ooo.x~ HHV eeeaaoea ex.N ea.o Nm.H ex.am coo.mm Au: Roma oo.m ea.o mo.~ mm.NHH ooo.me meeeeeeu ow.~ ma.o mw.H ex.ao Nooo.mm Au: Roma Ho.m ma.o oo.~ em.NHH Hooo.me aaaee use: Peace aHQaeees eexec eoe\a a Peace Pena Npcwoa-cop\m .mumou co>m xmmcn Hmacc< geesemesee Heepaea .n.z am.mH op NmN sage HHHazccm csou Ho Hpmzv use» can mcngu cw Ammuwca mmev mcmaoca on umcaasou mHmaa mmquHn pcmcmmmwu eo mqummc mmmaHmcm upsocoum11.eH.m mHamp 213 system ($1.05 and $0.82 at annual fuel cost inflation of 10% and 2%, respectively). The total costs for the biomass system are higher due to the higher investment in the CVCF furnace. The fixed costs are 68% of the total break-even costs in the case of the biomass system, and 57% in the case of the propane system (with 10% annual fuel cost inflation). There is no significant difference between the drying costs for wood chips and corncobs under the conditions presented in Tables 6.12 and 6.13. As mentioned earlier, the cost of a biomass fuel depends on the distance from which it is obtained, and the handling costs involved in drying and storage. Therefore, it should be remembered that the costs presented in Table 6.14 are affected by the transportation and handling costs, and thus, the costs would be slightly different if the transportation and the handling costs changed (see section 6.6.1). 6.7 Corn Contamination in Drying by Direct Biomass Heating In the MSU concentric vortex-cell biomass furnace, the flue gases are diluted with ambient air and passed directly through the drain. The products of incomplete combustion and fly ash are potential problems with a direct-heated drying system. The primary combustion products are carbon dioxide, carbon monoxide, ash, soot, aldehydes, ketones, sulful dioxide, oxides of nitrogen, and polycyclic aromatic hydrocarbons (PAH). The ash contains inert compounds of heavy metals. The ash and the soot particles contain a high percentage of the other combustion compounds and may be condensed and absorbed on the grain surface. 214 Table 6.15 shows the contamination data of the corn dried with the MSU CVCF/IBCF system. The following substances were detected in the corn: naphthalene (Nph), fluorine (F1), anthracene (An), benz [a] anthracene (BaA), and benzo [a] pyrene. During the steady-state operation of the MSU biomass furnace, there is no visible smoke, and only a small amount of fly ash is observed. The fly ash is filtered out of the drying air by a thin layer of corn (approximately 3 cm deep) at the bottom of the drying bin. This layer isa residue which is not transferred to the cooling bin. It is removed when the bin is cleaned at the end of the drying season. The corn dried with wood chips did not contain a measurable amount of any of the PAH before drying. However, the corn dried with corncobs did show measurable amounts of fluorine and anthracene before drying because of its growth environment, possibly close proximity to a highway (Sailor, 1978). The PAH concentrations in the biomass dried corn appear to increase slightly with time of exposure. Corn is subjected to hot drying air for a maximum of about 8-12 hours in the in-bin drying system (depending on the grain depth in the bin, the initial moisture content, and the drying temperature). The PAH concentration was in the range of 0.06 ppb (BaP) to 8.6 ppb (Nph) after a 6- to 9-hour drying period. The residual corn at the bottom of the drying bin showed a considerable higher concentration of anthracene (0.6 ppb, benz [a] anthracene (1.60 ppb), and benzo [a] pyrene (0.56 ppb) compared 215 .commom mcwxgu ms» we use ms» nu umcamHu mp :Hn any ems; um>osmc m? can .cHa mcquu ecu mo eouHon map Hm mcHasmc Hag» Ase m.nv ccoo we cmxaH :thx mm.o o.H om.o NH.o m.m 8 «mauwmma : 1. :.o 25 -1 o; 2362 11 - NH.o NH.o m.m N.N - 1- - NH.o - o.e -1 - oH.o mm.o - o.o mnou :sou oo.o mm.o NH.o NH.o m.H o.m 1- NN.o mH.o eH.o 1- e.o - - 11 1- 11 o.o mapgo too: xeeev xeeav mcmcxa mcmuacgpc< Anaav Anaav Haaav Hgv mam» max» Hana nag oNcmm mag Ncmm mcmumczuc< mcwcozpa mcmHmzusamz mucmuwmma .mma-ag use xmcmcm mmmEoHn saw: uchu ccou co mconsmuocux: umpasosm uHHoauaHoa mo :oHposucoucou-1.mH.m mHnwb 216 to the corn transferred to the cooling bin. The concentration of the BaA and BaP was less than 0.4 ppb in the corn transferred to the cool- ing bin. BaA and BaP have shown some carcinogenic activities in test animals (Souci, 1968). The range of concentrations of these compounds is well below that considered to be dangerous to human or animal health (Anderson et al., 1981). Moreover, since most of the corn dried in the USA is used for livestock, the risks to human health is minimal. The extent to which PAHs from contaminated animal feeds are transferred into meat for human consumption is not clear (Deutsche Forschungsgemein- schaft, 1976). The range of concentration values agrees with those measured by Hfitt et al. (1978) who studied corn contamination in high-temperature drying systems using a variety of fossil fuels. The PAH levels are of the same order of magnitude as those found by Joe et al. (1982) for barley drying. Wood chips and corncobs are similar biomass fuels with respect to their effect on the PAH condensation/absorption by corn. There was one exception. Corncobs resulted in a very small quantity of naphtha- lene in the dried corn, wood chips did not. Chakraborty and Long (1968) studied the formation of soot and PAHs in flames. They concluded that high temperatures in the pyrolysis zone favor the formation of carbonaceous residues and lead to a reduc- tion of PAHs. Also, PAHs decrease with increasing oxygen concentra- tion during pyrolysis, almost reaching zero at 20% excess oxygen. They also found that the optimum temperatdre fer' PHA synthesis in the pyrolysis zone is 700°C. At temperatures above 1000°C there is no 217 formation of PAHs. In the MSU biomass furnace the primary combustion zone (where pyrolysis occurs) temperature varied from 870°C to 1100°C at steady-state due to the preheating of the primary combustion air. This temperature range is above the optimum range for PAH forma- tion. Thus, the low levels of PAH contamination. Duplicate corn samples run on different days showed no detectable levels of the following metals: arsenic, cadmium, chromium, mercury, selenium, lead, and thalium. The dectection limits for these metals are given in section 5.1.2. 6.8 General Observations Some general observations on the CVCF/IBCF system are presented in this section. The adequacy of design of the major system components and maintenance requirements are discussed. 6.8.1 Fuel Feed System The fuel feed system consists of a self-unloading wagon and a feed hopper. The fuel feed system performed satisfactorily. Wood chips and corncobs without husks flow very smoothly through the system. The fuel hopper is coated with graphite-based paint. This feature, as well as the vibrator attached to the hopper, facilitates the smooth flow through the hopper. Corncobs with a large percentage of husks do not flow as smoothly as wood chips and husk-free cobs. There is no bridging of fuel at the bottom of the hopper. Backfiring through the feed hopper is effectively prevented by maintaining a slight vacuum in the combustion chamber. The vacuum is maintained by the eductor system and the drying fan. 218 The fuel feed rate is manually controlled by varying the speed of the -stoker-auger. The fuel feed rate is adjusted to maintain a specified drying temperature. When properly set, the feed rate is equal to the fuel combustion rate. There is no accumulation of unburned fuel in the combustion chamber during steady-state furnace operation. The drying air temperature fluctuates by i 5°C from the set temperature. The fuel feed rate control can be automated to respond to the drying air temperature. However, this increases the cost of the furnace without a corresponding improvement in the furnace operation. The fuel feed system requires little maintenance. 6.8. 2 Grate Performance The grate in the CVCF furnace requires periodic maintenance to remove clinker from the grate surface. The ash receptacle should be cleaned at least once every other day. Accumulation of ash is about 1 to 2 kg/h depending on the fuel type and the fuel feed rate. Most types of wood contain 0.1 to 2% ash, while corncobs contain approximately 1.6% ash (see Table 2.2). Husklage (a mixture of corncobs and husk) has a higher ash content (about 3%). The ashes should be removed daily. The formation of clinker is a serious problem in biomass furnaces. The clinker blocks the grate openings and therefore, limits the undergrate combustion air supply. Husklage produces more clinker than wood chips and corncobs. The wood chips used during the testing of the furnace produced relatively small amounts of clinker compared to 219 corncobs and husklage. The clinker should be removed at least once a week if wood chips are burned and more often when husklage is used. The firebricks randomly spaced on the grate were effective in insulating the grate from the high temperature in the combustion cham- ber. Also, the undergrate airflow cooled the grate from the underside. Thus, the grate design is satisfactory. However, the grate required replacement due to oxidation and other chemical reactions on the grate. 6.8.3 The Combustion Chamber The lower combustion chamber is lined with firebricks; the upper chamber and the inner lid are covered with a blanket of insula- tion with desirable refractory characteristics (the price of the insulation is 25% firebrick). The firebricks maintain dimensional integrity and desirable refractory and heat capacity properties without deleterious effect from biomass fuel abrasion. The insulation and the firebricks effectively protect the metal parts (the inner steel cylin- der and the inner lid) from the high combustion flame temperature (1,000-1,500°C) and the oxidizing atmosphere. The maintenance of the combustion chamber includes replacement of the tuyeres and bricks. It is not known how long these components will last. 6.8.4 Furnace-to-Dryer Duct The heat from the furnace is recovered by means of heated air blown around the furnace and mixed with the flue gases. The hot air 220 is blown into the bin plenum by a centrifugal fan via a rectangular duct. Approximately 95% of the heat loss from the system is through the duct linking the furnace to the dryer. The heat loss varies from 300 to 500 MJ/hr depending on the ambient and the drying air tempera- tures. The heat loss from the duct can be reduced considerably by insulating the duct and the bin plenum. 6.8.5 Safety Moving components of the furnace are adequately covered to prevent operator contact with injury. Carbon monoxide poisoning is a potential hazard. Exposure to the drying air inside the drying bin should be limited. Some sparks and a small quantity of fly ash reach the plenum and may result in some occasional smoldering of the collected corn dust. 6.8.6 Miscellaneous Unprotected bin and furnace components are likely to undergo accelerated corrosion due to exposure to the products of combustion. The rate of deterioration of these components is unknown. After two seasons of furnace operation, the corrosion and soot deposits in the bin interior are minimal. CHAPTER 7 SUMMARY A direct combustion concentric vortex-cell biomass furnace has been designed, built, and incorporated in an in-bin counterflow dryer (IBCF). The furnace is 1.2 m in diameter and 3.7 m high and has a maximum energy output of 690 kW. It is made of two concentric steel cylinders. The inside cylinder is lined with firebricks for insula- tion. The furnace/dryer system was tested under different ambient and drying conditions to determine the technical and economic feasi- bility as an alternative system for grain drying. The experimental results were compared with the performance of the propane fueled IBCF dryer. The results demonstrate that the concentric vortex-cell biomass furnace (CVCF) is a technically feasible alternative to fossil fuel burners for grain drying. The corn drying capacity of the experimental biomass fueled system is 2.0-3.0 wet tons per hour depending upon the drying air temperature, the grain depth, and upon the initial and the final grain moisture content. The drying capacity of the biomass fueled system is approximately 90% of the propane system. The reduction in the drying capacity is a result of the reduction in the flow rate of the drying air mass due to the heating of the air before the fan entrance. 221 222 The CVCF/IBCF system operates at about 70% efficiency in con- verting biomass fuels into energy for grain drying. Most of the heat losses (about 95% of the total) occur between the furnace and the dry- ing bin plenum. The heat loss from the furnace is about 5% of the total heat loss. This value was minimized by covering the top one- third of the furnace (1.2 m) with a shroud which reduces the radiation and convective heat loss from the furnace. Also, the lower half of the furnace is located below the ground level, thereby further reducing the low heat loss from the furnace. The average net specific energy consumption (SECO, kJ/kg of water removed) of the biomas fueled system is about 6,100 kJ/kg com- pared to an average of approximately 4,600 kJ/kg for a propane fueled system. Approximately 140 metric wet tons of corn was dried during the testing of the CVCF/IBCF drying system. The biomass furnace performed satisfactorily. There is no visible discoloration of or objectionable odor in the corn dried with the CVCF/IBCF system. Chemical analyses of the corn did not show dangerous levels of polycyclic aromatic hydro- carbons (PAHs) or heavy metal concentration; the levels detected were of the same order of magnitude as found in corn dried with propane or natural gas. The CVCF furnace is dependable. Drying air temperature control is stable (within i 5°C of the desirable temperature) and is maintained by regulating the fuel feed rate. The vibrator installed on the feed hopper prevents fuel bridging, and maintains a continuous flow of biomass fuel to the combustion chamber. 223 The grate system requires periodic maintenance to remove clinker, especially after burning corncobs. The period between cleanings depends upon the fuel feed rate and upon the ash content of the fuel. 0n the average, the grate should be cleaned at least once a week. The ash accumulation in the receptable is on the order of 1 to 2 kg/h depending on the type of fuel used and the fuel feed rate. This is due to the low ash content (0.2 to 3 percent) of wood and corncobs. The energy drying costs (exclusive of labor and fixed costs) are 30 to 40% of the energy costs of a propane system. However, the total break-even costs of the biomass system (including energy costs, labor, maintenance, depreciation, and taxes) are about 20% higher than of a propane system (at 1983 prices). The high costs are attributable to the high capital investment (approximately $16,000) required to convert a propane system into a biomass system. If the propane costs increase by about 70%, the biomass system becomes competitive with the propane fueled system. The concentric vortex-cell biomass furnace does not have an economic advantage as a grain dryer heat source at the current fossil fuel prices and interest rates. CHAPTER 8 CONCLUSIONS The main conclusions of this research are: 1. The concentric vortex-cell biomass furnace (CVCF) is a technically viable alternative to propane and natural gas burners for grain drying. 2. The CVCF furnace requires slightly more operator atten- tion than a propane system. 3. The drying air temperature is stable during the steady- state Operation of the furnace, and can be maintained at 70 to 95°C by regulating the fuel feed rate. 4. The drying capacity of the CVCF coupled to an in-bin counterflow (IBCF) grain dryer is about 90% of a propane fueled IBCF system operating at the same drying conditions. 5. The average net specific energy consumption of a CVCF/IBCF systemisabout 30% higher than of a propane fueled system. This is due to the reduced capacity and additional heat loss in a biomass system. 6. The energy costs (exclusive of labor and fixed costs) of the CVCF/IBCF system are 30 to 40% lower than the energy costs of a propane fueled IBCF system. The total break-even cost of operating a biomass system is 20% higher than of the propane system, due to the 224 225 high investment of converting a propane fueled system into a biomass fueled system. 7. The CVCF furnace does not have immediate economic feasi- bility at the current propane prices and interest rates. The propane prices have to increase by about 70% for the biomass system to become competitive with propane systems. 8. The corn dried by the CVCF/IBCF system has no visible dis- coloration or objectionable odor; and does not contain dangerous levels of carcinogenic polycyclic aromatic hydrocarbons or heavy metals. 9. The CVCF furnace model developed in this thesis accurately predicts the required biomass fuel feed rate and heat loss at differ- ent drying airflow rates, drying air temperatures, ambient temperatures, and fuel moisture contents. The model can be modified to simulate the performance of other furnace types. 10. The in-bin counterflow grain drying model presented simu- lates the intermittent action of the IBCF system. The values of the simulated parameters are close to the experimental values. CHAPTER 9 SUGGESTIONS FOR FURTHER RESEARCH The following recommendations are suggested as topics for further research: 1. Design of automatic furnace shut down in the event of: a. drying bin empty, 0' drying bin failure, c. extremely high or low drying air temperature, d. accidental fire. Determination of the optimum dimensions and configura- tion of biomass furnace eductors. Determination of the economic feasibility of insulat- ing the CVCF furnace, the duct work between the furnace and the drying bin, and the drying bin plenum. Investigation of alternative insulating materials for the combustion chamber of a concentric vortex-cell biomass furnace (CVCF). Determination of the effect of tuyere size and tuyere angle on particulate emission. Testing of other biomass materials as fuel, such as coal, apple pomice, rice hulls, etc. 226 10. 11. 227 Application of CVCF furnace to other dryer types. Adaptation of the furnace model to simulate other biomass furnaces. Dimensional analysis and similitude modeling of the CVCF and the IBCF systems. Statistical analysis to determine the variations in the experimental and the simulated values of fuel feed rate, heat loss, grain moisture content, and grain dry- ing capacity of the CVCF/IBCF system. The effect of precipitation and wind on the heat loss from the CVCF/IBCF drying system. APPENDICES 228 APPENDIX A SAMPLE 0F CVCF COMPUTER INPUT/OUTPUT DATA 229 Sample of CVCF Computer Input/Output Data. FUEL TYPE: CCRNCUES=1,MCCD=8.E. HUUDCHIFS FUEL FEED RATE,RGKHR:188. 188 88 FUEL iEIETEEE, PERCENTCNEFEE. 28.68 AMBIENT TEMFEEETEEE, C:4. 4.88 PERCENT EELETIEE HUMIDITY:7S. 75.88 DRYING EEEEEEEE, EEEI. 1.:8 EezEe- EIF, PERCENT:EE EB.88 COMBUSTIDN EIF. 1F’H: 757 7? AIR:FUEL EETIE- 7 58 T EM F‘ E RHT Ll F! E '5‘ : AMBIENT: 4.88: FEEHEETEE EEME. EIE: 123.9EE FEEHEEE FLAME: 1243.4EE FEEEILETIEN FLUE GEE: 783.54C EILETEE FLUE GEE: 181.87C DFrIvS (BIN PLENUM}: 7?.62C ULzTER FEENEEE EELL: 14E.EEE MEEE FLEEE: FLEL FEEE KG/H: 188.88 FEEL MEIETEEE, we: 28.88 FURNACE AIR SUPPLY, RGJHG-FUEL: 14.43 CMM: 19.32 EWCESS AIR, PERCENT: 50.88 PRECILI jTIDN FLUE GQS, RBEH: 1547.85 CNN: 77.14 DILUTED FLUE GAS, KGKH: 16388.67 CNN: 28?. 02 DRYING AIR, CMM: 274.11 DRYING HIR HE‘EDLUT E HLIV CITY: .8876 TOTéL HEAT OUTPUT, MJEH: 167 7.20 TCTQL HEAT LOSS, MJ./H: 44S. 95 BIN PLENUM HEAT LOSS, MJKH: 58.24 FURNACE EFFICIENCY PERCENT: 95.38 SYSTEM EFFICIENCY, PERCENT: 73.23 ASH ACCUMULATIUN, KG/H: .94 UNBURNED FUEL ACCUMULATION, RG/H: .88 230 APPENDIX B EXPERIMENTAL RESULTS 231 new zuwu_szz m>_ampmm gov.- mszmsmaeoh uncomppucou “cmanu mmmgm>< N.N m.mH mam N.HN -- EN Hm.o mmoLm>< N.N m.m~ mmm m.~o «.0 mm mm.o : HE N.N -- mmm 8.8” 8.0 m“ mo.~ o“ N.N -- omm ~.m~ 0.9 mm om.o m N.N m.mH own m.m~ N.o Nu no.0 m N.N m.mfi mHm m.o~ m.o Em mm.o N N.N -- mam m.mH o.H on mm.o m N.N -- ofim o.mH N.H mu Em.o m m.~ m.mH mom fl.- m.~ LE NE.o 8 m.~ m.m~ mom m.m m.H EL «8.0 m H.N -- HEN m.h E.H NE mm.o N m.H m.mH mum m.o m.H EN om.H H .EEE HEW. ...... ..-... ._ E”... 28...... 5% ...E. mcvzgo mgzumpoz mum; zomew< . mcwzgo .H .oz “nah "mppzmog FmpcmeLmaxmuu.H.m mpaup 232 EEE nguwe=s m>wpmpmm oom.m~ assaMLanmh ”mcowp_vcou ucmwnEm ommgm>< "wuoz 233 m.~ m.NH cum e.NH .: mm mm.o mmmLm>< m.~ H.mH oem m.mm ¢.o gm mw.o w m.~ N.eH «mm m.e~ m.o mm mm.o N ¢.N N.eH omm H.m~ n.o mm oo.H o ¢.~ m.m~ mum m.mH m.o om mm.o m ¢.~ m.m~ emm ~.mH H.H mm oo.H v ¢.N m.m~ mHm m.HH N.H mm oo.H m m.N o.o~ NHm m.m ¢.H mm mm.o N m.H o.m~ Rom m.w m.fi mm m~.H H .Ammw.wwn wmwn.w. ...... ......\.E . .mwmw. .....wm.... . ..... ...z mcwxgo mgspmmoz mum; zoPELm< . w mcwzgo mpuau mpuxu .m .oz vamp "mupammL Fmpcmspgngm--.m.m mpaME 234 won a» EEEE=E m>meme mom mgzumgmasm» "mcowuvucou acmpasm mmmgm>< "maoz N.N «Hm mu mmmLm>< N.N m.mH mam e.NH m.~ no.~ om mo.H o N.N o.m~ Hum ~.mH mN.~ om.o an oo.H m H.N :. mam e.NH m.~ No.H om so.H e o.~ e.mH «mm ~.mH N.N mn.o mm m~.H m a.“ :: cam m.eH mn.~ mm.o mu m~.~ N m.H n.mH com m.HH o.m mo.H mu me.“ ~ . . crE\mx ElcwE\ E a 00 a . ;\cou pm; a 3 u m m E guamu ; «Eva oz xupumamu pcmucou cwmgo mgmwa»masmu mpuao mpuzu mcwxgo mgzpmpoz mumL zopmgw< P o m .oz ummh "mupamog Pmpcmswgmaxu:u.m.m «Pam» 235 non zuwuwssz m>wumpmm oom.m mgaumgmaewp "meoppwucou ucmwnsm mmmLm>< ¢.N N.eH emu E.“ mm mo.H mmmLm>< ¢.N In How 0.0 mo.~ mm mH.H m ¢.N ¢.¢H mwN ¢.m MN.N mm m~.~ m m.N m.¢m mam v.o mm.H em mm.H v m.N m.¢~ mow ¢.m om.H mm Nw.o m m.N a: mom m.n N0.H om No.H N ~.m o.MH mom ©.oH NN.H. mm mw.o H FEEOE pa: .a.3 a cws\mx me-=*s\ms . u. g .85?“ .oz zupumamu ucmwcou s spamu mLsumLmaEmu mpuau m—uxu mcwzgo mgsumpoz mumL zapmgv< :PmLu mcwzgo .8 .oz pmmh "mupammL paucmswgmaxmu-.¢:m anmp 236 Ema nguwszg m>Ppmpwm oom.o: .mL:HMLmnEmE "copuwucou pcmwaEE mamgm>< "muoz o.N N.eH com mm.oN -: Nm Hm.H mmmLm>< o.N :- mHm m.oe mm.o NN om.fl o H.N H.mH Nam m.mN mm.o em mN.H m H.N -- Non m.mH o~.o Hm NH.H e o.N N.eH Nom m.mH mm.o om mN.H m o.N .. omN m.NH mo.“ Nm mo.H N m.H m.m~ HmN m.oH NN.H mm mm.H H ;\EOE um: .n.2 a cps\mx me-=ps\me s .Epamc mszmwwnsmp ; .85?» .oz ...... ...... m .oz pmmb “mapsmmL paucmE—Lmnxm--.m.m m—gmh 237 gem Euwu.s=; 8,,pmpmm oom.NH mgzumgogsap uncowuwucou pcmvnEn mmmgm>< "muoz m.H N.eH mmN HN.oH mm HN.H mmmsm>< m.H .. mwN N.mN mm.o mm mN.H m m.~ m.mH mNN m.NH NN.o Em N¢.o w o.H .: HNN o.eH mm.o Nm mm.~ N o.H N.mH NoN N.NN eo.H om mo.N o N.N :. NmN o.o~ 0N.H Hm «N.N m N.N :. NmN N.m mm.~ om mm.~ e N.N m.mH mmN N.N Hm.fi em me.“ m m.~ «.mfi omN w.m mm.H mm NM.” N m.H N.NN omN N.o mm.fl mm NN.~ a «NEE E... ...... ..-...... .. ..E. ......8 mcwzgo mgsumwoz mumL 3o—ELF< . a mcwxgo mpuxu mpuau .0 .oz ummh "mupammL Pmpcmevgmaxmuu.m.m m—nm» 238 Ema EEEEPEEE m>_p-mm oom.o mgapmgmaemp "cowawucoo pcmpneu mmmgm>< "mpoz m.H c.8H Ham m.m -- Nm mE.H mmmLo>< m.H -- com o.oH mN.o Nm mN.H N m.H o.¢H emN m.NH Nm.o Nm Nm.H m m.“ :: mwN w.oH NH.H am mm.H m m.H H.mH HmN H.m Hm.H em mm.H v m.H :. mNN m.N w¢.H Nm mm.H m m.~ H.m~ mmN m.m mo.H om mo.N N N.N m.m~ emN H.m mm.H mm Nm.H H ;\=ou pm; .5.3 E cwe\m¥ me-cre\me s .gugmc mLsugwmasou E .os_p .oz xuwuwamu pcmucoo :ngw map»: meuau mpuzo mcwxgo mLsgmmoz mum; somew< . o .N .oz ummh "mp_=mmL _mp=msmeaxmu-.N.m mpnmp APPENDIX C CVCF COMPUTER PROGRAM 239 C. C* C‘I' c. C* C* c. c* c. C* c* CI‘ c. Cfl’ C* C* C. C* c* C. C. C. Ci c* c. C* C'I' c* C. C* C. C* CVCF COMPUTER PROGRAM PROGRAM BURNER(OUTPUT,INPUT) THIS PROGRAM COMPUTES THE HEAT LOSSES FROM THE MSU CONCENTRIC VORTEX-CELL BIOMASS FURNACE. IT ALSO COMPUTES THE TEMPERATURES AND THE MASS FLOWS IN THE VARIOUS SECTIONS OF THE FURNACE. THE PROGRAM CONSISTS OF THE MAIN PROGRAM 'BURNER', SUBROUTINES 'PROPATY', 'AIRPTS', 'FLUPTY', AND THE FUNCTIONS 'RUTA', 'SIMPSON', ENTHAL, 'TABLE', HEATF', 'TADI', AND 'GMOW'. THE SUBROUTINES FUNCTIONS ARE AS FOLLOWS: PROPATY: CONTAINS TABLES OF THERMODYNAMIC PROPERTIES OF THE MAJOR GASES INVOLVED IN COMBUSTION. AIRPTS: COMPUTES THE PROPERTIES OF AIR AT A GIVEN TEMPERATURE CALLS THE FUNCTION TABLE WHICH INTERPORATES THE DATA CONTAINED IN THE SUBROUTINE PROPATY. FLUPTY: COMPUTES THE DYNAMIC VISCOSITY AND THERMAL CONDUCTIVITY OF FLUE GASES. RUTA: SOLVES FOR THE JACKET INNER.WALL TEMPERATURE BY DOING THE RADIATIVE, CONVECTIVE, AND CONDUCTIVE HEAT BALANCE. SIMPSON: COMPUTES THE ENTHALPY CHANGE OF FLUE GASES AT A GIVEN TEMPERATURE BY THE SIMPSON INTERGRATION OF THE SPECIFIC HEAT FUNCTIONS IN THE SUBROUTINE ENTHAL ENTHAL: CONTAINS SPECIFIC HEAT EQUATIONS FOR GASES. TABLE: INTERPORATES THE TABULATED DATA. HEATF: DETERMINES THE HEAT REQUIRED TO HEAT THE AIR AND THE FUEL TO THE COMBUSTION TEMPERATURE ABOVE 25C. TADI: DETERMINES THE ADIABATIC FLAME TEMPERATURE GIVEN THE FUEL HEATING VALUE, THE INITIAL FUEL AND THE AIR TEMPERATURES, AND THE PRODUCTS OF COMBUSTION. GMOW: COMPUTES THE MOLECULAR WEIGHT, SPECIFIC HEAT, AND THE TOTAL MASS OF FLUE GAS. COMMON/FRACTNS/COM(20),GAS(20),WMOL(20),CP(20),CM2,CM3,CM4,CMO COMMON/PROPETY/AK(30),ACP(30),APR(30),ADEN(30),AV(30),CFMF(10) COMMON/FLUE/DVCOZ(20),CKCOZ(20),DVCO(20),CKCO(ZO),DVH20(20), +CKH20(20),DVH2(20),CRH2(20),DVN2(20),CKN2(20),DVOZ(20),CKOZ(20) COMMON/SMS/SMLA,SMLFLU,DIFA,DIFLU,KA,KFLU,SMST,DIFST,KST,COMED(10) cmn/mssnam DIMENSION NAME(20),A(20),R(30),X(50),TFUR(50),VISH(20):QL(20) DIMENSION GASHI(10) 240 C. c. C. C. c. C. C. C. 33 43 c. C. c. 80 241 EXTERNAL HEATF,FLUPTY,TABLE,PROPATY,RUTA DATA PATM/14.696/ DATA C,H,OXY,CN,SUL,ASH/46.83,5.9S,45.025,.42,0.,1.6/ DATA FUEL,XMFW,TAMB,RH,PFF,BINP/90.,12.l,5.79,70,2.125,3.44/ THE A'S ARE THE AREAS OF THE VARIOUS SECTIONS OF THE FURNACE. DATA (A(L),L=l,6)/.732,10.167,1.167,14.009,2.627,14.009/ THE R'S ARE THE THERMAL RESISTANCES ACROSS VARIOUS SECTIONS OF THE FURNACE INCLUDING CONVECTIVE, RADIATIVE AND CONDUCTIVE THEMAL RESISTANCES. DATA (R(M),M=1,18)/.527,15.814,.978,S.,.312,2.385,.286, +7.88,9.362,4.029,.125,.125,3.75,15.228,1.0075,.157,.15,.008/ DATA AAO,AA2,AA3,AA4,AP/.09931,.00456,.0324,.0243,1.459/ DATA (WMOL(K),K=l,B)/44.01,28.0l,28.016,46.,30.,18.,32.,2./ TRK(R)=R/l.8 THR(C)=1.8*C TKF(Z)=1.8*Z-459.67 TFK(Z)'(Z+459.67)/1.8 CALL PROPATY Is THE PRACTIONAL MOLECULAR COMPOSITION OP COMPONENT I IN TEE PLUE GAS. QNA=COM(1)+COM(3)+COM(6)+COM(7) DO 11 NsI,7 IP(COM(N) .EQ. o.)GOTO 11 GAS(N)=COM(N)/QNA CONTINUE COMED(1)-COM(1) COMED<3>=.7899*CMOML+CC/2*PUELD/Ioo. COMED(6)-xMDa*PUELD/Ia.OIS+I.6081*CMOML*N+EE/2.*PUELD/Ioo. COMED(7)-.2101*ExT2 ONA2=COMED+COMED<6)+COMED(7) DO 39 N=I,7 IP(COMED(N) .Eo. o.)GOTO 39 GASED=COMED/ONA2 CONTINUE AIRMAss-CMz KCYCLE=1 TADR IS THE PIRST ESTIMATE OP ADIAEATIC PLAME TEMPERATURE AS GIVEN BY TILMAN ET AL. (1981). TAD2 IS THE ADIAEATIC PLAME TEMPERATURE TAKING INTO ACCOUNT THE REACTANTS INITIAL TEMPERATURE AND THE HEAT REQUIRED POR PRENEATING THE REACTANTS (FUEL AND AIR). TADR=1920.-2.*xMDE*IOO.-s.2*AES(ExCS) TADEs-TADN 1300 CONTINUE C. C. C. 15 244 IF(KCYCLE .EQ. I)CPA=TAELE(ACP,SMLA,DIPA,NA,TPRE) CON3=NEATP CONRAsGAS(J)/Ioo.*PNI+CONRA CONTINUE COND-PLUPTT(TP,I,2) 245 CONDMINsCONDMIN+ EEZeSE*T2**4 TP7=.2449*TPP-259.41 T7=TE7=TPK TsaCMG*CPG*TP+CMED*TEN T8=T8/(CMG*CPG+CMED) EBB=SB*T8**4 10 CONTINUE TPM= IP(NIT .EO. O)T67¢T7 TDI-T67-TAME TDI-AES(TD1) TD2-T57-TJAR TD2=AES TD3-TD1-TD2 TDB-ABS(TD3) TLM-TD3/ALOG(TD1/TD2) 20 23 1401 246 DE*.254 CALL AIRPTS(TLM,DEN,SRS,VIS,PRN,CPS) H67=SKS*.023/DE*(DE*V67/VIS)**.8*PRN**(l./3.) QG7=12.842*H67*(T7-TLM) EBA=SB*TAMB**4 QS7=QCO7+Q67-EBA/R(17) QC6=Q67*1.5 IF(KIT .EQ. 1)QC6=10.16667*H67*(T6-TLM) Q567=QC6+QS7*R(l7)/(R(16)+R(l7)) RSl=R(l7)/(R(16)*(R(16)+R(l7)))-1./R(16) IF(KIT .EQ. 0)T6=T7*1.5 K82 X(1)=T6*.925 X(2)=T6*l.5 CONTINUE 38K+1 FX1=RUTA(QSG7,RSI,CF,TF,R,X(H-2)) FX2=RUTA(QSG7,RSl,CF,TF,R,X(K-l)) X(K)=(FX2*X(K-2)-FX1*X(K-l)>/(FX2-FX1) ERROR=ABS(X(K)-X(K-l))/2. IF(ERROR .LE. 1E-10)GOTO 23 GOTO 20 CONTINUE T68ABS(X(K)) KITSKIT+1 EBG=SB*T6**4 TW1=(Q567-EBG*RSI+T6/R(18))*R(18) EB7=T7**4.*SB IF(T7 .LT. TE7)T7=TE7 T67'(T6+T7)/2. IF(KIT .EQ. 1)GOTO 10 QR7=.75*(EB7-EBA)/R(l7)+.25*(EB7-EBS)/R(l7) TF7-THF(T7) AA4=1.1675 TF481.0605*TF7+15.3076 TAF=.OB*TKF(T8)+26. V4'2.413 T4-TFK(TF4) EBGISB*T4**4 TA-TFMTAF) DEI.61 TW‘(T4+TAMB)/2. TDIITO-TAMB TDlIABS(TD1) KEYl=0 CONTINUE TDZ-T4-TA TDZSABS(TD2) DTM=(TDl-TD2)/ALOG(TD1/TD2) CALL AIRPTS TD2=T4-TIN4 TD2=AES DTM=(TDl-TD2)/ALOG(TDl/TD2) CALL AIRPTS+1/R(7)+1/R(8)+1/R(10) RT5=1/)+1/R<8) RT8=1/(RT1*R(10) *R( 4) )+1/R(6) RT9=RT8-RT6*RT7/RT5 DRIzEEII/R(I)+D34/R(IO> DN2=-(E32/R(3)+DJ4/R(8)> DN3=DJ4*RT4-EE4/R(7) DR4=DN2-DxL/(RTI*R(2)) DR5-DN3-DN1/(RTI*R(IO)) DJ3=(DK5-DK4*RT7/RTS)/RT9 DJZ=(DK4-DJ3*RTG)/RTS DJI= GR:.2775*EETA*AES(Ts-T13>/VIS**2 C* EOUTIONS 8-74 TO 8-76 GEEEART POR PREE CONVECTION C* BETWEEN HORIZONTAL PLATES, BOTTOM NOTTER. HNU=.068*GR**(1./3.) IF(GR .LT. IE+5>HNU=.I95*GR**.25 IP(GR .LT. IE+4)NNU=I. HSIHNU*SKS/ . 3048 AA3-1.0681 913=AA3*1.5*H3*(T3-T13) CALL AIRPTS RT13-R(11)/(R<13)*R(14))+RT12 DK6=DJS*RT10-EBB/R(5) DK7=-(Ene/R(15)+DJ5/R(13)) DN8=DN7-DN6*R(11)/R(14) DJ7s-Dx8/RT13 DJ6=(DK6-DJ7/R(13))*R(11) EBS=ABS(DJ6*RTll-DJ5/R(11)-DJ7/R(14))*R(12) Ts=(EES/SE)**.25 T5=(TA*5+T5)/6. IP(TS .LE. TAME)T5=TA 0T03=AT*NT*(T5-TAP) TA-((OT3+QTD3)*.06+CM3*CPS*TAME)/(CM3*CPS) KEY1=KEY1+1 IF(KEYl .LT. 2)GOTO 1401 TP5=TKF(T5) AT-2.6268 ORTsSE*AT*(T5**4—T5RT**4) TTP=(T5+TAME)/2. CALL AIRPTS(TTF,DEN,SKS,VIS,PRN,CPS) IF(VWIND .LT. 3.)GOTO 45 RE-VWIND*1.6/VIS HT5=SKS/l.6*.5*RE**.5*PRN**(1./3.) GOTO 47 CONTINUE EETA=1./TTP GR-43.6955*EETA*(T5—TAME)/VIS**2 RA-PRN*GR Ev3=7. ETSsSNS/1.6459*EV3*RA**.2 CONTINUE X(KCYCLE)-TF KCYCLE=KCYCLE+1 OCTuAT*ETs*(T5-TAME) QLT=QR7+QRT+QCT+OCO7+QT3+OTD3 QLN=(OR7+QRT+OCT)*3.6 IP(NCICLE .GE. 10)GOTO 1310 CALL AIRPTS(TA,DEN,SKS,VIS,PRN,CPS) QLPT=QLT+CM2*CPS*(TJAN-TAME)*60. QEVApstDE*PUELD*1.353E+s QLFT=QLPT+QEVAP TP-(CPG*CMC*60.*TAnz-OLPT)/(60.*CPG*CMG) x(NCTCLE)-TP ERR-ASS(x(xCICLE)-x(NCICLE-1))/2. IP(ERR .LE. .05>GOTO 1310 GOTO 1300 1310 CONTINUE C. c. c. C. C. C. c. c. 439 433 c. 249 QINPUT=HHV*FUELD*60. FUEP=<1.-QLN/QINPUT)*100. GM=GMOW=COMED<7>+.2101*CMOL3 CMD=CPD=0. QMD,CPD,CMD ARE THE TOTAL MOLES, SPECIFIC HEAT A MASS OF THE DILLUTED PLUE GASES. ' OMD=COM<1)+COM(3)+COM(6)+COM(7) GM=GMOWGOTO 1129 DUCT HEAT LOSS TH=TXT TFILM= COMPUTE DATA FOR PRINTING HHVW=QINPUT/FUEL/1000. HHVD=HHV/IOOO. ASHF=ASH/100. NKGLE=2.204622 ASN=ASH*FUELD*60./100. ASHLE=ASH*WKGLE TOC=TAME—273.15 TOF=TKF TPC=TJAK-273.15 TFF=TKF TFC=TF-273.15 TPP=TKF(TF) TXTC=TXT-273.15 =TKP(TXT) TEDC=T8-273.15 TEDF=TKP(T8) TENC=TEP-273.1s TWC=T7~273.15 TWPeTKF(T7> PRINT 150,TOC,TOF,TFC,TFF,TFC,TFF,TEDC,TEDF,TxTC,TXTF, +TBNC,TBIN,TWC,TWF FULB=FUEL*WKGLB*60. VWIND=VWIND*3.6 QINPUT=QINPUT/1000. WM?H8VWIND*5./8. CMD'CMD/FUEL CHMCBPCFM*.02837 PDPG=(CM2+CMED+FUEL*(1.-ASEP))*60. PDLBBPDPG'WKGLB CALL AIRPTS(T8,DEN,SKS,VIS,PRN,CPS) CMMED-PDFG/DEN/GO. CPMEDICMMED*35.3134 CMDLB=CMD*WKGLB*60. FUEL‘FUEL*60. CALL AIRPTSCTXT,DEN,SKS,VIS,PRN,CPS) CHMDBCMD/DEN CPMDSCMMD*35.3134 CMDBCMD'GO. BTUM‘QINPUT'9.4783E-4 QLH‘QLH/IOOO. QLBTUBQLB’9.4783E-l PRINT 25,FUEL,PULB,XHFW,CHO,CMMC,FCFM,EXCS,PDPG,PDLB, +CHHED,CPMED,CMD,CHDLB,CHHD,CFHD,DAIRV,CFHT,HIP PRINT 37,QINPUT,BTUM,QLH,QLBTU,FUEP,SYEP,ASH,ASHLB 91 111 109 25 ca 37 1400 200 83 329 93 899 900 252 FNLB=FUELNB*NNGLB IF(FUELNB .GT. 0.)PRINT 1400,FUELNB,FWLB CONTINUE PRINT 109 READ 329,TERM IF(TERM .GE. 1.)GOT0 43 FORMAT(* *,*EXCESS AIR, PERCENT:*) - FORMAT(*0*,/,* *,*CONTINUE? YES=1.; NO=0.*) FORMAT(* *,*PERCENT RELATIVE HUMIDITY:*) FORMAT(* *,*FUEL MOISTURE, PERCENT(NB)*) FORMAT(* *,*AMBIENT TEMPERATURE, C:*) FORMAT(* *,*FUEL FEED RATE,xG/HR:*) FORMAT(* *,*DRYING PRESSURE, RPA*) FORMAT(*1*,*FUEL TYPE: CORNCOBS=1,w00D-0.*) FORMAT(*O*,/*O*,“TEMPERATURES:*, +/* * * AMBIENT:*,23x,FB.2,*C *,F8.2,*F*, +/* * * PREHEATED COMB. AIR:*,11x,F8.2,*C *,F8.2,*F*, +/* * * FURNACE FLAME:*,17x,FB.2,*C *,P8.2,*P*, +/* * * PREDILUTION FLUE GAS:*,10x,F8.2,*C *,F8.2,*F*, +/* * * DILUTED FLUE GAs:*,14x,FB.2,*C *,F8.2,*F*, +/* * * DRYING (BIN PLENUM):*,11X,F8.2,*C *,F8.2,*P*, +/* *,* OUTER FURNACE WALL:* 12x,F8.2,*C *,F8.2,*F*) FORMAT(*0*,*MASS FLOWS:*, +/* *,* WIND SPEED, RPB:*,1sx,FB.2,sx,FB.2,* MPE*, +/* *,* FUEL FEED, KG/H:*,15x,F8.2,5X,F8.2,* LB/H*, +/* *,* FUEL MOISTURE, WB:* ,12x, F8. 2, +/- *, * FURNACE AIR SUPPLY, NG/NG-FUEL:* ,FB.2, +/* * ,29x,*CMM:*,FB. 2, 5x, FB. 2,* CFM*, +/* *,* EXCESS AIR, PERCENT. * ,11x, F8. 2, +/* *,* PREDILUTION FLUE GAS, KG/N:*,4x,FB.2,5x,FB.2,* LB/N*, +/* *,24x,*CMM:*,5x,F8.2,sx,FB.2,* CFM*, +/* *,* DILUTED FLUE GAS, NG/H:*,Bx,F8.2,sx,FB.2,* LB/H*, +/* *,20x,*CHM:*,9X,F8.2,5X,F8.2,* CFM*, +/* *,* DRTING AIR, CMM:*,15x,FB.2,5x,FB.2,* CFM*, +/* *,* DRTING AIR ABSOLUTE HUMIDITI:*,4x,FB.4) FORMAT(*0*,*TOTAL HEAT OUTPUT, MJ/N:*,9x,FB.2,sx,FB.2,* MBTU/E*, +/* *,*T0TAL MEAT LOSS, MJ/B:*,11x,F8.2,5x,Fa.2,* MBTU/N*, +/* *,*FURNACE EFFICIENCY, PERCENT:*,5X,P8.2, +/* *,*SISTEM EFFICIENCY, PERCENT:*,6x,FB.2, +/* *,*ASH ACCUMULATION, KG/H:*,10X,F8.2,5X,P8.2,* LB/H*) FORMAT(* *,*UNBURNED FUEL ACCUMULATION, NG/N:*,FB.2,5x, +P8.2,* LB/H*) FORMAT<1E+,35X,F7.2> FORMAT DIMENSION R(20) SB-s.669E-8 ..‘Q‘Q 253 EBG‘SB*T**4 QUOD=(QS-EBG*RS+T/R(18))*R(18) RUTA=SB*QUOD**4/R(19)+QUOD*(1./CP+1./R(18)) RUTABRUTA-(T/R(18)+SB*TP**4/R(19)*TF/CF RETURN » END c....................... 10 20 FUNCTION SIMPSON(TEM,REACT,GM,GN,C0M,CP) DIMENSION C0M(20),CP(20),WMOL(20) DATA (NMOL(R),x=1,8)/44.01,28.01,28.016,46.,30.,18.,32.,2./ MEATso. D0 20 L=l,7 IF(C0M(L) .EQ. 0.)GOT0 20 N=20 M=(TEM-298.)/N HALF-H/Z. NNsN-1 CM=COM(L)*WMOL(L) SN1=ENTNAL(298.,REACT,GM,GN,C0M,CP,L) SNZ=ENTHAL zERO(A.B)-A-B Bc-ENv-CON x-z x<1>-TEST-600. x(2)s‘1‘ES'r+100. NMcGM0w(COM,TEST,CPG,CMG,GN) CONTINUE R-x+1 EN1-SIMPSON(X(x-2),2. ,NM,GN,COM,CP) Fx1-zER0(EN1,BC) EN2-SIMPSON(x(N-1),2.,NM,GN,COM,CP) Fx2-zER0(EN2,BC) x(E>-(Fx2*x(N-2)-Fx1*N(N-1))/(Fx2-Fx1) ERR-ABS(X(K)-X(K-1))/2. IP(ERR .LE. 1.E-10)GOTO 10 10 10 257 GOTO 5 CONTINUE TADI=X(K) WM=GMOW PRINT 201,1NAME,IPRODU,IPRODU1,IEMC READ 200,TIN PRINT 199,TIN PRINT 190 ‘ READ 200,TAMB PRINT 199,TAMB PRINT 202 READ 200,RBAMB PRINT 199,RHAMB MIsMADBRH.RNCN) CONTINUE PRINT 207 261 READ 200,XMF PRINT 199,XMF XMF=XMF/100. IF(COOLER .EQ. 1. .AND. HOUR .NE. 0.)GOTO 1307 PRINT 208 READ 200,DEPTH PRINT 199,DEPTH 1307 CONTINUE IF(COOLER .NE. 1.)PRINT 203 IF(COOLER .NE. 1.)READ 200,8CFMF IF(COOLER .NE. 1.)PRINT 199,8CFMF IF(COOLER .EQ. 1.)BCFMF=30. IF(COOLER .EQ. 1. .AND. DEPTH .GT. 8.)BCFMF=20. CALL SECANT(DEPTH,BCFMF,CFM,PG,COOLER) C***** PRINT HEADER PAGE OF CONDITIONS AND PROPERTIES GA'60.*CFM/VSDBHA(F(TIN),HIN) PRINT 215,CFM,GA,XMO,PG IF(COOLER .EQ. 0.)PRINT 65 IF(COOLER .EQ. O.)READ 200, ANSWER IF(COOLER .EQ. 0.)PRINT 199,ANSWER IF(COOLER .EQ. 1.)ANSWER=1. IF(ANSWER .LE. 0.)GOTO 1307 PRINT 218 READ 200,TIME PRINT 199,TIME PRINT 209 READ 200,DBTPR PRINT 199,DBTPR HRPT=10. IF(SYSTEM .EQ. 2.)PRINT 49 IF(SYSTEM .EQ. 2.)READ 200,HRPT IF(SYSTEM .EQ. 2.)PRINT 199,HRPT DELPR=DBTPR ' IF(COOLER .NE. 1.)PRINT 212 IF(COOLER .NE. 1.)READ 200,HDF IF(COOLER .EQ. 0.)PRINT 199,EDF IF(COOLER .EQ. 1.)HDF=.75 C***** COMPUTE STEP AND ARRAY SIZES DELX=.25 IF(SYSTEM .LT. 2.)DELXIBPC*1.24/508.938 IND=INT(DEPTH/DELX) INDIBIND+1 INDII-IND1+1 C***** COMPUTE INLET RH AND INITIALIZE ALL ARRA! POSITIONS NECESSARY RHIN=RHDBHA(F(TIN).HIN) RHT=RHIN DO 1 I31,IND11 XM(I)‘XMO H(I,1)'HINIT T(I,1)=TIN TE(I,1)'THIN RH(I)'RHCN c... c... 1400 25 160 74 75 45 65 1300 77 49 170 190 198 199 200 201 202 262 1 CONTINUE 8(101)‘HI RH(1)-RHIN ** CONVERT AIRFLOW TO LB/HR IF(GA .GE. 500.)NC=.363*GA**.59 IF(GA .LT. 500)HC-.69*GA**.49 CON1=2.*GA*CA CON2=2.*GA*CV CON3=HC*SA*DELX CON4=RHOP*CP CON5=RHOP*CW DELT=2.*DELX*(CON4+CONS*XM(1))/(CON1+CON2*H(IND1,1))*.9 IF(COOLER .EQ. 1.)DELT=1. IF(COOLER .EQ. 1.)EOUR-0. ** CALL IBCFLW FOR SIMULATING IN-BIN COUNTERFLOW DRTING CONTINUE CALL IBCFLW4(TAMB,HRPT,TDAVE,XMDAVE) IF(COOLER .EQ. 1.)GOT0 25 PRINT 1300 READ 200,COOLER PRINT 199,C00LER IF(COOLER .EQ. 0.)GOTO 25 IP(SYSTEM .EQ. 2.)GOT0 1400 RHCN=.9*ERH(TDAVE.XMD) NINIT=EADBRE,RNCN) TIN=TAMB TEIN=TDAVE MIsEINaMADBRE,RHAMB) SISTEM=2. GOTO 1310 PRINT 260 DEP1=0. FORMAT<*0*,*COUNTERFLOW? N0=2; TES=1., REFILL-0.*) F0RMAT(*0*,*TIME BETWEEN REFILLS, BOURs:*) FORMAT(* *,*BUSHELS PER REFILL:*) F0RMAT(*0*,*BUSHELS PER CYCLB:*) FORMAT(*O*,*SATISPIED WITH DEPTH/PRESSURE COMBINATION?*, +/* *,*FEs-1.; NO=O.*) F0RMAT(*0*,*COOL THE GRAIN? IE5-1., N0-0.*) F0RMAT(*0*,5x,*ABSOLUTE HUMIDITY,DECIMAL:*) FORMAT(*0*,5x,*OUTPUT INTERVAL: HOUR;*) F0RMAT(1x*EQTN FOR LOW TEMP;(1 RUG,2 MISRA,3 SABBAH):*) F0RMAT(5x*AMBIENT AIR TEMP, F :*) F0RMAT(I1) F0RMAT(1N+,40x,F10.4) FORMAT(F10.2> FORMAT(*0COUNTERFL0W GRAIN DRIER SIMULATION*/ +* USING THE *A10,A10* EQUATION FOR *A10 / +* AND EMC BY *A10// . +* INPUT CONDITIONS :*/ +sx*DRTING AIR TEMP, F :*) FORMAT(SX*AMBIENT REL HUM, DEC :* 203 204 205 207 208 209 212 213 215 218 230 240 250 260 300 c.. 263 F0RMAT(*OFINES FACTOR; 1—2(1 CLEAN, 2 DIRTY):*) F0RMAT(sx*INLET GRAIN TEMP, F:*) F0RMAT(5x*INITIAL MOISTURE, W.B.PERC.:*) .FORMAT(sx*FINAL MOISTURE,W.B.PERC.:* F0RMAT(5x*BED DEPTE,FT:*) F0RMAT(5x*OUTPUT INTERVAL,FT:*) F0RMAT(5x*NYBRID DRYING FACTOR,DEC.:* F0RMAT(///1x*FAN Is DOWN FOR, :*) F0RMAT(//* PRELIMINARY CALCULATED VALUEs*// +* AIRFLOW, CFM/SQ FT *FB.4/ +* DRY AIRFLOW RATE, LB/MR-FTz *FB.4/ +* INLET MC(DRY BASIS DECIMAL) *FB.4,/ +* PLENUM PRESSURE, IN-HZO: *,F8.4) FORMAT<5x*MAx.DRYING TIME, HR:*) F0RMAT(sx*TYPE 0F FUEL USED (1=NO.2 FUEL*/ +5X,*2=NAT.GAS; 3-L.P.GAS; 4=BIOMASS):*) FORMAT(SX,*CALCULATED AMBIENT ABS NUM=*9x,F10.4) F0RMAT(5x*CALCULATED INLET ABS NUM=*11x,F10.4) F0RMAT(5x,*TNIS IS THE END OF COUNTERFLOW*) FORMAT(5X*DRYING AIR TEMP, F :*) END c * I...B-C.888888=388338. “m8 SUBROUTINE IDCFLW4 (TAMB, HRPT, TDAVE , XMDAVE) C * .I'muiulu'38-‘88-“888'38888338888388I...“ c . . . . . DIMENSION RHN(50),CYCLE(50),TDRY(40),xMDRY(40),X(50) COMMON/MAIN/XMT,THT,RHT,DELT,CFM,XMO,KAB,TOTEN,TOTHZO,XMS, +TOTSP,IPROD,FM,NDF,xMF,TIME,NEQ,NE01,NOUR,REFHR,REFBU,TOFENY COMMON/INPT/BPH,GP,IND1,DELX,DBPTH,DBTPR,XTEMPER COMMON/PRPRTY/SA,CA,CV,CW,RNOP,CP COMMON/ARRAYS/MGO).RH(50),'1‘(50,2),H(50,2),TH(SO,2),GA COMMON/HLATENT/HA,NB,NFG COMMON/IFLAGS/K,JM,ICON,DELXM COMMON/PREss/PATM,BCFMF,PG COMMON/COUNT/ITERCT,TIN,THIN,HIN COMMON/HRS/HR COMMON/COLD/COOLER,SYSTEM EXTERNAL ZEROIN EXTERNAL SOLVE4 DATA PATM/14.696/ DATA RHC,AREA/.999999,254.469/ F(T)-T+459.69 ERHCI'JIM) II 1. - EXP(- .38195*('r + 50.)*XM**2) C..... c..... IF(COOLER .EQ. 1. )REFHR’O . IF(REFHR .GT . 0 .)SYSTEM=1 . I'IRPTS'BO . EMBIN=XDD DRYDUT‘O . PWATER‘PBU'O . 264 GLEVEL1=DEPTH HZOOUTao. YL=0.0 8380. CTIME=25./60. KKIO CYCLE1=0. KBU=O RAB=0 REALT=0. RC0N=0 DRM=0. DEEP=REFIN-0. ITERCT=0 NODESsINT(.3/DELX)+1 RAB=0 IND11=IND1+1 IND2=IND1 TOTBTUW=0.0 TINN=TIN=T(1,1) RNINeRHDBNA,HIN) DPCN=2.*DELx xMBTstO c..... C***** BEGIN TIME LOOP C..... 40 YL'YL+DELT IF(CYCLEl .GT. 0. .OR. SYSTEM .GE. 2.)REALT=.1 IF(COOLER .EQ. 1.)REALT=0. HR‘HR+DELT+REALT HOUR=HOUR+REALT+DELT CYCLE2=HOUR HRPTS'HRPTS+DELT+REALT REFINBREFIN+DELT+REALT C..... C***** COMPUTE MC FOR DEPTHaO cut... DEP=DEP1=0. TN(1,1)-(14.*TE(1,1)+T(1,1))/15. IF(SYSTEM .EQ. 2.)TN(1,1)-(1.5*T(1,1)+TN(1,1))/2.s THT=T8(1,1) XMT=XM(1) RHT=RHIN CALL LAYEQ XM(1)-xMT H<1.2)-H(1.1) T(1,2)-TINN TU(1.2)-TN(1,1) RNR<1>=RNT*100. c..... C***** BEGIN DEPTH LOOP C..... 265 382 XMVBO. 102 CONTINUE K33 JM‘J-l TH RHN(LM)=RH(LM)*100. 105 CONTINUE ITERCT=ITERCT+1 c***** C***** CHECK IF LONG ENOUGH OR DRY ENOUGH OR TIME TO SAVE VALUES C***** FOR PRINTING. IF NONE OF THESE GO TO THE BEGINNING OF THE LOOP Cttttt IF(RBU .EQ. O)XMBOT=(XM(1)+XM(2))/2. IF(KBU .EQ. 0.)GTEMP=(TH(3,1)+TH(2,1)+TH<1,1))/3. XMBTM=XMBOT/(1.+XMBOT) xMBWBstBTM*100. C... CFMTOT=CFM*AREA xMV=xMV/FL0AT(IND1-1) m-XMV/(l. +xMV) IF(SYSTEM .EQ. 2. .AND. HRPTS .GE. HRPT>GOTO 400 IF(HOUR .GE. TIME)GOT0 400 IF(xMV. LE. xMF) GO TO 400 IF(SYSTEM .EQ. 2.)GOTO 40 IF(KBU .GT. 0)GOTO 400 IF(xMBTM .LE. XMF)GOTO 400 GOTO 40 c. 400 CONTINUE C* SECTION SHIFTS ARRAYS BY DEPTH MODES REMOVED Ct IF(SYSTEM .GE. 2.)GOTO 12 IF(KBU .GT. O)GOTO 5 IP(XMBTM .GT. EMF)GOT0 12 s IF(SYSTEM .EQ. 1.)CALL VARYD/2. C * C* TERMINATING CONDITIONS C * xMFs=xM5/100. IF(XMFS .LE. XMF)GOTO 73 IF(DEPTH .LE. .995)GOTO 73 IF(DEPTH .GE. 12. .AND. COOLER .NE. 1.)PRINT 193 193 FORMAT(*0*.*BIN 15 FULL, STOP PILLING!*) IF(DEPTH .GE. 12. .AND. COOLER .NE. 1.)GOTO 73 IF(HOUR .GE. TIME)GOTO 73 GOTO 40 73 CONTINUE IF(SYSTEM .GE. 2.)RETURN IND=INDI-1 SUMT=SUMM=0. DO 55 J=1.KK SUMT=SUMT+TDRY SUMM=SUMM+HMDRY 55 CONTINUE TDAVE=SUMT/xx XMDAVE=SUMM/KK DEPTH=2.*DELx*xH IF(DEPTH .GT. 12.)DEPTH=12. xMWD=xMDAVE/(1.+xMDAVE)*100. XMO=XMDAVE PRINT 1370,xMWD,TDAVE 1370 FORMAT(//* FINAL AVERAGE MOISTURE:*,1Bx,F12.4/ +* FINAL GRAIN TEMPERATURE:*.17X.F12.4) RETURN END c . t c * Ins:smalmzss-unmmssmswusssnaacuse SUBROUTINE CRSPR(NMAVE.TAVE.THAVE.XMEINW.XMEOUTW.RHK) C * 8‘883B88II..8ICIMISIWC:=HW c. . . . . DIMENSION RHK(50).NMW(50) COMMON/ARRAYS/XM(50),RH(50).T(50.2).N(50.2).TH(50.2).GA COMMON/MAIN/DUM(5).XMO.KAB.SKIP(4).IPROD COMMON/INPT/BPH.GP.INDI.DELX.DEPTH.DBTPR.XTEMPER C..... INDBINDl-l DBPR=DEP=0. SUM=SUMT*SUMTH=0.0 IF(DBTPR .GT. DEPTH)GOTO 5 PRINT 50 269 D0 10 J=1.IND1 DEPsDEP+DELx DBPR=DBPR+DELX XMW(J)=XM(J)/(l.+xM(J))*100. . IF(DBPR .GE. DBTPR)PRINT 100,DEP,HMW(J),TH(J,1),T(J,1),H(J,1),RHH +J) IF(DBPR .GE. DBTPR)DBPR-O 10 CONTINUE 5 CONTINUE DO 20 1:2.IND1 SUMT=SUMT+T(I,1) SUMTH=SUMTH+TH(I.1) 20 SUMtSUM+XM(I) XMAVE=SUM/FLOAT(IND1-1) TAVE=SUMT/FLOAT(IND1-1) THAVE=SUMTH/FLOAT(INDl-l) XMEINsEMC(RH(1),TH(1.1)) xMEOUT=EMC(RH(IN01),TH(INDl,1)) HMEINW=HMEIN/(1.+HMEIN)*100. XMEOUTW=XMEOUT/(l.+XMEOUT)*100. RETURN 50 FORMAT(//* DEPTH M c PROD.TEMP AIR TEMP AIR U.R. REL.HUM.*) 100 FORMAT(1X,F5.2,PB.2.F9.2.P9.2,F9.S.P10.2) END c.. C*.-.88.:sxssssxsssxsaclns:In:ascsaszasxasnssxssllac SUBROUTINE DATE(IPROD) C*I un=s=ssnaclcx C***** SUBROUTINE USED FOR INITIALIZING CONSTANTS FOR PRODUCTS COMMON/PRPRTY/ SA.CA.CV.CW.RHOP.CP COMMON/HLATENT/HA.HB th COMMON /NAME/INAME.IPRODU.IPRODU1.IEMC DATA INAME.IPRODU.IPRODU1.IEMC/IONTROETHOM .IOHTHINLAYER . +108 CORN .108 DEBOER / cattwt C***** INITIALIZE CONSTANTS FOR CORN SA'239. CA'0.242 CV30.45 CW=1.0 RHOP'38.71 HA-4.349 BBB-28.25 CP'.268 RETURN END c.. c * cuss-“ucsmmuumusmn-uc SUBROUTINE ABSN(NIN.TAMB.TIN.NI.FUEL.IPROD) C*833:.38:28.38sass:ass-ssssssssszsscsscsss3.338.383 cttata 270 C***** FUEL=1 STANDS FOR No.2 FUEL C***** FUEL=2 STANDS FOR NATURAL GAS C***** FUEL=3 STANDS FOR LIQUID PROPANE GAS C . . . . . IFUELsIFIx(FUEL) IF(IFUEL.EQ.1) A=7.0143E—5 IF(FUEL.EQ.2) A=8.17SE-5 IF(IFUEL.EQ.3) A=7.593E—5 IF(IPROD.EQ.2) CP=0.400 IF(IPROD.EQ.1) CP-0.268 HFUEL=A*(1.+HI)*CP*(TIN-TAMB) HIN=HI+HFUEL FUEL=FLOAT IP(RH(J)-RHC) 150,200,200 c..... C***** CONDENSATION SIMULATOR c..... c..... 200 CONTINUE TS¢T(J:2) HS=HADBRH(F(T(J,2)),RHC) H(J,2)-HS DHDTBHS-HADBRH(F(T(J,2))-l.,RHC) AIGA'CV BR(GA*CA+CCON2)-DHDT*GA*CV*(T(JM,2)-TS)-HS*GA*CV C=(GA*CA+CCON2)*(DHDT*TS-HS)-DHDT*(GA'CA*T(JM,2)+CCON2*THETA) 274 E(J,2)=(-E+(E*E—4.*A*C)**0.5)/(2.*A) CCONIsGA*(CA+Cv*N(J,2)> DELH=H(J,2)-H(JM,2) CCON3==RHDEHA(E(T(J,2)).H CALL LAYEQ c..... SOLVE4=XMT-XM(J)+CCON6*(HJZ-H(JM,2)) C* M EQUATION XMT=XM(J)-CCON6*(H(J,2)-H(JM,2)) H(J,2)=(XM(J)-XMT)/CCON6+H(JM,2) IF(COOLER .EQ. l.)RETURN IF(H(2,2) .GE. H(1,1))RETURN IF(SYSTEM .EQ. 2.)RETURN HTR=H(202) 3(111)‘HTR END C*sag 3:1 Bussasnlass-uuug= c.. C*.:Inssusnscss-l Icacgsnagsssaaxssuanuca: SUBROUTINE LAYEQ C* asses=8asaansssscssaasxnesssszssszssssss C***** DESCRIPTION C***** SUBROUTINE TO FIND THE MOISTURE CONTENT BASED ON EQUA- C***** TIONS BY J.M. TROEGER AND P.M. DEL GIUDICE c..... C***** USAGE C***** USED IN THE FIXED BED AND CROSSFLOW MODELS WITH GRAIN C***** TEMPERATURES BETWEEN 80 F AND 160 F C..... COMMON/MAIN/XMC,TH,RH,DELT,CFH,XMD,KA3,TOTEN,TOTHZO,XHS, +TOTSP,IPROD,FM,HDF,XMP,TIME,NEQ,NEQI,HOUR,REFHR,REPBU,TOFENY COMMON /NAME/INAME,IPRODU,IPRODU1,IEMC COMMON/COLD/COOLER,SYSTEM DATA INAME,IRRODU,IPRODUI,IEMC/IOHTROETUOM ,IOHTHINLAYER , +IOU CORN ,ION DEBOER / C..... C***** S T A T E M E N T F U N C T I O N S 275 C..... P1(XM,R,T)=EXP(-2.45+6.42*XM**1.25-3.lS*R+9.62*XM*SQRT(R)+.O3*T-.0 102*CFM) P2(R,T)'EXP(2.82+7.49*(R+.01)**.67-.Ol79*T) P3(P,Q)=-.12*(XMD-XME)**(Q+1.)*P*Q Ql(XM,R,T)=-3.98+2.87*XM-(.019/(R+.015))+.016*T Q2(R)'-EXP(.81-3.11*R) TF(P,Q,XO,XF,TO)=P*(XF-XME)**Q-P*(X0 -XME)**Q+TO M(P,Q,XO,TI,TO)=( (TI-T0)/P+(XO-M)**Q)**(1-/Q)+m c..... C***** P R O G R A M c..... IF(RH .LT. 0.)RH=.OOOl C***** CALL READYTH FOR PRELIMINARY CHECKS AND CALCULATIONS CALL READYTH(TXMO,DELM,XME,IOOPS,XMR) C***** CHECK ABSORPTION FLAG...IF SET GO TO ABSORPTION SIMULATION IF(IOOPS-1)1,6,1 1 IF(COOLER .EQ. 1.) GOTO 100 C***** COMPUTE TRANSITION M,P1,Ql, AND FIRST TRANSITION TIME X1M=.4*DELM+KME XZM=.12*DELM+XME TINC=DELT*60. P=P1(TXMD,RH,TH) M1(m0.RH.m) TXBTF(P,Q,TKMO,X1M,O.O) C***** CHECK IF PRESENT M IS IN FIRST REGION...IF IS IS COMPUTE C***** EQUIVALENT TIME AND ADD TINC IF(KMC.LT.X1M) GO TO 3 TI=TF(P,Q,TXMO,XMC,0.0)+TINC C***** CHECK IF EQUIVALENT TIME+TINC IS LESS THAN TRANSITION TIME.. C***** IF IT IS COMPUTE NEW M AND RETURN IF(TI.GT.TX) GO TO 2 KMC= HDF*XMN(P,Q,TRMD,TI,0.0) RETURN C***** EQUIVALENT TIME+TINC IS IN SECOND REGION--COMPUTE P2, QZ AND C***** NEW M THEN RETURN 2 P=P2(RH,TH) Q'QZ(RH) XMC‘ HDF*XMN(P,Q,X1M,TI,TX) RETURN C*'*** M IS NOT IN FIRST REGION--COMPUTE P2, QZ AND SECOND C***** TRANSITION TIME 3 P=P2(RH,TH) Q'QZ(RH) TXl=TX TX'TF(P,Q,XIM,X2M,TXI) C***** CHECK IF PRESENT M IS IN SECOND REGION...IF IT IS COMPUTE C***** EQUIVALENT TIME AND ADD TINC IF(XMC.LT.X2M) GO TO 5 TI'TF(P,Q,XIM,XMC,TK1)+TINC C***** CHECK IF EQUIVALENT TIME+TINC IS LESS THAN TRANSITION TIME.. C****' IF IT IS COMPUTE M AND RETURN 276 IF(TI.GT.TX) GO TO 4 KMC= HDF'XMN(P,Q,X1M,TI,TX1) RETURN C***** EQUIVALENT TIME+TINC IS IN THIRD REGION--COMPUTE P3, Q3 AND C***** NEW M THEN RETURN 4 P=P3(P,Q) Qt-l. XMC= HDF'XMN(P,Q,X2M,TI,TX) RETURN C***** M IS NOT IN SECOND REGION-~COMPUTE P3, Q3, EQUIVALENT TIME+ C***** TINC AND NEW M THEN RETURN S P-P3(P,Q) QI-l. TISTF(P,Q,X2M,XMC,TX)+TINC XMCg HDF*XMN(P,Q,XZM,TI,TX) RETURN ct**** C***** ABSORPTION SIMULATION C***** FIND NEW M AND INCREMENT COUNTER 6 DIV=-.625'PSDB(TH+459.69)**(.466*RH)*RH*RH*RH XMC‘ HDF*((XMC-XME)*EXP(DIV*DELT)+XME) KAB=KAB+1 RETURN 100 IF(IOOPS-l)10,20,10 10 ALMR=ALOG(KMR) A8-1.86178+0.0048843*TH 88427.364*EXP(-0.03301*TH) C**** FIND EQUIVALENT TIME BASED ON CURRENT TEMP AND MC C**** ADD DELT AND SOLVE FOR NEW MC TI!ALMR*(A+B*ALMR)+DELT ALMR=(-A-SORT(A*A+4 . 0*B'TI ) )/( 2 . 0*E) XMC= HDF*(DELM*EXP(ALMR)+XME) RETURN 20 KABBKAB+1 RETURN END C** C*nnun.ssssznnxnt Inuussacsnnznsscsscnsus: SUBROUTINE READYTH(TXMO,DELM,XME,IOOPS,XMR) C*s:xnuznsaasal IE ......... ctuatt C..... C***** DESCRIPTION C***** SUBROUTINE MAKES PRELIMINARY CHECKS AND CALCULATIONS C***** FOR THIN LAYER EQUATIONS c..... c..... USAGE C***** US!“ WITH LATEQ IN FIXED AND CROSS FLOW DRIER MODELS c..... COMMON/MAIN/XMC,TH,RH,DELT,CFM,XMD,KAB,TOTEN,TOTHZO.EMS, +TOTSP,IPROD,FM,HDF,XMF,TIME,NEQ,NEQI,HOUR,REFHR,REFBU,TOFENY COMMON /NAME/INAME,IPRODU,IPRODUI,IEMC C..... c..... 1 C..... C..... 2 3 4 C..... 5 C.. c. 277 IOOPS 3 0 COMPUTE EQUILIBRIUM MOISTURE CONTENT, COMPARE TO PRESENT MOISTURE CONTENT... IF GREATER SET IOOPS = 1 XME 3 EMC(RHITH) IF(XME - XMC) 2,1,1 IOOPS 3 1 COMPARE PRESENT MOISTURE CONTENT TO INITIAL MOISTURE CONTENT SET TXMO 3 THE LARGER VALUE IF(XMO - XMC) 3:494 TXMO 3 EMC GO TO 5 TXMO 3 HMO COMPUTE MOISTURE RATIO DELM 3 TXMO - XME xMR . (EMC - :IMEVDELM RETURN END c. c..... C..... c..... c..... C..... C..... c..... c..... c..... c..... c..... C..... 234 c..... C..... 300 c..... c..... 301 302 303 A 3 FUNCTION EMC(RH,T) E “888888“: S.F.DEBOER, PROGRAMMER IN ENGLISH UNITS AND EDISON W. RUGUMAYO, CONVERTER TO SI UNITS DESCRIPTION FUNCTION COMPUTES EQUILIBRIUM MOISTURE CONTENT OF CORN FROM A RELATIVE HUMIDITY AND TEMPERATURE CHECK TEMPERATURE TO DETERMINE EQUATION TO BE USED TC.(T-320 )/108 IF(TC - 112.778) 234,235,235 DEBOER EQUATIONS CHECK IF RH IS GREATER THAN IF(RH - .50) 300,300,309 PART ONE RH .LE. .50 ONLY COMPUTE CONSTANTS F1 3 -.00070596*TC + .0874496 F2 3 -.00078354*TC + .1188704 F3 3-.00096462*TC + .1474512 51 3 13.838*(-9.*Fl + 6.*F2 - F3) 52 3 13.838*(4.*F3 - 9.*F2 + 6.*F1) B 3 RH - .17 FIND INTERVAL IN WHICH RH LIES AND COMPUTE EQUILIBRIUM MOISTURE CONTENT IF (B) 301,301,302 EMC ' (Sl*RH*RH*RH/l.02 + (Fl/.17 - 51*.02833)*RH) RETURN IF(RH - .34) 303,303,304 .34 - RH EMC 3 (51*A*A*A/1.02 + 52*B*B*B/1.02 + (F2/.17 - 52*.02833)*B + .50 ..IF SO GO TO SECOND PART 278 +(Fl/.17 - 51*.02833)*A) RETURN 304 A 3 .51 - RH EMC 3 52*A*A*A/1.02 + (F3/.17)*(RH - .34) + (F2/.17 - 52*,02333) +*A RETURN C***** PART TWO---RH.GT. .50 ONLY C***** COMPUTE CONSTANTS 309 F0 3 -.000967l4*TC + .1452064 F1 3 -.00127350*TC + .1848600 F2 3 -.00134082*TC + .2293632 F3 3 -.00192780*TC + .3588280 51 - 13.838*(4.*F0 52 c 13.838*(4.*F3 E - RB - .66 IP(B) 305,305,306 C***** FIND INTERVAL IN WHICH RE LIES AND COMPUTE EQUILIBRIUM MOISTURE C***** CONTENT 305 A . RH - .49 EMC c Sl*A*A*A/1.02 + (Fl/.17 - 51*.02833)*A + (F0/.17)*(.66 - RH) RETURN 306 IF(RH - .83) 307,307,308 307 A - .83 - RH EMC . Sl*A*A*A/l.02 + 52*B*B*B/1.02 + (32/.17 - 52*.02833)*B + +(Pl/.17 - 51*.028333)*A RETURN 308 A - 1.0 - RH EMC - 52*A*A*A/1.02 + (F3/.17)*(RH - .83) + (F2/.l7 - 52*.028333) 1*A RETURN cttttt C***** THOMPSON EQUATION C***** TC.GE. 113 DEGREES CELCUS C***** COMPUTE EQUILIBRIUM MOISTURE CONTENT 235 EMC - .01*SQRT((-ALOG(1. - RH))/(.0000382*(1.8*TC + 82.))) RETURN END SUBROUTINE REFILL(REFIN,GLEVELI,XMAVE,COOLER) COMMON/MAIN/NMT,TUT,RHT,DELT,CEM,NMO,RAE,TOTEN,TOTH20,NMS, +TOTSP,IPROD,FM,HDF,XMF,TIME,NEQ,NEQI,HOUR,REFHR,REFBU,TOFENY COMMON/INPT/BPH,GP,INDl,DELX,DEPTH,DBTPR,XTEMPER COMMON/PRPRTY/SA,CA,CV,CW,RHOP,CP COMMON/ARRAxS/XM(50),RE(50),T<50,2),E(50,2),TU(50,2),GA COMMON/HLATENT/HA,HB,HFG COMMON/IELAGS/R,JM,ICON,DELxM COMMON/PRESS/PATM,BCFMF,PG COMMON/COUNT/ITERCT,TIN,THIN,BIN DATA pATM/14.696/ DATA RHC,AREA/.999999,254.469/ F(T)3T+459.69 REPFT3REFBU*I.24/AREA DEPTHF3REFFT+DEPTH 9.*F1 + 6.*F2 - F3) 9.*F2 + 6.*F1 - F0) 279 INDF=INT(DEPTHF/DELX)+1 INS=IND1+1 INDEX3INDF+1 DO 10 I3INS,INDEX XM(I)3XMO TH(I,1)3TH(I,2)3THIN T(I,1)3T(I,2)3T(IND1,1) H(I,1)3H(I,2)3H(IND1,1) RH(I)=RH(IND1) 10 CONTINUE SUM30. IND1=INDF DO 30 I31,IND1 SUM=SUM+XM(I) 30 CONTINUE XMAVE=SUM/IND1 GLEVEL1=DEPTH3DEPTHF PRINT 20,REFBU,DEPTH,HOUR CALL SECANT(DEPTH,BCFMF,CFM,PG,COOLER) GA360.*CFM/VSDBHA(F(TIN),HIN) REFIN=0. ' 20 FORMAT(*0*,*BIN REFILLED WITH *,F5.1,* BU TO A DEPTH OF *,F4.1, +* FT*/* *,*AFTER *,F4.1, * HOURS’) RETURN END REFERENCES 280 REFERENCES Adams, T.N. 1980. A simple fuel bed model for predicting particulate emissions from a wood-waste boiler. Combustion and Flame, Adams, T.N. 1979. Combustible carry over predictions for a wood-waste boiler. Combustion and Flame, 34: 47-61. Anderson, D.G.; and Pfost, D. 1978. Small Holder Grain Storage Problems in Kenya: Problems and Proposed Solutions. USAID, Nairobi, Kenya. Anderson, M.E.; Bern, C.J.; and Baker, J.L. 1983. Corn drying with biomass combustion products. Paper No. 83-3005. ASAE, St. Joseph, MI. Anderson, M.E.; Claar, P.W; and Bern, C.J. 1981. Corn drying evaluation utilizing a concentric-vortex biomass furnace system. Paper No. 81-3015. ASAE, St. Joseph, MI. Antal, M.J.; Edwards, 2.; Friedman, 8.C.; and Roger, P.E. 1970. A study of the steam gasification of organic wastes. Final Report. EPA University Grant No. R-804836010. Washington, DC. ASAE. 1978. Energy-A Vital Resource for the US Food Systems. Cost and Policy Inputs on Agriculture and the Consumer. ASAE, St. Joseph, MI. ASAE. 1982. 1982-1983 Agricultural Engineers Yearbook. ASAE, St. Joseph, MI. Babcock and Wilcox. 1978. Steam: Its Generation and Use. The Babcock and Wilcox Co., New York, NY. Bakker-Arkema, P.W.; Brook, R.C.; and Lerew, L.E. 1978. Cereal grain drying. In Cereal Science and Technology. Vol. II. Pomeranz, 2.; ed. Am. Assoc. Cereal Chemists Inc., St. Paul, MN. Bakker-Arkema, P.W.; Fontana. C.; and Westlaken, C.M. 1982. Concurrent flow drying of rice. Paper No. 82-3068. ASAE, St. Joseph, MI. 281 282 Bakker-Arkema, F.W.; Fontana, C.; Fedewa, G.L.; and Schisler, I.P. 1983. A comparison of drying rates of different grains. Paper No. 83-3009. ASAE, St. Joseph, MI. Bakker-Arkema, F.W.; Lerew, L.B.; DeBoer, S.F.; and Roth, M.G. 1974. Grain Drying Simulation. Research Report No. 224. Agr. Exp. Station, Michigan State University, East Lansing, MI. Bakker-Arkema, F.W.; Silva, J.S.; Mwaura, 8.13.; Rodriguez, J.C.; and Brook, R.C. 1980. Testing of alternative on-farm grain drying systems. Paper No. 80-3017. ASAE, St. Joseph, MI. Barkshi, A.S.; and Singh, R.P. 1979. Drying of high moisture parboiled rice. Paper No. 79-6006. ASAE, St. Joseph, MI. Barrett, J.R.; Jacko, R.B.; and Summer, B.R. 1980. Particulate emissions from biomass furnaces. Part 2. Characterization. Beall, F.C. 1968. Specific heat of wood—further research required to obtain meaningful data. Forest Products Laboratory. US Forest Service, Research Note FPL-0184. Madison, WI. Bliss, C.; and Black, D.O. 1977. Silvicultural Biomass Farms. Vol. 5, Conversion Processes and Costs. BRDA contract No. Bx-76-C-01-2081. The MITRE Corporation, McLean, VA. Bradbury, A.G.W.; and Shafizadeh, F. 1980. Role of oxygen chemisorption in low-temperature ignition of cellulose. Combustion and Flame, 37: 85-89. Brekke, O.L.; Griffin, E.L.; and Shove, G.C. 1973. Dry milling of corn artificially dried at various temperatures. Trans. ASAE, 16(4): 761-765. Brink, D.L. 1976. Pyrolysis-gasification-combustion: a process for ulitization of plant material. Applied Polymer, A Symposium, 28: 1377-1382. Brink, D.L.; Massoudi, M.S.; and Sawyer, R.P. 1973. A flow reactor technique for the study of wood pyrolysis. Presented at Fall 1973 meeting of the Western States Section, Combustion Institute. Brokaw, R.S. 1958. Thermophysical properties of hydrocarbon gas mixtures. J. Chem. Phys., 29: 391-406. Brooker, D.B.; Bakker-Arkema, F.W.; and Hall, C.W. 1974. Drying Cereal Grains. The Avi Publishing Co., Westport, CT. Brooker, D.B.; McKenzie, B.A.: and Johnson, 3.x. 1978. The present status of on-farm grain drying. Paper No. 78-3007. ASAB, St. Joseph, MI. 283 Browne, F.L. 1958. Theories of the Combustion of Wood and its Control. USDA Forest Service, Forest Products Laboratory. Report No. 2136. Madison, WI. Browning, C.W.; Brooker, D.B.; George, R.M.; and Browning, C.B. 1971. Batch in-bin drying by alternating heated and unheated air. Trans. ASAE, 14(1): 193-194. Buchele, W.F.; Claar, P.W.; and Marley, S.J. 1981. Development of a Biomass Energy System for Drying Corn. US Dept. of Energy. Contract No. BM-78-S-02-4916. Washington, DC. Burwell, C.C. 1978. Solar biomass energy. An overview of US potential. Science, 199: 1041-1048. Chakraborty, 8.8.; and Long, R. 1978. The formation of soot and polcycyclic hydrocarbons in diffusion flames-Part l. Combustion and Flame, 12: 226-235. Chapman, A.J. 1974. Heat Transfer. McMillan Publishing Co., New York, NY. Chu, S.T.; and Hustrulid, A. 1968. Numerical solution of diffusion equations. Trans. ASAE, 11: 705-710, 715. Claar, P.W.; Colvin, T.S.; and Marley, S.J. 1980a. Economic and energy analysis of potential corn-residue harvesting systems. ASAE, Agricultural Energy-Biomass Energy, Crop Production, 2: 273-279. Claar, P.W.; Buchele, W.F.; and Marley, S.J. 1980b. Development of a concentric-vortex, agricultural-residue furnace. ASAE, Agricultral Energy-Biomass Energy, Crop Production, 2: 349-356. Claar, P.W.; Buchele, W.F.: and Marley, S.J. 1981. Crop-residue-fired furnace for drying grain. Paper No. 81-1032. ASAE, St. Joseph, MI. Connors, W.J.; Johnson, L.N. Sarkanen, K.V.; and Winslow, P. 1980. Thermal degradation of kraft lignin in tetralin. Holsforschung. 34: 29-37. Conter, S.D.; and de Boor, C. 1980. Elementary Numerical Analysis. McGraw-Hill Book Co., New York, NY. Copeland, L.0. 1976. Principles of Seed Science and Technology. Burges Publishing Co., Minneapolis, MN. Crank, J. 1979. The Mathematics of Diffusion. Claredon Press, Oxford, England. Dahlberg, R.W. 1977. Corncobs-an energy source for drying seed. 32nd Corn and Sorghum Research Conference. American Seed Trade Association. 284 Dalpasquale, v.A. 1981. Drying of soybeans in continuous-flow dryers. Unpublished Ph.D. thesis. Agr. Eng. Dept., Michigan State University, East Lansing, MI. Davies, 6.5.; Mitra, A.N.; and Roy, A.N. 1967. Moment transfer studies in ejectors: correlations for single phase and two phase systems. Ind. and Eng. Chem.: Process Design and Development, 6(3): 293-299. Deutshe Forshungsgemeinschaft. 1976. Bewertung van Ruckstanden in Fleisch and Fleischerzeugnissen (Evaluation of Residue in meat and meat products). Rommission zur Prufung Von Ruchstanden in Lebensmitteln. Mitteilung 2. Verlag Boldt, Boppard, West Germany. Duffie, J.A.; and Beckman, W.A. 1980. Solar Engineering of Thermal Processes. John Wiley and Sons, New York, NY. Ebeling, J.M.; and Jenkins, B.M. 1983. Physical and chemical properties of biomass fuels. Paper No. 83-3546. ASAE, St. Joseph, MI. Edwards, J. 1974. Combustion: Formation and Emission of Trace Species. Ann Arbor Publ., Ann Arbor, MI. Fang, P.; and Mcbinnis, 6.0. 1976. Flash pyrolysis of holocellulose from loblolly pine bark. 13 Thermal Uses and Properties of Carbohydrates and Lignins, A Sympossium. Shafizadeh, Sarkanen and Tillman, eds. Academic Press, New York, NY. FAQ. 1983. The 1982 Production Year Book. FAD. Rome, Italy. Fisher, J.C.; 1974. Energy Crises in Perpective. John Wiley and Sons, New York, NY. Fletcher, T.L.: and Harris, 8.8. 1952. Products from the destructive distillation of Douglas fir lignin. TAPPI, 35: 536-541. Flood, C.A.; Sabbah, D.;'Meeker, D.: and Peart, R.M. 1972. Simulation of natural air drying systems. Trans. ASAB, 15: 156-159, 162. Foley, R.M.: and Hooven, D.I.B.v. 1981. Properties and industrial uses of corncobs. In Cereals: A Renewabe Resource, Theory and Practice. Pomeranz, Y.; and Munck, L., eds. Am. Assoc. Cereal Chemists Inc., St. Paul, MN. Fontana, C. 1983. Concurrent-flow versus conventional drying of rice. Unpublished Ph.D. thesis. Agr. Eng. Dept., Michigan State University, East Lansing, MI. Foster, 6.8. 1964. Dryeration-a corn drying process. USDA Agr. Marketing Service Bull. No. 532. USDA, washington, DC. 285 Foster, G.H.; Baker, K.D.; and Tuite, J. 1982. Combination drying with renewable energy. Paper No. 82-3522. ASAE, St. Joseph, MI. Freeman, J.E. 1978. Quality factors affecting value of corn for wet milling. In Proceedings 1977 Corn Quality Conference. University of Illinois, Urbana-Champaign, IL. Friedrich, R.A. 1978. Energy Conservation for American Agriculture. Ballinger Publishing Co., Cambridge, MA. Gebhart, B. 1971. Heat Transfer. McGraw-Hill Book Co., New York, NY. Glassman, I. 1977. Combustion. Academic Press, New York, NY. Goheen, D.W.; Orle, J.V.; and Wither, R.P. 1976. Indirect pyrolysis of kraft black liguors. In Thermal Uses and Properties of Carbohydrates and Lignins, A Symposium. Shafizadeh, Sarkanen and Tillman, eds. Academic Press, New York, NY. Graboski, M.; and Bain, R. 1979. Properties of Biomass relevant to gasification. In A Survey of Biomass Gasification. Volume II. Principles of gasification. SERI, US Dept. of Energy Contract No. EG-77-C-01-4042. Gray, V.H.; Gelder, T.F.; Cochran, R.P.; and Goodykoontz, J.R. 1960. Bonded and Sealed External Insulations for Liquid Hydrogen Fueled Rocket Tanks During Atmospheric Flight. AD 244287; Div. 14. Gray, W.A.; Kilham, J.R.; and Muller, R. 1976. Heat Transfer from Flames. Elek Science, London, England. Gustafson, R.J.; Morey, R.V.; Christensen, C.M.; and Meromick, R.A. 1978. Quality changes during high-low temperature drying. Trans. ASAE, 21(1): 162-169. Hall, D.O. 1979. Biological solar energy conversion for fuels. Nature, 278: 114-117. Handbook of Chemistry and Physics. 1966. Chemical Rubber Co., Cleveland, OH. Handerson, S.M.; and Perry, R.L. 1976. Agricultural Process Engineering. The Avi Publishing Co., Westport, CT. Hangebrauck, R.P.; von Lehmden, D.J.; Meeker, J.R. 1964. Emissions of polynuclear hydrocarbons and other pollutants from heat generation and incineration processes. J. Air Pollution Control Assoc., 14(7): 267-278. Harsh, S.B. 1983. Capital Investment Model-Including Buy or Custom Hire-a Telplan Program. Dept. of Agr. Economics, Michigan State University, East Lansing, MI. 286 Hathaway, I.L.; Yung, F.D.; and Kresselbach, T.T. 1952. The effect of drying temperature on the nutritive value and commercial grade of corn. J. Animal Science, 11: 430-440. Hay, J.M. 1974. Reactive Free Radicals. Academic Press, New York, NY. Helsel, 2.R.; and Wedin, W.F. 1983. Direct combustion energy from crops and crop residues produced in Iowa. Energy in Agr., 1(4): 317-329. Hill, L.D. 1975. Test weight of corn as a measure of quality. In Corn Quality in World Markets. Hill, L.D., ed. Interstate Printers and Publishers, Dansville, IL. Holman, J.P. 1981. Heat Transfer. McGraw-Hill Book Co., New York, NY. Holman, J.P. 1980. Thermodynamics. McGraw-Hill Book Co., New York, NY. Hottel, H.C.; and Sarofim, A.F. 1967. Radiative Heat Transfer. McGraw-Hill Book Co., New York, NY. Howell, J.R.; 1982. A Catalog of Radiation Configuration Factors. McGraw-Hill Book Co., New York, NY. Howlett, K.; and Gamache, A. 1977. Forest and Mill Residues as Potential Source of Biomass. Vol. VI. Final Report. ERDA Contract No. E(49-18)2081. The MITRE Corporation, McLean, VA. Hukill, W.V. 1954. Drying of grain. In Storage of Cereal Grains and Their Products. Anderson, J.A.; and Alcock, A.W., eds. Am. Assoc. of Cereal Chemists, St. Paul, MN. Hukill, W.V.; and Shedd, C.K. 1955. Nonlinear airflow in grain drying. Agr. Eng., 36: 462-466. Hutt, W.; and E. Winkler. 1978. Nontamination van Getreide durch Schadstoffe bei der Directen Trocknung (Contamination of cereals by pollutants during direct drying). Getreide, Mehl und Brot, 32(10): 260-263. Hutt, W.A.; Meiering, A.; 0e1schlager, W.; and Winkler, E. 1978. Grain contamination in drying with direct heating. Canadian Agr. Eng., 20(2): 103-107. Institute of Gas Technology. 1978. Coal Conversion Systems: Technical Data Book. DOE contract Ex-76-C-01-2286. NTIS, Springfield, VA. IRS. 1983. Farmer's Tax Guide. Publ. No. 225. Department of the Treasury, Internal Revenue Service, Richmond, VA. Jacko, R.B.; and Barrett, T.R. 1981. Particulate emissions from biomass furnaces. Part 1. Equipment and procedures. Paper No. 81-3561. ASAE, St. Joseph, MI. 287 Jacko, R.B.; Foster, G.H.; and Barrett, J.R. 1982. Contaminant residues on corn dried by combustion gases from biomass furnaces. Paper No. 82-3523. ASAE, St. Joseph, MI. Jensen, A.H. 1978. The effect of processing (roasting) on the nutritive value of corn for swine. In Proceedings 1977 Corn Quality Conference. University of Illinnois, Urbana-Champaign, IL. Jensen, A.H.; Terril, S.W.; and Becker, D.E. 1960. Nutritive value of corn dried at 140, 180 and 220°F for swine of different ages. J. Animal Science, 19: 629-643. Jensen, D.E.; and Jones, G.M. 1978. Reaction rate coefficent for flame calculation. Combustion and Flame, 32(1): 1-34. Joe, F.L.; Salemme, J.; and Fazio, T. 1982. High performance liquid chromatography with flourescence and ultraviolet detection of polynuclear aromatic hydrocarbons in barley malt. J. Assoc. of Anal. Chem., 65(6): 1395-1402. Junge, D.C. 1977. Boilers Fired with Wood and Bark Residues. Forest Research Laboratory. Research Bulletin 17. Oregon State University, Corvallis, OR. Junge, D.C. 1977. Investigations of the Direct Combustion of Wood Residue Fuel. Progress Report No. 1. Oregon State University, Corvallis, OR. Junge, D.C. 1979. Design Guideline Handbook for Industrial Spreader Stoker Boilers Fired with Wood and Bark Residue Fuels. Contract No. E.Y. 76-C-06-2227. Task Agreement No. 22. US Dept. of Energy, washington, DC. Hajewski, A.; Marley, S.J.; and Buchele, W.F. 1977. Drying corn with a corp residue-fired furnace. Paper No. 77-3525. ASAE, St. Joseph, MI. Kalchik, S.J.: Silva, J.S.; Bakker-Arkema, F.W.; and Miller, 8.5. 1979. An engineering-economic comparison of five drying techniques for shelled corn on Michigan Farms. Paper No. 79-3518. ASAE, St. Joseph, MI. Kalchik, S.J.: Silva, J.S.; Mwaura, E.N.: Rodriguez, J.C.; and Bakker-Arkema, F.W. 1981. On-farm corn drying comparison. Paper No. 81-3011. ASAE, St. Joseph, MI. Kalchik, S.J. 1984. Personal communication. Kalchik Farms, Bellaire, MI. Kanury, A.M. 1975. Introduction to Combustion Phenomena. Gordon and Breach Sci. Publ., New York, NY. 288 Kanury, A.M. 1972. Thermal decomposition kinetics of wood pyrolysis. Combustion and Flame, 18: 75-83. Kastner, L.J.; and Spooner, J.R. 1950. An investigation of the performance and design of the air ejector employing low pressure as the driving fluid. Proc. Inst. Mech. Eng., 162(2): 149-159. Keenan, J.H.; Newman, E.P.; and Lustwerk, F. 1950. An investigation of ejector design by analysis and experiment. J. Applied Mech. September, 1950: 299-309. Kenzler, G.A. Stone, M.L. 1982. Performance of a direct combustion biomass furnace. Paper No. 82-3604. ASAE, St. Joseph, MI. Kirk, D.E. 1959. Column thickness for shelled corn dryers. Kirk, R.R.; and 0thmer, D.F., eds. 1963. Encyclopedia of Chemical Technology. Wiley Interscience, New York, NY. Kroll, A.E. 1974. The design of jet pumps. Chem. Eng. Progress, 1(2): 21-24. Kung, H.H. 1972. A mathematical model of wood pyrolysis. Combustion and Flame, 18: 185-195. Lijinsky, W.; and Ross A. 1967. Production of carcinogenic polynuclear hydrocarbons in cooking of food. Food and Cosmetics, 5: 343-351. Liu, K.T.; Stambaugh, R.P.; Nack, H.; and Oxley, H. 1977. Pyrolytic gasification of kraft black liquors. In Fuels from Wastes. Tillman, D.A., ed. Academic Press, New York, NY. Mason, E.A.; and Saxena, S.C. 1958. Thermal properties of gas mixtures. Physics of Fluids, 1: 361-369. MacCallum, C. 1979. The sloping grate as an alternative to the travelling grate on hog fueled boilers. In Hardware for Energy Generation in the Forest Products Industry. Forest Products Res. Soc., Madison, WI. McAdams, W.H. 1954. Heat Transmission. McGraw-Hill Book Co., New York, NY. McCoy, E.S.; Barrett, J.R.; and Richey, C.B. 1981. Analysis of mass and energy transfers in a biomass gasifier/combustor furnace. Paper No. 81-3017. ASAE, St. Joseph, MI. McKenzie, B.A.; Foster, G.H.; Noyes, R.T.; and Thompson, R.A. 1972. Dryeration-better corn quality with high speed drying. Co-op. Ext. Serv. Bull. AE-72. Purdue University, West Lafayette, IN. 289 MDA. 1983. Michigan Agricultural Statistics. Mich. Dept. of Agr., Lansing, MI. Merril, R.; and Gage, T. 1978. Energy Primer, Solar, Water, Wind and Biofuels. Delta, New York, NY. Milne, T. 1979. Pyrolysis-the thermal behavior of biomass below 600°C. In a survey of Biomass Gasification. Vol. II. Principles of Gasification. SERI, Golden, CO. Misra, M.K. 1978. Thin layer drying and rewetting equations for shelled yellow corn. Unpublished Ph.D. thesis. Agr. Eng. Dept., University of Missouri, MO. MDA. 1981. Request For Technical Proposals: On-Farm Grain Storage. AID Project No. 615-0190. Ministry of Agriculture, Government of Kenya, Nairobi, Kenya. Mohsenin, N.N. 1980. Thermal Properties of Foods and Agricultural Materials. Gordon and Breach Sci. Publ., New York, NY. Moraczewski, B. 1980. Grow your own fuel, grain dryers now on the market. Farm Industries News, 14(3): 90-101. Morey, R.V.; Cloud, H.A.; and Lueschen, W.E. 1976. Practices for efficient utilization of energy for drying corn. Trans. ASAE, 19(14): 151-155. Morey, R.V.; Thimsen, D.J.; Lang, J.P.; and Hansen, D.J. 1982. A corncob fueled drying system. Paper No. 82-3518. ASAE, St. Joseph, MI. Morey, R.V.; and Thimsen, D.R. 1980. Two stage combustor to provide heat for drying corn. Paper No. 80-3507. ASAE, St. Joseph, MI. Muh, K.K. 1974. Determination of the coefficients of a thin layer equation for corn. Unpublished M.S. thesis. Agr. Eng. Dept., Michigan State University, East Lansing, MI. Muhlbauer, W.; Kuppinger, H.; and Isaacs, G.W. 1978. Design and operating conditions of single stage concurrent-counter flow grain dryers. Appendix 82. ASAE, St. Joseph, MI. Mulcahy, M.F.R.; and Young, C.C. 1957. The reaction of hydroxyl radicals with carbon at 298 K. Carbon, 13(2): 115-124. Nicholls, P.; and Eilers, M.G. 1934. The principles of underfeed combustion and the effects preheated air on overfeed and underfeed fuel beds. Trans. ASME., 56: 321-336. Ogan, K.; Katz, E.; and Slavin, W. 1979. Determination of polycyclic aromatic hydrocarbons in aqueous samples by reverse-phase liquid chromatography. Anal. Chem., 51(8): 1315-1323. 290 Palmer, H.B. 1974. Equilibrium and kinetics in flames. In Combustion Technology: Some modern Developments. Palmer, H.B.; and Beer, J.M., eds. Academic Press, New York, NY. Parkes, D.H. 1982. Personal communication. Shivvers Corporation, Corydon, IA. Payne, F.A.; Alphin, 3.6.; and Bunn, J.M. 1981. A cordwood gasification and gas combustion system. Paper No. 81-3589. ASAE, St. Joseph, MI. Payne, F.A., Ross, I.J.; Walker, J.N.; and Brashear, R.S. 1980. Gasification-combustion of corncobs and analysis of exhaust. Paper No. 80-3025. ASAE, St. Joseph, MI. Peill, J. 1980. Utilizing Biomass (Straw) for Crop Drying and Heating. Report No. 80-03-8-1801. Canadian Dept. of Agr., ERSI. Perry, R.H.; and Chilton, C.H. 1973. Chemical Engineers Handbook. McGraw-Hill Book Co., New York, NY. Pfost, H.B.; Maurer, 5.; Chung, D.S.; and Millikan, G. 1976. Summarizing and reporting equilibrium moisture data for grains. Paper No. 76-3520. ASAE, St. Joseph, MI. Pimentel, D.; Moran, M.A.; Fast, S.; Weber, G.; Bukantis, R.; Balliett, L.; Boveng, P.; Cleveland, C.; Hindman, 5.; and Young, M. 1981. Biomass Energy from crop and forest residues. Science, 212: 1110-1115. Prahacs, 5.; Barclay, H.G.; and Bhaba, S.P. 1971. A study of the possibilities of producing synthetic tonnage chemicals from lignocellulosic residues. Pulp and Paper Magazine of Canada, 72: 69-75. Richey, C.B.; Barrett, J.R.: and Foster, G.H. 1980. A biomass gasification furnace. Paper No. 80-3512. ASAE, St. Joseph, MI. Richey, C.B.; Barrett, J.R.; Foster, G.H.; and Kutz, L.J. 1981. Biomass down draft channel gasifier-furnace for drying corn. Paper No. 81-3590. ASAE, St. Joseph, MI. Roberts, A.F. 1970. A review of kinetics data for the pyrolysis of wood and related substances. Combustion and Flame, 14: 261-272. Rodriguez, J.C. 1982. Energy and grain quality characteristics of cross-flow and concurrent-flow dryers. Unpublished Ph.D. thesis. Agr. Eng. Dept., Michigan State University, East Lansing, MI. Ross, I.J.: and White, G.M. 1972. Discoloration and stress cracking of white corn as affected by overdrying. Trans. ASAE, 15(2): 327-329. 291 Rugumayo, E.W. 1979. Corn drying with solar heated air. Unpublished Ph.D. thesis. Agr. Eng. Dept., Michigan State University, East Lansing, MI. Rutledge, J.H. 1978. The value of corn quality to the dry miller. In Proceedings 1977 Corn Quality Conference. University of Illinois, Urbana-Champaign, IL. Sabbah, M.A. 1971. Convective heat transfer coefficient and tempering effects on cooling rates of shelled corn. Unpublished Ph.D. thesis. Agr. Eng. Dept., Purdue University, West Lafayette, IN. Sabbah, M.A.; Meyer, G.E.; Keener, H.M.; and Roller, W.L. 1979. Simulation studies of reversed-direction airflow drying method for soybean seed in a fixed bed. Trans. ASAE, 22(5): 1162-1166. Sailor, D. 1978. Dietary carcinogens. Fortsehr. Med. 96(9): 446-449. Santodonato, J.; Basu, D.; and Howard, P. 1980. Multimedia human exposure on carcinogenic risk assessment for environmental PAH. In Polynuclear Aromatic Hydrocarbons: Chemistry and Biological Effects. Fourth International Symposium on Polynuclear Aromatic Hydrocarbons. Battle Press, Columbus, OH. SERI. 1979. A Survey of Biomass Gasification. Vol. II. Principles of Gasification. Solar Energy Res. Inst., Golden, CO. Shafizadeh, F. 1968. Pyrolysis and combustion of cellulosic materials. Advances in Carbohydrate Chemistry, 23: 419-428. Shafizadeh, F.; and Chin, R.S. 1977. Thermal deterioration of wood. In wood Technology: Chemical Aspects. Goldstine, I.S., ed. ACS Press, Washington, DC. Shafizadeh, F.: and DeGroot, W.F. 1977. Thermal analysis of forest fuels. In Fuels and Energy from Renewable Resources. Tillman, D.A.; Sarkanen, K.V.; and Anderson, L.L., eds. Academic Press, New York, NY. Shafizadeh, F.; and DeGroot, W.F. 1976. Combustion characteristics of cellulosic fuels. In Thermal Uses and Properties of Carbohydrates and Lignins. Shafizadeh, F.; Sarkanen, K.V.; and Tillman, D.A., eds. Academic Press, New York, NY. Shafizadeh, F.; and Lai, Y.z. 1972. Thermal degradation of 1-6-anhydro-8-D-glucopyranuse. J. Org. Chem., 37: 278-284. Shaw, L.N.; Post, D.M.; and Eoff, K.M. 1982. A biomass energized crop dryer. Paper No. 82-3521. ASAE, St. Joseph, MI. 292 Shove, G.C. 1978. Corn drying with low temperature, high temperature combination. Paper No. 78-305. ASAE, St. Joseph, MI. Shove, G.C.; and White, G.M. 1977. Corn quality as affected by drying procedures. In Proceedings 1977 Corn Quality Conference, University of Illinois, Urbana-Champaign, IL. Shrader, W.D. 1977. Impact on soils of residue removal. In Proc. Conference on the Production of Crop Residues, Forage Legumes and Grasses for Conversion to Fuels and Chemicals. Midwest Res. Inst., Kansas City, MO. Siegel, R.; and Howell, J.R. 1981. Thermal Radiation Heat Transfer. McGraw-Hill Book Co., New York, NY.' Silva, J.S. 1980. An engineering economic comparison of five drying techniques for shelled corn on Michigan farms. Unpublished Ph.D. thesis. Agr. Eng. Dept., Michigan State University, East Lansing, MI. Singh, R.; Maheshulari, R.C.; and tha, T.P. 1980. Development of a husk fired furnace. J. Agr. Eng. Res., 25: 109-120. Skaar, C. 1972. Water in Wood. Syracuse Wood Science Series No. 4. Syracuse University Press, New York, NY. Souci, S.W. 1968. Uber den Einfluss von Verbrennungsgasen auf Lebensmittel und Futtermittel. Deutsche Lebensmittel Rundshau, 64(8): 235-241. Spiers, H.N. 1962. Technical Data on Fuel. John Wiley and Sons, New York, NY. Srivastava, A.H.; and Posselius, J.H. 1981. Dimentional analysis and similitude modeling of the gasification process. Paper “00 81-355880 ASAE, Ste JOSEph, MI. Stamm, A.J.; and Loughborough, W.H. 1935. Thermodynamics of the swelling of wood. J. Phys. Chem., 39: 121-132. Steffe, J.F. 1979. Moisture diffusion and tempering on the drying of rough rice. Unpublished Ph.D. thesis. Agr. Eng. Dept., University of California, Davis, CA. Steffgen, F.W. 1974. Energy from agricultural products. In A New Look at Energy Sources. McCloud, D.E., ed. Am. Soc. Agron. Madison, WI. Stobaugh, R.; and Yergin, D., eds. 1980. Energy Future. Ballantine Books, New York, NY. Stout, B.A. 1979. Agricultural biomass for fuels. Agr. Res. Inst., Washington, DC. 293 Strehler, A. 1976. Straw as fuel-theoretical possibilities and practical use. Deutshe Tiersuchter Adenauer Allee, Bonn, W. Germany Sukup, C.E.; Bern, C.J.; and Buchele, W.F. 1982. Performance of a biomass furnace for grain drying. Paper No. 82-3524. ASAE, St. Joseph, MI. Sullivan, J.E.; Costa, P.M.; Owens, F.N.; Jensen, A.H.; Wikoff, K.E.; and Hatfield, E.E. 1975. The effect of heat on nutritional value of corn. In Corn Quality in World Markets. Hill, L.D., ed. Interstate Printers and Publishers, Danville, IL. Sutherland, J.W. 1975., Batch grain dryer design and performance prediction. J. Agr. Eng. Res., 20: 423-432. Thimsen, D.P.; and Morey, R,V. 1981. Exhaust analysis of two-stage downdraft biomass combustion . Paper No. 81-3591. ASAE, St. Joseph, MI. Thompson, R.A.; and Foster, G.H. 1963. Stress cracks and breakage in artificially dried corn. USDA Marketing Research Report No. 631. USDA, Washington, DC. Thompson, T.L.; Peart, R.M.; and Foster, G.H. 1968. Mathematical simulation of corn drying-a new model. Trans. ASAE., 11(4): 582. Tillman, D.A.; Rossi, A.J.; and Kitto, W.D. 1981. Wood Combustion Principles, Processes, and Economics. Academic Press, New York, NY. Tillman, D.A. 1980. Fuels from waste. Kirk-Othmer Encyclopedia of Chemical Technology. Vol. II. John Willey and Sons, New York, NY. Troeger, J.M.; and Hukill, W.V. 1970. Mathematical description of the drying rate of fully exposed corn. Paper No. 70-324. ASAE, St. Joseph, MI. Tuttle, K.L. 1978. Combustion mechanisms in wood fired boilers. Unpublished Ph.D. thesis. Agr. Eng. Dept., Oregon State University, Corvallis, OR. US Forest Products Laboratory 1974. wood Handbook: Wood as an Engineering Material. USGPO, Washington, DC. USDA. 1982. Crop Production, 1981 Annual Summary. Statistical Reporting Service, USDA, Washington, DC. USDA Forest Service. 1982. Wood-Fired Boiler Systems for Space Heating. Vol. I. Design and Evalution. USDA Forest Service, Publ. No. EM 7180-2. washington, DC. 294 Wahby, M.F.I.; Claar, P.W.; and Buchele W.F. 1981. Combustion evaluation of a concentric-vortex cell furnace. Paper No. 81-3596. ASAE, St. Joseph, MI. Watson, S.A.; and Mirata, Y. 1962. Some wet-milling properties of artificially dried corn. Cereal Chem., 39: 35-44. Wenzel, H. 1970. The Chemical Technology of Wood. Academic Press, New York, NY. Westelaken, C.M.; and Bakker-Arkema, F.W. 1978. Concurrent-flow grain drying. Paper No. 78-0712. ASAE, St. Joseph, MI. Wiggins, D.; and Krieger, B.B. 1980. Effect of heating and quenching rates on volatiles produced from combustion-level-heat-flux pyrolysis of biomass. Proc. AICHE. 89th Natl. Mtg., Portland, OR 0 Williamson, J.L. 1975. Nutritional requirement of Livestock as related to corn quality. In Corn Quality in World Markets. Hill, L.D., ed. Interstate Printers and Publishers, Danville, IL. Winkler, E. 1977. Untersuchungen zur Spurenanalyse von Polyzyklischen aromatischen Kohlenwasserstoffen. Dissertation, Inst. Fur Organiche Chemie. Univ. Stuttgart, West Germany. Zink, H.; Brook, R.C.; and Peart, R.M. 1978. Engineering analysis of energy source for low temperature drying. Paper No. 78-3517. ASAE, St. Joseph, MI. Zink, H.R.; Bozdech, S.L.; and Popp, J.L. 1982. Seed corn drying with cob gasification. Paper No. 82-3607. ASAE, St. Joseph, MI.